Download molecular-genetic study of polygenic diseases with a special focus

Document related concepts

Hygiene hypothesis wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Gene therapy of the human retina wikipedia , lookup

Gene therapy wikipedia , lookup

Race and health wikipedia , lookup

Genetic engineering wikipedia , lookup

Public health genomics wikipedia , lookup

Epidemiology of metabolic syndrome wikipedia , lookup

Transcript
THIRD FACULTY OF MEDICINE
CHARLES UNIVERSITY
Division of Cell and Molecular Biology
Institute of Biochemistry, Cell and Molecular Biology
MOLECULAR-GENETIC STUDY OF POLYGENIC
DISEASES WITH A SPECIAL FOCUS ON DIABETES
MELLITUS
Čejková Pavlína
Ph.D. Thesis
Prague, July 2nd, 2008
This thesis was created with the assent of Division of Cell and Molecular Biology, Institute of
Biochemistry, Cell and Molecular Biology, Third Faculty of Medicine, Charles University at
Prague within the scope of Doctoral study of Biomedicine under the division council Molecular
and Cell Biology, Genetics and Virology.
Corresponding address of the author:
Division of Cell and Molecular Biology
Institute of Biochemistry, Cell and Molecular Biology
Third Faculty of Medicine, Charles University
Ruska 87
Prague 10
100 00
Tel.: +420 267 102 540
Fax: +420 267 102 650
Email: [email protected]
Tutors:
RNDr. Emanuel Žďárský, CSc.
Doc. MUDr. Marie Černá, CSc.
Division of Cell and Molecular Biology
Institute of Biochemistry, Cell and Molecular Biology
Third Faculty of Medicine
Charles University
ACKNOWLEDGEMENTS
This study was performed during the years 2000-2008 in the laboratories of Division of
Cell and Molecular Biology, Charles University in Prague, Czech Republic, and at the
Department of Molecular Diagnostics, Roswell Park Cancer Institute, Buffalo NY, U.S.A. I am
thankful to these institutions for providing excellent facilities and scientific background for my
doctoral study.
I would like to thank my tutors, RNDr. Emanuel Žďárský, CSc. and Doc. MUDr. Marie
Černá, CSc., for giving me the unique opportunity for professional growth and for their
extraordinary guidance and incentive advices during my doctoral study. My grateful thanks
belong also to Petr Starostik, M.D., the chief of the Department of Molecular Diagnostics,
RPCI, Buffalo NY, who was my supervisor during my two-year research experience at his lab.
He showed me a beauty of cancer research and thought me the basics of scientific cogitation.
Many thanks belong to all coauthors and colleagues from the Division of Cell and
Molecular Biology for creating an excellent working atmosphere, and for their professional help
and encouragement.
My special thanks go to my family, my parents and sister above all, for their patience
and support in the tough moments of this period.
This study was financially supported by the Research Program of the Czech Ministry of
Education and Youth: Identification code: MSM 0021620814: Prevention, diagnostics and
therapy of diabetes mellitus, metabolic and endocrine damage of organism, and the Research
Programs of Ministry of Health of the Czech Republic, IGA MZ CR: NB/7432-3/2003 and IGA
MZ CR: NR/8191-3/2004.
CONTENTS
ACKNOWLEDGEMENTS .......................................................................................................................3
CONTENTS ................................................................................................................................................4
ABBREVIATIONS .....................................................................................................................................6
LIST OF TABLES ......................................................................................................................................9
LIST OF FIGURES .................................................................................................................................. 10
CHAPTER 1 ................................................................................................................................................1
OVERVIEW OF CURRENT KNOWLEDGE .........................................................................................1
1.1 POLYGENIC DISEASES............................................................................................. 2
1.2 CLASSIFICATION OF VARIATION ............................................................................. 3
1.2.1
Physiological variabilty.................................................................................................................. 3
1.2.2
Mutations.......................................................................................................................................... 9
1.2.2.1
Classification of mutations ................................................................... 9
1.2.2.2
Pathological changes in tumors ......................................................... 10
1.3 QUANTITATIVE TRAIT LOCI (QTL) ....................................................................... 13
1.3.1
Methods for QTL analysis .......................................................................................................... 15
1.3.1.1
Experimental species .......................................................................... 15
1.3.1.2
Mapping QTLs in humans .................................................................. 19
1.3.1.3
Advantages and drawbacks of linkage analysis and association study
............................................................................................................ 20
1.3.1.4
QTL in human cancer disease ............................................................ 21
1.4 DISEASES WITH UNKNOWN QTLS ........................................................................ 27
1.4.1
Enteropathy-type T-cell lymphoma ......................................................................................... 27
1.4.1.1
Classification of ETL .......................................................................... 27
1.4.1.2
Environmental risk factors ................................................................. 28
1.4.1.3
Genetic risk factors ............................................................................. 30
1.5 DISEASES WITH KNOWN QTL............................................................................... 35
1.5.1
Diabetes mellitus ........................................................................................................................... 35
4
1.5.1.1
Classification ...................................................................................... 35
1.5.1.2
Type 1 diabetes mellitus (T1D) ........................................................... 37
1.5.1.2.1
Incidence and prevalence of T1D ....................................................... 37
1.5.1.2.2
Immunology of T1D ............................................................................ 38
1.5.1.2.3
Environmental risk factors.................................................................. 39
1.5.1.2.4
Genetic risk factors .............................................................................. 41
1.5.1.3
1.5.2
Type 2 diabetes mellitus (T2D) ........................................................... 51
1.5.1.3.1
Incidence and prevalence of T2D ....................................................... 51
1.5.1.3.2
Metabolic syndrome............................................................................. 52
1.5.1.3.3
Environmental risk factors.................................................................. 52
1.5.1.3.4
Genetic risk factors .............................................................................. 54
Other autoimmune diseases........................................................................................................ 56
1.5.2.1 Systemic lupus erythematosus ................................................................................................... 56
1.5.2.2
Rheumatoid arthritis ........................................................................... 62
1.5.2.3
Psoriatic arthritis................................................................................ 65
1.6 THE JOINT MARKER OF ENTEROPATHY-TYPE T-CELL LYMPHOMA AND
AUTOIMMUNE DIABETES MELLITUS ...................................................................... 68
CHAPTER 2 .............................................................................................................................................. 72
AIMS OF THE STUDY AND METHODOLOGY ................................................................................ 72
2.1 AIMS .................................................................................................................... 72
2.2 METHODOLOGY ................................................................................................... 73
CHAPTER 3 ............................................................................................................................................ . 74
RESULTS .................................................................................................................................................. 74
3.1 LIST OF ORIGINAL PAPERS .................................................................................... 74
3.2 COMMENTARY TO THE ORIGINAL PAPERS ............................................................. 76
3.3 ORIGINAL PAPERS ................................................................................................ 80
CHAPTER 4 ........................................................................................................................................ 15543
CONCLUSIONS ................................................................................................................................... 1553
REFERENCES........................................................................................................................................ 157
5
ABBREVIATIONS
ABL1
Abelson murine leukemia viral oncogene homolog 1
A-CGH
Array-based comparative genome hybridization
ADA
American Diabetes Association
AGA
Anti-gliadin antibody
ALCL
Anaplastic large cell lymphoma
ANA
Antinuclear autoantibody
APC
Antigen-presenting cell
BAC-CGH
Bacterial artificial chromosome-based comparative genome hybridization
YAC-CGH
Yeast artificial chromosome- based comparative genome hybridization
BMI
body mass index
CAPN10
Calpain 10
CD
Celiac disease
cDNA-CGH cDNA- based comparative genome hybridization
CF
Cystic fibrosis
CFTR
Cystic fibrosis transmembrane conductance regulator
CGH
Comparative genome hybridization
CTLA4
Cytotoxic T-lymphocyte associated protein 4
DM
Diabetes mellitus
DNA
Deoxyribonucleic acid
ddNTP
Dideoxy nucleotide triphosphates
EBV
Epstein-Barr virus
EMA
Anti-endomysium antibody
ETL
Enteropathy-type T-cell lymphoma
F1, F2
Filial generation 1, filial generation 2
FISH
Fluorescein in situ hybridization
FITC
Fluorescein isothiocyanate
GAD
Glutamic acid decarboxylase
6
GADA
Glutamic acid decarboxylase autoantibody
HLA
Human leukocyte antigen
HNF
Hepatocyte nuclear factor
HS
Heterogeneous stock
Htg
High-temperature growth
HTLV-1
Human T-lymphotropic virus
ICA
Islet-cell autoantibody
ICA512/IA-2 Protein thyrosine phosphatase
IDDM
insulin-dependent diabetes mellitus
IEL
Intraepithelial lymphocytes
IGF2
Insulin-like growth factor 2
IL
Interleukin
KCNJ11
Potassium inwardly-rectifying channel, subfamily J, member 11
KIR
Activating killer immunoglobulin-like receptor
Kir6.2
Inwardly rectifying K+ channel
LADA
Latent autoimmune diabetes in adults
LCR
Low-copy repeats
LD
Linkage disequilibrium
LINE
Long interspersed nuclear elements
LOH
Loss of heterozygosity
MHC
Major histocompatibility complex
MIC-A
MHC class I chain – related gene A
MODY
Maturity-onset diabetes of young
mRNA
Messenger ribonucleic acid
MSI
Microsatellite instability
NAHR
Non-allelic homologous recombination
NFKB1
Nuclear factor kappa B, subunit 1
NFKBIA
Nuclear factor of kappa light chain gene enhancer in B–cells inhibitor α
NHL
Non-Hodgkin lymphoma
NIDDM
non-insulin-dependent diabetes mellitus
NK cells
Natural killer cells
NKG2D
C-type lectin like NK receptor
NOTCH1
NOTCH, Drosophila homolog 1
PCR
Polymerase chain reaction
7
PPARγ
Peroxisome proliferator-activated receptor γ
PsA
Psoriatic arthritis
PTPN22
Protein thyrosine phosphatase non-receptor type 22
QPCR
Quantitative polymerase chain reaction
QTL
Quantitative trait locus
RA
Rheumatoid arthritis
RB
Retinoblastoma
RCD
Therapy-refractory celiac disease
RNA
Ribonucleic acid
rRNA
Ribosomal ribonucleic acid
SINE
Short interspersed nuclear elements
SLE
Systemic lupus erythematosus
SNP
Single nucleotide polymorphism
SSR
Simple sequence repeats
STR
Short tandem repeats
SUR1
Sulphonylurea receptor-1
T1D
Type 1 diabetes mellitus
T2D
Type 1 diabetes mellitus
TALL
T-cell acute lymphoblastic leukemia
TCR
T-cell receptor
TGFβ
Transforming growth factor β
TNF
Tumor necrosis factor
TP53
Transcription protein 53
tRNA
Transfer ribonucleic acid
TSG
Tumor suppressor gene
tTG
Tissue transglutaminase
VNTR
Variable number tandem repeats
WHO
World Health Organization
8
LIST OF TABLES
Table 1: Number of recurrent chromosomal gains and losses published so far ............. 31
Table 2: Autoantigens in type 1 diabetes mellitus ......................................................... 39
Table 3: List of IDDM loci ............................................................................................ 42
Table 4: Some of the non-IDDM loci associated with both types of diabetes .............. 43
Table 5: HLA class II haplotypes associated with susceptibility for or protection against
type 1 diabetes .................................................................................................. 46
Table 6: Environmental factors that may be relevant in the pathogenesis of systemic
lupus erythematosus........................................................................................ 59
Table 7: Genes involved in human systemic lupus erythematosus ............................... 60
Table 8: List of some genes and potential candidates involved in etiology of rheumatoid
arthritis ............................................................................................................ 65
9
LIST OF FIGURES
Figure 1: Deletions, duplications, inversions and gene conversions ............................... 8
Figure 2: Comparison of cumulative data for human Mendelian trait genes and
complex trait genes ......................................................................................... 14
Figure 3: Reciprocal hemizygosity ................................................................................ 18
Figure 4: Comparative genome hybridization – procedure ........................................... 22
Figure 5: Comparative genome hybridization – results................................................. 23
Figure 6: Fluorescence in situ hybridization – schematic illustration ........................... 24
Figure 7: Suggested classification of two ETL subtypes based on the clinical,
morphologic, immunophenotypic and genetic features ................................ 34
Figure 8: Overview of the extended HLA complex on the chromosome 6 ................... 44
Figure 9: Insulin VNTR minisatellite repeats................................................................ 47
Figure 10: Insulin gene variants that affect pre-mRNA splicing................................... 48
Figure 11: Kir6.2/SUR1 pancreatic β-cell ATP-sensitive K+ channel complex ........... 62
Figure 12: Genomic organization of prolactin gene ...................................................... 70
Figure 13: Multistage development of ETL .................................................................. 68
10
CHAPTER 1
OVERVIEW OF THE CURRENT KNOWLEDGE
The biological mechanisms that link together genetic variation and its phenotypic
outcome have been for long decades one of the most challenging puzzles of biology,
especially in the field of genetics. This is based on the faith that the genetic determinants
behind complex traits are tractable and that knowledge of genetic variation and its
contribution to a pathological phenotype will improve the diagnosis, treatment and
prevention of a significant portion of cases of the diseases that constitute major public
health burden of industrialized nations. (Un)fortunately, it is not only about gene(s) that
contribute to a specific phenotype, but also some additive effects as well as epistatic
interactions need to be taken into account since majority of the common human diseases,
like a cancer, heart disease, diabetes, schizophrenia etc. have complex etiologies involving
the action of many genes, as well as dynamic gene-environment interactions.
To elucidate the mechanisms underlying disease susceptibility and progression and
to improve diagnosis and treatment, an important strategy is to use proper genetic methods
that enable identifying of the causative DNA variants, and to use this knowledge as the first
step towards the eventual unraveling of the complex interplay between genes and
environment. Geneticists have approached this problem by trying to uncover genetic
variants that underlie the trait in question. It is possible to study both common diseases
(because of their direct medical relevance) and phenotypic traits (because they are easier to
accurately measure, often have a stronger genetic component, and often represent
themselves risk factors for the disease).
1
1.1 POLYGENIC DISEASES
A polygenic disorder (also termed multifactorial or complex) is caused by combined
effect of multiple genes and lifestyle and environmental factors. Examples of polygenic
diseases include coronary heart disease, hypertension, schizophrenia and diabetes.
Compared to its counterpart - monogenic disorders, which are associated with defect in a
single gene and whose inheritance follows simple Mendelian patterns, a pattern of
inheritance of complex disorders is not explicit. It is thus difficult to determine person’s
risk of inheriting or passing on such a disease. It is a simultaneous presence of several
genes responsible for the disease manifestation and the fact that specific factors causing
most of these disorders have not yet been identified what makes the complex diseases
difficult to study and treat. Moreover, multiple genes participating in pathological
phenotype may interact with each other with an unpredictable effect on the expression
(epistasis).
One of the basic problems may be the disease classification itself: Even single-gene
disorders inheritance of which can be explained by the classical Mendelian laws may lack
genotype-phenotype correlation and display complex phenotype. This discrepancy can be
enlightened by a “threshold model” [1] suggesting a presence of other independently
inherited factors, such as modifier genes, that influence the clinical phenotype by turning
"simple" Mendelian disorders into complex traits. This phenomenon may perhaps be
documented by cystic fibrosis, an autosomal recessive disease caused by mutations in the
cystic fibrosis transmembrane conductance regulator (CFTR). The phenotypic differences
described in this illness seem to be caused by modifier genes influencing the function of
CFTR gene and thus also patients’ phenotypes.
2
1.2 CLASSIFICATION OF VARIATION
1.2.1 Physiological variability
Variation at the DNA level is the basic source of phenotypic diversity. A DNA
sequence polymorphism is considered any variation present in the population at frequency
greater than 1% [2]. Although the majority of DNA changes have no impact on the final
protein structure and function, there are some that result in changes in the quality or
quantity of expressed protein. Until recently interest was focused only on the
polymorphisms located in the coding regions. Nowadays, however, there is a growing
evidence that regulatory polymorphisms in the non-coding DNA sequences located outside
exon regions modulate gene expression and even more, may be involved in processes of
evolutionary changes [3]. Regulatory polymorphisms may be “cis-acting” or “trans-acting”
depending on the copy of gene where they are located relate to affected gene. The cisacting polymorphisms are acting on the copy of the gene present on that allele and
typically, they can be found in or near the locus of the gene that they regulate. For example,
sequence in gene promoter is changed and this alters the binding affinity of transcription
factor, which in turn regulates expression of the gene. In contrast, the trans-acting
polymorphism in one gene regulates the expression of another gene that can be located
even on a different chromosome. As an example a gene encoding some transcription factor
that regulates the expression of other gene can serve.
Single nucleotide polymorphism (SNP)
The most common polymorphisms are single nucleotide polymorphisms (SNPs);
they have one of four possible nucleotides substituted for another. SNPs occur
approximately every 800 nucleotides across the genome [2]. Simple substitution of one or
more nucleotides in a sequence may change the amino acid in resulting protein depending
on the position of nucleotide in a triplet; substitution of the third nucleotide is usually silent
and does not express due to a degeneracy of the genetic code.
Overwhelming variability at the DNA level represented mainly by single nucleotide
polymorphisms is of a great importance even in an evolutionary context. Increased levels of
genetic variation between alleles or haplotypes in a species can be found in a significant
3
rate. There is a form of natural selection – called balancing selection – that works to
maintain such a genetic polymorphism (or multiple alleles) within a population. A balanced
polymorphism is a situation in which balancing selection within a population is able to
maintain stable frequencies of two or more phenotypic forms. The reason for maintaining
genetic polymorphism in such a considerable rate is that it brings a selective advantage at
both, population and individual levels. As an example, for the huge diversity of human
leukocyte antigens (HLA), which belong to the most polymorphic system in human
genome, is believed to be responsible the positive selection of new alleles that confer the
advantage of heterozygosity thus increasing the ability to optimally bind a broader range of
epitopes and keep up with the evolution of pathogens.
Two of several major mechanisms by which natural selection preserves this
variation and consequently may produce a balanced polymorphism, are best understood:
(A) heterozygote advantage (overdominance) and (B) frequency-dependent selection.
(A)
In heterozygote advantage, an individual who is heterozygous at a particular gene
locus has a greater fitness than a homozygous individual. A well-studied case of
heterozygote advantage is that of sickle cell anemia (SCA). Individuals who are
homozygous for the recessive allele at this locus are inflicted with sickle-cell disease, a
dissease in which red blood cells are grossly misshapen and which often results in a
reduced lifespan. Individuals heterozygous at this locus will not suffer from sickle-cell
disease but because of slightly irregularly shaped blood cells they are resistant to malaria.
This resistance is favored by natural selection in tropical regions where malaria is present
and so the heterozygote has an evolutionary edge. It is in this way that natural selection
preserves stable frequencies of both the heterozygote and the homozygote dominant
phenotypes. Cystic fibrosis (CF), hereditary disease of the lungs, sweat glands and
digestive system, may serve as another example. It was shown that the presence of a single
CFTR mutation (heterozygous state) may influence survivorship of people affected by
diseases involving loss of body fluid, typically due to diarrhea. The most common of these
maladies is cholera, which throughout history has killed many Europeans. A mouse model
of CF was used to study resistance to cholera, and the results were published in Science in
1994 [4]. Nevertheless, recent work evaluating candidate agents of selective pressure for
CF [5] suggests that from the three historically most probable death-causing epidemics,
4
cholera, typhoid and tuberculosis (TB), it was the latter one that enabled to explain the CF
gene persistence. And last but certainly not least example of advantageous heterozygosity
represents a certain mutation within triose phosphate isomerase (TPI), central enzyme of
glycolysis. In humans, this mutation affects the dimerisation of the protein and is causal for
a rare disease, triose phosphate isomerase deficiency. However, cells having reduced TPI
activity are more resistant against oxidative stress.
Frequency-dependent selection is a form of selection in which the relative fitness
of a specific phenotype declines if the frequency of that phenotype becomes too high. An
example of this type of selection is between parasites and their hosts: A certain parasite can
recognize one of two receptors in its host, receptor a or receptor β. If many parasites with
receptor a exist then hosts with receptor β will be selected for, and this will subsequently
increase the selective pressure on parasites using receptor β. This relationship will continue
rocking back and forth. Frequency-dependent selection has been observed in the banding
and color polymorphism in the European land snails, Cepaea nemoralis where thrushes
preferentially predate the most common morph, and also appears in the form of mate
preference, a type of sexual selection.
Repetitive elements
Other factors that contribute to a considerable proportion of variability of a human
genome are repetitive elements that make up over 45% of the human genome [6]. The term
“repetitive elements” stands for various DNA sequences that are present in multiple copies
in the genomes. They can be subdivided into tandemly repeated like microsatellites,
minisatellites and telomeres; or interspersed, for example mobile elements (e.g.
retroelements) and processed pseudogenes.
·
Tandem repetitions
o
Microsatellites
Microsatellites are short tandem repeated motifs of one to six nucleotides (Simple
Sequence Repeats, SSRs, or Short Tandem Repeats, STRs). SSRs comprise about 3% of
the human genome, with the greatest single contribution coming from dinucleotide repeats
(0.5%). The most frequent dinucleotide repeats are AC and AT (50% and 35% of
5
dinucleotide repeats, respectively) [7]. The repeating number of sequence unit can be
highly variable (10 to more than several hundreds), which is a characteristic that makes
microsatellites useful as genetic markers used for identification of a specific chromosome
or locus. The microsatellites are thought to arise by slippage during DNA replication [8, 9].
The majority of them occur in introns or other non-coding regions of the genome. Since
usually many alleles are present at a microsatellite locus, the microsatellite repeats may be
used in pedigrees for paternity testing, in forensic genetics to trace inheritance patterns
[10], for population genetic studies, linkage analysis and recombination mapping [11].
o Minisatellites (VNTR)
Longer tandem repeats are termed minisatellites or variable number tandem repeats,
VNTRs; their variant repeat units range in length from 10 to over 100 nucleotides and span
500 – 20.000 bp [6]. Similarly to microsatellites, the extreme polymorphism of
minisatellites has been extensively used for DNA fingerprinting [12] as well as for genetic
markers in linkage analysis and population studies. Moreover, minisatellites have been
intensively investigated in studies focused on mechanisms that evoke their instability.
·
Interspersed repetitive elements
Recent analyses show that more than half of the human genome may consist of
interspersed repetitive element [7, 13]. They can be subdivided based on their size, with
short interspersed nuclear elements (SINEs) being less than 500 bp long [6] and long
interspersed nuclear elements (LINEs) reaching more than 6-8 kb in size [14].
o Short Interspersed Nuclear Elements
SINEs represent reverse-transcribed RNA molecules originally transcribed by RNA
polymerase III into rRNA, tRNA and other small nuclear RNAs. The most abundant SINEs
in human chromosomes are mobile Alu elements. They represent around 10% of all SINEs
[7, 15]. Alu elements are about 300 bp long and are commonly found in introns, 3′
untranslated regions of genes and intergenic regions and do not contain any coding
sequence. They can be recognized by the restriction enzyme AluI (thus the name).
Previously, Alu repeats were believed to be "junk DNA", however, recent research suggests
they have a significant role in gene evolution, structure and transcription levels [16]. At the
6
molecular level, Alu elements are capable of insertion mutations, recombination between
elements, gene conversion and gene expression by altering the distribution of methylation
of genes throughout the genome. Based on the above, it is not surprising that Alu repeats
have been proposed to predispose the genome to instability [15] and their distribution seems
to be implicated in some genetic diseases including breast cancer, acute myelogenous
leukemia and Ewing sarcoma, but also neurofibromatosis, hemophilia, familial
hypercholesterolemia and T2D [6].
o Long Interspersed Nuclear Elements
Similarly to SINEs, LINEs constitute a considerable fraction of the human genome
(about 17%) [7, 14]. LINEs represent reverse-transcribed RNA molecules originally
transcribed by RNA polymerase II into mRNA. They code for 2 proteins; one that has the
ability to bind single stranded RNA, and another one with known reverse transcriptase and
endonuclease activity enabling them to copy both themselves and non-coding SINES such
as Alu elements [17]. It was shown that LINEs can generate processed pseudogenes [18].
It is now recognized that chromosomal sequence context plays a role in monoallelic gene
expression and may involve the recognition of long repeats or other features, since LINEs
were found significantly less truncated and more abundant around random monoallelically
expressed genes [19]. There are two major LINE families in mammalian genome: an old
LINE family, called LINE2, which apparently stopped transposing before the mammalian
radiation, and a younger family, called L1 or LINE1. Many members of L1 were inserted
after the mammalian radiation (and are still being inserted). A human genome contains
about 60.000 to 100.1000 L1 elements.
·
Segmental duplications ~ Low-Copy Repeats
In contrast to SINE and LINE repeats for which there are thousands of copies in the
genome, segmental duplications or low-copy repeats (LCR) occur, as the name says, only
in few copies. They are blocks of the genome that have been duplicated, once or several
times. The LCRs can contain genes; such genes are as a result of the duplications members
of a gene family with high homology between the individual members. The first LCRs
identified were very large (in contrast to microsatellites and LINE repeats; LCRs may reach
up 1 - 200 kb in size [7]) but as the technology is advancing, smaller LCRs are also being
7
identified. Because of their size and a high degree of sequence homology (95-97%) [20],
LCRs are prone to non-allelic homologous recombination (NAHR) leading to diseasecausing DNA rearrangements such as deletions (e.g. male infertility), duplications (e.g.
Charcot-Marie Tooth disease - a peripheral neuropathy), inversions (e.g. Haemophilia) and
gene conversions (generating pathogenic mutations by introducing non-functional
mutations from a pseudogene into an expressed gene (e.g. Congenital adrenal hyperplasia))
[20-22].
Figure 1: Deletions, duplications, inversions and gene
conversions.
Nevertheless, the gene conversions and rearrangements generated by NAHR do not
need to be pathogenic; they contribute to sequence variation and structural polymorphism
in the human genome.
8
RNA variability
Next to DNA, mRNA variation is being intensively studied. Since gene expression
is the primary mechanism how the information encoded in genome projects into phenotype,
it is of a fundamental importance to understand how gene-expression variability influences
pathogenesis of numerous diseases. Variability in the gene expression has already been
used as a quantitative trait for linkage disequilibrium analysis in human lymhoblastoid cell
line (Morley et al., 2004). The approach that is called “genetical genomics” experiences
considerable boom in these days. In genetical genomics, expression levels are treated as
quantitative phenotypes and genetic variants that influence gene expression are sought [23,
24].
1.2.2 Mutations
Genetic variation may under some circumstances cause disease development.
Pathological changes in genetic information are called mutations. Mechanisms similar to
those responsible for physiological DNA variability are applied in induction of mutations;
however, the effect is in the majority of cases pathological.
1.2.2.1
Classification of mutations
There are many ways of classification. The basic ordination compromises three
main categories [25]: (A) Genome mutations that affect the number of chromosomes in the
cell; (B) chromosome mutations altering the structure of individual chromosomes; and
finally (C) gene mutations responsible for alterations of individual genes.
A. Alteration in number of intact chromosomes (aneuploidy) is a consequence of
nondisjunction, the failure of chromosome pairs to separate properly during cell
division.
B. Large-scale mutations in chromosomal structure can occur spontaneously or may
result from abnormal segregation of translocated chromosomes during meiosis.
They include:
o Amplification (duplication)
o Deletions of large chromosomal regions
o Chromosomal translocations
9
o Inversion and interstitial deletion
o Loss of heterozygosity (loss of one allele in an individual who previously
had two different alleles, either by a deletion or recombination event; LOH)
C. Small-scale mutations affecting a gene in one or a few nucleotides. These mutations
involve:
o Single nucleotide substitution (point mutation, exchange of one nucleotide
for another)
§
Silent mutation (no effect on the coding amino acid)
§
Missense mutation (alteration of the sense of the gene coding strand
by specifying a different amino acid)
§
Nonsense mutation (chain termination mutation; it generates one of
the three “stop” codons into the sequence)
o Insertion (addition of one or more extra nucleotides into the DNA): shift of
reading frame, splice site mutation
o Deletion (removal one or more nucleotides from the DNA): alteration of
reading frame, splice site mutation
Based on the cause two classes of mutations are known, spontaneous and induced,
which can be further distinguished by the nature of inductor to physical, chemical or
biological. Other classifications point out the effect on function (loss-of-function mutations,
gain-of-function mutations, dominant negative mutations, lethal mutations); highlight the
aspect of affected phenotype (morphological mutations vs. biochemical mutations), or
underline the cause of inheritance:
o Wild type or homozygous non-mutated (neither allele is mutated)
o Heterozygous (only one allele is mutated)
o Homozygous (both the paternal and maternal alleles have an identical mutation)
o Compound heterozygous mutations (genetic compound; paternal and maternal
alleles have two different mutations)
1.2.2.2
Pathological changes in tumors
Compared to non-cancer complex diseases (type 1 diabetes, atherosclerosis, obesity,
coronary heart disease, hypertension, schizophrenia and many other), tumor development
10
generally undergoes a characteristic process. Three types of genes are involved in the
tumorigenesis: proto-oncogenes, tumor suppressor genes (TSG, anti-oncogenes), and
mismatch repair genes. Activation of these key players happens as a consequence of
various mutations. Chromosomal aberration is defined as a condition that leads to
chromosome rearrangement affecting both chromosome structure and chromosome
number. Such chromosomal changes cause gain or loss of chromosomal material [26].
Although this specific condition may affect whatever DNA region, microsatellites pay for
their variability by increased rate of mutations compared to other DNA regions. This high
rate of mutations is explained by slipped strand mispairing (slippage) during DNA
replication or by errors during recombination during meiosis [7, 27-29]. Recently, impact
of low-copy repeats on the generation of balanced and unbalanced chromosomal
aberrations has been recognized. Homologous recombinations have been proposed to be a
major cause of loss of heterozygozity involving tumor suppressor genes and its significance
and frequency may be more central to the process of carcinogenesis than described
previously [30-32]. It has been also speculated that at least in some case, NAHR involving
LCRs or other repeats may also be responsible for (onco)gene amplification [33].
Homologous recombination also plays a fundamental role in the somatic cells of the body
during the repair of damaging DNA lesions known as Double Strand Breaks (DSB) and in
various stages in carcinogenesis by implementing recombination events leading (next to
deletions, translocations and gene conversion) possibly also to telomere maintenance [34,
35].
There are two main types of mechanisms that activate proto-oncogenes:
1) Specific gene mutations (SNP, point mutation; affects oncogenes from ras family:
H-ras, K-ras, N-ras)
2) Chromosomal aberrations (specific for a given tumor)
a. Amplification (gain); duplication or multiplication of chromosomal material
at a given locus causing increase in number of proto-oncogene copies per
genome
b. Translocation
The most frequent tumor types fated by the above described pathological changes are
leukemias and lymphomas.
11
Many malignancies grow up after repression of anti-oncogenes. The following
mechanisms can inactivate them:
1) Specific gene mutations (deletion, SNP - either inherited or de novo)
2) Chromosomal aberrations
a. Chromosome loss
b. Translocation
The most frequent tumor types having TSGs with altered (decreased) or no function, are
solid tumors. Since the TSGs manage to sustain its function even in a recessive state (in
heterozygotes), the two hit hypothesis has been postulated explaining tumor development
[36].
Normal cells possess mechanisms for DNA reparation and mechanisms enabling the
cell to destroy itself by apoptosis if DNA damage is too severe. Genes, whose protein
products are involved in DNA replication or code for enzymes responsible for repair of
base-pair mismatches introduced during DNA replication, are termed mismatch repair
genes (or “caretakers”). Mutations that repress these genes can also eventually lead to
cancer.
Nevertheless, a mutation in one oncogene or one tumor suppressor gene is usually
not enough for a tumor to occur and a combination of a number of mutations is necessary.
12
1.3 QUANTITATIVE TRAIT LOCI (QTL)
Many complex phenotypes (e.g. height, weight, intelligence) are quantitative and
express continuous variation. Such a quantitative trait is controlled by multiple genes and
therefore is called multigenic or genetically complex. Such a complex genetics of
quantitative traits requires different approaches than the study of diseases caused by
monogenic traits. Genes that contribute to complex traits bring special challenges that make
gene discovery more difficult: locus heterogeneity1, epistasis2, low penetrance3, variable
expressivity4 and pleiotropy5, limited statistical power [37, 38].
A quantitative trait locus (QTL) is a term defining a region of a genome that is
responsible for variation in the quantitative trait of interest. Although identifying all such
regions that are associated with a specific complex phenotype might seem quite simple,
especially with all the genomic and computational tools available, the task is rather
difficult. Not only the entire number of QTL complicates the situation, but also the above
mentioned additional sources of variation bedevil the search. Ere now, more or less only
monogenic traits supplied encouraging results in associating genotype with phenotype not
only in humans [39] but also in animals [40] or plants [41]. The following figure (from
Glazier et al., [42]) indicates how different is a situation in identification of genes
controlling Mendelian traits versus those underlying genetically complex traits in the last
20 years.
1
A single disorder, trait, or pattern of traits caused by mutations in genes at different chromosomal loci.
Retinitis pigmentosa may have autosomal dominant, autosomal recessive or X-linked origin
2
Term used for interaction between genes/loci; epistasis occurs when the action of one gene is modified by
one or several other genes (sometimes called modifier genes). The combined phenotypic effect of two or more
genes/loci exceeds the sum of effects at individual genes/loci.
3
Penetrance is the proportion of individuals with a given genotype that express it at the phenotype level
4
Term coined for the level of phenotypic expression for a particular genotype
5
The situation in which a single gene affects two or more characteristics
13
Figure 2: Cumulative data for human Mendelian trait genes (to 2001) include all major
genes causing a Mendelian disorder in which causal variants have been
identified. This reflects mutations in a total of 1336 genes. Complex trait genes
were identified by the whole-genome screen approach and denote cumulative
year-on-year data described in this review. From Glazier, 2002
As already emphasized, genetic background of complex disease or trait cannot be
explained by single causative gene. Instead of that, terms like candidate genes (loci),
genetic linkage, association and linkage disequilibrium are frequently used.
Genetic linkage occurs when particular genetic loci or alleles are inherited jointly. Genetic
loci on the same chromosome are physically connected and tend to segregate together
during meiosis and are thus genetically linked. Gene alleles localized on different
chromosomes are usually not linked, due to independent assortment of chromosomes
during meiosis. Genetic linkage is measured by frequency of recombination between the
two loci.
Linkage disequilibrium (LD): Is defined as non-random distribution of alleles at two (or
more) loci. In other words, LD describes situation when some combination of alleles or
genetic markers occurs more or less frequently in population than it would be expected if
the formation of haplotypes from these alleles or markers is random. LD is generally
caused by genetic linkage and rate of recombination, but also by microevolution process in
14
population (mutation rate, random drift, natural selection, non-random mating, and
population structure).
Genetic association: When we cannot prove the causality of some gene(s) but we do
observe its/their presence together with some trait (they occur in the population
simultaneously) more often than could be explained by the chance, we talk about the
association of a given gene(s) with a trait or disease in question. Association between the
two genetic traits is called genetic linkage.
1.3.1 Methods for QTL analysis
The primary and major difficulty in identifying QTLs is that any single QTL
explains only a small fraction of the phenotypic variation and thus the phenotype–genotype
correlation is low. To identify chromosomal regions that affect the complex traits, various
approaches were implemented that may be divided into two basic stages [43]: The first
stage involves the QTL identification based on phenotype and genotype and correlation of
gene markers (e.g. DNA polymorphisms as SNPs or microsatellites) with a phenotype is
carried out. According to Darvasi and Pisante-Shalom, this stage consists of three
subsequent steps, suggested accordingly also by others [42]: (a) QTL detection, (b) QTL
mapping and (c) QTL fine mapping [43]. The main aim is to narrow down the region
correlating with the trait in question so that it contains as few genes as possible.
The analysis of previously identified QTL down to a single gene is a second stage,
which can be performed e.g. by functional assays; they provide an evidence that variation at
a particular gene has an effect on phenotype.
1.3.1.1
Experimental species
The fact that genetic loci contributing to quantitative traits have only small effect on
phenotype slows down the progress in identifying single genes. First, association studies
seemed to be the most appropriate method for finding the genes that influence complex
traits [44]. However, family-based studies have to be very large to provide the resolution
that is needed for positional cloning (mapping a disease gene to 1cM (or approximately
1Mb) will require a median of 700 sib pairs if the gene causes a twofold increase in risk to
offspring [45]), and population-based association studies may be confounded because of
15
environmental or genetic differences between cases and controls [46]. These difficulties
have led to the study of animal models of human traits.
Indispensable role in QTL detection play mice and other experimental species (both
plants and animals); works made on Drosophila melanogster or even yeasts contributed
greatly to genetic research of complex traits as well, especially in the methodological area
[47, 48]. Regardless of all methods, however, resolution is limited by recombination [49].
Recombination is the natural crossing over between the two copies of each chromosome
that occurs during meiosis; it is important for mapping, because it breaks down the linkage
between QTL and molecular markers on the same chromosome. The result is that markers
closer to a QTL tend to show greater association with the trait of interest than markers that
are further away. If there have been many generations of recombination, only markers that
are extremely close to the QTL will show association with phenotype. Nevertheless, in
most human pedigrees, it is impossible to get sufficient number of recombination events,
which causes poor resolution of the trait locus position [50]. That is why experimental
animals such as mice are intensively studied:
Homozygous inbred parental lines are being crossed to create F1 population
heterozygous at all markers and QTLs. F1 population is then used for further crosses such
as backcross, F2 intercrosses and crosses to generate recombinant inbred lines that serve for
statistical analysis between DNA polymorphic traits and phenotypes. Nevertheless,
although the inbred lines mating is advantageous compared to study of QTLs directly in
humans, it has its drawback as well. When inbred lines are used for QTL mapping, only
limited number of crosses in which recombination may occur is available (in F2 cross there
is only one generation in that recombination may happen and thus disrupt the linkage
between QTL and nearby marker).
In an attempt to circumvent the difficulties encountered with inbred crosses, Mott
and colleagues used a genetically heterogeneous stock (HS) of mice for which the ancestry
was known [51]. Their heterogeneous stock was established from cross of eight inbred lines
30 years ago and thus each chromosome from an HS animal represents a fine genetic
mosaic of the founder strains.
Mapping small-effect QTLs with a high degree of precision was enabled by another
approach that used commercially available outbred mice [52]. As already mentioned, an
advantage of HS and outbred strain is that many previous generations of recombination
16
effectively break down the linkage between QTLs and their neighboring markers. However,
there is a problem with studying marker–phenotype association in outbred strains because a
given marker allele may be located next to different QTL alleles and vice versa. To evade
this potential weakness of the method, Yalcin et al. took an advantage of the knowledge
that MF1 outbred mice can be treated as mosaics of standard inbred strains and instead of
analyzing a single SNP they analyzed the association between phenotype and haplotypes
(sets of closely linked SNPs) of commonly used inbred strains in order to model the origins
of the SNP alleles in MF1 outbred population [52]. The QTL fine-mapping was carried out
using HAPPY software, as progenitors were used eight inbred strains.
A final step in QTL investigation and the stage 2 according to Darvasi and PisanteShalom [43] should be the functional study that confirms the candidate gene(s) of the
detected QTL. To approaches that can be used for functional validation of suspected
gene(s) belong among others altering gene expression (either knockout or transgenic) and
quantifying natural gene-expression levels.
Novel method to resolve the underlying genetic architecture of QTLs was
demonstrated in 2002 year in yeasts [48]. First, by strategy called selective genotyping the
authors detected regions where studied QTL (high-temperature growth, Htg) was mapped.
Nevertheless, neither detailed sequence analysis of several strains nor a comparison of
mRNA transcript of the parental strains led to any result in fine mapping. This was
predictable because in quantitative complex traits, a single variation will probably never be
sufficient to display a given phenotype. Since functional validation of the designated
genomic region was still missing the authors established a novel approach termed
reciprocal-hemizygosity. They produced pairs of heterozygous strains from the original
Htg+ and Htg− parents, each member of a pair carrying reciprocal short deletions within
the candidate interval (Figure 3).
17
Figure 3: Reciprocal hemizygosity is the novel strategy introduced by Steinmetz et al. 2002.
To examine the high temperature growth phenotype, hybrids are formed between
an Htg+ strain (red) and an Htg− strain (blue). Each block represents a gene in
the region identified by quantitative trait analysis as containing genes that affect
the Htg phenotype. A pair of hemizygote hybrids is made for each gene in the
region: one with a gene deleted from the chromosome of Htg+ origin (Htg+Δ)
and one with a gene deleted from the chromosome of Htg− origin (Htg−Δ). By
deleting the gene in question from the Htg+ and the Htg− parents, and replacing
it with two different dominant drug markers (white boxes), the two hemizygotes
can be distinguished. From Darvasi and Pisante-Shalom, 2002 [43]
With reciprocal-hemizygosity analysis, it is tested whether an allele from one
genetic background was advantageous over that from the other by measuring competitive
growth between reciprocal strains. This approach thus allows observing the effect of
individual alleles in an otherwise uniform genetic background.
Another method was applied by Yalcin and co-workers [52]. They identified the
candidate genes by technique termed as quantitative complementation testing, originally
used with Drosophila melanogaster [47]. The test was re-named by Darvasi to “QTLknockout interaction test” [53] since it better describes its principle. The test requires a
strain carrying a mutation at a known candidate gene, a strain with wild type sequence in
that gene, and strains with different QTL alleles. Four combinations are generated by
18
crossing the lines and the individuals are measured for the trait of interest. If the gene
underlying the QTL is not the same as the knocked-out gene and the two genes do not
interact, then there will be an effect of the QTL and potentially of the mutation, but these
two effects will be independent and the interaction between the two factors will not be
significant. In contrast, a significant interaction indicates that the gene underlying the QTL
is the same as the known candidate gene (allelism), or that the two genes interact
(epistasis). However, this test does not distinguish between these two possibilities. So, the
“QTL-knockout interaction approach” tests the presence or absence of an interaction effect
between the null allele and the QTL.
1.3.1.2
Mapping QTLs in humans
Although in human linkage disequilibrium has been successfully used for mapping
simple Mendelian traits, it has been suggested that under certain settings even quantitative
trait loci may be mapped using LD [54]. The authors proposed the use of threshold-defined
cases for mapping loci influencing quantitative traits using LD at the population level. Two
approaches were suggested by Tenesa and.colleagues: Marker-based versus trait-based
method [50]. These two methods differ in the way how they deal with the phenotype factor.
Trait-based (dichotomizing) approach uses the phenotypic information only for the
selection step and thus creates two populations to be compared: one that has a trait and the
other that does not. Since the trait is continuous, only subjects who fall below some defined
cut-off are treated as positive (e.g. suffering from the disease; cases) and the rest as
negative (controls). The advantage of such a dichotomizing approach is that it resembles
the traditional disease mapping methods and similar ways of interpretation can be applied.
However, the information within the two classes is lost and the price paid – in terms of loss
of statistical power – might be very high. The trait-based approach for mapping quantitative
trait loci using LD is being frequently applied in association studies.
In the non-dichotomizing (in Tenesa et al. termed marker-based) approach is the
mean phenotypic value (quantitative trait) for each marker allele or genotype compared in
the same individuals. The study showed that significant amount of information is lost when
analyzing a quantitative trait as a threshold-defined binary trait, as opposed to analyzing it
using all the information available. The lost information is reflected in a loss of statistical
power to detect an association between marker genotype and phenotype.
19
1.3.1.3
Advantages and drawbacks of linkage analysis and association
study
Linkage analysis and association studies represent two major approaches that have
been used to map genetic variants influencing disease risk.
In linkage analysis, a genome-wide set of a few hundred or a few thousand markers
spread millions of bases apart is typed in families with multiple affected relatives. Markers
that segregate with the disease/trait in relatives more often than expected are than used for
localization of the disease genes. This approach is considered to be unbiased and
comprehensive search across the genome for susceptibility alleles. Although it has been
successfully applied to find the genes for many single-gene disorders, linkage analysis has
been much less successful for polygenic diseases and quantitative traits [38]. This may be
perhaps because of a limited power to detect the effect of common alleles with modest
effect on the disease.
Association studies look for a particular marker and correlate it with a disease (or
trait values) across a population rather than within families. For example, the insulin VNTR
class I allele, which has a frequency of approximately 29%, has definitively shown to
modestly affect the risk of type 1 diabetes, using association studies [55]. Association
studies require many more markers than linkage analysis and have much bigger power than
linkage analysis; they require that the markers are in linkage disequilibrium with a disease
allele, whereas in linkage analysis it is sufficient to have markers only in a linkage with the
disease allele. The association studies involve a simple comparison of the frequencies of an
allele or haplotype between individuals with and without a trait of interest. If compelling
evidence for frequency differences exists, then either the locus (or loci) in question harbors
alleles that causally or directly influence the trait, or the alleles are in linkage
disequilibrium with alleles at a neighboring locus that directly influence the trait in
question. Nevertheless, despite the bigger power, the association studies still require huge
sample size (thousands of individuals) and thus they are currently limited to candidate
genes or regions. The search throughout the genome would be too expensive.
20
1.3.1.4
QTL in human cancer disease
Detection of QTLs and sequentially genes involved in tumorigenesis (process of
tumor formation in a body) engage a whole cascade of methods. In general, similarly to
what was described up to here, it covers an identification of affected chromosome (1)
followed by delineation and downsizing of the region (2), to finally map and identify the
candidate gene(s) (3). Methods most frequently used for searching genes participating in
pathological cell growth classified according to their order in the multi-step process are:
1)
1)
CGH and its modifications, FISH
2)
Microsatellite analysis, SNaPshot analysis, gDNA quantitative PCR
3)
Sequencing, expression and functional studies
Methods listed in step 1 are suitable for the primary superficial location of mutation,
with a resolution at a chromosome/chromosomal band level. Comparative Genome
Hybridization (CGH) is a molecular-cytogenetic technique for analysis of copy number
changes (gains/losses) in the DNA content of tumor cells. The method is based on the
hybridization of fluorescently labeled tumor (frequently with Fluorescein - FITC) and
normal DNA (frequently with Rhodamine or Texas Red) to normal (meaning diploid)
human metaphase chromosome. After in situ hybridization, regional differences in the
fluorescence ratio of tumor vs. control DNA corresponding to the loss or gain of genetic
material along the length of chromosome can be detected by microscope. CGH will detect
only unbalanced chromosomal changes; structural aberrations such as balanced reciprocal
translocations and inversions can not be detected. CGH has limited mapping resolution
(approximately 20 Mb) therefore modifications of the hybridization method have been
recently invented that help in the routine research praxis to track down smaller and smaller
DNA copy number variations associated with a wide range of diseases, from developmental
disorders to cancer. These array-based comparative genome hybridization techniques (ACGH) use probes (printed in on a glass slide) to which the DNA hybridizes. The size of the
probes determines the resolution: BAC-CGH (Bacterial Artificial Chromosome-based
comparative genome hybridization), YAC-CGH (Yeast Artificial Chromosome) and
cDNA-CGH use for physical mapping genomic DNA sequences cloned in vectors such as
cosmids, bacterial artificial chromosomes (BACs) or yeast artificial chromosomes (YACs)
21
and thus enable to detect chromosome copy number changes at a higher resolution level (~
0.5 Mb) than conventional chromosome-based CGH.
Figure 4: Comparative genome hybridization - procedure
22
Figure 5: Comparative genome hybridization - results. Green lines on the right side
of chromosomes represent amplification of genetic material at a given
region (overrepresented loci); red lines on the left are deleted
(underrepresented) regions; yellow rectangles: high level of amplification
Fluorescence in situ hybridization (FISH) employs fluorescently labeled sequences
of single-stranded DNA (probes), which are complementary to the DNA sequences the
researchers wish to visualize and examine. The technique provides similar results as CGH
but with higher resolution. Moreover, compared to CGH, it can recognize balanced
reciprocal translocations. Unlike most other techniques used for chromosomes study that
require the cells to be actively dividing, FISH can be performed on non-dividing cells,
making it a highly versatile procedure.
23
Figure 6: Fluorescence in situ hybridization – schematic
illustration
There are three different types of FISH probes, each of which has a specific
application:
Whole chromosome probes are actually collections of smaller probes, each of which
hybridizes to a different sequence along the length of the same chromosome. Using these
libraries of probes, it is possible to paint an entire chromosome and generate a spectral
karyotype. This full color image of the chromosomes allows distinguishing between the
chromosomes based on their colors, rather than based on their dark and light banding
patterns, viewed in black and white through traditional karyotyping. Whole chromosome
probes are particularly useful for examining chromosomal abnormalities, e.g. when a piece
of one chromosome is attached to the end of another chromosome.
Alphoid or centromeric repeat probes are generated from repetitive sequences found at
the centromeres of chromosomes. Because each chromosome can be painted in a different
color, this technique serves to determine whether an individual has the correct number of
chromosomes or, for example, whether a person has an extra copy of a chromosome.
Locus specific probes hybridize to a particular region of a chromosome. This type of probe
is useful when a small portion of a gene has been isolated and used to determine on which
chromosome that gene is located.
24
2)
After the affected chromosome is known, methods allowing screening of the whole
chromosome to narrow down the aberrant region may be applied. Microsatellite analysis
takes an advantage of frequent pathological errors that occur during cell division and
detects resulting change in number of microsatellite repetitions. The variation in number of
repeats affects the overall length of the microsatellite, characteristics that can be measured
by various laboratory techniques. The regions containing the microsatellite repeats are
amplified by polymerase chain reaction (PCR) with oligonucleotides (primers) that flank
the microsatellite region. The size of amplicon corresponds with the number of repeats and
can be detected on a gel electrophoresis that distinguishes the DNA fragments based on
their size. Nevertheless, the most precise results can be obtained from fragment analysis
method: One oligonucleotide of each PCR primer pair is labeled with fluorescent dye,
usually FAM or HEX, and PCR product than subjected to electrophoresis on a sequencer.
The automatically collected data are analyzed using specialized software. Only genotypes
heterozygous for a given locus are regarded as informative; homozygosity and
microsatellite instability (MSI) render the particular locus unevaluable for loss of
heterozygosity (LOH) or amplification. In heterozygous genotypes, ratio of both alleles in
normal and tumor tissues is calculated. If these ratios show a difference of more than 40%,
the genotype is determined to display allelic imbalance.
Next to the microsatellite repeats, SNPs may be used for detection of loss or gain of
chromosomal material as well. The technique is called SNaPshot and uses labeled
dideoxynucleotides (ddNTPs), common in sequencing. Each ddNTP (ddATP, ddGTP,
ddCTP, ddTTP) is labeled by different fluorescent dye. Except for 2 oligonucleotides used
for PCR amplification, one extra primer for SNapShot reaction is designed whose 3’end
terminates just one nucleotide before the SNP is located. The procedure is same as in
sequencing with labeled ddNTPs except for one difference: in SNapShot reaction only 1
nucleotide-extension occurs since no free dNTPs are present. The reaction is carried out on
a sequencer; data are collected and analyzed using specialized software. Similarly to
microsatellite analysis, only heterozygous genotype is informative; the areas of peaks
representing two alleles are measured and the ratio is compared between patient’s tumor
and normal DNA. Although the SNaPshot chemistry was originally developed for allelic
discrimination, it may have much broader utilization.
25
In case there are no microsatellite repeats or suitable SNPs within the locus of
interest available, quantitative PCR (QPCR) that measures copy number of genomic
DNA may be used. This technique compares the DNA copy number between tumor and
normal sample (it does not require matched normal and tumor DNA from one patient). It
takes an advantage of real-time PCR instruments and chemistry (both SYBRGreen and
TaqMan are suitable) that allow to monitor an increasing amount of amplified PCR product
in real-time, which is dependent on the template input. The DNA quantities are normalized
using several reference loci located on chromosomes usually unaltered in the studied cancer
type and compared with standard DNA (calibrator) from unrelated normal individuals.
Therefore, this technique can answer the question regarding copy number changes that
might have occurred in tumor DNA.
3)
As soon as the microsatellite or SNP analysis narrows down the affected region up
to several kb, all genes and expressed sequence tags (ESTs) need to be considered carefully
with regards to their potential tumorigenic effect. With a help of publicly available
databases (NCBI - http://www.ncbi.nlm.nih.gov/, Sanger - http://www.ncbi.nlm.nih.gov/,
PANTHER - http://www.pantherdb.org/, USCS - http://genome.ucsc.edu/ etc.) genes of
known function (or based on protein similarity) are examined and those ones most probably
not involved in the malignant process are excluded. The remaining genes and ESTs of a
given locus are to be analyzed in details – e.g. by sequencing, expression or functional
studies.
Sequencing is a technique that enables to determine the exact DNA sequence.
There are several ways how to sequence target DNA: next to the Maxam-Gilbert and in
these days more often used Sanger’s sequencing it is possible to sequence directly using
polymerase and modified dNTPs, by annealing to the oligonucleotides attached to the
chips, by pyrophosphate sequencing, MALDI-TOF sequencing etc.
26
1.4 DISEASES WITH UNKNOWN QTL
1.4.1 Enteropathy-type T-cell lymphoma
1.4.1.1
Classification of ETL
Enteropathy-type T-cell lymphoma (ETL or EATL), a rare intestinal lymphoma
accounting for less than 1% of non-Hodgkin lymphomas (NHL) [56] and fewer than 5% all
gastrointestinal tract lymphomas [57], was characterized by Isaacson and Wright only in
1978 [58, 59]. In 1994 was this cancer included as “intestinal T-cell lymphoma (with or
without enteropathy)” in Revised European American classification of lymphoid neoplasms
(REAL) [60]. The present term, enteropathy-type T-cell lymphoma, is recommended by
World Health Organization (WHO) [56] that defines ETL as a tumor of intraepithelial Tlymphocytes with various degrees of transformation.
·
Occurrence
ETL occurs mostly in jejunum or ileum; presentation in duodenum, stomach, colon
or outside the gastrointestinal tract has also been recorded but is rare [56]. Malignant ETL
cells arise after clonal proliferation of phenotypically abnormal intraepithelial lymphocytes
(IELs) [56, 61] that might have already persisted for years [62]. Interleukin-15, whose
physiological function is to launch lymphocyte proliferation, seems to be responsible for
neoplastic expansion of aberrantly changed monoclonal IELs. Resulting accumulation of
phenotypically aberrant IELs is probably the first step in pathogenesis of ETL [63].
