ArticlePDF Available

Silencing of Glutathione Peroxidase 3 through DNA Hypermethylation Is Associated with Lymph Node Metastasis in Gastric Carcinomas

Authors:

Abstract and Figures

Gastric cancer remains the second leading cause of cancer-related death in the world. H. pylori infection, a major risk factor for gastric cancer, generates high levels of reactive oxygen species (ROS). Glutathione peroxidase 3 (GPX3), a plasma GPX member and a major scavenger of ROS, catalyzes the reduction of hydrogen peroxide and lipid peroxides by reduced glutathione. To study the expression and gene regulation of GPX3, we examined GPX3 gene expression in 9 gastric cancer cell lines, 108 primary gastric cancer samples and 45 normal gastric mucosa adjacent to cancers using quantitative real-time RT-PCR. Downregulation or silencing of GPX3 was detected in 8 of 9 cancer cell lines, 83% (90/108) gastric cancers samples, as compared to non-tumor adjacent normal gastric samples (P
Content may be subject to copyright.
Silencing of Glutathione Peroxidase 3 through DNA
Hypermethylation Is Associated with Lymph Node
Metastasis in Gastric Carcinomas
Dun-Fa Peng
1
, Tian-Ling Hu
1
, Barbara G. Schneider
2
, Zheng Chen
1,5
, Ze-Kuan Xu
5,6
, Wael El-Rifai
1,3,4
*
1Department of Surgery, Vanderbilt University Medical Center, Nashville, Tennessee, United States of America, 2Department of Medicine, Vanderbilt University Medical
Center, Nashville, Tennessee, United States of America, 3Department of Cancer Biology, Vanderbilt University Medical Center, Nashville, Tennessee, United States of
America, 4Department of Veterans Affairs Tennessee Valley Healthcare System, Nashville, Tennessee, United States of America, 5Department of General Surgery, the First
Affiliated Hospital of Nanjing Medical University, Nanjing, China, 6Institute of Tumor Biology, Jiangsu Province Academy of Clinical Medicine, Nanjing, China
Abstract
Gastric cancer remains the second leading cause of cancer-related death in the world. H. pylori infection, a major risk factor
for gastric cancer, generates high levels of reactive oxygen species (ROS). Glutathione peroxidase 3 (GPX3), a plasma GPX
member and a major scavenger of ROS, catalyzes the reduction of hydrogen peroxide and lipid peroxides by reduced
glutathione. To study the expression and gene regulation of GPX3, we examined GPX3 gene expression in 9 gastric cancer
cell lines, 108 primary gastric cancer samples and 45 normal gastric mucosa adjacent to cancers using quantitative real-time
RT-PCR. Downregulation or silencing of GPX3 was detected in 8 of 9 cancer cell lines, 83% (90/108) gastric cancers samples,
as compared to non-tumor adjacent normal gastric samples (P,0.0001). Examination of GPX3 promoter demonstrated DNA
hypermethylation ($10% methylation level determined by Bisulfite Pyrosequencing) in 6 of 9 cancer cell lines and 60% of
gastric cancer samples (P= 0.007). We also detected a significant loss of DNA copy number of GPX3 in gastric cancers
(P,0.001). Treatment of SNU1 and MKN28 cells with 5-Aza-29Deoxycytidine restored the GPX3 gene expression with a
significant demethylation of GPX3 promoter. The downregulation of GPX3 expression and GPX3 promoter hypermethyla-
tion were significantly associated with gastric cancer lymph node metastasis (P= 0.018 and P= 0.029, respectively). We also
observed downregulation, DNA copy number losses, and promoter hypermethylation of GPX3 in approximately one-third of
tumor-adjacent normal gastric tissue samples, suggesting the presence of a field defect in areas near tumor samples.
Reconstitution of GPX3 in AGS cells reduced the capacity of cell migration, as measured by scratch wound healing assay.
Taken together, the dysfunction of GPX3 in gastric cancer is mediated by genetic and epigenetic alterations, suggesting
impairment of mechanisms that regulate ROS and its possible involvement in gastric tumorigenesis and metastasis.
Citation: Peng D-F, Hu T-L, Schneider BG, Chen Z, Xu Z-K, et al. (2012) Silencing of Glutathione Peroxidase 3 through DNA Hypermethylation Is Associated with
Lymph Node Metastasis in Gastric Carcinomas. PLoS ONE 7(10): e46214. doi:10.1371/journal.pone.0046214
Editor: Regine Schneider-Stock, Institute of Pathology, Germany
Received April 16, 2012; Accepted August 29, 2012; Published October 10, 2012
This is an open-access article, free of all copyright, and may be freely reproduced, distributed, transmitted, modified, built upon, or otherwise used by anyone for
any lawful purpose. The work is made available under the Creative Commons CC0 public domain dedication.
Funding: This work was funded by National Institutes of Health R01CA106176 and R01CA93999. The funders had no role in study design, data collection and
analysis, decision to publish, or preparation of the manuscript.
Competing Interests: Co-author Wael El-Rifai is a PLOS ONE Editorial Board member. This does not alter the authors’ adherence to all the PLOS ONE policies on
sharing data and materials.
* E-mail: wael.el-rifai@Vanderbilt.edu
Introduction
Gastric cancer (GC) is the fourth most common cancer in the
world [1,2] with about 900,000 new cases diagnosed in the world
each year [2]. Gastric cancer remains the second leading cause of
cancer-related deaths worldwide [2]. Although there has been a
decline in the overall incidence of distal gastric cancer during the
past decades, the incidence is rising for adenocarcinomas of the
proximal part of the stomach in the Western world [2]. It is
generally agreed that gastric cancer is a multifactor disease in
which Helicobacter pylori (H. pylori) infection plays a crucial role,
especially for distal gastric cancer [3–5]. Accumulating data
indicate that H. pylori infection generates high levels of reactive
oxygen species (ROS) from multiple sources [6]. Activated
neutrophils, for example, are the main source of ROS production
in the H. pylori-infected stomach; however, H. pylori itself can also
produce ROS [7]. In addition, extensive recent studies have
revealed that H. pylori-induced ROS production in gastric
epithelial cells might affect gastric epithelial cell signal transduc-
tion, resulting in gastric carcinogenesis [8–10]. Excessive ROS
production in the stomach promotes DNA damage in gastric
epithelial cells, suggesting its involvement in gastric carcinogenesis
[7,11,12].
Normal cells have intact antioxidative properties that protect
cells from ROS-induced DNA damage and cell injury [13–15].
Among these systems, the glutathione peroxidase family (GPXs) is
a major antioxidative enzyme family that catalyzes the reduction
of hydrogen peroxide, organic hydroperoxide, and lipid peroxides
by reduced glutathione [15–19]. Glutathione peroxidase 3
(GPX3), also named plasma glutathione peroxidase, is the only
known selenocysteine containing an extracellular antioxidant
isoform [20]. The human GPX3 gene is approximately 10 kb in
length, spanning 5 exons on chromosome 5q32. [20,21]. It has
been reported that GPX3 catalyzes the reduction of hydrogen
peroxide and lipid peroxides and is a major scavenger of ROS
produced during normal metabolism or after oxidative insult
PLOS ONE | www.plosone.org 1 October 2012 | Volume 7 | Issue 10 | e46214
[22,23]. GPX3 is selectively expressed in normal human tissues,
including the gastrointestinal tract. However, downregulation of
GPX3 has been recently reported in multiple human cancers such
as prostate, esophageal, and bladder cancer [24–26], suggesting its
importance in human tumorigenesis. In the present study, we
examined GPX3 gene expression, gene promoter methylation
status, and copy number in a panel of primary gastric cancers and
correlated it with clinicopathological parameters.
Materials and Methods
Ethics Statement
De-identified human tissue samples were obtained from the
archives of pathology at Vanderbilt University (Nashville, TN,
USA) and from the National Cancer Institute Cooperative Human
Tissue Network (CHTN). The use of specimens was approved by
the Institutional Review Board at Vanderbilt University Medical
Center. All patients provided written consent, and samples were
collected after surgical resection. All tissue samples that were
included in this study were collected from tissues that remained
after the completion of diagnosis and are otherwise discarded.
Tissue samples
All tissue samples were obtained from the pathology archives at
Vanderbilt University (Nashville, TN, USA) and from the
National Cancer Institute Cooperative Human Tissue Network
(CHTN). The use of specimens from the archival tissue repository
was approved by the Institutional Review Board. All tissue samples
included in this study were coded and collected from tissues that
remained after the completion of diagnosis and that were
otherwise discarded. Among 108 patients, 70 male and 38 female,
ages ranging from 38–87 years of age with a median age of 64
years. All tumors were histologically verified. The gastric
adenocarcinomas ranged from well-differentiated (WD) to poorly
differentiated (PD), stages I to IV, with a mix of intestinal- and
diffuse-type tumors. The ‘‘normal’’ samples in this study were the
gastric mucosal epithelial tissues adjacent to the cancers without
neoplastic changes.
