ArticlePDF Available

Overexpression of a splice variant of DNA methyltransferase 3b, DNMT3b4, associated with DNA hypomethylation on pericentromeric satellite regions during human hepatocarcinogenesis

Authors:

Abstract and Figures

DNA hypomethylation on pericentromeric satellite regions is an early and frequent event associated with heterochromatin instability during human hepatocarcinogenesis. A DNA methyltransferase, DNMT3b, is required for methylation on pericentromeric satellite regions during mouse development. To clarify the molecular mechanism underlying DNA hypomethylation on pericentromeric satellite regions during human hepatocarcinogenesis, we examined mutations of the DNMT3b gene and mRNA expression levels of splice variants of DNMT3b in noncancerous liver tissues showing chronic hepatitis and cirrhosis, which are considered to be precancerous conditions, and in hepatocellular carcinomas (HCCs). Mutation of the DNMT3b gene was not found in HCCs. Overexpression of DNMT3b4, a splice variant of DNMT3b lacking conserved methyltransferase motifs IX and X, significantly correlated with DNA hypomethylation on pericentromeric satellite regions in precancerous conditions and HCCs (P = 0.0001). In particular, the ratio of expression of DNMT3b4 to that of DNMT3b3, which is the major splice variant in normal liver tissues and retains conserved methyltransferase motifs I, IV, VI, IX, and X, showed significant correlation with DNA hypomethylation (P = 0.009). Transfection of human epithelial 293 cells with DNMT3b4 cDNA induced DNA demethylation on satellite 2 in pericentromeric heterochromatin DNA. These results suggest that overexpression of DNMT3b4, which may lack DNA methyltransferase activity and compete with DNMT3b3 for targeting to pericentromeric satellite regions, results in DNA hypomethylation on these regions, even in precancerous stages, and plays a critical role in human hepatocarcinogenesis by inducing chromosomal instability.
Content may be subject to copyright.
Overexpression of a splice variant of DNA
methyltransferase 3b, DNMT3b4, associated
with DNA hypomethylation on pericentromeric
satellite regions during human hepatocarcinogenesis
Yoshimasa Saito*
, Yae Kanai*, Michiie Sakamoto*, Hidetsugu Saito
, Hiromasa Ishii
, and Setsuo Hirohashi*
*Pathology Division, National Cancer Center Research Institute, 5-1-1 Tsukiji, Chuo-ku, Tokyo 104-0045, Japan; and Department of Internal Medicine,
Keio University School of Medicine, 35 Shinanomachi, Shinjyuku-ku, Tokyo 160-8582, Japan
Edited by Stanley M. Gartler, University of Washington, Seattle, WA, and approved May 31, 2002 (received for review March 1, 2002)
DNA hypomethylation on pericentromeric satellite regions is an early
and frequent event associated with heterochromatin instability dur-
ing human hepatocarcinogenesis. A DNA methyltransferase,
DNMT3b, is required for methylation on pericentromeric satellite
regions during mouse development. To clarify the molecular mech-
anism underlying DNA hypomethylation on pericentromeric satellite
regions during human hepatocarcinogenesis, we examined muta-
tions of the DNMT3b gene and mRNA expression levels of splice
variants of DNMT3b in noncancerous liver tissues showing chronic
hepatitis and cirrhosis, which are considered to be precancerous
conditions, and in hepatocellular carcinomas (HCCs). Mutation of the
DNMT3b gene was not found in HCCs. Overexpression of DNMT3b4,
a splice variant of DNMT3b lacking conserved methyltransferase
motifs IX and X, significantly correlated with DNA hypomethylation
on pericentromeric satellite regions in precancerous conditions and
HCCs (P0.0001). In particular, the ratio of expression of DNMT3b4
to that of DNMT3b3, which is the major splice variant in normal liver
tissues and retains conserved methyltransferase motifs I, IV, VI, IX,
and X, showed significant correlation with DNA hypomethylation
(P0.009). Transfection of human epithelial 293 cells with DNMT3b4
cDNA induced DNA demethylation on satellite 2 in pericentromeric
heterochromatin DNA. These results suggest that overexpression of
DNMT3b4, which may lack DNA methyltransferase activity and com-
pete with DNMT3b3 for targeting to pericentromeric satellite regions,
results in DNA hypomethylation on these regions, even in precan-
cerous stages, and plays a critical role in human hepatocarcinogenesis
by inducing chromosomal instability.
DNA methylation plays important roles in gene silencing,
chromatin remodeling, and genome stability (1– 4). Aber-
rant DNA methylation is one of the most consistent epigenetic
changes in human cancers (1–4). Generally, the overall level of
DNA methylation is lower in cancer cells than in normal cells (5,
6), although a number of tumor suppressor genes are silenced by
DNA methylation on CpG islands around their promoter regions
in cancer cells (1–4, 7, 8).
We have carefully examined alterations of DNA methylation
status on pericentromeric satellite regions and CpG islands of
specific genes, and expression of DNA methyltransferases and
methyl-CpG-binding proteins, in noncancerous liver tissues show-
ing chronic hepatitis and cirrhosis, which are considered to be
precancerous conditions (9, 10), and in hepatocellular carcinomas
(HCCs) (11–17). Among these alterations, DNA hypomethylation
on pericentromeric satellite regions was detected even in precan-
cerous conditions and appears to be one of the earliest epigenetic
changes during human hepatocarcinogenesis (17). Satellite regions
are located in pericentromeric heterochromatin DNA, and DNA
hypomethylation on these regions is known to result in centromeric
decondensation, enhancing chromosome recombinations (18, 19).
In fact, frequent chromosome 1q copy gain with a pericentromeric
breakpoint was reported in HCCs showing DNA hypomethylation
on satellite 2 (20). DNA hypomethylation on pericentromeric
satellite regions may induce chromosomal instability during hepa-
tocarcinogenesis, even in precancerous conditions.
A newly identified DNA methyltransferase, DNMT3b (21), is
specifically required for DNA methylation on pericentromeric
satellite regions in embryonic stem cells and early mouse embryos
(22). Germ-line mutations of the DNMT3b gene have been re-
ported in patients with immunodeficiencycentromeric instability
facial anomalies syndrome (23, 24), a rare recessive autosomal
disorder characterized by DNA hypomethylation on pericentro-
meric satellite regions (25). These findings encouraged us to
examine genetic alterations of the DNMT3b gene in HCCs.
Down-regulation of DNMT3b is unlikely to underlie DNA
hypomethylation on pericentromeric satellite regions during human
hepatocarcinogenesis, because we and other groups have reported
increased expression of the mRNA for DNMT3b in human cancers,
including HCCs (17, 26, 27). However, four splice variants of human
DNMT3b in the C-terminal catalytic domain are known (ref. 26,
Fig. 1). DNMT3b4 and DNMT3b5, without conser ved methyltrans-
ferase motifs IX and X, probably lack DNA methyltransferase
activity. So far, no studies of the expression of DNMT3b in human
cancers have discriminated between the splice variants.
To clarify the molecular mechanism underlying DNA hypom-
ethylation on pericentromeric satellite regions during human hepa-
tocarcinogenesis, we examined mutations of the DNMT3b gene and
expression levels of splice variants of DNMT3b in noncancerous
liver tissues showing chronic hepatitis and cirrhosis and in HCCs.
Materials and Methods
Patients and Tissue Specimens. Fifty-nine primary HCCs and the
corresponding noncancerous liver tissues were obtained from sur-
gically resected materials from 49 patients (cases H1 to H49) who
were treated at the National Cancer Center Hospital, Tokyo.
Twelve patients were hepatitis B virus surface antigen (HBs-Ag)-
positive, 29 were anti-hepatitis C virus antibody (anti-HCV)-
positive, one was both HBs-Ag and anti-HCV positive, and seven
were both HBs-Ag and anti-HCV negative. Histological examina-
tion of noncancerous liver tissues from HCC patients revealed no
remarkable findings, findings compatible with chronic hepatitis,
and findings compatible with cirrhosis in 1, 24, and 24 tissues,
respectively. For comparison, normal liver tissues showing no
remarkable histological findings were also obtained from eight
patients (cases C1 to C8) who were both HBs-Ag and anti-HCV
negative and underwent partial hepatectomy for liver metastasis of
primary colon cancer.
This paper was submitted directly (Track II) to the PNAS office.
Abbreviations: DNMT3b, DNA methyltransferase 3b; HCC, hepatocellular carcinoma; RT, re-
verse transcription; GAPDH, glyceraldehyde-3-phosphate dehydrogenase.
To whom reprint requests should be addressed. E-mail: shirohas@ncc.go.jp.