·
Histology and immunology
Histologically, wide range of cytomorphological appearances was recognized,
ranging from highly pleiomorphic or anaplastic tumors to those composed of monomorphic
small- to medium-sized tumor cells. However, some discrepancies in published percentages
of particular cell types do appear: As described in Chott et al. [64, 65], the majority of
ETLs are composed of pleiomorphic medium-sized to large, sometimes anaplastic cells that
are usually CD3+, CD7+, CD4- and CD5-. Approximately 20% of the recorded cases are
monomorphic, small- to medium-sized and express CD3, CD7, CD8 and CD56.
In
contrary, according to WHO histological description the most frequently observed tumor
cells are relatively monomorphic of a medium to large size with round or angulated
27
vesicular nuclei, prominent nucleoli and moderate to abundant, pale-staining cytoplasm.
Less commonly, the tumor exhibits marked pleiomorphism with multinucleated cells that
resemble anaplastic large cell lymphoma [56]. Nevertheless, the recent data on ETL suggest
that morphological classification of this lymphoma by itself is of lesser importance [66].
Regarding immunophenotype, WHO classification points out to expression of CD3,
CD7, CD103 and in a subset of cells CD8, and lack of expression of CD4 and CD5. Almost
all ETL cases express in varying proportion of tumor cells also CD30 [56], [61]].
Moreover, in the subset comprised of small to medium-sized cells the tumor cells were
CD8+ and CD56+, as reported elsewhere [56], [64], which has been used in a proposed
ETL sub-classification [64, 67]: Approximately 80%–90% of ETLs (frequently referred to
as non-monomorphic or type 1 ETL, see below) are composed of pleiomorphic, anaplastic,
or immunoblastic tumor cells that usually have a CD3+CD4-CD8-CD7+CD5-CD56immunophenotype, a phenotype similar to that of the majority of normal intestinal
intraepithelial T-cell receptor a/β T lymphocytes. About a third of these ETLs are
associated with a previous clinical history of celiac disease. In 85% of these tumors there is
histological evidence of enteropathy-like alterations in the gastrointestinal mucosa adjacent
to the invasive tumor. In contrast, about 10% – 20% of ETL cases (subsequently referred to
as monomorphic or type 2 ETL, see below) are characterized by frequent expression of
CD8 and CD56, and are composed commonly of monomorphic small- to medium-sized
tumor cells. These likely are derived from activated CD8+CD56+ intraepithelial T
lymphocytes, which make up around 15% of intraepithelial T-cell receptor a/β T
lymphocytes. This subtype appears to be associated rarely with a previous clinical history
of celiac disease. In addition, only 50% of tumors show histological evidence of
enteropathy adjacent to the invasive neoplasm.
1.4.1.2
Environmental risk factors
T-cell neoplasms are in general relatively uncommon. Together with NK-cell
neoplasms they account for only 12 % of all non-Hodgkin lymphomas [56]. These
lymphomas show significantly different incidence variations in different geographical
regions, which suggests environmental factors may have an impact on the tumorigenesis.
T-cell lymphomas are reported more often in Asia where is also simultaneously high
28
incidence of the HTLV-1 virus (in endemic regions of Japan the seroprevalence of HTLV-1
is 8-10%). It is thus considered as proved that one of the main risk factors for T-cell
lymphoma in Japan is the virus [56]. Next to this, Epstein-Barr virus (EBV) has been
detected in a minority of cases, but the viral infection is often restricted to a small
subpopulation of tumor cells and again, is to a certain extent epidemiologically dependent
[68, 69]. As another factor deemed to influence the incidence of T- and NK-cell
lymphomas is a racial predisposition. It was shown that aggressive NK/T-cell leukemias are
more frequent in Asia inhabitants than in the rest of the world.
ETL is often described as the most serious complication of celiac disease (CD) [70].
Clinically, there is an evidence of three possible associations with celiac disease: (1) ETL is
a long-standing complication of celiac disease; (2) A short history of malabsorption (weeks
or months) clinically reminding celiac disease (this is the most frequent condition; such a
form, however, may already arise as a consequence of lymphoma rather than celiac disease
alone); (3) The clinical picture presenting no malabsorption at all or the history of
malabsorption that is not clear. Nevertheless, the histological analysis of the resected tissue
adjacent to invasive tumor shows increased intraepithelial lymphocytes, villous atrophy and
crypt hyperplasia, features similar to celiac disease.
With regard to pathogenesis, a chronic stimulation of intraepithelial lymphocytes by
nutritional gluten may be an important trigger of lymphomagenesis (reviewed by [61]).
Nevertheless, specific genetic events are probably required to drive the process of
lymphomagenesis further – as evident in long-term celiac disease patients who never
developed ETL although they displayed clonal T-cell population [71]. Recently, two types
of the therapy-refractory celiac disease (RCD) have been distinguished [72]. Except for
persistent features of celiac disease, type I RCD does not exhibit any atypical histology or
immunology of intraepithelial T lymphocytes, and shows polyclonal pattern. In contrast,
type II RCD is CD3+, does not express surface T-cell markers such as CD8 or CD4, and
commonly shows monoclonal T-cell receptor (TCR) rearrangement. Moreover,
intraepithelial lymphocytes of type II RCD are often described with clonal cytogenetic
alterations [73]. Type II RCD thus may as a “cryptic ETL” represent link between celiac
disease and enteropathy-type T-cell lymphoma [74].
29
1.4.1.3
Genetic risk factors
With regards to genetic alterations occurring in ETL and compared to other
gastrointestinal lymphomas, only few data are available. Next to the factors that constitute
physiological genetic polymorphisms and whose variances are considered prerequisite for
inconsiderable amount of polygenic disorders (such as HLA haplotypes), specific features –
genetic aberrations – representing hallmark in ETL were found.
·
Most frequent chromosomal aberration
The most frequently changed locus in ETL is the highly recurrent 9q gain (70%)
[67]. The conventional CGH [66] and microsatellite marker study [75] allowed for more
precise determination of the locus and narrowed it down to 9q33-q34 region. Till the date,
all studies designated this amplified region as a unifying characteristic feature present in a
very high (40 – 70%) number of patients [57, 66, 67, 75]. The 9q33-q34 amplification was
observed in ETL irrespective of lymphoma morphology or immunology, which only
highlights the presumptive pivotal role of this aberration in ETL pathogenesis. In the
vicinity of amplified locus flanked by D9S290 and D9S1847 microsatellite markers
(positioned 4cM apart) at least two candidate genes are located: Proto-oncogene ABL1
(whose involvement in translocation t(9;22)(q34;q11) and consequent fusion to BCR in
chronic myelogenous leukemia, CML, is notoriously known) [76, 77] and transmembrane
receptor involved in T-cell development, NOTCH1. Also NOTCH1 has an attractive
suspect-like record for tumorigenesis: Its overexpression observed in anaplastic large cell
lymphoma (ALCL) [78] or high expression in T-cell acute lymphoblastic leukemia (TALL)
caused by t(7;9)(q34;q34.3) [79] is a clear evidence of its potential oncogenic property.
·
Further frequent chromosomal changes
Among additional chromosomal imbalances frequently detected in ETL by the
conventional CGH dominate recurrent chromosomal gains of 7q, 5q and 1q and recurrent
losses of 8p, 13p, and 9p loci. All chromosomal aberrations detected in more than 15% of
ETL cases analyzed and published in 4 key studies are summarized in Table 1.
30
Table 1
Gains
Losses
Diverse
aberrations
Chromosomal
loci
9q31.3-q33.2
9q33.2-q33.3
9q34-qter
9q33-q34
1q23.1-q23.3
1q25.3-q31.2
1q32.2-q41
5p15.33
5q33-q34
5q34-q35
6p25.2-pter
7p22.3
7q11.22-q11.23
7q11.23-q21.3
7q22
7q31
7q33-q34
7q36-qter
8q13.3-q21.11
8q24
12p13.31-p13.33
16p13.3
17q25.3
22q12.3-q13.2
21q22.3
11p15.5
20q11.21-q12
22q11.21-q11.23
2q31.1-q31.3
2q24.3-q25.1
8p22-p23
8cen-p21
9p21
9p11.2-p13.3
10q26.2-q26.3
11q14.1-q14.2
13q22
18q22
17q23-25
17p13.1 (p53)
19q13.2-q13.31
3p21.31
5q21 (APC)
7p21
16q12.1
Zettl 2002
(CGH, FISH)
Baumgärtner
2003 (MA)
Zettl 2007
(CGH, FISH)
De Leeuw 2007
(a-CGH)
60%
67%
63%
58%
16%
40%
64%
26%
40%
43%
47%
37%
30.8%
18%
21%
24%
31%
32%
18%
21%
24%
18%
15%
24%
50%
23%
27%
37%
37%
43%
43%
47%
43%
43%
43%
27%
23%
23%
23%
23%
20%
20%
20%
20%
17%
36.7%
23%
20%
30%
20%
25%
18%
34.6%
26.9%
16.6%
24%
17q23-25
42.3%
5q33-q34
7q31
1q41
7p21
38.5%
40%
36.4%
32%
31
23%
20%
20%
20%
23%
Table 1: Number of recurrent chromosomal gains and losses published so far (in % of all
cases). Only aberrations with frequency ≥ 15% are shown. Frequency above 30%
depicted in bold.
As shown, several regions containing loci of known tumor suppressor genes or
oncogenes suffer from consistent abnormalities: The p53 (TP53) locus in 17p13.1, the APC
gene locus in 5q21, the INK4a/ARF (p15/p16) locus in 9p21, and the retinoblastoma (RB1)
gene locus in 13q14 revealed LOHs (although the last two loci in frequency lower than
15%) [75]. Similar to other lymphomas, the minimal overlapping region on 7q was 7q21q22 [80, 81]. It harbors genes playing a role in cell cycle control, CDK6 and ASK/DBF4, as
well as gene encoding transcriptional co-activator TRAPP and multidrug resistance gene
MDR1. Interestingly, gains of 1q22-q44 had been reported previously as a highly recurrent
genetic alteration in refractory celiac disease [73].
Certain genetic alterations presented a close correlation with tumor morphology.
Notably, monomorphic ETL more frequently showed gains of the MYC oncogene locus at
8q24 (73% vs. 27%, P < 0.03) and lacked gains of 1q32.2-q41 (20% vs. 73%, P < 0.01) and
5q34-q35.2 (20% vs. 80%, P < 0.01) [67]. In contrast, non-monomorphic ETL was
characterized genetically by frequent gains of 1q32.2-q41 (75% vs. 32%, P < 0.01).
It is noteworthy that, in contrast to gains of 9q chromosomal region, amplifications
of 7q and 1q are among the most frequent genetic alterations detected both in classical
cytogenetic studies and CGH analysis of different types of T- and B-cell lymphomas [8082]. It thus seems that genetic alterations shared among different T-cell lymphoma entities
(such as gains of 1q and 7q or losses of 13q) may abrogate the function of genetic loci that
control basic regulatory mechanisms of T lymphocytes proliferation or may be acquired
during lymphoma progression. In contrast, entity-characteristic genetic alterations (such as
9q amplification in ETL) may overcome growth regulation mechanisms in specific T-cell
subpopulation (such as in intraepithelial T lymphocytes), which gives rise to distinct T-cell
lymphoma entities (ETL).
·
HLA class II
Most patients show the HLA-DQA1*0501-DQB1*0201 (HLA-DQ2) genotype [83].
The HLA analysis in the 2 histological types of ETL revealed significant differences.
Compared with monomorphic ETL, non-monomorphic ETL patients more frequently
32
carried HLA-DQ2 (73% vs. 33%; P = 0.03) and were homozygous significantly more often
for HLA-DQ2 (40% vs. 6%; P = 0 .03). Homozygosity for HLA-DQ2 is associated with
almost 5-fold increase in risk of disease development than HLA-DQ2 heterozygosity [84,
85]. The study by de Leeuw et al. [67] reported in non-monomorphic ETL cases the overall
HLA-DQB1 genotype pattern closely resembling genotypes commonly published for ETL,
whereas genotypes detected in monomorphic ETLs did not significantly differ from those
recognized in normal Caucasian controls (P = 0.677).
Next to this, correlation of the HLA genotypes with genetic aberrations identified an
association between 5q34-q35.2 gains and DQB1*02 genotype (79% vs. 31%, P = 0.014)
[67].
·
Minimum two pathways of lymphomagenesis
Next to the gain of 9q34, which is the unifying characteristic feature present in 40%
of both, monomorphic and pleiomorphic histological types of lymphoma, Baumgärtner et al
[75] observed another, much smaller group of cases that showed 3q27 amplification. These
two pathological features were mutually exclusive. Interestingly, although de Leeuw and
colleagues suggest an existence of two ETL types as well; they discovered that 9q31.3q33.2 gains were predominantly absent in ETLs with 16q12.1 losses (56% vs. 10%, P <
0.01) [67]. Similar to 9q gains, 16q losses did not correlate to any clinical, morphologic or
immunophenotypic features, thus suggesting that both amplification of 9q34 and deletion
of 16q12.1 may result in similar biological effects in ETL lymphomagenesis.
To conclude, based on the all correlative findings 2 types of ETL could be distinguished
(Figure 7) [67]:
Type 1 ETL is characterized by non-monomorphic cytomorphology (pleiomorphic,
anaplastic, immunoblastic), CD56 negativity, common gains of chromosomes 1q and 5q,
but rarely that of the MYC oncogene locus. In view of shared HLA genotypes and genetic
alterations, type 1 ETL may arise from, and be pathogenetically linked to, refractory celiac
disease by a stepwise acquisition of genetic alterations, among which gains of 1q may be
the hallmark genetic alteration at the stage of refractory celiac disease (see chapter 1.6).
Type 2 ETL displays monomorphic small- to medium-sized tumor cell morphology
and is characterized by frequent CD56 expression, MYC oncogene locus gain, and rare
33
gains of 1q and 5q regions. Given the genetic alterations and HLA-DQB1 genotype
patterns, celiac disease may not be causal to type 2 ETL. These data suggest that a different
mechanism of lymphomagenesis is prevalent in type 2 ETL.
Both types, however, are characterized by a high prevalence of 9q gains/16q losses
as the unifying feature.
Figure 7: Suggested classification of two ETL subtypes based on the clinical, morphologic,
immunophenotypic and genetic features. From de Leeuw 2007 [67]
34
1.5 DISEASES WITH KNOWN QTL
1.5.1 Diabetes mellitus
Diabetes mellitus is one of the most common endocrine disorders affecting almost
6% of the world’s population. The prevalence of this chronic metabolic disease is on the
increase. According to the projections of Amos et al. [86], the total number of people with
diabetes mellitus all over the world will reach 221 million in the year 2010 in contrast to
the 124 million estimated for 1997. Based on other estimates, this number of diabetic
patients is expected to reach 300 million in 2025 year [87].
1.5.1.1
Classification
Diabetes mellitus (DM) is a complex of several disorders with wide range of
pathogenesis affecting the glucose, lipid and protein metabolism in the body. The main
characteristic represents hyperglycemia resulting from defects in insulin secretion, insulin
action or both. The chronic hyperglycemia of diabetes is associated with long-term damage,
dysfunction, and failure of various organs, especially the eyes, kidneys, nerves, heart, and
blood vessels [88].
Several pathogenic processes are involved in the development of diabetes. These
range from autoimmune destruction of the β-cells of the pancreas with consequent insulin
deficiency to abnormalities that result in resistance to insulin action. The basis of the
abnormalities in carbohydrate, fat, and protein metabolism in diabetes is deficient action of
insulin on target tissues. Deficient insulin action results from inadequate insulin secretion
and/or diminished tissue responses to insulin at one or more points in the complex
pathways of hormone action. Impairment of insulin secretion and defects in insulin action
frequently coexist in the same patient, and it is often unclear which abnormality, if either
alone, is the primary cause of the hyperglycemia.
Based on the etiopathogenesis, there are two main types of diabetes: Autoimmune
type 1 diabetes (T1D) (or insulin-dependent DM, IDDM) is caused by an absolute
deficiency of insulin secretion. The cause of type 2 diabetes (T2D) (or formerly noninsulin-dependent, NIDDM) is a combination of resistance to insulin action and an
inadequate compensatory insulin secretory response. Nevertheless, there has been a long
history of discussion on the proper diabetes classification and it is not absolutely clear
35
whether these two most frequent forms of diabetes should be viewed, understood and
treated as one disorder with partly different etiopathogenesis, especially when a large
proportion of diabetic patients may have both type 1 and type 2 processes contributing to
their diabetic phenotype [89], or whether the unifying hypothesis for the mechanisms of βcell death in T1D and T2D should rather be ignored [89, 90].
A specific subtype of T1D is Latent Autoimmune Diabetes in Adults (LADA).
LADA represents controversial type of diabetes sometimes designated type 1.5 (T1.5D)
[91] or “slow-progressing type 1 diabetes”. A fraction of individuals with chronic
destructive process of insulin-secreting β-cells associated with immunological changes
have clinical diabetes that initially is not insulin-requiring. The onset of diabetes in these
patients is usually in adult life, and because their diabetes is at least initially not insulinrequiring, they appear clinically to be affected by type 2 diabetes. The Immunology of
Diabetes Society proposed that patients are ≥ 30 years of age, positive for at least one of the
four autoantibodies commonly found in T1D patients (ICAs and autoantibodies to GAD65,
IA-2, and insulin), and not treated with insulin within the first six months after diagnosis
(which is the criteria supposed to distinguish between T1D occurring in patients ≥ 30 years
and LADA). Special attention should be paid to diagnose such patients because therapy
may influence the speed of progression toward insulin dependency, and in this respect,
efforts should be made to protect residual C-peptide secretion [92]. Only the studies of
immunologic, metabolic and genetic profiles of late-onset T1D and LADA [93-95] are
believed to answer the question how close regarding the pathophysiology adult-onset T1D
and LADA are.
Distinct forms of type 2 diabetes that are connected to the impaired β-cell
function due to monogenetic defect are referred to as Maturity-Onset Diabetes of Young
(MODY) [88]. They are characterized by impaired insulin secretion with minimal or no
defects in insulin action and an autosomal dominant inheritance pattern. Mutations at six
genetic loci on different chromosomes have been identified and confirmed to date: (HNF)1α (MODY3), glucokinase (MODY 2), HNF-4α (MODY 1), HNF-1β (MODY 5), insulin
promoter factor (IPF)-1 (MODY 4), and NeuroD1 (MODY 6) [96]. Recently, new three
potentially MODY-related loci are under close testing.
36
1.5.1.2
Type 1 diabetes mellitus (T1D)
1.5.1.2.1
Incidence and prevalence of T1D
Several lines of evidence support a critical role of exogenous factors in the
pathogenesis of T1D. Studies in monozygotic twins indicate that only 13-33% are pairwise
concordant for T1D [97, 98] suggesting that there is either acquired postconceptional
genetic discordance or differential exposure to putative environmental factor(s). It has also
been observed that if the first twin of a monozygotic twin pair develops diabetes prior to
age 5, the risk for the second twin to develop diabetes within 20 years of follow up is 50%,
while if the first twin develops diabetes after age 20, the risk for the co-twin is less than 5%
[99]. In contrast to monozygotic twins, the risk for developing diabetes or anti-islet
autoantibodies does not appear to differ between dizygotic twins and siblings of patients
with T1D. Early reports of increased prevalence of anti-islet autoantibodies in dizygotic
twins have been retracted [100], and a combined series indicates a relatively low prevalence
of autoantibody-positivity in dizygotic twins. For monozygotic twins, expression of any of
the “biochemical” anti-islet autoantibodies (GAD65, ICA512 (IA-2) or insulin
autoantibodies, see below) is associated with a very high risk of progression to T1D [99].
Based on decades of follow up, the development of diabetes in autoantibody-positive
monozygotic twins is preceded by progressive loss of insulin secretion [101].
Other evidence of genetic involvement in the etiology of T1D is that 95% of type 1
diabetes patients carry HLA-DR3, HLA-DR4, or both. The relative risk of type 1 diabetes
conferred by DR3 is seven fold, by DR4 nine fold and by DR3/DR4 is more than fourteen
fold of that conferred by other alleles. Recently, mathematical model has been introduced
that tries to explain the rising incidence of T1D and the excess of HLA-DR3/HLA-DR4
heterozygotes among patients [102].
The incidence of type 1 diabetes follows two-peak distribution: it begins sharply to
rise at about 9 months of age, continues to rise until age 12–14 years, and then declines
[103]. The second, less significant increase is documented around 40 years of age. A rise of
incidence in the old age (above 70 years) is rather a consequence of natural exhaustion of
the immune system. A similar pattern is seen in many countries irrespective of whether the
overall incidence of type 1 diabetes is low or high. The incidence of T1D is lowest in
Japan, the Caribbean, and southern Europe, while the highest incidence rates are in the
Scandinavian countries [104], particularly in Finland, and in Sardinia in Italy [105].
37
Accumulated data on the incidence of T1D during the last 20 years show that type 1
diabetes occurs in most racial and ethnic groups but the risk is highest among white
population [104]. While these differences among races may be determined in part by genes,
the variation within the white population is almost as large as that observed among all
races. All of these examples support the role of genetic and environmental factors in the
etiology of diabetes. A more recent study shows that also climate has a role in the etiology
of T1D [106].
1.5.1.2.2
Immunology of T1D
Despite the fact that the trigger leading to the pathogenesis of type 1 diabetes is
currently unknown, the pathophysiology of the disease is believed to be biphasic. The first
stage is represented by leukocytes infiltration to the pancreatic islets; this response does not
cause damage. Nevertheless, in the second phase, which occurs only in diabetes-prone
individuals, autoreactive T cells acquire aggressive potential and destroy the majority of the
pancreatic islets [107]. The cellular autoimmunity in T1D responsible for the destruction of
pancreatic β-cells is mediated by T cells [108]. B cells produce autoantibodies to β-cell
antigens, and in turn present these self-antigens to T cells, but β-cells and autoantibodies
are not strictly required in this autoimmune process [109].
In most cases a preclinical period characterized by the presence of autoantibodies to
pancreatic β-cell antigens precedes the onset of hyperglycemia by a few years. This long
pre-clinical period provides an opportunity for prevention. Known autoantigens include
insulin, glutamic acid decarboxylase (GAD), and antibodies against the islet cell antigen
512 phosphatase (IA-2), of which only insulin is cell specific (Table 2).
Insulin autoantibodies occur more in DR4 haplotype patients and are useful if
measured prior to administering exogenous insulin. Glutamic acid decarboxylase antibodies
persist the longest time following diagnosis and are useful in confirming autoimmune
etiology in long-standing cases. The presence of more than one type of antibody is highly
predictive of disease, and occurs years before clinical manifestations [110].
38
Major autoantigens
Minor autoantigens
Shared autoantigens
Ÿ Glutamic acid decarboxylase (GAD65)
Ÿ Protein thyrosine phosphatase
(ICA512/IA-2)
Ÿ Insulin (IAA)
Ÿ ICA 12 (SOX13)
Ÿ ICA 69
Ÿ Carboxypeptidase H
Ÿ Gangliosides (GM-1)
Ÿ Sulfatides
Ÿ Tissue transglutaminase (tTG)
Ÿ 21 – hydroxylase (21- OH)
Ÿ Gliadin
Ÿ Endomysium
Ÿ Other antigens
Table 2: Autoantigens in type 1 diabetes mellitus
T1D occurs often in association with other autoimmune diseases (thyroid disease,
Addison’s disease, celiac disease, rheumatic disease and others). The celiac disease (CD) is
the second most prevalent autoimmune condition accompanying type 1 diabetes, after
autoimmune thyroid disease. Co-existence of T1D and CD has been reported in 1-7% of
diabetic cases [111, 112], ten times more frequent than in general population [113], and the
anti-gliadin, anti-endomysium and anti-tissue transglutaminase antibodies are detected in
sera of DM patients [114].
1.5.1.2.3
Environmental risk factors
A series of evidence support a critical role of exogenous factors in the development
of type 1 diabetes, such as (1) the fact that < 10% of individuals with HLA-conferred
diabetes susceptibility do progress to clinical disease, (2) a pairwise concordance of type 1
diabetes of < 40% among monozygotic twins, (3) a more than 10-fold difference in the
disease incidence among Caucasians living in Europe, (4) a several fold increase in the
incidence over the last 50 years, and (5) migration studies indicating that the disease
incidence has increased in population groups who have moved from a low-incidence to a
high-incidence region.
Environmental factors have a role in ethnic differences in the prevalence of diabetes
mellitus. However, there are clear differences in the clinical phenotype of diabetes between
39
different ethnic groups, which are pinpointing to a genetic role. The annual rate of
incidence of childhood T1D in France, between 1988 and 1995, was comparable to that
observed in the Nordic countries. In spite of the similarity in the annual rate of incidence of
childhood type 1 diabetes, the overall incidence of T1D in childhood is still smaller in
France, suggesting that environmental factors play a major role in the pathogenesis of type
1 diabetes.
o Ethnic factors
Major ethnic and geographical differences in prevalence and incidence of T1D are
known [115]. The most frequently reported cases of T1D are in the Nordic countries, with
the highest incidence in Finland (35/100,000/year) of the population in children aged 0–14
years. The lowest incidence is seen in Asia (0.5–1.3/100,000/year). Low rates are also
reported in Africa and Latin America [116]. These significant differences in the incidences
of type 1 diabetes are possibly due to variations in genetic and environmental factors [117].
Another example of how ethnicity may influence the disease development was published in
1996 [118]: The Chile study showed a significant difference between the incidence of T1D
in native Chileans (0.42/100,000) compared to Caucasian Chileans (1.58/100,000) between
years 1983 and 1993.
o Nutrition
Nutrition during the early months of life (neonatal period and early infancy) is
considered of a high importance with regards to potential DM development. Early dietary
exposure to cow’s milk containing feed, short duration of exclusive breast feeding, and
high intake of cow’s milk protein in recent diet [119] may increase susceptibility to type 1
diabetes mellitus [120]. Prolonged breast feeding (minimum recommended is six months
period, more beneficial is one year) has shown to be uncoverable for proper development of
fully functioning child’s immune system. Immediately after birth, environmental stimuli
such as intestinal flora and microbial infections provide antigens for neonatal T cells and
TH1/TH2 balance, which triggers the education of the infant’s immune system. Maternal
milk provides am important source of immunoglobulin A (SIgA) antibodies. Before
infant’s own immunity is fully established, this SIgA from mother were shown to help to
protect against Vibrio cholerae, Campylobacter, Shigella and Giardia [121]. Also protection
40
of breast milk against enterovirus infections has been reported, and cases of enterovirus
infection in early life followed by T1D development documented [122].
Furthermore, milk is rich in receptor analogues for certain epithelial structures,
which microbes need for attachment to host tissues as an initial step in infections. Next to
that, human milk was shown to have the infant anti-inflammatory property. An early
cessation of breast feeding thus may have negative impact on the newborn health and
proper development of immunity.
o Viruses and infection
High incidence of diabetes was found to be associated with congenital rubella [123].
Based on the WHO analysis [124], 20% of children with congenital rubella develop T1D.
Further, antibodies against Coxsackie B4 have been reported in 20–30% of new cases of
T1D [125]. Rotavirus and herpes viruses have been also associated with β-cell
autoimmunity and T1D [126]. It is hypothesized that particular genotypes may predispose
their carrier to development of a stronger humoral response to infection. It is plausible that
early infection with a non-diabetogenic strain of a virus can induce immunity to
antigenetically similar diabetogenic strains and protects from T1D.
1.5.1.2.4
Genetic risk factors
To better understand the pathophysiology of T1D, it is necessary to strike to a
molecular basis of impaired regulation of glucose metabolism – to DNA. Nevertheless,
several obstacles slow down the process of deciphering the exact role of genetic factors in
the etiology of diabetes mellitus. These include the modification of the expression of the
diabetic genotype by environmental factors and the variability in the age of onset of
diabetes [105]. The T-cell mediated autoimmune process that destroys pancreatic β-cells in
T1D is a complex phenotype influenced by multiple genetic and environmental factors.
Based on the positional approach (linkage studies), 18 loci throughout the whole
genome were identified and designated IDDM1 – IDDM18 [127]. Unfortunately, it was
shown that many of the loci are false positive; these statistical artifacts originated from
underestimates of the sample size required for meaningful statistical power [128]. Table 3
summarizes up to date recognized IDDM loci.
41
Locus
Chromosome
Gene/marker
IDDM 1
IDDM 2
IDDM 3
IDDM 4
IDDM 5
IDDM 6
6p21
11p15.5
15q26
11q13
6q25
18q12-q21
IDDM 7
2q31
HLA
INS
D15S107
ZFM1 (Zinc finger protein 162)
SOD2 (superoxid dismutase)
DCC associated gene
ZNF236
bcl-2
NEUROD1
IGPR (islet specific glucose 6 – phosphatase)
IL-1 gene cluster, HOXD8, GAD1, GALNT3
IDDM 8
IDDM 9
IDDM 10
IDDM 11
IDDM 12
6q27
3q21-q25
10p11.2-q11.2
14q24.3-q31
2q33
IDDM 13
IDDM 14
IDDM 15
IDDM 16
IDDM 17
IDDM 18
2q34
GAD2
ENSA (alpha-endosulphine)
CTLA-4
CD28
ICOS
NRAMP1
6q21
14
10q25
5q31.1-33.1
Transient neonatal diabetes
Ig heavy chain
D10S1750-D10S1773
ILB12
Table 3: List of IDDM loci
The presented work addresses in greater detail only two IDDM loci, HLA (IDDM1)
and INS-VNTR (IDDM2), which are discussed in depth below. Nevertheless, the large
body of evidence suggests that also other important disease-associated loci contribute to the
T1D risk, although with much smaller effects. The IDDM12 locus (localized on
chromosome 2q33) contains CTLA-4 gene encoding cytotoxic T-lymphocyte associated
protein 4 and CD28. These are likely candidates for development of T1D and other
autoimmune diseases because of their important role in the T-cell proliferative response
[129]. Moreover, the effect of IDDM12 seems to be independent of HLA alleles or the
IDDM2 risk genotype [130].
Next to aforementioned IDDM loci, several other non-IDDM regions of genome
have been proven so far to be associated with diabetes mellitus as well. Among the
susceptibility loci of non-IDDM nature may be mentioned genes such as MIC-A (Major
histocompatibility complex (MHC) class I chain – related gene A) [131, 132] believed to be
42
recognized by one subpopulation of intestinal γδ T cells [133] and suggested to play a role
in the activation of a subpopulation of NK cells that express the NKG2D receptor [134];
PTPN22 (Protein thyrosine phosphatase non-receptor type 22) [135], gene encoding a
member of tyrosine phosphatases (PTPs) protein family, important down-regulator of T cell
activity [136, 137]; NFKB and NFKBIA genes coding for transcription factor nuclear
factor κB (NFκB) and its inhibitor IκB, respectively, whose role in inflammation process
and autoimmune disease etiology has been proposed [138]; gene for proinflammatory
cytokine interleukin-18 (IL-18) playing a key role in autoimmune, inflammatory, and
infectious diseases [139-141]. Table 4 provides list of some of the identified non-IDDM
loci.
Chromosome
location
Gene
symbol
Gene name
Type of
diabetes
11p15.1
Sulphonylurea receptor-1
T2D
2q37.3
19q13.3
11q22.2.-q22.3
SUR1
(ABCC8)
CAPN10
GYS1
IL-18
Calpain 10
Glycogen synthase
Interleukin-18
T2D
T2D
T1D
2q36
11p15.1
IRS-1
KCNJ11
6p21
MIC-A
T2D
T2D, neonatal
diabetes
T1D
4q24
14q13
3p25
NFKB
NFKBIA
PPARγ
1p13
PTPN22
10q25.3
TCF7L2
Insulin receptor substrate-1
Potassium inwardly-rectifying
channel, subfamily J, member 11
Major histocompatibility
complex (MHC) class I chain –
related gene A
Nuclear factor κB
Inhibitor of nuclear factor κB
Peroxisome proliferator –
activated receptor γ
Protein thyrosine phosphatase
non-receptor type 22
Transcription factor 7-like 2
(T-cell specific, HMG-box)
T1D, T2D
T1D, T2D
T2D
T1D
T2D
Table 4: Some of the non-IDDM loci associated with both types of diabetes
o IDDM1 locus: HLA
The first candidate locus studied and found to be strongly associated with T1D was
the human leukocyte antigen (HLA) region on chromosome 6p21.3 [142]. It accounts for
40% - 50% of the genetic susceptibility [143], through a large variety of protective and
43
predisposing haplotypes. Other important loci associated with T1D have much smaller
effects than HLA.
Human leukocyte antigens belong to the most polymorphic genes in the human
genome [144]. Such diversity brings a positive evolutionary advantage over the pathogens.
The human MHC is divided into three regions: class I, class II, and class III. The
class I genes are located telomeric in the complex. They have single polypeptide chain
associated with β2-microglobuline. The class I region includes HLA-A, -B and -C genes.
The antigens are expressed on the surface of almost all tissue cells and their function is to
present peptides to CD8+ cytotoxic T cells. HLA class II genes are located at the
centromeric end of MHC [145]. Class II molecules composed of heavy α-chain and light βchain are expressed on antigen presenting cells, APC (monocytes, macrophages, B
lymphocytes, dendritic cells and activated T lymphocytes). The HLA class II region
includes HLA-DR, -DQ and -DP subregions. Major function of antigens is to present
peptides to CD4+ T cells. The MHC class III region contains many genes with varying
functions in immune system.
Figure 8: Overview of the extended HLA complex on the chromosome 6 (modified from
Louka and Solid, 2003).
44
Most T1D-relevant polymorphisms are amino acid changes in exon 2 of the βchain of DR and both α and β-chains of DQ. Thus, the most common T1D-predisposing
haplotypes in Caucasians are DRB1*0301-DQA1*0501-DQB1*0201 and DRB1*0401DQA1*0301-DQB1*0302, abbreviated as DR3-DQ2 and DR4-DQ8, respectively [128,
146]. Interestingly, heterozygosity for DQ2/DQ8 which (because of linkage disequilibrium
almost always implies DR3/DR4 heterozygosity) confers the highest T1D risk in
Caucasians [147]. This genotype is present only in 3% of the general population, but in
30% of T1D patients, conferring a 15-fold relative risk and an earlier onset of disease.
Majority of the Caucasians suffering from T1D has at least one of these two haplotypes.
Conversely, a subset of HLA alleles provides dominant protection against
diabetes, the most common is the HLA-DQ6 haplotype, DRB1*1501-DQA1*0102DQB1*0602. It is found in less than 1% of diabetic children and in almost 20% of the
general population [145]. Another dramatically protective allele is DRB1*1401, associated
with DQA1*0101-DQB1*0503. In combination with a predisposing haplotype, the carrier
of the protective haplotype remains at low risk [148].
At a molecular level, risk alleles differ structurally from protective alleles. Most
characteristic is the absence of an aspartic acid molecule at position 57 of the β-chain of the
DQ molecule. This reverses the electric charge of the peptide-binding groove of the HLADQ8 molecule, thereby possibly altering the binding of peptide epitopes [149]. Published
crystal structure of DQ8 with an insulin peptide B (residues 9–23 that represent an
important T cell epitope) draws to the attention the hypothesis that these class II alleles
primarily influence diabetes risk by the peptides they bind and present to T lymphocytes
[150].
45
DRB1
*0301
*0401
*0405
*0801
*1601
*0101
*0403
*0701
*1501
*1401
*0701
DQA1
DR3
DR4
DR4
DR8
DR2
DR1
DR4
DR7
DR2
DR6
DR7
*0501
*0301
*0301
*0401
*0102
*0101
*0301
*0201
*0102
*0101
*0201
DQB1
*0201
*0302
*0302
*0402
*0502
*0501
*0302
*0201
*0602
*0503
*0303
Diabetes association
DQ2
DQ8
DQ8
DQ4
DQ5
DQ5
DQ8
DQ2
DQ6
DQ5
DQ9
Strongly positive
Strongly positive
Strongly positive
Positive
Positive
Positive
Neutral or weakly negative
Negative
Strongly negative
Strongly negative
Strongly negative
Table 5: Class II HLA haplotypes associated with susceptibility for or protection against type
1 diabetes
o IDDM2 locus: INS-VNTR
The second strongest association with T1D has been that of the INS-VNTR
(variable number of tandem repeats) polymorphism at the promoter region of the insulin
gene (INS), designated IDDM2 [55]. IDDM2 was mapped to the short arm of chromosome
11 between the genes for tyrosine hydroxylase (TH) and insulin-like growth factor 2 (IGF2)
on 11p15.5 [55, 151-153] (Figure 9). Alleles of the INS-VNTR are classified based on their
size (which corresponds to number of the 14-15 nucleotides long repeats) into three groups:
class I (represented by short alleles with 28 – 63 repetitions), class III (long alleles, 141 –
209 repeats) and middle class II alleles being exclusively of African/Asian origin with an
average length 80 repeat units [55, 154]. Next to this, also sequence polymorphism within
the repeating motive has been identified [154-157] that enables finer sub-classification of
each of the class. The role of INS-VNTR in the development of type 1 diabetes mellitus is
generally accepted [55, 158-160] and the risk associated with the INS-VNTR class I alleles
as well as the dominant protective effect of class III alleles for the T1D phenotype are for
the most part unquestioned.
46
Figure 9: Insulin VNTR minisatellite repeats. The INS -VNTR is located 363 bp upstream
from the insulin gene (INS) transcription starting site. It is also 5 kb upstream
from a second potential target for regulation, the insulin-like growth factor 2
(IGF2), and 10 kb downstream of the tyrosine hydroxylase gene (TH).
The accepted scheme of insulin expression in the thymus and pancreas explains the
process of development of antibodies against insulin by the finding that the class I/I
genotype is responsible for lower insulin mRNA expression in the fetal thymus but
increased expression in both the fetal and adult pancreas and, the class III/III genotype is
associated with higher expression levels in the thymus, which could more efficiently induce
negative selection of insulin-specific T-lymphocytes [161, 162].
It might also be noted that the –23HphI mutation affects the consensus
sequence for the intron splicing machinery. A similar mutation in the CFTR (cystic fibrosis
transmembrane conductance regulator) gene is able to cause serious disease [163],
therefore, the potential affect of the mutation should not be underestimated. Moreover, just
recently Kralovicova and colleagues have shown that variants of the INS –23HphI SNP
47
(termed by the authors as IVS1-6A/T) can affect pre-mRNA splicing [164]. This suggests a
functional role for –23HphI SNP.
IVS+5insTTGC
IVS2+11C/T
IVS1-6A/T
E3+168C/A
E3+155C/T
E2+53G/A
Figure 10: Insulin gene variants that affect pre-mRNA splicing. A: Schematic
representation of the INS reporter construct, tested DNA variants, and
alternatively spliced mRNA isoforms. Exons 1–3 are shown as black, gray, and
white boxes, respectively. Thick lines denote introns, colorful lines represent
mRNA isoforms identified by sequencing (numbered in circles). The designation
of DNA variants involves the distance (in nucleotides) from the splice sites.
From Kralovicova, 2006
In light of this finding, an intriguing question arises of which one of the two
polymorphisms, VNTR or –23HphI (or perhaps neither?) is the primary functional
polymorphism at the IDDM2 locus that confers risk for diabetes mellitus. The latest
experiments leave little doubt that the intragenic variability in the insulin gene itself is
responsible for various defects in insulin “processing and handling” and may play a role in
insulin resistance, diabetes mellitus and/or obesity.
o Non-IDDM locus: MIC-A
Major histocompatibility complex (MHC) class I chain –related gene A (MIC-A) is
located within the MHC class I region of the chromosome 6. It contains the long open
reading frame encoding three distinct extracellular domains (α1, α2 and α3) a
transmembrane segment, and a cytoplasmatic tail. It has been reported that MIC-A may be
recognized by a subpopulation of intestinal γδ T cells [133] and may play a role in the
activation of a subpopulation of NK cells that express the NKG2D receptor [134].
48
Sequence analysis of the MIC-A gene showed a trinucleotide repeat (GCT)
microsatellite polymorphism within the transmembrane region [165, 166]. So far, six
alleles of the exon 5 of the MIC-A gene, which consist of 4, 5, 6, 9 and 10 repetitions of
GCT, or five repetitions of GCT with an additional nucleotide insertion (GGCT), have been
identified [166, 167]. The polymorphism of the MIC-A gene and its location in the HLA
region warrant studies aimed at identifying an association with the risk for autoimmune
diseases. The MIC-A gene has already been found to confer genetic risk for Behçet’s
disease [168], ankylosing spondylitis [169], and autoimmune Addison’s disease [170].
Recent works also support the findings that MIC-A is associated with autoimmune diabetes
mellitus [132, 171].
o Non-IDDM locus: PTPN22
Lymphoid-specific phosphatase (LYP) encoded by PTPN22 gene belongs to a
family of tyrosine phosphatases (PTPs) involved in preventing spontaneous T cell
activation by dephosphorylating and inactivating T cell receptor–associated kinases and
their substrates [136, 137].
A functional PTPN22 polymorphism, C1858T, was found to be associated with type
1 diabetes in different Caucasian populations. The minor allele of this variant confers
predisposition to T1D, as observed in a case-control study with European American and
Sardinian cohorts [172]. The C1858T substitution changes amino acid from arginine to
tryptofan at codon 620 (R620W). This amino acid is a part of proline-rich motif
participating in binding to SH3 domain of Csk kinase that suppresses the mediators of T
cell signaling [137]. It has been shown that mutant 620W protein does not bind to Csk
[172].
The PTPN22 1858T allele has been found to be associated with several autoimmune
disorders: rheumatoid arthritis [173], systemic lupus erythematosus [174], Wegener's
granulomatosis [175] and myasthenia gravis [176] among them. Confirmation of the
PTPN22 association with T1D was performed in several independent populations [177179].
49
o Non-IDDM loci: NFKB and NFKBIA
Nuclear factor κB (NFκB) is a transcription factor that has been shown to be
involved in the regulation of many genes that encode mediators of the immune response,
embryo and cell lineage development, cell apoptosis, inflammation, cell cycle, oncogenesis,
viral replication, and a variety of autoimmune diseases [138, 180, 181]. NF-κB is formed
by homo or heterodimers of five NF-κB family members [182]. These dimers are usually
present in an inactive form in the cytoplasm, where they remain bound to a group of related
inhibitory κB (IκB) proteins [182, 183]. The activation occurs upon various stimuli leading
to a range of different reactions even specific for different cell types [138, 184]. At the
molecular level, NFκB is activated through phosphorylation of an inhibitor of NFκB (IκB).
Phosphorylated IκB is released from NFκB/IκB complex, allowing the translocation of
NFκB molecule into the nucleus. Once in the nucleus, NFκB binds to the DNA consensus
sequence (5’ GGGACTTTCC 3’) of various genes, thereby activating their transcription
[185, 186]. The activating stimuli include reactive oxygen species and advanced glycation
end products, which are toxic products of nonenzymatic glycation caused by long-term
hyperglycemia and oxidative stress.
Based on several studies it has been suggested that NFκB may work as a potential
predisposing factor in T1D. It was shown in human and rat primary β-cells that after
blocking of NF-κB activation [187] apoptosis has been prevented and thus NFκB-induced
β-cell destruction confirmed in several studies [188-191]. Paradoxically, NF-κB-regulated
genes have been shown to inhibit apoptosis in diverse cell types [192-195].
Genes coding for both NFκB and IκB proteins are interesting with regards to their
plausible role in β-cell death. Near the NFKB1 gene located on the chromosome 4q23-q24
[196], a polymorphic CA dinucleotide repeat microsatellite polymorphism with 18
described alleles has been identified [197]. In the NFKBIA that mapps to 14q13 locus, an
A/G single nucleotide polymorphism at the 3’ UTR (untranslated region) has been
identified [198]. Both of the polymorphisms have been tested for their potential impact on
amino acid sequence, structure, mRNA stability or function of resulting proteins.
50
1.5.1.3
Type 2 diabetes mellitus (T2D)
1.5.1.3.1
Incidence and prevalence of T2D
Type 2 diabetes prevalence is increasing in an epidemic fashion worldwide [87],
with a predicted doubling of the prevalence to 325 million affected persons in 25 years.
This increase has been ascribed to a collision between genes and increasing globalization of
the Western lifestyle, but few genetic variants have been consistently associated with T2D
[199-201].
The majority (~97%) of all diabetic patients have type 2 diabetes. The association of
genetic factors with T2D is well documented in many parts of the world. A strong genetic
predisposition to T2D was suggested by 3 lines of evidence [202]: (1) 58% of monozygotic
co-twins of diabetic twins were themselves diabetic compared with an expected prevalence
of 10%; (2) only 1 of 15 originally disease-discordant, monozygotic twin pairs remained
discordant for diabetes; and (3) 65% of non-diabetic monozygotic co-twins of diabetic
twins had elevated glucose values. Because concordance for diabetes was less than 100%
for twins aged 52-65 years and because twins varied in age at onset of disease, non-genetic
factors may also influence diabetes development, as proposed also by a population-based
study in Denmark [203].
Physiological variability in glucose metabolism and insulin action and secretion
even in nondiabetics has been demonstrated. Intrapopulation studies have shown that there
may be an important but delicate phenotypic difference in glucose metabolism in a given
population. In a study performed on a population of Swedish male nondiabetics, the
subjects were found to be either low or high insulin responders, a difference that was
attributed in part to physical fitness. In Australian Aborigines, major differences in insulin
response have been reported for populations living in the desert and in the coastal regions
[204].
Comparison of the phenotypic characteristics of the disease has given more
information on the nature of the genetic factors in this disease. Studies on Asian Indians
with T2D showed that they are more insulin resistant than are Caucasians with T2D, even
when the degree of obesity is comparable [205]. This suggests that there may be distinct
subtypes of type 2 diabetes in different ethnic groups depending on the phenotype of the
diabetes gene. All of these examples suggest physiological differences in glucose
51
metabolism, which may in turn be due to genetic differences even within a single
population.
1.5.1.3.2
Metabolic syndrome
Metabolic syndrome is the term frequently coined for type 2 diabetes and obesity
[206]. Metabolic syndrome is defined by American Heart Association and the US National
Heart, Lung and Blood Institute as “multiple, interrelated risk factors of metabolic origin
that appear to directly promote the development of atherosclerotic cardiovascular disease”.
It is believed that this constellation of risk factors (such as obesity, dyslipidemia,
hypertension, elevated plasma glucose and proinflammatory state) is strongly associated
with the occurrence of type 2 diabetes. Metabolic syndrome is also known as metabolic
syndrome X, syndrome X, insulin resistance syndrome, Reaven’s syndrome or CHAOS
(Australia). Nevertheless, concept of metabolic syndrome has been disputed recently by the
American Diabetes Association (ADA) and the European Association for the Study of
Diabetes.
1.5.1.3.3
Environmental risk factors
o Location
Several studies have demonstrated the importance of residence on the prevalence of
T2D. Japanese living in Brazil have higher prevalence of type 2 diabetes compared to those
in Japan [207]. Similarly, Japanese Americans living in Hawaii and Los Angeles have a
higher prevalence of type 2 diabetes than native Japanese [208]. Also differences in
prevalence in terms of rural vs. urban population have been observed: The age-adjusted
prevalence of diabetes mellitus was significantly higher in the urban population compared
to rural population [209-211]. There seem to be two reasons for this rural vs. urban
differences in prevalence: First, people in the urban setting may be more affluent and eat
more junk food in contrast to people in the rural setting. Second, people in the rural setting
are more likely to be involved in more physical activity compared to their urban
counterparts, thereby reducing the likelihood of developing type 2 diabetes.
52
o Obesity
Obesity has long been recognized as one of the risk factors for T2D development.
Its prevalence is rapidly rising to epidemic proportions around the world at an alarming
rate. According to the WHO more than 1 billion people are overweight and at least 300
million of these are obese [206]. The WHO has concluded that the fundamental causes of
this epidemic are sedentary lifestyles, high-fat, energy-dense diets, and increased
urbanization. In the U.S., results from the National Health and Nutrition Examination
Survey conducted in 1999 – 2000 indicate that a staggering 64.5% of US adults are either
overweight or obese, defined as having body mass index (BMI) of 25 or more. From 1960
to 2000, the prevalence of obesity in US adults (defined as BMI greater or equal to 30) has
risen from 13.3% to 30.9%, with the prevalence of extreme obesity (BMI greater than or
equal to 40) increasing from 0.8% to 4.7% in this 40-year period. Moreover, this trend is
expected to continue.
o Physical activity
It has been shown that regular physical activity increases insulin sensitivity and
glucose tolerance [212]. Besides that, it has been recently confirmed that physical activity
reduces the risk of type 2 diabetes [213]. A study performed on an African American
population in the United States reported that the prevalence of diabetes increases with the
degree of inactivity and obesity [208].
o Aging. Severe and prolonged stress. Drugs
Aging was proposed as one of recently identified factors playing a role in diabetes
pathogenesis. It has been previously hypothesized [214] that defects in mitochondrial
oxidative and phosphorylation capacity might be a contributing factor to the increased
triglyceride content in muscle and liver [215] and that mitochondrial function is reduced
with aging and may contribute to age-related development of insulin resistance [216].
Moreover, malfunctioning of biological clocks (located in suprachiasmatic nucleus, SCN)
was used to explain a growing number of younger subjects and even children prone to
develop type 2 diabetes. While elderly patients have an impaired function of the SCN due
to the degeneration of neurons, it is believed that in younger subjects the clock loses its
53
"feeling" for internal and external rhythms owing to modern lifestyle, which may lead to
the metabolic syndrome and type 2 diabetes [217].
Severe and prolonged stress related to today’s modern life may also be associated
with glucose intolerance and may hence increase the risk of diabetes [218]. The
development of diabetes after severe and prolonged stress seems to be caused by the
activation of the adrenal hormones, notably the glucocorticoids that have been noted to
cause glucose intolerance.
Drugs like corticosteroids and some oral contraceptive steroids may cause glucose
intolerance and T2D in susceptible individuals [219]. The role of other drugs like diuretics
and β-adrenoceptor blocking agents in the etiology of glucose intolerance and T2D is,
however, questionable.
1.5.1.3.4
Genetic risk factors
The genetic factors contributing to T2DM are much less explored than those
involved in T1D etiology; though not all of the molecules involved in the action of insulin
have been studied, some recent large-scale T2D genetic studies have succeeded in
identifying several “T2D-loci” (Table 4). Among the well-established candidates are genes
encoding peroxisome proliferator-activated receptor γ (PPARG) [200], which is a receptor
playing crucial regulatory role in adipogenesis, lipid and carbohydrate metabolism, cell
differentiation and proliferation [220], in glucose homeostasis and insulin sensitivity [221,
222]; calpain 10 (CAPN10) [223] gene, polymorphism of which was found to be associated
with reduced muscle mRNA levels and insulin resistence [224]. Among other susceptibility
loci can be mentioned TCF7L2 (Transcription factor 7-like 2) [225-228] that after
heterodimerization with β-catenin induces transcription of number of genes including
intestinal proglucagon [228], and NFKB and NFKBIA genes. Next to their role in
autoimmune processes leading to T1D, a possible link between NFκB and the development
of insulin resistance and type 2 diabetes has also been discussed [229-233]. Other candidate
genes with potential role in the pathology of T2D resulting from their physiological
functions are IRS-1 and GYS1. IRS-1 codes for insulin receptor substrate-1 protein linking
the tyrosine-phosphorylated insulin receptor to the downstream elements of the insulin
signaling pathway [234]. The Gly972Arg polymorphism was found to be associated with
decreased fasting insulin levels in T2D patients [235] and impaired insulin secretion [236].
54
XbaI polymorphism within GYS1, gene for glycogen synthase, has been described as
associated with insulin resistance [237, 238]. Interestingly, INS-VNTR has been
intensively studied not only for its association with T1D, but also for its involvement in
T2D, birth size, obesity and polycystic ovary syndrome [55, 239, 240]. These associations
stem from the potential impact that VNTR has on the transcriptional regulation of the
insulin gene in both the thymus and pancreas [161, 162]. Even though INS-VNTR has been
reported to play a role in the development of T2D and discussed the risk of class III alleles,
some studies show conflicting data [241-243].
More and more evidence is pointing to genomic mutations in various ionconducting channels as being responsible for many genetic diseases [244-246]. The
inwardly rectifying Kir6.2 channel (KCNJ11) and regulatory sulfonylurea receptor SUR1
(ABCC8) represent two subunits of the pancreatic β-cell ATP-sensitive K+ channel
complex [199, 247-249]. This complex plays a key role in glucose-stimulated insulin
secretion, which makes it a potential candidate for a genetic defect contributing to the
development of T2D [250, 251].
After several initial reports with inconsistent data [247, 252] the association
between the E23K polymorphism of KCNJ11 and the susceptibility to T2D was recently
suggested [199] and finally confirmed by meta-analysis [249]. Moreover, in cases of
permanent neonatal diabetes, several mutations responsible for amino acid substitutions in
both, the Kir6.2 (KCNJ11) and SUR1 (ABCC8) genes have been described, as well as in
congenital hyperinsulinaemia in humans, which is a rare genetic disorder characterized by
higher insulin levels in parallel with low blood glucose [253]. These findings illustrate the
importance of both genes for the development of diabetes caused by a defect in insulin
secretion [254-256]. Furthermore, functional studies show that E23K SNP is involved in
increasing the “open probability” of the Kir6.2 channel, which should lead to diminished
insulin secretion [257], and provide supporting evidence for KCNJ11 being a diabetogene.
55
Figure 11: The SUR1 receptor consists of two polypeptides, a potassium channel that is
encoded by the KIR6.2 gene (KCNJ11) and an ATP-cassette binding protein
encoded by the actual SUR1 gene (ABCC8). From Brinkmann, 2006
1.5.2 Other autoimmune diseases
1.5.2.1
Systemic lupus erythematosus
Systemic lupus erythematosus (SLE) is an autoimmune rheumatic disease,
characterized by the presence of autoantibodies. Virtually every organ or system can be
involved, but commonly, SLE affects the skin, joints, haematopoietic system, kidneys,
lungs and central nervous system.
The American College of Rheumatology has developed classification criteria for
SLE. For a diagnosis to be made, four of the following 11 criteria must be met: malar rash,
discoid
rash,
photosensitivity,
oral
ulcers,
arthritis,
serositis,
renal
disorder,
neuropsychiatric disorder, hematologic disorder, immunologic disorder, and antinuclear
antibody [258].
·
Incidence and prevalence
Worldwide, a conservative estimate states that over 5 million people have lupus.
The prevalence of SLE worldwide varies greatly. The disease, similar to many rheumatic
56
disorders, affects predominantly females (4:1 over males). From population-based
epidemiologic studies, it has been estimated that one in 3,450 women (independent of race)
in the United Kingdom, one in 250 black women in the United States, one in 1,000 Chinese
women, and one in 4,200 white women in New Zealand may have SLE [259-262].