Cell lines
Nine gastric cancer cell lines (AGS, MKN28, MKN45,
MKN75, KATO3, SNU1, SNU5, SNU16, and RF1) were
purchased from American Type Culture Collection (Manassas,
VA, http://www.atcc.org) and Riken (Ibaraki, Japan; http://
www.brc.riken.go.jp/lab/cell/english). Cells were maintained in
either DMEM medium or RPMI 1640 medium with a supplement
of 10% fetal bovine serum and antibiotics. All cell lines were
cultured at 37uC with 5% CO
2
.
Quantitative Real-Time Reverse Transcription PCR (qRT-
PCR) Analysis of GPX3
Total RNA was isolated using the RNeasy Mini-kit (Qiagen,
Valencia, CA, USA). Single-stranded cDNA was subsequently
synthesized using the iScript cDNA Synthesis Kit (Bio-Rad,
Hercules, CA, USA). Expression of GPX3 was evaluated for a set
of 153 frozen primary human samples including 108 samples of
gastric carcinoma and 45 samples of normal gastric mucosa
adjacent to cancers. For 24 tumors, matching normal mucosa was
available from the same patients. The GPX3 primers (forward 59-
GCCGGGGACAAGAGAAGT-39and reverse 59-GAGGACG-
TATTTGCCAGCAT-39) were designed using the online soft-
ware, Primer 3 (http://frodo.wi.mit.edu/). The qRT-PCR was
performed using an iCycler (Bio-Rad) with the threshold cycle
number determined by use of iCycler software, version 3.0.
Reactions were performed in triplicate and the threshold numbers
were averaged. Results for the GPX3 gene were normalized to
HPRT1 gene (forward 59-TTGGAAAGGGTGTT
TATTCCTCA-39and reverse 59-TCCAGCAGGTCAGC AAA-
GAA-39), which had minimal variation in all normal and tumor
samples tested, and is therefore considered to be a reliable and
stable reference gene for RT-PCR. Expression was calculated by
use of the formula 2
(RtEt)
/2
(RnEn)
as previously described [27,28].
For all of the primary gastric carcinoma samples, the gene was
considered downregulated, if the relative mRNA expression was
#0.5 [29].
DNA Bisulfite Treatment and Pyrosequencing Analysis
DNA was purified using a DNeasy Tissue Kit (Qiagen). The
bisulfite modification of the DNA from cell lines and tissues was
performed using an EZ DNA Methylation-Gold Kit (Zymo
Research, Orange, CA), according to the manufacturer’s protocol.
The GPX3 promoter CpG island was identified by using a CpG
island online search tool (http://www.uscnorris.com), as previous-
ly described [28]. The Pyrosequencing primers were designed
using PSQ Assay Design Software (Biotage, Uppsala, Sweden).
The forward primer sequence was AGGTGGGGAGTT-
GAGGGTAA, the reverse biotin-labeled primer sequence was
Biotin-TCCCAACCACCTTTCAAAC, and the sequencing
primer was GGGAGTTGAGGGTAAGT. A 40 ng aliquot of
modified DNA was subjected to polymerase chain reaction (PCR)
amplification of the specific promoter region using the above
primers and the Platinum PCR SuperMix High Fidelity Enzyme
Mix (Invitrogen, Carlsbad, CA). The PCR products were checked
by gel electrophoresis to confirm the size of the product and rule
out the formation of primer dimers. The specific PCR products
were then subjected to quantitative Pyrosequencing analysis using
a Biotage PyroMark MD System (Biotage), following the protocol
provided by the manufacturer. The results were analyzed by Pyro
Q-CpG 1.0.9 software (Biotage). Based on the methylation levels
in the normal samples, we used 10% methylation as a cutoff for
the identification of DNA hypermethylation of the GPX3
promoter. Statistical analysis was performed to detect significant
changes in the frequencies of DNA methylation of the CpG sites
between tumor and normal samples.
5-Aza-29Deoxycytidine and Trichostatin-A Treatment
For validation of the role of promoter DNA hypermethylation
in transcriptional regulation of GPX3 in vitro, gastric cancer cell
lines SNU1 and MKN28 were used. SNU1 and MKN28 cells
were maintained in Dulbecco’s modified Eagle’s medium
(DMEM) supplemented with 10% fetal bovine serum (FBS) and
antibiotics (Invitrogen). Cells were seeded at low density for
24 hours and then treated with 5 mM 5-Aza-29deoxycytidine (5-
Aza, Sigma-Aldrich, St. Louis, MO) for 72 hours or 300 nM
Trichostatin-A (TSA, Wako, Osaka, Japan) for 24 hours. Total
RNA and DNA were isolated and purified by RNeasy and
DNeasy Tissue kits (Qiagen), as described above. DNA methyl-
ation levels of the CpG nucleotides of the GPX3 promoter were
determined by Pyrosequencing. The GPX3 mRNA expression
levels were determined by qRT-PCR, as described above.
Immunofluorescence staining of GPX3 protein in SNU1
cells
To check GPX3 protein expression after 5-Aza treatment, we
performed immunofluorescence staining against GPX3 in SNU1.
SNU1 cells treated with 5 mM 5-Aza (Sigma-Aldrich) for 72 hours
and/or 300 nM Trichostatin-A (TSA, Wako) for 24 hours, fixed
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 2 October 2012 | Volume 7 | Issue 10 | e46214
with fresh 4% paraformaldehyde for 45 min at room temperature,
followed by permeabilization with 0.1% Triton X-100 in 0.1%
sodium citrate for 2 min on ice. Cells were then incubated with
10% normal goat serum (Invitrogen) for 20 min at room
temperature. Cells were incubated with primary antibody against
GPX3 (Rabbit, 1:500; Abnova, Taiwan) overnight at 4uC followed
by secondary goat anti-rabbit antibody conjugated with Alexa
Fluor 488 (1:1000, Invitrogen) at room temperature for 45 min.
The slides were mounted using Vectashield with DAPI (Vector
Laboratories, Burlingame, California, USA) and viewed under a
fluorescence microscope.
Measurement of DNA Copy Number by Quantitative PCR
(qPCR)
For evaluation of relative DNA copy numbers, we performed
the qPCR amplifications using an iCycler (Bio-Rad). PCR
reactions were prepared in a total volume of 20 ml containing
template DNA (40 ng), with the threshold cycle number deter-
mined by use of iCycler software version 3.0. Primers were
designed using the online software Primer 3 (http://frodo.wi.mit.
edu/). The forward and reverse primers for GPX3 genomic DNA
were 59-CCCCTTCAGTAGGGCCTAAG-39and 59-
TTCTTCAGGACCAGGACCAC-39, respectively. The primers
were obtained from Integrated DNA Technologies (Coralville,
Iowa). Reactions were performed in triplicate, and the threshold
numbers (CT) were averaged. The results were normalized to the
average CT of both b-Actin and GAPDH, which generated
similar results and had minimal variation in all normal and tumor
samples tested. DNA copy number was calculated in the same way
as the quantification of qRT-PCR for mRNA expression [29] and
normalized to the average level of 10 blood DNA samples of
normal individuals in which the GPX3 copy number are not
affected ( 2 copies, equal to copy number ratio 1.0). Loss of DNA
copy number was considered at a relative cutoff ratio #0.5.
Construction of GPX3 expression adenoviral system
The full length of GPX3 coding sequence with Flag-tag was
amplified from normal cDNA by PCR using Platinum PCR
SuperMix High Fidelity (Invitrogen) and was cloned into the
pACCMV.pLpA plasmid. The pACCMV.pLpA-GPX3 plasmid
was then co-transfected with pJM17 plasmid into 293 AD cells to
generate and propagate the recombinant GPX3-expressing
adenoviral particles as previously described [30]. The viruses
were plaque purified, and the titer of the virus was determined
using the Adeno-X qPCR Titration Kit (Clontech, California,
USA) following the manufacturer’s instructions.
Colony formation assay
AGS cells were infected with 25 MOI control or GPX3-
expressing adenoviral particles. 48 hours after infection, cells were
split and seeded (500 cells/well) in 6-well plates. Cells were
cultured at 37uC for another 2 weeks. Cells then were stained with
0.05% crystal violet. The images of the plates were analyzed using
Image J software. Each experiment was set in triplicate and
statistical analysis was done using Prism software. The relative
number of colonies in GPX3-expressing cells was adjusted to
control cells.
Colony formation in soft agar
To check if GPX3 was also involved in anchorage-independent
growth, soft agar colony formation assay was performed. In brief,
AGS and MKN28 cells were infected with 25 MOI control or
GPX3-expressing adenoviral particles. 48 hours after infection,
cells were split and seeded (1.0610
5
cells/well) in a 0.35% noble
agar (Sigma) mixed with culture medium (on the top of 0.5%
noble agar with medium) in 6-well plates. Cells were cultured at
37uC for another 2 weeks. The images of the plates were captured
under a microscope and analyzed using Image J software. Each
experiment was set in triplicate and statistical analysis was done
using Prism software. The relative number of colonies in GPX3-
expressing cells was adjusted to control cells.