10060–10065
PNAS
July 23, 2002
vol. 99
no. 15 www.pnas.orgcgidoi10.1073pnas.152121799
PCR–Single-Strand Conformation Polymorphism Analysis and Direct
Sequencing. Genomic DNA was amplified by PCR with 22
rhodamine-labeled intronic oligonucleotide primer sets encom-
passing all coding exons of the DNMT3b gene (GenBank accession
no. AL035071 and Table 1). Primers were designed on the basis of
the exon-intron boundaries as described (28). PCR products were
electrophoresed on 6% polyacrylamide gel. DNA was recovered
from the mobility-shifted bands and sequenced in both directions
with the ABI PRISM BigDye Terminator kit and an A BI PRISM
310 Genetic Analyzer (Applied Biosystems).
Reverse-Transcription (RT)-PCR and Cloning Sequencing. First-strand
cDNA was prepared from total RNA with random
hexadeoxynucleotide primers and SuperScript RNase H
reverse
transcriptase (GIBCOBRL). cDNA derived from human testis
total RNA (CLONTECH) was used as control. Subsequent PCR
with a primer set of (forward) 5-CCT GCT GAA TTA CTC ACG
CCC C-3and (reverse) 5-GTC TGT GTA GTG CAC AGG AA A
GCC-3(26) amplified all four splice variants in the C-terminal
catalytic domain of the DNMT3b gene. For visual confirmation, the
PCR products were separated electrophoretically on 1.5% agarose
gel. The PCR products were then cloned into the pCRII vector (TA
cloning kit; Invitrogen). The inserts were amplified by colony PCR
and sequenced with both vector primers: M13 forward, 5-GTA
AAA CGA CGG CCA G-3; M13 reverse, 5-CAG GAA ACA
GCT ATG AC-3.
Splice Variant-Specific Quantitative RT-PCR. Oligonucleotide primer
sets specific for DNMT3b3 and DNMT3b4 were designed. Each
primer spans the splice variant-specific exon-exon boundary: 3b3
forward, 5-GAT GAA CAG GAT CT T TGG CTT T-3(exon
2023); 3b3 reverse, 5-GCC TGG CTG GAA CTA TTC ACA-3
(exon 23); 3b4 forward, 5-CGG GAT GAA CAG T TA AAG
AAA GTA C-3(exon 2022); 3b4 reverse, 5-CCA AAG ATC
CTT TCG AGC TC-3(exon 2223). The PCRs were performed
with the SYBR Green PCR Core Reagents kit (Applied Biosys-
tems). Real-time detection of the emission intensity of SYBR
Green bound to double-stranded DNAs was performed with the
ABI PRISM 7700 Sequence Detection System. cDNAs derived
from the HCC cell line Alexander (29) were used as the calibrator
samples. Quantitative PCRs were performed in triplicate for each
sample-primer set, and the mean of the three experiments was used
as the relative quantification value. At the end of 40 PCR cycles, the
reaction products were separated electrophoretically on 3% aga-
rose gel to confirm that no nonspecific product was obtained at each
amplification.
Transfection with DNMT3b4 cDNA. Full-length human DNMT3b4
cDNA was prepared by RT-PCR of total RNA from the Alexander
cell line. The primer set (forward, 5-CGC GGA TCC TGG AA A
GCA TGA AGG GAG ACA C-3; reverse, 5-GCT CTA GAA
CTG TTC ATC CCG GGT AGG TTG-3) was designed such that
after a BamHIXbaI double digest the coding sequence of
DNMT3b4 could be ligated in-frame into the ex pression vector with
a myc tag at the C terminus (pcDNA3.1-myc; Invitrogen). To
confirm the fidelity of the PCR, the insert was fully sequenced. The
cloned cDNA was linearized and transfected into 293 cellshuman
embryonic kidney cells transformed with adenovirus type 5 DNA
(30)using FuGENE 6 Transfection Reagent (Roche Diagnos-
tics). After G418 selection of resistant cells, myc-tagged
DNMT3b4-expressing clones were chosen by Western blotting.
Mock-transfected 293 cells were generated by transfection with
vector alone.
Western Blotting. Total cell extracts were separated by SDSPAGE,
transferred electrophoretically onto poly(vinylidene dif luoride) fil-
Fig. 1. mRNA structure and protein structure of splice variants of human
DNMT3b. The four splice variants in the C-terminal catalytic domain (26) and the
numbers of exons and the exon-intron boundaries (28) have been described.
Table 1. The primer sets for PCRsingle-strand conformation polymorphism to detect mutation of the DNMT3b gene
Target exons Primer sets Target exons Primer sets
Exon 2 5-TCCCTGCTTCCCTTTCACC-3(sense) Exon 13 5-ACCCCAGGCTTTAGCAGCT-3(sense)
5-TTGTCTGCAGTGACCGCTC-3(antisense) 5-GGAGTTAGAGGAGGCAAGGG-3(antisense)
Exon 3 5-TTAGCAAGGCCGTTCCC-3(sense) Exon 14 5-CCCTTCTCTGGTCTCCGATT-3(sense)
5-CCACGTGATGAAAGCCAAAG-3(antisense) 5-GACTGCAGGAACGTAGGAGC-3(antisense)
Exon 4 5-TGACTTGCTGATACCCTGGG-3(sense) Exon 15 5-CAGGAGACCAGCTCTGACAAAG-3(sense)
5-GAGTGTTGGCCACAAGTGCT-3(antisense) 5-TTTCCAACACCCTGTGCC-3(antisense)
Exon 5 5-AGGCCTCCAGTCACCTAAGG-3(sense) Exon 16 5-GTCTTTGCCCTGTGCCTTC-3(sense)
5-TGCAGTGAGTGAGGCATATCTC-3(antisense) 5-CCTGGCTACCCTGTTGTGAC-3(antisense)
Exon 6 5-TTGCTCTGGCCCAAACTATG-3(sense) Exon 17 5-GAACTGTTCATTTTACCATAGCAGG-3(sense)
5-GGTCCTCAGTCACCCTGG-3(antisense) 5-GGGAAAAAGACAGGAAGAGATG-3(antisense)
Exon 7 5-CCTCTCCTCACTGGGATTTCTT-3(sense) Exon 18 5-TCTAGAAGTGGGTCCAGCTCTC-3(sense)
5-TTCAAAGGGAGGCAGGC-3(antisense) 5-AAGCAAGTGGCTCCTTCTCAG-3(antisense)
Exon 8 5-AGACATGGCACCTGGGACA-3(sense) Exon 19 5-GAACCTGCTGGTCTCAGGGA-3(sense)
5-TCTTGCTTCATCCCTGCTCT-3(antisense) 5-CCTGCCCTGGCTGTCTGT-3(antisense)
Exon 9 5-CACCCCCCCATTCATCA-3(sense) Exon 20 5-CTGCAAAGGTCTGGTTGACAC-3(sense)
5-GACTCTCCCAAGAAGTGGTCC-3(antisense) 5-CCAGGTCTTTCTAGGAAGGCTT-3(antisense)
Exon 10 5-TGGGTGACAGAGCAAGACC-3(sense) Exon 21 5-TGTGCCTAGCAGAGGACCCT-3(sense)
5-CAGAAGAAAGTGCATAGAAAACAGG-3(antisense) 5-AGTCCCCACTTGGAGGTCAC-3(antisense)
Exon 11 5-ACCCAGGCATAGCATGGTCT-3(sense) Exon 22 5-TTGAGCCTTGACTCCCCAG-3(sense)
5-TGATCTGCAGCGTCCTCCT-3(antisense) 5-GCCCATGTCTGCCCATTT-3(antisense)
Exon 12 5-GGAATTGATCTGTACCCGGC-3(sense) Exon 23 5-GGTTGAGGCTGTCAACATCC-3(sense)
5-TGGGTTAAACCCTGACAGGG-3(antisense) 5-ATCACACCTCCTGGGTCCTG-3(antisense)
Saito et al. PNAS
July 23, 2002
vol. 99
no. 15
10061
MEDICAL SCIENCES
ters, and then incubated with anti-myc mAb 9E10 (Santa Cruz
Biotechnology).
Southern Blotting. High molecular weight DNA (5
g) was digested
for 24 h with 10 units of either MspIorHpaII, which cut at the
sequence CCGG, per
g of DNA. HpaII does not cut when the
internal cytosine is methylated. The DNA fragments were sepa-
rated by electrophoresis, transferred to nitrocellulose membranes,
and hybridized with
32
P-labeled oligonucleotide probe for satellite
2 (31). Signal intensities were measured with an image analyzer
(model BAS-2500; Fujifilm, Tokyo).
Statistics. Correlations between mRNA levels for DNMT3b splice
variants and DNA methylation status were analyzed with the
KruskalWallis test. Differences in ratios for signal intensities
of Southern blotting were analyzed with the MannWhitney U
test. Differences with Pvalues of less than 0.05 were considered
significant.
