ArticlePDF Available

Role of promoter hypermethylation in Cisplatin treatment response of male germ cell tumors

Authors:

Abstract and Figures

Male germ cell tumor (GCT) is a highly curable malignancy, which exhibits exquisite sensitivity to cisplatin treatment. The genetic pathway(s) that determine the chemotherapy sensitivity in GCT remain largely unknown. We studied epigenetic changes in relation to cisplatin response by examining promoter hypermethylation in a cohort of resistant and sensitive GCTs. Here, we show that promoter hypermethylation of RASSF1A and HIC1 genes is associated with resistance. The promoter hypermethylation and/or the down-regulated expression of MGMT is seen in the majority of tumors. We hypothesize that these epigenetic alterations affecting MGMT play a major role in the exquisite sensitivity to cisplatin, characteristic of GCTs. We also demonstrate that cisplatin treatment induce de novo promoter hypermethylation in vivo. In addition, we show that the acquired cisplatin resistance in vitro alters the expression of specific genes and the highly resistant cells fail to reactivate gene expression after treatment to demethylating and histone deacetylase inhibiting agents. Our findings suggest that promoter hypermethylation of RASSF1A and HIC1 genes play a role in resistance of GCT, while the transcriptional inactivation of MGMT by epigenetic alterations confer exquisite sensitivity to cisplatin. These results also implicate defects in epigenetic pathways that regulate gene transcription in cisplatin resistant GCT.
Content may be subject to copyright.
BioMed Central
Page 1 of 12
(page number not for citation purposes)
Molecular Cancer
Open Access
Research
Role of promoter hypermethylation in Cisplatin treatment
response of male germ cell tumors
Sanjay Koul
†1
, James M McKiernan
†2
, Gopeshwar Narayan
1
,
Jane Houldsworth
3,4
, Jennifer Bacik
5
, Deborah L Dobrzynski
4
,
Adel M Assaad
1
, Mahesh Mansukhani
1
, Victor E Reuter
6
, George J Bosl
4
,
Raju SK Chaganti
3,4
and Vundavalli VVS Murty*
1,7
Address:
1
Department of Pathology, College of Physicians & Surgeons of Columbia University, 630 West 168th Street, New York, NY 10032, USA,
2
Department of Urology, College of Physicians & Surgeons of Columbia University, 630 West 168th Street, New York, NY 10032, USA,
3
The Cell
Biology Program, Memorial Sloan-Kettering Cancer Center, New York, NY 10021, USA,
4
Department of Medicine, Memorial Sloan-Kettering
Cancer Center, New York, NY 10021, USA,
5
Department of Epidemiology and Biostatistics, Memorial Sloan-Kettering Cancer Center, New York,
NY 10021, USA,
6
Department of Pathology, Memorial Sloan-Kettering Cancer Center, New York, NY 10021, USA and
7
Institute for Cancer
Genetics, College of Physicians & Surgeons of Columbia University, 630 West 168th Street, New York, NY 10032, USA
Email: Sanjay Koul - sk1276@columbia.edu; James M McKiernan - jmm23@columbia.edu; Gopeshwar Narayan - gn110@columbia.edu;
Jane Houldsworth - houldswj@mskcc.org; Jennifer Bacik - bacik1@MSKCC.ORG; Deborah L Dobrzynski - dobrzynd@MSKCC.ORG;
Adel M Assaad - aa2071@columbia.edu; Mahesh Mansukhani - mm322@columbia.edu; Victor E Reuter - reuterv@mskcc.org;
George J Bosl - boslg@mskcc.org; Raju SK Chaganti - chagantr@mskcc.org; Vundavalli VVS Murty* - vvm2@columbia.edu
* Corresponding author †Equal contributors
Abstract
Background: Male germ cell tumor (GCT) is a highly curable malignancy, which exhibits exquisite
sensitivity to cisplatin treatment. The genetic pathway(s) that determine the chemotherapy
sensitivity in GCT remain largely unknown.