Scratch wound healing assay
To check if GPX3 has a role in cell migration, we performed a
scratch wound healing assay in AGS cells [31]. AGS cells were
infected with 25 MOI GPX3-expressing or control adenoviruses.
When cells were 100% confluent, a scratch was made across the
plates using a pipette’s tip. Cells were cultured in complete
medium and images were taken every 24 hours to monitor the
wound healing process.
Statistical Analysis
GraphPad Prism software version 4.0 (GraphPad Prism
Software, La Jolla, CA) was used for all of the statistical analyses.
The Student ttest was used to compare the DNA methylation level
and mRNA ratios between normal and gastric cancers in matched
samples (paired t-test) and unmatched samples (unpaired t-test). In
addition, we analyzed the association between DNA methylation
and clinicopathological factors. The correlation between the DNA
methylation level and mRNA expression was determined by
Spearman’s rank correlation. The comparison of GPX3 gene
expression or DNA methylation level with clinicopathological
factors was made by Chi-square/Fisher’s exact tests or unpaired t-
test. All Pvalues were based on two-tailed tests and differences
were considered statistically significant when the Pvalue was
#.05.
Results
GPX3 was frequently downregulated or silenced in
gastric carcinomas
The qRT-PCR analysis demonstrated that GPX3 mRNA
expression was frequently downregulated in gastric cancer cell
lines (8/9 cell lines examined, Table 1) and in primary gastric
cancer tissue samples (90/108, 83%) as compared to 45 normal
samples (P,.0001; Figure 1A). Further analysis of 24 paired tumor
and normal samples confirmed the significant downregulation of
GPX3 mRNA expression in tumors as compared with their
corresponding normal samples (Figure 1B). Moreover, about one-
third (36/108, 33%) of primary gastric cancers and 6/9 of the
cancer cell lines (Table 1) showed complete silencing of GPX3
mRNA expression, as indicated by the absence of a detectable
signal. Notably, about 33% (15/45) of the ‘‘normal’’ gastric tissues
adjacent to cancers without neoplastic changes also displayed
downregulation (relative expression, #0.5) of GPX3 mRNA, as
normalized to the value of the average of all normal samples.
Promoter DNA hypermethylation of the GPX3 gene
correlates with downregulation of mRNA expression
The DNA methylation changes, which were determined by
Pyrosequencing technology, were not associated with the patients’
ages. A representative DNA methylation scheme of each CpG sites
in the GPX3 promoter of 4 matched normal and tumors is shown
in Figure 2A. Quantitative analysis of GPX3 promoter DNA
methylation, indicated increased promoter DNA methylation
levels of all tested CpG nucleotides in tumor samples compared
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 3 October 2012 | Volume 7 | Issue 10 | e46214
with normal samples (Figure 2B). The average methylation level
for the GPX3 gene promoter in gastric cancers was significantly
higher than that in normal samples (P= 0.007, Figure 2C).
Hypermethylation of the GPX3 promoter ($10%) was detected in
60% (36/60) of the gastric cancers and 6/9 cancer cell lines
(Table 1), consistent with our results showing GPX3 downregu-
lation. We observed that 39% (9/23) of the normal gastric tissues
adjacent to cancers showed an increased DNA methylation level of
the GPX3 promoter from 10% to 15%. Analysis of DNA
methylation in 22 tumor samples and their matching tumor-
adjacent normal gastric mucosae demonstrated a similar signifi-
cant increase in the level of DNA methylation in tumors compared
with their controls (P= 0.0086, Figure 2D). We next analyzed the
promoter DNA methylation against mRNA expression levels in all
samples. As shown in Figure 3A, samples with hypermethylation
(.10%) had significantly lower levels of GPX3 expression
(P= 0.035) compared with samples with absent or low promoter
methylation levels (#10%). Using the Spearman rank correlation,
we found a significant inverse correlation between promoter
methylation and mRNA expression of GPX3 (coefficient
r=20.32, P= 0.024; Figure 3B). These results suggest that the
hypermethylation of the GPX3 promoter region is one of the
factors involved in the suppression of its mRNA expression in
gastric cancers.
5-Aza-29Deoxycytidine (5-Aza) and Trichostatin-A (TSA)
treatment restored GPX3 expression in silenced gastric
cancer cell lines
As shown in Figure 4, the 5-Aza treatment of the SNU1 (A) and
MKN28 (B) cell lines with its fully methylated GPX3 promoter
(Table 1) restored GPX3 mRNA expression, and this restoration
of gene expression was associated with promoter demethylation.
TSA treatment alone had no effect in restoring the GPX3
expression or in altering the methylation levels. However,
administration of TSA following 5-Aza had a significant additive
effect in restoring gene expression. Furthermore, TSA treatment
following 5-Aza led to additional gene expression with a further
decrease in the methylation level of GPX3. To check if 5-Aza
treatment also restored GPX3 protein level, we performed an
immunofluorescence assay using antibody against GPX3 in SNU1
cells. As shown in Figure 4C, there was a significant increase in the
GPX3 green immunofluorescence signal after 5-Aza and 5-Aza-
TSA treatments as compared to DMSO control, suggesting that 5-
Aza and 5-Aza-TSA can restore the GPX3 protein expression in
these cells.
Figure 1. GPX3 is downregulated in gastric cancers. 108 gastric cancer samples and 45 tumor-adjacent histologically normal gastric mucosa
samples were analyzed by real-time RT-PCR for GPX3 expression, which was significantly downregulated in gastric cancers as compared to adjacent
normal samples (A). A similar result was confirmed in 24 matched normal and tumor samples from the same patients (B).
doi:10.1371/journal.pone.0046214.g001
Table 1. DNA copy number, methylation level and gene
expression of GPX3 in gastric cancer cell lines.
Sample
DNA Copy
number ratio % Methylation
mRNA expression
ratio
Normal 1.0 5 1
AGS 0.5 3 0.1
MKN28 0.6 91 0
MKN45 0.3 92 0
MKN75 0.6 91 0
KATO3 0.2 88 0
SNU1 0.4 94 0
SNU5 0.8 4 0.9
SNU16 0.3 6 0.1
RF1 0.6 24 0
Copy number of the GPX3 gene was determined using qPCR and normalized to
10 normal blood DNA samples as described in the Methods section. The real
copy number of each sample equals to the above copy number ratio times 2.
DNA methylation levels were determined using Pyrosequencing and are shown
as an average level of the 8 CpG sites in the GPX3 promoter region. GPX3 mRNA
expression fold was determined using real-time RT-PCR and normalized to the
average value of normal gastric samples as described in the Methods section.
doi:10.1371/journal.pone.0046214.t001
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 4 October 2012 | Volume 7 | Issue 10 | e46214
Loss of DNA copy numbers cooperates with DNA
hypermethylation for silencing GPX3 mRNA expression in
gastric cancer
Although GPX3 promoter hypermethylation correlated statis-
tically with low gene expression levels, we detected silencing of
GPX3 mRNA expression in 83% (90/108) of gastric cancers,
whereas promoter hypermethylation was seen in only 60% (36/60)
of gastric cancers. These findings prompted us to find out whether
loss of copy numbers could be a contributing factor in silencing
GPX3 expression. Evaluation of relative DNA copy numbers of
the 9 gastric cancer cell lines clearly demonstrated that 5 of the 9
cell lines have lost GPX3 copy number (copy number ratio#0.5)
as compared to the normal cells, which have a copy number of 2
(Table 1). In particular, in AGS and SNU16, loss of DNA copy
number seems to account for GPX3 downregulation. While in the
other 6 cell lines (MKN28, MKN45, MKN75, KATO3, SNU1,
and RF1), both loss of copy number and DNA hypermethylation
are associated with the silencing. The only cell line (SNU5) lacking
both DNA hypermethylation and copy number loss also expresses
GPX3 similar to the level in normal samples. The DNA copy
number of twenty-two gastric cancer samples and 22 tumor-
adjacent ‘‘normal’’ stomach samples were compared with the
average copy number of 10 blood samples from normal
individuals. The GPX3 copy numbers in the 10 blood samples
were used as an accurate reference for calculating the ratio and
normalization in gastric tissues. As shown in Figure 5, there was a
significantly lower level of GPX3 DNA copy numbers in the
gastric cancers and the tumor-adjacent ‘‘normal’’ stomach tissue
samples, as compared to that in normal blood samples (P= 0.001).