Results
No Mutation of the
DNMT3b
Gene in HCCs. Fig. 2 shows examples of
PCRsingle-strand conformation polymorphism analysis of the
DNMT3b gene in HCC cases. Because all of the shifted bands were
detected in both the noncancerous liver tissue and the HCC of a
particular case, they were considered to be polymorphisms. Se-
quencing revealed that the polymorphisms were located in introns
2 and 13 near the exon-intron boundaries (Table 2). No mutation
of any coding exon of the DNMT3b gene was detected in 59 HCCs.
DNMT3b3 and DNMT3b4 Are Major Splice Variants in HCCs. Fig. 3A
shows examples of RT-PCR products obtained with a primer set
that amplifies all four splice variants of DNMT3b in the C-terminal
catalytic domain. cDNA derived from total RNA of human testis,
in which significant expression of all four splice variants has been
reported (26), was used as control. At the end of 35 cycles of PCR
amplification, PCR products of about 230 bp were visible in almost
all samples (band d in Fig. 3A). PCR products of about 350 bp were
clearly visible in HCCs and Alexander cells, but control testis
contained little of this product (band b in Fig. 3A). RT-PCR
products of sample H35T1 were then cloned into the pCRII vector
and the inserts were amplified by colony PCR (Fig. 3B). Almost all
clones examined possessed inserts of about 230 bp (band d) or 350
bp (band b). Sequencing confirmed that the inserts of about 230 bp
(band d) and 350 bp (band b) corresponded to DNMT3b3, skipping
exons 21 and 22, and DNMT3b4, skipping exon 21, respectively (ref.
26, Figs. 1 and 3C).
Overexpression of DNMT3b4 Significantly Correlates with DNA Hy-
pomethylation on Pericentromeric Satellite Regions in Precancerous
Conditions and HCCs. To accurately quantify levels of mRNA for
DNMT3b3 and DNMT3b4, which were the major splice variants in
HCCs (Fig. 3), splice variant-specific quantitative RTPCR was
performed in eight normal liver tissues from patients with liver
metastasis of primary colon cancer and in 49 noncancerous liver
tissues and 59 HCCs from 49 HCC patients.
We have previously reported the DNA methylation status on
pericentromeric satellite regions in this cohort (17). In the eight
normal liver tissues, satellites 2 and 3 were heavily methylated (17).
We then categorized the degree of DNA hypomethylation in
Southern blotting as ,,or2:indicates that the HpaII digest
showed the same hybridization pattern as normal liver tissues,
indicates that smaller fragments were detected in the HpaII digest
Fig. 2. Examples of PCRsingle-strand conformation polymorphism analysis
of the DNMT3b gene in HCC cases with the intronic primer set encompassing
exon 13. N, noncancerous liver tissue; T, HCC. The shifted bands of H15N and
H15T (arrows) were sequenced and revealed to be a polymorphism in intron
13 (Table 2).
Table 2. Polymorphisms of the DNMT3b gene
Position Genotype
HCC patients
(n49)
Patients with liver
metastasis of primary
colon cancer (n8)
Intron 2
(22 bp upstream of the 3splice site)
38 5
c9 3
cc2 0
Intron 13
(1618 bp downstream of the 5splice site)
actact 48 8
act10
Fig. 3. Examples of RT-PCR with a primer set that will amplify all four splice
variants in the C-terminal catalytic domain of DNMT3b in HCCs (H) and Alexander
cells. (A) Alexander cells were derived from human HCC (29) and showed DNA
hypomethylation on pericentromeric satellite regions (data not shown). PCR
products of about 420 bp (a), 350 bp (b), 300 bp (c), and 230 bp (d) will correspond
to DNMT3b1, DNMT3b4, DNMT3b5, and DNMT3b3, respectively. (B) RT-PCR
products of H35T1 were cloned into the pCRII vector, and the inserts were
examined by colony PCR. Almost all clones examined possessed inserts of about
350 bp (b) and 230 bp (d). (C) Sequencing conrmed that the insert of about 350
bp (b) corresponded to DNMT3b4, which lacks the conserved methyltransferase
motifs IX and X (Fig. 1) because of the premature stop codon (underlined) arising
from skipping of exon 21.
10062
www.pnas.orgcgidoi10.1073pnas.152121799 Saito et al.
compared with normal liver tissues, and 2indicates that the HpaII
digest showed the same hybridization pattern as the MspIdigestof
the same sample and normal liver tissues (17). Nine (18%) of 49
noncancerous liver tissues from HCC patients and 39 (66%) of 59
HCCs showed or 2DNA hypomethylation (17).
Fig. 4Ashows mRNA expression levels for DNMT3b3 normal-
ized to glyceraldehyde-3-phosphate dehydrogenase (GAPDH)
mRNA. The average levels of mRNA for DNMT3b3, normalized
to GAPDH mRNA, in tissue samples from HCC cases with ,,
and 2DNA hypomethylation were 0.72 1.20, 0.87 0.85, and
0.81 1.01 (mean SD), respectively. There was no significant
correlation between mRNA levels for DNMT3b3 and DNA meth-
ylation status on pericentromeric satellite regions.
Fig. 4Bshows mRNA expression levels for DNMT3b4 normal-
ized to GAPDH mRNA. The average levels of mRNA for
DNMT3b4, normalized to GAPDH mRNA, in tissue samples from
HCC cases with ,, and 2DNA hypomethylation were 0.31
0.23, 0.78 0.79, and 0.83 0.94, respectively. There was a
Fig. 4. Splice variant-specic quantitative RTPCR for
DNMT3b in HCC cases (H) and cases with liver metastasis of
primary colon cancer (C).N, noncancerous liver tissue; T,
HCC. (A) mRNA levels for DNMT3b3 normalized to GAPDH
mRNA. (B) mRNA levels for DNMT3b4 normalized to
GAPDH mRNA. (C) The ratio of DNMT3b4 mRNA to
DNMT3b3 mRNA. We have previously reported DNA meth-
ylation status on pericentromeric satellite regions for these
cases (17). In normal liver tissues (C1 to C8), satellites 2 and
3 were heavily methylated (17). The samples indicated by
white, hatched, and black bars show no (), moderate (),
and strong (2) DNA hypomethylation, respectively, as
dened in ref. 17. mRNA levels for DNMT3b4 normalized to
GAPDH mRNA (B,P0.0001, KruskalWallis test) and the
ratios of DNMT3b4 mRNA to DNMT3b3 mRNA (C,P
0.009, KruskalWallis test) each were signicantly corre-
lated with DNA methylation status on pericentromeric sat-
ellite regions.
Saito et al. PNAS
July 23, 2002
vol. 99
no. 15
10063
MEDICAL SCIENCES
significant correlation between mRNA levels for DNMT3b4 and
DNA methylation status on pericentromeric satellite regions (P
0.0001, KruskalWallis test). Fig. 4Cshows the ratio of DNMT3b4
mRNA to DNMT3b3 mRNA. The averages of the ratios of
DNMT3b4 mRNA to DNMT3b3 mRNA in tissue samples from
HCC cases with ,, and 2DNA hypomethylation were 0.71
0.52, 1.40 1.40, and 1.65 1.70, respectively. There was a
significant correlation between the ratio of DNMT3b4 mRNA to
DNMT3b3 mRNA and DNA methylation status on pericentro-
meric satellite regions (P0.009, KruskalWallis test).
Expression levels of mRNA for DNMT3b1 and DNMT3b5 in all
tissue samples were too low to be detected by this quantitative
RTPCR system.
Transfection of Human Epithelial 293 Cells with DNMT3b4 cDNA
Induces DNA Demethylation on Satellite 2. Five clones expressing
myc-tagged DNMT3b4, 3b41to3b45, were obtained. The ex-
pression of myc-tagged DNMT3b4 was weak in 3b41, whereas it
was particularly strong in 3b45 (Fig. 5A). About 50 days after
transfection, DNA methylation status on satellite 2 was analyzed by
Southern blotting. As shown in Fig. 5B, in 293 parent cells (indi-
vidually cultured parent-1 and parent-2), mock transfectants
(mock-1 to mock-3), and the DNMT3b4-transfected 3b41 clone,
which only weakly expresses myc-tagged DNMT3b4, larger DNA
fragments were detected in the HpaII digest compared with the
MspI digest. On the other hand, in clones 3b42to3b45, which
strongly express myc-tagged DNMT3b4, smaller fragments were
detected in the HpaII digest compared with parent cells, mock
transfectants, and clone 3b41. The signal intensity of the 6.6-kbp
HpaII digest was strong in all parent cells, mock transfectants, and
DNMT3b4 transfectants. The ratios of the signal intensity of the
2.0-kbp HpaII digest to that of the 6.6-kbp HpaII digest in clones
3b42to3b45 were significantly higher than those in parent cells,
mock transfectants, and clone 3b41 (Fig. 5B,P0.0001, Mann
Whitney Utest). Moreover, as shown in Fig. 5, there was a
correlation between expression level of myc-tagged DNMT3b4 and
the ratio of signal intensity of the 2.0-kbp HpaII digest to that of the
6.6-kbp HpaII digest among all DNMT3b4 transfectants. These
results suggest that overexpression of DNMT3b4 induces DNA
demethylation on satellite 2 in human epithelial cells.