·
Autoantibodies
B lymphocytes from patients with SLE display a lack of self-tolerance, and an
inappropriate overproduction of antibodies. The presence of antinuclear autoantibodies
(ANA) is the immunological hallmark of SLE. ANA are directed against single-stranded
and double-stranded DNA, histones, small nuclear ribonucleoproteins (snRNPs), the Ro
(SS-A) and the Lo (SS-B) ribonucleoprotein particles, Sm (Smith antigen), ENA
(extractable nuclear antigen). A positive ANA is a sensitive test found in 98% of patients
with SLE. Anti-DNA antibodies are seen in approximately 60% of patients with SLE [263].
Next to ANA, 10-65 % of patients are positive for anti-erythrocyte antibodies, 10-15% of
patients develop anti-phospholipids and anticoagulants antibodies and almost 30% express
rheumatoid factor (RF), which are immunoglobulins specific to the Fc region of IgG.
Complexes of antibodies and their antigens can damage tissues by activating complement
and by engaging Fc receptors on macrophages and other inflammatory cells.
·
Immune dysfunction, hormonal risk factors
There is a presence of overactive B cells in SLE, which produce an abundance of
autoantibodies. Nevertheless, the development and survival of these cells is dependent on
and supported by help of T cells and may also require an inappropriate lack of T cell
suppression. Some B cell activators, such as B-lymphocyte stimulator (BLyS), are reported
to be upregulated as well [264]. Further, the association between genetic complement
deficiencies and lupus development has triggered speculations about a possible role for
complement in SLE pathogenesis [265]. Defective clearance of apoptotic fragments may
provide the link between complement dysfunction and SLE [266].
There is a whole body of evidence that cytokines may be involved in lupus etiology.
Interleukin 10 (IL-10) that stimulates B cell proliferation and antibody production seems to
be a perfect candidate [267, 268]; on the other hand, in some studies [269] but not others
57
[270] tumor necrosis factor a (TNF-a) investigation suggested a protective role of this
cytokine in SLE development. Next to this, impaired secretion of tumor necrosis factor β
(TNF-β) is believed to explain at least in part the overproduction of antibodies seen in lupus
[271-273].
The SLE affects predominantly women, most frequently in the years between their
menarche and menopause, which suggests the immunomodulatory role of sex hormones in
the development of the SLE. Estrogen was found to act as a potent disease stimulator in
lupus-prone mice [274]. Moreover, an opposite, protective role of androgens in
autoimmunity development has been confirmed [275]. From other hormonal dysbalances in
SLE patients, increased serum prolactin levels were detected [276].
·
Environmental risk factors
The search for environmental triggers looked for culprits among viruses.
Nevertheless, neither Epstain-Barr virus, nor cytomegalovirus, parvovirus B19 nor the
retroviruses have been proved guilty [277].
In contrast, ultraviolet light has been proposed to function as one of the potential
disease activator. As hypothesized, anti-Ro antibodies are particularly associated with the
development of a photosensitive rash. UV light exposure causes the release of
proinflammatory cytokines and increases the rate of keratinocyte apoptosis, which leads to
extracellular exposure of autoantigens including Ro, and subsequent keratinocyte
cytotoxicity [278].
58
Chemical/physical factors
Dietary factors
Infectious agents
Hormones and
environmental oestrogens
Ÿ Aromatic amines
Ÿ Hydrazines
Ÿ Drugs (procainamide, hydralazine, chlorpromazine,
isoniazid, phenytoin, penicillamine)
Ÿ Tobacco smoke
Ÿ Hair dyes
Ÿ Ultraviolet light
Ÿ L-canavanine (alfalfa sprouts)
Ÿ High intake of saturated fats
Ÿ ?Bacterial DNA/endotoxins
Ÿ ?Retroviruses
Ÿ Hormonal replacement therapy, oral contraceptive pills
Ÿ ?Prenatal exposure to oestrogens
Table 6: Environmental factors that may be relevant in the pathogenesis of systemic lupus
erythematosus (from Mok CC and Lau CS, 2003)
·
Genetic risk factors
The presence of genetic factors contributing to SLE pathogenesis is clearly
documented by the 25% - 50% concordance between affected monozygotic twins compared
to only 2% - 5% in dizygotic pairs [279, 280]. It was estimated that at least four
susceptibility genes are required for SLE to occur [281]. In the Caucasian population an
association of HLA-DR2 and HLA-DR3 with SLE has been observed, with a relative risk
for the development of diseases of approximately 2-5 [280, 282]. A second area of potential
linkage, derived from affected sibling pair studies, was mapped to the chromosome 1q
region [283]. In general, genes expected to be involved in SLE pathogenesis are primarily
those of immune functions. Areas of interest include genes encoding proteins participating
in antigen presentation (HLA genes), apoptosis, Fc receptor, B and T cell function, and
production of cytokines and complement [284-289].
59
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
HLA genes
Non-HLA genes
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
Ÿ
DR2, DR3 (relative risk 2–5)
DR2, DR3, DR7, DQw1, DQw2, DQA1, DQB1, B8 (anti-Ro)
DR3, DR8, DRw12 (anti-La)
DR3, DQw2, DQA1, DQB1, B8 (anti-Ro and anti-La)
DR2, DR3, DR7, DQB1 (anti-DNA)
DR2, DR4, DQw5, DQw8, DQA1, DQB1 (anti-U1 ribonuclear protein)
DR2, DR4, DR7, DQw6, B61 (anti-Sm)
DR4, DR7, DQ6, DQ7, DQw7, DQw8, DQw9 (anticardiolipin or
lupus anticoagulant)
Complement genes (C2, C4, C1q)
Mannose binding lectin
Tumour necrosis factor α (TNFa)
T cell receptor (TCR)
Interleukin 6 (IL-6)
CR1
Immunoglobulin Gm and Km
FcγRIIA (IgG Fc receptor)
FcγRIIIA (IgG Fc receptor)
poly-ADP ribose polymerase (PARP)
Heat shock protein 70 (Hsp70)
Humhr 3005
Runt-related transcription factor 1 (RUNX1)
Prolactin (PRL)
Protein tyrosine phosphatase nonreceptor 22 (PTPN22)
Cytotoxic T lymphocyte-associated protein 4 (CTLA4)
Programmed cell death 1 gene (PDCD1)
Table 7: Genes involved in human systemic lupus erythematosus. Updapted from Mok CC
and Lau CS, 2003
·
The role of prolactin in the (auto)immunity
Neuropeptide hormone prolactin (PRL) is produced mainly by lactotroph
cells of the anterior pituitary gland. Nevertheless, extrapituitary production has been
described as well, especially by immune cells such as lymphocytes and monocytes,
endometrium and epithelial cells [290].
The interplay between neuroendocrine and immune system has been confirmed in
both humans and experimental animal models. The association of the prolactin and the
immune system was first proposed based on the atrophy of the thymus after
hypophysectomy when the hypophysectomized rats preserved 10 - 20% residual lactogenic
activity [291]. Nowadays, the role of PRL in maintaining and restoring immune
60
homeostasis
is
well
documented
[292,
293].
PRL
is
a
member
of
the
cytokine/haematopoietic peptide family and via its downstream target IRF-1 (interferon
regulatory factor 1) participates in the maturation and differentiation of immune cells [294296].
Prolactin and its production by lymphocytes have been suggested to play a distinct
role in the pathogenesis of systemic autoimmune diseases such as rheumatoid arthritis and
systemic lupus erythematosus [297-299]. Moreover, high serum prolactin levels were
observed in group of patients with psoriasis and link between keratinocytes
hyperproliferation and prolactin has been proposed [300]. Good response to bromocriptine
therapy (dopamine agonist) that decreased prolactin levels in patients with SLE and PsA
[301, 302] has been demonstrated.
PRL gene is located on the short arm of chromosome 6 (6p22.2 – p21.3), telomeric
to HLA region. The peripheral PRL production is driven by an alternative promoter located
5,840 base pairs (bp) upstream of the pituitary transcription start site and transcription in
lymphocytes requires a specific binding site [303]. The extrapituitary promoter contains a
G/T single nucleotide polymorphism at the position -1149 [304]. The -1149 G/T SNP is a
part of a conserved GATA sequence, site important for binding of GATA family
transcriptional factor. Higher PRL mRNA expression associated with the G allele in
lymphocytes was reported [304]. Moreover, high frequency of the G allele was described in
patients with SLE [305] but was not confirmed in other works [306].
Figure 12: Genomic organization of prolactin gene. Multiple promoters and multiple
transcription start sites (®) are indicated as well as SNP polymorphisms
61
tested for their potential association with SLE diseases. From Mellai M et al.,
2003. Exons are represented by black boxes, untranslated regions are shaded.
1.5.2.2
Rheumatoid arthritis
The rheumatoid arthritis is a chronic, recurrent, systemic inflammatory disease
primarily involving joints. Both environmental and genetic risk factors are reported to be
implicated [307, 308].
·
Autoantibodies
Autoantibodies are detected in approximately 2/3 of patients with RA and predict
more severe disease [309-311]. Two major types of autoantibodies used clinically to
characterize RA are rheumatoid factor and anti–cyclic citrullinated peptide (CCP) antibody
directed against peptides, in which arginine has been posttranslationally modified to
become citrulline [312]. These autoantibodies are strongly correlated but may represent
distinct clinical subsets of RA.
·
Immunologic pathogenesis
The cause of rheumatoid arthritis is still unknown. It was proposed that RA requires
susceptibility to the disease through genetic endowment with specific markers and an
infectious event that triggers an autoimmune response. As in other autoimmune disorders,
the "mistaken identity" theory suggests that an offending organism causes an immune
response that leaves behind antibodies specific to that organism. This subsequently leads to
the activation of the serum complement cascade and the recruitment of the phagocytic
macrophages that begin immune attack against synovium.
T cells recognition of antigen plays an important role in pathogenesis of RA as
evidenced by T cells prominent presence in the inflammatory infiltrate in the rheumatoid
synovium. Surprisingly, the synovial tissue does not present much of T cell-derived
cytokines. In contrast, there is a wide range of macrophage-derived products including
proinflammatory cytokines such as TNFa and IL-1, IL-6, IL-8 and transforming growth
factor β (TGFβ). These cytokines further exacerbate the inflammation of the synovium,
leading to edema, vasodilatation, and can activate synovial fibroblasts and other cells to
62
produce matrix metalloproteinases involved in the degradation of cartilage [313]. Once the
inflammatory reaction is established, the synovium thickens, the cartilage and the
underlying bone begin to disintegrate and evidence of joint destruction appears. Based on
the abovementioned facts, it is suspected that T cells do play a critical role in the initiation
of the synovitis; however, macrophages represent the major agent involved in its
propagation.
·
Environmental risk factors
Some infectious organisms mentioned in the context of RA triggers have been
Mycoplasma, Erysipelothrix, parvovirus B19 and rubella, but these associations have not
been supported in epidemiological studies. Nor has convincing evidence been presented for
other types of triggers such as food allergies. There is also no clear evidence that physical
and emotional effects, stress and improper diet could be a trigger for the disease. Many
negative findings suggest that either the trigger is different from patient to patient, or that
the trigger might in fact be a chance event, as hypothesized by Edwards et al. [314].
Epidemiological studies have confirmed a potential association between RA and
two herpesvirus infections: Epstein-Barr virus (EBV) and Human Herpes Virus 6 (HHV-6)
[315]. Individuals with RA are more likely to exhibit an abnormal immune response to the
Epstein-Barr virus [316] [317]. The HLA-DRB1*0404 was shown to be associated with low
frequencies of T cells specific for the EBV glycoprotein 110 and thus predisposing to RA
development Balandraud [318].
Further, cigarette smoking seems to be one of the most clearly defined
environmental risk factor for rheumatoid arthritis [319].
·
Genetic risk factors
Rheumatoid arthritis has a complex mode of inheritance. The genetic contribution to
a susceptibility to rheumatoid arthritis has been shown in studies of twins [320] and
families [321] and in genomewide linkage scans [322]; the relative risk of RA detected in
monozygotic twins is 2.0–62.0 and in siblings of affected individuals 2.0–17.0. Rheumatoid
arthritis is strongly associated with the HLA-DR4 (most specifically DR*0401 and
DR*0404) [323-325]. Actually, 90% of patients with RA have been shown to express the
HLA-DR4/DR1 alleles compared to only 40% of unaffected controls. The genetic
63
association with HLA-DR4 is believed to play a major role in switch of acute inflammation
into permanent chronic inflammatory state.
Based on the newly discovered associations, a similar role has been hypothesized
for the PTPN22 gene [173, 326] and for rs3761847 SNP that is in linkage disequilibrium
with two additional genes, TRAF1 (encoding tumor necrosis factor receptor-associated
factor 1) and C5 (complement component 5) located on chromosome 9, all involved in
regulating immune responses. Also other SNPs characterized by high level of statistical
significance in genomewide scan were detected; some of these loci contained candidate
genes of known biologic relevance to rheumatoid arthritis, including CD40, bradykinin
receptor 1 (BDKRB1), and the 17q chemokine gene cluster containing CCL1, CCL8, and
CCL13 [327]. Nonetheless, their significance in RA pathogenesis remains to be tested.
Peripheral production of prolactin and its recognized cytokine-like role in immunity made
the PRL a great candidate for RA-associated gene [297, 299].
Analysis of mRNA expression profile of immortalized lymphoblastoid cell lines
derived from 11 pairs of RA-discordant monozygotic twins revealed 747 over- and 416
underexpressed genes in RA twin compared to healthy co-twin [328]. Among three most
significantly overexpressed genes were FLJ90650 (LVRN) encoding laeverin (a novel
enzyme with sequence homology to CD13), HSD11B2 gene for 11β-hydroxysteroid
dehydrogenase type 2 (a steroid pathway enzyme), and CYR61 coding for cysteine-rich,
angiogenic inducer 61, which is a known angiogenic factor. Moreover, products of all these
genes were abundantly expressed in RA synovial tissues. Next to these, some other genes
known to be involved in RA pathogenesis were differentially expressed: proapoptotic
CASP7 and FADD were underexpressed, while antiapoptotic TIAF1 gene was found to be
overexpressed [328].
Several additional genes in association with RA reported recently and promising
candidate genes are listed in Table 8.
64
Chromosome
location
6p21.3
9q33.q34.1
2q32.2-q32.3
2q33
1p36.13
Gene symbol
Gene name
HLA-DRB1
TRAF1 – C5
locus
STAT4
CTLR4
PADI4
Human leukocyte antigen, class II
Tumor necrosis factor receptor-associated factor
1; Complement component 5
Signal transducer and activator of transcription 4
Cytotoxic T-lymphocyte associated protein 4
Citrullinating enzyme peptidylarginine
deiminase 4
Protein thyrosine phosphate non-receptor type 22
CD40 antigen
Chemokine, CC motif, ligand 1; Chemokine, CC
motif, ligand 8;
Chemokine, CC motif, ligand 13
Bradykinin receptor 1
Solute carrier family 22 (organic transporter),
member 4
Runt-related transcription factor 1
1p13
20q12-q13.2
17q11.2
PTPN22
CD40
CCL1 – CCL8 –
CCL13 locus
14q32.1-q32.2
5q31.1
BDKRB1
SLC22A4
(OCT1)
RUNX1 (AML1)
21q22.3
Table 8:
List of some genes and potential candidates involved in etiology of
rheumatoid arthritis
1.5.2.3
Psoriatic arthritis
Psoriasis is a chronic and relapsing inflammatory disease of the skin associated with
various immunologic abnormalities. Approximately 10% - 30% of psoriasis patients suffer
from joint afflictions, indicative of psoriatic arthritis (PsA), which is a chronic, recurrent,
asymmetric, erosive polyarthritis.
·
Incidence and prevalence
Prevalence of psoriasis varies from ~ 0.2% of the population in Japan; the
commonest is in Scandinavia and Northern Europe where it reaches 3% [329]. Genetic
factors appear to play a role in disease caution. Psoriasis is found in family members of
approximately 15% of patients. The concordance of psoriasis in monozygotic twins is 65–
72%, versus 15–30% in dizygotic twins. Determination of concordance in older twin pairs
from a national twin registry in Denmark revealed nearly 90–100% heritability [330]. In an
Australian study the monozygotic twin concordance rate is lower (35% for monozygotic
twins and 12% for dizygotic twins), giving an estimated heritability of 80% [331].
Psoriatic arthritis can develop at any age, however it tends to appear about 10 years
after the first signs of psoriasis, usually between the ages of 30 and 50, but it can also affect
children. Men and women are equally affected by this condition. The onset of the arthritis
65
may be acute or insidious and is usually preceded by skin disease. In about one in seven
cases the arthritis symptoms may occur before any skin involvement [313].
·
Immunologic pathogenesis
The cause of psoriasis and PsA is unknown. Evidence for immunopathogenesis in
psoriatic arthritis includes the presence of antibodies directed against skin antigens and
activated T cells in skin and synovium, however, the rheumatoid factor is absent. Psoriatic
arthritis has histopathologic features that are more characteristic of other forms of
spondyloarthritis than rheumatoid arthritis indicating distinct disease mechanisms. Matrix
metalloproteinases are strongly expressed in psoriatic arthritis synovium. Animal models
with features of psoriatic arthritis were developed to help to study precise mechanisms
involved in the etiology of PsA [332-336].
·
Environmental risk factors
It is hypothesized that psoriasis occurs due to a combination of genetic
predisposition and environmental assaults. These can include mechanical, ultraviolet and
chemical injury, infection, psychological stress and smoking or certain medications [329,
337]. The most compelling of these factors is infection with group A streptococci [338].
Streptococcal throat infections frequently precede occurrences of guttate psoriasis that can
then lead to chronic plaque psoriasis. A recent study revealed that all patients with guttate
psoriasis carried the HLA-C*0602 allele. There are also claims that chronic plaque psoriasis
may be aggravated by infection [339].
Psoriasis also occurs in association with human immunodeficiency virus (HIV)
infection [340].
·
Genetic risk factors
A variety of genetic factors have been associated with PsA [329, 341]. These
include alleles in the MHC locus such as HLA-B13, B17, B27 (almost 50% of patients
with psorasis and spondylitis [313]), B57, HLA-C6, MHC class I chain-related gene A,
allele A9 (MICA-A9), the caspase-activating recruitment domain 15 (CARD15) on
chromosome 16q, and polymorphisms in TNF gene. The stress releasing hormone and T
cell mediated cytokine prolactin may be involved in the immunological background of PsA
66
[302]. Alenius and colleagues found an association with the TNFB localized in the HLA
region, but they did not uncover any associations with any of the HLA-B or HLA-C alleles
previously shown to be associated with PsA [342].
Nevertheless, the most impressive genetic association recently described is the
strong link between the presence of certain activating killer immunoglobulin-like receptors
(KIR) on natural killer (NK) cells and PsA [343]. It was shown that subjects were most
susceptible to PsA if they had certain activating KIRs combined with the absence of HLA
ligands for corresponding homologous inhibitory KIRs. Activating KIRs have also been
associated with an increased risk of psoriasis [344]. In a revised model Nelson et al.
reported that the presence of the two activating receptors KIR2DS1 and KIR2DS2
increases the risk for PsA, and this risk is enhanced if the inhibitory HLA class I ligands for
the inhibitory KIR2DL are missing [345].
Some other studies brought controversial data [346-348], which may reflect the
multifactorial pattern of PsA inheritance [349]. Moreover, PsA is a heterogeneous disease
with multiple manifestations (erosive polyarthritis, axial disease, oligoarthritis, distal joint
arthritis with or without spondylitis) and contributing genetic factors may differ in the
subgroups.
67
1.6 THE JOINT MARKER OF ENTEROPATHY-TYPE T-CELL
LYMPHOMA AND AUTOIMMUNE DIABETES MELLITUS
During the last several years some evidences appear that lymphoproliferative
disorders and autoimmune diseases can share some aspects of their pathogenesis. Involved
mechanisms are frequently the same in both diseases: environmental factors such as the use
of cytotoxic or immunosuppressive agents, genetic factors favoring a common genetic
susceptibility, and immunologic factors such as immunoregulatory disturbances of the
immune system.
In this study, such a common mechanism for enteropathy-type T-cell lymphoma and
autoimmune diabetes mellitus is the occurrence of celiac disease (CD) [113, 350-354].
Celiac disease is a condition characterized by an increased immunologic
responsiveness to gluten, dietary wheat gliadin. Recent research revealed that CD is an
autoimmune disorder that may affect as many as 1% people in Western countries [355,
356].
As far as pathogenesis is concerned, not only an extrinsic trigger (the gluten
peptides), but also the autoantigen, the enzyme tissue transglutaminase (tTG), has been
identified [357]. tTG facilitates the breakdown of proline-rich, therefore proteolysis by
gastric and pancreatic enzymes-resistant gluten. When antibodies to tTG are generated,
enterocytes are destroyed and the common CD symptoms diagnosed. Other associated
antibodies frequent in CD are the anti-gliadin antibodies (AGA) and antibodies to
endomysium (EMA).
Celiac disease’s histological features resemble those typically found in ETL:
increased intraepithelial lymphocytes (often defined as more than 40 lymphocytes per 100
erythrocytes) [57], villous atrophy and crypt hyperplasia. Refractory celiac disease (RCD)
is clinically defined as persistent villous atrophy with crypt hyperplasia and increased
intraepithelial lymphocytes despite a strict gluten-free diet for more than 12 months or,
when there are severe persistent symptoms requiring clinical intervention regardless of the
duration of gluten-free diet [72]. Both phenotypic and immunologic features indicate RCD
may be a pre-malignant phenotype.
Further evidences for the association of CD and ETL include the demonstration of
gluten sensitivity in ETL patients [358] and also the observations that in patients with
68
celiac disease, a gluten-free diet may be protective against the development of lymphoma
[359].
Moreover, in ETL, TCR β and γ genes are clonally rearranged [56], which again
confirms the relationship between the two conditions.
It has been long recognized that several autoimmune disorders are more prevalent in
patients with celiac disease than in general population. T1D is reported as the most frequent
condition associated with celiac disease [113, 360]. On the other hand, the CD-specific
autoantibodies often accompany autoimmune process in type 1 diabetes as well. Next to
T1D, additional autoimmune diseases show gluten-sensitivity: SLE, arthritis, polymyositis,
thyroiditis and others.
There are numerous immunogenetic theories trying to explain this association by
antigenic mimicry, damage-induced neoantigen exposure, altered intestinal permeability,
idiotype network dysregulation or epitope spreading [360]. It is worth of noting that the
genetic susceptibility to CD, T1D and also ETL significantly overlaps: a certain degree of
susceptibility to these diseases is associated with the HLA-DQA1*0501-DQB1*0201 (HLADQ2) and also DQA1*0301-DQB1*0302 (HLA-DQ8) haplotypes [361]. Additionally, nonHLA genes seem to participate in common etiology as well. Of candidate genes for CD
investigated during the recent years, locus on chromosome 2q33 containing CTLA4, CD28,
and ICOS genes has given a constant association [362-365]. Among other genes that seem
to be shared in T1D and CD etiology are CCR3, MIC-A, IL-2/IL-12 locus [366], and also
analyses of several polymorphisms within the genes encoding cytokines brought positive
results.
The reason for the association of ETL with CD has not yet been fully elucidated. A
potential mechanism that drives the autoimmune celiac disease into lymphoma
development has been reported by Reines [367] who suggests that autoimmunity in
autoimmune diseases is basically antineoplastic. According to the author and based on
several theories and models, individuals suffering from autoimmune disease have inherited
many prematurely aging cells. These damaged cells adapt to in vivo conditions by
beginning to transform into cancer cells. As long as they have not fully transformed, these
cells will continue to signal ‘danger’ to the innate immune system. Thus, the clinical
69
outcome of this competition between immunity and incipient neoplasia varies upon the
degree of tumor-proneness or resistance of the individual.
In the ETL, IL-15 (highly expressed in the CD involved epithelium) has been shown
to preferentially stimulate the survival and expansion of the clonal abnormal intraepithelial
T cells from patients with RCD [368] and invoke INFγ and cytotoxic agencies such as
granzyme B and perforin production by these cells, which are directed against intestinal
epithelial cells. Eventually, stepwise appearance of genetic alteration(s) characteristic for
ETL seals doom.
Figure 13: Multistage development of ETL. The percentage in parentheses is the
approximate proportion of CD4− and CD8− or CD8+ and CD56+ IELs at
various clinicopathological stages of the development of ETL. From Isaacson &
Du, 2002
It thus seems that gluten may represent some common point of concurrence between
ETL and T1D. It is apparent that one environmental trigger may under specific stimulating
extrinsic and intrinsic conditions lead to different pathological phenotypes. This divergence
70
could be caused by several factors starting from the amount of gluten exposure, the degree
of inflammatory cytokine response, the number and type of produced antibodies, and
ending by genetic background of the individual.
71
CHAPTER 2
AIMS OF THE STUDY AND METHODOLOGY
2.1 AIMS
The aim of presented work was to contribute to the understanding of molecular
mechanisms underlying pathogenesis of several polygenic disorders by
1) identifying (new) quantitative trait loci in polygenic disease(s) with up to now
insufficiently explored genetic component contributing to disease etiology
2) testing the contribution of previously identified candidate genes to pathogenesis of
diseases with known QTLs and finding new interactions and subtyping associations
72
2.2 METHODOLOGY
1)
To further delineate 9q34 region, previously established as the most frequent
aberration occurring in enteropathy-type T-cell lymphoma, and to identify probable
candidate genes (paper No. 1), microsatellite analysis and gDNA-QPCR were
employed.
2)
A.
To characterize the genetic background of patients with different types of
diabetes mellitus (T1D in children, T1D in adults, LADA, T2D) by
performing association studies and investigating polymorphisms in HLA,
insulin (INS), Kir6.2 (KCNJ11), NFKB1 and NFKBIA genes (papers No. 2 –
8), PCR-SSP, PCR-fragment analysis and PCR-RFLP methods were used.
In our research, we were particularly interested in
i) age-related differences in immunogenetic predisposition to DM
ii) genetic discrimination of adult-onset T1D and LADA to eventually
confirm heterogeneity of group of autoimmune diabetes mellitus in
adults
iii) investigation of association of quantitative trait loci mentioned above
and biochemical and immunological markers specific for DM
B.
Determination of gene expression of diabetes associated HLA-DRB1*04 and
NFKB1 genes in antigen presenting (APC) cells of T1D patients (paper No.
8) was performed by RT-QPCR using TaqMan chemistry.
C.
A potential association between prolactin gene and three system
autoimmune disorders (systemic lupus erythematosus, rheumatoid arthritis,
psoriatic arthritis) was studied by PCR-RFLP (papers No. 9 and 10).
73
CHAPTER 3
RESULTS
3.1 LIST OF ORIGINAL PAPERS
1. Cejkova P, Zettl A, Baumgärtner AK, Chot A, Ott G, Müller-Hermelink HK,
Starostik P: Amplification of NOTCH1 and ABL1 gene loci is a frequent aberration
in enteropathy-type T-cell lymphoma. Virchows Arch 2005 Apr; 446(4):416-20.
2. Cejkova P, Novota P, Cerna M, Kolostova K, Novakova D, Kucera P, Novak J,
Andel M, Weber P, Zdarsky E: HLA DRB1, DQB1 and insulin promoter VNTR
polymorphisms: interactions and the association with adult-onset diabetes mellitus
in Czech patients. Int J Immunogenet 2008 Apr; 35(2):133-40.
3. Cejkova P, Novota P, Cerna M, Kolostova K, Novakova D, Kucera P, Novak J,
Andel M, Weber P, Zdarsky E: KCNJ11 E23K polymorphism and diabetes mellitus
with adult onset in Czech patients. Folia Biol (Praha) 2007 53(5):173-5.
4. Cerna M, Kolostova K, Novota P, Romzova M, Cejkova P, Pinterova D Pruhova S,
Treslova L, Andel M: Autoimmune Diabetes Mellitus with Adult Onset and Type 1
Diabetes Mellitus in Children Have Different Genetic Predispositions. Ann N Y
Acad Sci 2007 Sep;1110:140-50
5. Cerna M, Novota P, Kolostova K, Cejkova P, Zdarsky E, Novakova D, Kucera P,
Novak J, Andel M: HLA in Czech adult patients with autoimmune diabetes
mellitus: comparison with Czech children with type 1 diabetes and patients with
type 2 diabetes. Eur J Immunogenet 2003 30:401-407
6. Novota P, Cerna M, Kolostova K, Cejkova P, Zdarsky E, Novakova D, Kucera P,
Novak J, Andel M: Diabetes mellitus in adults: association of HLA DRB1 and
74
DQB1 diabetes risk alleles with GADab presence and C-peptide secretion. Immunol
Lett 2004; 95:229-232
7. Novota P, Cejkova P, Cerna M, Andel M: Genetic risk factors in autoimmune
diabetes mellitus, their significance and fiction. Cas Lek Cesk 2004; 143(3):159-63.
Review in Slovak.
8. Kolostova K, Pinterova D, Novota P, Romzova M, Cejkova P, Pruhova S, Lebl J,
Treslova L, Andel M, Cerna M: HLA, NFKB1 and NFKBIA Gene Polymorphism
Profile in Autoimmune Diabetes Mellitus Patients. Exp Clin Endocrinol Diabetes
2007 115:124-129
9. Fojtikova M, Cerna M, Cejkova P, Ruzickova S, Dostal C: Extrapituitary prolactin
promoter polymorphism in Czech patients with systemic lupus erythematosus and
rheumatoid arthritis. Ann Rheum Dis 2007 May;66(5):706-7
10. Stolfa J, Fojtkova M, Cejkova P, Cerna M, Sedova L, Dostal C: Polymorphism of
the prolactin extrapituitary promoter in psoriatic arthritis. Rheumatol Int 2007 Sep;
27(11):1095-6.
75
3.2 COMMENTARY ON THE ORIGINAL PAPERS
Paper No. 1 focused on the amplification of 9q34 that was identified as the most
frequent aberration occurring in ETL, and on further delineation of the amplified region and
identification of candidate gene(s). Though in multiple cases the allelic imbalances in the
9q34 locus were discontinuous, not showing any consistent pattern, detailed microsatellite
analyses on 26 ETL cases revealed allelic imbalance in ABL1 and NOTCH1 gene loci (the
former flanked by microsatellites D9S290–D9S1847, the latter by D9S158). Similar results
provided TaqMan-based quantitative real-time PCR (QPCR) showing amplification of
ABL1 and NOTCH1 exons in 50% and 65% of cases, respectively. Amplifications of the
NOTCH1 gene were more frequent than of the ABL1 gene; moreover, the analyzed
NOTCH1 exon consistently displayed higher levels of amplification than ABL1 coding
sequences.
Both ABL1 and NOTCH1 are hematopoietically significant genes and their
participation in several tumor diseases is documented, which was further confirmed by our
results indicating that both genes in question are targeted by the amplification; nevertheless,
the data favor NOTCH1 as the primary target of genomic DNA amplification in ETL.
Papers No. 2 – 8 deal with determination of genetic predisposition to different
types of diabetes mellitus. The strongest association with autoimmune diabetes mellitus has
been ascribed to HLA genes designated IDDM1 (papers No. 2 and 4 – 8; paper No. 7 is a
review in Slovak language).
Autoimmune diabetes patients manifested after 35 years of age can be divided
into two different groups having partly different HLA class II predisposition: adult onset
type 1 diabetes mellitus (T1D), carrying DRB1*03 and DRB1*04 risk alleles (P < 0.05, OR
= 3.1 for both), and increased frequency DQB1*0302 compared to healthy control subjects
(20.5% vs. 12.8%, P = NS), and a latent autoimmune diabetes in adults (LADA) with a sole
DRB1*03 risk allele (P < 0.01, OR = 3.4).
Further, we can confirm the difference in HLA allele frequencies between
glutamic acid decarboxylase antibodies (GADA) positive/negative groups and C-peptide
(CP) negative/positive groups in DM patients diagnosed after 35 years of age. The
DRB1*03, DRB1*04 and DQB1*0302 alleles are associated with decrease of CP levels and
76
hence may represent risk for CP (~ insulin) production, while DRB1*06 seems to be
protective (P < 0.05, OR = 0.4). Next to this, we found DRB1*03 allele as a significant
marker of autoantibody (GADA) development (P < 0.0001, OR = 4.2).
Like in adult-onset T1D, the HLA analysis in T1D manifested in childhood
revealed three associated alleles: HLA-DRB1*03, HLA-DRB1*04 and HLA-DQB1*0302.
These alleles confer much higher relative risk for childhood than adult onset (DR*03: OR =
3.3, P < 0.0001; DR*04: OR = 22.9, P < 0.0001; DQ*0302: OR = 46.5, P < 0.0001).
According to data presented in our studies it can be concluded that HLA-association is agedependent.
As expected, the HLA investigation did not show any association with type 2
diabetes mellitus.
Analysis of gene expression of HLA-DRB1*04 in circulating antigen presenting
cells (Paper No. 8) revealed significant increase of DRB1*04 mRNA in adult T1D
compared to T1D manifested in children. We can hypothesize either that the increased
expression of HLA-DR transcript is an age dependent factor and its elevated levels may
reflect an accumulation of environmental stress or, that the HLA-DR molecules function as
a protective factor and in comparison to the HLA-DQ molecules suppress the autoimmune
process. Nonetheless, other regulatory factors (also on the epigenetic basis) may be
involved in the selective regulation of HLA genes [369].
In addition to the HLA, Paper No. 2 studied an association of IDDM2 locus
(INS-VNTR) with three types of DM: adult-onset T1D, LADA and T2D. Rather than an
association with disease itself an association with a specific QTL has been outlined: we
were able to show that A allele and AA genotype are risk factors involved in fasting Cpeptide secretion since we have observed increased allele and genotype frequencies in CP
negative patient group (P < 0.004, OR = 3.3 for A allele and OR = 3.54 for AA genotype).
Conversely, the T allele and TT genotype frequencies were decreased in a group of diabetic
patients with low or absent C peptide secretion; however, only T allele frequency reached
statistical significance (13.8% in CP negative vs. 34.6% in CP positive patient group, P <
0.008). The INS –23HphI A allele thus appears to be associated with a decrease in secretory
activity of pancreatic β-cells.
77
With regards to autoantibody production, we have not observed any association
between INS-VNTR (–23HphI) polymorphism and levels of GADA.
The analysis of potential interaction between the HLA and the INS-VNTR loci
suggested that the simultaneous effect of the two gene polymorphisms predisposes affected
individuals to an increased risk of diabetes development. The strongest additive effect was
notable in T1D for combination of –23HphI AA genotype and HLA DRB1*03 allele. Their
simultaneous presence was found in 37.5% cases, whereas in only 9.2% in healthy control
subjects (P < 0.007).
Next to the HLA, Paper No. 8 studies the association of NFκB (NFKB1) and its
inhibitor IκB (NFKBIA) gene polymorphisms with three subgroups of autoimmune
diabetes mellitus: T1D in children, T1D in adults, and LADA. We detected 15 alleles of the
NFKB1 polymorphism (CA-repeats) in the Czech population. The alleles ranged in size
from 114 -142 bp corresponding to 10 – 25 CA dinucleotides. Only the frequency of the
A7 allele of NFKB1 (size 132 bp, 20 repeats) has been significantly increased in T1D in
adults when compared to healthy subjects (OR = 10.69, P < 0.01).
With regards to the A/G variation in 3’UTR region of the NFKBIA, gene
encoding IκB, which is a inhibitor of NFκB, significant increase in the frequency of AA
genotype was found in LADA group (OR =2.7, P < 0.001). There is also an evidence that
heterozygous AG genotype may protect against autoimmune diabetes with late onset,
according to the results acquired in the groups of LADA (non-significantly) and adult T1D
patients (OR = 0.44, P < 0.01).
For the lack of any evidence of functional significance of polymorphisms tested in
this study, we assayed the NFKB1 gene expression on the mRNA level in circulating
peripheral blood mononuclear cells in diabetic patients with different disease onset.
However, the TaqMan-based quantitative real-time PCR did reveal no difference in
expression of NFKB1 gene in APC between groups with different NFKBIA and NFKB1
genotypes.
Our data reflect the high probability of association Rel/NF-κB family genes with an
autoimmune diabetes course, but the function of the genetic variations remains to be further
elucidated.
78
Paper No. 3 describes genetic analysis of the E23K polymorphism of KCNJ11 gene
in the Czech diabetic patients diagnosed after the age of 35. In spite of several works
reporting association of K allele with type 2 diabetes mellitus, we failed to confirm it.
Moreover, we were not able to prove any significant correlation between the polymorphism
in question and C-peptide secretion. This would be of particular interest given to
physiological function of the K channel of pancreatic β-cells. Although we did detect an
increase in frequencies of K allele and KK genotype in CP negative patient group, these
differences were below statistical significance. Possible explanation for the results observed
is that the study suffers from a low statistical power to detect a possible effect.
Nevertheless, with some level of caution it might be proposed that the Kir6.2 E23K
SNP is a genetic marker associated rather directly with secretory activity of β-cells of
pancreas than with the whole disease entity.
Case-control approach has been employed also in Papers No. 9 and 10 addressing a
role of prolactin and its -1149 G/T SNP in the extrapituitary promoter in the pathogenesis
of systemic autoimmune diseases.
Paper No. 9 does not report any differences in genotype and allele frequencies of
the PRL -1149 G/T polymorphism in patients suffering from systemic lupus erythematosus
compared to group of healthy people, and no correlation with common SLE autoantibodies
was found. Nonetheless, we were able to detect an association between G allele and
articular affection (OR = 2.56, P < 0.01) and besides, age-related differences in genetic
predisposition have been observed.
In contrast to the above, increased GT heterozygous genotype was detected in
patients wit rheumatoid arthritis (OR = 1.8, P < 0.05).
Similarly to SLE, we did not identify the association of the PRL -1149 G/T SNP
with psoriatic arthritis and its characteristic clinical features (see Paper No. 10), which
suggests that non-physiological stress-induced serum prolactin levels found in PsA patients
might be due to prolactin release from the pituitary gland rather than its lymphocyte
production.
79
3.3 ORIGINAL PAPERS
80
Original paper No. 1
AMPLIFICATION OF NOTCH1 AND ABL1 GENE LOCI IS A FREQUENT ABERRATION IN
ENTEROPATHY-TYPE T-CELL LYMPHOMA
Cejkova P, Zettl A, Baumgärtner AK, Chot A, Ott G, Müller-Hermelink HK, Starostik P
Virchows Arch. 2005 Apr;446(4):416-20
81
Virchows Arch (2005) 446: 416–420
DOI 10.1007/s00428-005-1214-6
ORIGINA L ARTI CLE
Pavlina Cejkova . Andreas Zettl .
Anne K. Baumgärtner . Andreas Chott . German Ott .
Hans Konrad Müller-Hermelink . Petr Starostik
Amplification of NOTCH1 and ABL1 gene loci is a frequent
aberration in enteropathy-type T-cell lymphoma
Received: 21 October 2004 / Accepted: 12 January 2005 / Published online: 9 March 2005
# Springer-Verlag 2005
Abstract We have shown previously that amplification of
chromosomal region 9q34 is the most frequent aberration
in enteropathy-type T-cell lymphoma (ETL). To determine
the minimum amplified 9q34 region and identify possible
candidate gene(s), we performed a detailed microsatellite
screening and quantitative real-time PCR (QPCR) on 26
ETL cases. Microsatellite analysis revealed allelic imbalance in both ABL1 and NOTCH1 gene loci (microsatellites
D9S290–D9S1847 and D9S158 flanking the former and
latter genes, respectively) localized in the band 9q34. The
results were confirmed by TaqMan-based QPCR showing
amplification of ABL1 and NOTCH1 exons in 50% and
65% of cases, respectively. Amplifications of the NOTCH1
gene were more frequent than of the ABL1 gene; moreover,
the analyzed NOTCH1 exon consistently displayed higher
levels of amplification than ABL1 coding sequences. From
9q34 known genes, NOTCH1 could thus be the primary
target of genomic DNA amplification in ETL.
Keywords T-cell lymphoma . Genetic analysis .
Lymphomagenesis . Coeliac disease . Chromosome 9 .
Amplification
P. Cejkova . P. Starostik (*)
Department of Pathology and Laboratory Medicine,
Roswell Park Cancer Institute,
Elm & Carlton Streets,
Buffalo, NY, 14263, USA
e-mail: [email protected]
Tel.: +1-716-8458136
Fax: +1-716-8457638
A. Zettl . A. K. Baumgärtner . G. Ott .
H. K. Müller-Hermelink
Institute of Pathology,
Würzburg University,
Würzburg, Germany
A. Chott
Clinical Institute of Pathology,
Vienna General Hospital,
Vienna, Austria
Introduction
Enteropathy-type T-cell lymphoma (ETL) is a malignant
intestinal lymphoma associated with late presentation, rapid
progression, severe side effects (bowel perforation), and
absence of specific therapy. Intriguingly, this disease displays the same genetic background as celiac disease [10],
namely the HLA genotype DQA1*0501, DQB1*0201 [13],
and occurs solely as a sequel of either refractory celiac
disease or ulcerative colitis. Little is known about specific
aberrations in this disease. Unlike the majority of lymphomas presenting with a typical primary translocation,
e.g., t(14;18) in follicular lymphoma, no translocations have
yet been found in ETL [4, 24, 30]. In previous studies, we
identified the amplification of 9q34 to be the most frequent
aberration occurring in ETL. Besides several putative geneencoding sequences, the 9q34 region also encompasses
gene loci of two important hematopoietic regulatory genes:
ABL1 and NOTCH1. In the presented work, we further delineated and narrowed the region in 9q34 implicated in the
amplifications and show that both ABL1 and NOTCH1
genes are affected by genomic DNA amplification in 9q34.
Materials and methods
Patients and samples
For the study, 26 ETL cases from cancer registries at the
Institute of Pathology of the Würzburg University in Würzburg, Germany, and the Clinical Institute of Pathology,
Vienna General Hospital, Vienna, Austria, on which sufficient both tumor and normal (mostly muscle) tissue was
available, were selected. The samples were exclusively paraffin-embedded/formalin-fixed blocks. The diagnosis was
established according to the criteria of the World Health
Organization (WHO) Classification of Lymphoid Tumors
[15] by means of morphological and immunophenotypic
analyses of tissue sections using standard staining methods.
417
Microscopic dissection and DNA extraction
In each case, 16 serial 10-μm-thick tissue sections were cut.
The first and last cuts were stained with hematoxylin and
eosin to assure high tumor content and as guidance for the
following microdissection. Remaining tissue sections were
stained with Nuclear-Fast Red to precisely delineate tumorcontaining areas. Under microscope, areas showing ETL
were scraped to be used for microsatellite analysis. The
collected tissue was deparaffinized with xylene before digestion. DNA extraction was performed using proteinase K
and phenol-chloroform according to routine molecular biology protocols. A polymerase chain reaction (PCR)-based
clonality analysis assay for TCR-γ gene rearrangement [29]
was performed on all DNA samples to assure the control
normal DNA did not contain the lymphoma clone.
Microsatellite analysis
Microsatellite analysis was performed using a panel of six
repeats: D9S1821, D9S290, D9S1847, D9S1818, D9S1826,
and D9S158 (Ensembl, Sanger Institute; http://www.sanger.
ac.uk). Primer sequences for the amplification of microsatellites were retrieved from Genome Database (http://www.
gdb.org). PCR primers were synthesized at MWG Biotech
(High Point, NC), and one oligonucleotide of each primer
pair was labeled with fluorescent dye phosphoramidites
FAM or HEX. Thirty cycles were carried out in a PE-9600
thermal cycler (ABI, Foster City, CA) in a total volume of
20 μL. Aliquots of the PCR reactions were then mixed
with size standard and formamide, denatured, and subjected
to electrophoresis on an ABI 3100 DNA Sequencer (ABI,
Foster City, CA). The automatically collected data were
analyzed using GeneScan and Genotyper software as described in the manufacturer’s manual. Only genotypes heterozygous for a given locus were regarded as informative;
homozygosity and microsatellite instability (MSI) rendered
the particular locus unevaluable for loss of heterozygosity
(LOH) or amplification. In heterozygous genotypes, ratios
of both alleles in normal and tumor tissues were calculated.
If these ratios showed a difference of more than 40%, the
genotype was determined to display allelic imbalance. All
aberrations were confirmed by repeating the experiments.
For further determination of LOH or amplification in a
locus, first the unchanged allele was identified (by comparison with other microsatellites showing no change in the
same multiplex PCR) and then the ratios of the allele showing decreased or increased signal to the unchanged allele
were calculated; first for control DNA and then for the
tumor. Increase of the ratio by 40% in the tumor (compared
with the control) was considered amplification; a decrease
by 40% was considered LOH.
Quantitative PCR
Quantitative PCRs were performed in triplicate on the ABI
7700 Sequence Detection System instrument (ABI, Foster
City, CA) under standard conditions using the TaqMan
Universal PCR Master Mix (ABI, Foster City, CA). DNA
quantities were normalized using two reference loci located
on chromosomes usually unaltered in ETL: D6S310 and
D14S79. Primers for target loci (ABL1, NOTCH1) were
designed to amplify regions spanning across exon–intron
boundaries. TaqMan probes labeled on the 5′ end with 6′
carboxy fluorescein (FAM) and on the 3′ end with TAMRA
and primers were synthesized by MWG Biotech (High
Point, NC). Primer sequences: ABL1 (intron 6–7/exon 7)
forward primer 5′AGGAGCTCTCATGGGTGAACA 3′,
reverse primer 5′ GTTCTCCCCTACCAGGCAGTT 3′,
probe (FAM)-5′ CCTTTCTTAGAGATCTTGCTGCCC 3′(TAMRA); NOTCH1 (intron 22–23/exon 23) forward
primer 5′ GCTTGGGCCACTGACGAA 3′, reverse primer
5′ AGTCGTCCACGTTGATCTCACA 3′, probe (FAM)-5′
TGCACACCTGCGGGGCCA 3′-(TAMRA); D6S310 forward primer 5′ CCAGGGCTACACTTTCAGGTTT 3′, reverse primer 5′ GGAAGGGCTAGGAAGTGCTATG 3′,
probe (FAM)-5′ TGCCTTCTGCAATCTGGATCCCAAA
3′-(TAMRA); D14S79 forward primer 5′ TGAGACGG
AGTCTCGCTCTGT 3′, reverse primer 5′ TCCAAAAGT
CAGAGGTTTCAGTGA 3′, probe (FAM)-5′ CACGCC
ATTGCACTCCAGCCTG 3′-(TAMRA).
Sample DNAs were compared with standard DNA
(calibrator) pooled from 23 unrelated normal individuals
using the relative quantification ΔΔCt technique [19].
DNA copy numbers greater than 2.8 were considered to be
gains and those less than 1.2 losses.
Results
We previously performed comparative genomic hybridization (CGH) and microsatellite studies on a group of 26 ETL
specimens and detected several consistent abnormalities.
The most frequent aberration identified in this lymphoma
was the amplification of chromosomal region 9q34. This
amplification was detected by CGH in 58% of the examined
lymphomas [32] and in 40% by microsatellite analysis
[2]. To analyze this amplification further we carried out a
detailed analysis of the 9q34 region using six microsatellite markers: D9S1821, D9S290, D9S1847, D9S1818,
D9S1826, and D9S158. These repeats cover approximately
10 Mbp of the amplified region (Ensembl, Sanger Institute;
http://www.sanger.ac.uk) and are on average about 2 Mbp
apart. Microsatellite results revealed allelic imbalance in
16 of 26 studied cases (62%) (Fig. 1). To reliably distinguish LOH from genomic amplification, multiplex PCR
reactions with the marker showing allelic imbalance and
minimally one other marker used as an internal control
were performed. Additionally, the microsatellite analysis
results were compared with CGH data, which were available on 58% of these tumors. None of the amplifications
as determined by microsatellite analysis appeared as loss
of signal by CGH. Genotypes displaying allelic imbalance
in the 9q34 region centered around the ABL1 gene locus
(D9S290–D9S1847) in a subgroup of ETL. However, in
418
other cases, multiple discontinuous allelic imbalances involving both the ABL1 and NOTCH1 loci could be seen.
There are several genes or putative gene-coding sequences localized in the 9q34 chromosomal band. Two of
them, ABL1 and NOTCH1, are genes very well known for
their key roles in hematopoietic malignancies and would
be thus primary candidates for gene(s) dysregulated in
ETL. To further ascertain which of the known genes located between or in close vicinity to repeats showing the
most aberrations would be a good candidate for genomic
DNA amplification and to confirm the microsatellite analysis results we performed a TaqMan-based QPCR. Using
primers spanning exon–intron boundaries, we analyzed
ABL1 exon 7 and NOTCH1 exon 23 for copy number amplifications. The results revealed that NOTCH1 genomic
DNA was more frequently amplified then the ABL1 gene
locus (65% and 50%, respectively, Fig. 1). NOTCH1 also
consistently showed higher levels of amplification than
ABL1 (median 4.1 and 2.82 DNA copies, respectively,
Fig. 2). It thus seems that the NOTCH1 gene locus could be
the primary target of genomic DNA amplification in the
region 9q34 in ETL.
Discussion
Fig. 1 Genetic aberrations in 9q34 by microsatellite analysis and
quantitative real-time polymerase chain reaction (QPCR). Rows individual cases, columns microsatellites or QPCR assay targets analyzed (with their position on chromosome 9 depicted according to
the Ensembl database). Black genomic DNA amplification; falling
stripes homozygote; ascending stripes heterozygote; vertical stripes
microsatellite instability; NA no PCR amplificate. Only cases displaying allelic imbalances or those positive for genomic DNA amplification by QPCR are shown
Fig. 2 Amplification of ABL1
and NOTCH1 gene loci by
quantitative real-time polymerase chain reaction (QPCR). X
axis represents each case; Y axis
shows genomic DNA copy
number of NOTCH1 (exon 23)
and ABL1 (exon7) gene loci
(empty boxes and black
triangles, respectively; log
scale). S=standard pooled DNA
showing two DNA copies for
both analyzed loci
Previously, we reported amplification of 9q34 to be the
most frequent aberration found in ETL. In this study, we
further narrowed down the minimum amplified region in
the chromosomal band 9q34. The results of microsatellite
analysis performed on the lymphomas actually favored the
ABL1 as primary target of the amplifications. However, in
multiple cases, the allelic imbalances were discontinuous,
not showing any consistent pattern. We therefore analyzed
ABL1 exon 7 and NOTCH1 exon 23 for copy number amplifications by means of QPCR using TaqMan probes. Interestingly, the number of amplification-positive cases went up
as the QPCR detected the ABL1 or NOTCH1 gene amplifications in cases negative by microsatellite analysis. The
percentage of cases positive for the ABL1 amplification in-
419
creased from 38.5% (D9S290) or 30.8% (D9S1847) to 50%;
the NOTCH1 amplification-positive cases went up from
15.4% (D9S158) to 65.4%. Both these hematopoietically
significant genes thus seem to be targeted by the amplifications. Such amplifications of genomic DNA could be
one mechanism of how to induce their permanent overexpression, resulting in constitutive signaling that could predispose to ETL.
ABL1 is a ubiquitously expressed tyrosine kinase that is
fused to BCR in CML and precursor B-ALL cases with the t
(9;22)(q34;q11) translocation [6, 7], to ETV6 in leukemias
with the t(9;12)(q34;p13) translocation [11] and to NUP214
in episomes consisting of ABL1 and NUP214 amplicons in
T-cell acute lymphoblastic leukemia (T-ALL) [12]. Exclusive amplification of ABL1 was described in the CML-derived cell line K562 [5], in a patient with CML in lymphoid
blast crisis [18], and in three cases of secondary AML [28].
Recently, amplification involving the ABL1 gene was described in several cases of T-ALL. The fusion gene products
of the above-mentioned translocations deregulate pathways
controlling cell growth and cell cycling. In normal cells, the
ABL1 oncogene has been implicated in processes of cell
differentiation, cell division, cell adhesion, and stress response. Activity of the ABL1 protein is negatively regulated by its SH3 domain, and deletion of the SH3 domain
turns ABL1 into an oncogene. Its DNA-binding activity is
regulated by CDC2-mediated phosphorylation, suggesting
a cell cycle function for ABL1. So far, it is not completely
clear how ABL1 contributes to oncogenesis. Several possible ways of signaling through downstream targets involving PI3-kinase, RAS, RAF and the proapoptotic protein
BAD [27], JAK2 tyrosine kinase [31], MYC and anti-apoptotic BCL2 [16] were outlined. Up to this day, it has not
been fully elucidated whether only a quantitative change
in expression (overexpression) of wild-type ABL1 is sufficient to render the same or similar phenotype as fused
BCR-ABL1 or other mutants of the ABL1 gene itself.
Unlike ABL1, the NOTCH1 gene encodes a transmembrane receptor, a critical regulator of multiple developmental pathways. It has the capacity to induce apoptosis in B
cells [20], whereas it promotes survival in T cells [8]. Its
role in leukemogenesis was identified as early as in 1991
by Ellisen et al. [9] who described a t(7;9)(q34;q34.3) in TALL, causing overexpression of truncated NOTCH1 mRNA
missing the anchoring amino terminus under the influence
of the T-cell receptor promoter. Such NOTCH1 activation leads to leukemia induction as shown previously [1,
25]. Activated NOTCH1 signaling mediated by the wildtype protein has been shown to play a role in pathogenesis of anaplastic large cell lymphoma (ALCL) [17], where
NOTCH1 is highly expressed and its interaction with
Jagged1 through homotypic or heterotypic cell–cell interactions was shown to directly contribute to lymphomagenesis. Signaling through nuclear transcriptional repressor
CSL, known also as RBP-Jκ, stimulates the transcription of
downstream target genes such as HES1 and HES5 hairy
enhancer of split and HRT/HERP [14, 22]. However, genetic evidence points also to the existence of a so-far poorly
characterized CSL-independent signaling pathway [3, 23].
Recently, it was shown that activated NOTCH1 inhibits
p53-induced apoptosis through a PI3K-PKB/Akt-dependent pathway [21, 26]. PI3-kinase thus could represent a
common pathway involved in both ABL1 and NOTCH1
downstream signaling. Synergistic activation of this pathway by both ABL1 and NOTCH1 could potentiate their
impact on inhibition of apoptosis and facilitate cell cycle
deregulation by ABL1. However, such hypotheses require
further investigations. The answer will surely yield further
insights into molecular mechanisms underlying the development of ETL.