The comparison of GPX3 copy number in 22 matched tumor and
Figure 2. GPX3 promoter was hypermethylated in gastric cancers. DNA methylation level of 8 CpG sites in the GPX3 promoter was
quantitated by Pyrosequencing. (A) A schematic profile of GPX3 methylation of the 8 CpG sites was examined in 4 matched normal (N) and tumor (T)
samples. (B) The DNA methylation level of each of the 8 CpG sites in 23 normal samples and 60 tumor samples were measured. * P,0.01. (C) The
average DNA methylation level of the 8 CpG sites in 23 normal samples and 60 tumor samples is shown. (D) The average DNA methylation level of the
8 CpG sites in 22 matched normal and tumor samples from the same patients is indicated.
doi:10.1371/journal.pone.0046214.g002
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 5 October 2012 | Volume 7 | Issue 10 | e46214
Figure 3. GPX3 promoter methylation level inversely correlates with GPX3 gene expression. (A) GPX3 expression level in gastric cancers
with GPX3 hypermethylation ($10%) was significantly lower than that in gastric cancers without hypermethylation (,10%). (B) Spearman analysis of
all normal and tumor samples with GPX3 gene expression and DNA methylation demonstrated a significant inverse correlation(r = 20.32) between
DNA hypermethylation and gene expression.
doi:10.1371/journal.pone.0046214.g003
Figure 4. 5-Aza treatment restored GPX3 gene expression in gastric cancer cell lines with silenced GPX3. SNU1 and MKN28 cancer cells
were treated with 5-Aza and TSA as described in the Methods section. In A (SNU1 cells) and B (MKN28 cells), relative GPX3 expression ratios
normalized to HPRT are shown in the left panel. DNA methylation levels of corresponding samples are shown on the right panel. C shows
immunofluorescence staining of GPX3 protein in SNU1 cells. 5-Aza, 5-Aza-29deoxycytidine. TSA, Trichostatin-A.
doi:10.1371/journal.pone.0046214.g004
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 6 October 2012 | Volume 7 | Issue 10 | e46214
normal samples demonstrated lower levels of DNA copy numbers
in tumors (P= 0.058).
Loss of GPX3 correlated with tumor lymph node
metastasis
Based on the available clinicopathological information, we
found both GPX3 hypermethylation and downregulation of
GPX3 expression significantly correlated with lymph node
metastasis (P= 0.029 and P= 0.018 for DNA methylation and
gene expression, respectively; Figure 6). No significant correlation
was found between either DNA methylation or gene expression,
and other parameters examined, such as patient age, sex, tumor
grade, and invasion depth. GPX3 DNA copy number did not
correlate with lymph node metastasis and other parameters,
possibly due to the small number of samples.
Reconstitution of GPX3 in gastric cancer cells did not
suppress tumor cell growth but inhibited tumor cell
migration
Because GPX3 has been suggested to be a potential tumor
suppressor in prostate cancer [26], we hypothesized that GPX3
may have the similar tumor suppressor function in gastric cancer.
We reconstituted GPX3 gene expression in AGS and MKN28 cell
lines and performed colony formation and soft agar colony
formation assays. Interestingly, we did not observe any significant
difference in the number of colonies between GPX3-expressing
AGS cells and the control cells (Figure 7A). Similar results were
obtained by soft agar assay in AGS cells and MKN28 cells
(Figures 7B&7C). In contrast, scratch wound healing assay data
indicated that GPX3-expressing AGS cells exhibited significantly
slower repair of the scratched wound as compared to control cells,
suggesting that GPX3 expression could impair cell migration
(Figure 8).
Discussion
The glutathione peroxidase family (GPXs) is a major anti-
oxidative enzyme family that catalyzes the reduction of hydrogen
peroxide, organic hydroperoxide, and lipid peroxides [15–19]. In
this process, GPXs convert oxygen superoxide and hydrogen
peroxide, a major ROS which have been reported to induce
oxidative DNA damage [30], to harmless intermediate products,
and therefore play critical roles in protecting cells from DNA
damage. GPX3 is a secreted form of the GPX family that is readily
detectable in plasma and mucosal surfaces, and detoxifies ROS
before it can enter into cells [17]. H. pylori infection is the main risk
factor for gastric tumorigenesis which leads to pro-inflammatory
response and generation of high levels of ROS in stomach with a
significant increase in oxidative DNA damage [7]. ROS-induced
DNA damage leading to disruption of genomic integrity has been
shown to be an important cause of human cancers [14]. The loss
of GPX-mediated activities may be associated with early stages of
inflammation-mediated carcinogenesis [16]. Therefore, our results
showing frequent loss of GPX3 expression in gastric cancer may
underscore the failure in the cellular antioxidant system which is
the first line of defense against detrimental ROS activity.
Figure 5. GPX3 copy number loss in gastric cancers. GPX3 copy
number was determined by qPCR and normalized to both beta-actin
and GAPDH of the same samples. The results were then compared to 10
blood DNA samples from normal individuals in which GPX3 copy
number was assumed to be 2.0 (copy number ratio equals to 1.0 in
Table 1). Loss of GPX3 copy number was detected in both gastric cancer
samples and adjacent normal samples.
doi:10.1371/journal.pone.0046214.g005
Figure 6. Dysfunction of GPX3 correlated with lymph node metastasis. (A) Loss of GPX3 gene expression correlated with lymph node
metastasis. (B) Increase in GPX3 methylation correlated with lymph node metastasis. (C) GPX3 copy number change did not correlate with lymph
node metastasis.
doi:10.1371/journal.pone.0046214.g006
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 7 October 2012 | Volume 7 | Issue 10 | e46214
Downregulation of GPX3 has been reported in human cancers
such as lung, ovarian, bladder, esophageal, and prostate cancer
[24–26,28,32], as well as in gastric cancers [33,34]. Mono-allelic
hypermethylation and inactivation of GPX3 in benign precursor
lesions, metaplasia, and dysplasia of the esophagus has been
reported; while inactivation of both alleles was detected in invasive
carcinoma [25]. Complete inactivation of the GPX3 gene by
genetic loss of one allele and methylation-mediated silencing of the
remaining allele was also reported in prostate cancer [26]. In the
current study, we have demonstrated promoter hypermethylation
and DNA copy number loss as possible mechanisms mediating
silencing of GPX3 in gastric cancers, although other mechanisms
such as gene mutation and miRNA regulation may also be
involved and need to be further studied. The occurrence of both
genetic and epigenetic alterations in GPX3 suggests that silencing
of this gene could be a critical event in the multi-step gastric
tumorigenesis cascade. Of note, we observed mRNA downregu-
lation, promoter hypermethylation and DNA copy number loss of
GPX3 in a number of tumor-adjacent ‘‘normal’’ gastric mucosae
samples. These tumor-adjacent ‘‘normal’’ tissues, although histo-
logically normal, they usually have some degree of inflammation
and could possibly have changes at the molecular level. Therefore,
our data suggest that silencing of GPX3 is possibly an early event
in gastric tumorigenesis. Given the known function of GPX3, its
Figure 7. Reconstitution of GPX3 did not suppress tumor cells growth. (A) Colony formation assay in AGS cells. (B) Soft agar colony
formation assay in AGS cells. (C) Soft agar colony formation assay in MKN28 cells. Right panels in A, B and C demonstrate the quantitative data using
Image J software.
doi:10.1371/journal.pone.0046214.g007
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 8 October 2012 | Volume 7 | Issue 10 | e46214
downregulation is expected to result in impairment in the
antioxidant capacity of cells with possible accumulation of
oxidative DNA damage at the early stage of gastric tumorigenesis.
Further studies on GPX3 methylation, inflammation and H. pylori
infection are needed to fully elucidate the relationship among
them.
A recent study has demonstrated that GPX3 can suppress
prostate cancer growth and metastasis [26]. Unfortunately, we did
not confirm the similar tumor suppression role of GPX3 in two
gastric cancer cell lines (AGS and MKN28), suggesting that GPX3
may have differential functional roles in different organs.
Nonetheless, our scratch wound healing assay data indicated that
GPX3-expressing cells have impairment in their migration
capacity. This is in accordance with our findings showing a
significant correlation of GPX3 downregulation and hypermethy-
lation with lymph node metastasis. Further studies are necessary to
confirm this finding using alternative assays and to elucidate the
underlying mechanisms of how GPX3 potentially regulates cancer
metastasis.
In conclusion, our results suggest that GPX3 gene inactivation
by promoter methylation and/or copy number loss is a frequent
finding in gastric cancer that correlates with increased incidence of
lymph node metastasis. This loss may account for a reduced
antioxidant capacity and accumulation of oxidative DNA damage
observed in gastric tumorigenesis. Further studies to explore the
functions of GPX3 in gastric tumorigenesis are required to address
its potential role in the development and progression of gastric
cancer, in particular, its potential roles in tumor cell migration and
metastasis.
Author Contributions
Conceived and designed the experiments: DFP WER ZKX BS. Performed
the experiments: TLH ZC DP. Analyzed the data: DFP WER.
Contributed reagents/materials/analysis tools: DFP WER. Wrote the
paper: DFP WER BS.
References
1. Parkin DM, Pisani P, Ferlay J (1999) Estimates of the worldwide incidence of 25
major cancers in 1990. International journal of cancer Journal international du
cancer 80: 827–841.
2. Crew KD, Neugut AI (2006) Epidemiology of gastric cancer. World journal of
gastroenterology : WJG 12: 354–362.