Discussion
We have reported DNA hypomethylation on pericentromeric
satellite regions, DNA hypermethylation on CpG islands of
genes such as p16,E-cadherin, and HIC-1 (hypermethylated-in-
cancer), overexpression of DNA methyltransferases, and re-
duced expression of methyl-CpG-binding proteins during hepa-
tocarcinogenesis (1217). Among these findings, DNA
hypomethylation on pericentromeric satellite regions was ob-
served even in precancerous conditions and appears to be one of
the earliest events during hepatocarcinogenesis. To prevent the
development of HCCs in hepatitis virus carriers suffering from
chronic hepatitis or cirrhosis, clarification of the molecular
mechanisms underlying early events during hepatocarcinogen-
esis is needed. We have focused on the mechanism of DNA
hypomethylation on pericentromeric satellite regions.
DNMT3b is required for de novo DNA methylation on pericen-
tromeric satellite regions in embryonic stem cells and during mouse
development (22). In adult cells, the role of DNMT3b remains
unclear and it is unknown how DNA methylation on pericentro-
meric satellite regions is maintained. DNMT1 has so far been
recognized as the ‘‘maintenance’’ DNA methyltransferase. Once
established by DNMT3b during development, DNA methylation on
pericentromeric satellite regions is likely to be maintained by
DNMT1 in adult cells. In cancer cells, however, targeting of
DNMT1 to substrate DNA may be disrupted by mechanisms such
as dysfunction of p21
WAF1
(32), which competes with DNMT1 for
binding to proliferating cell nuclear antigen (33). In fact, in human
cancers, overexpression of DNMT1 is significantly correlated with
CpG island methylation phenotype, which is defined by frequent
DNA hypermethylation on CpG islands that are not methylated in
normal cells (27). If DNMT1 does not maintain DNA methylation
on pericentromeric satellite regions because of disturbance of its
targeting in cancer cells, DNMT3b must rescue DNA methylation
on these regions. Moreover, some recent studies have proposed that
all active DNA methyltransferases, DNMT1 and members of the
DNMT3 family, probably possess both de novo and maintenance
DNA methyltransferase activity in vivo, regardless of their prefer-
ence for hemimethylated or unmethylated substrates in vitro (4,
3436). Therefore, it is likely that DNMT3b is required to maintain
DNA methylation on pericentromeric satellite regions even in
somatic cells, including cancer cells. Therefore, we focused on
alterations of DNMT3b during hepatocarcinogenesis.
When analyzed previously by quantitative RT-PCR w ith a primer
set that cannot discriminate between splice variants of DNMT3b,
expression of total mRNA for DNMT3b did not correlate with
DNA hypomethylation on pericentromeric satellite regions in HCC
cases (17). Thus, it is unlikely that reduced expression of DNMT3b
simply causes DNA hypomethylation on these regions during
hepatocarcinogenesis. Although germ-line mutations of the
DNMT3b gene have been reported in patients with immunodefi-
ciencycentromeric instabilityfacial anomalies syndrome (23, 24),
in HCCs we detected no somatic mutations of the DNMT3b gene
in any coding exons.
There have been a small number of studies on expression of the
four splice variants in the C-terminal catalytic domain of DNMT3b
in vivo. DNMT3b3 possesses the N-terminal region and conserved
methyltransferase motifs I, IV, VI, IX, and X, but lacks 63 aa
residues between motifs VI and IX. Because mouse DNMT3b3 did
not methylate an applied unmethylated substrate in an in vitro study
(37), DNA methyltransferase activity of human DNMT3b3 should
also be carefully discussed. However, it has been reported that
Fig. 5. Transfection of human epithelial 293 cells with DNMT3b4 cDNA. (A)
Western blotting with monoclonal anti-myc antibody. Band a, myc-tagged
DNMT3b4; band b, endogenous c-myc. Among DNMT3b4 transfectants, expres-
sion of myc-tagged DNMT3b4 was weak in 3b41, whereas it was particularly
strong in 3b45. About 50 days after transfection, DNA methylation status on
satellite 2 was analyzed by Southern blotting (B). In parent cells (individually
cultured parent-1 and parent-2), mock transfectants, and 3b41, larger DNA
fragments were detected in the HpaII (H) digest compared with the MspI (M)
digest. In clones 3b4 2to3b45, smaller fragments were detected in the H digest
compared with parent-1 to parent-2, mock-1 to mock-3, and 3b4 1.The ratios of
signal intensity of the 2.0-kbp H digest (arrow) to that of the 6.6-kbp H digest
(arrowhead), whose intensity was strong in all parent cells, mock transfectants,
and DNMT3b4 transfectants, are shown at the bottom (P0.0001, Mann
Whitney Utest).
10064
www.pnas.orgcgidoi10.1073pnas.152121799 Saito et al.
DNMT3b3 is ubiquitously expressed in normal human tissues (26).
Our data also indicate that the major variant in normal liver tissues
is DNMT3b3. Therefore, we cannot rule out the possibility that
DNMT3b3 carries the DNMT3b activity at least in vivo, for example
in human liver tissue. DNMT3b4 probably does not show methyl-
transferase activity because it lacks the conserved methyltrans-
ferase motifs IX and X, although it retains the N-terminal domain
required for targeting to heterochromatin sites through binding to
RP58 (38, 39). In vivo, it has been reported that normal human
tissues, with the exception of testis, do not express significant levels
of DNMT3b4 (26). Here, we confirmed the trace level of
DNMT3b4 expression in normal liver tissues.
Overexpression of DNMT3b4 and elevation of the ratio of
DNMT3b4 mRNA to DNMT3b3 mRNA were both significantly
correlated with the degree of DNA hypomethylation on pericen-
tromeric satellite regions in precancerous conditions and HCCs.
DNMT3b4 may compete with the major variant, DNMT3b3, for
targeting to pericentromeric satellite regions. This may be the
reason overexpression of DNMT3b4, especially in relation to the
expression of DNMT3b3, results in DNA hypomethylation on
pericentromeric satellite regions in precancerous conditions and
HCCs. To confirm this possibility, we introduced DNMT3b4 into
human epithelial 293 cells, which express a significant level of
DNMT3b3 mRNA and a trace of endogenous DNMT3b4 mRNA
(data not shown). DNA demethylation on satellite 2 was observed
in DNMT3b4 transfectants, depending on the expression level of
myc-tagged DNMT3b4. Although the overexpression levels in
DNMT3b4 transfectants may exceed the levels of DNMT3b4 in
tissue samples from HCC patients, the results of the transfection
studies support the possibility that overexpression of DNMT3b4
induces DNA demethylation on pericentromeric satellite regions in
human epithelial cells. On the other hand, a few exceptional tissue
samples from HCC patients without overexpression of DNMT3b4
or an elevated ratio of DNMT3b4 mRNA to DNMT3b3 mRNA
also showed DNA hypomethylation on pericentromeric satellite
regions. Therefore, overexpression of DNMT3b4 should not be
considered as the only mechanism for DNA hypomethylation on
pericentromeric satellite regions.
We and other groups have shown that chromosomal instability,
e.g., allelic imbalance on chromosomes 1 and 16, accumulates even
in noncancerous liver tissues showing chronic hepatitis and cirrho-
sis, and is more frequent in HCCs (15, 16, 40). Satellite 2 is the major
sequence of the heterochromatin region adjacent to the centro-
meres of chromosomes 1 and 16. In fact, frequent 1q copy gain with
a breakpoint in heterochromatin DNA was reported in HCCs with
DNA hypomethylation on satellite 2 (20). Overexpression of
DNMT3b4 may lead to chromosomal instability through induction
of DNA hypomethylation on pericentromeric satellite regions, even
in the precancerous stages of HCCs. On the other hand, it was
recently shown that CpG islands and genes are relatively commonly
located in heterochromatin regions (41). The growth rate of
DNMT3b4 transfectants became about double that of mock trans-
fectants soon after introduction of DNMT3b4, when allelic imbal-
ance may not yet have accumulated (unpublished data). This
growth change may be caused by alterations in the expression of
genes in regions in which DNA methylation status is affected by
DNMT3b activity. In fact, the expression levels of several genes that
potentially participate in signal transduction pathways andor cell
growth were altered in the DNMT3b4 transfectants (unpublished
data).