Results: We studied epigenetic changes in relation to cisplatin response by examining promoter
hypermethylation in a cohort of resistant and sensitive GCTs. Here, we show that promoter
hypermethylation of RASSF1A and HIC1 genes is associated with resistance. The promoter
hypermethylation and/or the down-regulated expression of MGMT is seen in the majority of
tumors. We hypothesize that these epigenetic alterations affecting MGMT play a major role in the
exquisite sensitivity to cisplatin, characteristic of GCTs. We also demonstrate that cisplatin
treatment induce de novo promoter hypermethylation in vivo. In addition, we show that the
acquired cisplatin resistance in vitro alters the expression of specific genes and the highly resistant
cells fail to reactivate gene expression after treatment to demethylating and histone deacetylase
inhibiting agents.
Conclusions: Our findings suggest that promoter hypermethylation of RASSF1A and HIC1 genes
play a role in resistance of GCT, while the transcriptional inactivation of MGMT by epigenetic
alterations confer exquisite sensitivity to cisplatin. These results also implicate defects in epigenetic
pathways that regulate gene transcription in cisplatin resistant GCT.
Published: 18 May 2004
Molecular Cancer 2004, 3:16
Received: 06 May 2004
Accepted: 18 May 2004
This article is available from: http://www.molecular-cancer.com/content/3/1/16
© 2004 Koul et al; licensee BioMed Central Ltd. This is an Open Access article: verbatim copying and redistribution of this article are permitted in all
media for any purpose, provided this notice is preserved along with the article's original URL.
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 2 of 12
(page number not for citation purposes)
Background
Adult male germ cell tumors (GCTs) are considered to be
a model system for a curable malignancy because of their
exquisite sensitivity to cisplatin (CDDP)-based combina-
tion (cisplatin, etoposide, with or without bleomycin)
chemotherapy. Histologically, GCTs present as a germ cell
(GC)-like undifferentiated seminoma (SGCT) or a differ-
entiated nonseminoma (NSGCT). NSGCTs display com-
plex differentiation patterns that include embryonal,
extra-embryonal, and somatic tissue types [1]. Further-
more, embryonal lineage teratomas differentiate into var-
ious somatic cell types that may undergo malignant
transformation to epithelial, mesenchymal, neurogenic,
or hematologic tumors [2]. Seminomas are exquisitely
sensitive to radiation therapy while NSGCTs are highly
sensitive to treatment with CDDP-based chemotherapy.
Despite this sensitivity to chemotherapy, 20–30% of met-
astatic tumors are refractory to initial treatment, requiring
salvage therapy and accounting for high mortalitiy. Such
patients are treated with high dose and experimental
chemotherapy protocols [3]. The underlying molecular
basis of this exquisite drug responsiveness of GCT remains
to be fully understood [4].
Little is known about the genetic basis of chemotherapy
response in GCT. Studies have previously identified that
TP53 mutations and gene amplification may play a role in
GCT resistance [5,6]. It has also been recently shown that
microsatellite instability is associated with the treatment
resistance in GCT [7]. An epigenetic alteration by pro-
moter hypermethylation that plays a role in inactivating
tumor suppressor genes in a wide-variety of cancers also
has been shown to occur in GCT [8-10]. We previously
showed the absence of promoter hypermethylation in
SGCT and acquisition of unique patterns of promoter
hypermethylation in NSGCT [8]. However, the role of
such epigenetic changes in GCT resistance and sensitivity
remains unknown.
In the present study, we evaluated the status of hyper-
methylation in 22 gene promoters in 39 resistant and 31
sensitive NSGCTs. We found that RASSF1A and HIC1 pro-
moter hypermethylation was associated with highly resist-
ant tumors. Evidence was also obtained suggesting that
promoter hypermethylation is induced against the initial
CDDP treatment and that this hypermethylation plays a
crucial role in further treatment response. We show that
changes in the patterns of gene expression occur during
the in vitro acquisition of a highly refractory tumor to
CDDP, which irreversibly affects the response to demeth-
ylating and histone deacetylase inhibiting agents.