Nevertheless, there is still the possibility remaining that
a different unknown oncogene is the target of amplifications in the 9q34 chromosomal region. The amplifications
in the 9q34 are concentrated around the ABL1 gene between the D9S290 and D9S1847 microsatellites (3.8 Mbp
apart). This interval contains numerous genes and ETSs.
However, we found multiple discontinuous amplifications
in the 9q34 chromosomal band in 27% of analyzed ETLs.
These randomly appearing amplifications do not pinpoint
any particular specific interval in 9q34 as a locus of the
amplified oncogene.
We have shown previously that amplification of 9q34 is
the most frequent allelic imbalance in ETL. In the present
work, we identified two genes affected by the 9q34 amplification, ABL1 and NOTCH1, with the latter the most
frequently amplified (two-thirds of cases). This finding is
of interest, as it would point future investigations toward
exploration of the role of NOTCH1 in the pathogenesis of
this disease. Additional RNA/protein-based studies are required to confirm the primary target status of ABL1 and
NOTCH1 and possibly identify new, so-far unknown gene
(s) located in 9q34 being involved in the tumorigenesis of
ETL.
References
1. Aster JC, Xu L, Karnell FG et al (2000) Essential roles for ankyrin repeat and transactivation domains in induction of T-cell
leukemia by notch1. Mol Cell Biol 20:7505–7515
2. Baumgartner AK, Zettl A, Chott A et al (2003) High frequency
of genetic aberrations in enteropathy-type T-cell lymphoma.
Lab Invest 83:1509–1516
3. Bush G, diSibio G, Miyamoto A et al (2001) Ligand-induced
signaling in the absence of furin processing of Notch1. Dev
Biol 229:494–502
4. Carbonnel F, Lavergne A, Messing B et al (1994) Extensive
small enteropathy-type T-cell lymphoma of low-grade malignancy associated with a new chromosomal translocation. Cancer
73:1286–1291
5. Collins SJ, Groudine MT (1983) Rearrangement and amplification of c-abl sequences in the human chronic myelogenous
leukemia cell line K-562. Proc Natl Acad Sci U S A 80:4813–
4817
6. De Klein A, van Kessel AG, Grosveld G et al (1982) A cellular
oncogene is translocated to the Philadelphia chromosome in
chronic myelocytic leukaemia. Nature 300:765–767
7. De Klein A, Hagemeijer A, Bartram CR et al (1986) bcr
rearrangement and translocation of the c-abl oncogene in Philadelphia positive acute lymphoblastic leukemia. Blood 68:
1369–1375
420
8. Deftos ML, Bevan MJ (2000) Notch signaling in T cell development. Curr Opin Immunol 12:166–172
9. Ellisen LW, Bird J, West DC et al (1991) TAN-1, the human
homolog of the Drosophila notch gene, is broken by chromosomal translocations in T lymphoblastic neoplasms. Cell 66:
649–661
10. Ferguson A, Kingstone K (1996) Coeliac disease and malignancies. Acta Paediatr Suppl 412:78–81
11. Golub TR, Goga A, Barker GF et al (1996) Oligomerization of
the ABL tyrosine kinase by the Ets protein TEL in human
leukemia. Mol Cell Biol 16:4107–4116
12. Graux C, Cools J, Melotte C et al (2004) Fusion of NUP214 to
ABL1 on amplified episomes in T-cell acute lymphoblastic
leukemia. Nat Genet 36:1084–1089
13. Howell WM, Leung ST, Jones DB et al (1995) HLA-DRB,
-DQA, and -DQB polymorphism in celiac disease and enteropathy-associated T-cell lymphoma. Common features and additional risk factors for malignancy. Hum Immunol 43:29–37
14. Iso T (2001) HERP, a novel heterodimer partner of HES/E(spl)
in Notch signaling. Mol Cell Biol 21:6080–6089
15. Jaffe ES, Harris NL, Stein H et al (2001) Pathology and genetics
of tumours of haematopoietic and lymphoid tissues. World
Health Organization classification of tumours. IARC Press, Lyon,
pp 157–160
16. Jaiswal S, Traver D, Miyamoto T et al (2003) Expression of
BCR/ABL and BCL-2 in myeloid progenitors leads to myeloid
leukemias. Proc Natl Acad Sci U S A 100:10002–10007
17. Jundt F, Anagnostopoulos I, Forster R et al (2002) Activated
Notch1 signaling promotes tumor cell proliferation and survival
in Hodgkin and anaplastic large cell lymphoma. Blood 99:
3398–3403
18. Kenner L, Beham-Schmid C, Kerbl R et al (1999) ABL
amplification in a patient with lymphoid blast crisis of chronic
myelogenous leukaemia. Virchows Arch 434:255–257
19. Livak KJ, Schmittgen TD (2001) Analysis of relative gene
expression data using real-time quantitative PCR and the 2
(-Delta Delta C(T)) method. Methods 25:402–408
20. Morimura T, Goitsuka R, Zhang Y et al (2000) Cell cycle arrest
and apoptosis induced by Notch1 in B cells. J Biol Chem
275:36523–36531
21. Nair P, Somasundaram K, Krishna S (2003) Activated Notch1
inhibits p53-induced apoptosis and sustains transformation by
human papillomavirus type 16 E6 and E7 oncogenes through a
PI3K-PKB/Akt-dependent pathway. J Virol 77:7106–7112
22. Nakagawa O (2000) Members of the HRT family of basic
helix–loop–helix proteins act as transcriptional repressors
downstream of Notch signaling. Proc Natl Acad Sci U S A 97:
13655–13660
23. Ordentlich P, Lin A, Shen CP et al (1998) Notch inhibition of
E47 supports the existence of a novel signaling pathway. Mol
Cell Biol 18:2230–2239
24. Ott G, Katzenberger T, Siebert R et al (1998) Chromosomal
abnormalities in nodal and extranodal CD30+ anaplastic large
cell lymphomas: infrequent detection of the t(2;5) in extranodal
lymphomas. Genes Chromosomes Cancer 22:114–121
25. Pear WS, Aster JC, Scott ML et al (1996) Exclusive
development of T cell neoplasms in mice transplanted with
bone marrow expressing activated Notch alleles. J Exp Med
183:2283–2291
26. Sade H, Krishna S, Sarin A (2004) The anti-apoptotic effect of
Notch-1 requires p56lck-dependent, Akt/PKB-mediated signaling in T cells. J Biol Chem 279:2937–2944
27. Skorski T, Bellacosa A, Nieborowska-Skorska M et al (1997)
Transformation of hematopoietic cells by BCR/ABL requires
activation of a PI-3k/Akt-dependent pathway. EMBO J 16:
6151–6161
28. Tanaka K, Arif M, Eguchi M et al (1997) Frequent jumping
translocations of chromosomal segments involving the ABL
oncogene alone or in combination with CD3-MLL genes in
secondary leukemias. Blood 89:596–600
29. Trainor KJ, Brisco MJ, Wan JH et al (1991) Gene rearrangement in B- and T-lymphoproliferative disease detected by the
polymerase chain reaction. Blood 78:192–196
30. Wright DH, Jones DB, Clark H et al (1991) Is adult-onset
coeliac disease due to a low-grade lymphoma of intraepithelial
T lymphocytes? Lancet 337:1373–1374
31. Xie S, Wang Y, Liu J et al (2001) Involvement of Jak2 tyrosine
phosphorylation in Bcr–Abl transformation. Oncogene 20:6188–
6195
32. Zettl A, Ott G, Makulik A et al (2002) Chromosomal gains at
9q characterize enteropathy-type T-cell lymphoma. Am J Pathol
161:1635–1645
Original paper No. 2
HLA DRB1, DQB1 AND INSULIN PROMOTER VNTR POLYMORPHISMS:
INTERACTIONS AND THE ASSOCIATION WITH ADULT-ONSET DIABETES MELLITUS IN
CZECH PATIENTS
Cejkova P, Novota P, Cerna M, Kolostova K, Novakova D, Kucera P, Novak J,
Andel M, Weber P, Zdarsky E
Int J Immunogenet. 2008 Apr;35(2):133-40.
87
doi: 10.1111/j.1744-313X.2008.00749.x
HLA DRB1, DQB1 and insulin promoter VNTR polymorphisms: interactions
and the association with adult-onset diabetes mellitus in Czech patients
Blackwell Publishing Ltd
P. Cejkova,* P. Novota,* M. Cerna,* K. Kolostova,* D. Novakova,† P. Kucera,† J. Novak,‡
M. Andel,‡ P. Weber§ & E. Zdarsky¶
Summary
Both the human leucocyte antigen (HLA) DRB1 and the
HLA DQB1 gene loci play a role in the development and
progression of autoimmune diabetes mellitus (T1DM).
Similarly, the insulin promoter variable number tandem
repeats (INS-VNTR) polymorphism is also involved in the
pathogenesis of diabetes mellitus (DM). We studied the
association between each of these polymorphisms and
DM diagnosed in patients older than age 35 years.
Furthermore, we analysed possible interactions between
HLA DRB1/DQB1 and INS-VNTR polymorphisms.
Based on C-peptide and GADA levels we were able to
distinguish three types of diabetes: T1DM, latent autoimmune diabetes in adults (LADA) and T2DM. INS-VNTR
was genotyped indirectly by typing INS –23HphI A/T
polymorphism. The genotype and allele frequencies of
INS –23HphI did not differ between each of the diabetic
groups and group of healthy subjects. We did, however,
observe an association between the INS –23HphI alleles,
genotypes and C-peptide secretion in all diabetic patients:
A allele frequency was 86.2% in the C-peptide-negative
group vs. 65.4% in the C-peptide-positive group (Pcorr. <
0.005); AA genotype was found to be 72.4% in the
C-peptide-negative group vs. 42.6% in the C-peptidepositive groups (Pcorr. < 0.01). The HLA genotyping revealed
a significantly higher frequency of HLA DRB1*03 allele
in both T1DM and LADA groups when compared to
healthy subjects: T1DM (25.7%) vs. control group
* Department of Cell and Molecular Biology, Third Faculty of Medicine,
Charles University, Prague, Czech Republic, † Department of Cell and
Molecular Immunology, Third Faculty of Medicine, Charles University,
Prague, Czech Republic, ‡ Diabetes Center and Department of Internal
Medicine, Third Faculty of Medicine, Charles University, Prague, Czech
Republic, § Department of Internal Medicine, Geriatrics and General
Medicine, University Hospital Brno, Brno, Czech Republic, and
¶ Department of Pharmacology, Third Faculty of Medicine, Charles
University, Prague, Czech Republic
Received 4 November 2007; revised 4 November 2007; accepted 9
December 2007
Correspondence: Pavlina Cejkova, Department of Cell and Molecular
Biology, Institute of Biochemistry, Cell and Molecular Biology, Third
Faculty of Medicine, Charles University, Ruska 87, 10034 Prague 10,
Czech Republic. Tel: +420 267 102540; Fax: +420 267 102650;
E-mail: [email protected]
(10.15%), odds ratio (OR) = 3.06, P < 0.05; LADA
(27.6%) vs. control (10.15%), OR = 3.37, P < 0.01. The
simultaneous presence of both HLA DRB1*04 and
INS –23HphI AA genotype was detected in 37.5% of the
T1DM group compared to only 9.2% of the healthy
individuals group (OR = 5.9, Pcorr. < 0.007). We summarize
that in the Central Bohemian population of the Czech
Republic, the INS –23HphI A allele appears to be
associated with a decrease in pancreatic beta cell secretory
activity. HLA genotyping points to at least a partial
difference in mechanism, which leads to T1DM and
LADA development as well as a more diverse genetic predisposition in juvenile- and adult-onset diabetes. The
simultaneous effect of HLA and INS-VNTR alleles/
genotypes predispose individuals to an increased risk of
diabetes development.
Introduction
The aetiology of diabetes mellitus (DM) is very complex,
as this disease has both polygenetic and epigenetic factors,
which participate in its development. The contribution of
each of several genes to this complex disease state is likely
to be small; only the combined effect of several susceptibility
genes, often in concert with predisposing environmental
factors, will lead to the disease phenotype. Moreover,
interactions at the gene–gene as well as gene–environment
level will play a significant role in development of polygenetic disorders.
Diabetes mellitus is represented by two major types:
autoimmune DM and metabolic diabetes. Autoimmune
diabetes is characterized by insulin deficiency relating to
an autoimmune-destruction of pancreatic β-cells (type 1
diabetes mellitus (T1DM) or insulin-dependent, IDDM)
(Atkinson & Maclaren, 1994). The strongest genetic
association with T1DM identified so far, is that of human
leucocyte antigen (HLA) class I and class II, more specifically
DRB1*03-DQA1*0501-DQB1*0201 (HLA-DR3 haplotype) and/or HLA DRB1*04-DQA1*0301-DQB1*0302
(HLA-DR4 haplotype) (Huang et al., 1996; Cox et al.,
2001). The HLA gene complex was mapped to the short
arm of chromosome 6 and designated as IDDM1. The
second strongest association has been that of the INS-VNTR
(variable number of tandem repeats) polymorphism at the
promoter region of the insulin gene (INS), designated as
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
133
134
P. Cejkova et al.
IDDM2 (Bennett & Todd, 1996). The risk associated with
the INS-VNTR class I alleles for the T1DM phenotype is
for the most part unquestioned (Bennett & Todd, 1996).
Metabolic diabetes, which is characterized by insulin
resistance, is also known as type 2 diabetes mellitus
(T2DM) or non-insulin-dependent diabetes mellitus
(NIDDM) (Porte & Kahn, 2001; Kahn, 2003). The
genetic factors contributing to T2DM are much less
explored; though not all of the molecules involved in the
action of insulin have been studied, some recent
large-scale T2DM genetic studies have succeeded in
establishing several other ‘DM-loci’. Among the wellestablished candidates are genes encoding peroxisome
proliferator-activated receptor γ (PPARG) (Altshuler
et al., 2000), calpain 10 (CAPN10) (Horikawa et al.,
2000) and Kir6.2 (KCNJ11), representing one of the two
subunits of the pancreatic β-cell ATP-sensitive K+ channel
(Hani et al., 1998; Yamada et al., 2001; Gloyn et al.,
2003). Interestingly, INS-VNTR has been intensively
studied not only for its association with T1DM, but also
for its association with T2DM, birth size, obesity and
polycystic ovary syndrome (Bennett & Todd, 1996;
Waterworth et al., 1997; Dunger et al., 1998). These associations stem from the potential impact that VNTR has on
the transcriptional regulation of the insulin gene in both
the thymus and the pancreas (Vafiadis et al., 1996, 1997).
Even though INS-VNTR has been reported to play a role
in the development of T2DM, some studies show conflicting data (Huxtable et al., 2000; Hansen et al., 2004;
Meigs et al., 2005).
The aim of this study is to investigate the INS-VNTR,
HLA DQB1 and HLA DRB1 loci with regard to their
roles in the development and progression of DM in
patients from Central Bohemia of the Czech Republic
with disease onset occurring after age 35. Furthermore,
we aimed to explore potential interactions between the
HLA and the INS-VNTR loci.
age or older when first diagnosed with DM. Patients were
selected from an epidemiological study conducted at the
Center for Diabetes at the Third Faculty of Medicine,
Charles University in Prague, and from several outpatient
diabetic clinics in Prague and Melnik. The diagnosis of
diabetes was established according to the current WHO
criteria, which consider both the patients’ clinical features
and laboratory data; including the presence of antibodies
against glutamic acid decarboxylase GAD65 (GADA) and
serum C-peptide levels (American Diabetes Association,
2006). Patients were divided into three groups: (i)
late-onset T1DM, which required insulin therapy within
6 months after diagnosis and showed a C-peptidenegative (CP–), GADA–/+profile; (ii) latent autoimmune
diabetes in adults (LADA), defined as C-peptide-positive
(CP+), GADA+ and having a minimum 6 months between
diagnosis and initiation of insulin therapy; (iii) T2DM
with CP+, GADA– values. Patients’ clinical data are
presented in Table 1.
The control group consisted of 124 healthy individuals
randomly selected from volunteers at the Blood Donors
Center. This study was approved by the Ethical Committee from the Third Faculty of Medicine, Charles
University in Prague, and all patients signed the appropriate
informed consent forms before being allowed to participate.
Fasting C-peptide assessment
Concentrations of fasting C-peptide were measured by
radioimmunoassay (IRMA, Immunotech Czech Republic)
in plasma samples stored at –20 °C. Plasma samples from
49 healthy individuals were analysed to establish control
values: average 486 pmol L–1, minimum–maximum values
206–934 pmol L–1. Based on these control test findings, a
200 pmol L–1 or higher was considered physiological and
designated CP+.
GADA detection
Methods
Subjects
Our study included 300 patients with diabetes recruited
from the Central Bohemia region of the Czech Republic.
In order to participate, study subjects must be 35 years of
Antibody assessment was performed as described
previously (Kucera et al., 2003). Briefly, the presence of
IgG antibodies against GAD65 was confirmed by commercial ELISA test (Roche Applied Science, Mannheim,
Germany), which has a 98% specificity and 69% sensitivity
levels. Values lower than 32 ng mL–1 were considered
Table 1. Clinical features of diabetic patients
BMI (kg m–2)
Age at onset (years)
Duration of diabetes (years)
Mean C-peptide at onset (pmol L–1)a
Mean GADA (ng mL–1)b
T1DM
n = 44
LADA
n = 43
T2DM
n = 213
27.4 (22.4–36.7)
43 (36–55)
18.6 (2–50)
47.3 (0.5–194)
265.9 (2.76–3000)
31.6 (26.6–45.5)
53 (35–71)
15.4 (4–32)
542.2 (204.5–1522)
554.5 (50.4–2000)
31.4 (18.4–50.1)
53 (35–81)
13.4 (1–65)
796 (202–3369)
7.56 (0.29–49.87)
All data shown as mean with minimum and maximum value; a C-peptide cut-off level: 200 pmol L–1; b GADA cut-off level: 50 ng mL–1. BMI, body mass
index; GADA, glutamic acid decarboxylase GAD65; LADA, latent autoimmune diabetes in adults; T1DM, type 1 diabetes mellitus; T2DM, type 2
diabetes mellitus.
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
HLA DRB1, DQB1 and INS-VNTR in diabetes
negative. In order to increase the precision, the value of
50 ng mL–1 was used as the lower level cut-off and values
higher than 50 ng mL–1 (referred to as GADA+) were
considered a reliable sign of autoimmune insulitis.
135
by using an ethidium bromide staining method. The
sequences of the original primers are available from the
authors upon request.
Statistical evaluation
Genomic DNA preparation
Genomic DNA from peripheral blood samples was
extracted by QiaAmp DNA Maxi Kit spin columns (QIAGEN GmbH, Hilden, Germany) according to the
manufacturer’s instructions.
Analysis of HLA DRB1 and DQB1 alleles
HLA DRB1 and DQB1 genotyping was performed by
polymerase chain reaction (PCR) with sequence-specific
primers (SSP-PCR) (Zetterquist & Olerup, 1992; Olerup
et al., 1993) supplied by Genovision (Oslo, Norway).
INS-VNTR genotyping
The INS –23HphI SNP polymorphism, which is in a
99.7% linkage disequilibrium with both the INS-VNTR
class I allele (A allele of –23HphI) and the class III allele
(T allele of –23HphI) (Bennett & Todd, 1996), is often
used for indirect subtyping of INS-VNTR. The PCRbased restriction fragment length polymorphism (RFLP)
method employed originally designed oligonucleotides
INSM and INSO, which amplified product 159 bp long
under the following PCR conditions. The PCR was carried
out using a total volume of 20 μL, of which contained
0.1 μg of genomic DNA, 1.5 mm Mg2+, 1000 nm of each
primer, 10 mm dNTPs, 10× PCR buffer with (NH4)2SO4
and 1 U Taq DNA polymerase (Fermentas GmbH, St. LeonRot, Germany). The PCR process started with denaturation
at a temperature of 94 °C for 2 min, followed by a 3-step
cycle of 94 °C for 20 s, 64 °C for 20 s and 72 °C for 20 s.
This process was repeated 35 times. After PCR amplification, Alw26I digestion of the amplified DNA was
carried out at a temperature of 37 °C for 2 h and was
stopped by heat inactivation (65 °C for 20 min).
Following electrophoresis, both the digested and the
undigested products were visualized in a 4% agarose gel
EpiInfo software (version 3.3, October 2004, Atlanta,
Georgia) was used for statistical analysis. Allele and
genotype frequencies were determined by using a direct
counting method. Statistical differences in allele or
genotype distributions were analysed by either a χ2 or a
Fisher’s exact probability test. Significance was defined by
using a Bonferroni-corrected P-value < 0.05. The strength
of the observed associations was estimated by calculating
the odds ratios (OR), according to Wolf’s method with a
Holdane’s correction. Briefly, it was calculated as (a × d)/
(b × c), where a, b, c and d are the numbers of markerpositive patients, marker-negative patients, markerpositive controls and marker-negative controls, respectively
(Thomson, 1981).
Results
Genotype distribution of INS –23HphI in patients with
diabetes and control subjects
The case–control study showed no significant difference
in INS-VNTR allele frequencies or genotype distribution
(tested indirectly by typing –23HphI SNP) between
T1DM patients (frequency of A allele ~ INS-VNTR
class I allele: 86.2%, Pcorr. = not significant, NS), LADA
(64.5%, Pcorr. = NS), T2DM (65.9%, Pcorr. = NS) and the
control group (73.7%). All genotype distributions
(shown in Table 2) were consistent with Hardy–Weinberg
equilibrium.
We also compared allele and genotype frequencies
within all groups with diabetes and found statistically
significant differences between T1DM and the two other
types of diabetes: A allele frequency in T1DM (86.2%) vs.
LADA (64.5%, Pcorr. = 0.03) and vs. T2DM (64.9%,
Pcorr. = 0.01); AA genotype frequency in T1DM (72.4%)
vs. LADA (41.9%; Pcorr. = NS) and vs. T2DM (42.7%,
Pcorr. = 0.02).
Table 2. Frequency of alleles and genotypes of INS –23HphI polymorphism in adult patients with diabetes and control subjects
Alleles and genotypes
Locus
INS –23HphI (A/T)
A
T
AA
AT
TT
Control subjects
T1DM
Pa
LADA
Pb
T2DM
Pc
174 (73.7%)
62 (26.3%)
64 (54.2%)
46 (39%)
8 (6.8%)
50 (86.2%)
8 (13.8%)
21 (72.4%)
8 (27.6%)
0 (0%)
NS
NS
NS
NS
NS
40 (64.5%)
22 (35.5%)
13 (41.9%)
14 (45.2%)
4 (12.9%)
NS
NS
NS
NS
NS
216 (65.9%)
112 (34.1%)
70 (42.7%)
76 (46.3%)
18 (11%)
NS
NS
NS
NS
NS
For statistical analysis 2 × 2 contingency tables comparing type 1 diabetes mellitus with control subjects were used.
For statistical analysis 2 × 2 contingency tables comparing latent autoimmune diabetes in adults with control subjects were used.
c
For statistical analysis 2 × 2 contingency tables comparing type 2 diabetes mellitus with control subjects were used.
a
b
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
136
P. Cejkova et al.
either allele or genotype frequencies when we compared
the GADA+ and the GADA– groups (Table 3).
INS –23HphI and levels of C-peptide and GADA
We studied the INS –23HphI genotype distribution and
allele frequencies in patients, who were divided according
to their fasting C-peptide and GADA levels (Table 3).
There was a statistically significant difference in both
allele and genotype frequencies between the CP+ and the
CP– patient groups. This effect was independent of (lack
of) antibody production. The predisposing A allele was
detected in 86.2% of CP– patients vs. only 65.4% of CP+
patients (Pcorr. = 0.004, OR = 3.27). The AA genotype was
identified in 72.4% of CP– patients vs. 42.6% in CP+
patients (Pcorr. = 0.008, OR = 3.54). In contrast to the
above findings, we observed no significant difference in
HLA DRB1 and HLA DQB1 genes in patients with diabetes
and control subjects
Table 4 presents DQB1 and DRB1 allele frequencies
and compares each of the groups with diabetes with the
control group. Only the HLA DQB1*03 alleles were
subjected to further subtyping. We observed an increase in
the frequency of DQB1*0302 alleles in T1DM (20.5%)
when comparing to the control group (12.8%). However,
this difference did not reach statistical significance after
Bonferroni correction was applied. Nevertheless, we saw
Table 3. Frequency of alleles and genotypes of INS –23HphI polymorphism between patients with diabetes characterized by C-peptide and GADA
levels
Alleles and genotypes
Locus
INS –23HphI (A/T)
A
T
AA
AT
TT
CP– (n = 29)
CP+ (n = 195)
P (OR)
GADA– (n = 180)
GADA+ (n = 44)
P (OR)
50 (86.2%)
8 (13.8%)
21 (72.4%)
8 (27.6%)
0 (0%)
256 (65.4%)
134 (34.6%)
83 (42.6%)
90 (46.1%)
22 (11.3%)
0.004; OR = 3.27 (1.44–7.7)
0.004; OR = 0.31 (0.13–0.69)
0.008; OR = 3.54 (1.4–9.2)
NS
NS
245 (68.1%)
115 (31.9%)
83 (46.1%)
79 (43.9%)
18 (10%)
61 (69.3%)
27 (30.7%)
21 (47.7%)
19 (43.2%)
4 (9.1%)
NS
NS
NS
NS
NS
CP–, C-peptide-negative; CP+, C-peptide-positive; GADA, glutamic acid decarboxylase GAD65; NS, not significant; OR, odds ratio.
Table 4. Frequency of alleles and genotypes of HLA DQB1 and HLA DRB1 in control subjects and patients with T1DM, LADA and T2DM diagnosed
after the of age 35
Locus
Control subjects
(n = 124)
T1DM
(n = 41)
30 (37.5%)
25 (31.25%)
15 (20.5%)d
0 (0%)
14 (17.5%)
11 (13.75%)
(n = 37)
10 (13.5%)
19 (25.7%)
HLADQB1 alleles
DQB1*02
DQB1*03
DQB1*0302
DQB1*04
DQB1*05
DQB1*06
HLADRB1 alleles
DRB1*01
DRB1*03
61 (24.8%)
88 (35.8%)
24 (12.8%)d
3 (1.2%)
38 (15.4%)
56 (22.8%)
(n = 107)
19 (9.2%)
21 (10.15%)
DRB1*04
21 (10.15%)
19 (25.7%)
DRB1*07
DRB1*08
DRB1*09
DRB1*10
DRB1*11
DRB1*12
DRB1*13
DRB1*14
DRB1*15
DRB1*16
33 (15.9%)
6 (2.9%)
2 (1.0%)
1 (0.5%)
29 (14.0%)
2 (1.0%)
29 (14.0%)
7 (3.4%)
33 (15.9%)
4 (1.9%)
8 (10.8%)
2 (2.7%)
0 (0%)
2 (2.7%)
5 (6.8%)
0 (0%)
4 (5.4%)
0 (0%)
4 (5.4%)
1 (1.3%)
Pa
NS
NS
NSd
NS
NS
NS
NS
< 0.05; OR = 3.06
(1.45–6.44)
< 0.05; OR = 3.06
(1.45–6.44)
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
LADA
(n = 31)
18 (30%)
20 (33.4%)
4 (8.2%)d
2 (3.3%)
12 (20%)
8 (13.3%)
(n = 29)
5 (8.6%)
16 (27.6%)
7 (12.1%)
4 (6.9%)
3 (5.2%)
1 (1.7%)
0 (0%)
10 (17.3%)
0 (0%)
4 (6.9%)
1 (1.7%)
6 (10.3%)
1 (1.7%)
Pb
NS
NS
NSd
NS
NS
NS
NS
< 0.01; OR = 3.37
(1.53–7.45)
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
T2DM
(n = 158)
Pc
73 (23.2%)
106 (33.7%)
23 (8.6%)d
8 (2.5%)
60 (19.0%)
68 (21.6%)
(n = 90)
16 (8.9%)
15 (8.3%)
NS
NS
NSd
NS
NS
NS
24 (13.3%)
NS
29 (16.1%)
8 (4.4%)
3 (1.7%)
3 (1.7%)
25 (13.9%)
2 (1.1%)
16 (8.9%)
15 (8.3%)
19 (10.6%)
5 (2.8%)
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
NS
For statistical analysis 2 × 2 contingency tables comparing T1DM with control subjects were used.
For statistical analysis 2 × 2 contingency tables comparing LADA with control subjects were used.
c
For statistical analysis 2 × 2 contingency tables comparing T2DM with control subjects were used.
d
For statistical analysis 2 × 2 contingency tables were used that compared HLADQB1*0302 allele vs. non-HLADQB1*0302 alleles (only subtyped alleles).
LADA, latent autoimmune diabetes in adults; NS, not significant; T1DM, type 1 diabetes mellitus; T2DM, type 2 diabetes mellitus.
a
b
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
HLA DRB1, DQB1 and INS-VNTR in diabetes
137
a significant increase in the frequency of HLA DRB1*03
and HLA DRB1*04 alleles in patients with T1DM (both
alleles detected in 25.7% in T1DM vs. only 10.15% in
control, Pcorr. < 0.05, OR = 3). Furthermore, the LADA
group showed higher levels of DRB1*03 when compared
to the control group of healthy individuals (27.6% vs.
10.15%, respectively; Pcorr. < 0.01, OR = 3.4).
genotype was detected in the healthy control group in
39.8% of subjects, whereas only 16.7% of individuals
showed this absence in the T1DM group (OR = 0.3,
P < 0.05, Pcorr. = NS). Those subjects with T1DM possessing
a combination of both risk markers DRB1*04 and
–23HphI AA showed a significant risk increase (OR = 5.9,
Pcorr. < 0.007). Data are summarized in Table 6.
HLA allele frequencies and levels of C-peptide and GADA
Discussion
In order to study HLA DQB1 and DRB1 allele frequencies
in association with these clinical features, all patients with
diabetes had to be divided according to C-peptide and
GADA levels. We did a separate comparison with each of
the patient groups and the healthy control subjects and
our findings highlighted two risk alleles in the CP– patient
group: DRB1*03 (27.9% vs. 10.15% in control; Pcorr. <
0.001, OR = 3.43) and DRB1*04 (25.6% vs. 10.15% in
control; Pcorr. < 0.001, OR = 3.04). In the GADA+ group
there was one predisposing allele detected, DRB1*03
(Pcorr < 0.0001, OR = 4.24) (Table 5).
We performed extensive genotyping for two HLA genes
and the INS –23HphI single nucleotide polymorphism in
300 diabetic patients, all of whom developed their disease
after age 35. This association was studied by comparing
patients’ allele and genotype frequencies with that of a
healthy control group consisted of 124 subjects. The contributing coauthors have published their preliminary findings
elsewhere (Cerna et al., 2003; Novota et al., 2004).
The INS –23HphI promoter polymorphism did not
show to have any association with any group of patients
with DM diagnosed after age 35. Since the role of INSVNTR in the development of T1DM is generally accepted
(Bennett & Todd, 1996; Perez et al., 2003; Barratt et al.,
2004; Guja et al., 2004), a possible explanation for the
above discrepancy centres around the age of onset. It is
likely that the late onset of the disease may be caused, or
at least modified, by gene loci different from the ones associated with juvenile onset, as shown in HLA (Cerna et al.,
2003; Murao et al., 2004; Cerna et al., in press). HLA
DQB1 is considered to be the main T1DM predisposing
HLA DRB1*03 and HLA DRB1*04 interactions with
INS –23HphI
The simultaneous combination of the DRB1*03 and the
INS –23HphI T allele was found in 26.9% of the LADA
patients and in only 7.1% of the control subjects
(OR = 4.8, P < 0.005, Pcorr. = NS). The simultaneous
absence of the DRB1*03 allele and the INS –23HphI AA
Table 5. Allele frequencies of HLA DQB1 and HLA DRB1 polymorphisms in diabetic patients characterized by C-peptide and GADA levels
Locus
HLADQB1 alleles
DQB1*02
DQB1*03
DQB1*0302
DQB1*04
DQB1*05
DQB1*06
Locus
HLADRB1 alleles
DRB1*01
DRB1*03
DRB1*04
DRB1*07
DRB1*08
DRB1*09
DRB1*10
DRB1*11
DRB1*12
DRB1*13
DRB1*14
DRB1*15
DRB1*16
CP– (n = 49)
CP+ (n = 153)
GADA– (n = 182)
GADA+ (n = 57)
Controls (n = 124)
35 (37.2%)
30 (31.9%)
19 (24.7%)b
1 (1.1%)
17 (18.1%)
11 (11.7%); OR = 0.4c
92 (24.3%)
127 (33.5%)
34 (12.1%)b
10 (2.6%)
73 (19.3%)
77 (20.3%)
85 (23.6%)
121 (33.6%)
39 (12.4%)b
8 (2.2%)
75 (20.85%)
71 (19.7%)
42 (37.8%)
35 (31.5%)
13 (14.9%)b
3 (2.7%)
14 (12.6%)
17 (15.4%)
61 (24.8%)
88 (35.8%)
24 (12.8%)b
3 (1.2%)
38 (15.4%)
56 (22.8%)
CP– (n = 43)
CP+ (n = 122)
GADA– (n = 111)
GADA+ (n = 51)
Controls (n = 107)
8 (7.8%)
33 (32.3%); OR = 4.24e
16 (15.7%)
6 (5.9%)
5 (4.9%)
1 (1%)
0 (0%)
14 (13.7%)
0 (0%)
8 (7.8%)
1 (1%)
9 (8.8%)
1 (1%)
19 (9.2%)
21 (10.15%)
21 (10.15%)
33 (15.9%)
6 (2.9%)
2 (1.0%)
1 (0.5%)
29 (14.0%)
2 (1.0%)
29 (14.0%)
7 (3.4%)
33 (15.9%)
4 (1.9%)
11 (12.8%)
24 (27.9%); OR = 3.43d
22 (25.6%); OR = 3.04d
8 (9.3%)
3 (3.5%)
0 (0%)
2 (2.3%)
6 (7%)
0 (0%)
4 (4.6%)
0 (0%)
4 (4.6%)
2 (2.3%)
21 (8.6%)
31 (12.7%)
33 (13.5%)
33 (13.5%)
12 (4.9%)
4 (1.6%)
4 (1.6%)
35 (14.4%)
2 (0.8%)
22 (9%)
16 (6.6%)
25 (10.3%)
6 (2.5%)
24 (10.8%)
22 (9.9%)
36 (16.2%)
34 (15.3%)
10 (4.5%)
3 (1.3%)
5 (2.2%)
27 (12.2%)
2 (0.9%)
17 (7.7%)
15 (6.8%)
20 (9%)
7 (3.2%)
a
P-values are corrected (Bonferroni).
For statistical analysis 2 × 2 contingency tables were used that compared HLA DQB1*0302 allele vs. non HLA DQB1*0302 alleles (only subtyped
alleles).
c
Pcorr. < 0.05; d Pcorr. < 0.001; e Pcorr. < 0.0001.
CP–, C-peptide-negative; CP+, C-peptide-positive; GADA–, glutamic acid decarboxylase GAD65 negative; GADA+, glutamic acid decarboxylase
GAD65 positive; HLA, human leucocyte antigen.
b
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
138
P. Cejkova et al.
Table 6. Combination of –23HphI T and HLA DRB1*03 in LADA and healthy control subjects
None of the two markers
Only one of the two markers
Both of the two markers
Control subjects
(n = 98)
LADA
(n = 26)
OR (95% CI)
P-value
Pcorr.
41 (41.8%)
39 (39.8%)
7 (7.1%)
7 (26.9%)
15 (57.7%)
7 (26.9%)
0.5 (0.2–1.45)
2.1 (0.8–5.4)
4.8 (1.3–17.7)
NS
NS
< 0.005
NS
NS
NS
Correction factor used: 15.
Combination of –23HphI AA and HLA DRB1*03 in T1DM and healthy control subjects
None of the two markers
Only one of the two markers
Both of the two markers
Control subjects
(n = 98)
T1DM
(n = 24)
OR (95% CI)
P-value
Pcorr.
39 (39.8%)
48 (49%)
11 (11.2%)
4 (16.7%)
14 (58%)
6 (25%)
0.3 (0.1–1.0)
1.4 (0.5–4.0)
2.6 (0.8–9.1)
< 0.05
NS
NS
NS
NS
NS
Correction factor used: 15.
Combination of –23HphI AA and HLA DRB1*04 in T1DM and healthy control subjects
None of the two markers
Only one of the two markers
Both of the two markers
Control subjects
(n = 98)
T1DM
(n = 24)
OR (95% CI)
P-value
Pcorr.
35 (35.7%)
54 (55.1%)
9 (9.2%)
6 (25%)
9 (37.5%)
9 (37.5%)
0.6 (0.2–1.8)
0.5 (0.2–1.3)
5.9 (1.8–19.9)
NS
NS
< 0.001
NS
NS
< 0.007
Correction factor used: 15.
CI, confidence interval; HLA, human leucocyte antigen; LADA, latent autoimmune diabetes in adults; NS, not significant; OR, odds ratio;
T1DM, type 1 diabetes mellitus.
factor in children. However, with the progression of age
the genetic predisposition is switched to HLA DRB1,
which has been confirmed by our data. The HLA profile
in the tested groups with diabetes revealed only a slight
increase in DQB1*02 and DQB1*0302 alleles, and a
decrease in DQB1*06 allele in T1DM. The strongest
predisposing factors was shown to be DRB1*03 and
DRB1*04 alleles. The results obtained from our work are,
for the most part, in agreement with data published
elsewhere and support the hypothesis that there are
different genetic predisposing factors for juvenile and
adult-onset DM.
Nevertheless, despite the lack of association between
the INS –23HphI SNP and diabetes, we did observe a statistical significant difference in allele and genotype frequencies between adult-onset T1DM and other types of
diabetes (LADA, T2DM). This finding contributes to a
long history of discussion on genetic profiles of lateonset T1DM and LADA. It is believed that genetics may
answer the question, whether T1DM in adults with rapid
progression and LADA diabetes with a less aggressive
progress are two distinct forms of autoimmune diabetes
with late onset (Tuomi et al., 1999; Hosszufalusi et al.,
2003; Fourlanos et al., 2005; Palmer et al., 2005).To date,
results have shown discrepancies (Tuomi et al., 1999;
Desai et al., 2006). In this work, the INS –23HphI A allele
was detected more often in T1DM patients than in the
LADA group, and the difference was statistically significant (Pcorr. = 0.03). The disparity between the two diabetic
groups reflects the pancreatic beta cells secretory activity,
shown by C-peptide levels, since the presence or absence
of C-peptide secretion is the major criteria for discrimination between these two types of autoimmune diabetes.
With regard to HLA, DRB1*03 was identified as the main
predisposing allele in autoimmune LADA diabetes
(Pcorr. < 0.01). DRB1*03 and DRB1*04 alleles were identified as the genetic risk factors in adult-onset T1DM
(both Pcorr. < 0.05). The DRB1*04 is in linkage disequilibrium with DQB1*0302 allele, for which we observed
higher (but not significant) frequencies. Data presented
here thus support the generally accepted theory that there
are partially different mechanisms in the development
and/or progression of T1DM and LADA diabetes.
A quantitative trait study revealed an association
between C-peptide secretion and INS-VNTR polymorphism alleles. We observed a positive correlation between
the INS-VNTR class III allele (–23HphI T allele) and the
fasting C-peptide levels (34.6% in CP+ compared to
13.8% in the CP– group of patients). This finding correlates with the association between the INS-VNTR and the
anti-insulin antibody production detected in our study
(data not shown). This further suggests a protective effect
by the INS-VNTR class III allele (–23HphI T allele) and
an increased risk effect by the INS-VNTR class I allele
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
HLA DRB1, DQB1 and INS-VNTR in diabetes
(–23HphI A allele) on insulin peptide structure and
secretion. Our findings fit the accepted scheme of insulin
expression in the thymus and pancreas, explaining the
process of development of antibodies against insulin.
The class I/I genotype was found to be responsible for
lower insulin mRNA expression in the fetal thymus, but
increased expression in both the fetal and the adult
pancreas. The class III/III genotype was associated with
higher antigen levels in the thymus, which could more
efficiently induce negative selection of insulin-specific
T-lymphocytes (Vafiadis et al., 1996, 1997). Other than
this, we were unable to find any association between
INS-VNTR and the presence of the GAD65 antibody.
We also analysed the distribution of the HLA DQB1
and HLA DRB1 alleles with regards to their association
with GADA and C-peptide levels. The two risk HLA
alleles were detected in the CP– patient group. Both
DRB1*03 and DRB1*04 alleles showed similar OR (3.43
and 3.04, respectively), suggesting equal predisposing
effects. The DRB1*03 allele was also identified as a risk
allele in GADA+ patients with the OR value higher than
4. These results are in agreement with published data
(Gambelunghe et al., 2000).
Finally, we aimed to investigate the additive effect of the
two candidate loci, IDDM1 and IDDM2. Data presented
in Table 6 show that the combined effect of both the HLA
class II allele (DRB1*04) and the INS –23HphI AA
genotype should increase the risk of developing T1DM by
almost twofold. Therefore, one can say that both the HLA
class II allele DRB1*04 and the INS –23HphI AA
genotype are important in conferring genetic susceptibility
to T1DM.
To conclude, it seems that the IDDM2 quantitative trait
locus represented by the INS-VNTR does not associate by
itself with the disease, but rather with a specific phenotype. Our results suggest that either INS-VNTR, or
INS –23HphI or both are genetic factor(s) that influence
pancreatic beta cell secretory activity, which can eventually
result in a pathological phenotype. HLA class II genotyping
confirmed different immunogenetic aetiopathogenesis of
T1DM and LADA disease. It also suggested a diverse
mechanism that launches clinical signs of autoimmune
diabetes in children and adults. We have shown that the
simultaneous effect of IDDM1 and IDDM2 loci predispose
to an increase in risk of disease development.
Acknowledgements
This study was supported by the Research Program of
the Czech Ministry of Education and Youth, MSM
0021620814: Prevention, diagnostics and therapy of
diabetes mellitus, metabolic and endocrinal affections
of organism. The authors would like to thank Ashley
Nicholson for language and style correction.
References
Altshuler, D., Hirschhorn, J.N., Klannemark, M., Lindgren, C.M.,
Vohl, M.C., Nemesh, J. et al. (2000) The common PPARgamma
139
Pro12Ala polymorphism is associated with decreased risk of type
2 diabetes. Nature Genetics, 26, 76.
American Diabetes Association (2006) Diagnosis and classification
of diabetes mellitus. Diabetes Care, 29, S43–S48.
Atkinson, M.A. & Maclaren, N.K. (1994) The pathogenesis of
insulin-dependent diabetes mellitus. New England Journal of
Medicine, 331, 1428.
Barratt, B.J., Payne, F., Lowe, C.E., Hermann, R., Healy, B.C.,
Harold, D. et al. (2004) Remapping the insulin gene/IDDM2
locus in type 1 diabetes. Diabetes, 53, 1884.
Bennett, S.T. & Todd, J.A. (1996) Human type 1 diabetes and the
insulin gene: Principles of mapping polygenes. Annual Review of
Genetics, 30, 343.
Cerna, M., Novota, P., Kolostova, K., Cejkova, P., Zdarsky, E.,
Novakova, D., Kucera, P., Novak, J. & Andel, M. (2003) HLA in
Czech adult patients with autoimmune diabetes mellitus:
comparison with Czech children with type 1 diabetes and patients
with type 2 diabetes. European Journal of Immunogenetics, 30,
401.
Cerna, M., Kolostova, K., Novota, P., Romzova, M., Cejkova, P.,
Pinterova, D., Pruhova, S., Treslova, L. & Andel, M. (2007)
Autoimmune diabetes mellitus with adult onset and type 1
diabetes mellitus in children have different genetic
predispositions. Annals of the New York Academy of
Sciences, 1110, 140.
Cox, N.J., Wapelhorst, B., Morrison, V.A., Johnson, L., Pinchuk, L.,
Spielman, R.S., Todd, J.A. & Concannon, P. (2001) Seven regions
of the genome show evidence of linkage to type 1 diabetes in a
consensus analysis of 767 multiplex families. American Journal of
Human Genetics, 69, 820.
Desai, M., Zeggini, E., Horton, V.A., Owen, K.R., Hattersley, A.T.,
Levy, J.C. et al. (2006) The variable number of tandem repeats
upstream of the insulin gene is a susceptibility locus for latent
autoimmune diabetes in adults. Diabetes, 55, 1890.
Dunger, D.B., Ong, K.K., Huxtable, S.J., Sherriff, A., Woods, K.A.,
Ahmed, M.L. et al. (1998) Association of the INS VNTR with size
at birth. ALSPAC Study Team. Avon Longitudinal Study of
Pregnancy and Childhood. Nature Genetics, 19, 98.
Fourlanos, S., Dotta, F., Greenbaum, C.J., Palmer, J.P., Rolandsson,
O., Colman, P.G. & Harrison, L.C. (2005) Latent autoimmune
diabetes in adults (LADA) should be less latent. Diabetologia, 48,
2206.
Gambelunghe, G., Forini, F., Laureti, S., Murdolo, G., Toraldo, G.,
Santeusanio, F., Brunetti, P., Sanjeevi, C.B. & Falorni, A. (2000)
Increased risk for endocrine autoimmunity in Italian type 2
diabetic patients with GAD65 autoantibodies. Clinical
Endocrinology, 52, 565.
Gloyn, A.L., Weedon, M.N., Owen, K.R., Turner, M.J., Knight,
B.A., Hitman, G. et al. (2003) Large-scale association studies of
variants in genes encoding the pancreatic beta-cell KATP channel
subunits Kir6.2 (KCNJ11) and SUR1 (ABCC8) confirm that the
KCNJ11 E23K variant is associated with type 2 diabetes.
Diabetes, 52, 568.
Guja, C., Guja, L., Nutland, S., Rance, H., Todd, J.A. &
Ionescu-Tirgoviste, C. (2004) Strong association of insulin gene
INS-VNTR polymorphisms with type 1 diabetes in the
Romanian population. Romanian Journal of Internal Medicine,
42, 313.
Hani, E.H., Boutin, P., Durand, E., Inoue, H., Permutt, M.A., Velho,
G. & Froguel, P. (1998) Missense mutations in the pancreatic islet
beta cell inwardly rectifying K+ channel gene (KIR6.2/BIR): a
meta-analysis suggests a role in the polygenic basis of type II
diabetes mellitus in Caucasians. Diabetologia, 41, 1511.
Hansen, S.K., Gjesing, A.P., Rasmussen, S.K., Glumer, C.,
Urhammer, S.A., Andersen, G. et al. (2004) Large-scale studies
of the HphI insulin gene variable-number-of-tandem-repeats
polymorphism in relation to type 2 diabetes mellitus and
insulin release. Diabetologia, 47, 1079.
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
140
P. Cejkova et al.
Horikawa, Y., Oda, N., Cox, N.J., Li, X., Orho-Melander, M.,
Hara, M. et al. (2000) Genetic variation in the gene encoding
calpain-10 is associated with type 2 diabetes mellitus. Nature
Genetics, 26, 163.
Hosszufalusi, N., Vatay, A., Rajczy, K., Prohaszka, Z., Pozsonyi, E.,
Horvath, L. et al. (2003) Similar genetic features and different
islet cell autoantibody pattern of latent autoimmune diabetes in
adults (LADA) compared with adult-onset type 1 diabetes with
rapid progression. Diabetes Care, 26, 452.
Huang, W., Connor, E., Rosa, T.D., Muir, A., Schatz, D., Silverstein,
J., Crockett, S., She, J.X. & Maclaren, N.K. (1996) Although
DR3-DQB1*0201 may be associated with multiple component
diseases of the autoimmune polyglandular syndromes, the human
leukocyte antigen DR4-DQB1*0302 haplotype is implicated only
in beta-cell autoimmunity. Journal of Clinical Endocrinology and
Metabolism, 81, 2559.
Huxtable, S.J., Saker, P.J., Haddad, L., Walker, M., Frayling, T.M.,
Levy, J.C., Hitman, G.A., O’Rahilly, S., Hattersley, A.T. &
McCarthy, M.I. (2000) Analysis of parent-offspring trios
provides evidence for linkage and association between the insulin
gene and type 2 diabetes mediated exclusively through paternally
transmitted class III variable number tandem repeat alleles.
Diabetes, 49, 126.
Kahn, S.E. (2003) The relative contributions of insulin resistance and
beta-cell dysfunction to the pathophysiology of type 2 diabetes.
Diabetologia, 46, 3.
Kucera, P., Novakova, D., Behanova, M., Novak, J., TlaskalovaHogenova, H. & Andel, M. (2003) Gliadin, endomysial and
thyroid antibodies in patients with latent autoimmune diabetes of
adults (LADA). Clinical and Experimental Immunology, 133,
139.
Meigs, J.B., Dupuis, J., Herbert, A.G., Liu, C., Wilson, P.W. &
Cupples, L.A. (2005) The insulin gene variable number tandem
repeat and risk of type 2 diabetes in a population-based sample of
families and unrelated men and women. Journal of Clinical
Endocrinology and Metabolism, 90, 1137.
Murao, S., Makino, H., Kaino, Y., Konoue, E., Ohashi, J., Kida, K.
et al. (2004) Differences in the contribution of HLA-DR and -DQ
haplotypes to susceptibility to adult- and childhood-onset type 1
diabetes in Japanese patients. Diabetes, 53, 2684.
Novota, P., Cerna, M., Kolostova, K., Cejkova, P., Zdarsky, E.,
Novakova, D., Kucera, P., Novak, J. & Andel, M. (2004)
Diabetes mellitus in adults: association of HLA DRB1 and DQB1
diabetes risk alleles with GADab presence and C-peptide
secretion. Immunology Letters, 95, 229.
Olerup, O., Aldener, A. & Fogdell, A. (1993) Hla-Dqb1 and
Hla-Dqa1 typing by PCR amplification with sequence-specific
primers (Pcr-Ssp) in 2 hours. Tissue Antigens, 41, 119.
Palmer, J.P., Hampe, C.S., Chiu, H., Goel, A. & Brooks-Worrell,
B.M. (2005) Is latent autoimmune diabetes in adults distinct from
type 1 diabetes or just type 1 diabetes at an older age? Diabetes,
54 (Suppl. 2), S62.
Perez, D.N., Bilbao, J.R., Calvo, B., Vitoria, J.C., Vazquez, F. &
Castano, L. (2003) 5′-Insulin gene VNTR polymorphism is
specific for type 1 diabetes: no association with celiac or Addison’s
disease. Annals of the New York Academy of Sciences, 1005, 319.
Porte, D. Jr & Kahn, S.E. (2001) beta-cell dysfunction and failure in
type 2 diabetes: potential mechanisms. Diabetes, 50 (Suppl. 1),
S160.
Thomson, G. (1981) A review of theoretical aspects of HLA and
disease associations. Theoretical Population Biology, 20, 168.
Tuomi, T., Carlsson, A., Li, H., Isomaa, B., Miettinen, A., Nilsson,
A. et al. (1999) Clinical and genetic characteristics of type 2
diabetes with and without GAD antibodies. Diabetes, 48, 150.
Vafiadis, P., Bennett, S.T., Colle, E., Grabs, R., Goodyer, C.G. &
Polychronakos, C. (1996) Imprinted and genotype-specific
expression of genes at the IDDM2 locus in pancreas and
leucocytes. Journal of Autoimmunity, 9, 397.
Vafiadis, P., Bennett, S.T., Todd, J.A., Nadeau, J., Grabs, R.,
Goodyer, C.G., Wickramasinghe, S., Colle, E. & Polychronakos,
C. (1997) Insulin expression in human thymus is modulated by
INS VNTR alleles at the IDDM2 locus. Nature Genetics, 15, 289.
Waterworth, D.M., Bennett, S.T., Gharani, N., McCarthy, M.I.,
Hague, S., Batty, S. et al. (1997) Linkage and association of
insulin gene VNTR regulatory polymorphism with polycystic
ovary syndrome. Lancet, 349, 986.
Yamada, Y., Kuroe, A., Li, Q., Someya, Y., Kubota, A., Ihara, Y.,
Tsuura, Y. & Seino, Y. (2001) Genomic variation in pancreatic ion
channel genes in Japanese type 2 diabetic patients. Diabetes/
Metabolism Research and Reviews, 17, 213.
Zetterquist, H. & Olerup, O. (1992) Identification of the
Hla-Drb1-*-04, Hla-Drb1-*-07, and Hla-Drb1-*-09 alleles by
PCR amplification with sequence-specific primers (PCR-SSP)
in 2 hours. Human Immunology, 34, 64.
© 2008 The Authors
Journal compilation © 2008 Blackwell Publishing Ltd, International Journal of Immunogenetics 35, 133–140
Original paper No. 3
KCNJ11 E23K POLYMORPHISM AND DIABETES MELLITUS WITH ADULT ONSET IN CZECH
PATIENTS
Cejkova P, Novota P, Cerna M, Kolostova K, Novakova D, Kucera P, Novak J,
Andel M, Weber P, Zdarsky E
Folia Biol (Praha). 2007 53(5):173-5
96
Original Article
KCNJ11 E23K Polymorphism and Diabetes Mellitus
with Adult Onset in Czech Patients
(diabetes mellitus / Kir6.2 / insulin / C-peptide)
P. ČEJKOVÁ1, P. NOVOTA1, M. ČERNÁ1, K. KOLOŠTOVÁ1, D. NOVÁKOVÁ2,
P. KUČERA2, J. NOVÁK3, M. ANDĚL3, P. WEBER4, E. ŽDÁRSKÝ5
1
Department of Cell and Molecular Biology, 2Department of Cell and Molecular Immunology, 3Diabetes
Center and Department of Internal Medicine, Charles University in Prague, 3rd Faculty of Medicine, Prague,
Czech Republic
4
Department of Internal Medicine, Geriatrics and General Medicine, University Hospital Brno,
Czech Republic
5
Department of Pharmacology, Charles University in Prague, 3rd Faculty of Medicine, Prague,
Czech Republic
Abstract. In this work, we studied the association of
the E23K polymorphism of the Kir6.2 ATP-sensitive potassium channels in 212 Czech patients with
diabetes mellitus who were diagnosed after the age
of 35. Patients were classified into T1DM, LADA
and T2DM groups based on C-peptide and GADA
levels. Carriers of the predisposing Kir6.2 E23K K
allele showed no increased risk of either type of diabetes mellitus development. On the other hand, we
found a correlation between E23K SNP of the
KCNJ11 gene and C-peptide levels, which may be
considered a measure of pancreatic β-cell activity,
although this correlation was not statistically significant. In conclusion, we failed to confirm the Kir6.2
E23K as a genetic marker for T1DM, LADA and
T2DM in the Central Bohemian population of the
Czech Republic.