We conclude, therefore, that overexpression of DNMT3b4 in-
duces DNA demethylation on pericentromeric satellite regions
even in precancerous stages and may play critical roles in the
development of HCC through chromosomal instability and aber-
rant expression of cancer-related genes. It would be of further
interest to investigate whether aberrant expression of DNMT3b
splice variants is of general significance during carcinogenesis, even
in organs other than the liver.
This study was supported by a Grant-in-Aid for the Second Term
Comprehensive 10-Year Strategy for Cancer Control and a Grant-in-Aid
for Cancer Research from the Ministry of Health, Labor, and Welfare
of Japan. Y.S. is a recipient of a Research Resident Fellowship from the
Foundation for Promotion of Cancer Research in Japan.
1. Jones, P. A. & Laird, P. W. (1999) Nat. Genet. 21, 163167.
2. Jones, P. A. & Takai, D. (2001) Science 293, 10681070.
3. Baylin, S. B., Esteller, M., Rountree, M. R., Bachman, K. E., Schuebel, K. & Herman, J. G.
(2001) Hum. Mol. Genet. 10, 687692.
4. Robertson, K. D. (2001) Oncogene 20, 31393155.
5. Gama-Sosa, M. A., Slagel, V. A., Trewyn, R. W., Oxenhandler, R., Kuo, K. C., Gehrke, C. W.
& Ehrlich, M. (1983) Nucleic Acids Res. 11, 68836894.
6. Vachtenheim, J., Horakova, I. & Novotna, H. (1994) Cancer Res. 54, 11451148.
7. Yoshiura, K., Kanai, Y., Ochiai, A., Shimoyama, Y., Sugimura, T. & Hirohashi, S. (1995) Proc.
Natl. Acad. Sci. USA 92, 7416 7419.
8. Eguchi, K., Kanai, Y., Kobayashi, K. & Hirohashi, S. (1997) Cancer Res. 57, 49134915.
9. Tsuda, H., Hirohashi, S., Shimosato, Y., Terada, M. & Hasegawa, H. (1988) Gastroenterology 95,
16641666.
10. Yasui, H., Hino, O., Ohtake, K., Machinami, R. & Kitagawa, T. (1992) Cancer Res. 52,
68106814.
11. Kanai, Y., Ushijima, S., Tsuda, H., Sakamoto, M., Sugimura, T. & Hirohashi, S. (1996) Jpn. J.
Cancer Res. 87, 12101217.
12. Kanai, Y., Ushijima, S., Hui, A. M., Ochiai, A., Tsuda, H., Sakamoto, M. & Hirohashi, S. (1997)
Int. J. Cancer 71, 355359.
13. Sun, L., Hui, A. M., Kanai, Y., Sakamoto, M. & Hirohashi, S. (1997) Jpn. J. Cancer Res. 88,
11651170.
14. Kanai, Y., Hui, A. M., Sun, L., Ushijima, S., Sakamoto, M., Tsuda, H. & Hirohashi, S. (1999)
Hepatology 29, 703709.
15. Kondo,Y., Kanai, Y., Sak amoto, M., Mizokami, M., Ueda, R. & Hirohashi, S. (2000) Hepatology
32, 970979.
16. Kanai,Y., Ushijima, S., Tsuda, H., Sak amoto, M. & Hirohashi, S. (2000) Cancer Lett. 148, 7380.
17. Saito, Y., Kanai, Y., Sakamoto, M., Saito, H., Ishii, H. & Hirohashi, S. (2001) Hepatology 33,
561568.
18. Kokalj-Vokac, N., Almeida, A., Viegas-Pequignot, E., Jeanpierre, M., Malfoy, B. & Dutrillaux,
B. (1993) Cytogenet. Cell Genet. 63, 1115.
19. Jeanpierre, M., Turleau, C., Aurias, A., Prieur, M., Ledeist, F., Fischer, A. & Viegas-Pequignot,
E. (1993) Hum. Mol. Genet. 2, 731735.
20. Wong, N., Lam, W. C., Lai, P. B., Pang, E., Lau, W. Y. & Johnson, P. J. (2001) Am. J. Pathol.
159, 465471.
21. Okano, M., Xie, S. & Li, E. (1998) Nat. Genet. 19, 219220.
22. Okano, M., Bell, D. W., Haber, D. A. & Li, E. (1999) Cell 99, 247257.
23. Hansen, R. S., Wijmenga, C., Luo, P., Stanek, A. M., Canfield, T. K., Weemaes, C. M. & Gartler,
S. M. (1999) Proc. Natl. Acad. Sci. USA 96, 1441214417.
24. Xu, G. L., Bestor, T. H., Bourchis,D., Hsieh, C. L., Tommer up, N., Bugge, M., Hulten, M., Qu,
X., Russo, J. J. & Viegas-Pequignot, E. (1999) Nature (London) 402, 187191.
25. Maraschio, P., Zuffardi, O., Dalla Fior, T. & Tiepolo, L. (1988) J. Med. Genet. 25, 173180.
26. Robertson, K. D., Uzvolgyi, E., Liang, G., Talmadge, C., Sumegi, J., Gonzales, F. A. & Jones,
P. A. (1999) Nucleic Acids Res. 27, 22912298.
27. Kanai, Y., Ushijima, S., Kondo, Y., Nak anishi, Y. & Hirohashi, S. (2001) Int. J. Cancer 91,
205212.
28. Xie, S., Wang, Z., Okano, M., Nogami, M., Li, Y., He, W. W., Okumura, K. & Li, E. (1999) Gene
236, 8795.
29. Alexander, J. J., Bey, E. M., Geddes, E. W. & Lecatsas, G. (1976) S. Afr. Med. J. 50, 21242128.
30. Graham, F. L., Smiley, J., Russell, W. C. & Nairn, R. (1977) J. Gen. Virol. 36, 5974.
31. Tagarro, I., Fernandez-Peralta, A. M. & Gonzalez-Aguilera, J. J. (1994) Hum. Genet. 93,
383388.
32. Baylin, S. B. (1997) Science 277, 19481949.
33. Chuang,L. S., Ian, H. I., Koh, T. W., Ng, H. H., Xu, G. & Li, B. F. (1997) Science 277, 1996 2000.
34. Vertino, P. M., Yen, R. W., Gao, J. & Baylin, S. B. (1996) Mol. Cell. Biol. 16, 45554565.
35. Rhee, I., Jair, K. W., Yen, R. W., Lengauer, C., Herman, J. G., Kinzler, K. W., Vogelstein, B.,
Baylin, S. B. & Schuebel, K. E. (2000) Nature (London) 404, 10031007.
36. Liang, G., Chan, M. F., Tomigahara, Y., Tsai, Y. C., Gonzales,F. A., Li, E., L aird, P. W. & Jones,
P. A. (2002) Mol. Cell. Biol. 22, 480 491.
37. Aoki, A., Suetake, I., Miyagawa, J., Fujio, T., Chijiwa, T., Sasaki, H. & Tajima, S. (2001) Nucleic
Acids Res. 29, 35063512.
38. Bachman, K. E., Rountree, M. R. & Baylin, S. B. (2001) J. Biol. Chem. 276, 3228232287.
39. Fuks, F., Burgers, W. A., Godin, N., Kasai, M. & Kouzarides, T. (2001) EMBO J. 20, 25362544.
40. Kishimoto, Y., Shiota, G., Wada, K., Kitano, M., Nakamoto, K., Kamisaki, Y., Suou, T., Itoh,
T. & Kawasaki, H. (1996) J. Cancer Res. Clin. Oncol. 122, 585589.
41. Barry, A. E., Howman, E. V., Cancilla, M. R., Saffery, R. & Choo, K. H. (1999) Hum. Mol. Genet .
8, 217227.
Saito et al. PNAS
July 23, 2002
vol. 99
no. 15
10065
MEDICAL SCIENCES
... The overexpression of DNMT3B isoforms has been reported in several types of cancer cells; for example, DNMT3B3 and DNMT3B4 in gastric cancer [33], DNMT3B4 in the liver and clear cells renal cancers [33][34][35], DNMT3B7 in breast cancer [36,37], and ∆DNMT3B4 in lung cancer [20]; however, the expression patterns and consequences of DNMT3B isoforms' overexpression in EOC are still unknown. Therefore, this study aimed to analyze the expression patterns of DNMT3B isoforms and their association with clinical variables such as tumor grade, disease stage, and prognostic value in EOC. ...