Results
Promoter hypermethylation in relation to chemotherapy
resistance and sensitivity
Based on our previous observations in GCT, we studied 22
gene promoters for hypermethylation in 70 NSGCTs
derived from 60 patients [8]. Promoter hypermethylation
was found in nine of 22 genes examined. One or more
genes were methylated in 41 (59%) tumors. The fre-
quency of hypermethylation for each of the genes was:
RASSF1A (35.7%), HIC1 (31.9%), BRCA1 (26.1%), APC
(24.3%), MGMT (20%), RARB (5.7%), FHIT (5.7%),
FANCF (5.7%), and ECAD (4.3%). This frequency was
similar to our previously published observations on unse-
lected patients with NSGCTs [8].
The frequency of overall promoter hypermethylation (one or
more of the 22 genes methylated) was similar in the sensitive
(18 of 29 patients; 62%) and resistant (21 of 31 patients;
68%) tumors. However, the frequency of promoter hyper-
methylation of individual genes differed between sensitive
and resistant tumors. RASSF1A (52% in resistant vs. 28% in
sensitive) and HIC1 (47% in resistant vs. 24% in sensitive)
genes showed higher frequency of promoter hypermethyla-
tion in resistant tumors (Table 2, Fig. 1). These differences
were not statistically significant due to small number of
tumors studied. However, the differences were more pro-
nounced when the sensitive and highly resistant tumors were
compared (discussed below). On the other hand, the sensitive
tumors exhibited higher frequency of promoter hypermethyl-
ation compared to resistant tumors in MGMT (31% vs. 13%)
and RARB (14% vs. 0%; P = 0.05) (Table 2, Fig. 1). Other
genes that exhibited frequent hypermethylation showed no
significant differences (APC, 24% vs. 29%; BRCA1, 31% vs.
30%) between the sensitive and resistant groups. These data,
thus, suggest that promoter hypermethylation of RASSF1A
and HIC1 is associated with chemotherapy resistance pheno-
type, while promoter hypermethylation of MGMT and RARB
genes is commonly seen in sensitive tumors.
Table 1: Histologic and phase characteristics of sensitive and
resistant NSGCTs
Sensitive
(N = 31)
Resistant
(N = 39)
Histology
Teratoma 14 15
Embryonal carcinoma 6 2
Yolk sac tumor 3 6
Mixed tumor/malignant transformation 8 16
Phase at tissue collection
Primary untreated (P) 1 5
Metastatic untreated (M1) 12 4
One regimen of chemotherapy (C1) 18 14
Two regimens of chemotherapy (C2) - 8
Three or more regimens of
chemotherapy (C3)
-8
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 3 of 12
(page number not for citation purposes)
CDDP treatment induces de novo promoter
hypermethylation in vivo
To assess the effect of CDDP-treatment on promoter
hypermethylation, we examined tumor tissues that were
collected at different phases of resistance (Table 1). The
frequency of hypermethylation at different phases was: P,
16.7%; M1, 37.5%; C1, 75%; C2, 62.5%, and C3, 62.5%.
Tumors from patients who underwent one or more regi-
mens of chemotherapy (C1, C2, or C3 phases) exhibited
a significantly (P = 0.001) higher (34 of 48 patients; 71%)
frequency of promoter hypermethylation compared to
those from untreated (P and M1) (7 of 22 tumors; 32%)
patients after adjusting for sensitive/resistance status. The
frequency of promoter hypermethylation was also
significantly higher in tumors from treated patients when
sensitive and resistant groups were analyzed separately (P
0.02). The differences in overall promoter hypermethyl-
ation between untreated tumors (P/M1; 32%) and C1
tumors (75%) were highly significant (P = 0.004), while
the differences between untreated tumors and C2/C3
Promoter hypermethylation in patients with sensitive and resistant GCTs in response to cisplatin combination chemotherapyFigure 1
Promoter hypermethylation in patients with sensitive and resistant GCTs in response to cisplatin combination chemotherapy.
RASSF1A and HIC1 genes showed frequent methylation in resistant tumors, while MGMT and RARB promoters were more
commonly methylated in sensitive tumors.