3. Pandey R, Misra V, Misra SP, Dwivedi M, Kumar A, et al. (2010) Helicobacter
pylori and gastric cancer. Asian Pacific journal of cancer prevention : APJCP 11:
583–588.
4. Piazuelo MB, Epplein M, Correa P (2010) Gastric cancer: an infectious disease.
Infectious disease clinics of North America 24: 853–869, vii.
5. Pritchard DM, Crabtree JE (2006) Helicobacter pylori and gastric cancer.
Current opinion in gastroenterology 22: 620–625.
6. Augusto AC, Miguel F, Mendonca S, Pedrazzoli J Jr, Gurgueira SA (2007)
Oxidative stress expression status associated to Helicobacter pylori virulence in
gastric diseases. Clinical biochemistry 40: 615–622.
7. Handa O, Naito Y, Yoshikawa T (2010) Helicobacter pylori: a ROS-inducing
bacterial species in the stomach. Inflammation research : official journal of the
European Histamine Research Society [et al] 59: 997–1003.
8. Handa O, Naito Y, Yoshikawa T (2007) CagA protein of Helicobacter pylori: a
hijacker of gastric epithelial cell signaling. Biochemical pharmacology 73: 1697–
1702.
9. Smoot DT, Elliott TB, Verspaget HW, Jones D, Allen CR, et al. (2000)
Influence of Helicobacter pylori on reactive oxygen-induced gastric epithelial
cell injury. Carcinogenesis 21: 2091–2095.
10. Lamb A, Yang XD, Tsang YH, Li JD, Higashi H, et al. (2009) Helicobacter
pylori CagA activates NF-kappaB by targeting TAK1 for TRAF6-mediated Lys
63 ubiquitination. EMBO reports 10: 1242–1249.
11. Bagchi D, McGinn TR, Ye X, Bagchi M, Krohn RL, et al. (2002) Helicobacter
pylori-induced oxidative stress and DNA damage in a primary culture of human
gastric mucosal cells. Digestive diseases and sciences 47: 1405–1412.
12. Papa A, Danese S, Sgambato A, Ardito R, Zannoni G, et al. (2002) Role of
Helicobacter pylori CagA+infection in determining oxidative DNA damage in
gastric mucosa. Scandinavian journal of gastroenterology 37: 409–413.
13. Halliwell B (1999) Antioxidant defence mechanisms: from the beginning to the
end (of the beginning). Free radical research 31: 261–272.
14. Strange RC, Spiteri MA, Ramachandran S, Fryer AA (2001) Glutathione-S-
transferase family of enzymes. Mutation research 482: 21–26.
15. Hayes JD, McLellan LI (1999) Glutathione and glutath ione-dependent enzymes
represent a co-ordinately regulated defence against oxidative stress. Free radical
research 31: 273–300.
16. Brigelius-Flohe R, Kipp A (2009) Glutathione peroxidases in different stages of
carcinogenesis. Biochimica et biophysica acta 1790: 1555–1568.
17. Brigelius-Flohe R (2006) Glutathione peroxidases and redox-regulated tran-
scription factors. Biological chemistry 387: 1329–1335.
18. Wingler K, Brigelius-Flohe R (1999) Gastrointestinal glutathione peroxidase.
BioFactors 10: 245–249.
19. Herbette S, Roeckel-Drevet P, Drevet JR (2007) Seleno-independent glutathione
peroxidases. More than simple antioxidant scavengers. The FEBS journal 274:
2163–2180.
20. Yoshimura S, Suemizu H, Taniguchi Y, Watanabe K, Nomoto Y, et al. (1991)
Molecular cloning of human plasma glutathione peroxidase gene and its
expression in the kidney. Nucleic acids symposium series: 163–164.
21. Yoshimura S, Suemizu H, Taniguchi Y, Arimori K, Kawabe N, et al. (1994)
The human plasma glutathione peroxidase-encoding gene: organization,
sequence and localization to chromosome 5q32. Gene 145: 293–297.
22. Takebe G, Yarimizu J, Saito Y, Hayashi T, Nakamura H, et al. (2002) A
comparative study on the hydroperoxide and thiol specificity of the glutathione
peroxidase family and selenoprotein P. The Journal of biological chemistry 277:
41254–41258.
23. Comhair SA, Erzurum SC (2005) The regulation and role of extracellular
glutathione peroxidase. Antioxidants & redox signaling 7: 72–79.
24. Chen B, Rao X, House MG, Nephew KP, Cullen KJ, et al. (2011) GPx3
promoter hypermethylation is a frequent event in human cancer and is
associated with tumorigenesis and chemotherapy response. Cancer letters 309:
37–45.
25. Lee OJ, Schneider-Stock R, McChesney PA, Kuester D, Roessner A, et al.
(2005) Hypermethylation and loss of expression of glutathione peroxidase-3 in
Barrett’s tumorigenesis. Neoplasia 7: 854–861.
26. Yu YP, Yu G, Tseng G, Cieply K, Nelson J, et al. (2007) Glutathi one peroxidase
3, deleted or methylated in prostate cancer, suppresses prostate cancer growth
and metastasis. Cancer research 67: 8043–8050.
27. El-Rifai W, Moskaluk CA, Abdrabbo MK, Harper J, Yoshida C, et al. (2002)
Gastric cancers overexpress S100A calcium-binding proteins. Cancer research
62: 6823–6826.
28. Peng DF, Razvi M, Chen H, Washington K, Roessner A, et al. (2009) DNA
hypermethylation regulates the expression of members of the Mu-class
Figure 8. Reconstitution of GPX3 expression suppressed tumor
cell migration. AGS cells were infected with Ad-GPX3 and Ad-CTRL.
When cells were confluent (48 hours after viral infection), a scratch
wound healing assay was performed. The scratch gap was periodically
monitored and recorded.
doi:10.1371/journal.pone.0046214.g008
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 9 October 2012 | Volume 7 | Issue 10 | e46214
glutathione S-transferases and glutathione peroxidases in Barrett’s adenocarci-
noma. Gut 58: 5–15.
29. Soutto M, Peng D, Razvi M, Ruemmele P, Hartmann A, et al. (2010) Epigenetic
and genetic silencing of CHFR in esophageal adenocarcinomas. Cancer 116:
4033–4042.
30. Peng D, Belkhiri A, Hu T, Chaturvedi R, Asim M, et al. (2011) Glutathione
peroxidase 7 protects against oxidative DNA damage in oesophageal cells. Gut.
31. Liang CC, Park AY, Guan JL (2007) In vitro scratch assay: a convenient and
inexpensive method for analysis of cell migration in vitro. Nature protocols 2:
329–333.
32. He Y, Wang Y, Li P, Zhu S, Wang J, et al. (2011) Identification of GPX3
epigenetically silenced by CpG methylation in human esophageal squamous cell
carcinoma. Digestive diseases and sciences 56: 681–688.
33. Jee CD, Kim MA, Jung EJ, Kim J, Kim WH (2009) Identification of genes
epigenetically silenced by CpG methylation in human gastric carcinoma.
European journal of cancer 45: 1282–1293.
34. Zhang X, Yang JJ, Kim YS, Kim KY, Ahn WS, et al. (2010) An 8-gene
signature, including methylated and down-regulated glutathione peroxidase 3, of
gastric cancer. International journal of oncology 36: 405–414.
GPX3 Silencing and Metastasis
PLOS ONE | www.plosone.org 10 October 2012 | Volume 7 | Issue 10 | e46214
... Furthermore, it should also be noted that clonogenicity is higher for lower sowing densities, probably due to competitive inhibition phenomena. The clonogenicity results are consistent with the results found in the literature [41][42][43][44]. This method shows that cancer stem cells can survive and form colonies in an anchorage-independent culture model. ...
Article
Full-text available
Gastric cancer (GC) ranked as the fifth most incident cancer in 2020 and the third leading cause of cancer mortality. Surgical prevention and radio/chemotherapy are the main approaches used in GC treatment, and there is an urgent need to explore and discover innovative and effective drugs to better treat this disease. A new strategy arises with the use of repurposed drugs. Drug repurposing coupled with drug combination schemes has been gaining interest in the scientific community. The main objective of this project was to evaluate the therapeutic effects of alternative drugs in GC. For that, three GC cell lines (AGS, MKN28, and MKN45) were used and characterized. Cell viability assays were performed with the reference drug 5-fluororacil (5-FU) and three repurposed drugs: natamycin, nitazoxanide, and benztropine. Nitazoxanide displayed the best results, being active in all GC cells. Further, 5-FU and nitazoxanide in combination were tested in MKN28 GC cells, and the results obtained showed that nitazoxanide alone was the most promising drug for GC therapy. This work demonstrated that the repurposing of drugs as single agents has the ability to decrease GC cell viability in a concentration-dependent manner.