... In the present study, we found that the overexpression of DNMT3B was associated with differentiation grades and stages of disease; specifically, DNMT3B3 was associated with disease progression, low DFS, and notably with OS in EOC. Furthermore, in agreement with our results, other studies have also reported the overexpression of DNMT3B3 in gastric cancer patients [33] and a correlation between DNMT3B3 expression and DNA hypomethylation in liver cancer cells [35]. ...
... Interestingly, the structure of DNMT3B3 possesses methyltransferase motifs I, IV, VI, IX, and X, suggesting that it has a potential methyltransferase activity, in contrast to DNMT3B4, where most of the methyltransferase motifs are missing [27]. In this sense, we found that overexpression of DNMT3B3 in the OVCAR3 cell line induces hypomethylation of DNA in the locus of satellite 2. Similar to our results, it has been reported that overexpression of DNMT3B4 causes hypomethylation on pericentromeric satellite regions in the epithelial 293 cell line [35]. Besides, it has been observed that DNMT3B isoforms compete with each other; for example, DNMT3B3 competes with DNMT3B4 to target DNA regions in human epithelial cancer cells [26]. ...
Article
Full-text available
Alterations in DNA methylation are critical for the carcinogenesis of ovarian tumors, especially ovarian carcinoma (OC). DNMT3B, a de novo DNA methyltransferase (DNMT), encodes for fifteen spliced protein products or isoforms. DNMT3B isoforms lack exons for the catalytic domain, with functional consequences on catalytic activity. Abnormal expression of DNMT3B isoforms is frequently observed in several types of cancer, such as breast, lung, kidney, gastric, liver, skin, leukemia, and sarcoma. However, the expression patterns and consequences of DNMT3B isoforms in OC are unknown. In this study, we analyzed each DNMT and DNMT3B isoforms expression by qPCR in 63 OC samples and their association with disease-free survival (DFS), overall survival (OS), and tumor progression. We included OC patients with the main histological subtypes of EOC and patients in all the disease stages and found that DNMTs were overexpressed in advanced stages (p-value < 0.05) and high-grade OC (p-value < 0.05). Remarkably, we found DNMT3B1 overexpression in advanced stages (p-value = 0.0251) and high-grade serous ovarian carcinoma (HGSOC) (p-value = 0.0313), and DNMT3B3 was overexpressed in advanced stages (p-value = 0.0098) and high-grade (p-value = 0.0004) serous ovarian carcinoma (SOC). Finally, we observed that overexpression of DNMT3B isoforms was associated with poor prognosis in OC and SOC. DNMT3B3 was also associated with FDS (p-value = 0.017) and OS (p-value = 0.038) in SOC patients. In addition, the ovarian carcinoma cell lines OVCAR3 and SKOV3 also overexpress DNMT3B3. Interestingly, exogenous overexpression of DNMT3B3 in OVCAR3 causes demethylation of satellite 2 sequences in the pericentromeric region. In summary, our results suggest that DNMT3B3 expression is altered in OC.
... However, DNMT3B6 has been shown to interact and allosterically activate the full-length catalytically active enzyme DNMT3B1. [12][13][14][15][16][17] Aberrant expression and splicing of DNMT3B are linked to the loss of methylation at oncogenes and repetitive elements in diverse cancers, including colorectal, lung, and breast cancers. [18][19][20][21] Several observations indicate that the highly regulated spatiotemporal expression and alternative splicing of DNMT3B are critical for cell differentiation, homeostasis, and survival. ...
... Dnmt3b transcripts comprise two major alternatively spliced isoforms: Dnmt3b1, a fulllength transcript, and a catalytically inactive Dnmt3b6, in which exons 22 and 23 are excluded. [12][13][14][15][16][17] RT-PCR was used to amplify both isoforms pre-and post-differentiation of ESCs ( Figure 1B). As shown previously, 54 both isoforms are expressed equally in s-ESCs. ...
Article
Full-text available
Embryonic expression of DNMT3B is critical for establishing de novo DNA methylation. This study uncovers the mechanism through which the promoter-associated long non-coding RNA (lncRNA) Dnmt3bas controls the induction and alternative splicing of Dnmt3b during embryonic stem cell (ESC) differentiation. Dnmt3bas recruits the PRC2 (polycomb repressive complex 2) at cis-regulatory elements of the Dnmt3b gene expressed at a basal level. Correspondingly, Dnmt3bas knockdown enhances Dnmt3b transcriptional induction, whereas overexpression of Dnmt3bas dampens it. Dnmt3b induction coincides with exon inclusion, switching the predominant isoform from the inactive Dnmt3b6 to the active Dnmt3b1. Intriguingly, overexpressing Dnmt3bas further enhances the Dnmt3b1:Dnmt3b6 ratio, attributed to its interaction with hnRNPL (heterogeneous nuclear ribonucleoprotein L), a splicing factor that promotes exon inclusion. Our data suggest that Dnmt3bas coordinates alternative splicing and transcriptional induction of Dnmt3b by facilitating the hnRNPL and RNA polymerase II (RNA Pol II) interaction at the Dnmt3b promoter. This dual mechanism precisely regulates the expression of catalytically active DNMT3B, ensuring fidelity and specificity of de novo DNA methylation.
... Aberrant alternative splicing in the DNMT3B gene, which gives rise to DNMT3B4, is mainly detectable in hepatocellular carcinoma. The non-functional DNMT3B4 __ which lacks DNMT motifs __ induces hypomethylation in pericentromeric satellite regions, increasing the sensitivity of the genome to additional mutations [162]. Mutations within MBD domains of DNA methylation readers are other abnormalities of the DNA methylation machinery in different solid tumors. ...
... DNMT3B is the main de novo methyltransferase overexpressed in different types of cancer and has been described as necessary for tumor development 13,14,[72][73][74] . With a general approach, we explore a little about the frequency of DNMT3B dysregulation in tumors of epithelial origin. ...
Article
Full-text available
DNA methylation is a key epigenetic modification to regulate gene expression in mammalian cells. Abnormal DNA methylation in gene promoters is common across human cancer types. DNMT3B is the main de novo methyltransferase enhanced in several primary tumors. How de novo methylation is established in genes related to cancer is poorly understood. CpG islands (CGIs), common sequences, and transcription factors (TFs) that interact with DNMT3B have been associated with abnormal de novo methylation. We initially identified cis elements associated with DNA methylation to investigate the contribution of DNMT3B overexpression to the deregulation of its possible target genes in an epithelial cell model. In a set of downregulated genes (n = 146) from HaCaT cells with DNMT3B overexpression, we found CGI, common sequences, and TFs Binding Sites that interact with DNMT3B (we called them P-down-3B). PPL1, VAV3, IRF1, and BRAF are P-down-3B genes that are downregulated and increased their methylation in DNMT3B presence. Together these findings suggest that methylated promoters aberrantly have some cis elements that could conduce de novo methylation by DNMT3B.
... Although the individual differences between the obese and the control groups were not significant, both HDACs, DNMT1and DNMT3b underwent obvious downward trends in the overweight/obese group. Generally, high expression levels of HDACs, DNMT1 and DNMT3b can inhibit gene transcription [25][26][27][28][29]. Moreover, these results were consistent with our finding that methylation levels of cytosine in the promoter and the CGI were significantly lower in overweight/obese group compared to those in their normal weight counterpart. ...
Article
Full-text available
Background Over the past few decades, global maternal obesity prevalence has rapidly increased. This condition may induce long-lasting pathophysiological effects on either fetal or infant health that could be attributable to unknown unique changes in the umbilical blood composition. Methods A total of 34 overweight/obese and 32 normal-weight pregnant women were recruited. Fifteen umbilical blood samples including 8 overweight/obese subjects and 7 normal weight women were sequenced using Targeted Bisulfite Sequencing technology to detect the average methylation level of cytosine and identify the differentially methylated region (DMR). GO and KEGG analyses were then employed to perform pathway enrichment analysis of DMR-related genes and promoters. Moreover, the mRNA levels of methylation-related genes histone deacetylases (HDACs) and DNA methyltransferases (DNMTs) were characterized in the samples obtained from these two groups. Results Average methylated cytosine levels in both the CpG islands (CGI) and promoter significantly decreased in overweight/obese groups. A total of 1669 DMRs exhibited differences in their DNA methylation status between the overweight/obese and control groups. GO and KEGG analyses revealed that DMR-related genes and promoters were enriched in the metabolism, cancer and cardiomyopathy signaling pathways. Furthermore, the HDACs and DNMTs mRNA levels trended to decline in overweight/obese groups. Conclusions Decreased methylated cytosine levels in overweight/obese women induce the gene expression activity at a higher level than in the control group. DMRs between these two groups in the fetal blood may contribute to the changes in gene transcription that underlie the increased risk of metabolic disorders, cancers and cardiomyopathy in their offspring.