Table 2: Frequency of promoter methylation of individual genes in sensitive and resistant NSGCT
Gene Sensitive
1
(N = 29) (%) Resistant
2
(N = 31) (%) P-value
APC 7 (24) 9 (29) 0.77
BRCA1
3
9 (31) 9 (30) 1.0
ECAD 1 (3) 2 (6) 1.0
FANCF 2 (7) 2 (6) 1.0
FHIT 2 (7) 2 (6) 1.0
HIC1
3
7 (24) 14 (47) 0.10
MGMT 9 (31) 4 (13) 0.12
RARB 4 (14) 0 0.05
RASSF1A 8 (28) 16 (52) 0.07
1
Consists of tumors, with or without retroperitoneal lymph node, sensitive for one cycle of chemotherapy
2
Consists of tumors, with or without
retroperitoneal lymph node, required of two or more cycles of chemotherapy, all resistant tumors, and all patients died of disease
3
Only 30
resistant tumors studied for methylation status
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 4 of 12
(page number not for citation purposes)
Table 3: Promoter hypermethylation in various phases of treatment in NSGCT
Gene Phase
1
p-value
3
P/M1 (N = 22) C1 (N = 32) C2/C3 (N = 16) P/M1 vs C1 P/M1 vs C2/C3
APC 2 (9.1) 8 (25) 7 (43.8) 0.25 0.03
BRCA1
2
4 (18.2) 12 (37.5) 2 (13.3) 0.22 1.0
HIC1
2
3 (13.6) 11 (34.4) 8 (53.5) 0.17 0.09
MGMT 0 11 (34.4) 3 (18.8) 0.003
4
RARB 0 4 (12.5) 0 0.14
4
RASSF1A 5 (22.7) 10 (31.3) 10 (62.5) 0.72 0.09
1
See Table 1 for definition of phase
2
Only 15 tumors studied in C2/C3 phase
3
Adjusted for sensitive/resistant status
4
Model cannot be fit due to
sparse data; Fisher's exact p-value given
Comparisons of promoter hypermethylation frequencies in different phases of treated patients with GCTsFigure 2
Comparisons of promoter hypermethylation frequencies in different phases of treated patients with GCTs. Phase definitions
are shown in Table 1. P, untreated primary tumor; M1, untreated metastatic tumor; C1, one regimen of chemotherapy; C2/C3,
two or more regimens of chemotherapy. Promoter methylation of RASSF1A, HIC1, and APC genes was significantly high in
resistant tumors. The MGMT, BRCA1, and RARB genes show higher frequency of promoter methylation in tumors exposed to
one cycle of chemotherapy.
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 5 of 12
(page number not for citation purposes)
tumors (62.5%) were less pronounced (P = 0.09). These
data, thus, suggest that the higher frequency of promoter
hypermethylation seen in treated tumors may be due in
response to CDDP treatment. This de novo increase in
promoter hypermethylation was evident in most analyzed
genes. Comparison among tumors that were not treated
(P/M1), those that were treated with one cycle of chemo-
therapy (C1) and those that were treated with two or more
cycles (C2/C3) showed notable differences for APC,
MGMT, HIC1, RARB and RASSF1A (Table 3). Promoter
hypermethylation of RASSF1A, HIC1, and APC genes was
higher in the treated tumors, with highly resistant (C2/
C3) tumors exhibiting the highest incidence (Fig. 2).
However, this trend was different for BRCA1, MGMT, and
RARB genes. These genes exhibited either no methylation
or low frequency of hypermethylation in untreated
tumors, while the C1 tumors had the highest incidence of
promoter hypermethylation (Fig. 2). However, this was
decreased or absent in highly resistant (C2/C3) tumors.
These data, thus, strongly suggest that promoter hyper-
methylation was induced in response to the first exposure
to CDDP and the tumors harboring promoter hypermeth-
ylation responded differently to further treatment in a
gene specific manner. Thus, our data indicate that the
tumors with promoter hypermethylation of RASSF1A,
HIC1, and APC resulted in failure to respond to further
treatment, while the tumors harboring promoter hyper-
methylation of MGMT, RARB, and BRCA1 responded
favorably.