... Despite the fact that these findings were made in tumor cells that spontaneously overexpressed GPX3 and that high GPX3 expression was associated with poorer overall survival in patients with ovarian cancer and with increased tumor stage [109], all direct manipulations with GPX3 overexpression conversely indicate the potential safety of such manipulation from the perspective of carcinogenesis. The suppression of proliferation, migration, and invasion in vitro [326,327,332,333] as well as of tumorigenicity in vivo [328][329][330][331] were shown in models overexpressing GPX3. Additionally, several other studies have provided evidence that in tumor tissues and cancer cell lines, GPX3 expression is suppressed or completely blocked due to promoter hypermethylation or gene deletion [329,331,[344][345][346]. ...
Article
Full-text available
Reactive oxygen species (ROS) are normal products of a number of biochemical reactions and are important signaling molecules. However, at the same time, they are toxic to cells and have to be strictly regulated by their antioxidant systems. The etiology and pathogenesis of many diseases are associated with increased ROS levels, and many external stress factors directly or indirectly cause oxidative stress in cells. Within this context, the overexpression of genes encoding the proteins in antioxidant systems seems to have become a viable approach to decrease the oxidative stress caused by pathological conditions and to increase cellular stress resistance. However, such manipulations unavoidably lead to side effects, the most dangerous of which is an increased probability of healthy tissue malignization or increased tumor aggression. The aims of the present review were to collect and systematize the results of studies devoted to the effects resulting from the overexpression of antioxidant system genes on stress resistance and carcinogenesis in vitro and in vivo. In most cases, the overexpression of these genes was shown to increase cell and organism resistances to factors that induce oxidative and genotoxic stress but to also have different effects on cancer initiation and promotion. The last fact greatly limits perspectives of such manipulations in practice. The overexpression of GPX3 and SOD3 encoding secreted proteins seems to be the “safest” among the genes that can increase cell resistance to oxidative stress. High efficiency and safety potential can also be found for SOD2 overexpression in combinations with GPX1 or CAT and for similar combinations that lead to no significant changes in H2O2 levels. Accumulation, systematization, and the integral analysis of data on antioxidant gene overexpression effects can help to develop approaches for practical uses in biomedical and agricultural areas. Additionally, a number of factors such as genetic and functional context, cell and tissue type, differences in the function of transcripts of one and the same gene, regulatory interactions, and additional functions should be taken into account.
... 5-Aza-2′ deoxycytidine compounds include 5-Aza-2′ deoxycytidine and 5-Aza-2′ deoxycytidine, which may be found in a broad range of chemical forms (Figure 3). To see this, we used 5-Aza to demethylate the GPX3 promoter in SNU1 and the 5-Aza treatment resulted in restored GPX3 mRNA levels ( Figure 2) [19]. ere was no difference in the methylation levels after TSA treatment alone. ...
Article
Full-text available
Stomach cancer is the second largest cause of cancer-related mortality globally, and it continues to be a reason for worry today. Inhalation of the stomach cancer risk factor H. pylori produces large levels of reactive oxygen species (ROS). When combined with glutathione reductase, glutathione peroxidase 3 (GPX3) catalyzes the reduction of hydrogen peroxide and lipid peroxides. To get a better understanding of the GPX3 gene’s role in the illness, the researchers used quantitative real-time RT-PCR to examine the gene’s expression and regulation in gastric cancer cell lines, original gastric cancer samples, and 45 normal stomach mucosa adjacent to malignancies. According to the research, GPX3 expression was decreased or silenced in eight of nine cancer cell lines and 83 percent of gastric cancer samples (90/108) as compared to normal gastric tissues in the vicinity of the tumor (P<0.0001). It was found that 60 percent of stomach cancer samples exhibited DNA hypermethylation after analyzing the GPX3 promoter (P=0.007) (a methylation level of more than 10 percent, as measured by bisulfite pyrosequencing). In stomach tumors, we found a statistically significant reduction in the amount of GPX3 DNA copies (P<0.001). The gene expression of SNU1 and MKN28 cells was restored after treatment with 5-Aza-2′ Deoxycytidine to reduce GPX3 promoter methylation. Genetic and epigenetic alterations lead GPX3 to be dysfunctional in gastric cancer. This indicates that the systems that regulate ROS have been disrupted, and GPX3 may be implicated in the development of gastric cancer, as shown by our results when evaluated alone and in combination.
... GPx3 is also down-regulated in hepatocellular carcinoma and esophageal squamous cell carcinoma through promoter hypermethylation, which may lead to cancer development and progression (He et al., 2011;Cao et al., 2015). Silencing of GPx3 through DNA hypermethylation is associated with lymph node metastasis in gastric cancer and cervical cancer (Peng et al., 2012;Zhang et al., 2014). However, the functional research of GPx3 in gastric cancer needs to be further explored. ...
Article
Full-text available
Gastric cancer is one of the most heterogeneous tumors with multi-level molecular disturbances. Sustaining proliferative signaling and evading growth suppressors are two important hallmarks that enable the cancer cells to become tumorigenic and ultimately malignant, which enable tumor growth. Discovering and understanding the difference in tumor proliferation cycle phenotypes can be used to better classify tumors, and provide classification schemes for disease diagnosis and treatment options, which are more in line with the requirements of today’s precision medicine. We collected 691 eligible samples from The Cancer Genome Atlas (TCGA) and Gene Expression Omnibus (GEO) database, combined with transcriptome data, to explore different heterogeneous proliferation cycle phenotypes, and further study the potential genomic changes that may lead to these different phenotypes in this study. Interestingly, two subtypes with different clinical and biological characteristics were identified through cluster analysis of gastric cancer transcriptome data. The repeatability of the classification was confirmed in an independent Gene Expression Omnibus validation cohort, and consistent phenotypes were observed. These two phenotypes showed different clinical outcomes, and tumor mutation burden. This classification helped us to better classify gastric cancer patients and provide targeted treatment based on specific transcriptome data.
... BGN, NOX4, SPARC, HEYL, SPP1, CTHRC1, SFRP4, FBXO32, KIF4A, MMP9, TIMP1 were related to the proliferation, invasion, prognosis, and recurrence of GC in different level and validated in many kind of tumors [25][26][27][28][29][30][31][32][33][34][35][36][37]. The key gene GPX3 is downregulation in GC cell lines and samples because of abnormal promoter methylation, and related to lymph node metastasis [38]. Cai et al. study proved that GPX3 affected migration and invasion through NFкB/Wnt5a/JNK pathway in GC [39]. ...
Article
Full-text available
Background: To date, the pathogenesis of gastric cancer (GC) remains unclear. We combined public database resources and bioinformatics analysis methods, explored some novel genes and verified the experiments to further understand the pathogenesis of GC and to provide a promising target for anti-tumor therapy. Methods: We downloaded the chip data related to GC from the Gene Expression Omnibus (GEO) database, extracted differentially expressed genes (DEGs), and then determined the key genes in the development of GC via PPI networks and model analysis. Functional annotation via GO and KEGG enrichment of DEGs was used to understand the latent roles of DEGs. The expression of the triggering receptor expressed on myeloid cells 2 (TREM2) gene in GC cell lines was verified via RT-PCR and western blotting. Moreover, the CCK-8, wound healing assay, and transwell migration and invasion assays were used to understand the changes in the proliferation, migration, and invasion abilities of GC cells after silencing TREM2. Western blotting verified the interaction between TREM2 and PI3K predict of the string website, as well as the effect of TREM2 on EMT. Finally, a lung metastasis model was used to explore the relationship between TREM2 and metastasis. Results: Our study identified 16 key genes, namely BGN, COL1A1, COL4A1, COL5A2, NOX4, SPARC, HEYL, SPP1, TIMP1, CTHRC1, TREM2, SFRP4, FBXO32, GPX3, KIF4A, and MMP9 genes associated with GC. The EMT-related pathway was the most significantly altered pathway. TREM2 expression was higher in GC cell lines and was remarkably associated with tumor invasion depth, TNM stage, histological grade, histological type, anatomic subdivision, and Helicobacter pylori state. Knockdown of TREM2 expression inhibited the proliferation, migration, and invasion of GC cells as well as the progression of EMT by PI3K/AKT signaling in vitro. In addition, lung metastasis were decreased in vivo. Conclusions: We identified some important genes associated with the progression of GC via public database analysis, explored and verified the effects of proto-oncogene TREM2 on EMT via the PI3K/AKT pathway. TREM2 may be a novel target in the GC therapy.
... Notably, some of these anti-oxidant enzymes such as GPX3 and GPX7 possess unique tumor suppressor functions, in addition to their anti-oxidant properties. For example, GPX3 has a tumor suppressor role in esophageal adenocarcinoma [59], gastric cancer [60,61], breast cancer [62], prostate cancer [63], and colorectal cancer [64]. Similarly, GPX7 has anti-tumorigenic functions in esophageal [57], and gastric adenocarcinomas [65]. ...