Introduction
Diabetes mellitus (DM) with manifestation after the
age of 35 is a very heterogeneous disease in which both
Received January 31, 2007. Accepted June 22, 2007.
This study was supported by the Research Programme of the
Ministry of Education, Youth and Sports of the Czech Republic
MSM 0021620814.
genetic and environmental factors contribute to its development. The inwardly rectifying K+channel (Kir) is
one of two subunits in the pancreatic β-cell ATP-sensitive potassium channel complex. This complex plays a
key role in glucose-stimulated insulin secretion, which
makes it a potential candidate for a genetic defect contributing to the development of type 2 DM (T2DM).
KCNJ11, a gene encoding Kir6.2 subunit, therefore belongs to well-established candidate genes of T2DM
(Porte and Kahn, 2001; Kahn, 2003). Furthermore, functional studies show that E23K SNP is involved in increasing the “open probability” of the Kir6.2 channel,
which should lead to diminished insulin secretion
(Schwanstecher et al., 2002), and provide supporting
evidence for KCNJ11 being a diabetogene. After several
initial reports with inconsistent data (Hani et al., 1998;
Yamada et al., 2001) the association between the E23K
polymorphism and the susceptibility to T2DM was recently confirmed (Gloyn et al., 2003).
The aim of this study was to investigate the Kir6.2
E23K gene polymorphism with regard to its possible
role in the development and progression of DM in patients with disease onset occuring after 35 years of age,
who were collected from the Central Bohemian population of the Czech Republic.
Corresponding author: Pavlína Čejková, Department of Cell and
Molecular Biology, 3rd Faculty of Medicine, Charles University,
Ruska 87, 100 34 Prague 10, Czech Republic. Phone: (+420) 261
102 540; Fax: (+420) 261 102 650; e-mail: pavlina.cejkova@lf3.
cuni.cz
Material and Methods
Abbreviations: CP – C-peptide, DM – diabetes mellitus, GADA
– glutamic acid decarboxylase antibodies, LADA – latent autoimmune diabetes in adults, PCR – polymerase chain reaction, RFLP
– restriction fragment length polymorphism, SNP – single nucleotide polymorphism, T1DM – type 1 diabetes mellitus, T2DM –
type 2 diabetes mellitus.
Two hundred and twelve diabetic patients (117 women and 95 men) from Central Bohemia were recruited,
who were at least 35 years old at the time of DM diagnosis. The diagnosis of diabetes was established according to the current WHO criteria including the pres-
Folia Biologica (Praha) 53, 173-175 (2007)
Subjects
174
Vol. 53
P. Čejková et al.
Table 1. Clinical features of diabetic patients genotyped for Kir6.2 gene E23K polymorphism
BMI [kg/m2]
Age at onset [years]
Duration of diabetes [years]
Mean C-peptide at onset [pmol/L]¹
Mean GADA [ng/mL]²
T1DM
N = 16
LADA
N = 24
T2DM
N = 172
28 (22.4-36.7)
43 (36-55)
16.7 (3-32)
66.2 (4-197.0)
223.8 (2.76-3000)
32.5 (26.6-45.5)
54 (35-71)
14 (4-29)
677 (277.8-1752.3)
375.5 (50.6-2800)
31.2 (18.4-50.1)
54 (35-81)
15.1 (1-65)
756.2 (202-3082)
7.69 (0.29-49.9)
All data shown as mean with min and max value; ¹C-peptide cut-off level: 200 pmol/l; ²GADA cut-off level: 50 ng/ml
ence of antibodies against glutamic acid decarboxylase
GAD65 (GADA) and serum C-peptide levels (CP) (Report, 2003). Patients were divided into three groups:
1) late onset T1DM which required insulin therapy
within 6 months after diagnosis: CP-, GADA-/+; 2) latent autoimmune diabetes in adults (LADA): CP+,
GADA+ and defined as having a minimum 6-month
long period between diagnosis and initiation of insulin
therapy; 3) T2DM: CP+, GADA- (Table 1). The control
group consisted of 113 healthy individuals randomly selected from volunteers at the Blood Donors Center. Genomic DNA from peripheral blood samples was extracted by QiaAmp DNA Mini Kit spin columns (QIAGEN
GmbH, Hilden, Germany).
original primers used in the Kir6.2 assays are available
from the authors on request.
PCR-RFLP
We did not detect any statistically significant difference (P = 0.05) between T2DM patients and the control
group. Similar results were obtained for both T1DM and
LADA patient groups (Table 2a): the frequency of the K
allele was 46.9% (Pcorr = NS) in T1DM, 29.2% (Pcorr =
NS) in LADA and 36.9% (Pcorr = NS) in T2DM patients
as compared to 36.7% in the control group. Likewise,
the genotype distribution did not differ between the
three groups and was consistent with Hardy-Weinberg
equilibrium.
Table 2b presents comparison of allele and genotype
frequencies of the KCNJ11 E23K polymorphism in all
diabetic patients classified based on C-peptide levels
(cut-off level 200 pmol/l). The K allele was increased in
CP- compared to CP+ diabetic patients (46.9% vs. 36%,
respectively), and similarly, the KK homozygous genotype was found more often in CP- (25%) than CP+
Genotyping was performed using the PCR-based restriction fragment length polymorphism (RFLP) method. Originally designed oligonucleotides ATPB and
ATPC amplified product of the size of 169 bp under the
following PCR conditions: initial denaturation at 94°C
for 3 min, then 40 times denaturation at 94°C for 15 s,
annealing at 56°C for 15 s and extension at 72°C for 15
s. To determine the presence of the E (glutamic acid) or
K (lysine) allele, a HinfI digestion was performed at
37°C for 1 h followed by heat inactivation. Within the
reverse primer ATPC, an artificial HinfI site was introduced and the resulting fragments were: undigested 169
bp (K/K homozygote), digested 145 + 24 bp (E/E homozygote). The digested and undigested products were
visualized using ethidium bromide staining after electrophoresis on 4% agarose gel. The sequences of the
Statistical evaluation
Epi Info software (Version 3.3, October 2004, Atlanta, Georgia) was used for statistical analysis. Statistical
differences in allele or genotype distributions were analysed by either the χ2 or Fisher’s exact probability test.
Significance was defined using a Bonferroni corrected P
value lower than 0.05. The strength of the observed associations was estimated by calculating the odds ratios
(OR).
Results
Table 2a. Frequency of alleles and genotypes of Kir6.2 E23K polymorphism in control subjects and patients with T1DM,
LADA and T2DM diagnosed after the of age 35
Genes
Kir6.2 E23K
(E/K)
E
K
EE
EK
KK
Alleles and genotypes
Control subjects
T1DM
P
LADA
P
T2DM
P
143 (63.3%)
83 (36.7%)
48 (42.5%)
47 (41.6%)
18 (15.9%)
NS
NS
NS
NS
NS
34 (70.8%)
14 (29.2%)
12 (50%)
10 (41.7%)
2 (8.3%)
NS
NS
NS
NS
NS
217 (63.1%)
127 (36.9%)
66 (38.4%)
85 (49.4%)
21 (12.2%)
NS
NS
NS
NS
NS
17 (53.1%)
15 (46.9%)
5 (31.2%)
7 (43.8%)
4 (25%)
Kir6.2 E23K Polymorphism and Diabetes Mellitus in Adults
Vol. 53
Table 2b. Frequency of alleles and genotypes of Kir6.2
E23K polymorphism between C-peptide-positive and C-peptide-negative diabetic patients
Locus
Kir6.2 E23K
(E/K)
E
K
EE
EK
KK
Alleles and genotypes
CP- (N = 16) CP+ (N = 196)
P
17 (53.1%)
15 (46.9%)
5 (31.2%)
7 (43.8%)
4 (25%)
NS
NS
NS
NS
NS
251 (64%)
141 (36%)
78 (39.8%)
95 (48.5%)
23 (11.7%)
(11.7%) patients. Nevertheless, these differences were
not statistically significant.
Discussion
Despite reports published elsewhere (Willer et al.,
2007), with regard to the allele and genotype frequencies of the Kir6.2 E23K polymorphism, we cannot confirm recently published data (Love-Gregory et al., 2003;
Florez et al., 2004; Willer et al., 2007) and the association between the genomic variations of Kir6.2 E23K
SNP and the development of common T2DM. Although
the E23K variant in KCNJ11 showed no association
with diabetes in Czech (for the K allele, OR 1.01 [95%
CI 0.71–1.43], P = NS), 95% CI around the OR overlaps in meta-analysis of European populations (Gloyn et
al., 2003; van Dam et al., 2005), suggesting that our results are not inconsistent with the previous studies.
Considering the physiological function of the ATPdependent K channel of pancreatic β cells, we tested to
see whether there was a correlation between C-peptide
levels and genotypes/alleles of the KCNJ11 E23K polymorphism. As shown in Table 2b, the K allele and KK
genotype were observed more frequently in patients
with low C-peptide levels. Although this finding did not
reach statistical significance, it suggests an association
of the K allele of E23K Kir6.2 polymorphism with decreased pancreatic β-cell secretory activity.
In conclusion, the presented study did not confirm
that the Kir6.2 E23K SNP is a genetic marker of diabetes mellitus in Czech diabetic patients and rather proposes its direct impact on the secretory activity of pancreatic β cells.
Aknowledgements
The authors would like to thank Mr. Thomas Secrest
for a professional language and style correction.
References
Florez, J. C., Burtt, N., de Bakker, P. I., Almgren, P., Tuomi, T.,
Holmkvist, J., Gaudet, D., Hudson, T. J., Schaffner, S. F.,
175
Daly, M. J., Hirschhorn, J. N., Groop, L., Altshuler, D. (2004)
Haplotype structure and genotype-phenotype correlations of
the sulfonylurea receptor and the islet ATP-sensitive potassium
channel gene region. Diabetes 53, 1360-1368.
Gloyn, A. L., Weedon, M. N., Owen, K. R., Turner, M. J.,
Knight, B. A., Hitman, G., Walker, M., Levy, J. C.,
Sampson, M., Halford, S., McCarthy, M. I., Hattersley,
A. T., Frayling, T. M. (2003) Large-scale association studies
of variants in genes encoding the pancreatic β-cell KATP
channel subunits Kir6.2 (KCNJ11) and SUR1 (ABCC8)
confirm that the KCNJ11 E23K variant is associated with
type 2 diabetes. Diabetes 52, 568-572.
Hani, E. H., Boutin, P., Durand, E., Inoue, H., Permutt, M.
A., Velho, G., Froguel, P. (1998) Missense mutations in
the pancreatic islet β-cell inwardly rectifying K+ channel
gene (KIR6.2/BIR): a meta-analysis suggests a role in the
polygenic basis of type II diabetes mellitus in Caucasians.
Diabetologia 41, 1511-1515.
Kahn, S. E. (2003) The relative contributions of insulin
resistance and β-cell dysfunction to the pathophysiology of
type 2 diabetes. Diabetologia 46, 3-19.
Love-Gregory, L., Wasson, J., Lin, J., Skolnick, G., Suarez,
B., Permutt, M. A. (2003) E23K Single nucleotide
polymorphism in the islet ATP-sensitive potassium channel
gene (Kir6.2) contributes as much to the risk of type II
diabetes in caucasians as the PPARγ Pro12Ala variant.
Diabetologia 46, 136-137.
Porte, D., Kahn, S. E. (2001) β-cell dysfunction and failure in
type 2 diabetes - potential mechanisms. Diabetes 50, S160S163.
Report of the Expert Committee on the Diagnosis and
Classification of Diabetes Mellitus (2003) Diabetes Care
26, S5-S20.
Schwanstecher, C., Meyer, U., Schwanstecher, M. (2002)
K(IR)6.2 polymorphism predisposes to type 2 diabetes by
inducing overactivity of pancreatic β-cell ATP-sensitive
K(+) channels. Diabetes 51, 875-879.
van Dam, R. M., Hoebee, B., Seidell, J. C., Schaap, M. M., de
Bruin, T. W. A., Feskens, E. J. M. (2005) Common variants
in the ATP-sensitive K+ channel genes KCNJ11 (Kir6.2) and
ABCC8 (SUR1) in relation to glucose intolerance: populationbased studies and meta-analyses. Diabet. Med. 22, 590-598.
Willer, C. J., Bonnycastle, L. L., Conneely, K. N., Duren,
W. L., Jackson, A. U., Scott, L. J., Narisu, N., Chines, P.
S., Skol, A., Stringham, H. M., Petrie, J., Erdos, M. R.,
Shift, A. J., Enloe, S. T., Sprau, A. G., Smith, E., Tong, M.,
Doheny, K. F., Pugh, E. W., Watanabe, R. M., Buchanan,
T. A., Valle, T. T., Bergman, R. N., Tuomilehto, J., Mohlke,
K. L. Collins, F. S., Boehnke, M. (2007) Screening of
134 single nucleotide polymorphisms (SNPs) previously
associated with type 2 diabetes replicates association with
12 SNPs in nine genes. Diabetes 56, 256-264
Yamada, Y., Kuroe, A., Li, Q., Someya, Y., Kubota, A., Ihara,
Y., Tsuura, Y., Seino, Y. (2001) Genomic variation in
pancreatic ion channel genes in Japanese type 2 diabetic
patients. Diabetes Metab. Res. Rev. 17, 213-216.
Original paper No. 4
AUTOIMMUNE DIABETES MELLITUS WITH ADULT ONSET AND TYPE 1 DIABETES
MELLITUS IN CHILDREN HAVE DIFFERENT GENETIC PREDISPOSITIONS
Cerna M, Kolostova K, Novota P, Romzova M, Cejkova P, Pinterova D, Pruhova S,
Treslova L, Andel M
Ann. N. Y.Acad. Sci 2007 Sep;1110:140-50
100
Autoimmune Diabetes Mellitus
with Adult Onset and Type 1 Diabetes
Mellitus in Children Have Different
Genetic Predispositions
MARIE CERNA,a KATARINA KOLOSTOVA,a PETER NOVOTA,a
MARIANNA ROMZOVA,a PAVLINA CEJKOVA,a DANIELA PINTEROVA,a
STEPANKA PRUHOVA,c LUDMILA TRESLOVA,b AND MICHAL ANDELb
a Department of Cellular and Molecular Biology, Centrum for Diabetes,
Metabolism and Nutrition Research of Third Faculty of Medicine, Charles
University in Prague, Ruska 87, Praha, Czech Republic
b Second Internal Clinics, Centrum for Diabetes, Metabolism and Nutrition
Research of Third Faculty of Medicine, Charles University in Prague, Ruska 87,
Praha, Czech Republic
c Department
of Pediatrics, Centrum for Diabetes, Metabolism and Nutrition
Research of Third Faculty of Medicine, Charles University in Prague, Ruska 87,
Praha, Czech Republic
ABSTRACT: Type 1 diabetes with manifestation after 35 years of age is
defined by CP <200 pmol/L and institution of insulin therapy within 6
months after diagnosis. Latent autoimmune diabetes mellitus in adults
(LADA) manifesting after 35 years of age is defined by minimum 6
months after diagnosis without insulin therapy and C peptide (CP) >200
pmol/L and antiGAD > 50 ng/mL. We aimed to find a possible genetic
discrimination among different types of autoimmune diabetes. To accomplish this goal, we analyzed DNA samples from 31 LADA patients,
75 patients with adult-onset type 1 diabetes mellitus, 188 type 1 diabetic
children, and 153 healthy adult individuals. We studied five genetic loci
on chromosomes 6, 11, 4, and 14: HLA DRB1 and DQB1 alleles, major histocompatibility complex (MHC) class I–related gene-A (MIC-A)
microsatellite polymorphism, interleukin (IL)-18 single nucleotide polymorphism, the microsatellite polymorphism of nuclear factor kappa B
gene (NF-B1), and the single nucleotide polymorphism of a gene for
its inhibitor (NF-BIA). HLA-DR3 was detected as the predisposition
allele for LADA (OR = 4.94, P < 0.0001). Further we found a statistically
significant increase of NF-BIA
AA genotype (OR = 2.68, ∗P < 0.01).
∗
On the other hand, DRB1 04, which is linked with DQB1 0302, was
Address for correspondence: Marie Cerna, Department of Cellular and Molecular Biology, Third
Faculty of Medicine, Charles University in Prague, Ruska 87, 100-00, Praha 10, Czech Republic. Voice:
+420-267-102-657; fax: +420-267-102-650.
[email protected]
C 2007 New York Academy of Sciences.
Ann. N.Y. Acad. Sci. 1110: 140–150 (2007). doi: 10.1196/annals.1423.016
140
CERNA et al.
141
observed as a risk factor in patients with type 1 diabetes mellitus (T1DM)
onset after 35 years of age (OR = 10.47, P < 0.0001 and OR = 9.49,
P < 0.0001, respectively). There was also an association with MIC-A5.1
(OR = 2.14, P < 0.01). Statistically significant difference was found in
the distribution of IL-18 promoter –607 (C/A) polymorphism between
LADA and T1DM in adults (P < 0.01). We conclude that all subgroups
of autoimmune diabetes have partly different immunogenetic predisposition.
KEYWORDS: T1DM; LADA; IDDM; HLA; autoantibodies
INTRODUCTION
Autoimmune diabetes mellitus, also called type 1 diabetes mellitus (T1DM),
is usually caused by autoimmune destruction of islet -cells in the pancreas.1
The destruction of islet -cells causes insulin deficiency and thereby a dysregulation in anabolism and catabolism. T1DM has features characteristic of most
of the autoimmune diseases: (1) It is associated with specific human leukocyte
antigens (HLA) class II haplotypes (particularly HLA DRB1∗ 04–DQB1∗ 0302,
but also HLA DRB1∗ 03–DQB1∗ 0201). (2) Autoantibodies against autoantigens of -cells circulating in the body are found (GAD65, IA-2, ICA, IAA). (3)
Mononuclear infiltration of the pancreatic islet can be detected histologically.
There are indications that -cell damage may be induced at any age and
that the timing of the clinical presentation is highly variable, with the youngest
patients diagnosed in infancy and the oldest at a later age.2 Epidemiological
data suggest that in 30–50% of cases clinical signs of T1DM may develop
after the age of 20 years. Such patients have many features of classic T1DM,
but previous studies indicate that subjects presenting with the disease in adulthood are characterized by a longer symptomatic period before diagnosis, better
preserved -cell function, and a reduced frequency of insulin autoantibodies
(IAAs). Genetic predisposition to T1DM is less marked (decreased prevalence
of HLA class II susceptibility alleles), and there is probably also a possible contribution of protective genes toward -cell destruction.3 Some studies
have suggested that later disease onset is characterized by a higher frequency
of DRB1∗ 03.4 Other studies found DRB1∗ 04 associated with later onset of
T1DM.5 It was observed that the prevalence of the HLA- DQB1∗ 0201/0302
(DR3/DR4) genotype reflects the age at onset of diabetes, and the 0302 genotype is associated with insulin dependency.
Aims of the Study
The aims of this study are to analyze the genetic predisposition for autoimmune diabetes mellitus with onset after 35 years of age to discern (a) whether it
is a heterogeneous group of diseases, in which we can further identify subtypes
142
ANNALS OF THE NEW YORK ACADEMY OF SCIENCES
of diabetes, or (b) whether diabetes with adult onset if it has a different immunogenetic predisposition compared to T1DM with childhood onset.
PATIENTS
Two hundred eighty-one patients suffering from diabetes mellitus with manifestation after 35 years of age were divided into three groups:
1. T1DM: CP ≤200 pmol/L. insulin <6 m. (N = 64)
2. Latent autoimmune diabetes in adults (LADA): CP ≥200 pmol/L. GADA
≥ 50 mg/mL insulin >6 m. (N = 31)
3. T2DM: CP ≥200 pmol/L, GADA ≤ 50 ng/mL (N = 131)
LADA is considered to be a subtype of T1DM. Other tested groups were
188 children withT1DM and 159 healthy adult individuals.
METHODS
We have analyzed five genetic loci localized on chromosomes 4, 6, 11, and
14 in patients with childhood- and adult-onset autoimmune diabetes compared
to healthy control subjects:6–11
Genes
•
•
•
•
•
Alleles
HLA class II
MIC-A (MHC class I-related gene-A)
NF-B1 (nuclear factor- B)
NF-BIA (inhibitor I-B)
IL-18 (interleukin-18)
DRB1 and DQB1
A microsatellite polymorphism GCT in exon 5
CA repeating polymorphism in regulation region
A single nucleotide polymorphism A/G in 3 UTR
A promoter polymorphism –137 G/C, –607 C/A
PREDISPOSITION GENES AND THEIR PRODUCTS
HLA Class II Alleles
An association between HLA and autoimmune diabetes mellitus has been
recognized for more than 20 years. Many studies have verified that DQB1∗ 0302
is a strong susceptibility gene and that the heterozygous combination of
DQA1∗ 0301-DQB1∗ 0302 on the HLA-DR4 haplotype and DQA1∗ 0501DQB1∗ 0201 on the HLA-DR3 haplotype results in a synergistically increased
risk of T1DM. DQA1∗ 0301-DQB1∗ 0302 is the most prevalent risk-conferring
haplotype in Caucasians, detected in 74% of T1DM patients followed by
CERNA et al.
143
DQA1∗ 0501-DQB1∗ 0201, detected in 52% of Caucasian patients.12 Also,
there is agreement that the DQB1∗ 0602 allele, linked with DR2, is a strong protective gene.13 The protective effect of DQB1∗ 0602 is dominant to the susceptibility effects of DQB1∗ 0302-DQA1∗ 0301 and DQB1∗ 0201-DQA1∗ 0501.14
A single copy of this DQB1∗ 0602 allele is adequate to confer significant
negative association. HLA-DQ genes are of primary importance, but HLADR genes modify the risk conferred by HLA-DQ.15 The risk associated with
an HLA genotype is defined by the particular combination of susceptible
and protective alleles. The frequencies of predisposition alleles, DQB1∗ 0302
and DQB1∗ 0201, are usually increased, whereas frequency of the protective
DQB1∗ 0602 allele is usually decreased.16,17
Major Histocompatibility Complex (MHC) Class I
Chain-Related Gene A (MIC-A)
A number of observations indicate that class II genes cannot explain all of
the MHC associations with T1DM. A role for MHC complex genes other than
class II genes has been suggested. Several studies have supported a role for
both MHC HLA class III and class I genes in T1DM predisposition. The most
studied class III genes are tumor necrosis factor and lymphotoxin (TNF- and
TNF-). The strongest evidence for susceptibility genes in the class I region
comes from recent systematic assessment of microsatellite markers spanning
this region. Among them, one is the gene MIC-A, which is located within the
MHC class I region of chromosome 6. It contains long open reading frames
encoding three distinct extracellular domains (1, 2, and 3), a transmembrane segment, and a cytoplasmatic tail. It has been reported that MIC-A may
be recognized by a subpopulation of intestinal T cells,18 and may play a
role in the activation of a subpopulation of NK cells that express the NKG2D
receptor.19
Sequence analysis of the exon 5 encoding the transmembrane segment
showed a trinucleotide repeat (GCT) microsatellite polymorphism within this
region.20 So far, six alleles of the exon 5 of the MIC-A gene, which differ
in number of (GCT) repetitions at position 296, have been identified. These
alleles contain 4, 5, 6, 9, 10 repetitions of GCT, and five repetitions of GCT
with an additional nucleotide insertion (GGCT). The MIC-A alleles have been
accordingly named A4, A5, A6, A9, A10, and A5.1, respectively.21 Because
of the proximity of the MIC-A locus to HLA-B and HLA-C, MIC-A polymorphisms have been investigated for associations with a variety of HLA class
I-associated diseases. Some of these studies show a positive association between MIC-A and the disease in question, but this may be due to its linkage disequilibrium with HLA-B. However, most disease association studies have only
concentrated on the MIC-A transmembrane region, focusing on the six variants
144
ANNALS OF THE NEW YORK ACADEMY OF SCIENCES
so far identified. Recent studies of MIC-A transmembrane polymorphism focused on the investigation of its relationship to autoimmune diabetes mellitus
showed an association of the A5 allele with the susceptibility to young-onset
T1DM.22,23 Conversely, this allele does not confer increased risk for adultonset T1DM.24
NF- B and Its Inhibitor
The gene encoding NF-B, NF-B1, on chromosome 4 with a microsatellite polymorphism (CA) inregulation region (18 alleles), and the gene encoding inhibitor of NF-B, NF-BIA, on chromosome 14 with single nucleotide
polymorphism (A/G) in 3 UTR, are other candidates of genetic predisposition to diabetes mellitus for the reason of their relationship to inflammatory
processes.25,26 NF-B is a transcription factor, which is implicated in the regulation of many genes that encode mediators of immune response, embryo and
cell lineage development, cell apoptosis, inflammation, cell cycle progression,
oncogenesis, viral replication, and various autoimmune diseases. The activation of NF-B is thought to be part of a stress response as it is activated by
a variety of stimuli, including reactive oxygen species (ROS) and advanced
glycation end products (AGEs), which are toxic products of nonenzymatic
glycation due to long-term hyperglycemia and oxidative stress. At the cellular
level NF-B is activated through phosphorylation of inhibitor of NF-B (IB),
what triggers subsequent translocation of NF-B molecules into the nucleus,
where they are bound with a consensus sequence (5 GGGACTTTCC-3 ) of
various genes, activating their transcription.
Interleukin (IL)-18
On chromosome 11 is located the gene encoding IL-18, with two single
nucleotide polymorphisms inpromoter region at position –607 and –137. The
cytokine IL-18, the Th1 cytokine, and a member of the IL-1 superfamily,
elicits several biological activities that initiate and promote host defense and
inflammation following infection or injury.27,28 It is a pleiotropic cytokine
predominantly secreted by activated monocytes/macrophages involved in the
regulation of innate and acquired immune response, playing a key role in
autoimmune, inflammatory, and infectious diseases. IL-18 acts as a proinflammatory factor and, in synergy with IL-12, promotes development of Th1
lymphocyte response by induction of interferon- (IFN- ) production, modulates activity of natural killer cells, increases TNF- and IL-1 production by
macrophages, upregulates the expression of adhesion molecules, and induces
nitric oxide production in the area of chronic inflammation.29 It was reported
by Nicoletti et al.30 that IL-18 serum levels are increased in the subclinical
stage of T1DM in first-degree relatives of T1DM patients. The importance
CERNA et al.
145
of IL-18 for the occurrence of T1DM is underscored by the observation that
the murine IL-18 gene maps to an interval on chromosome 9 in the vicinity
of the diabetes susceptibility gene from the nonobese diabetic (NOD) mouse,
idd2.31 Several polymorphisms in the promoter region of human IL-18 gene,
located at chromosome 11q22.2–q22.3, have been identified.32 Kretowski et
al.33 reported the role of IL-18 gene polymorphism at positions –607 and –137
in the predisposition to T1DM in Polish population.
RESULTS
The most significant risk factor is on chromosome 6, human leukocyte antigen class II alleles, HLA DRB1 and DQB1. Odds ratio for single predisposition
HLA alleles has been significantly different in different groups of patients. For
DRB1∗ 03 allele, it has been the highest in LADA patients (OR = 4.94 with
CI 95% between 1.83 and 13.49), for DRB1∗ 04 and DQB1∗ 0302 alleles, it
has been the highest in diabetic children (OR = 22.87 with CI 95% between
10.89 and 48.87, and OR = 46.48 with CI 95% between 16.15 and 140.23,
respectively). In the case of T1DM with adult onset, predisposition alleles
have been similar as in children, but odds ratio has been lower and proved a
stronger influence of DR alleles than DQ alleles (OR = 10.47 with CI 95%
between 4.53 and 24.65, and OR = 9.49 with CI 95% between 3.91 and 23.59,
respectively).
Analysis of gene expression of HLA-DRB1∗ 04 in antigen-presenting cells
of peripheral blood has described significantly increased quantity of mRNA of
this allele in T1DM with adult onset in comparison to T1DM in children and to
control groups of adults and children (FIG. 1). One HLA hypothesis points to a
possible role of HLA-DR molecules as a protective factor, which in comparison
to HLA-DQ molecules suppresses autoimmune reactions. Higher quantity of
DR molecules in adult patients may suggest breaking the autoimmune process.
On chromosome 6 is also located the MIC-A gene. Its exon 5, coding transmembrane region, has been tested for a microsatellite polymorphism (GCT)
with six alleles. The results in this study suggest that MIC-A5.1 allele is associated with genetic susceptibility to adult onset of T1DM (corrected P =
0.009; OR = 2.14 with CI 95% 1.29–3.55), whereas it is increased, but not
significantly in LADA. A different allele, MIC-A5, is associated with T1DM
in children.
Since NF-B is a transcription factor which is implicated in the regulation
of many genes involved in inflammatory processes, a microsatellite polymorphism (CA) in the regulation region of NF-B1 gene and a single nucleotide
polymorphism (A/G) in 3 UTR of the NF-BIA gene are other candidates of
genetic predisposition to autoimmune diabetes mellitus. A statistically significant difference in allele frequencies of the NF-B1 gene has been observed
only in T1DM with adult onset. The frequency of A7 allele (size 132 bp, 20
146
ANNALS OF THE NEW YORK ACADEMY OF SCIENCES
4.5
Relative RNA amount
4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
T1DM Children
T1DM Adults
Controls
FIGURE 1. Relative quantification data from quantitative analysis of HLA-DRB1∗ 04
mRNA levels in the tested groups of patients and controls.The expression of HLA-DRB1∗ 04
was found significantly higher in T1DM adult patients. The comparison was done with two
specific control groups (adult and child controls with HLA-DRB1∗ 04 allele). The relative
RNA amount for tested groups of type 1 diabetes (T1DM) with adult onset and in childhood
was significantly different (P-value 0.034; comparison by Kruskall–Wallis test and Dunn’s
multiple comparison test). The difference between patient groups and control groups was
not significant by this type of testing.
repeats) has been significantly increased in comparison to that of a control
healthy group (P < 0.01; OR = 10.0 with CI 95% 1.25–215.6).
There was no difference in A and G allele frequencies of the NF-BIA gene
between the control group and patients; the difference has been observed in
distribution of genotypes. There is evidence that the heterozygous genotype, in
comparison to a control healthy group, is protective for autoimmune diabetes
with onset in adulthood: 55.6% in controls, 45.2% in LADA (nonsignificant
decrease), and 35.9% in T1DM with adult onset (P < 0.01; OR = 0.4 with
CI 95% 0.24–0.86). A statistically significant increase of the AA genotype
frequency of the NF-BIA gene has been detected in autoimmune diabetes
type LADA (32.3%, P < 0.01; OR = 2.7 with CI 95% 1.06–6.75) and in
type 2 diabetes (42.3%, P < 0.001; OR = 2.6 with CI 95% 1.52–4.41) in
comparison to a control healthy group (16.9%). In T1DM with adult onset, the
genotype frequency has been also increased, but it has not reached a statistical
significance. On the other hand, the values for diabetes in children are not
different from control ones. The fact that polymorphism of the NF-B system
CERNA et al.
147
TABLE 1. Overview of predisposition genes in autoimmune diabetes with different onset
Gene
LADA
after 35 years of age
HLA
DRB1∗ 03
MIC-A
NF-BIA
NF-B1
IL-18
(OR = 5)
Allele A5.1a
Homozygote AA
−
Homozygote −607 CC ↑
(increased frequency)
a Observation
T1DM
after 35 years of age
DRB1∗ 04
DQB1∗ 0302
(OR = 10)
Allele A5.1
Homozygote AAa
Allele A7
Homozygote −607 CC ↓
(decreased frequency)
T1DM
in children
DQB1∗ 0302
(OR = 46)
Allele A5
−
−
Allele −137 C
was not statistically significant.
is associated not only with autoimmunity, but also with metabolic disease,
documents its central role in cell signaling.
Promoter polymorphism of IL-18 gene is not associated with any group
of diabetes, but a statistical significant difference has been found between
autoimmune diabetes type LADA and adult-onset T1DM.
CONCLUSIONS
We can summarize our observations in these findings:
The main and unique HLA predisposition allele in autoimmune diabetes
type LADA has been detected as HLA-DR3 (OR = 5.0, P < 0.0001). Other
HLA alleles have not been found to have a significant association with the
disease. Furthermore, the statistically significant increase of the AA genotype
frequency of the NF-BIA gene (OR = 2.7, P < 0.01) has been present in
comparison to a control healthy group. Contrary to this, the genetic risk factor
in adult-onset T1DM has been detected as the DRB1∗ 04 allele (OR = 10.0,
P < 0.0001), which is in linkage disequilibrium with other predisposition
allele, DQB1∗ 0302 (OR = 10.0, P < 0.0001). This group is also associated
with microsatellite polymorphisms of MIC-A and NF-B1 genes (MIC-A5.1
allele: OR = 2.14, P < 0.01; NF-B1-A7 allele: OR = 10.0, P < 0.01).
The statistically significant difference has been observed in the distribution
of the promoter polymorphism at position –607 (C/A) of IL-18 gene between
both groups of patients (P < 0.01), which supports our conclusions about
partly different etiopathogenesis of these two groups of adult patients (TABLE 1).
Taken together, these observations suggest a generally accepted conclusion
that the presence of predisposing genes in autoimmune diabetes decreases with
age, probably because of the increasing influence of environmental factors.
Autoimmune diabetes with manifestation in adults (after 35 years ) may have
148
ANNALS OF THE NEW YORK ACADEMY OF SCIENCES
a partly different immunogenetic etiopathogenesis than autoimmune diabetes
with manifestation in childhood. Allele polymorphism of the NF-B system
presents a genetic risk factor only in autoimmune diabetes with adult onset,
but not in diabetes of childhood.
Furthermore, we can hypothesize that T1DM in children or adults might
have different predisposition genes than those of LADA. Compared to
fast-progressing adult-onset T1DM, slowly progressing adult-onset T1DM
(LADA) can involve protective genes leading to slow, progressive -cell destruction and insulin deficiency. The different mechanisms reflect different
HLA association in these diseases. Allele polymorphism of the HLA system
is associated with autoimmune diabetes in both categories, but for each group
of patients the specific association has been described. The most significant
association is with DQB1 alleles in T1DM, but with DRB1 alleles in LADA.
ACKNOWLEDGMENTS
This survey was funded by the Research Design of the Ministry of Education
and Youth of the Czech Republic (Identification code: MSM 0021620814: Prevention, diagnostics and therapy of diabetes mellitus, metabolic and endocrine
damage of organism).
REFERENCES
1. AMERICAN DIABETES ASSOCIATION. 2006. Diagnosis and classification of diabetes
mellitus. Diabetes Care 29 Suppl 1: S43–S48.
2. SABBAH, E., K. SAVOLA, T. EBELING, et al. 2000. Genetic, autoimmune, and clinical
characteristics of childhood- and adult-onset type 1 diabetes. Diabetes Care 23:
1326–1332.
3. POZZILLI, P. & U. DI MARIO. 2001. Autoimmune diabetes not requiring insulin at
diagnosis (latent autoimmune diabetes of the adult). Diabetes Care 24: 1460–
1467.
4. LOHMANN, T., J. SESSLER, H.J. VERLOHREN, et al. 1997. Distinct genetic and immunological features in patients with onset of IDDM before and after age 40.
Diabetes Care 20: 524–529.
5. TUOMI, T., A. CARLSSON, H. LI, et al. 1999. Clinical and genetic characteristics of
type 2 diabetes with and without GAD antibodies. Diabetes 48: 150–157.
6. CERNA, M., M.A. FERNANDEZ-VINA, E. IVASKOVA, et al. 1992. Comparison of HLA
class II alleles in Gypsy and Czech populations by DNA typing with oligonucleotide probes. Tissue Antigens 39: 111–116.
7. CERNA, M., P. NOVOTA, K. KOLOSTOVA, et al. 2003. HLA in Czech adult patients
with autoimmune diabetes mellitus: comparison with Czech children with type
1 diabetes and patients with type 2 diabetes. Eur. J. Immunogenet. 30: 401–407.
8. NOVOTA, P., M. CERNA, K. KOLOSTOVA, et al. 2004. Diabetes mellitus in adults:
association of HLA DRB1 and DQB1 diabetes risk alleles with GADab presence
and C-peptide secretion. Immunol. Lett. 95: 229–232.
CERNA et al.
149
9. NOVOTA, P., L. KOLESÁR, A. SLAVČEV, et al. 2004. Fluorescence-based automated
fragment analysis of microsatellite polymorphism within the transmembrane
region of the MIC-A gene. Folia Biol. (Praha) 50: 21–23.
10. NOVOTA, P., K. KOLOSTOVA, D. PINTEROVA, et al. 2005. Association of MHC class
I chain related gene-A microsatellite polymorphism with the susceptibility to
T1DM and LADA in Czech adult patients. Int. J. Immunogenet. 32: 273–275.
11. NOVOTA, P., K. KOLOSTOVA, D. PINTEROVA, et al. Interleukin IL-18 gene promoter
polymorphisms in adult patients with type 1 diabetes mellitus and latent autoimmune diabetes in adults. Immunol. Lett. 96: 247–251.
12. SANJEEVI, C.B., M. LANDIN-OLSSON, I. KOCKUM, et al. 1995. Several residues in
the DQA1 and DQB1 chain are either positively or negatively associated with
insulin-dependent diabetes mellitus. Diabetes 44: S1–S241.
13. KOCKUM, I., R. WASSMUTH, E. HOLMBERG, et al. 1993. HLA-DQ primarily confers
protection and HLA-DR susceptibility in type I (insulin-dependent) diabetes
studied in population-based affected families and controls. Am. J. Hum. Genet.
53: 150–167.
14. PUGLIESE, A., R. GIANANI, R. MOROMISATO, et al. HLA-DQB1∗ 0602 is associated
with dominant protection from diabetes even among islet cell antibody-positive
first-degree relatives of patients with IDDM. Diabetes 44: 608–613.
15. SANJEEVI, C.B., A. FALORNI, I. KOCKUM, et al. 1996. HLA and glutamic acid
decarboxylase in human insulin-dependent diabetes mellitus. Diabet. Med. 13:
209–217.
16. RONINGEN, K.S., A. SPURKLAND, B.D. TAIT, et al. 1991. HLA class II association in
insulin dependent diabetes mellitus among Blacks, Caucasoids and Japanese. In
HLA 1991—Proceedings of the 11th International Histocompatibility workshop
and conference. Oxford University Press.
17. SANJEEVI, C.B., T.P. LYBRAND, M. LANDIN-OLSSON, et al. 1994. Analysis of antibody markers, DRB1, DRB5, DQA1 and DQB1 genes and modeling of DR2
molecules in DR2-positive patients with insulin-dependent diabetes mellitus.
Tissue Antigens 44: 110–119.
18. GROH, V., A. STEINLE, S. BAUER, et al. 1998. Recognition of stress-induced MHC
molecules by intestinal epithelial gammadelta T cells. Science 279: 1737–1740.
19. BAUER, S., V. GROH, J. WU, et al. 1999. Activation of NK cells and T cells by
NKG2D, a receptor for stress-inducible MICA. Science 285: 727–729.
20. FODIL, N., L. LALOUX, V. WANNER, et al. 1996. Allelic repertoire of the human
MHC class I MICA gene. Immunogenetics 44: 351–357.
21. PEREZ-RODRIGUEZ, M., A. CORELL, J.R. ARGUELLO, et al. 2000. A new MICA
allele with ten alanine residues in the exon 5 microsatellite. Tissue Antigens 55:
162–165.
22. GAMBELUNGHE, G., M. GHADERI, A. COSENTINO, et al. 2000. Association of MHC
class I chain-related A (MIC-A) gene polymorphism with type I diabetes. Diabetologia 43: 507–514.
23. GUPTA, M., L. NIKITINA-ZAKE, M. ZARGHAMI, et al. 2003. Association between
the transmembrane region polymorphism of MHC class I chain related gene-A
and type 1 diabetes mellitus in Sweden. Hum. Immunol. 64: 553–561.
24. GAMBELUNGHE, G., M. GHADERI, C. TORTOIOLI, et al. 2001. Two distinct MICA
gene markers discriminate major autoimmune diabetes types. J. Clin. Endocrinol.
Metab. 86: 3754–3760.
25. HEGAZY, D.M., D.A. O’REILLY, B.M. YANG, et al. 2001. NFkappaB polymorphisms
and susceptibility to type 1 diabetes. Genes Immun. 2: 304–308.
150
ANNALS OF THE NEW YORK ACADEMY OF SCIENCES
26. GYLVIN, T., R. BERGHOLDT, J. NERUP, et al. 2002. Characterization of a nuclearfactor-kappa B (NFkappaB) genetic marker in type 1 diabetes (T1DM) families.
Genes Immun. 3: 430–432.
27. DINARELLO, C.A. 1999. IL-18: A TH1-inducing, proinflammatory cytokine and
new member of the IL-1 family. J. Allergy Clin. Immunol. 103: 11–24.
28. NAKANISHI, K., T. YOSHIMOTO, H. TSUTSUI, et al. 2001. Interleukin-18 is a unique
cytokine that stimulates both Th1 and Th2 responses depending on its cytokine
milieu. Cytokine Growth Factor Rev. 12: 53–72.
29. NAKAHIRA, M., H.J. AHN, W.R. PARK, et al. 2002. Synergy of IL-12 and IL-18 for
IFN-gamma gene expression: IL-12-induced STAT4 contributes to IFN-gamma
promoter activation by up-regulating the binding activity of IL-18-induced activator protein 1. J. Immunol. 168: 1146–1153.
30. NICOLETTI, F., I. CONGET, R. DI MARCO, et al. 2001. Serum levels of the interferongamma-inducing cytokine interleukin-18 are increased in individuals at high risk
of developing type I diabetes. Diabetologia 44: 309–311.
31. ROTHE, H., N.A. JENKINS, N.G. COPELAND, et al. 1997. Active stage of autoimmune
diabetes is associated with the expression of a novel cytokine, IGIF, which is
located near Idd2. J. Clin. Invest. 99: 469–474.
32. GIEDRAITIS, V., B. HE, W.X. HUANG, et al. 2001. Cloning and mutation analysis
of the human IL-18 promoter: a possible role of polymorphisms in expression
regulation. J. Neuroimmunol. 112: 146–152.
33. KRETOWSKI, A., K. MIRONCZUK, A. KARPINSKA, et al. 2002. Interleukin-18 promoter polymorphisms in type 1 diabetes. Diabetes 51: 3347–3349.
Original paper No. 5
HLA IN CZECH ADULT PATIENTS WITH AUTOIMMUNE DIABETES MELLITUS:
COMPARISON WITH CZECH CHILDREN WITH TYPE 1 DIABETES AND PATIENTS
WITH TYPE 2 DIABETES
Cerna M, Novota P, Kolostova K, Cejkova P, Zdarsky E, Novakova D, Kucera P,
Novak J, Andel M
Eur J Immunogenet 2003 30:401-407
111
HLA in Czech adult patients with autoimmune diabetes mellitus:
comparison with Czech children with type 1 diabetes and patients
with type 2 diabetes
Blackwell Publishing Ltd.
M. Cerna,* P. Novota,* K. Kolostova,* P. Cejkova,* E. Zdarsky,* D. Novakova,† P. Kucera,†
J. Novak‡ & M. Andel‡
Summary
Type 1 diabetes results from an autoimmune insulitis,
associated with HLA class II alleles. The evidence about
HLA allele association is not clear in patients diagnosed
after 35 years of age. In this study we have analyzed HLA
alleles of DQB1 and DRB1 genes by sequence specific
primer (SSP)-PCR technique in adult patients with disease
onset after 35 years of age. Two hundred and eighty-one
patients were divided into three groups according to
the insulin therapy, the level of C peptide (CP), and GAD
antibodies (anti-GAD). Group 1 (type 1 diabetes in adults)
was characterized by CP less than 200 pmol / L and antiGAD more or less than 50 ng /mL (n = 80). All of them had
insulin therapy within 6 months after diagnosis. Group 2
latent autoimmune diabetes mellitus in adults (LADA)
was defined by a minimum 6-month-long phase after
diagnosis without insulin therapy, and was characterized
by CP more than 200 pmol / L and anti-GAD more than
50 ng/mL (n = 70). Group 3 (type 2 diabetes) was characterized by CP more than 200 pmol / L and anti-GAD less
than 50 ng/mL (n = 131). None ever had insulin therapy.
In group 1, there was increased frequency of DRB1*04
(45.0% vs. controls 14.1%, OR = 5.0, P < 0.0005) and
DQB1*0302 alleles (43.3% vs. controls 11.1%, OR = 6.1,
P < 0.00005). There was increased frequency of
DRB1*03 and DQB1*0201, and decreased frequency of
DQB1*0602 (3.3% vs. controls 20.2%), but it was not
significant. In group 2, there was a significantly increased
frequency of DRB1*03 only (50.0% vs. controls 21.2%,
OR = 3.7, P < 0.05). Compared with children with type 1
diabetes and adults with type 2 diabetes (group 3), we
* Department of Cell and Molecular Biology, † Department of Cell and
Molecular Immunology, ‡ Diabetes Center and Department of
Medicine II, Third Medical Faculty of Charles University, Prague,
Czech Republic
Received 22 October 2001; revised 21 July 2003;
accepted 15 September 2003
Correspondence: Marie Cerna, Department of Cell and Molecular
Biology, Third Medical Faculty of Charles University, Prague, Ruská 87,
100 34 Praha 10, Czech Republic. Tel.: + 420 2 67 102667;
Fax: + 420 2 67 102650; E-mail: [email protected]
conclude that the presence of predisposing DQB1 alleles
in adults with type 1 diabetes decreases with the age,
probably due to environmental factors. Only the
DRB1*03, but not the DQB1 gene, becomes the main
predisposing allele in LADA patients. These findings
suggest that the presence of HLA-DQB1*0302 identifies
patients at high risk of requiring insulin treatment. Type 1
diabetes mellitus (DM) in children or adults may have
partly different immunogenetic etiopathogenesis than LADA.
Introduction
Autoimmune diseases form a complex group of approximately 40 diseases affecting 5–7% of the general
population. Type 1 diabetes mellitus (DM) is one of the
most frequent and potentially severe with a prevalence of
approximately 0.5% (Sinha et al., 1990).
Type 1 diabetes mellitus develops as a result of selective
autoimmune destruction of pancreatic β-cells, a process
called insulitis. The presence of antibodies against isletcells (ICA), insulin (IAA), glutamic acid decarboxylase
(GAD-Ab) and tyrosine-phosphatase (IA2-Ab), and
infiltration of pancreatic islands by B and T lymphocytes
are all evidence of organ specific autoimmunity.
The exact cause of insulitis is not completely understood.
The etiology is multifactorial and includes both genetic
and environmental factors. Many papers have shown that
HLA genes, the human main histocompatibility locus on
the chromosome 6, are key genetic factors (Thomson
et al., 1988). The studies have proved that type 1 diabetes
is associated with certain HLA class II alleles, namely
DRB1*0301, DRB1*0401, DRB1*0405, DQB1*0201,
DQB1*0302, DQA1*0301. Some alleles are also protective, namely DQB1*0602, DQB1*0603, DQB1*0301.
The HLA frequencies in Czech population are similar
as those of the other Caucasian populations (Cerna et al.,
1992; Drabek et al., 1998; Loudova et al., 1999). Serological typing of the HLA system in young patients with
type 1 diabetes in the Czech population has also proved
HLA predisposition antigens similar to other Caucasians
(Dvorakova & Majsky, 1979, Vondra et al., 1996).
The association of HLA alleles with type 1 diabetes
changes with age of onset. Generally, the younger the
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401–407
401
402
M. Cerna et al.
patients the more significant is the HLA association (Tait
et al., 1995; Lohmann et al., 1997; Graham et al., 1999).
The evidence about HLA allele association is not clear,
however. Some studies have suggested that later disease
onset is characterized by a higher frequency of DRB1*03
(Lohmann et al., 1997; Zevaco-Mattei et al., 1999;
Gambelunghe et al., 2000). Other studies found DRB1*04
associated with later onset of type 1 diabetes mellitus (Tait
et al., 1995; Tuomi et al., 1999). Analyses in non-diabetic
offspring of probands with latent autoimmune diabetes
mellitus in adults (LADA) concluded that LADA is a
familial disease involving gene defects leading to a slow
progressive β-cells destruction and insulin deficiency
(Vauhkonen et al., 2000).
The aims of our study were to analyze the frequency
distribution of HLA DRB1 and DQB1 alleles in the adult
onset diabetes in a Czech population and compare this
data with those from controls, diabetic children and the
healthy Czech population.
Patients and methods
Patients
We investigated 281 patients (172 women) with diagnosis
of diabetes after 35 years of age. Their average age at
disease onset was 53 years (range 35–81). The patients
came consecutively for consultation to our diabetes centre
and collaborating diabetes clinics. All patients had fasting
C peptide, GAD and IA2 antibodies measured at the
time of investigation. Body mass index (BMI) was
calculated from weight and height. The therapy was
recorded (Table 1).
We divided our patients into three groups: type 1
diabetes in adults, LADA, and type 2 diabetes. Patients
with autoimmune diabetes (group 1 and 2) were negative
for C peptide or positive for autoantibodies. All of them
had insulin therapy. LADA (group 2) was defined by a
minimum 6-month-long phase after diagnosis without
insulin therapy. Patients with type 2 diabetes (group 3)
were positive for C peptide and negative for autoantibodies.
None ever had insulin therapy.
Group 1, type 1 diabetes in adults, comprised 80 patients (56 women) with an average BMI of 27 (SD = 4.6)
and average duration of diabetes 16 years (SD = 14.2).
All patients had fasting C peptide secretion less than
200 pmol/L. Anti-GAD more than 50 ng/mL was present
in 50% of patients, anti-IA2 more than 0.9 U/mL was
present in 15% of patients. Ten per cent of patients
were positive for both antibodies, but 45% of patients
were without these two antibodies.
Group 2, LADA, comprised 70 patients (40 women)
with an average BMI of 32 (SD = 4.8) and average
duration of diabetes 14 years (SD = 7.0). All patients had
fasting C peptide secretion more than 200 pmol/L.
Anti-GAD more than 50 ng/mL was present in 100% of
patients and anti-IA2 more than 0.9 U/mL was present in
11% of patients.
Type 2 diabetic patients (group 3) comprised 131 persons
(76 women) with an average BMI of 31 (SD = 5.5)
and average duration of diabetes 13 years (SD = 6.6),
with fasting C peptide secretion more than 200 pmol / L
and anti-GAD less than 50 ng /mL. None ever had insulin
therapy.
For comparison we used the results of HLA class II
analysis of Czech diabetic children with type 1 diabetes
mellitus (Cinek et al., 2001). HLA typing performed in
children was done by the same technique as in this study
(SSP-PCR supplied by Genovision, Oslo, Norway).
We have compared the patient data with the data of the
healthy Czech population, which was published earlier
(Cerna et al., 1992). The controls were genotyped in
a previous study by oligonucleotide hybridization. We
tested several control samples by the sequence specific
primer (SSP)-PCR method used in this study to prove that
we get the same results by both techniques. The study was
approved by the ethical committee of Third Medical
Faculty of Charles University at Prague. All patients gave
Features
Group 1
(n = 80)
Group 2
(n = 70)
Group 3
(n = 131)
Female
Age at disease onset (years)a
Duration of DM (years)a
BMI (kg/m2)a
Insulin therapy
(number)
Fasting C peptideb
(pmol/L)a
Anti-GADc
(ng/mL)a
61%
43 (36 –56)
16 (4 –27)
27 (22–37)
Yes
80/80
–
63 (4–197)
+/–
193 (3–3000)
57%
52 (35 –71)
14 (4 –29)
32 (27 –46)
Yes
70/70
+
609 (51–2800)
+
379 (210 –1753)
58%
53 (35 –81)
13 (1–22)
31 (18 –50)
No
0/131
+
772 (1–50)
–
8 (202–3370)
Table 1. Clinical features in 281 Czech adult
patients with diabetes
a
Average value with ranges in parentheses;
+ positive value ≥ 200 pmol / L, – negative value < 200 pmol /L;
c
+ positive value ≥ 50 ng /mL, – negative value < 50 ng /mL;
Group 1 (type 1 diabetes in adults): C peptide < 200 pmol /L, anti-GAD > or < 50 ng /mL;
Group 2 (LADA): C peptide > 200 pmol / L, anti-GAD > 50 ng /m, > 6 months without insulin;
Group 3 (type 2 diabetes): C peptide > 200 pmol / L, anti-GAD < 50 ng /mL.
b
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401– 407
HLA in Czech patients with diabetes mellitus
informed consent to the use of their blood samples in
this study.
403
Analysis of HLA DRB1 and DQB1 alleles
HLA DRB1 and DQB1 genes were typed using PCR with
sequence specific primers (SSP-PCR) (Zetterquist &
Olerup, 1992; Olerup et al., 1993) supplied by Genovision.
Fasting C peptide
We used an immunoradiometric method (Immunotech,
Prague, Czech Republic). The serum samples from 49
healthy individuals were analyzed and the results are as
follows: average value 486 pmol / L, median 449 pmol/L,
standard deviation 170 pmol / L, and minimal-maximal
value 206–934 pmol / L. The level less than 200 pmol/L
has been taken as decreased C peptide secretion.
GAD antibodies
The presence of IgG antibodies against GAD has been
detected by the ELISA method (Roche Molecular
Biochemicals, Mannheim, Germany). The test has 98% of
specificity and 69% of sensitivity (Verge et al., 1998). A level
less than 32 ng/mL is considered as negative. To increase
the precision, in our study, a level less than 50 ng /mL has
been taken as normal. A level more than 50 ng /mL has
been considered as a reliable sign of autoimmune insulitits.
IA2 antibodies
The presence of IgG antibodies against IA2 has been
detected by the ELISA method (Roche Molecular Biochemicals). A level less than 0.9 U/mL is considered as
negative.
Statistics
Allele, gene, and haplotype frequencies were calculated
using standard methods. The significance of the differences in allele frequencies were evaluated by χ2 with Yates’
correction and P-values were calculated. When a hypothesis
that had not been tested was analyzed, the P-values were
corrected by multiplying by the number of alleles tested.
The strength of the observed associations was estimated
by calculating odds ratios [relative risk (RR)] using the
method by Woolf. When one of the RR values is 0 or close
to 0, RR can be calculated by using Woolf’s method with
Haldane’s correction.