Since we previously showed that yolk sac tumor (YST)
exhibit higher frequency of hypermethylation compared
to other histologic types among the genes tested [8], we
included histology as an additional covariate in the above
analyses in an attempt to account for histological differ-
ences in promoter hypermethylation. We found that the
differences in overall promoter hypermethylation
between untreated and treated tumors were no longer
significant when adjusted for histology (data not shown).
The small number of observations in each histological
group, however, prevents us from making any meaningful
conclusions from this analysis. Although the data is indic-
ative of histologic differences in promoter methylation,
further analysis of gene specific promoter hypermethyla-
tion on a larger panel of tumors is needed to satisfactorily
address this issue.
No effect of CDDP treatment in vitro on promoter
methylation
Since we showed CDDP treatment induces promoter
hypermethylation in tumors in vivo, we wanted to test
whether a similar phenomenon occurs in vitro. To investi-
gate this, we exposed four NSGCT cell lines to two differ-
ent concentrations of CDDP for various time periods as
described in the methods. The specimens from which the
169A, 218A, and 240A cell lines derived were also
included in the panel of tumors studied for hypermethyl-
ation. In addition, two independent clones derived from
833K-E and 240A as D1 and D4 resistant cells (see mate-
rials and methods) were also examined. We did not detect
changes in promoter hypermethylation in 18 (APC,
GSTP1, BRCA1, DAPK, p16, p14, MGMT, APAF1,
RASSF1A, HIC1, RB, TIMP3, FANCF, RARB, CDH1, TP53,
FHIT, and MLH1) genes examined. These results clearly
indicate that CDDP treatment in vitro, within the tested
concentrations, does not cause promoter hypermethyla-
tion. However, we found methylation of BRCA1,
RASSF1A, or HIC1gene promoters in three cell lines but
not in their corresponding primary tumors (Fig. 3). The
genes methylated in cultured tumor cells were the same
genes that were also frequently methylated in NSGCT
patients. The gene promoters that did not exhibit frequent
methylation in primary tumors were not methylated in
cultured tumor cells.
Loss of activation of gene expression to inhibitors of
methylation and histone deacetylation in acquired CDDP
resistance in vitro
To further examine the effect of CDDP treatment in vitro,
we then studied expression of MGMT, HIC1, and FANCF,
the genes that were either promoter hypermethylated or
down regulated in GCT, in D1 and D4 clones derived
from the cell lines 833K-E and 240A.
MGMT is a DNA repair enzyme that protects cells against
the effects of alkylating agents by removing adducts
formed at the O
6
position of guanine in DNA [11]. Tumor
sensitivity to alkylating agents has been shown to depend
on MGMT expression [12]. Resistant tumors are generally
Promoter methylation changes in vitro in GCT cell linesFigure 3
Promoter methylation changes in vitro in GCT cell lines.
Appearance of de novo promoter hypermethylation in cell
lines. T, tumor; CL, cell line; CDDP, cisplatin-treated cell line.
M, methylated DNA; U, unmethylated DNA; Cell lines exam-
ined are shown below. Note absence of methylation in pri-
mary tumor in both the cell lines and appearance of novel
methylated allele in cell line DNA. Note both alleles of
RASSF1A are methylated in 218A cell line.
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 6 of 12
(page number not for citation purposes)
shown to express high levels of the MGMT gene [12]. To
examine the role of MGMT in in vitro-acquired CDDP
resistance of GCT, we studied mRNA levels in 833K-E and
240A cell lines that harbored methylated promoters. Both
showed low levels of detectable mRNA by RT-PCR. How-
ever, only the 833K-E showed reactivation of expression
upon treatment with 5-Aza-C or TSA, while these agents
had no effect on the 240A cell line. The levels of MGMT
mRNA in D1 and D4 refractory cells were either remained
at the levels similar to untreated cells in D1 cells or were
slightly increased in D4 cells in both the cell lines (Fig. 4).