Article
Full-text available
Esophageal adenocarcinoma (EAC) is the dominant form of esophageal malignancies in the United States and other industrialized countries. The incidence of EAC has been rising rapidly during the past four decades. Barrett’s esophagus (BE) is the main precancerous condition for EAC, where a metaplastic columnar epithelium replaces normal squamous mucosa of the lower esophagus. The primary risk factor for BE and EAC are chronic gastroesophageal reflux disease (GERD), obesity and smoking. During the BE-dysplasia-EAC sequence, esophageal cells are under a tremendous burden of accumulating reactive oxygen species (ROS) and oxidative stress. While normal cells have intact antioxidant machinery to maintain a balanced anti-tumorigenic physiological response, the antioxidant capacity is compromised in neoplastic cells with a pro-tumorigenic development antioxidant response. The accumulation of ROS, during the neoplastic progression of the GERD-BE-EAC sequence, induces DNA damage, lipid peroxidation and protein oxidation. Neoplastic cells adapt to oxidative stress by developing a pro-tumorigenic antioxidant response that keeps oxidative damage below lethal levels while promoting tumorigenesis, progression, and resistance to therapy. In this review, we will summarize the recent findings on oxidative stress in tumorigenesis in the context of the GERD-BE-EAC process. We will discuss how EAC cells adapt to increased ROS. We will review APE1 and NRF2 signaling mechanisms in the context of EAC. Finally, we will discuss the potential clinical significance of applying antioxidants or NRF2 activators as chemoprevention and NRF2 inhibitors in treating EAC patients.
Article
Full-text available
Reactive oxygen species (ROS) and reactive nitrogen species (RNS) are well recognized for playing a dual role, since they can be either deleterious or beneficial to biological systems. An imbalance between ROS production and elimination is termed oxidative stress, a critical factor and common denominator of many chronic diseases such as cancer, cardiovascular diseases, metabolic diseases, neurological disorders (Alzheimer’s and Parkinson’s diseases), and other disorders. To counteract the harmful effects of ROS, organisms have evolved a complex, three-line antioxidant defense system. The first-line defense mechanism is the most efficient and involves antioxidant enzymes such as superoxide dismutase (SOD), catalase (CAT), and glutathione peroxidase (GPx). This line of defense plays an irreplaceable role in the dismutation of superoxide radicals (O2•−) and hydrogen peroxide (H2O2). The removal of superoxide radicals by SOD prevents the formation of the much more damaging peroxynitrite ONOO⁻ (O2•− + NO• → ONOO⁻) and maintains the physiologically relevant level of nitric oxide (NO•), an important molecule in neurotransmission, inflammation, and vasodilation. The second-line antioxidant defense pathway involves exogenous diet-derived small-molecule antioxidants. The third-line antioxidant defense is ensured by the repair or removal of oxidized proteins and other biomolecules by a variety of enzyme systems. This review briefly discusses the endogenous (mitochondria, NADPH, xanthine oxidase (XO), Fenton reaction) and exogenous (e.g., smoking, radiation, drugs, pollution) sources of ROS (superoxide radical, hydrogen peroxide, hydroxyl radical, peroxyl radical, hypochlorous acid, peroxynitrite). Attention has been given to the first-line antioxidant defense system provided by SOD, CAT, and GPx. The chemical and molecular mechanisms of antioxidant enzymes, enzyme-related diseases (cancer, cardiovascular, lung, metabolic, and neurological diseases), and the role of enzymes (e.g., GPx4) in cellular processes such as ferroptosis are discussed. Potential therapeutic applications of enzyme mimics and recent progress in metal-based (copper, iron, cobalt, molybdenum, cerium) and nonmetal (carbon)-based nanomaterials with enzyme-like activities (nanozymes) are also discussed. Moreover, attention has been given to the mechanisms of action of low-molecular-weight antioxidants (vitamin C (ascorbate), vitamin E (alpha-tocopherol), carotenoids (e.g., β-carotene, lycopene, lutein), flavonoids (e.g., quercetin, anthocyanins, epicatechin), and glutathione (GSH)), the activation of transcription factors such as Nrf2, and the protection against chronic diseases. Given that there is a discrepancy between preclinical and clinical studies, approaches that may result in greater pharmacological and clinical success of low-molecular-weight antioxidant therapies are also subject to discussion.
Article
Full-text available
Epigenetic alterations, including DNA methylation, histone modification, loss of genome imprinting, chromatin remodeling and noncoding RNAs, are associated with human carcinogenesis. Among them, DNA methylation is a fundamental epigenetic process to modulate gene expression. In cancer cells, altered DNA methylation includes hypermethylation of site-specific CpG island promoter and global DNA hypomethylation. Detection of aberrant gene promoter methylation has been applied to clinic to stratify risk in cancer development, detect early cancer and predict clinical outcomes. Environmental factors associated with carcinogenesis are also significantly related to aberrant DNA methylation. Importantly, epigenetic changes, including altered DNA methylation, are reversible and thus, used as targets for cancer therapy or chemoprevention. An increasing number of recent studies reported DNA methylation level to be a useful biomarker for diagnosis, risk assessment and prognosis prediction for gastrointestinal cancers. This review summarized the accumulated evidence for clinical application to use aberrant DNA methylation levels in gastrointestinal cancers, including colorectal, gastric and esophageal cancer.
Article
The phenotypical and molecular heterogeneity of gastric cancer has hampered the introduction in clinical practice of a unifying classification of the disease. However, as next generation sequencing (NGS) technologies enhanced the comprehension of the molecular landscape of gastric cancer, novel molecular classification systems have been proposed, allowing the dissection of molecular tumor heterogeneity and paving the way for the development of new targeted therapies. Moreover, the use of NGS analyses in the molecular profiling of formalin-fixed paraffin-embedded (FFPE) specimens will improve patient selection for the enrolment in novel clinical trials. In conclusion, the application of NGS in precision oncology will revolutionize the diagnosis and clinical management in gastric cancer patients.
Article
Full-text available
Glutathione peroxidase 3 (GPX3) is the main antioxidant enzyme in plasma. Its biological roles are to protect cells from oxidative stress-induced damage. Several studies have been reported the association between GPX3 expression and its correlation with cancer carcinogenesis including breast cancer. The aim of this research was to investigate the GPX3 messenger ribonucleic acid (mRNA) expression in 82 breast tumors and paired normal breast tissues by SYBR green quantitative real-time reverse transcription-polymerase chain reaction and the association with clinicopathological data. Our results show that GPX3 reduced expression was found significantly associated with number of metastatic lymph nodes (odds ratio [OR] = 3.41, 95% confidence interval [CI] = 1.35–8.64, p = 0.01), no distant metastasis (OR = 5.52, 95% CI = 3.74–11.89, p = 0.04), and nonhormone usage breast cancer patients (OR = 0.19, 95% CI = 0.04–0.93, p = 0.04). This finding suggested that GPX3 plays a role in breast carcinogenesis, and might serve as a prognostic biomarker in breast cancer patients.
Article
Full-text available
The incidence and mortality of gastric cancer have fallen dramatically in US and elsewhere over the past several decades. Nonetheless, gastric cancer remains a major public health issue as the fourth most common cancer and the second leading cause of cancer death worldwide. Demographic trends differ by tumor location and histology. While there has been a marked decline in distal, intestinal type gastric cancers, the incidence of proximal, diffuse type adenocarcinomas of the gastric cardia has been increasing, particularly in the Western countries. Incidence by tumor sub-site also varies widely based on geographic location, race, and socioeconomic status. Distal gastric cancer predominates in developing countries, among blacks, and in lower socioeconomic groups, whereas proximal tumors are more common in developed countries, among whites, and in higher socio-economic classes. Diverging trends in the incidence of gastric cancer by tumor location suggest that they may represent two diseases with different etiologies. The main risk factors for distal gastric cancer include Helicobacter pylori (H pylori) infection and dietary factors, whereas gastroesophageal reflux disease and obesity play important roles in the development of proximal stomach cancer. The purpose of this review is to examine the epidemiology and risk factors of gastric cancer, and to discuss strategies for primary prevention.
Article
Full-text available
Exposure of the oesophageal mucosa to gastric acid and bile acids leads to the accumulation of reactive oxygen species (ROS), a known risk factor for Barrett's oesophagus and progression to oesophageal adenocarcinoma (OAC). This study investigated the functions of glutathione peroxidase 7 (GPX7), frequently silenced in OAC, and its capacity in regulating ROS and its associated oxidative DNA damage. Using in-vitro cell models, experiments were performed that included glutathione peroxidase (GPX) activity, Amplex UltraRed, CM-H(2)DCFDA, Annexin V, 8-oxoguanine, phospho-H2A.X, quantitative real-time PCR and western blot assays. Enzymatic assays demonstrated limited GPX activity of the recombinant GPX7 protein. GPX7 exhibited a strong capacity to neutralise hydrogen peroxide (H(2)O(2)) independent of glutathione. Reconstitution of GPX7 expression in immortalised Barrett's oesophagus cells, BAR-T and CP-A led to resistance to H(2)O(2)-induced oxidative stress. Following exposure to acidic bile acids cocktail (pH4), these GPX7-expressing cells demonstrated lower levels of H(2)O(2), intracellular ROS, oxidative DNA damage and double-strand breaks, compared with controls (p<0.01). In addition, these cells demonstrated lower levels of ROS signalling, indicated by reduced phospho-JNK (Thr183/Tyr185) and phospho-p38 (Thr180/Tyr182), and demonstrated lower levels of apoptosis following the exposure to acidic bile acids or H(2)O(2)-induced oxidative stress. The knockdown of endogenous GPX7 in immortalised oesophageal squamous epithelial cells (HET1A) confirmed the protective functions of GPX7 against pH4 bile acids by showing an increase in the levels of H(2)O(2), intracellular ROS, oxidative DNA damage, double-strand breaks, apoptosis, and ROS-dependent signalling (p<0.01). The dysfunction of GPX7 in oesophageal cells increases the levels of ROS and oxidative DNA damage, which are common risk factors for Barrett's oesophagus and OAC.