Results
Analysis of HLA DRB1 alleles
DRB1*04 was significantly increased in group 1
compared with controls (OR = 5.0, P < 0.0005) as was
DRB1*03, but not significantly. DRB1*03 was significantly associated with group 2 compared with controls
(OR = 3.7, P < 0.05). No differences in HLA DRB1 frequencies from the healthy population were observed in
the patients with type 2 diabetes — group 3 (Table 2).
Genomic DNA preparation
A salting out method (Miller et al., 1988) was used to
prepare genomic DNA from peripheral blood leukocytes.
The QIAamp spin columns (QIAGEN GmbH, Hilden,
Germany) were used for rapid purification.
Analysis of HLA DQB1 alleles
DQB1*0302 was significantly increased in group 1
compared with controls (OR = 6.1, P < 0.00005), as was
DQB1*0201, but not significantly. In the same group,
DRB1
alleles
Children
(n = 261)
Group 1
(n = 80)
Group 2
(n = 70)
Group 3
(n = 131)
Control
(n = 99)
01
03
04
07
08
09
10
11
12
13
14
15
16
NT
NT
NT
NT
NT
NT
NT
NT
NT
NT
NT
NT
NT
11.7
40.0
45.0**(5.0)
16.7
5.0
0
5.0
11.7
1.7
11.7
1.7
11.7
3.3
16.7
50.0*(3.7)
23.3
13.3
13.3
3.3
0
30.0
0
13.3
3.3
20.0
3.3
16.7
13.3
21.1
26.7
7.8
3.3
3.3
25.6
2.2
15.6
13.3
20.0
4.4
22.2
21.2
14.1
36.4
5.1
1.0
2.0
23.2
10.1
22.2
5.1
23.2
5.1
Data from children with diabetes obtained from Cinek et al. with permission for comparison.
Data from healthy control obtained from Cerna et al. are compared with patient groups.
Group 1 (type 1 diabetes in adults): C peptide < 200 pmol / L, anti-GAD > or < 50 ng /mL.
Group 2 (LADA): C peptide > 200 pmol / L, anti-GAD > 50 ng /mL, > 6 months without insulin.
Group 3 (type 2 diabetes): C peptide > 200 pmol / L, anti-GAD < 50 ng /mL.
NT, not tested, * P < 0.05, ** P < 0.0005, OR values are in brackets.
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401–407
Table 2. Frequency (%) of DRB1 alleles in
Czech patients with diabetes
404
M. Cerna et al.
DQB1
alleles
Children
(n = 261)
Group 1
(n = 80)
Group 2
(n = 70)
Group 3
(n = 131)
Control
(n = 99)
0201
0202
0301
0302
0303
0304
0401
0402
0501
0502
0503
0504
0601
0602
0603
0604
61.0*10(3.4)
NP
5.0*15(0.08)
65.0*15(9.0)
NP
NP
NP
NP
NP
NP
1.0*(0.13)
NP
NP
1.0*15(0.02)
4.0*5(0.20)
NP
41.7
6.7
18.3
43.3**(6.1)
5.0
0
3.3
1.7
33.3
3.3
1.7
0
1.7
3.3
6.7
8.3
42.9
14.3
28.6
14.3
14.3
0
0
3.6
21.4
3.6
0
0
0
10.7
0
7.1
24.4
19.1
31.3
13.7
3.8
2.3
0
5.3
23.7
6.1
6.9
1.5
2.3
16.8
11.5
6.9
30.3
19.2
38.4
11.1
8.1
0
0
4.0
23.2
5.1
5.1
0
3.0
20.2
13.1
5.1
Table 3. Frequency (%) of DQB1 alleles in
Czech patients with diabetes
Data from children with diabetes used from Cinek et al. with permission for comparison.
Data from healthy control used from Cerna et al. are compared with patient groups.
Group 1 (type 1 diabetes in adults): C peptide < 200 pmol /L, anti-GAD > or < 50 ng /mL.
Group 2 (LADA): C peptide > 200 pmol / L, anti-GAD > 50 ng /mL, > 6 months without insulin.
Group 3 (type 2 diabetes): C peptide > 200 pmol / L, anti-GAD < 50 ng /mL.
NP, not published, * P < 0.005, ** P < 0.00005, *5 P < 10–5, *10 P < 10−10, *15 P < 10−15.
OR values are in brackets.
Children
(n = 261)
DRB1 genotype
03/04
NP
DQB1 genotype
0201/0302
36.0*15(22.0)
0201/0301
1.0**(0.1)
Group 1
(n = 80)
Group 2
(n = 70)
Group 3
(n = 131)
Control
(n = 99)
17.0**(8.3)
6.7
0
2.0
15.0*(7.4)
0
3.6
3.6
0.8
6.1
2.0
8.1
Table 4. Frequency (%) of DQB1 and DRB1
genotypes associated with type 1 diabetes
Data from children with diabetes used from Cinek et al. with permission for comparison.
Data from healthy control used from Cerna et al. are compared with patient groups.
Group 1 (type 1 diabetes in adults): C peptide < 200 pmol /L, anti-GAD > or < 50 ng /mL.
Group 2 (LADA): C peptide > 200 pmol / L, anti-GAD > 50 ng /mL, > 6 months without insulin.
Group 3 (type 2 diabetes): C peptide > 200 pmol / L, anti-GAD < 50 ng /mL.
NP, not published, * P < 0.05, ** P < 0.01, *15 P < 10−15, OR values are in brackets.
DQB1*0602 was decreased compared with controls, but
not significantly. In group 2, DQB1*0201 was increased
compared with controls, but not significantly. DQB1*0602
was non-significantly decreased compared with controls
up to 10.7 vs. 20.2%; all carriers had the second DQB1
allele 0201 (30% of all 0602) or 0302 (70% of all 0602)
(Table 3).
HLA DR4 subtyping
In patients with diabetes, we have distinguished three
DRB1*04 alleles: DRB1*0401 (50% of all DR4),
DRB1*0403 (25% of all DR4) and DRB1*0404 (25% of
all DR4). For comparison, in controls, we distinguished
five DRB1*04 alleles: DRB1*0401 (50% of all DR4),
DRB1*0402 (7% of all DR4), DRB1*0404 (21% of all
DR4), DRB1*0405 (14% of all DR4) and DRB1*0408
(7% of all DR4). Comparing the patients with the healthy
population, we have not found any association of
DRB1*04 subtypes with disease phenotypes.
Heterozygosity for DRB1 or DQB1 genotypes
We have observed a statistically significant increased prevalence of high risk DRB1*03/DRB1*04 and
DQB1*0201/DQB1*0302 heterozygotes among patients
in group 1 compared with controls (OR = 8.3, P < 0.01
and OR = 7.4, P < 0.05). This risk became more significant when we divided this group into anti-GAD-positive
and anti-GAD-negative patients, because predisposing
heterozygotes were particularly associated with the
anti-GAD-positive subgroup (25%, OR = 12.4, P < 0.005
both). The protective DQB1*0201/DQB1*0301 genotype has not been found in group 1 of patients with type 1
diabetes mellitus. Other DQB1 genotypes have not been
significantly associated with the disease (Table 4).
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401– 407
HLA in Czech patients with diabetes mellitus
DQB1 genotype
0201/0302
Children
Group 1
Control
0 –4 years
(n = 82)
5 –9 years
(n = 98)
10 –14 years
(n = 81)
(n = 80)
(n = 99)
45.0
36.0
27.0
15.0*(7.4)
2.0
405
Table 5. Frequency (%) of DQB1* 0201/0302
heterozygotes in dependence on the age of
type 1 diabetes manifestation
Data from diabetic children used from Cinek et al. with permission for comparison.
Data from healthy controls used from Cerna et al.
Group 1 (type 1 diabetes in adults): C peptide < 200 pmol / L, anti-GAD > or < 50 ng /mL.
Data from each group are compared with each other.
* P < 0.05, OR values are in brackets.
The HLA class II association and age of onset
Comparing our patients with type 1 diabetes with
children with type 1, diabetes (Cinek et al., 2001), we
found decreasing significance of predisposing alleles with
increasing age of onset. The frequency of DQB1*0302
allele in children was 65%, in adult patients of group 1
43.3%, and in healthy controls 11.1% (Table 3). The
frequency of DQB1*0201 allele in children was 61%, in
adult patients of group 1 41.7%, and in healthy controls
30.3% (Table 3). A decreasing prevalence of genetic
factors with increasing age of onset was not only found
for predisposing single alleles, but also for DQB1* 0201/
0302 heterozygotes (Table 5). While the prevalence of
0201/0302 heterozygotes among children diagnosed
at the age of 0–4, 5–9, and 10 –14 years was 45, 36 and
27%, respectively, the prevalence of 0201/0302 heterozygotes among adults diagnosed after the age of 35 years
was 15% in the patients of group 1.
Discussion
The classification of diabetes mellitus discriminate type 1,
type 2, other specific types and gestational diabetes
(Report, 2001). It has been reported that these types of
diabetes involve separate genetic factors, suggesting that,
in addition to the perceived differences in clinical manifestations and course, their etiologic mechanisms are
distinct. The presence of islet cell autoantibodies in adult
diabetic subjects who do not require insulin treatment for
at least 6 months after the initial clinical diagnosis identifies
the so-called latent autoimmune diabetes in the adult
(LADA). Glutamic acid decarboxylase autoantibodies
(GAD65Ab) are the best immune marker to identify
LADA patients. The main goal of our study was to
analyze HLA class II antigens in adult patients with autoimmune diabetes mellitus, to prove, if the basic data
known from children are valid also in type 1 adult
patients, and to compare these results with those from
LADA patients. In view of the long and accurate followup in our study, we were able to further subdivide diabetic
patients diagnosed after 35 years of age into three groups
that differ in the presence of fasting C peptide and GAD
antibodies, and insulin therapy, respectively.
The association of DQB1*0302 and DRB1*04 that
was observed with the group of adult patients with type 1
diabetes with negative C peptide (group 1) suggests that
these alleles represent a marker or risk factor for requiring
insulin treatment, as was reported (Horton et al., 1999).
The highest risk in this group was for DQB1*0302.
The odds ratio for DQB1*0302 was 6.1 (ratio 2.6–14.9). The
odds ratio for DRB1*04 was lower, 5.0 (ratio 2.2–11.4).
The frequencies of DQB1*0201 and DRB1*03 were increased, but not significant. The frequency of DQB1*0201/
DQB1*0302 and DRB1*03/DRB1*04 genotypes in the
C-peptide-negative group was statistically different from
controls (OR = 7.4, ratio 1.7–33.2 and OR = 8.3, ratio
1.9–36.4), but lower than in children with diabetes. The
pattern of predisposing alleles in adults with type 1
diabetes is similar with that in children with type 1
diabetes, but the prevalence seems to be age related.
Whether these age-dependent differences reflect different
environmental influences remains to be proven. Our
observations correspond with results in Sweden, where
the prevalence of the HLA-DQB1*0201/0302 (DR3/
DR4) genotype reflects the age at onset of diabetes, and
the 0302 genotype is associated with insulin dependency
(Tuomi et al., 1999).
The association of DRB1*03 only was found with the
group of LADA patients with normal C peptide (group 2).
The odds ratio for DRB1*03 was 3.7 (ratio 1.8–8.0).
What was amazing, but very logical, was the exact match
of the presence of C peptide with non-insulin-requirement.
Previous studies proved that for early onset of type 1
diabetes, DQB1 alleles are a better marker of genetic
predisposition than DRB1 alleles (Todd et al., 1987). Our
findings support the observations of some investigators
that DRB1 alleles are the main marker for late onset
(Lohmann et al., 1997). Certain research groups have
described DRB1*04 and DQB1*0302 as the main marker
for LADA (Hosszufalusi et al., 2003). Why different
predisposing DR alleles are detected in different ethnic
groups is not clear. The hypothesis been suggested that
HLA-DRB1 does not predispose to autoimmune disease
per se, but rather fails to provide efficient protection
(Zanelli et al., 2000). According to these facts we can
hypothesize that type 1 diabetes mellitus in children or
adults may have partly different immunogenetic etiopathogenesis than LADA. The different mechanism reflects
different HLA association in these diseases. The most
significant association is with DQB1 alleles in type 1
diabetes, but with DRB1 alleles in LADA.
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401–407
406
M. Cerna et al.
Patients with type 2 diabetes without signs of autoimmune insulitis (fasting CP normal, GAD and IA2
antibodies negative, without insulin therapy) did not
differ in HLA frequencies from healthy Czech controls.
The type of autoantibodies to islet cell antigens distinguish between acute-onset type 1 diabetes and LADA
because GAD antibodies and ICA indicate slow disease
progression, whereas the presence of IA2 antibodies is
associated with an acute-onset clinical phenotype (Seissler
et al., 1998). That is probably why 50% of patients with
type 1 diabetes in adults are negative for anti-GAD. These
patients could be positive for other non-tested antibodies
or their antibody production has burnt-out over the course
of years. Additional testing for IA2 antibodies revealed
that 15% of patients in group 1 (20% in anti-GAD+ and
10% in anti-GAD–) and 11% of patients in group 2 are
positive for this autoantibody.
We conclude that Czech patients with autoimmune
diabetes manifested after 35 years of age can be divided
into two groups having different HLA class II associations:
type 1 diabetes in adults with DQB1*0302 and DRB1*04
risk alleles, and LADA with DRB1*03 risk allele. The
occurrence of HLA-DQB1*0302 (0201) and DRB1*04
(03) alleles in adults with type 1 diabetes was less frequent
than in children, but more frequent than in patients with
type 2 diabetes and healthy controls.
Acknowledgements
This study was financially supported by the Research
Programme of Ministry of Education, Youth, and Sports
of Czech Republic, VZ: J13/98: 111200001: The initial
stages of diabetes, metabolic, endocrine and environmental
damages of the organism (2001). We thank Anezka
Koubova and Jana Potockova for excellent technical
assistance.
References
Cerna, M., Fernandez-Vina, M., Ivaskova, E. & Stastny, P. (1992)
Comparison of HLA class II alleles in gypsy and Czech
populations by DNA typing with oligonucleotide probes.
Tissue Antigens, 39, 111.
Cinek, O., Kolouskova, S., Snajderova, M., Sumnik, Z.,
Sedlakova, P., Drevinek, P., Vavrinec, J. & Ronningen, K.S. (2001)
HLA class II genetic association of type 1 diabetes mellitus in
Czech children. Pediatric Diabetes, 2, 98.
Drabek, J., Vranova, K., Ambruzova, Z., Petrek, M., Bartova, A.
& Weigl, E. (1998) Class I typing by PCR-SSP in Olomouc.
Bone Marrow Transplant, 22, S23.
Dvorakova, L. & Majsky, A. (1979) HLA antigens in offspring of
both diabetic parents. Casopis Lekaru Ceskych, 118, 1294.
Expert Committee on the Diagnosis and Classification of Diabetes
Mellitus. Report. (2001) Diabetes Care, 24, S5.
Gambelunghe, G., Forini, F., Laureti, S., Murdolo, G., Toraldo, G.,
Santeusanio, F., Brunetti, P., Sanjeevi, C.B. & Falorni, A. (2000)
Increased risk for endocrine autoimmunity in Italian type 2
diabetic patients with GAD65 autoantibodies. Clinical
Endocrinological (Oxford), 52, 565.
Graham, J., Kockum, I., Sanjeevi, C.B., Landin-Olsson, M.,
Nystrom, L., Sundkvist, G. et al. (1999) Negative association
between type 1 diabetes and HLA DQB1*0602-DQA1*0102 is
attenuated with age at onset. Swedish Childhood Diabetes
Study Group. European Journal of Immunogenetics,
26, 117.
Horton, V., Stratton, I., Bottazzo, G.F., Shattock, M., Mackay, I.,
Zimmet, P., Manley, S., Holman, R. & Turner, R. (1999) Genetic
heterogeneity of autoimmune diabetes: age of presentation in
adults is influenced by HLA DRB1 and DQB1 genotypes
(UKPDS 43). UK Prospective Diabetes Study (UKPDS) Group.
Diabetologia, 42, 608.
Hosszufalusi, N., Vatay, A., Rajczy, K., Prohaszka, Z., Pozsonyi, E.,
Horvath, L. et al. (2003) Similar genetic features and different
islet cell autoantibody pattern of latent autoimmune diabetes in
adults (LADA) compared with adult-onset type 1 diabetes with
rapid progression. Diabetes Care, 26, 452.
Lohmann, T., Seissler, J., Verlohren, H.J., Schroder, S., Rotger, J.,
Dahn, K., Morgenthaler, N. & Scherbaum, W.A. (1997)
Distinct genetic and immunological features in patients
with onset of IDDM before and after age 40. Diabetes Care, 20,
524.
Loudova, M., Sramkova, I., Cukrova, V., Sieglova, Z. & Brdicka, R.
(1999) Frequencies of HLA-DRB1-DQB1 and -DPB1 alleles in
Czech population. Folia Biologica (Praha), 45, 27.
Miller, S.A., Dykes, D.D. & Polesky, H.F. (1988) A simple salting
out procedure for extracting DNA from human nucleated
cells. Nucleic Acids Research, 16, 1215.
Olerup, O., Aldener, A. & Fogdell, A. (1993) HLA-DQB1
and -DQA1 typing by PCR amplification with sequencespecific primers (PCR-SSP) in 2 hours. Tissue Antigens,
41, 119.
Seissler, J., de Sonnaville, J.J., Morgenthaler, N.G., Steinbrenner, H.,
Glawe, D., Khoo-Morgenthaler, U.Y., Lan, M.S., Notkins, A.L.,
Heine, R.J. & Scherbaum, W.A. (1998) Immunological
heterogeneity in type I diabetes: presence of distinct autoantibody
patterns in patients with acute onset and slowly progressive
disease. Diabetologia, 41, 891.
Sinha, A.A., Lopez, M.T. & McDevitt, H.O. (1990)
Autoimmune diseases: the failure of self tolerance. Science,
248, 1380.
Tait, B.D., Harrison, L.C., Drummond, B.P., Stewart, V.,
Varney, M.D. & Honeyman, M.C. (1995) HLA antigens and age
at diagnosis of insulin-dependent diabetes mellitus. Human
Immunology, 42, 116.
Thomson, G., Robinson, W.P. & Kuhner, M.K. (1988) Genetic
heterogeneity, modes of inheritance, and risk estimates for a joint
study of Caucasians with insulin-dependent diabetes mellitus.
American Journal of Human Genetics, 43, 799.
Todd, J.A., Bell, J.I. & McDevitt, H.O. (1987) HLA-DQβ gene
contributes to susceptibility and resistance to insulin-dependent
diabetes mellitus. Nature, 329, 599.
Tuomi, T., Carlsson, A., Li, H., Isomaa, B., Miettinen, A.,
Nilsson, A. et al. (1999) Clinical and genetic characteristics of
type 2 diabetes with and without GAD antibodies. Diabetes,
48, 150.
Vauhkonen, I., Niskanen, L., Knip, M., Ilonen, J., Vanninen, E.,
Kainulainen, S., Uusitupa, M. & Laakso, M. (2000)
Impaired insulin secretion in non-diabetic offspring of probands
with latent autoimmune diabetes mellitus in adults.
Diabetologia, 43, 69.
Verge, C.F., Stenger, D., Bonifacio, E., Colman, P.G., Pilcher, C.,
Bingley, P.J. & Eisenbarth, G.S. (1998) Combined use of
autoantibodies (IA-2 autoantibody, GAD autoantibody, insulin
autoantibody, cytoplasmic islet cell antibodies) in type 1 diabetes:
Combinatorial Islet Autoantibody Workshop. Diabetes,
47, 1857.
Vondra, K., Vrbikova, J., Ivaskoval, E., Pobisova, Z., PorsovaDutoit, I., Skibova, J., Stolba, P., Voborska, M. & Sterzl, I.
(1996) Thyroglobulin and microsome autoantibodies and
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401– 407
HLA in Czech patients with diabetes mellitus
their clinical significance in adult type I diabetics. Vnitrni
Lekarstvi, 42, 767.
Zanelli, E., Breedveld, F.C. & de Vries, R.R. (2000) HLA association
with autoimmune disease: a failure to protect? Rheumatology, 10,
1060.
Zetterquist, H. & Olerup, O. (1992) Identification of the
HLA-DRB1*04-DRB1*07, and -DRB1*09 alleles by PCR
407
amplification with sequence-specific primers (PCR-SSP) in 2
hours. Human Immunology, 34, 64.
Zevaco-Mattei, C., Reviron, D., Atlan-Gepner, C., Botti, G.,
Simonin, G., Lassmann-Vague, V., Vague, P., Mercier, P. &
Vialettes, B. (1999) Relationship between HLA-DQ and -DR
genotypes and clinical characteristics in a French population of
type 1 diabetic patients. Diabetic Medicine, 16, 201.
© 2003 Blackwell Publishing Ltd, European Journal of Immunogenetics 30, 401–407
Original paper No. 6
DIABETES MELLITUS IN ADULTS: ASSOCIATION OF HLA DRB1 AND DQB1 DIABETES
RISK ALLELES WITH GADAB PRESENCE AND C-PEPTIDE SECRETION
Novota P, Cerna M, Kolostova K, Cejkova P, Zdarsky E, Novakova D, Kucera P,
Novak J, Andel M
Immunol Lett. 2004; 95:229-232
119
Immunology Letters 95 (2004) 229–232
Diabetes mellitus in adults: association of HLA DRB1
and DQB1 diabetes risk alleles with GADab presence
and C-peptide secretion
Peter Novotaa,∗ , Marie Cernaa , Katarina Kolostovaa , Pavlina Cejkovaa , Emanuel Zdarskya ,
Dana Novakovab , Petr Kucerab , Jan Novakc , Michal Andelc
a Department of Cell and Molecular Biology, The 3rd Faculty of Medicine, Charles University at Prague, Czech Republic
Department of Cell and Molecular Immunology, The 3rd Faculty of Medicine, Charles University at Prague, Czech Republic
Diabetes Center and Department of Internal Medicine II, The 3rd Faculty of Medicine, Charles University at Prague, Czech Republic
b
c
Received 29 June 2004; received in revised form 21 July 2004; accepted 29 July 2004
Available online 23 August 2004
Abstract
In our study, we investigated the relationship of HLA class II alleles to antibody production against glutamic acid decarboxylase (GADab)
and to C-peptide secretion (CP) in diabetic patients. A group of 334 patients (190 women) diagnosed after 35 years of age and 99 control
subjects were studied. Patients were divided into four groups according to concentrations of CP and GADab, respectively (CP high/low,
GADab positive/negative). HLA DQB1 and DRB1 alleles were genotyped by SSP-PCR. The significance of DQB1 and DRB1 risk alleles
was evaluated by examination of their odds ratios computed by testing 2 × 2 tables considering Bonferonis’ corrected P < 0.05 as significant. We found strong association between the HLA DRB1*03 risk allele and presence of GADab, and close relationship of the HLA
DRB1*04 and HLA DQB1*0302 risk alleles with decreased CP level. Taken together we conclude that the DRB1*04 and DQB1*0302
alleles are associated with progressive decrease of CP level, while DRB1*03 is a significant genetic marker of autoantibody (GADab)
development.
© 2004 Elsevier B.V. All rights reserved.
Keywords: HLA/DQ/DR; Glutamic acid decarboxylase (GAD); C-peptide (CP); Diabetes mellitus (DM)
1. Introduction
Diabetes mellitus with onset after the age of 35 years [1]
is a heterogeneous group representing at least two major
groups of diseases: insulin deficient (type 1 diabetes mellitus, T1DM) and insulin resistant diabetes mellitus (type 2
diabetes mellitus, T2DM) [2,3]. The etiology of the former
is an autoimmune disorder and the etiology of the latter is a
metabolic disorder. However, about 10% of patients initially
∗ Corresponding author. Present address: Department of Cell and
Molecular Biology, Department of Immunogenetics (Z4), DNA laboratory,
Institute for Clinical and Experimental Medicine, Vı́deňská 1958/9, 14021
Prague 4, Czech Republic. Tel.: +420 261 362 350; fax: +420 241 721 603.
E-mail address: [email protected] (P. Novota).
0165-2478/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.imlet.2004.07.016
diagnosed as patients with type 2 diabetes manifest autoimmune markers of ␤-cell destruction. These subjects have a
slowly progressive form of autoimmune diabetes mellitus and
are considered having latent autoimmune diabetes mellitus in
adults (LADA) [4–7].
The most important gene loci defining risk to T1DM are
located within the HLA gene region. HLA DQ molecules are
of primary importance but HLA DR gene products modify
the risk conferred by HLA DQ. The risk associated with an
HLA genotype is defined by the particular combination of
susceptible and protective alleles. The complicated analysis
of HLA genotypes is simplified by strong linkage disequilibrium between HLA DRB1, DQA1 and DQB1 loci. Childhood
T1DM is characterized by an abrupt onset and ketosis and is
230
P. Novota et al. / Immunology Letters 95 (2004) 229–232
associated with HLA DRB1*04-DQA1*0301-DQB1*0302,
and a high frequency of insulin and thyrosin-phosphatase autoantibodies [7,8].
The term latent autoimmune diabetes in adults was introduced to define adult diabetic patients initially non-insulinrequiring but with immune markers of type 1 diabetes
that, in a number of cases, progress to insulin dependency
[6]. It is characterized by the presence of glutamic acid decarboxylase 65 autoantibodies (GADab) and/or islet cell
antibodies (ICA), and is associated with HLA DRB1*03DQA1*0501-DQB1*0201 and/or with HLA DRB1*04DQA1*0301-DQB1*0302 haplotype [5,9].
In patients diagnosed after 35 years of age, the relation
of HLA risk alleles and GADab and/or C-peptide secretion
(CP), generally used as a marker of insulin production, is
not completely clear. In this study, we investigated the relationship between the presence of diabetes risk HLA alleles,
and the presence of GADab and/or CP in diabetic patients
diagnosed after 35 years of age.
2. Patients and methods
2.1. C-peptide level
Levels of CP were measured on an empty stomach (fasting C-peptide). For detection, we used an immunoradiometric method (Immunotech, Prague, Czech Republic). The
serum samples from 99 healthy individuals provided control values: average value 486 pmol/l; median 449 pmol/l;
standard deviation 170 pmol/l; and minimal–maximal value
206–934 pmol/l. According by values lower than 200 pmol/l
were taken as decreased C-peptide secretion (marked as low
CP, CP negative, or CP−).
2.2. GAD antibodies
The presence of IgG antibodies against GAD was detected by the ELISA method (Roche Molecular Biochemicals, Mannheim, Germany). The test had 98% of specificity
and 69% of sensitivity. The cut-off point of the test was
32 ng/ml. Levels lower than this cut-off point were considered
negative (marked as GADab negative, or GADab−), levels
higher than 32 ng/ml were considered positive (marked as
GADab positive, or GADab+) [10].
2.3. Genomic DNA preparation
The QIAamp spin columns (QIAGEN GmbH, Hilden,
Germany) were used for extraction and purification of genomic DNA from peripheral blood leukocytes.
2.4. Analysis of HLA DRB1 and DQB1 alleles
HLA DRB1 and DQB1 alleles were genotyped using PCR
with sequence specific primers (SSP-PCR) supplied by Genovision (Oslo, Norway) [11,12].
2.5. Statistics
Allele and gene frequencies were calculated using standard methods. The significance of the differences in allele
frequencies were evaluated by χ2 and P values were calculated with Bonferoni correction. The strength of the observed
associations was estimated by calculating odds ratios (OR)
using the method by Woolf.
2.6. Group of patients
We investigated 334 patients (190 women) with diabetes
mellitus with disease onset after 35 years of age. All of them
were Caucasians and residents of the Middle region of Czech
Republic. Diagnosis of diabetes was established according to
the current WHO definitions and criteria for diagnosing diabetes [13]. The T1DM (49), LADA (48), and T2DM (237)
patients were included. Patients with autoimmune diabetes
(T1DM and LADA) were negative for C-peptide and/or positive for autoantibodies. All of them had insulin therapy.
LADA was defined by a minimum 6 months long phase after
diagnosis without insulin therapy. Patients with T2DM were
positive for C-peptide. None ever had insulin therapy. The
age at disease onset and duration of DM were recorded. For
detailed characterization of patient groups see Table 1.
Table 1
Clinical characteristics of investigated patient groups
Number of patients
Number of female
Body mass index (BMI)
Age of onset of DM (years)
Duration of DM (years)
Diagnosis
T1DM, N = 49
LADA, N = 48
T2DM, N = 237
CP+ (≥200 pmol/l)
CP− (≤200 pmol/l)
GADab+ (≥32 ng/ml)
GADab− (≤32 ng/ml)
285
153
31.3 (22.2; 50.1)
53 (35; 81)
12.35 (0; 31)
49
32
28.0 (22.4; 36.7)
44 (36; 56)
14.46 (1; 29)
97
58
31.1 (22.4; 45.5)
52 (35; 71)
13.27 (1; 31)
237
136
31.1 (22.2; 50.1)
53 (35; 81)
12.48 (0; 31)
50.5% (49/97)
49.5% (48/97)
0.0% (0/97)
0.0% (0/237)
0.0% (0/237)
100.0% (237/237)
0.0% (0/285)
16.8% (48/285)
83.2% (237/285)
100.0% (49/49)
0.0% (0/49)
0.0% (0/49)
Values of BMI, age of onset of disease, and duration of DM are calculated as average. In brackets are minimal and maximal values.
P. Novota et al. / Immunology Letters 95 (2004) 229–232
231
Table 2
HLA DQB1 frequencies in investigated patient groups (GADab+, GADab−, CP+, and CP−)
DQB1*0201
DQB1*0202
DQB1*0301
DQB1*0302
DQB1*0303
DQB1*0304
DQB1*0402
DQB1*0501
DQB1*0502
DQB1*0503
DQB1*0504
DQB1*0601
DQB1*0602
DQB1*0603
DQB1*0604
GADab+ (N = 45)
GADab− (N = 146)
CP+ (N = 153)
CP− (N = 38)
Controls (N = 99)
0.467
0.133
0.267
0.289
0.111
0.022
0.067
0.178
0.022
0.000
0.000
0.022
0.067
0.022
0.111
0.267
0.185
0.315
0.205
0.048
0.021
0.048
0.281
0.068
0.062
0.014
0.021
0.164
0.110
0.062
0.288
0.190
0.333
0.170
0.065
0.020
0.059
0.248
0.059
0.059
0.013
0.020
0.170
0.098
0.065
0.447
0.105
0.184
0.447 (6.48)∗
0.053
0.026
0.026
0.316
0.053
0.000
0.000
0.026
0.026
0.053
0.079
0.303
0.192
0.384
0.111
0.081
0.000
0.040
0.232
0.051
0.051
NT
0.030
0.202
0.131
0.051
Statistically significant difference in frequency of allele DQB1*0302 in group CP− in comparison with healthy control. Data from healthy control used from
Cerna et al. [14].
∗ P < 0.001; OR values in brackets.
c
3. Results
4. Discussion
We investigated the distribution of HLA DRB1 and DQB1
alleles and the odds ratios of risk and protective alleles in the
four patient groups (Table 1). Subsequently, we explored the
relationship between HLA risk or protective alleles, and positivity (negativity) for GADab and values of CP, respectively.
In the GADab+ group, we found only one predisposing
allele (DRB1*03, Pc = 0.0002). In the GADab− group, we
did not find any association with HLA alleles. In the CP−
group, we detected two risk alleles, DRB1*04 (Pc = 0.0001)
and DQB1*0302 (Pc = 0.0002). In the CP+ group, we did
not detected any HLA association. None protective alleles
were ascertained in all these groups (see Tables 2 and 3).
Odds ratios of the predisposing alleles were following: in
the GADab+ group, OR = 4.90 (CI 95% = 2.21; 10.94) was
calculated for allele DRB1*03; in the CP−group, OR = 5.80
(CI 95% = 2.36; 14.39) was calculated for allele DRB1*04;
and OR = 6.48 (CI 95% = 2.43; 17.54) for allele DQB1*0302.
We analyzed the frequencies of HLA class II alleles in
patients with diabetes mellitus manifested after 35 years of
age. We found two risk HLA alleles in the CP− group and
one in the GADab+ group of patients. The frequencies of
the DRB1*04 and DQB1*0302 alleles were significantly increased in the CP− group (Pc < 0.001 for both alleles). The
odds ratio was higher for DQB1*0302 (OR = 6.48) than for
DRB1*04 (OR = 5.80), what suggests DQB1*0302 as main
risk allele in the CP−group. The frequency of DRB1*03 allele was significantly higher in the GADab+ group (Pc <
0001). Its odds ratio was the lowest in the study (OR = 4.90).
Our results are similar to data published by Caillat-Zucman
et al. [7] or Krokowski et al. [8]. The only difference is in
the frequency of the known risk allele DQB1*0201. The frequency of this allele was increased in our GADab+ group,
but it did not reach statistically significant difference when
compared with controls.
Table 3
HLA DRB1 frequencies in investigated patient groups (GADab+, GADab−, CP+, and CP−)
DRB1*01
DRB1*03
DRB1*04
DRB1*07
DRB1*08
DRB1*09
DRB1*10
DRB1*11
DRB1*12
DRB1*13
DRB1*14
DRB1*15
DRB1*16
GADab+ (N = 51)
GADab− (N = 111)
CP+ (N = 122)
CP− (N = 43)
Controls (N = 99)
0.157
0.569 (4.90)∗
0.314
0.118
0.098
0.020
0.000
0.255
0.000
0.137
0.020
0.176
0.020
0.198
0.162
0.270
0.252
0.081
0.027
0.045
0.225
0.018
0.135
0.108
0.171
0.054
0.164
0.221
0.230
0.230
0.090
0.033
0.033
0.262
0.016
0.156
0.107
0.197
0.041
0.233
0.465
0.488 (5.80)∗
0.163
0.070
0.000
0.047
0.140
0.000
0.093
0.000
0.093
0.047
0.222
0.212
0.141
0.363
0.050
0.010
0.020
0.232
0.101
0.222
0.050
0.232
0.050
Statistically significant difference in frequency of allele DRB1*03 in group GADab+ and allele DRB1*04 in group CP− in comparison with healthy control.
Data from healthy control used from Cerna et al. [14].
∗ Pc < 0.001; OR values in brackets.
232
P. Novota et al. / Immunology Letters 95 (2004) 229–232
Our study showed that there is a difference in HLA
class II allele frequencies between GADab+ group and
CP−group. GADab positivity is strongly associated with
the HLA DRB1*03 allele and CP negativity is linked to the
DQB1*0302 and HLA DRB1*04 alleles. These results are
supported by Gambelunghe et al. [9] who stated that GADab
is closely associated with the presence of HLA DRB1*03DQB1*02 haplotype. Finally, we conclude that the DRB1*04
and DQB1*0302 alleles are associated with decrease of CP
levels while DRB1*03 is a significant genetic marker of autoantibody (GADab) development.
Acknowledgements
This study was financially supported by the Research Program of Ministry of Health of Czech Republic, IGA MZ CR:
NB/7432-3/2003.
References
[1] Newman B, Selby JV, King MC, Slemenda C, Fabsitz R, Friedman GD. Concordance for type 2 (non-insulin-dependent) diabetes
mellitus in male twins. Diabetologia 1987;30:763–8.
[2] Niskanen L, Karjalainen J, Siitonen O, Uusitupa M. Metabolic evolution of type 2 diabetes: a 10-year follow-up from the time of
diagnosis. J Int Med 1994;236:263–70.
[3] Vauhkonen I, Niskanen L, Vanninen E, Kainulainen S, Uusitupa M,
Laakso M. Defects in insulin secretion and insulin action in noninsulin-dependent diabetes mellitus are inherited. Metabolic studies
on offspring of diabetic probands. J Clin Invest 1998;101:86–96.
[4] Groop LC, Bottazzo GF, Doniach D. Islet cell antibodies identify
latent type I diabetes in patients aged 35–75 years at diagnosis.
Diabetes 1986;35:237–41.
[5] Kobayashi T, Tamemoto K, Nakanishi K, Kato N, Okubo M, Kajio H, Sugimoto T, Murase T, Kosaka K. Immunogenetic and
clinical characterization of slowly progressive IDDM. Diab Care
1993;16:780–8.
[6] Tuomi T, Groop LC, Zimmet PZ, Rowley MJ, Knowles W, Mackay
IR. Antibodies to glutamic acid decarboxylase reveal latent autoimmune diabetes mellitus in adults with a non-insulin-dependent onset
of disease. Diabetes 1993;42:359–62.
[7] Caillat-Zucman S, Garchon HJ, Timsit J, Assan R, Boitard C, DjilaliSaiah I, Bougneres P, Bach JF. Age-dependent HLA genetic heterogeneity of type 1 insulin-dependent diabetes mellitus. J Clin Invest
1992;90:2242–50.
[8] Krokowski M, Bodalski J, Bratek A, Boitard C, Caillat-Zucman S.
HLA class II-associated predisposition to insulin-dependent diabetes
mellitus in a Polish population. Hum Immunol 1998;59:451–5.
[9] Gambelunghe G, Forini F, Laureti S, Murdolo G, Toraldo G, Santeusanio F, Brunetti P, Sanjeevi CB, Falorni A. Increased risk for endocrine autoimmunity in Italian type 2 diabetic patients with GAD65
autoantibodies. Clin Endocrinol (Oxf) 2000;52:565–73.
[10] Verge CF, Stenger D, Bonifacio E, Colman PG, Pilcher C, Bingley
PJ, Eisenbarth GS. Combined use of autoantibodies (IA-2 autoantibody, GAD autoantibody, insulin autoantibody, cytoplasmic islet
cell antibodies) in type 1 diabetes: Combinatorial Islet Autoantibody
Workshop. Diabetes 1998;47:1857–66.
[11] Olerup O, Zetterquist H. HLA-DR typing by PCR amplification
with sequence-specific primers (PCR-SSP) in 2 h: An alternative
to serological DR typing in clinical practice including donorrecipient matching in cadaveric transplantation. Tissue Antigens
1992;39:225–35.
[12] Olerup O, Aldener A, Fogdell A. HLA-DQB1 and -DQA1 typing
by PCR amplification with sequence-specific primers (PCR-SSP) in
2 h. Tissue Antigens 1993;41:119–34.
[13] World Health Organization. Definition, Diagnosis and Classification of Diabetes Mellitus and its Complications. Geneva, Switzerland: Department of Noncommunicable Disease Surveillance, WHO;
1999.
[14] Cerna M, Fernandez-Vina M, Ivaskova E, Stastny P. Comparison of
HLA class II alleles in gypsy and Czech populations by DNA typing
with oligonucleotide probes. Tissue Antigens 1992;39:111.
Original paper No. 7
GENETIC RISK FACTORS IN AUTOIMMUNE DIABETES MELLITUS, THEIR SIGNIFICANCE
AND FUNCTION
Novota P, Cejkova P, Cerna M, Andel M
Cas Lek Cesk. 2004; 143(3):159-63. Review in Slovak
124
Original paper No. 8
HLA, NFKB1 AND NFKBIA GENE POLYMORPHISM PROFILE IN AUTOIMMUNE
DIABETES MELLITUS PATIENTS
Kolostova K, Pinterova D, Novota P, Romzova M, Cejkova P, Pruhova S, Lebl J,
Treslova L, Andel M, Cerna M
Exp Clin Endocrinol Diabetes 2007 115:124-129
130
124 Article
HLA, NFKB1 and NFKBIA Gene Polymorphism Profile in
Autoimmune Diabetes Mellitus Patients
Authors
K. Katarina1, P. Daniela1, N. Peter1, R. Marianna1, C. Pavlina1, P. Stepanka2, L. Jan2, T. Ludmila3, A. Michal3,
C. Marie1
Affiliations
1
2
3
Key words
䉴 T1DM
䊉
䉴 HLA
䊉
䉴 NFKB1
䊉
䉴 NFKBIA
䊉
䉴 LADA
䊉
Funding sources: The project
was supported by MSM
0021620814
received
1. 3. 2006
first decision 12. 7. 2006
accepted
26. 7. 2006
Bibiliography
DOI 10.1055/s-2007-949589
Exp Clin Endocrinol Diabetes
2007; 115:124–129
© J.A. Barth Verlag in Georg
Thieme Verlag KG · Stuttgart ·
New York ·
ISSN 0947-7349
Correspondence
K. Katarina
Medical Faculty Charles
University · Department of Molecular and Cell Biology · Ruská
87 · 10034 Prague 10 · Czech
Republic
Tel.: + 420/26 71 02 66 2
Fax: + 420/26 71 02 65 0
[email protected].
cuni.cz
Department of Molecular and Cell Biology, 3rd Medical Faculty, Charles University, Prague Czech Republic
Department of Pediatrics, Centrum for Diabetes, metabolism and nutrition research of 3rd Medical Faculty,
Charles University, Prague Czech Republic
Second Internal Clinics, Centrum for Diabetes, metabolism and nutrition research of 3rd Medical Faculty,
Charles University, Prague Czech Republic
Type 1 diabetes mellitus (T1DM) is one of the
long-time studied autoimmune disorders. The
triggering of the autoimmune process has been
ascribed to various genes active in the regulation of the cytokine gene transcription including
the Rel/NF-B gene family. In our study the gene
polymorphism of HLA class II, NFKB1 (nuclear
factor of kappa light polypeptide gene enhancer
in B-cells 1) and NFKBIA (inhibitor of nuclear factor kappa B) was tested. Patients were divided
into the subgroups in relation to the disease
type: T1DM in children, T1DM in adults, and
Latent Autoimmune Diabetes in Adults (LADA).
HLA-DRB1*04 and HLA-DQB1*0302 have been
detected as risk factors for T1DM in adults and
particularly in children (P < 0.0001, OR = 22.9
and 46.5 respectively). HLA-DRB1*03 has been
found as a single risk factor for LADA (P < 0.0001,
OR = 4.9). We detected 15 alleles for the NFKB1
gene polymorphism (CA-repeats) in the Czech
population. The alleles were ranging in size from
114-142 bp corresponding to 10–25 CA repeats.
Frequency of the A7 allele of NFKB1 gene has
been significantly increased in T1DM adults
(P < 0.01). There was no difference in A and a G
allele frequency of NFKBIA gene between the
control group and patients, but the association of
the AA genotype of NFKBIA gene has been found
for LADA (P < 0.05). Summarizing our results
we concluded that there is a high probability of
association of gene polymorphism from Rel/NFB family with an autoimmune diabetes course.
Due to the results obtained in the epidemiological study we have been looking also for the function significance of the genetic predisposition.
No significant changes have been observed by
real time PCR testing of HLA-DRB1*04 gene and
NFKB1 gene expression between T1DM diabetic
group with different HLA, NFKB1, NFKBIA genetic
background.
Abbreviations
&
Introduction
&
Abstract
&
LADA
Latent autoimmune diabetes in
adults
NFB
Nuclear factor kappa B
NFKB1
Nuclear factor of kappa light
polypeptide gene enhancer in Bcells 1 gene
NFKBIA
Inhibitor of nuclear factor kappa B
gene
T1DM
Type 1 diabetes mellitus
Type 1 diabetes mellitus (T1DM) is a chronic
autoimmune disease characterized by a loss of
tolerance towards own antigenic structures of
beta-pancreatic cells. The destruction of beta
cells subsequently leads to the loss of insulin
production. T1DM is not a homogeneous disease,
since several of its clinical features are different
in children up to 6 years of age as compared to
older patients (Csorba & Lyon, 2005). There are
more factors, which trigger the autoimmune
response in susceptible individuals; however,
they are only partially known so far.
One of the predisposing factors is the genotype
background, ascribed mainly to genes of the HLA.
Out of extensive genetic and epidemiological
studies, the Caucasoid population is known to
have a significant association of insulin-depend-
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
Article 125
Table 1
Characteristics of the tested groups
Features
80
Female
Age at disease onset (years)
Duration of DM (years)
Insulin therapy
Fasting C peptide (pmol/L)
Anti-GAD (ng/ml)
*
Group 1
Group 2
Group 3
Group 4
T1DM children
n=179
60 %
7.7 (1–16)
4.9 (1–16)
yes
T1DM adults
n=75
52 %
22.1 (20–45)
17.0 (1–50)
yes
Control
n=153
*
*
*
*
LADA
n=31
44 %
47.0 (25–64)
15.0 (3–32)
yes
478.0 (4.4–1522)
392.0 (5–2800)
*
*
*
no
*
*
data not available.
ent diabetes mellitus with the increased frequencies of haplotypes HLA-DRB1*04-DQA1*0301-DQB1*0302 and DRB1*0301DQA1*0501-DQB1*0201. The strength of associations is ethnically
variable. The HLA genotype DRB1*03-DQB1*0201/DRB1*04DQB1*0302 confers a 25-fold increase in the risk of T1DM. (Jaini
et al., 2002; Rewers et al., 2003; Shawkatova et al., 2000).
It is well known that T1DM results from the breakdown of
peripheral tolerance. (Wheat et al., 2004) that is ended with
cytokine-induced beta-cell death. Since the nuclear factor
kappa-B (NFB) plays an important role in cytokine-induced
gene activation, it is a very attractive candidate for T1DM predisposition. NFB is a transcription factor which interacts with
kappa inhibitory proteins (IB) to regulate gene expression (Curran et al., 2002) of a variety of processes including innate and
adaptive immune responses, cell growth, apoptosis, tissue differentiation and inflammation. (Baldwin, 2001). The NFB transcription factor complex has two alternative DNA binding
subunits, NFKB1 and NFKB2. NFKB1 encodes two isoforms, the
cytoplasmatic non-DNA-binding p105, and the 50 kDa DNAbinding p50 (Heron et al., 1995). To exert its effect, p50 binds to
p65 (encoded by NFKB2) to form biologically active heterodimers, which activate transcription in promoter sequences of
inflammatory genes (e.g. IL-12, TNF-, IFN-), but alternatively
p50 able to form homodimers that block gene transcription by
binding to NFB response sites in nuclues. Recent studies have
investigated role of NFB in the pathogenesis of various human
diseases including immune deficiency, carcinogenesis and
atherogenesis. Lately the possible link between NFB and the
development of insulin resistance, type 2 diabetes (Arkan et al.,
2005; Cai et al., 2005; Chen, 2005; Evans et al., 2002) and in
diabetic polyneuropathies (Haslbeck et al., 2005) has also been
suggested.
This study investigated common variants within the genes coding for NFB and IB, NFKB1 [4q24] and NFKBIA [14q13] to test
the probable genetic predisposition to autoimmune diabetes.
The NFB complex is inhibited by IB proteins (NFKBIA or NFKBIB), which inactivate NFB by trapping it in the cytoplasm.
Phosphorylation of serine residues on the IB proteins by kinases
(IKBKA or IKBKB) marks them for destruction via the ubiquitination pathway, thereby allowing activation of the NFB complex.
NFKBIA is encoding for IkB, which binds preferentially to p65.
After degradation of IkB, p65 tranlocates to the nucleus where
it can form heterodimers with p50, released from p105 and following NFB heterodimer complex binds to decameric DNA
sequences and activates transcription of NFB regulated target
genes (Barnes & Karin, 1997; Bierhaus et al., 2001; Hayden &
Ghosh, 2004).
A polymorphic dinucleotide CA microsatellite repeat, with 18
described alleles, has been identified close to the coding region
of the human NFKB1 gene. NFKBIA genotypes (A/G in 3’UTR)
have been determined in previous studies as wild-type 424 bp;
variant 306 and 118 bp; and heterozygote 424, 306 and 118 bp
after restriction digestion (Curran et al., 2002).
Following the results obtained in our epidemiological studies
we have also focused on the probable functional significance of
the tested gene polymorphisms. It has been demonstrated that
variations of the density of HLA class II molecules on APCs influence the intensity and the nature of T cell response (Lechler et
al., 1985). Due to the fact that NFkB participates even if indirectly on the regulation of HLA transcription we have finally
decided to compare the mRNA level of the HLA-DRB1*04 allele
and NFKB1 gene on APCs of peripheral blood leucocytes in diabetic patients in connection to the genotype background tested
previously.
Methods
&
Subjects
The 267 individuals with a previous diabetes mellitus diagnosis,
159 ethnically matched controls for NFKBIA analysis and 58 controls for NFKB1 genotyping were involved into the study. Criteria
of the current WHO definitions for diagnosing diabetes were
applied (World Health Organization Definition, 1999), considering patients’ clinical features and laboratory data, including the
presence of anti-islet autoantibodies (autoantibodies to glutamic
acid decarboxylase - GAD65) and serum C-peptide (CP) level. All
of the patients had insulin therapy. LADA was defined by a minimum 6 months long phase after diagnosis without insulin therapy. All subjects were informed and consented to participate
before the study. The following parameters were recorded for
each patient at the time of study: sex, age, disease duration,
complications and family history (see Table 1). Both affected and
control populations were recruited from individuals residing in
the same geographical location in the Czech region and were
from Caucasian descent.
Genotyping
DNA was extracted from peripheral blood leukocytes using of
DNA blood mini isolation method (Qiagen, Hilden, Gemany).
Genotyping. of the NFKB1 dinucleotide repeat with use of fluorescently labeled primers was previously described by Ota et al.
(Ota et al., 1999). Polymerase chain reaction (PCR) amplification
was performed in a reaction mixture containing 25–50 ng
genomic DNA, 0.5 mM each primer, 0.2 mM dNTPs, 1xPCR buffer,
2.5 mM MgCl2 and DNA polymerase. PCR conditions were 4 min
at 94˚C; 30 cycles of 30 s at 94˚C, 30 s at 52˚C and 30 s at 72˚C; followed by 5 min at 72˚C. NFKB1 alleles were then detected by
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
126 Article
Table 2
The relative risk values and more additive statistics for T1DM associated HLA – alleles in the tested groups
Groups
LADA
T1DM children
T1DM adults
Controls
Alelle frequencies
HLA-DRB1*03
P
Relative risk
Odds ratio
HLA-DRB1*04
P
Relative risk
Odds ratio
HLA-DQB1*0201
P
Relative risk
Odds ratio
HLA-DQB1*0302
P
Relative risk
Odds ratio
n=27 %
29.6
0.0001
3.25 (1.69–6.23)
4.94 (1.83–13.49)
18.5
n.s.
–
–
25.0
n.s.
–
–
14.3
n.s.
–
–
n=188 %
25.0
0.0001
1.51 (1.25–1.81)
3.30 (1.78–6.17)
43.8
0.0001
3.12 (2.35–4.15)
22.87 (10.89–48.87)
21.7
n.s.
–
–
30.8
0.0001
8.98 (4.76–16.93)
46.48 (16.15–140.23)
n=62 %
21.0
n.s
–
–
34.6
0.0001
3.42 (2.32–5.04)
10.47 (4.53–24.65)
21.8
n.s.
–
–
29.0
0.0001
3.02 (2.12–4.30)
9.49 (3.91–23.59)
n=99 %
10.1
fragment analysis on the ALF express II detection system (Amersham Pharmacia Biotech, Uppsala, Sweden).
NFKBIA amplification was performed in a 20 l final volume containing genomic DNA, 1xPCR buffer, 3.75 mM MgCl2, 0.2 mM
dNTPs, 0.5 mM each primer and DNA polymerase. PCR conditions were initial denaturation at 94˚C for 5 min, followed by 30
cycles of 94˚C for 30 s, 52˚C for 30 s and 72˚C for 30 s, with a final
extension of 2 min at 72˚C. Following amplification, 10 l of product was digested with HaeIII at 37˚C overnight. Samples were
then loaded into an ethidium bromide stained 2 % agarose gel for
genotype determination. Genotypes of 268 patients and 159
matched controls were determined as type (AA), 424 bp; variant
(GG), 306 and 118 bp; and heterozygote (AG), 424, 306 and 118 bp
(Curran et al., 2002).
HLA class II typing was performed according to the standard protocol of Genovision (SSP-PCR) (Genovision, Oslo, Norway). PCR
conditions were initial denaturation at 94˚C for 5 min, followed
by 30 cycles of 94˚C for 30 s, 52˚C for 30 s, 72˚C for 30 s, with a
final extension of 10 min at 72˚C.
Gene expression testing
According to the genotyping study the independent sets of children and adults T1DM patients and sex- and age-matched
healthy controls were chosen for the expression studies. 28
T1DM patients included in this study were adults, and 55 T1DM
patients were children. The adult control set of patients was created in cooperation with blood transfusion stationary. The age of
adults was 36,4 ± 11,5 (mean ± SD). The children - age matched
control set was obtained from a phenylketonuria study. The age
of children control set was 11, 4 ± 8,2. (mean ± SD). The control
sets were matched by HLA-DRB1*04 appearance.
6.5
26.3
5.1
–
–
Biosystems, Foster City, CA, USA). The cDNA was amplified using
specific primers for HLA-DRB1*04 designed by Primer express®
as well as of Taqman®MGB (HLA-DRB1*04 gene forward primer
5´ACACCCGACCACGTTTCTTG 3´, HLA-DRB1*04 gene forward
primer
5´TCCGTCCCGTTGAAGAAATG
3´,
HLA-DRB1*04
Taqman®MGB probe 6-FAM-CACTCATGTTTAACCTGCT). Testing
of NFKB1 expression was done with assay on demand set of
primers and probes. As an internal control the human beta actin
was used (Applied Biosystems, Foster City, CA, USA). The 2 − Ct
[2-delta delta Ct] method was applied for relative quantification
(Livak & Schmittgen, 2001).
Statistical analysis
Allele and genotype frequencies, phenotype frequencies, and
the frequency of an allele in the chromosomal pool of population were determined by direct counting for all groups of patients
and controls. The genotype frequencies were tested for confirmation to Hardy-Weinberg equilibrium. For statistical establishment of significant differences genotype and allele distributions
were compared between two populations using 2 analysis, followed by Bonferroni correction for multiply comparisons. The
strength of the observed associations was estimated by calculating odds ratios [relative risk (RR)] using the method by Woolf.
The results from real time PCR were compared between the
groups by one-way ANOVA testing and also by nonparametric
Kruskall-Wallis statistics. The observed groups were also compared by Dunn’s multiple comparison tests. P-value < 0.05 was
considered significant.