The latter is consistent with the role of MGMT in effi-
ciently repairing DNA adducts in resistant cells and this
could be due to partial demethylation of the promoters in
highly resistant D4 cells in these cell lines. Concordant
with demethylation of promoter in D4 cells, these cells
fail to up-regulate gene expression after 5-Aza-C or TSA
treatment (Fig. 4).
The HIC1 gene showed hypermethylated promoters in
both 833K-E and 240A cell lines. Analysis of mRNA
showed that only 240A cell line exhibited a detectable
level of expression, while it was absent in the 833K-E cell
line. HIC1 was reactivated after treatment with 5-Aza-C or
TSA in 833K-E, while the treatment had no effect on 240A
cell line (Fig. 4). However, both the D1 and D4 clones
derived from these cell lines showed complete lack of
expression of HIC1 and failed to respond to either 5-Aza-
C or TSA (Fig. 4).
The FANCF gene belongs to the family of six Fanconi ane-
mia proteins that facilitate mono-ubiquitinylation of
FANCD2, which plays a role in a large multimeric protein
complex required for DNA repair [13]. Acquired CDDP
resistance in ovarian carcinoma correlates with subtle
methylation/demethylation of FANCF promoter leading
to the suggestion that demethylation of this gene causes
CDDP resistance [14]. Here, we examined the FANCF
gene expression in the development of in vitro CDDP
resistance. The FANCF promoter was not methylated in
both 833K-E and 240A cells and low levels of mRNA
expression were found in both. However, both cell lines
showed an up-regulated expression of FANCF mRNA after
5-Aza-C or TSA treatment. Although the levels of FANCF
expression in D1 cells of 833K-E remain at the levels in
untreated cells, the D1 cells of 240A showed an elevated
level of mRNA (Fig. 4). However, the D4 resistant clones
from both cell lines showed a decrease in expression com-
pared to untreated cells and lost the ability to respond to
5-Aza-C or TSA (Fig. 4). Analysis of a control gene did not
affect the pattern of expression in relation to in vitro
CDDP resistance in these cells (Fig. 4).
Taken together, the results obtained from all three genes
studied here indicate that changes in gene expression
occur in the development of low to high refractory CDDP
resistance. Highly resistant clones fail to respond to
demethylating or histone deacetylase inhibiting agents in
activating gene expression suggesting that irreversible
changes occur in a pathway that control gene
transcription.
MGMT is partially methylated and down regulated in most
GCTs
We showed earlier promoter hypermethylation of MGMT
in 21% NSGCT and complete lack of or down-regulated
gene expression by RT-PCR in 96% of tumors [8]. To
examine whether the down-regulated RNA levels reflected
in decreased protein, we performed an immunohisto-
chemical analysis of MGMT on a tissue array containing
18 SGCTs and 18 NSGCTs. The MGMT expression was
absent in 33 (91.7%) tumors. The remaining 3 tumors
(two yolk sac tumor and one immature teratoma) were
weakly positive compared to the controls (data not
shown). Thus, the combined data on analyses of RT-PCR
from our previous study and the levels of protein reported
here showed down-regulated levels of MGMT expression
in all histologic types of GCT. The cells from GCT have
been reported to exhibit reduced efficiency in the removal
of CDDP-induced adducts [15,16]. Concordantly, more
RT-PCR analysis of gene expression in relation to CDDP-induced resistance in GCT cell linesFigure 4
RT-PCR analysis of gene expression in relation to CDDP-
induced resistance in GCT cell lines. Aza, 5-Aza-2'-deoxycyti-
dine; TSA, trichostatin; D1, low refractory CDDP resistant
cells; D4, high refractory CDDP resistant cells; SEPT6, septin
6. Actin (empty arrow) was used as an internal control. Filled
arrowhead indicates the specific gene. Septin 6 gene was
used as another control.