Article
Full-text available
Reactive oxygen species (ROS) and reactive nitrogen species (RNS) have been reported to impact gastric inflammation and carcinogenesis. However, the precise mechanism by which Helicobacter pylori induces gastric carcinogenesis is presently unclear. This review focuses on H. pylori-induced ROS/RNS production in the host stomach, and its relationship with gastric carcinogenesis. Activated neutrophils are the main source of ROS/RNS production in the H. pylori-infected stomach, but H. pylori itself also produces ROS. In addition, extensive recent studies have revealed that H. pylori-induced ROS production in gastric epithelial cells might affect gastric epithelial cell signal transduction, resulting in gastric carcinogenesis. Excessive ROS/RNS production in the stomach can damage DNA in gastric epithelial cells, implying its involvement in gastric carcinogenesis. Understanding the molecular mechanism behind H. pylori-induced ROS, and its involvement in gastric carcinogenesis, is important for developing new strategies for gastric cancer chemoprevention.
Article
The annual incidence rates (crude and age-standardized) and numbers of new cases of 25 different cancers have been estimated for the year 1990 in 23 areas of the world. The total number of new cancer cases (excluding non-melanoma skin cancer) was 8.1 million, just over half of which occur in the developing countries. The most common cancer in the world today is lung cancer, accounting for 18% of cancers of men worldwide, and 21% of cancers in men in the developed countries. Stomach cancer is second in frequency (almost 10% of all new cancers) and breast cancer, by far the most common cancer among women (21% of the total), is third. There are large differences in the relative frequency of different cancers by world area. The major cancers of developed countries (other than the 3 already named) are cancers of the colon-rectum and prostate, and in developing countries, cancers of the cervix uteri and esophagus. The implications of these patterns for cancer control, and specifically prevention, ave discussed. Tobacco smoking and chewing are almost certainty the major preventable causes of cancer today. Int. J. Cancer 80:827-841, 1999. (C) 1999 Wiley-Liss, Inc.
Article
The gastrointestinal glutathione peroxidase (GI-GPx) is the fourth member of the GPx family. In rodents, it is exclusively expressed in the gastrointestinal tract, in humans also in liver. It has, therefore, been discussed to function as a primary barrier against the absorption of ingested hydroperoxides. A vital function of GI-GPx can be deduced from the unusual stability of its mRNA under selenium-limiting conditions, the presence of low amounts of GI-GPx protein in selenium deficiency where cGPx is absent, and the fast reappearance of the GI-GPx protein upon refeeding of cultured cells with selenium compared to the slower reappearance of cGPx protein. Furthermore, the Secis efficiency of GI-GPx is low when compared to cGPx and PHGPx. It is, however, almost independent of the selenium status of the cells tested. All these characteristics rank GI-GPx high in the hierarchy of selenoproteins and point to a role of GI-GPx which might be more crucial than that of cGPx, at least in the gastrointestinal system.
Article
The gastrointestinal glutathione peroxidase (GI-GPx) is the fourth member of the GPx family. In rodents, it is exclusively expressed in the gastrointestinal tract, in humans also in liver. It has, therefore, been discussed to function as a primary barrier against the absorption of ingested hydroperoxides. A vital function of GI-GPx can be deduced from the unusual stability of its mRNA under selenium-limiting conditions, the presence of low amounts of GI-GPx protein in selenium deficiency where cGPx is absent, and the fast reappearance of the GI-GPx protein upon refeeding of cultured cells with selenium compared to the slower reappearance of cGPx protein. Furthermore, the Secis efficiency of GI-GPx is low when compared to cGPx and PHGPx. It is, however, almost independent of the selenium status of the cells tested. All these characteristics rank GI-GPx high in the hierarchy of selenoproteins and point to a role of GI-GPx which might be more crucial than that of cGPx, at least in the gastrointestinal system.
Article
Glutathione peroxidase 3 (GPx3), a plasma antioxidant enzyme, maintains genomic integrity by inactivating reactive oxygen species (ROS), known DNA-damaging agents and mediators of cancer chemotherapy response. In this study, we demonstrate that loss of GPx3 expression by promoter hypermethylation is frequently observed in a wide spectrum of human malignancies. Furthermore, GPx3 methylation correlates with head and neck cancer (HNC) chemoresistance and may serve as a potential prognostic indicator for HNC patients treated with cisplatin-based chemotherapy. Our findings support the hypothesis that defects in the antioxidant system may contribute to tumorigenesis of a wide spectrum of human malignancies. GPx3 methylation may have implications in chemotherapy response and clinical outcome of HNC patients.
Article
Helicobacter pylori has been the subject of intense investigation since its culture from a gastric biopsy in 1982. From the beginning, this gram-negative bacterium has provoked the interest of bacteriologists, gastroenterologists, infectious disease specialists, cancer biologists, epidemiologists, pathologists, and pharmaceutical scientists. Pathologists were among the first groups of scientists to reevaluate their data in the context of the newly discovered bacterial etiological agent. Chronic inflammation elicited by the bacterium provided the missing link in the progression to gastric carcinoma; accordingly, H. pylori was named as a class 1 carcinogen by the World Health Organization. Two key papers published in 1991 in the Journal of the National Cancer Institute reported a positive association between gastric cancer and H. pylori infection. This fact provided a strong rationale to treat all who tested positive for H. pylori. Antibiotic regimens have been largely successful, but some agents such as metronidazole and clarithromycin have been rendered ineffective in several countries and geographical areas of the United States by the emergence of strains resistant to these compounds. Although there was some skepticism initially, within few years numerous research groups verified the association of H.pylori with gastric carcinoma. Host related factors for the development of disease can indicate genetic susceptibility (or resistance) or acquired influences, which may stimulate defenses of the host against environmental carcinogens like H.pylori. The present article is a mini-review of the history and epidemiology of the bacterium and its suggested association with the development and progression of gastric cancer.
Article
The role of infectious agents and chronic inflammation in carcinogenesis is being increasingly recognized. It has been estimated that about 18% of cancers are directly linked to infections, particularly gastric adenocarcinoma (Helicobacter pylori), cervical carcinoma (human papilloma viruses), and hepatocarcinoma (hepatitis B and C viruses). Multiple clinical trials of COX-2 inhibitors and other antiinflammatory agents have shown a beneficial effect on the development of diverse tumors, such as those of the colon, stomach, prostate, and breast. However, their mechanism of action is not completely understood and may differ among the infectious agents and tumor types. Because gastric adenocarcinomas account for more than 90% of all gastric malignancies, this review focuses on adenocarcinomas.
Article
Esophageal squamous cell carcinoma (ESCC) is one of the most common causes of cancer mortality in the gastrointestinal tract. Promoter hypermethylation of tumor suppressor genes contributes to gene inactivation during development of ESCC. To identify novel methylation-silenced genes in ESCC. Genome-wide microarrays were applied to search for genes that were markedly upregulated after treatment with 5-aza-2'-deoxycytidine (5-Aza-dC) and that were markedly decreased in tumor tissue compared with paired adjacent nontumor tissue. Reverse-transcription polymerase chain reaction (PCR), immunohistochemistry, methylation-specific PCR, and bisulfite genomic sequencing were employed to investigate expression and methylation of candidate genes in five human ESCC cell lines, two human immortalized normal esophageal epithelial cell lines, primary ESCC tumor tissues, and paired adjacent nontumor tissues. GPX3 was selected as a novel candidate hypermethylated gene in ESCC through microarray analysis. In most ESCC cell lines, GPX3 messenger RNA (mRNA) expression was downregulated and the CpG island of GPX3 promoter was methylated. Demethylation treatment with 5-Aza-dC restored GPX3 mRNA expression. Methylation of GPX3 promoter was more frequent in ESCC tumor tissues (71.4%) than in adjacent nontumor tissues (10.7%) (P < 0.001), and methylation of GPX3 promoter correlated significantly with GPX3 mRNA downregulation. Finally, GPX3 protein expression was also significantly lower in ESCC tumor tissues than in adjacent nontumor tissues. GPX3 is downregulated through promoter hypermethylation in ESCC, which may be a potential biomarker of ESCC.