Real time RT-PCR analysis
Results
&
The HLA association study
Peripheral blood with EDTA was collected by venipuncture, and
APCs were immediately separated by immunomagnetic separation by Dynabeads (Dynal HLA-class II, 210.04, Dynal, Oslo, Norway). Total RNA was extracted using the RNA blood mini kit
(Qiagen, Hilden, Germany). RNA was reverse transcribed by Taqman® real time PCR reagents. The quantitative real time RT-PCR
was performed in duplicates using the Taqman® PCR Kit in 96well microtiter plates on the ABI PRISM 7000 Sequence Detector
Systems, according to the manufacturer’s instructions (Applied
The HLA association study was performed to compare the risk
values between the tested groups. HLA-associations have shown
strong age-dependence (see Table 2). The relative risk values for
HLA-DQB1*0302 were almost three times higher in the group of
children with diabetes than the relative risk values in T1DM of
adults. On the other side HLA-DRB1*03 was associated with the
group of LADA patients two times stronger than with the group
of children diabetics. HLA-DRB1*04 was significantly increased
in both T1DM groups compared to controls and the power of
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
Article 127
Table 3
NFKB1
Frequencies of detected NFKB1 allele polymorphism in the tested groups
Lenght
Control Czech
Control Australia
LADA
T1DM children
T1DM adults
n=34 %
1.5
1.5
2.9
x
x
20.6
2.9
2.9
14.7
2.9
14.7
27.9
4.4
4.4
x
n=55 %
x
x
x
0.9
x
19.1
10.0
1.9
4.5
0.9
10.9
38.4
4.5
8.1
x
n=67 %
x
x
x
x
1.5
23.1
6.0
2.2
11.2
13.4*
12.7
24.6
2.5
0.7
x
(Curran et al., 2002)
gene
A01
A02
A03
A1
A2
A3
A4
A5
A6
A7
A8
A9
A10
A11
A12
(bp)
114
116
118
120
122
124
126
128
130
132
134
136
138
140
142
n=58 %
x
x
x
x
x
23.2
10.3
4.3
7.8
1.7
8.6
35.3
2.6
2.6
1.7
n=109 %
2.9
2.9
x
x
x
23.5
x
x
14.7
2.9
14.7
2.9
2.9
5.9
x
*
significant P value (P < 0.05)
association was comparable between these two groups. The
results of the study just confirm the previous results from the
HLA typing studies, but were needed in the relation to the further NFKB1 and NFKBIA polymorphism testing.
Distribution of NFKB1 polymorphism in Czech
population
15 alleles for the NFKB1 gene polymorphism (CA-repeats) in the
Czech population were detected so far. The alleles were ranging
in size from 114-142 bp corresponding to 10–25 CA repeats. The
complete results can be seen in Table 3. The most frequent allele
size in the patients and the control group was 136 bp (35.3 %)
corresponding to 22 CA repeats. The differences in allele frequencies, compared to healthy controls, were observed only in
the group of T1DM adult patients. The frequency of A7 allele
(size 132 bp, 20 repeats) was significantly increased in comparison with the control groups with a relative risk value of 1.88
(OR = 10.69, P < 0.01, CI = 95 %).
Distribution of 3´ UTR NFKBIA polymorphism in Czech
population
The genotype frequencies of the NFKBIA gene polymorphism in
the control and patient groups are presented in Table 4. There
was no difference in A and G allele frequency between the control group and patients, the difference was observed in distribution of genotypes between the patient groups. There is an
evidence that the heterozygous genotype of NFKBIA is protective
for diabetes onset in adulthood, according to the results in LADA
group (nonsignificantly) and T1DM adults (RR = 0.56, OR = 0.44,
P < 0.01, CI = 95 %). In the mentioned groups the frequency of the
heterozygotes is lowered and the frequency of the homozygous
genotypes rises. A significant difference was observed for AA
genotype in LADA group, with a relative risk value of 2.23
(OR = 2.68, P < 0.001, CI = 95 %) The AA and GG genotype frequencies in T1DM adults were increased, but with the border significance.
Expression studies of HLA-DRB1*04 and NFKB1
The results obtained in mRNA quantification studies can be seen
in Table 5 for HLA-DRB1*04 and in Table 6 for NFKB1 gene.
HLA-DRB1*04. gene expression was tested first among different
groups of patients. The significant difference was found only in
mRNA expression levels of HLA-DRB1*04 between the group of
T1DM children and the group of T1DM adults (P = 0.034). The
expression of HLA-DRB1*04 was significantly higher in T1DM
adult patients. The comparison was done with two specific control groups (positive for HLA-DRB1*04). Since there was no difference between the control groups with different age, we
compared the control groups as one control set. No difference in
the expression of HLA-DRB1*04 has been observed after analysis
among different HLA clas II, NFKBIA and NFKB1 genotypes.
No difference in the expression of NFKB1 gene has been observed
among different types of diabetic patient groups and even
after analysis based on the found HLA clas II, NFKBIA and NFKB1
genotypes.
Discussion
&
Development of type 1 diabetes requires coordinated expression
of genes responsible for initiation and progression of the disease, what is regulated by small number of transcription factors
including the Rel/NF-kB family. In our study we have compared
different groups of patients with autoimmune diabetes mellitus: 1) T1DM children 2) T1DM adults 3) LADA.
Based on the HLA - typing studies, we have concluded that HLA
– DR3 is significantly associated only with LADA (RR = 3.25) and
children diabetes (RR = 1.51). HLA – DQB1*0302 has three times
stronger relative risk for diabetes in childhood (RR = 8.98) than
for diabetes in adults (RR = 3.02). The results are in concordance
with previous testing in Czech population (Cerna et al., 2003;
Cinek et al., 2001; Novota et al., 2004). Surprisingly, there was
only a border significance of HLA-DQB1*0201 gene association in
our T1DM children group. This is a difference from the results
published in Czech population study by Cinek (Cinek et al.,
2001). Generally, the found difference could be explained by different number of patients involved into the study what could
change significance level.
Based on NFKB1 and NFKBIA polymorphism studies, we have
found out that A7 (132 bp) allele of NFKB1 gene presents a risk
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
128 Article
Table 4
Frequencies of NFKBIA genotypes in the tested groups
NFKBIA genotype
Control Czech
LADA
T1DM children
T1DM adults
GG (00)
AG (10)
AA (11)
n=159 %
27.0
55.6
16.9
n=31 %
22.6
45.2
32.3*
n=173 %
31.8
55.5
12.7
n=64 %
35.9
35.9*
28.0
*
significant P value (P < 0.05)
Table 5 Statistical analysis of the data from relative mRNA quantification of
the HLA-DRB1*04 gene
Table 6 Statistical analysis of the data from relative mRNA quantification of
the NFKB1 gene
Group
1
2
3
Group
1
2
3
DRB1*04
T1DM children
T1DM adults
Controls
NFKB1
T1DM children
T1DM adults
Controls
Number of
values
Minimum
Maximum
Mean
Std. Error
55
28
10
24
20
20
0,390
2,510
1,132
0,070
0,230
4,390
1,845*
0,221
0,600
1,410
1,050
0,105
Number of
values
Minimum
Maximum
Mean
Std. Error
0,458
2,460
1,120
0,127
0,150
3,370
0,991
0,274
0,400
3,000
1,335
0,155
*
significant P value (P&equals;0.034)
for our group of T1DM adults (RR = 1.88). In previous studies no
association with NFKB1 A7 allele was affirmed. Several reports
on the association study about NFKB1 and NFKBIA with T1DM,
T2DM, celiac disease, rheumatoid arthritis, systemic lupus erythematosus, breast cancer in a variety of ethnic groups (United
Kingdom, Denmark, Spain, Australia) exist (Curran et al., 2002;
Gylvin et al., 2002; Karban et al., 2004; Rueda et al., 2004; Smyth
et al., 2006). No T1DM association with any allele of the NFKB1
microsatellite marker could however be demonstrated in Danish
T1DM families as reported previously (Gylvin et al., 2002; Karban et al., 2004). In contrast, the frequency of the A10 (138 bp)
allele was significantly increased in patients with T1DM (0.17)
compared with the normal controls (0.02) (Hegazy et al., 2001).
However, the discrepancies in the frequencies of NFKB1 alleles
in control sets of populations with different genetic origin show
that it is not possible to compare the frequency of the risk alleles
among various ethnics. The exact mechanism underlying the
NFKB1 related disease susceptibility remains unknown. (Borm
et al., 2005).
It is known that 3´UTR is a regulatory region essential for the
appropriate gene expression of many genes. Any variation in
3´UTR of NFKBIA gene could alter the function of IB. A longlasting sustained activation of NFkB in the absence of decreased
IkB in mononuclear cells from patients with type 1 diabetes has
been reported (Bierhaus et al., 2001). The significant association
in NFKBIA polymorphism testing was found only for the AA
homozygous genotype in 3´UTR in the LADA group. Similarly the
heterozygous NFKBIA conformation seems to have significant
advantage for T1DM adults; otherwise it would be such frequent.
So far there is no evidence in literature about functional significance of our tested NFKB1 and NFKBIA polymorphisms. This was
a reason why we decided to test NFKB1 and HLA-DRB1*04-gene
expression on the mRNA level. There was not, however, any difference in the HLA-DRB1*04 and NFKB1 gene transcription
between groups with different NFKBIA and NFKB1 genotypes.
The mRNA transcript levels of HLA-DRB1*04 in circulating
peripheral blood mononuclear cells differ significantly only in
diabetic patients with different diabetes onset. We hypothesize
that the increased HLA-DRB1*04 mRNA expression could be a
protecting factor in a group of T1DM adults.
Summarizing our results we conclude that there is a high probability of association Rel/NF-B family genes with an autoimmune diabetes course, but the function of the genetic variations
needs to be examined further.
Acknowledgments
&
This survey was funded by the Research Program of Czech Ministry of Education and Youth: Identification code: MSM
0021620814 (Prevention, diagnostics and therapy of diabetes
mellitus, metabolic and endocrinal affections of organism). The
authors would also like to thank Anezka Koubova, Miroslava
Zaoralova and Ivana Spoljaricova for excellent technical assistance.
References
1 Arkan MC, Hevener AL, Greten FR, Maeda S, Li ZW, Long JM, WynshawBoris A, Poli G, Olefsky J, Karin M: IKK-beta links inflammation to obesity-induced insulin resistance. Nat Med 2005; 11: 191–198
2 Baldwin AS Jr.: Series introduction: the transcription factor NF-kappaB
and human disease. J Clin Invest 2001; 107: 3–6
3 Barnes PJ, Karin M: Nuclear factor-kappaB: a pivotal transcription factor in chronic inflammatory diseases. N Engl J Med 1997; 336: 1066–
1071
4 Bierhaus A, Schiekofer S, Schwaninger M, Andrassy M, Humpert PM,
Chen J, Hong M, Luther T, Henle T, Kloting I, Morcos M, Hofmann M,
Tritschler H, Weigle B, Kasper M, Smith M, Perry G, Schmidt AM, Stern
DM, Haring HU, Schleicher E, Nawroth PP: Diabetes-associated sustained activation of the transcription factor nuclear factor-kappaB.
Diabetes 2001; 50: 2792–2808
5 Borm ME, van Bodegraven AA, Mulder CJ, Kraal G, Bouma G: A NFKB1
promoter polymorphism is involved in susceptibility to ulcerative
colitis. Int J Immunogenet 2005; 32: 401–405
6 Cai D, Yuan M, Frantz DF, Melendez PA, Hansen L, Lee J, Shoelson SE:
Local and systemic insulin resistance resulting from hepatic activation
of IKK-beta and NF-kappaB. Nat Med 2005; 11: 183–190
7 Cerna M, Novota P, Kolostova K, Cejkova P, Zdarsky E, Novakova D,
Kucera P, Novak J, Andel M: HLA in Czech adult patients with autoim-
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
Article 129
8
9
10
11
12
13
14
15
16
17
18
19
mune diabetes mellitus: comparison with Czech children with type
1 diabetes and patients with type 2 diabetes. Eur J Immunogenet
2003; 30: 401–407
Chen F: Is NF-kappaB a culprit in type 2 diabetes? Biochem Biophys
Res Commun 2005; 332: 1–3
Cinek O, Kolouskova S, Snajderova M, Sumnik Z, Sedlakova P, Drevinek
P, Vavrinec J, Ronningen KS: HLA class II genetic association of type 1
diabetes mellitus in Czech children. Pediatr Diabetes 2001; 2: 98–
102
Csorba TR, Lyon AW: Abnormal proinsulin congeners as autoantigens
that initiate the pathogenesis of Type 1 diabetes. Med Hypotheses
2005; 64: 186–191
Curran JE, Weinstein SR, Griffiths LR: Polymorphic variants of NFKB1
and its inhibitory protein NFKBIA, and their involvement in sporadic
breast cancer. Cancer Lett 2002; 188: 103–107
Evans JL, Goldfine ID, Maddux BA, Grodsky GM: Oxidative stress and
stress-activated signaling pathways: a unifying hypothesis of type 2
diabetes. Endocr Rev 2002; 23: 599–622
Gylvin T, Bergholdt R, Nerup J, Pociot F: Characterization of a nuclearfactor-kappa B (NFkappaB) genetic marker in type 1 diabetes (T1DM)
families. Genes Immun 2002; 3: 430–432
Haslbeck KM, Schleicher E, Bierhaus A, Nawroth P, Haslbeck M, Neundorfer B, Heuss D: The AGE/RAGE/NF-(kappa)B pathway may contribute to the pathogenesis of polyneuropathy in impaired glucose
tolerance (IGT). Exp Clin Endocrinol Diabetes 2005; 113: 288–291
Hayden MS, Ghosh S: Signaling to NF-kappaB. Genes Dev 2004; 18:
2195–2224
Hegazy DM, O’Reilly DA, Yang BM, Hodgkinson AD, Millward BA,
Demaine AG: NFkappaB polymorphisms and susceptibility to type 1
diabetes. Genes Immun 2001; 2: 304–308
Heron E, Deloukas P, van Loon AP: The complete exon-intron structure
of the 156-kb human gene NFKB1, which encodes the p105 and p50
proteins of transcription factors NF-kappa B and I kappa B-gamma:
implications for NF-kappa B-mediated signal transduction. Genomics
1995; 30: 493–505
Jaini R, Kaur G, Mehra NK: Heterogeneity of HLA-DRB1*04 and its
associated haplotypes in the North Indian population. Hum Immunol
2002; 63: 24–29
Karban AS, Okazaki T, Panhuysen CI, Gallegos T, Potter JJ, Bailey-Wilson
JE, Silverberg MS, Duerr RH, Cho JH, Gregersen PK, Wu Y, Achkar JP,
Dassopoulos T, Mezey E, Bayless TM, Nouvet FJ, Brant SR: Functional
20
21
22
23
24
25
26
27
28
29
annotation of a novel NFKB1 promoter polymorphism that increases
risk for ulcerative colitis. Hum Mol Genet 2004; 13: 35–45
Lechler RI, Norcross MA, Germain RN: Qualitative and quantitative
studies of antigen-presenting cell function by using I-A-expressing L
cells. J Immunol 1985; 135: 2914–2922
Livak KJ, Schmittgen TD: Analysis of relative gene expression data using
real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 2001; 25: 402–408
Novota P, Cerna M, Kolostova K, Cejkova P, Zdarsky E, Novakova D,
Kucera P, Novak J, Andel M: Diabetes mellitus in adults: association of
HLA DRB1 and DQB1 diabetes risk alleles with GADab presence and
C-peptide secretion. Immunol Lett 2004; 95: 229–232
Ota N, Nakajima T, Shirai Y, Emi M: Isolation and radiation hybrid
mapping of a highly polymorphic CA repeat sequence at the human
nuclear factor kappa-beta subunit 1 (NFKB1) locus. J Hum Genet 1999;
44: 129–130
Rewers A, Babu S, Wang TB, Bugawan TL, Barriga K, Eisenbarth GS, Erlich
HA: Ethnic differences in the associations between the HLA-DRB1*04
subtypes and type 1 diabetes. Ann N Y Acad Sci 2003; 1005: 301–
309
Rueda B, Lopez-Nevot MA, Ruiz MP, Ortega E, Maldonado J, Lopez M,
Martin J: CA microsatellite polymorphism of the nuclear factor kappa
B1 gene in celiac disease. Eur J Immunogenet 2004; 31: 129–132
Shawkatova I, Fazekasova H, Michalkova D, Martinka E: Association of
type I diabetes mellitus with HLA alleles]. Bratisl Lek Listy 2000; 101:
624–625
Smyth DJ, Howson JM, Payne F, Maier LM, Bailey R, Holland K, Lowe CE,
Cooper JD, Hulme JS, Vella A, Dahlman I, Lam AC, Nutland S, Walker NM,
Twells RC, Todd JA: Analysis of polymorphisms in 16 genes in type 1
diabetes that have been associated with other immune-mediated
diseases. BMC Med Genet 2006; 7: 20
Wheat W, Kupfer R, Gutches DG, Rayat GR, Beilke J, Scheinman RI, Wegmann DR: Increased NF-kappa B activity in B cells and bone marrowderived dendritic cells from NOD mice. Eur J Immunol 2004; 34:
1395–1404
World Health Organization Definition Diagnosis and Classification of
Diabetes Mellitus and its Complications. Geneva, Switzerland: Department of Noncommunicable Disease Surveillance. WHO, 1999
Katarina K et al. HLA, NFKB1 and NFKBIA Gene Polymorphism … Exp Clin Endocrinol Diabetes 2007; 115:124–129
Original paper No. 9
EXTRAPITUITARY PROLACTIN PROMOTER POLYMORPHISM IN CZECH PATIENTS WITH
SYSTEMIC LUPUS ERYTHEMATOSUS AND RHEUMATOID ARTHRITIS
Fojtikova M, Cerna M, Cejkova P, Ruzickova S, Dostal C
Ann Rheum Dis. 2007 May;66(5):706-7
137
1
LETTER
Extrapituitary prolactin promoter polymorphism in Czech
patients with systemic lupus erythematosus and rheumatoid
arthritis
Markéta Fojtı́ková, Marie Černá, Pavlı́na Čejková, Šárka Růžičková, Ctibor Dostál
...................................................................................................................................
Ann Rheum Dis 2007;000:1–2. doi: 10.1136/ard.2006.061788
P
rolactin (PRL) and its production by lymphocytes have
been suggested to play a distinct role in the pathogenesis of
systemic lupus erythematosus (SLE) and rheumatoid
arthritis (RA).1 2 PRL acts as a cytokine and influences the
maturation and differentiation of immune cells.3 4
Extrapituitary PRL synthesis is regulated by an alternative
promoter,5 which contains a single-nucleotide polymorphism
(SNP) at the region 21149 G/T. Higher PRL mRNA expression
is associated with the G allele in lymphocytes.6 High frequency
of the G allele was described in patients with SLE,7 but was not
confirmed in other works.8
We investigated 21149 G/T SNP in 156 patients with SLE and
173 patients with RA, and in 123 healthy individuals (control
group). Patients with SLE consisted of: 134 (85.9%) women
and 22 (14.1%) men, with a mean age of 43.4 years. SLE
diagnosis was determined using the American College of
Rheumatology classification criteria.9 Patients with RA consisted of: 132 (76.3%) women and 41 (23.7%) men, with a
mean age of 57.4 years; all fulfilled the RA diagnostic criteria.10
The control group consisted of 40 (32.5%) women and 83
(67.5%) men, with a mean age of 38.7 years. The study was
approved by the ethics committee of The Third Medical Faculty,
Charles University, Prague, Czech Republic.
The PCR—reverse fragment length polymorphism method
was used for 21149 G/T SNP detection. During PCR, the 137
base pairs (bp) region of the PRL extrapituitary promoter was
amplified by using the following primers: forward 59GCAGGTCAAGATAACCTGGA
and
reverse
59CATCTCAGAGTTGAATTTATTTCCTT. For reverse fragment
length polymorphism, ApoI restriction endonucleases were
used. The genotypes identified were TT homozygote characterised by 120 and 17 bp, GG homozygote characterised by 85,
35 and 17 bp, and GT heterozygote characterised by 120, 85 and
35 bp + 17 bp DNA fragments. Results were evaluated by x2 test
with Bonferroni correction.
There was no difference in genotype and allele frequencies in
the SLE group compared with healthy Czech individuals
(table 1). Our results support previous Italian findings,8 but
differ from the UK report.7 However, with respect to the specific
organ manifestation of SLE (table 1), we detected an association between the G allele and articular involvement
(Pc = 0.0086, OR 2.56, 95% CI 1.51 to 4.33). Based on age
when SLE was diagnosed, we observed a GG genotype
frequency of 44.8% in the 21–40 years subgroup compared
with 15.8% and 24.0% in the ,20 years and .40 years
subgroup, respectively (Pc = 0.023, OR 2.94, 95% CI 1.43 to
5.96). Additionally, we correlated the presence of alleles and
genotypes of 21149 G/T SNP with antibodies against antinuclear antibody, double-stranded DNA, Sm, RNP, Ro and La,
but no connection was found (data not shown).
Annals of the Rheumatic Diseases
ar61788
Module 4 28/2/07 10:57:25
Significantly higher heterozygote GT genotype was detected
in the RA group compared with the controls (Pc = 0.039, OR
1.82, 95% CI 1.14 to 2.94). We observed no differences in
homozygote genotypes or in allele frequencies between patients
with RA and controls (table 1). Across the groups, no gender
differences in genotype distribution or allele frequencies were
identified (data not shown).
In conclusion, the presence of the G allele and GG genotype
of the PRL extrapituitary promoter 21149 G/T SNP is associated
with certain clinical features of SLE. The GT genotype is a
predisposing genetic factor for RA. Further investigation on this
is required.
ACKNOWLEDGEMENTS
This study was financially supported by the research programme of the
Ministry of Health of the Czech Republic—IGA MZ CR: NR/8191-3/
2004.
.......................
Authors’ affiliations
Markéta Fojtı́ková, Ctibor Dostál, Institute of Rheumatology, Na Slupi 4,
Prague, Czech Republic
Marie Černá, Pavlı́na Čejková, Department of Cell and Molecular Biology,
Third Faculty of Medicine, Charles University in Prague, Prague, Czech
Republic
Šárka Růžičková, Department of Molecular Biology and Immunogenetics,
Institute of Rheumatology, Na Slupi 4, Prague, Czech Republic
Competing interests: None declared.
Correspondence to: Dr M Fojtı́ková, Department of Cell and Molecular
Biology, Third Faculty of Medicine, Charles University in Prague, Ruská 87,
100 00 Prague 10, Czech Republic; [email protected]
Accepted 13 January 2007
REFERENCES
1 Leańos-Miranda A, Cárdenas-Mondragon G. Serum free prolactin
concentrations in patients with systemic lupus erythematosus are associated with
lupus activity. Rheumatology 2006;45:97–101.
2 Neidhart M, Gay RE, Gay S. Prolactin and prolactin-like polypeptides in
rheumatoid arthritis. Biomed Pharmacother 1999;53:218–22.
3 Matera L. Endocrine, paracrine and autocrine actions of prolactin on immune
cells. Life Sci 1996;59:599–614.
4 Grimaldi CM. Sex and systemic lupus erythematosus: the role of the sex
hormones estrogen and prolactin on the regulation of autoreactive B cells. Curr
Opin Rheumatol 2006;18:456–61.
5 Gellersen B, Kempf R, Telgmann R, DiMattia GE. Nonpituitary human prolactin
gene transcription is independent of Pit-1 and differentially controlled in
lymphocytes and in endometrial stroma. Mol Endocrinol 1994;8:356–73.
6 Stevens A, Ray D, Alansari A, Hajeer A, Thomson W, Donn R, et al.
Characterization of a prolactin gene polymorphism and its associations with
systemic lupus erythematosus. Arthritis Rheum 2001;44:2358–66.
Abbreviations: PRL, prolactin; RA, rheumatoid arthritis; SLE, systemic
lupus erythematosus; SNP, single-nucleotide polymorphism
www.annrheumdis.com
Topics:
2
Fojtı´ková , Černá , Čejková , et al
Table 1 Occurrence of genotype and allele frequencies of 21149 G/T single-nucleotide polymorphism of the extrapituitary
prolactin promoter in Czech patients with systemic lupus erythematosus (SLE) and rheumatoid arthritis and controls, and in patients
with SLE according to specific organ involvement
21149 G/T SNP of the extrapituitary prolactin promoter
Genotype
Subject groups
GG
SLE
Pc1
Specific organ involvement
Renal
With
Without
Pc2
With
Without
Pc2
With
Without
Pc2
With
Without
Pc2
With
Without
Pc2
With
Without
Pc2
Neuropsychiatric
Cardiac
Pulmonary
Articular
Dermal
RA
Pc3
Controls
Allele frequency
GT
TT
Number of cases %
%
%
156
34.60
NS
44.90
NS
63
93
31.10
37.60
NS
30.30
35.80
NS
28.60
36.40
NS
32.00
35.10
NS
41.00
15.40
NS
36.20
32.30
NS
32.40
NS
39.80
52.40
39.80
NS
51.50
43.10
NS
45.70
44.60
NS
48.00
44.30
NS
43.60
48.70
NS
45.70
43.50
NS
56.10
0.039*
41.50
33
123
35
121
25
131
117
39
94
62
173
123
G
T
20.50
NS
0.57
NS
0.43
NS
17.50
22.60
NS
18.20
21.10
NS
25.70
19.00
NS
20.00
20.60
NS
15.40
35.90
NS
18.10
24.20
NS
11.50
NS
18.70
0.56
0.57
NS
0.56
0.57
NS
0.51
0.59
NS
0.56
0.57
NS
0.63
0.40
0.0086
0.59
0.54
NS
0.60
NS
0.61
0.44
0.43
NS
0.44
0.43
NS
0.49
0.41
NS
0.44
0.43
NS
0.37
0.60
0.0086`
0.41
0.46
NS
0.40
NS
0.39
Pc, P corrected values; RA, rheumatoid arthritis; SLE, systemic lupus erythematosus; SNP, single-nucleotide polymorphism.
Pc values were calculated by the x2 test from 262 contingency tables of the separate genotype occurrence or allele frequency of (1) each patient with SLE and control
group (Pc1), (2) presence or absence for the clinical feature (Pc2) and (3) each patient with RA and control group (Pc3). Bonferroni correction for multiple comparisons
was used.
*OR 1.82, 95% CI 1.14 to 2.94.
OR 2.56, 95% CI 1.51 to 4.33.
`
OR 0.39, 95% CI 0.23 to 0.6.
7 Stevens A, Ray DW, Worthington J, Davis JR. Polymorphisms of the human
prolactin gene—implications for production of lymphocyte prolactin and systemic
lupus erythematosus. Lupus 2001;10:676–83.
8 Mellai M, Giordano M, D’Alfonso S, Marchini M, Scorza R, Giovanna Danieli M,
et al. Prolactin and prolactin receptor gene polymorphisms in multiple sclerosis
and systemic lupus erythematosus. Hum Immunol 2003;64:274–84.
9 Hochberg MC. Updating the American College of Rheumatology revised criteria
for the classification of systemic lupus erythematosus. Arthritis Rheum
1997;40:1725.
10 Arnett FC. Revised criteria for the classification of rheumatoid-arthritis. Bull
Rheum Dis 1989;38:1–6.
www.annrheumdis.com
Annals of the Rheumatic Diseases
ar61788
Module 4 28/2/07 10:57:25
Topics:
Original paper No. 10
POLYMORPHISM OF THE PROLACTIN EXTRAPITUITARY PROMOTER IN PSORIATIC
ARTHRITIS
Stolfa J, Fojtkova M, Cejkova P, Cerna M, Sedova L, Dostal C
Rheumatol Int. 2007 Sep;27(11):1095-6
140
Rheumatol Int (2007) 27:1095–1096
DOI 10.1007/s00296-007-0360-3
LE TTER TO THE EDIT OR
Polymorphism of the prolactin extrapituitary promoter
in psoriatic arthritis
Jilí Ktolfa · Markéta Fojtíková · Pavlína Bejková ·
Marie Berná · Liliana Kedová · Ctibor Dostál
Received: 24 January 2007 / Accepted: 20 April 2007 / Published online: 13 June 2007
© Springer-Verlag 2007
Psoriatic arthritis is characterized as seronegative arthritis
that aVects patients suVering from psoriasis [1]. Aetiology
of this condition is still not clear and therapeutic
approaches are not always successful. However, good
response to bromocriptine therapy decreasing prolactin levels in patients with psoriatic arthritis has been demonstrated
in some reports [2, 3]. Prolactin acts as a cytokine and plays
a role in pathogenesis of systemic autoimmune diseases
such as rheumatoid arthritis and systemic lupus erythematosus [4, 5]. Moreover, high serum prolactin levels were
observed in group of patients with psoriasis and link
between keratinocytes hyperproliferation and prolactin has
been proposed [6]. The peptide hormone prolactin is produced from pituitary lactotrophs and extrapituitary tissues
such as lymphocytes as well. Extrapituitary PRL production is regulated by an alternative promoter located 5.8 kb
upstream from the pituitary one [7, 8]. A functional single
nucleotide polymorphism (SNP) G/T at the position ¡1149
of this extrapituitary promoter has been observed and in
lymphocytes higher PRL mRNA expression found to be
connected with G allele [9]. In our work we studied
¡1149G/T SNP PRL in group of 83 Czech patients with
psoriatic arthritis (PsA).
We genotyped ¡1149G/T SNP in 83 PsA patients and
123 healthy individuals (control group). PsA patients: 43
(51.8%) females, 40 (48.2%) males, average age 54.1 years.
J. Ktolfa · M. Fojtíková (&) · L. Kedová · C. Dostál
Institute of Rheumatology, Na Slupi 4,
128 50 Prague 2, Czech Republic
e-mail: [email protected]
P. Bejková · M. Berná
Department of Cell and Molecular Biology,
Third Faculty of Medicine, Charles University in Prague,
Ruská 87, 100 00 Prague 10, Czech Republic
Control group: 40 (32.5%) females and 83 (67.5%) males,
average age 38.7 years. This study was approved by the
Ethical Committee of the Third Faculty of Medicine,
Charles University, Prague.
PCR-RFLP methodology was used for ¡1149G/T SNP
detection. PCR: The 137 bp region of the PRL extrapituitary promoter was ampliWed by employing following primers: Forward primer 5⬘ GCAGGTCAAGATAACCTGGA
and reverse primer 5⬘ CATCTCAGAGTTGAATTTATT
TCCTT. The PCR reaction was run under these conditions:
initial temperature 94°C for 2 min, then 32 cycles with
94°C for 17 s, 55°C for 17 s, 72°C for 17 s, and Wnal temperature 72°C for 1 min. RFLP: ApoI restriction endonuclease was used. The three genotypes were identiWed as
indicated: the homozygote TT was characterized by
120 + 17 bp, the homozygote GG by 85 + 35 + 17 bp, and
the heterozygote GT genotype by 120 + 85 + 35 + 17 bp
DNA fragments. Results were evaluated by Chi square test
with Bonferroni correction.
The genotypes and alleles frequency of ¡1149G/T SNP
PRL extrapituitary promoter were similar in group of PsA
patients and control group (Table 1). We did not Wnd any
diVerences in genotype and allele distribution between
healthy female and male and between female and male
patients with psoriatic arthritis, respectively (data not
shown). Moreover, we correlated genotype and allele distribution of ¡1149G/T SNP with clinical feature of psoriatic
arthritis, such as age of onset psoriasis and radiological
type of psoriatic arthritis. In Table 2 there are results of this
correlation, where no signiWcant association was detected.
In conclusion, ¡1149G/T SNP of the extrapituitary prolactin promoter is not associated with psoriatic arthritis (see
Table 1) and its characteristic such as the onset of skin
lesion and radiological type of arthritis (see Table 2). Our
Wndings could indicate that ¡1149G/T SNP prolactin
123
1096
Rheumatol Int (2007) 27:1095–1096
Table 1 Genotype and allele frequency of the ¡1149G/T SNP prolactin extrapituitary promoter in patients with psoriatic arthritis and
healthy controls
Genotype and allele
PsA N (%)
Controls N (%)
P corr.
GG
31 (37.3)
49 (39.8)
NS
GT
38 (45.8)
51 (41.5)
NS
TT
14 (16.9)
23 (18.7)
NS
G
100 (60.2)
149 (60.6)
NS
T
66 (39.8)
97 (39.4)
NS
SNP single nucleotide polymorphism, PRL prolactin, PsA psoriatic
arthritis, N number of cases, NS non signiWcant, P corr. P corrected
values were calculated by the Chi squared test from 2 £ 2 contingency
tables; Bonferroni correction for multiple comparisons was applied
Table 2 Genotype and allele frequency of the ¡1149G/T SNP
extrapituitary prolactin promoter and clinical features of psoriatic
arthritis
Type PsA
N
Genotype
Allele
frequency
GG (%)
GT (%)
TT (%)
G
T
Erosive
55
36.3
50.9
12.8
0.62
0.38
Non-erosive
28
39.3
35.7
25.0
0.57
0.43
P corr.
NS
NS
NS
NS
NS
Type Ia
58
36.2
50.0
13.8
0.61
0.39
Type IIb
22
40.9
31.8
27.3
0.57
0.43
P corr.
NS
NS
NS
NS
NS
Type PV
c
SNP single nucleotide polymorphism, PRL prolactin, PsA psoriatic
arthritis, PV psoriasis vulgaris, N number of cases, NS non signiWcant,
P corr. P corrected values were calculated by the Chi squared test from
2 £ 2 contingency tables; Bonferroni correction for multiple comparisons was used
a
Type I—onset of skin lesions before 40 years
b
Type II—onset of skin lesions after 40 years
c
3 patients were without skin lesions, but classiWed as psoriatic
arthritis
extrapituitary promoter does not belong to pathogenetic
factors of psoriatic arthritis, and thus conWrm the results of
the same analysis in juvenile idiopathic arthritis [10], but
123
not in SLE, that found G allele of ¡1149G/T SNP to be a
risk factor [8]. Possible alteration of prolactin levels in psoriatic arthritis might be caused by regulation of prolactin
release from pituitary gland. Nevertheless, further investigations for explaning the role of prolactin in psoriasis and
psoriatic arthritis pathogenesis are necessary.
Acknowledgment Supported by the Research program of the Ministry of health of Czech Republic: IGA MZ CR: NR/8191¡3/2004.
References
1. Wright V (1989) What is the best treatment approach for a patient
with the mutilating pattern of psoriatic peripheral arthritis. Br J
Rheumatol 28:382
2. Buskila D, Sukenik S, Holcberg G, Horowitz J (1991) Improvement of psoriatic arthritis in a patient treated with bromocriptine
for hyperprolactinemia. J Rheumatol 18:611–612
3. McMurray RW (2001) Bromocriptine in rheumatic and autoimmune diseases. Semin Arthritis Rheum 31:21–32
4. Giasuddin ASM, El Sherif AI, El Ojali SI (1998) Prolactin: does it
have a role in the pathogenesis of psoriasis? Dermatology
197:119–122
5. Leajos-Miranda A, Cárdenas-Mondragon G (2006) Serum free
prolactin concentrations in patients with systemic lupus erythematosus are associated with lupus activity. Rheumatology 45:97–
101
6. Nagafuchi H, Suzuki N, Kaneko A, Asai T, Sakane T (1999) Prolactin locally produced by synovium inWltrating T lymphocytes induces excessive synovial cell functions in patients with
rheumatoid arthritis. J Rheumatol 26:1890–1900
7. Gellersen B, Kempf R, Telgmann R, DiMattia GE (1994) Nonpituitary human prolactin gene transcription is independent of Pit¡1
and diVerentially controlled in lymphocytes and in endometrial
stroma. Mol Endocrinol 8:356–373
8. Stevens A, Ray D, Alansari A, Hajeer A, Thomson W, Donn R,
Ollier WE, Worthington J, Davis JR (2001) Characterization of a
prolactin gene polymorphism and its associations with systemic
lupus erythematosus. Arthritis Rheum 44:2358–2366
9. Stevens A, Ray DW, Worthington J, Davis JR (2001) Polymorphisms of the human prolactin gene—implications for production
of lymphocyte prolactin and systemic lupus erythematosus. Lupus
10:676–683
10. Donn RP, Farhan A, Stevans A, Ramanan A, Ollier WE, Thomson
W; British Paediatric Rheumatology Study Group (2002) Neuroendocrine gene polymorphisms and susceptibility to juvenile idiopathic arthritis. Rheumatology (Oxford) 41:930–936
CHAPTER 4
CONCLUSIONS
1.
Both NOTCH1 and ABL1 genes are afflicted by amplification of 9q34 locus in
ETL; however, NOTCH1 seems to be a primary target of genomic DNA
amplification. The ABL1 amplification detected by microsatellite analysis and
QPCR was positive in 38.5% (D9S290)/30.8% (D9S1847) and 50% cases,
respectively. The NOTCH1 amplification was positive in 15.4% (D9S158) and
65.4% cases (detected by QPCR).
2. A. i
The HLA alleles associated with T1D in children do not differ from those in
adult-onset T1D; however, in childhood they confer higher risk (HLADRB1*03: P < 0.0001, OR = 3.3 in children vs. P < 0.05, OR = 3.1 in adults;
HLA-DRB1*04: P < 0.0001, OR = 22.9 in children vs. P < 0.05, OR = 3.1 in
adults; HLA-DQB1*0302: P < 0.0001, OR = 46.5 in children vs. P = NS in
adults) suggesting a possible role of environmental factors, that increases with
age, and at least partly different immunogenetic etiology between autoimmune
DM manifested in childhood and in adults.
2. A. ii
Adult onset T1D and LADA have a partially different predisposition genes.
The HLA-DRB1*04, HLA-DRB1*03 (P < 0.05, OR = 3.1 for both) have been
detected as predisposing factors in T1D in adults, whilst the HLA-DRB1*03
allele (P < 0.01, OR = 3.4) has been a single risk factor for LADA.
There was the association of AA genotype of NFKBIA gene with LADA (P <
0.001 OR=2.7).
143
We did not confirm the association of the INS-VNTR (–23HphI A/T SNP)
with T1D. Similarly, there was no association of the Kir6.2 E23K SNP with
T2D.
The IDDM1 and IDDM2 diabetic loci have cumulative effect: the
simultaneous presence of the INS –23HphI AA genotype and HLA-DRB1*03
allele increases risk of T1D development almost two times (from OR = 3.1 for
presence of HLA-DRB1*03 to OR = 5.9 for presence of both markers).
2. A. iii
The INS –23HphI SNP A allele and AA genotype seem to be a risk factor for
production of C-peptide/insulin by β-cells of pancreas (P < 0.004, OR = 3.3
for A allele and OR = 3.54 for AA genotype). There was no association
between the polymorphism and GADA levels. In contrast, the HLA-DRB1*03
allele is a marker of GADA development (P < 0.0001, OR = 4.2).
2. B.
Analysis of the HLA-DRB1*04 expression in peripheral blood APCs revealed
significantly increased quantity of mRNA in T1D adults when compared to
T1D in children; however, no difference in HLA-DRB1*04 and NFKB1 gene
expression between T1D patients with different HLA, NFKB1, NFKBIA
genetic background was found.
2. C.
The analysis of PRL -1149 G/T polymorphism did not reveal any association
with systemic lupus erythematosus or psoriatic arthritis in the Czech
population, whereas significantly increased heterozygous genotype was
detected in patients with rheumatoid arthritis (P < 0.05). Nevertheless, in SLE,
G allele has been confirmed a risk factor for joint involvement (OR = 2.56, P
< 0.01).
144
REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
Dipple, K.M. and E.R. McCabe, Modifier genes convert "simple" Mendelian
disorders to complex traits. Mol Genet Metab, 2000. 71(1-2): p. 43-50.
Knight, J.C., Regulatory polymorphisms underlying complex disease traits. J Mol
Med, 2005. 83(2): p. 97-109.
Wray, G.A., Transcriptional regulation and the evolution of development. Int J
Dev Biol, 2003. 47(7-8): p. 675-84.
Gabriel, S.E., et al., Cystic fibrosis heterozygote resistance to cholera toxin in the
cystic fibrosis mouse model. Science, 1994. 266(5182): p. 107-9.
Poolman, E.M. and A.P. Galvani, Evaluating candidate agents of selective
pressure for cystic fibrosis. J R Soc Interface, 2007. 4(12): p. 91-8.
Batzer, M.A. and P.L. Deininger, Alu repeats and human genomic diversity. Nat
Rev Genet, 2002. 3(5): p. 370-9.
Lander, E.S., et al., Initial sequencing and analysis of the human genome.
Nature, 2001. 409(6822): p. 860-921.
Kruglyak, S., et al., Equilibrium distributions of microsatellite repeat length
resulting from a balance between slippage events and point mutations. Proc Natl
Acad Sci U S A, 1998. 95(18): p. 10774-8.
Toth, G., Z. Gaspari, and J. Jurka, Microsatellites in different eukaryotic
genomes: survey and analysis. Genome Res, 2000. 10(7): p. 967-81.
Blouin, M.S., et al., Use of microsatellite loci to classify individuals by
relatedness. Mol Ecol, 1996. 5(3): p. 393-401.
Goldstein, D.B., et al., An evaluation of genetic distances for use with
microsatellite loci. Genetics, 1995. 139(1): p. 463-71.
Gill, P., A.J. Jeffreys, and D.J. Werrett, Forensic application of DNA
'fingerprints'. Nature, 1985. 318(6046): p. 577-9.
Li, W.H., et al., Evolutionary analyses of the human genome. Nature, 2001.
409(6822): p. 847-9.
Smit, A.F., The origin of interspersed repeats in the human genome. Curr Opin
Genet Dev, 1996. 6(6): p. 743-8.
Stenger, J.E., et al., Biased distribution of inverted and direct Alus in the human
genome: implications for insertion, exclusion, and genome stability. Genome Res,
2001. 11(1): p. 12-27.
Santangelo, A.M., et al., Ancient exaptation of a CORE-SINE retroposon into a
highly conserved mammalian neuronal enhancer of the proopiomelanocortin
gene. PLoS Genet, 2007. 3(10): p. 1813-26.
Deininger, P.L. and M.A. Batzer, Mammalian retroelements. Genome Res,
2002. 12(10): p. 1455-65.
145
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
Esnault, C., J. Maestre, and T. Heidmann, Human LINE retrotransposons
generate processed pseudogenes. Nat Genet, 2000. 24(4): p. 363-7.
Allen, E., et al., High concentrations of long interspersed nuclear element
sequence distinguish monoallelically expressed genes. Proc Natl Acad Sci U S A,
2003. 100(17): p. 9940-5.
Bailey, J.A., et al., Segmental duplications: organization and impact within the
current human genome project assembly. Genome Res, 2001. 11(6): p. 1005-17.
Eichler, E.E., Recent duplication, domain accretion and the dynamic mutation of
the human genome. Trends Genet, 2001. 17(11): p. 661-9.
Samonte, R.V. and E.E. Eichler, Segmental duplications and the evolution of the
primate genome. Nat Rev Genet, 2002. 3(1): p. 65-72.
Jansen, R.C. and J.P. Nap, Genetical genomics: the added value from
segregation. Trends Genet, 2001. 17(7): p. 388-91.
Li, J. and M. Burmeister, Genetical genomics: combining genetics with gene
expression analysis. Hum Mol Genet, 2005. 14 Spec No. 2: p. R163-9.
Nussbaum, R.L.M., R. R.; Willard, H. F., THOMSON & THOMSON genetics in
medicine. 6th revised edition ed. Genetic Variation in Individuals: Mutation
and Polymorphism. 2004: Saunders
Nowell, P.C., The clonal evolution of tumor cell populations. Science, 1976.
194(4260): p. 23-8.
Leach, F.S., Microsatellite instability and prostate cancer: clinical and
pathological implications. Curr Opin Urol, 2002. 12(5): p. 407-11.
Wells, R.D., et al., Advances in mechanisms of genetic instability related to
hereditary neurological diseases. Nucleic Acids Res, 2005. 33(12): p. 3785-98.
Tippin, B., et al., To slip or skip, visualizing frameshift mutation dynamics for
error-prone DNA polymerases. J Biol Chem, 2004. 279(44): p. 45360-8.
Lupski, J.R., J.R. Roth, and G.M. Weinstock, Chromosomal duplications in
bacteria, fruit flies, and humans. Am J Hum Genet, 1996. 58(1): p. 21-7.
Bishop, A.J. and R.H. Schiestl, Homologous recombination as a mechanism of
carcinogenesis. Biochim Biophys Acta, 2001. 1471(3): p. M109-21.
Tutt, A., et al., Mutation in Brca2 stimulates error-prone homology-directed
repair of DNA double-strand breaks occurring between repeated sequences.
EMBO J, 2001. 20(17): p. 4704-16.
van Dartel, M. and T.J. Hulsebos, Amplification and overexpression of genes in
17p11.2 ~ p12 in osteosarcoma. Cancer Genet Cytogenet, 2004. 153(1): p. 77-80.
Jasin, M., Chromosome breaks and genomic instability. Cancer Invest, 2000.
18(1): p. 78-86.
Pfeiffer, P., W. Goedecke, and G. Obe, Mechanisms of DNA double-strand
break repair and their potential to induce chromosomal aberrations.
Mutagenesis, 2000. 15(4): p. 289-302.
Knutsen, T., Cytogenetic mechanisms in the pathogenesis and progression of
follicular lymphoma. Cancer Surv, 1997. 30: p. 163-92.
Lander, E.S. and N.J. Schork, Genetic dissection of complex traits. Science,
1994. 265(5181): p. 2037-48.
Risch, N.J., Searching for genetic determinants in the new millennium. Nature,
2000. 405(6788): p. 847-56.
146
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
Blumenfeld, A., et al., Localization of the gene for familial dysautonomia on
chromosome 9 and definition of DNA markers for genetic diagnosis. Nat Genet,
1993. 4(2): p. 160-4.
Georges, M., et al., Microsatellite mapping of a gene affecting horn development
in Bos taurus. Nat Genet, 1993. 4(2): p. 206-10.
Keim, P., et al., RFLP mapping in soybean: association between marker loci and
variation in quantitative traits. Genetics, 1990. 126(3): p. 735-42.
Glazier, A.M., J.H. Nadeau, and T.J. Aitman, Finding genes that underlie
complex traits. Science, 2002. 298(5602): p. 2345-9.
Darvasi, A. and A. Pisante-Shalom, Complexities in the genetic dissection of
quantitative trait loci. Trends Genet, 2002. 18(10): p. 489-91.
Risch, N. and K. Merikangas, The future of genetic studies of complex human
diseases. Science, 1996. 273(5281): p. 1516-7.
Weeks, D.E. and G.M. Lathrop, Polygenic disease: methods for mapping
complex disease traits. Trends Genet, 1995. 11(12): p. 513-9.
Roses, A.D., Pharmacogenetics and the practice of medicine. Nature, 2000.
405(6788): p. 857-65.
Long, A.D., et al., Genetic interactions between naturally occurring alleles at
quantitative trait loci and mutant alleles at candidate loci affecting bristle number
in Drosophila melanogaster. Genetics, 1996. 144(4): p. 1497-510.
Steinmetz, L.M., et al., Dissecting the architecture of a quantitative trait locus in
yeast. Nature, 2002. 416(6878): p. 326-30.
Christians, J.K. and P.D. Keightley, Behavioural genetics: finding genes that
cause complex trait variation. Curr Biol, 2005. 15(1): p. R19-21.
Tenesa, A., et al., Mapping quantitative trait loci using linkage disequilibrium:
marker- versus trait-based methods. Behav Genet, 2005. 35(2): p. 219-28.
Mott, R., et al., A method for fine mapping quantitative trait loci in outbred
animal stocks. Proc Natl Acad Sci U S A, 2000. 97(23): p. 12649-54.
Yalcin, B., et al., Genetic dissection of a behavioral quantitative trait locus shows
that Rgs2 modulates anxiety in mice. Nat Genet, 2004. 36(11): p. 1197-202.
Darvasi, A., Dissecting complex traits: the geneticists' "Around the world in 80
days". Trends Genet, 2005. 21(7): p. 373-6.
Schork, N.J., et al., Linkage disequilibrium analysis of biallelic DNA markers,
human quantitative trait loci, and threshold-defined case and control subjects.
Am J Hum Genet, 2000. 67(5): p. 1208-18.
Bennett, S.T. and J.A. Todd, Human type 1 diabetes and the insulin gene:
principles of mapping polygenes. Annu Rev Genet, 1996. 30: p. 343-70.
Isaacson, P.G.W., D.; Ralfkiaer, E.; Jaffe, E.S., Classification of Tumours:
Pathology and Genetics of Tumours of Haematopoietic and Lymphoid Tissues.
Enteropathy-type T-cell Lymphoma. 2001, Lyon: IARC Press. 208-209.
Zettl, A., et al., Enteropathy-type T-cell lymphoma. Am J Clin Pathol, 2007.
127(5): p. 701-6.
Isaacson, P. and D.H. Wright, Intestinal lymphoma associated with
malabsorption. Lancet, 1978. 1(8055): p. 67-70.
Isaacson, P. and D.H. Wright, Malignant histiocytosis of the intestine. Its
relationship to malabsorption and ulcerative jejunitis. Hum Pathol, 1978. 9(6): p.
661-77.
147
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
Harris, N.L., et al., A revised European-American classification of lymphoid
neoplasms: a proposal from the International Lymphoma Study Group. Blood,
1994. 84(5): p. 1361-92.
Isaacson, P.G. and M.Q. Du, Gastrointestinal lymphoma: where morphology
meets molecular biology. 2005/01/12 ed. J Pathol. Vol. 205. 2005. 255-74.
Carbonnel, F., et al., Are complicated forms of celiac disease cryptic T-cell
lymphomas? Blood, 1998. 92(10): p. 3879-86.
Halfdanarson, T.R., M.R. Litzow, and J.A. Murray, Hematologic
manifestations of celiac disease. Blood, 2007. 109(2): p. 412-21.
Chott, A., et al., Most CD56+ intestinal lymphomas are CD8+CD5-T-cell
lymphomas of monomorphic small to medium size histology. Am J Pathol, 1998.
153(5): p. 1483-90.
Chott, A., et al., Classification of intestinal T-cell neoplasms and their
differential diagnosis. Am J Clin Pathol, 1999. 111(1 Suppl 1): p. S68-74.
Zettl, A., et al., Chromosomal gains at 9q characterize enteropathy-type T-cell
lymphoma. Am J Pathol, 2002. 161(5): p. 1635-45.
Deleeuw, R.J., et al., Whole-genome analysis and HLA genotyping of
enteropathy-type T-cell lymphoma reveals 2 distinct lymphoma subtypes.
Gastroenterology, 2007. 132(5): p. 1902-11.
Pan, L., et al., Epstein-Barr virus (EBV) in enteropathy-associated T-cell
lymphoma (EATL). J Pathol, 1993. 170(2): p. 137-43.
Quintanilla-Martinez, L., et al., Primary non-Hodgkin's lymphoma of the
intestine: high prevalence of Epstein-Barr virus in Mexican lymphomas as
compared with European cases. Blood, 1997. 89(2): p. 644-51.
Ferguson, A. and K. Kingstone, Coeliac disease and malignancies. Acta
Paediatr Suppl, 1996. 412: p. 78-81.
Daum, S., et al., Refractory sprue syndrome with clonal intraepithelial
lymphocytes evolving into overt enteropathy-type intestinal T-cell lymphoma.
Digestion, 2000. 62(1): p. 60-5.
Daum, S., C. Cellier, and C.J. Mulder, Refractory coeliac disease. Best Pract
Res Clin Gastroenterol, 2005. 19(3): p. 413-24.
Verkarre, V., et al., Recurrent partial trisomy 1q22-q44 in clonal intraepithelial
lymphocytes in refractory celiac sprue. Gastroenterology, 2003. 125(1): p. 40-6.
Cellier, C., et al., Refractory sprue, coeliac disease, and enteropathy-associated
T-cell lymphoma. French Coeliac Disease Study Group. Lancet, 2000. 356(9225):
p. 203-8.
Baumgartner, A.K., et al., High frequency of genetic aberrations in enteropathytype T-cell lymphoma. Lab Invest, 2003. 83(10): p. 1509-16.
De Klein, A., et al., bcr rearrangement and translocation of the c-abl oncogene in
Philadelphia positive acute lymphoblastic leukemia. Blood, 1986. 68(6): p. 136975.
de Klein, A., et al., A cellular oncogene is translocated to the Philadelphia
chromosome in chronic myelocytic leukaemia. Nature, 1982. 300(5894): p. 765-7.
Jundt, F., et al., Activated Notch1 signaling promotes tumor cell proliferation
and survival in Hodgkin and anaplastic large cell lymphoma. Blood, 2002. 99(9):
p. 3398-403.
148
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
Ellisen, L.W., et al., TAN-1, the human homolog of the Drosophila notch gene, is
broken by chromosomal translocations in T lymphoblastic neoplasms. Cell, 1991.
66(4): p. 649-61.
Schlegelberger, B. and A.C. Feller, Classification of peripheral T-cell
lymphomas: cytogenetic findings support the updated Kiel classification. Leuk
Lymphoma, 1996. 20(5-6): p. 411-6.
Schlegelberger, B., et al., Cytogenetic findings in peripheral T-cell lymphomas as
a basis for distinguishing low-grade and high-grade lymphomas. Blood, 1994.
83(2): p. 505-11.
Tsukasaki, K., et al., Comparative genomic hybridization analysis in adult T-cell
leukemia/lymphoma: correlation with clinical course. Blood, 2001. 97(12): p.
3875-81.
Howell, W.M., et al., HLA-DRB, -DQA, and -DQB polymorphism in celiac
disease and enteropathy-associated T-cell lymphoma. Common features and
additional risk factors for malignancy. Hum Immunol, 1995. 43(1): p. 29-37.
Al-Toma, A., et al., Human leukocyte antigen-DQ2 homozygosity and the
development of refractory celiac disease and enteropathy-associated T-cell
lymphoma. Clin Gastroenterol Hepatol, 2006. 4(3): p. 315-9.
Vader, W., et al., The HLA-DQ2 gene dose effect in celiac disease is directly
related to the magnitude and breadth of gluten-specific T cell responses. Proc
Natl Acad Sci U S A, 2003. 100(21): p. 12390-5.
Amos, A.F., D.J. McCarty, and P. Zimmet, The rising global burden of diabetes
and its complications: estimates and projections to the year 2010. Diabet Med,
1997. 14 Suppl 5: p. S1-85.
Zimmet, P., The burden of type 2 diabetes: are we doing enough? Diabetes
Metab, 2003. 29(4 Pt 2): p. 6S9-18.
Diagnosis and classification of diabetes mellitus. Diabetes Care, 2007. 30 Suppl
1: p. S42-7.
Tuomi, T., Type 1 and type 2 diabetes: what do they have in common? Diabetes,
2005. 54 Suppl 2: p. S40-5.