... DNMT3L is a stimulator of de novo methylation by DNMT3A and DNMT3B [15,16], particularly at retrotransposons [14], and we observed a high number of gains with DNMT3L overexpression, particularly within open sea regions. Other catalytically inactive isoforms have been demonstrated to both stimulate [46][47][48] and inhibit [31,49] de novo methylation, underlining the complexity in understanding such processes. DNMT3B3 and DNMT3B4 are interact directly with DNMT3A2, DNMT3B1 and DNMT3B2 to inhibit de novo methylation by weakening their binding to DNA [31]. ...
Article
Full-text available
Background DNA methylation in the human genome is established and maintained by DNA methyltransferases (DNMTs). DNMT isoforms show differential expression by cell lineage and during development, but much remains to be elucidated about their shared and unique genomic targets. Results We examined changes in the epigenome following overexpression of 13 DNMT isoforms in HEK293T cells. We observed increased methylation (Δ β > 0.2) at 43,405 CpG sites, with expression of DNMT3A2, DNMTΔ3B4 and DNMTΔ3B2 associated with the greatest impact. De novo methylation occurred primarily within open sea regions and at loci with intermediate methylation levels ( β : 0.2–0.6). 53% of differentially methylated loci showed specificity towards a single DNMT subfamily, primarily DNMTΔ3B and DNMT3A and 39% towards a single isoform. These loci were significantly enriched for pathways related to neuronal development (DNMTΔ3B4), calcium homeostasis (DNMTΔ3B3) and ion transport (DNMT3L). Repetitive elements did not display differential sensitivity to overexpressed DNMTs, but hypermethylation of Alu elements was associated with their evolutionary age following overexpression of DNMT3A2, DNMT3B1, DNMT3B2 and DNMT3L. Differential methylation (Δ β > 0.1) was observed at 121 of the 353 loci associated with the Horvath ‘epigenetic clock’ model of ageing, with 51 showing isoform specificity, and was associated with reduction of epigenetic age by 5–15 years following overexpression of seven isoforms. Finally, we demonstrate the potential for dietary constituents to modify epigenetic marks through isoform-specific inhibition of methylation activity. Conclusions Our results provide insight into regions of the genome methylated uniquely by specific DNMT isoforms and demonstrate the potential for dietary intervention to modify the epigenome.
Article
Background & Aims Small nuclear ribonucleoprotein U1 subunit 70 (SNRNP70) as one of the components of the U1 small nuclear ribonucleoprotein (snRNP) is rarely reported in cancers. This study aims to estimate the application potential of SNRNP70 in hepatocellular carcinoma (HCC) clinical practice. Methods Based on the TCGA database and cohort of HCC patients, we investigated the expression patterns and prognostic value of SNRNP70 in HCC. Then, the combination of SNRNP70 and alpha-fetoprotein (AFP) in 278 HCC cases was analyzed. Next, western blotting and immunohistochemistry were used to detect the expression of SNRNP70 in nucleus and cytoplasm. Finally, Cell Counting Kit-8 (CCK-8) and scratch wound healing assays were used to detect the effect of SNRNP70 on the proliferation and migration of HCC cells. Results SNRNP70 was highly expressed in HCC. Its expression was increasingly high during the progression of HCC and was positively related to immune infiltration cells. Higher SNRNP70 expression indicated a poor outcome of HCC patients. In addition, nuclear SNRNP70/AFP combination could be a prognostic biomarker for overall survival and recurrence. Cell experiments confirmed that knockdown of SNRNP70 inhibited the proliferation and migration of HCC cells. Conclusion SNRNP70 may be a new biomarker for HCC progression and HCC diagnosis as well as prognosis. SNRNP70 combined with serum AFP may indicate the prognosis and recurrence status of HCC patients after operation.
Article
Careful microscopic observation of histopathological specimens, accumulation of large numbers of high‐quality tissue specimens, and analysis of molecular pathology in relation to morphological features are considered to yield realistic data on the nature of multistage carcinogenesis. Since the morphological hallmark of cancer is disruption of the normal histological structure maintained through cell−cell adhesiveness and cellular polarity, attempts have been made to investigate abnormalities of the cadherin‐catenin cell adhesion system in human cancer cells. It has been shown that the CDH1 tumor suppressor gene encoding E‐cadherin is silenced by DNA methylation, suggesting that a “double hit” involving DNA methylation and loss of heterozygosity leads to carcinogenesis. Therefore, in the 1990s, we focused on epigenomic mechanisms, which until then had not received much attention. In chronic hepatitis and liver cirrhosis associated with hepatitis virus infection, DNA methylation abnormalities were found to occur frequently, being one of the earliest indications that such abnormalities are present even in precancerous tissue. Aberrant expression and splicing of DNA methyltransferases, such as DNMT1 and DNMT3B, was found to underlie the mechanism of DNA methylation alterations in various organs. The CpG island methylator phenotype in renal cell carcinoma was identified for the first time, and its therapeutic targets were identified by multilayer omics analysis. Furthermore, the DNA methylation profile of nonalcoholic steatohepatitis (NASH)‐related hepatocellular carcinoma was clarified in groundbreaking studies. Since then, we have developed diagnostic markers for carcinogenesis risk in NASH patients and noninvasive diagnostic markers for upper urinary tract cancer, as well as developing a new high‐performance liquid chromatography‐based diagnostic system for DNA methylation diagnosis. Research on the cancer epigenome has revealed that DNA methylation alterations occur from the precancerous stage as a result of exposure to carcinogenic factors such as inflammation, smoking, and viral infections, and continuously contribute to multistage carcinogenesis through aberrant expression of cancer‐related genes and genomic instability. DNA methylation alterations at the precancerous stages are inherited by or strengthened in cancers themselves and determine the clinicopathological aggressiveness of cancers as well as patient outcome. DNA methylation alterations have applications as biomarkers, and are expected to contribute to diagnosis, as well as preventive and preemptive medicine.
Article
Aim The aim of this study was to clarify the significance of DNA methylation alterations of cryptogenic hepatocellular carcinomas (HCCs). Methods Using the Infinium assay, we performed genome‐wide DNA methylation analysis of 250 liver tissue samples, including non‐cancerous liver tissue (U‐N) and corresponding cancerous tissue (U‐T) from patients with cryptogenic HCC without a history of excessive alcohol use and hepatitis virus infection, and whose U‐N samples showed no remarkable histological features (no microscopic evidence of simple steatosis, any form of hepatitis including non‐alcoholic steatohepatitis [NASH], or liver cirrhosis). Results We identified 3,272 probes that showed significant differences of DNA methylation levels between U‐N and normal liver tissue samples from patients without HCC, indicating that a distinct DNA methylation profile had already been established at the precancerous U‐N stage. U‐Ns have a DNA methylation profile differing from that of non‐cancerous liver tissue of patients with NASH‐related, viral hepatitis‐related, and alcoholic liver disease‐related HCCs. Such DNA methylation alterations in U‐Ns were inherited by U‐Ts. The U‐Ns showed DNA methylation alteration of ADCY5 resulting in alteration of its mRNA expression, whereas non‐cancerous liver tissue of patients with NASH‐, viral hepatitis‐, or alcoholic liver disease‐related HCCs did not. DNA methylation levels of MICAL2 and PLEKHG2 in U‐Ts were correlated with larger tumor diameter and portal vein involvement, respectively. Conclusions U‐N‐specific DNA hypermethylation of ADCY5 may have significance even from the precancerous stage in liver showing no remarkable histological features. DNA hypomethylation of MICAL2 and PLEKHG2 may determine the clinicopathological features of cryptogenic HCC. This article is protected by copyright. All rights reserved.
Article
Full-text available
The Dnmt3a DNA methyltransferase is essential for mammalian development and is responsible for the generation of genomic methylation patterns, which lead to transcriptional silencing. Here, we show that Dnmt3a associates with RP58, a DNA-binding transcriptional repressor protein found at transcriptionally silent heterochromatin. Dnmt3a acts as a co-repressor for RP58 in a manner that does not require its de novo methyltransferase activity. Like other characterized co-repressors, Dnmt3a associates with the histone deacetylase HDAC1 using its ATRX-homology domain. This domain of Dnmt3a represents an independent transcriptional repressor domain whose silencing functions require HDAC activity. These results identify Dnmt3a as a co-repressor protein carrying deacetylase activity and show that Dnmt3a can be targeted to specific regulatory foci via its association with DNA-binding transcription factors.