Cnop, M., et al., Mechanisms of pancreatic beta-cell death in type 1 and type 2
diabetes: many differences, few similarities. Diabetes, 2005. 54 Suppl 2: p. S97107.
Palmer, J.P., et al., Is latent autoimmune diabetes in adults distinct from type 1
diabetes or just type 1 diabetes at an older age? Diabetes, 2005. 54 Suppl 2: p.
S62-7.
Pozzilli, P. and U. Di Mario, Autoimmune diabetes not requiring insulin at
diagnosis (latent autoimmune diabetes of the adult): definition, characterization,
and potential prevention. Diabetes Care, 2001. 24(8): p. 1460-7.
Hosszufalusi, N., et al., Similar genetic features and different islet cell
autoantibody pattern of latent autoimmune diabetes in adults (LADA) compared
with adult-onset type 1 diabetes with rapid progression. Diabetes Care, 2003.
26(2): p. 452-7.
Tuomi, T., et al., Clinical and genetic characteristics of type 2 diabetes with and
without GAD antibodies. Diabetes, 1999. 48(1): p. 150-7.
149
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
114.
Vatay, A., et al., Differences in the genetic background of latent autoimmune
diabetes in adults (LADA) and type 1 diabetes mellitus. Immunol Lett, 2002.
84(2): p. 109-115.
Weedon, M.N. and T.M. Frayling, Insights on pathogenesis of type 2 diabetes
from MODY genetics. Curr Diab Rep, 2007. 7(2): p. 131-8.
Barnett, A.H., et al., Diabetes in identical twins. A study of 200 pairs.
Diabetologia, 1981. 20(2): p. 87-93.
Kaprio, J., et al., Concordance for type 1 (insulin-dependent) and type 2 (noninsulin-dependent) diabetes mellitus in a population-based cohort of twins in
Finland. Diabetologia, 1992. 35(11): p. 1060-7.
Redondo, M.J., et al., Heterogeneity of type I diabetes: analysis of monozygotic
twins in Great Britain and the United States. Diabetologia, 2001. 44(3): p. 35462.
Gale, E.A., et al., Reanalysis of twin studies suggests that diabetes is mainly
genetic. BMJ, 2001. 323(7319): p. 997-8.
Srikanta, S., et al., Type I diabetes mellitus in monozygotic twins: chronic
progressive beta cell dysfunction. Ann Intern Med, 1983. 99(3): p. 320-6.
Awdeh, Z.L., et al., A genetic explanation for the rising incidence of type 1
diabetes, a polygenic disease. J Autoimmun, 2006. 27(3): p. 174-81.
Christau, B., et al., Incidence of insulin-dependent diabetes mellitus (0--29 years
at onset) in Denmark. Acta Med Scand Suppl, 1979. 624: p. 54-60.
Patrick, S.L., C.S. Moy, and R.E. LaPorte, The world of insulin-dependent
diabetes mellitus: what international epidemiologic studies reveal about the
etiology and natural history of IDDM. Diabetes Metab Rev, 1989. 5(7): p. 571-8.
Trevisan, R., M. Vedovato, and A. Tiengo, The epidemiology of diabetes
mellitus. Nephrol Dial Transplant, 1998. 13 Suppl 8: p. 2-5.
Yang, Z., et al., Epidemics of type 1 diabetes in China. Pediatr Diabetes, 2005.
6(3): p. 122-8.
Melanitou, E., Functional genomics in early autoimmunity. Ann N Y Acad Sci,
2005. 1050: p. 64-72.
Roep, B.O., The role of T-cells in the pathogenesis of Type 1 diabetes: from cause
to cure. Diabetologia, 2003. 46(3): p. 305-21.
Martin, S., et al., Development of type 1 diabetes despite severe hereditary Blymphocyte deficiency. N Engl J Med, 2001. 345(14): p. 1036-40.
Winter, W.E., N. Harris, and D. Schatz, Type 1 diabetes islet autoantibody
markers. Diabetes Technol Ther, 2002. 4(6): p. 817-39.
Esteban, L.M. and A.G. Baxter, Polyspecificity of autoimmune responses in type
1 (autoimmune) diabetes. Clin Exp Immunol, 2001. 126(2): p. 184-6.
Sigurs, N., et al., Prevalence of coeliac disease in diabetic children and
adolescents in Sweden. Acta Paediatr, 1993. 82(9): p. 748-51.
Collin, P., et al., Endocrinological disorders and celiac disease. Endocr Rev,
2002. 23(4): p. 464-83.
Kucera, P., et al., Gliadin, endomysial and thyroid antibodies in patients with
latent autoimmune diabetes of adults (LADA). Clin Exp Immunol, 2003. 133(1):
p. 139-43.
150
115.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
Karvonen, M., et al., A review of the recent epidemiological data on the
worldwide incidence of type 1 (insulin-dependent) diabetes mellitus. World Health
Organization DIAMOND Project Group. Diabetologia, 1993. 36(10): p. 883-92.
Rewers, M., et al., Trends in the prevalence and incidence of diabetes: insulindependent diabetes mellitus in childhood. World Health Stat Q, 1988. 41(3-4): p.
179-89.
LaPorte, R.E., et al., Geographic differences in the risk of insulin-dependent
diabetes mellitus: the importance of registries. Diabetes Care, 1985. 8 Suppl 1: p.
101-7.
Larenas, G., et al., Incidence of insulin-dependent diabetes mellitus in the IX
region of Chile: ethnic differences. Diabetes Res Clin Pract, 1996. 34 Suppl: p.
S147-51.
Verge, C.F., et al., Environmental factors in childhood IDDM. A populationbased, case-control study. Diabetes Care, 1994. 17(12): p. 1381-9.
Virtanen, S.M., et al., Infant feeding in Finnish children less than 7 yr of age
with newly diagnosed IDDM. Childhood Diabetes in Finland Study Group.
Diabetes Care, 1991. 14(5): p. 415-7.
Hanson, L.A., et al., Breast feeding: overview and breast milk immunology. Acta
Paediatr Jpn, 1994. 36(5): p. 557-61.
Lonnrot, M., et al., Onset of type 1 diabetes mellitus in infancy after enterovirus
infections. Diabet Med, 1998. 15(5): p. 431-4.
Syllaba, J., [New findings in type I diabetes]. Cas Lek Cesk, 1994. 133(2): p. 3740.
Prevention of diabetes mellitus. Report of a WHO Study Group. World Health
Organ Tech Rep Ser, 1994. 844: p. 1-100.
Roivainen, M., et al., Several different enterovirus serotypes can be associated
with prediabetic autoimmune episodes and onset of overt IDDM. Childhood
Diabetes in Finland (DiMe) Study Group. J Med Virol, 1998. 56(1): p. 74-8.
Honeyman, M.C., et al., Association between rotavirus infection and pancreatic
islet autoimmunity in children at risk of developing type 1 diabetes. Diabetes,
2000. 49(8): p. 1319-24.
Davies, J.L., et al., A genome-wide search for human type 1 diabetes
susceptibility genes. Nature, 1994. 371(6493): p. 130-6.
Cox, N.J., et al., Seven regions of the genome show evidence of linkage to type 1
diabetes in a consensus analysis of 767 multiplex families. Am J Hum Genet,
2001. 69(4): p. 820-30.
Nistico, L., et al., The CTLA-4 gene region of chromosome 2q33 is linked to, and
associated with, type 1 diabetes. Belgian Diabetes Registry. Hum Mol Genet,
1996. 5(7): p. 1075-80.
Van der Auwera, B.J., et al., CTLA-4 gene polymorphism confers susceptibility
to insulin-dependent diabetes mellitus (IDDM) independently from age and from
other genetic or immune disease markers. The Belgian Diabetes Registry. Clin
Exp Immunol, 1997. 110(1): p. 98-103.
Bahram, S. and T. Spies, The MIC gene family. Res Immunol, 1996. 147(5): p.
328-33.
Gambelunghe, G., et al., Association of MHC Class I chain-related A (MIC-A)
gene polymorphism with Type I diabetes. Diabetologia, 2000. 43(4): p. 507-14.
151
133.
134.
135.
136.
137.
138.
139.
140.
141.
142.
143.
144.
145.
146.
147.
148.
149.
Groh, V., et al., Recognition of stress-induced MHC molecules by intestinal
epithelial gammadelta T cells. Science, 1998. 279(5357): p. 1737-40.
Bauer, S., et al., Activation of NK cells and T cells by NKG2D, a receptor for
stress-inducible MICA. Science, 1999. 285(5428): p. 727-9.
Smyth, D., et al., Replication of an association between the lymphoid tyrosine
phosphatase locus (LYP/PTPN22) with type 1 diabetes, and evidence for its role
as a general autoimmunity locus. Diabetes, 2004. 53(11): p. 3020-3.
Hill, R.J., et al., The lymphoid protein tyrosine phosphatase Lyp interacts with
the adaptor molecule Grb2 and functions as a negative regulator of T-cell
activation. Exp Hematol, 2002. 30(3): p. 237-44.
Cloutier, J.F. and A. Veillette, Cooperative inhibition of T-cell antigen receptor
signaling by a complex between a kinase and a phosphatase. J Exp Med, 1999.
189(1): p. 111-21.
Hayden, M.S. and S. Ghosh, Signaling to NF-kappaB. Genes Dev, 2004. 18(18):
p. 2195-224.
Rothe, H., et al., IL-18 inhibits diabetes development in nonobese diabetic mice
by counterregulation of Th1-dependent destructive insulitis. J Immunol, 1999.
163(3): p. 1230-6.
Nakanishi, K., et al., Interleukin-18 is a unique cytokine that stimulates both Th1
and Th2 responses depending on its cytokine milieu. Cytokine Growth Factor
Rev, 2001. 12(1): p. 53-72.
McInnes, I.B., et al., Interleukin 18: a pleiotropic participant in chronic
inflammation. Immunol Today, 2000. 21(7): p. 312-5.
Permutt, M.A., J. Wasson, and N. Cox, Genetic epidemiology of diabetes. J Clin
Invest, 2005. 115(6): p. 1431-9.
Risch, N., Assessing the role of HLA-linked and unlinked determinants of
disease. Am J Hum Genet, 1987. 40(1): p. 1-14.
Trowsdale, J., J. Ragoussis, and R.D. Campbell, Map of the human MHC.
Immunol Today, 1991. 12(12): p. 443-6.
Hardy, D.A., et al., Mapping of the class II region of the human major
histocompatibility complex by pulsed-field gel electrophoresis. Nature, 1986.
323(6087): p. 453-5.
Huang, W., et al., Although DR3-DQB1*0201 may be associated with multiple
component diseases of the autoimmune polyglandular syndromes, the human
leukocyte antigen DR4-DQB1*0302 haplotype is implicated only in beta-cell
autoimmunity. J Clin Endocrinol Metab, 1996. 81(7): p. 2559-63.
Graham, J., et al., Negative association between type 1 diabetes and HLA
DQB1*0602-DQA1*0102 is attenuated with age at onset. Swedish Childhood
Diabetes Study Group. Eur J Immunogenet, 1999. 26(2-3): p. 117-27.
Pugliese, A., et al., HLA-DQB1*0602 is associated with dominant protection from
diabetes even among islet cell antibody-positive first-degree relatives of patients
with IDDM. Diabetes, 1995. 44(6): p. 608-13.
Todd, J.A., J.I. Bell, and H.O. McDevitt, HLA-DQ beta gene contributes to
susceptibility and resistance to insulin-dependent diabetes mellitus. Nature, 1987.
329(6140): p. 599-604.
152
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
Lee, K.H., K.W. Wucherpfennig, and D.C. Wiley, Structure of a human insulin
peptide-HLA-DQ8 complex and susceptibility to type 1 diabetes. Nat Immunol,
2001. 2(6): p. 501-7.
Harper, M.E., A. Ullrich, and G.F. Saunders, Localization of the human insulin
gene to the distal end of the short arm of chromosome 11. Proc Natl Acad Sci U
S A, 1981. 78(7): p. 4458-60.
Owerbach, D., et al., The insulin gene is located on the short arm of chromosome
11 in humans. Diabetes, 1981. 30(3): p. 267-70.
Junien, C. and V. van Heyningen, Report of the committee on the genetic
constitution of chromosome 11. Cytogenet Cell Genet, 1990. 55(1-4): p. 153-69.
Bell, G.I., M.J. Selby, and W.J. Rutter, The highly polymorphic region near the
human insulin gene is composed of simple tandemly repeating sequences.
Nature, 1982. 295(5844): p. 31-5.
Stead, J.D., et al., Influence of allele lineage on the role of the insulin
minisatellite in susceptibility to type 1 diabetes. Hum Mol Genet, 2000. 9(20): p.
2929-35.
Stead, J.D. and A.J. Jeffreys, Allele diversity and germline mutation at the
insulin minisatellite. Hum Mol Genet, 2000. 9(5): p. 713-23.
Vafiadis, P., et al., Class III alleles of the variable number of tandem repeat
insulin polymorphism associated with silencing of thymic insulin predispose to
type 1 diabetes. J Clin Endocrinol Metab, 2001. 86(8): p. 3705-10.
Barratt, B.J., et al., Remapping the insulin gene/IDDM2 locus in type 1 diabetes.
Diabetes, 2004. 53(7): p. 1884-9.
Perez De Nanclares, G., et al., 5'-Insulin gene VNTR polymorphism is specific
for type 1 diabetes: no association with celiac or Addison's disease. Ann N Y
Acad Sci, 2003. 1005: p. 319-23.
Guja, C., et al., Strong association of insulin gene INS-VNTR polymorphisms
with type 1 diabetes in the Romanian population. Rom J Intern Med, 2004.
42(2): p. 313-23.
Vafiadis, P., et al., Imprinted and genotype-specific expression of genes at the
IDDM2 locus in pancreas and leucocytes. J Autoimmun, 1996. 9(3): p. 397-403.
Vafiadis, P., et al., Insulin expression in human thymus is modulated by INS
VNTR alleles at the IDDM2 locus. Nat Genet, 1997. 15(3): p. 289-92.
Zielenski, J., et al., Skipping of exon 12 as a consequence of a point mutation
(1898 + 5G-->T) in the cystic fibrosis transmembrane conductance regulator gene
found in a consanguineous Chinese family. Clin Genet, 1995. 47(3): p. 125-32.
Kralovicova, J., et al., Variants in the human insulin gene that affect pre-mRNA
splicing: is -23HphI a functional single nucleotide polymorphism at IDDM2?
Diabetes, 2006. 55(1): p. 260-4.
Fodil, N., et al., Allelic repertoire of the human MHC class I MICA gene.
Immunogenetics, 1996. 44(5): p. 351-7.
Ota, M., et al., Trinucleotide repeat polymorphism within exon 5 of the MICA
gene (MHC class I chain-related gene A): allele frequency data in the nine
population groups Japanese, Northern Han, Hui, Uygur, Kazakhstan, Iranian,
Saudi Arabian, Greek and Italian. Tissue Antigens, 1997. 49(5): p. 448-54.
Perez-Rodriguez, M., et al., A new MICA allele with ten alanine residues in the
exon 5 microsatellite. Tissue Antigens, 2000. 55(2): p. 162-5.
153
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.
180.
181.
182.
183.
184.
185.
186.
Mizuki, N., et al., Triplet repeat polymorphism in the transmembrane region of
the MICA gene: a strong association of six GCT repetitions with Behcet disease.
Proc Natl Acad Sci U S A, 1997. 94(4): p. 1298-303.
Yabuki, K., et al., Triplet repeat polymorphism in the MICA gene in HLA-B27
positive and negative caucasian patients with ankylosing spondylitis. Hum
Immunol, 1999. 60(1): p. 83-6.
Gambelunghe, G., et al., Microsatellite polymorphism of the MHC class I chainrelated (MIC-A and MIC-B) genes marks the risk for autoimmune Addison's
disease. J Clin Endocrinol Metab, 1999. 84(10): p. 3701-7.
Gupta, M., et al., Association between the transmembrane region polymorphism
of MHC class I chain related gene-A and type 1 diabetes mellitus in Sweden.
Hum Immunol, 2003. 64(5): p. 553-61.
Bottini, N., et al., A functional variant of lymphoid tyrosine phosphatase is
associated with type I diabetes. Nat Genet, 2004. 36(4): p. 337-8.
Begovich, A.B., et al., A missense single-nucleotide polymorphism in a gene
encoding a protein tyrosine phosphatase (PTPN22) is associated with rheumatoid
arthritis. Am J Hum Genet, 2004. 75(2): p. 330-7.
Orozco, G., et al., Association of a functional single-nucleotide polymorphism of
PTPN22, encoding lymphoid protein phosphatase, with rheumatoid arthritis and
systemic lupus erythematosus. Arthritis Rheum, 2005. 52(1): p. 219-24.
Jagiello, P., et al., The PTPN22 620W allele is a risk factor for Wegener's
granulomatosis. Arthritis Rheum, 2005. 52(12): p. 4039-43.
Vandiedonck, C., et al., Association of the PTPN22*R620W polymorphism with
autoimmune myasthenia gravis. Ann Neurol, 2006. 59(2): p. 404-7.
Onengut-Gumuscu, S., et al., A functional polymorphism (1858C/T) in the
PTPN22 gene is linked and associated with type I diabetes in multiplex families.
Genes Immun, 2004. 5(8): p. 678-80.
Ladner, M.B., et al., Association of the single nucleotide polymorphism C1858T
of the PTPN22 gene with type 1 diabetes. Hum Immunol, 2005. 66(1): p. 60-4.
Zheng, W. and J.X. She, Genetic association between a lymphoid tyrosine
phosphatase (PTPN22) and type 1 diabetes. Diabetes, 2005. 54(3): p. 906-8.
Ghosh, S. and M. Karin, Missing pieces in the NF-kappaB puzzle. Cell, 2002.
109 Suppl: p. S81-96.
Bonizzi, G. and M. Karin, The two NF-kappaB activation pathways and their
role in innate and adaptive immunity. Trends Immunol, 2004. 25(6): p. 280-8.
May, M.J. and S. Ghosh, Rel/NF-kappa B and I kappa B proteins: an overview.
Semin Cancer Biol, 1997. 8(2): p. 63-73.
Karin, M., The beginning of the end: IkappaB kinase (IKK) and NF-kappaB
activation. J Biol Chem, 1999. 274(39): p. 27339-42.
Mercurio, F. and A.M. Manning, Multiple signals converging on NF-kappaB.
Curr Opin Cell Biol, 1999. 11(2): p. 226-32.
Evans, J.L., et al., Oxidative stress and stress-activated signaling pathways: a
unifying hypothesis of type 2 diabetes. Endocr Rev, 2002. 23(5): p. 599-622.
Ho, E. and T.M. Bray, Antioxidants, NFkappaB activation, and diabetogenesis.
Proc Soc Exp Biol Med, 1999. 222(3): p. 205-13.
154
187.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
203.
Baker, M.S., et al., Expression of a dominant negative inhibitor of NF-kappaB
protects MIN6 beta-cells from cytokine-induced apoptosis. J Surg Res, 2001.
97(2): p. 117-22.
Hegazy, D.M., et al., NFkappaB polymorphisms and susceptibility to type 1
diabetes. Genes Immun, 2001. 2(6): p. 304-8.
Bierhaus, A., et al., Diabetes-associated sustained activation of the transcription
factor nuclear factor-kappaB. Diabetes, 2001. 50(12): p. 2792-808.
Norlin, S., U. Ahlgren, and H. Edlund, Nuclear factor-{kappa}B activity in
{beta}-cells is required for glucose-stimulated insulin secretion. Diabetes, 2005.
54(1): p. 125-32.
Liuwantara, D., et al., Nuclear factor-kappaB regulates beta-cell death: a critical
role for A20 in beta-cell protection. Diabetes, 2006. 55(9): p. 2491-501.
Beg, A.A. and D. Baltimore, An essential role for NF-kappaB in preventing
TNF-alpha-induced cell death. Science, 1996. 274(5288): p. 782-4.
Karin, M., et al., NF-kappaB in cancer: from innocent bystander to major
culprit. Nat Rev Cancer, 2002. 2(4): p. 301-10.
Kucharczak, J., et al., To be, or not to be: NF-kappaB is the answer--role of
Rel/NF-kappaB in the regulation of apoptosis. Oncogene, 2003. 22(56): p. 896182.
Papa, S., et al., Linking JNK signaling to NF-kappaB: a key to survival. J Cell
Sci, 2004. 117(Pt 22): p. 5197-208.
Heron, E., P. Deloukas, and A.P. van Loon, The complete exon-intron structure
of the 156-kb human gene NFKB1, which encodes the p105 and p50 proteins of
transcription factors NF-kappa B and I kappa B-gamma: implications for NFkappa B-mediated signal transduction. Genomics, 1995. 30(3): p. 493-505.
Ota, N., et al., Isolation and radiation hybrid mapping of a highly polymorphic
CA repeat sequence at the human nuclear factor kappa-beta subunit 1 (NFKB1)
locus. J Hum Genet, 1999. 44(2): p. 129-30.
Klein, W., et al., A polymorphism of the NFKBIA gene is associated with Crohn's
disease patients lacking a predisposing allele of the CARD15 gene. Int J
Colorectal Dis, 2004. 19(2): p. 153-6.
Gloyn, A.L., et al., Large-scale association studies of variants in genes encoding
the pancreatic beta-cell KATP channel subunits Kir6.2 (KCNJ11) and SUR1
(ABCC8) confirm that the KCNJ11 E23K variant is associated with type 2
diabetes. Diabetes, 2003. 52(2): p. 568-72.
Altshuler, D., et al., The common PPARgamma Pro12Ala polymorphism is
associated with decreased risk of type 2 diabetes. Nat Genet, 2000. 26(1): p. 7680.
Florez, J.C., et al., Haplotype structure and genotype-phenotype correlations of
the sulfonylurea receptor and the islet ATP-sensitive potassium channel gene
region. Diabetes, 2004. 53(5): p. 1360-8.
Newman, B., et al., Concordance for type 2 (non-insulin-dependent) diabetes
mellitus in male twins. Diabetologia, 1987. 30(10): p. 763-8.
Poulsen, P., et al., Heritability of type II (non-insulin-dependent) diabetes
mellitus and abnormal glucose tolerance--a population-based twin study.
Diabetologia, 1999. 42(2): p. 139-45.
155
204.
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
220.
221.
O'Dea, K., et al., Impaired glucose tolerance, hyperinsulinemia, and
hypertriglyceridemia in Australian aborigines from the desert. Diabetes Care,
1988. 11(1): p. 23-9.
Sharp, P.S., et al., Insulin resistance in patients of Asian Indian and European
origin with non-insulin dependent diabetes. Horm Metab Res, 1987. 19(2): p. 845.
Smyth, S. and A. Heron, Diabetes and obesity: the twin epidemics. Nat Med,
2006. 12(1): p. 75-80.
Franco, L.J., Diabetes in Japanese-Brazilians--influence of the acculturation
process. Diabetes Res Clin Pract, 1996. 34 Suppl: p. S51-7.
Hara, H., G. Egusa, and M. Yamakido, Incidence of non-insulin-dependent
diabetes mellitus and its risk factors in Japanese-Americans living in Hawaii and
Los Angeles. Diabet Med, 1996. 13(9 Suppl 6): p. S133-42.
Ali, O., et al., Prevalence of NIDDM and impaired glucose tolerance in
aborigines and Malays in Malaysia and their relationship to sociodemographic,
health, and nutritional factors. Diabetes Care, 1993. 16(1): p. 68-75.
Al-Nuaim, A.R., Prevalence of glucose intolerance in urban and rural
communities in Saudi Arabia. Diabet Med, 1997. 14(7): p. 595-602.
Lerman, I.G., et al., The prevalence of diabetes and associated coronary risk
factors in urban and rural older Mexican populations. J Am Geriatr Soc, 1998.
46(11): p. 1387-95.
Kriska, A.M., et al., Association of physical activity and serum insulin
concentrations in two populations at high risk for type 2 diabetes but differing by
BMI. Diabetes Care, 2001. 24(7): p. 1175-80.
Helmrich, S.P., et al., Physical activity and reduced occurrence of non-insulindependent diabetes mellitus. N Engl J Med, 1991. 325(3): p. 147-52.
Kelley, D.E., et al., Dysfunction of mitochondria in human skeletal muscle in
type 2 diabetes. Diabetes, 2002. 51(10): p. 2944-50.
Shulman, G.I., Cellular mechanisms of insulin resistance. J Clin Invest, 2000.
106(2): p. 171-6.
Petersen, K.F., et al., Mitochondrial dysfunction in the elderly: possible role in
insulin resistance. Science, 2003. 300(5622): p. 1140-2.
Kreier, F., et al., "Diabetes of the elderly" and type 2 diabetes in younger
patients: possible role of the biological clock. Exp Gerontol, 2007. 42(1-2): p. 227.
Huang, Z., V. Cabanela, and T. Howell, Stress, bottlefeeding, and diabetes.
Lancet, 1997. 350(9081): p. 889.
Classification and diagnosis of diabetes mellitus and other categories of glucose
intolerance. National Diabetes Data Group. Diabetes, 1979. 28(12): p. 1039-57.
Guan, Y. and M.D. Breyer, Peroxisome proliferator-activated receptors
(PPARs): novel therapeutic targets in renal disease. Kidney Int, 2001. 60(1): p.
14-30.
Deeb, S.S., et al., A Pro12Ala substitution in PPARgamma2 associated with
decreased receptor activity, lower body mass index and improved insulin
sensitivity. Nat Genet, 1998. 20(3): p. 284-7.
156
222.
223.
224.
225.
226.
227.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
Hara, K., et al., The Pro12Ala polymorphism in PPAR gamma2 may confer
resistance to type 2 diabetes. Biochem Biophys Res Commun, 2000. 271(1): p.
212-6.
Horikawa, Y., et al., Genetic variation in the gene encoding calpain-10 is
associated with type 2 diabetes mellitus. Nat Genet, 2000. 26(2): p. 163-75.
Baier, L.J., et al., A calpain-10 gene polymorphism is associated with reduced
muscle mRNA levels and insulin resistance. J Clin Invest, 2000. 106(7): p. R6973.
Saxena, R., et al., Common single nucleotide polymorphisms in TCF7L2 are
reproducibly associated with type 2 diabetes and reduce the insulin response to
glucose in nondiabetic individuals. Diabetes, 2006. 55(10): p. 2890-5.
Lyssenko, V., et al., Mechanisms by which common variants in the TCF7L2 gene
increase risk of type 2 diabetes. J Clin Invest, 2007. 117(8): p. 2155-63.
Meigs, J.B., et al., Genome-wide association with diabetes-related traits in the
Framingham Heart Study. BMC Med Genet, 2007. 8 Suppl 1: p. S16.
Yi, F., P.L. Brubaker, and T. Jin, TCF-4 mediates cell type-specific regulation of
proglucagon gene expression by beta-catenin and glycogen synthase kinase3beta. J Biol Chem, 2005. 280(2): p. 1457-64.
Arkan, M.C., et al., IKK-beta links inflammation to obesity-induced insulin
resistance. Nat Med, 2005. 11(2): p. 191-8.
Boden, G., et al., Free fatty acids produce insulin resistance and activate the
proinflammatory nuclear factor-kappaB pathway in rat liver. Diabetes, 2005.
54(12): p. 3458-65.
Cai, D., et al., Local and systemic insulin resistance resulting from hepatic
activation of IKK-beta and NF-kappaB. Nat Med, 2005. 11(2): p. 183-90.
Itani, S.I., et al., Lipid-induced insulin resistance in human muscle is associated
with changes in diacylglycerol, protein kinase C, and IkappaB-alpha. Diabetes,
2002. 51(7): p. 2005-11.
Shoelson, S.E., J. Lee, and M. Yuan, Inflammation and the IKK beta/I kappa
B/NF-kappa B axis in obesity- and diet-induced insulin resistance. Int J Obes
Relat Metab Disord, 2003. 27 Suppl 3: p. S49-52.
White, M.F., The insulin signalling system and the IRS proteins. Diabetologia,
1997. 40 Suppl 2: p. S2-17.
Almind, K., et al., Aminoacid polymorphisms of insulin receptor substrate-1 in
non-insulin-dependent diabetes mellitus. Lancet, 1993. 342(8875): p. 828-32.
Porzio, O., et al., The Gly972-->Arg amino acid polymorphism in IRS-1 impairs
insulin secretion in pancreatic beta cells. J Clin Invest, 1999. 104(3): p. 357-64.
Groop, L.C., et al., Association between polymorphism of the glycogen synthase
gene and non-insulin-dependent diabetes mellitus. N Engl J Med, 1993. 328(1):
p. 10-4.
Orho-Melander, M., et al., A paired-sibling analysis of the XbaI polymorphism
in the muscle glycogen synthase gene. Diabetologia, 1999. 42(9): p. 1138-45.
Dunger, D.B., et al., Association of the INS VNTR with size at birth. ALSPAC
Study Team. Avon Longitudinal Study of Pregnancy and Childhood. Nat Genet,
1998. 19(1): p. 98-100.
157
240.
241.
242.
243.
244.
245.
246.
247.
248.
249.
250.
251.
252.
253.
254.
255.
Waterworth, D.M., et al., Linkage and association of insulin gene VNTR
regulatory polymorphism with polycystic ovary syndrome. Lancet, 1997.
349(9057): p. 986-90.
Huxtable, S.J., et al., Analysis of parent-offspring trios provides evidence for
linkage and association between the insulin gene and type 2 diabetes mediated
exclusively through paternally transmitted class III variable number tandem
repeat alleles. Diabetes, 2000. 49(1): p. 126-30.
Hansen, S.K., et al., Large-scale studies of the HphI insulin gene variablenumber-of-tandem-repeats polymorphism in relation to Type 2 diabetes mellitus
and insulin release. Diabetologia, 2004. 47(6): p. 1079-87.
Meigs, J.B., et al., The insulin gene variable number tandem repeat and risk of
type 2 diabetes in a population-based sample of families and unrelated men and
women. J Clin Endocrinol Metab, 2005. 90(2): p. 1137-43.
Ptacek, L.J., et al., Identification of a mutation in the gene causing hyperkalemic
periodic paralysis. Cell, 1991. 67(5): p. 1021-7.
Browne, D.L., et al., Episodic ataxia/myokymia syndrome is associated with point
mutations in the human potassium channel gene, KCNA1. Nat Genet, 1994. 8(2):
p. 136-40.
Ophoff, R.A., et al., Familial hemiplegic migraine and episodic ataxia type-2 are
caused by mutations in the Ca2+ channel gene CACNL1A4. Cell, 1996. 87(3): p.
543-52.
Hani, E.H., et al., Missense mutations in the pancreatic islet beta cell inwardly
rectifying K+ channel gene (KIR6.2/BIR): a meta-analysis suggests a role in the
polygenic basis of Type II diabetes mellitus in Caucasians. Diabetologia, 1998.
41(12): p. 1511-5.
Hart, L.M., et al., Reduced second phase insulin secretion in carriers of a
sulphonylurea receptor gene variant associating with Type II diabetes mellitus.
Diabetologia, 2000. 43(4): p. 515-9.
van Dam, R.M., et al., Common variants in the ATP-sensitive K+ channel genes
KCNJ11 (Kir6.2) and ABCC8 (SUR1) in relation to glucose intolerance:
population-based studies and meta-analyses. Diabet Med, 2005. 22(5): p. 590-8.
Porte, D., Jr. and S.E. Kahn, beta-cell dysfunction and failure in type 2 diabetes:
potential mechanisms. Diabetes, 2001. 50 Suppl 1: p. S160-3.
Kahn, S.E., The relative contributions of insulin resistance and beta-cell
dysfunction to the pathophysiology of Type 2 diabetes. Diabetologia, 2003. 46(1):
p. 3-19.
Yamada, Y., et al., Genomic variation in pancreatic ion channel genes in
Japanese type 2 diabetic patients. Diabetes Metab Res Rev, 2001. 17(3): p. 213-6.
Nichols, C.G., J.C. Koster, and M.S. Remedi, beta-cell hyperexcitability: from
hyperinsulinism to diabetes. Diabetes Obes Metab, 2007. 9 Suppl 2: p. 81-8.
Sagen, J.V., et al., Permanent neonatal diabetes due to mutations in KCNJ11
encoding Kir6.2: patient characteristics and initial response to sulfonylurea
therapy. Diabetes, 2004. 53(10): p. 2713-8.
Smith, A.J., et al., Molecular cell biology of KATP channels: implications for
neonatal diabetes. Expert Rev Mol Med, 2007. 9(21): p. 1-17.
158
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
267.
268.
269.
270.
271.
272.
273.
Vaxillaire, M., et al., Kir6.2 mutations are a common cause of permanent
neonatal diabetes in a large cohort of French patients. Diabetes, 2004. 53(10): p.
2719-22.
Schwanstecher, C., U. Meyer, and M. Schwanstecher, K(IR)6.2 polymorphism
predisposes to type 2 diabetes by inducing overactivity of pancreatic beta-cell
ATP-sensitive K(+) channels. Diabetes, 2002. 51(3): p. 875-9.
Guidelines for referral and management of systemic lupus erythematosus in
adults. American College of Rheumatology Ad Hoc Committee on Systemic
Lupus Erythematosus Guidelines. Arthritis Rheum, 1999. 42(9): p. 1785-96.
Johnson, A.E., et al., The prevalence and incidence of systemic lupus
erythematosus in Birmingham, England. Relationship to ethnicity and country of
birth. Arthritis Rheum, 1995. 38(4): p. 551-8.
Hart, H.H., R.R. Grigor, and D.E. Caughey, Ethnic difference in the prevalence
of systemic lupus erythematosus. Ann Rheum Dis, 1983. 42(5): p. 529-32.
McCarty, D.J., et al., Incidence of systemic lupus erythematosus. Race and
gender differences. Arthritis Rheum, 1995. 38(9): p. 1260-70.
Mok, C.C. and C.S. Lau, Lupus in Hong Kong Chinese. Lupus, 2003. 12(9): p.
717-22.
Worrall, J.G., et al., SLE: a rheumatological view. Analysis of the clinical
features, serology and immunogenetics of 100 SLE patients during long-term
follow-up. Q J Med, 1990. 74(275): p. 319-30.
Stohl, W., Systemic lupus erythematosus: a blissless disease of too much BLyS (B
lymphocyte stimulator) protein. Curr Opin Rheumatol, 2002. 14(5): p. 522-8.
Agnello, V., Association of systemic lupus erythematosus and SLE-like
syndromes with hereditary and acquired complement deficiency states. Arthritis
Rheum, 1978. 21(5 Suppl): p. S146-52.
Sturfelt, G., et al., Novel roles of complement in systemic lupus erythematosus-hypothesis for a pathogenetic vicious circle. J Rheumatol, 2000. 27(3): p. 661-3.
Lacki, J.K., et al., Cytokine concentration in serum of lupus erythematosus
patients: the effect on acute phase response. J Med, 1997. 28(1-2): p. 99-107.
Llorente, L., et al., Role of interleukin 10 in the B lymphocyte hyperactivity and
autoantibody production of human systemic lupus erythematosus. J Exp Med,
1995. 181(3): p. 839-44.
Jacob, C.O., et al., Heritable major histocompatibility complex class II-associated
differences in production of tumor necrosis factor alpha: relevance to genetic
predisposition to systemic lupus erythematosus. Proc Natl Acad Sci U S A, 1990.
87(3): p. 1233-7.
Herrera-Esparza, R., et al., Renal expression of IL-6 and TNFalpha genes in
lupus nephritis. Lupus, 1998. 7(3): p. 154-8.
Ohtsuka, K., et al., Cytokine-mediated down-regulation of B cell activity in SLE:
effects of interleukin-2 and transforming growth factor-beta. Lupus, 1999. 8(2):
p. 95-102.
Ohtsuka, K., et al., The relationship between defects in lymphocyte production of
transforming growth factor-beta1 in systemic lupus erythematosus and disease
activity or severity. Lupus, 1999. 8(2): p. 90-4.
Ohtsuka, K., et al., Decreased production of TGF-beta by lymphocytes from
patients with systemic lupus erythematosus. J Immunol, 1998. 160(5): p. 2539-45.
159
274.
275.
276.
277.
278.
279.
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
291.
292.
293.
Carlsten, H., et al., Oestrogen is a potent disease accelerator in SLE-prone MRL
lpr/lpr mice. Clin Exp Immunol, 1990. 80(3): p. 467-73.
Lucas, J.A., et al., Prevention of autoantibody formation and prolonged survival
in New Zealand black/New Zealand white F1 mice fed dehydroisoandrosterone. J
Clin Invest, 1985. 75(6): p. 2091-3.
Moszkorzova, L., et al., Hyperprolactinaemia in patients with systemic lupus
erythematosus. Clin Exp Rheumatol, 2002. 20(6): p. 807-12.
James, J.A., J.B. Harley, and R.H. Scofield, Role of viruses in systemic lupus
erythematosus and Sjogren syndrome. Curr Opin Rheumatol, 2001. 13(5): p.
370-6.
Sontheimer, R.D., Photoimmunology of lupus erythematosus and
dermatomyositis: a speculative review. Photochem Photobiol, 1996. 63(5): p. 58394.
Deapen, D., et al., A revised estimate of twin concordance in systemic lupus
erythematosus. Arthritis Rheum, 1992. 35(3): p. 311-8.
Pisetsky, D.S., G. Gilkeson, and E.W. St Clair, Systemic lupus erythematosus.
Diagnosis and treatment. Med Clin North Am, 1997. 81(1): p. 113-28.
Schur, P.H., Genetics of systemic lupus erythematosus. Lupus, 1995. 4(6): p.
425-37.
Woodrow, J.C., Immunogenetics of systemic lupus erythematosus. J Rheumatol,
1988. 15(2): p. 197-9.
Criswell, L.A. and C.I. Amos, Update on genetic risk factors for systemic lupus
erythematosus and rheumatoid arthritis. Curr Opin Rheumatol, 2000. 12(2): p.
85-90.
Chung, S.A. and L.A. Criswell, PTPN22: its role in SLE and autoimmunity.
Autoimmunity, 2007. 40(8): p. 582-90.
Vang, T., et al., Protein tyrosine phosphatase PTPN22 in human autoimmunity.
Autoimmunity, 2007. 40(6): p. 453-61.
Chistiakov, D.A. and A.P. Chistiakov, Is FCRL3 a new general autoimmunity
gene? Hum Immunol, 2007. 68(5): p. 375-83.
Prokunina, L., et al., A regulatory polymorphism in PDCD1 is associated with
susceptibility to systemic lupus erythematosus in humans. Nat Genet, 2002. 32(4):
p. 666-9.
Sullivan, K.E., Genetics of systemic lupus erythematosus. Clinical implications.
Rheum Dis Clin North Am, 2000. 26(2): p. 229-56, v-vi.
Atkinson, J.P., Complement activation and complement receptors in systemic
lupus erythematosus. Springer Semin Immunopathol, 1986. 9(2-3): p. 179-94.
Ben-Jonathan, N., et al., Extrapituitary prolactin: distribution, regulation,
functions, and clinical aspects. Endocr Rev, 1996. 17(6): p. 639-69.
Nagy, E. and I. Berczi, Hypophysectomized rats depend on residual prolactin for
survival. Endocrinology, 1991. 128(6): p. 2776-84.
Bole-Feysot, C., et al., Prolactin (PRL) and its receptor: actions, signal
transduction pathways and phenotypes observed in PRL receptor knockout mice.
Endocr Rev, 1998. 19(3): p. 225-68.
Yu-Lee, L.Y., Molecular actions of prolactin in the immune system. Proc Soc
Exp Biol Med, 1997. 215(1): p. 35-52.
160
294.
295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
308.
309.
310.
Matera, L., Endocrine, paracrine and autocrine actions of prolactin on immune
cells. Life Sci, 1996. 59(8): p. 599-614.
Grimaldi, C.M., Sex and systemic lupus erythematosus: the role of the sex
hormones estrogen and prolactin on the regulation of autoreactive B cells. Curr
Opin Rheumatol, 2006. 18(5): p. 456-61.
Yu-Lee, L.Y., et al., Interferon-regulatory factor 1 is an immediate-early gene
under transcriptional regulation by prolactin in Nb2 T cells. Mol Cell Biol, 1990.
10(6): p. 3087-94.
Nagafuchi, H., et al., Prolactin locally produced by synovium infiltrating T
lymphocytes induces excessive synovial cell functions in patients with rheumatoid
arthritis. J Rheumatol, 1999. 26(9): p. 1890-900.
Leanos-Miranda, A. and G. Cardenas-Mondragon, Serum free prolactin
concentrations in patients with systemic lupus erythematosus are associated with
lupus activity. Rheumatology (Oxford), 2006. 45(1): p. 97-101.
Neidhart, M., R.E. Gay, and S. Gay, Prolactin and prolactin-like polypeptides in
rheumatoid arthritis. Biomed Pharmacother, 1999. 53(5-6): p. 218-22.
Giasuddin, A.S., A.I. El-Sherif, and S.I. El-Ojali, Prolactin: does it have a role
in the pathogenesis of psoriasis? Dermatology, 1998. 197(2): p. 119-22.
Buskila, D., et al., Improvement of psoriatic arthritis in a patient treated with
bromocriptine for hyperprolactinemia. J Rheumatol, 1991. 18(4): p. 611-2.
McMurray, R.W., Bromocriptine in rheumatic and autoimmune diseases. Semin
Arthritis Rheum, 2001. 31(1): p. 21-32.
Gellersen, B., et al., Nonpituitary human prolactin gene transcription is
independent of Pit-1 and differentially controlled in lymphocytes and in
endometrial stroma. Mol Endocrinol, 1994. 8(3): p. 356-73.
Stevens, A., et al., Characterization of a prolactin gene polymorphism and its
associations with systemic lupus erythematosus. Arthritis Rheum, 2001. 44(10):
p. 2358-66.
Stevens, A., et al., Polymorphisms of the human prolactin gene--implications for
production of lymphocyte prolactin and systemic lupus erythematosus. Lupus,
2001. 10(10): p. 676-83.
Mellai, M., et al., Prolactin and prolactin receptor gene polymorphisms in
multiple sclerosis and systemic lupus erythematosus. Hum Immunol, 2003. 64(2):
p. 274-84.
Firestein, G.S., Evolving concepts of rheumatoid arthritis. Nature, 2003.
423(6937): p. 356-61.
Klareskog, L., et al., A new model for an etiology of rheumatoid arthritis:
smoking may trigger HLA-DR (shared epitope)-restricted immune reactions to
autoantigens modified by citrullination. Arthritis Rheum, 2006. 54(1): p. 38-46.
van Zeben, D., et al., Clinical significance of rheumatoid factors in early
rheumatoid arthritis: results of a follow up study. Ann Rheum Dis, 1992. 51(9):
p. 1029-35.
van Jaarsveld, C.H., et al., The prognostic value of the antiperinuclear factor,
anti-citrullinated peptide antibodies and rheumatoid factor in early rheumatoid
arthritis. Clin Exp Rheumatol, 1999. 17(6): p. 689-97.
161
311.
312.
313.
314.
315.
316.
317.
318.
319.
320.
321.
322.
323.
324.
325.
326.
327.
Kroot, E.J., et al., The prognostic value of anti-cyclic citrullinated peptide
antibody in patients with recent-onset rheumatoid arthritis. Arthritis Rheum,
2000. 43(8): p. 1831-5.
Schellekens, G.A., et al., The diagnostic properties of rheumatoid arthritis
antibodies recognizing a cyclic citrullinated peptide. Arthritis Rheum, 2000.
43(1): p. 155-63.
Sack, K.E.F., K. H., Medical Imunology. Rheumatic Diseases. 2001: The
McGraw-Hill Companies, Inc.
Edwards, J.C., G. Cambridge, and V.M. Abrahams, Do self-perpetuating B
lymphocytes drive human autoimmune disease? Immunology, 1999. 97(2): p.
188-96.
Alvarez-Lafuente, R., et al., Potential relationship between herpes viruses and
rheumatoid arthritis: analysis with quantitative real time polymerase chain
reaction. Ann Rheum Dis, 2005. 64(9): p. 1357-9.
Ferrell, P.B., et al., Seroepidemiological study of relationships between EpsteinBarr virus and rheumatoid arthritis. J Clin Invest, 1981. 67(3): p. 681-7.
Catalano, M.A., et al., Antibodies to Epstein-Barr virus-determined antigens in
normal subjects and in patients with seropositive rheumatoid arthritis. Proc Natl
Acad Sci U S A, 1979. 76(11): p. 5825-8.
Balandraud, N., J. Roudier, and C. Roudier, Epstein-Barr virus and rheumatoid
arthritis. Autoimmun Rev, 2004. 3(5): p. 362-7.
Padyukov, L., et al., A gene-environment interaction between smoking and
shared epitope genes in HLA-DR provides a high risk of seropositive rheumatoid
arthritis. Arthritis Rheum, 2004. 50(10): p. 3085-92.
MacGregor, A.J., et al., Characterizing the quantitative genetic contribution to
rheumatoid arthritis using data from twins. Arthritis Rheum, 2000. 43(1): p. 307.
Bali, D., et al., Genetic analysis of multiplex rheumatoid arthritis families. Genes
Immun, 1999. 1(1): p. 28-36.
Amos, C.I., et al., High-density SNP analysis of 642 Caucasian families with
rheumatoid arthritis identifies two new linkage regions on 11p12 and 2q33.
Genes Immun, 2006. 7(4): p. 277-86.
Plenge, R.M., et al., Replication of putative candidate-gene associations with
rheumatoid arthritis in >4,000 samples from North America and Sweden:
association of susceptibility with PTPN22, CTLA4, and PADI4. Am J Hum
Genet, 2005. 77(6): p. 1044-60.
Hall, F.C., et al., Influence of the HLA-DRB1 locus on susceptibility and severity
in rheumatoid arthritis. QJM, 1996. 89(11): p. 821-9.
Jawaheer, D., et al., Dissecting the genetic complexity of the association between
human leukocyte antigens and rheumatoid arthritis. Am J Hum Genet, 2002.
71(3): p. 585-94.
Lee, A.T., et al., The PTPN22 R620W polymorphism associates with RF positive
rheumatoid arthritis in a dose-dependent manner but not with HLA-SE status.
Genes Immun, 2005. 6(2): p. 129-33.
Plenge, R.M., et al., TRAF1-C5 as a risk locus for rheumatoid arthritis--a
genomewide study. N Engl J Med, 2007. 357(12): p. 1199-209.
162
328.
329.
330.
331.
332.
333.
334.
335.
336.
337.
338.
339.
340.
341.
342.
343.
344.
345.
346.
Haas, C.S., et al., Identification of genes modulated in rheumatoid arthritis using
complementary DNA microarray analysis of lymphoblastoid B cell lines from
disease-discordant monozygotic twins. Arthritis Rheum, 2006. 54(7): p. 2047-60.
Bowcock, A.M. and W.O. Cookson, The genetics of psoriasis, psoriatic arthritis
and atopic dermatitis. Hum Mol Genet, 2004. 13 Spec No 1: p. R43-55.
Wuepper, K.D., S.N. Coulter, and A. Haberman, Psoriasis vulgaris: a genetic
approach. J Invest Dermatol, 1990. 95(5): p. 2S-4S.
Duffy, D.L., L.S. Spelman, and N.G. Martin, Psoriasis in Australian twins. J
Am Acad Dermatol, 1993. 29(3): p. 428-34.
Hammer, R.E., et al., Spontaneous inflammatory disease in transgenic rats
expressing HLA-B27 and human beta 2m: an animal model of HLA-B27associated human disorders. Cell, 1990. 63(5): p. 1099-112.
Khare, S.D., H.S. Luthra, and C.S. David, Spontaneous inflammatory arthritis
in HLA-B27 transgenic mice lacking beta 2-microglobulin: a model of human
spondyloarthropathies. J Exp Med, 1995. 182(4): p. 1153-8.
Bardos, T., et al., Mice lacking endogenous major histocompatibility complex
class II develop arthritis resembling psoriatic arthritis at an advanced age.
Arthritis Rheum, 2002. 46(9): p. 2465-75.
Lories, R.J., et al., Ankylosing enthesitis, dactylitis, and onychoperiostitis in male
DBA/1 mice: a model of psoriatic arthritis. Ann Rheum Dis, 2004. 63(5): p. 5958.
Cook, P.W., et al., Transgenic expression of the human amphiregulin gene
induces a psoriasis-like phenotype. J Clin Invest, 1997. 100(9): p. 2286-94.
Peters, B.P., F.G. Weissman, and M.A. Gill, Pathophysiology and treatment of
psoriasis. Am J Health Syst Pharm, 2000. 57(7): p. 645-59; quiz 660-1.
Baker, B.S., et al., Induction of cutaneous lymphocyte-associated antigen
expression by group A streptococcal antigens in psoriasis. Arch Dermatol Res,
1997. 289(12): p. 671-6.
Mallon, E., et al., HLA-C and guttate psoriasis. Br J Dermatol, 2000. 143(6): p.
1177-82.
Breuer-McHam, J.N., et al., Distinct serum cytokines in AIDS-related skin
diseases. Viral Immunol, 1998. 11(4): p. 215-20.
Anandarajah, A.P. and C.T. Ritchlin, Pathogenesis of psoriatic arthritis. Curr
Opin Rheumatol, 2004. 16(4): p. 338-43.
Alenius, G.M., et al., Analysis of 6 genetic loci for disease susceptibility in
psoriatic arthritis. J Rheumatol, 2004. 31(11): p. 2230-5.
Martin, M.P., et al., Cutting edge: susceptibility to psoriatic arthritis: influence of
activating killer Ig-like receptor genes in the absence of specific HLA-C alleles. J
Immunol, 2002. 169(6): p. 2818-22.
Suzuki, Y., et al., Genetic polymorphisms of killer cell immunoglobulin-like
receptors are associated with susceptibility to psoriasis vulgaris. J Invest
Dermatol, 2004. 122(5): p. 1133-6.
Nelson, G.W., et al., Cutting edge: heterozygote advantage in autoimmune
disease: hierarchy of protection/susceptibility conferred by HLA and killer Ig-like
receptor combinations in psoriatic arthritis. J Immunol, 2004. 173(7): p. 4273-6.
Rahman, P., et al., CARD15: a pleiotropic autoimmune gene that confers
susceptibility to psoriatic arthritis. Am J Hum Genet, 2003. 73(3): p. 677-81.
163
347.
348.
349.
350.
351.
352.
353.
354.
355.
356.
357.
358.
359.
360.
361.
362.
363.
364.
365.
366.
Queiro-Silva, R., et al., The effect of HLA-DR antigens on the susceptibility to,
and clinical expression of psoriatic arthritis. Scand J Rheumatol, 2004. 33(5): p.
318-22.
Grubic, Z., et al., The MICA-A4 triplet repeats polymorphism in the
transmembrane region confers additional risk for development of psoriatic
arthritis in the Croatian population. Eur J Immunogenet, 2004. 31(2): p. 93-8.
Rahman, P. and J.T. Elder, Genetic epidemiology of psoriasis and psoriatic
arthritis. Ann Rheum Dis, 2005. 64 Suppl 2: p. ii37-9; discussion ii40-1.
Corrao, G., et al., Mortality in patients with coeliac disease and their relatives: a
cohort study. Lancet, 2001. 358(9279): p. 356-61.
Green, P.H., et al., Risk of malignancy in patients with celiac disease. Am J Med,
2003. 115(3): p. 191-5.
Catassi, C., et al., Risk of non-Hodgkin lymphoma in celiac disease. JAMA,
2002. 287(11): p. 1413-9.
Freeman, H.J., Lymphoproliferative and intestinal malignancies in 214 patients
with biopsy-defined celiac disease. J Clin Gastroenterol, 2004. 38(5): p. 429-34.
Cooper, B.T., et al., Celiac disease and malignancy. Medicine (Baltimore), 1980.
59(4): p. 249-61.
Fasano, A., et al., Prevalence of celiac disease in at-risk and not-at-risk groups in
the United States: a large multicenter study. Arch Intern Med, 2003. 163(3): p.
286-92.
Maki, M., et al., Prevalence of Celiac disease among children in Finland. N Engl
J Med, 2003. 348(25): p. 2517-24.
Dieterich, W., et al., Identification of tissue transglutaminase as the autoantigen
of celiac disease. Nat Med, 1997. 3(7): p. 797-801.
O'Farrelly, C., et al., Humoral response to wheat protein in patients with coeliac
disease and enteropathy associated T cell lymphoma. Br Med J (Clin Res Ed),
1986. 293(6552): p. 908-10.
Holmes, G.K., et al., Malignancy in coeliac disease--effect of a gluten free diet.
Gut, 1989. 30(3): p. 333-8.
Shaoul, R. and A. Lerner, Associated autoantibodies in celiac disease.
Autoimmun Rev, 2007. 6(8): p. 559-65.
Stepniak, D. and F. Koning, Celiac disease--sandwiched between innate and
adaptive immunity. Hum Immunol, 2006. 67(6): p. 460-8.
Djilali-Saiah, I., et al., CTLA-4 gene polymorphism is associated with
predisposition to coeliac disease. Gut, 1998. 43(2): p. 187-9.
Naluai, A.T., et al., The CTLA4/CD28 gene region on chromosome 2q33 confers
susceptibility to celiac disease in a way possibly distinct from that of type 1
diabetes and other chronic inflammatory disorders. Tissue Antigens, 2000. 56(4):
p. 350-5.
Hunt, K.A., et al., A common CTLA4 haplotype associated with coeliac disease.
Eur J Hum Genet, 2005. 13(4): p. 440-4.
Brophy, K., et al., Haplotypes in the CTLA4 region are associated with coeliac
disease in the Irish population. Genes Immun, 2006. 7(1): p. 19-26.
Rueda, B., et al., Association of MICA-A5.1 allele with susceptibility to celiac
disease in a family study. Am J Gastroenterol, 2003. 98(2): p. 359-62.
164
367.
368.
369.
Reines, B.P., Hypothesis. Bystanders or bad seeds? Many autoimmune-target
cells may be transforming to cancer and signalling "danger" to the immune
system. Autoimmunity, 2001. 33(2): p. 121-34.
Mention, J.J., et al., Interleukin 15: a key to disrupted intraepithelial lymphocyte
homeostasis and lymphomagenesis in celiac disease. Gastroenterology, 2003.
125(3): p. 730-45.
Chen, Y.G., et al., Cellular expression requirements for inhibition of type 1
diabetes by a dominantly protective major histocompatibility complex haplotype.
Diabetes, 2007. 56(2): p. 424-30.
165