Article
Full-text available
Gene function in cancer can be disrupted either through genetic alterations, which directly mutate or delete genes, or epigenetic alterations, which alter the heritable state of gene expression. The latter events are mediated by formation of transcriptionally repressive chromatin states around gene transcription start sites and an associated gain of methylation in normally unmethylated CpG islands in these regions. The genes affected include over half of the tumor suppressor genes that cause familial cancers when mutated in the germline; the selective advantage for genetic and epigenetic dysfunction in these genes is very similar. The aberrant methyla-tion can begin very early in tumor progression and mediate most of the important pathway abnormalities in cancer including loss of cell cycle control, altered function of transcription factors, altered receptor function, disruption of normal cell–cell and cell–substratum interaction, inactivation of signal transduction pathways, loss of apoptotic signals and genetic instability. The active role of the aberrant methylation in transcriptional silencing of genes is becoming increasingly understood and involves a synergy between the methylation and histone deacetylase (HDAC) activity. This synergy can be mediated directly by HDAC interaction with DNA methylating enzymes and by recruitment through complexes involving methyl-cytosine binding proteins. In the translational arena, the promoter hypermethylation changes hold great promise as DNA tumor markers and their potentially reversible state creates a target for cancer therapeutic strategies involving gene reactivation.
Article
Gene function in cancer can be disrupted either through genetic alterations, which directly mutate or delete genes, or epigenetic alterations, which alter the heritable state of gene expression. The latter events are mediated by formation of transcriptionally repressive chromatin states around gene transcription start sites and an associated gain of methylation in normally unmethylated CpG islands in these regions. The genes affected include over half of the tumor suppressor genes that cause familial cancers when mutated in the germline; the selective advantage for genetic and epigenetic dysfunction in these genes is very similar. The aberrant methylation can begin very early in tumor progression and mediate most of the important pathway abnormalities in cancer including loss of cell cycle control, altered function of transcription factors, altered receptor function, disruption of normal cell-cell and cell-substratum interaction, inactivation of signal transduction pathways, loss of apoptotic signals and genetic instability. The active role of the aberrant methylation in transcriptional silencing of genes is becoming increasingly understood and involves a synergy between the methylation and histone deacetylase (HDAC) activity. This synergy can be mediated directly by HDAC interaction with DNA methylating enzymes and by recruitment through complexes involving methyl-cytosine binding proteins. In the translational arena, the promoter hypermethylation changes hold great promise as DNA tumor markers and their potentially reversible state creates a target for cancer therapeutic strategies involving gene reactivation.
Article
DNA–(cytosine-5) methyltransferase (MCMT) methylates newly replicated mammalian DNA, but the factors regulating this activity are unknown. Here, MCMT is shown to bind proliferating cell nuclear antigen (PCNA), an auxiliary factor for DNA replication and repair. Binding of PCNA requires amino acids 163 to 174 of MCMT, occurs in intact cells at foci of newly replicated DNA, and does not alter MCMT activity. A peptide derived from the cell cycle regulator p21WAF1 can disrupt the MCMT-PCNA interaction, which suggests that p21WAF1 may regulate methylation by blocking access of MCMT to PCNA. MCMT and p21WAF1 may be linked in a regulatory pathway, because the extents of their expression are inversely related in both SV40-transformed and nontransformed cells.
Article
We have purified GST-fused recombinant mouse Dnmt3a and three isoforms of mouse Dnmt3b to near homogeneity. Dnmt3b3, an isoform of Dnmt3b, did not have DNA methylation activity. Dnmt3a, Dnmt3b1 or Dnmt3b2 showed similar activity toward poly(dG-dC)-poly(dG-dC) for measuring de novo methylation activity, and toward poly(dI-dC)-poly(dI-dC) for measuring total activity. This indicates that the enzymes are de novo-type DNA methyltransferases. The enzyme activity was inhibited by NaCl or KCl at concentrations >100 mM. The kinetic parameter, KmAdoMet, for Dnmt3a, Dnmt3b1 and Dnmt3b2 was 0.4, 1.2 and 0.9 µM when poly(dI-dC)-poly(dI-dC) was used, and 0.3, 1.2 and 0.8 µM when poly(dG‐dC)‐poly(dG-dC) was used, respectively. The KmDNA values for Dnmt3a, Dnmt3b1 and Dnmt3b2 were 2.7, 1.3 and 1.5 µM when poly(dI-dC)-poly(dI-dC) was used, and 3.5, 1.0 and 0.9 µM when poly(dG-dC)-poly(dG-dC) was used, respectively. For the methylation specificity, Dnmt3a significantly methylated CpG >> CpA. On the other hand, Dnmt3b1 methylated CpG > CpT ≥ CpA. Immuno-purified Dnmt3a, Myc‐tagged and overexpressed in HEK 293T cells, methylated CpG >> CpA > CpT. Neither Dnmt3a nor Dnmt3b1 methylated the first cytosine of CpC.
Article
The recessive autosomal disorder known as ICF syndrome (for immunodeficiency, centromere instability and facial anomalies; Mendelian Inheritance in Man number 242860) is characterized by variable reductions in serum immunoglobulin levels which cause most ICF patients to succumb to infectious diseases before adulthood. Mild facial anomalies include hypertelorism, low-set ears, epicanthal folds and macroglossia. The cytogenetic abnormalities in lymphocytes are exuberant: juxtacentromeric heterochromatin is greatly elongated and thread-like in metaphase chromosomes, which is associated with the formation of complex multiradiate chromosomes. The same juxtacentromeric regions are subject to persistent interphase self-associations and are extruded into nuclear blebs or micronuclei. Abnormalities are largely confined to tracts of classical satellites 2 and 3 at juxtacentromeric regions of chromosomes 1, 9 and 16. Classical satellite DNA is normally heavily methylated at cytosine residues, but in ICF syndrome it is almost completely unmethylated in all tissues. ICF syndrome is the only genetic disorder known to constitutive abnormalities of genomic methylation Here we show that five unrelated ICF patients have mutations in both alleles of the gene that encodes DNA methyltransferase 3B (refs 5, 6). Cytosine methylation is essential for the organization and stabilization of a specific type of heterochromatin, and this methylation appears to be carried out by an enzyme specialized for the purpose.
Article
In order to clarify the significance of DNA methylation in both earlier and later stages of hepatocarcinogenesis, the DNA methylation state on chromosome 16, on which loss of heterozygosity (LOH) has frequently been detected in human hepatocellular carcinomas (HCCs), was examined. DNA from primary HCCs and tissues showing chronic hepatitis and liver cirrhosis, which are considered to be precancerous conditions, was analyzed by digestion with methylation-sensitive and non-sensitive restriction enzymes. DNA hypermethylation at the D16S32, tyrosine aminotransferase (TAT) and D16S7 loci and hypomethylation at the D16S4 locus were detected in 18%, 58%, 20% and 48% of examined HCCs, respectively. Aberrant DNA methylation occurred more frequently in advanced HCCs than in early HCCs. Moreover, DNA hypermethylation at the D16S32, TAT and D16S7 loci was frequently observed in chronic hepatitis and liver cirrhosis. The incidence of DNA hypermethylation was higher than that of LOH (42% at the TAT locus). These data suggest that DNA hypermethylation might predispose the locus to allelic loss. Aberrant DNA methylation is a significant change which may participate in the early developmental stages of HCCs.
Article
To evaluate the significance of alterations in DNA methylation during human hepatocarcinogenesis, we examined levels of mRNA for DNA methyltransferases and methyl-CpG-binding proteins and the DNA methylation status in 67 hepatocellular carcinomas (HCCs). The average level of mRNA for DNMT1 and DNMT3a was significantly higher in noncancerous liver tissues showing chronic hepatitis or cirrhosis than in histologically normal liver tissues, and was even higher in HCCs. Significant overexpression of DNMT3b and reduced expression of DNMT2 were observed in HCCs compared with the corresponding noncancerous liver tissues. DNA hypermethylation on CpG islands of the p16 (8% and 66%) and hMLH1 (0% and 0%) genes and methylated in tumor (MINT) 1 (6% and 34%), 2 (24% and 58%), 12 (21% and 33%), 25 (0% and 5%), and 31 (0% and 23%) clones, and DNA hypomethylation on satellites 2 and 3 (18% and 67%), were detected in noncancerous liver tissues and HCCs, respectively. There was no significant correlation between the expression level of any DNA methyltransferase and DNA methylation status. Reduced expression of DNA repair protein, MBD4, was significantly correlated with poorer tumor differentiation and involvement of portal vein. Slightly reduced expression of MBD2 was detected in HCCs, and the expression of MeCP2 was particularly reduced in HCCs with portal vein involvement. These data suggest that overexpression of DNMT1 and DNMT3a, DNA hypermethylation on CpG islands, and DNA hypomethylation on pericentromeric satellite regions are early events during hepatocarcinogenesis, and that reduced expression of MBD4 may play a role in malignant progression of HCC.