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 7 of 12
(page number not for citation purposes)
than 85% of metastatic GCT can be cured with chemo-
therapy [3]. Since GCTs exhibit either promoter hyper-
methylation or down-regulated gene expression of the
MGMT gene in most tumors, we suspected that this gene
may play a role in CDDP sensitivity. Our previous data
suggest that down-regulated expression of MGMT occurs
by other mechanisms, in addition to complete promoter
hypermethylation. We previously ruled out mutational
inactivation of this gene in GCT [8]. The MSP method
only detects complete hypermethylation of CpG islands
and partial methylation will not be identified by this
method. Since the down regulated MGMT expression can
be reactivated after cellular exposure to 5-aza-C in
unmethylated GCT cell lines, we reasoned that partial
methylation might exist. To test this, we sequenced 98
CpG methylation sites spanning the promoter of MGMT
in genomic DNA from two normal testes, one methylated
cell line, seven NSGCTs and five SGCTs which have
unmethylated promoters by MSP (Fig. 5). Normal testes
showed a few random methylated CpGs in a small
number of clones. The MSP positive Tera-2 cell line had
dense methylation of the entire MGMT promoter. All
seven NSGCTs and five SGCTs showed partial methyla-
tion of the promoter region at specific CpG residues (Fig.
5). Three (T-228A, T-186B, and T-288A) of the seven
NSGCTs studied were also classified as resistant GCTs. The
frequency of methylated residues varied within clones
derived from the same tumor and between tumors. The
present data, thus, suggest, but does not prove, that partial
methylation of MGMT promoter accounts for down-regu-
lated gene expression in GCT. Overall, these results indi-
cate a role of MGMT promoter hypermethylation state,
either complete or partial, in determining the exquisite
sensitivity to CDDP in GCT.
Discussion
The molecular mechanisms that determine the curability
of GCT to CDDP-based combination chemotherapy are
unclear [17-20]. Understanding the genetic basis of this
exquisite sensitivity could lead to the development of a
more effective treatment for resistant tumors. A number of
genetic mechanisms for CDDP resistance, such as
enhanced adduct repair, drug inactivation, or tolerance to
DNA damage, have been proposed [21].
We and others previously reported that epigenetic altera-
tions in the promoters of specific genes occur in NSGCT
[8-10,22]. We also showed that promoter hypermethyla-
tion was associated with gene repression in NSGCT and
this down regulated expression is reactivated upon
demethylation suggesting a potential role for epigenetic
changes in GCT biology [8]. These results prompted us to
examine the possible involvement of epigenetic changes
in chemo-sensitivity and resistance in GCT. To achieve
this, we investigated epigenetic changes in resistant and
sensitive NSGCTs and found a high incidence of promoter
hypermethylation of RASSF1A and HIC1 in resistant
tumors, while promoter hypermethylation of MGMT and
RARB genes was associated with sensitive tumors.
RASSF1A has shown to be epigenetically inactivated in a
wide variety of tumor types suggesting a major role for this
gene in cancer [23]. In the present study, we demonstrated
that a higher frequency of resistant tumors carry promoter
methylation compared to sensitive GCT, suggesting that
RASSF1A hypermethylation is associated with the resist-
ance phenotype. RASSF1A represents a long isoform of
human RASSF1 gene, which encodes a diacylglycerol
(DAG)-binding domain at the NH
2
terminus, a RAS-asso-
ciation domain at COOH terminus, which interacts with
the XPA protein. RASSF1A gene functions as a negative
regulator of cell growth [23].
Hic1encoding a zinc finger transcription factor acts as a
tumor suppressor gene [24]. HIC1 is silenced by promoter
hypermethylation in several types of human cancer [25].
We found a higher frequency of resistant tumors harbor-
ing HIC1 promoter hypermethylation.
On the other hand, we showed promoter hypermethyla-
tion of MGMT and RARB genes associated with CDDP
sensitivity. MGMT gene encodes O(6)-methylguanine-
DNA methyltransferase and plays an important role in
removing DNA adducts formed by alkylating agents [11].
Epigenetic silencing of MGMT has been shown to confer
enhanced sensitivity on cancer cell to alkylating agents,
while the lack of methylation and high-levels of protein
expression contribute to drug-resistance phenotype
[12,26,27]. We showed here that either complete or par-
tial methylation of MGMT occurs in a majority of GCT.
These data suggest that the complete promoter methyla-
tion of MGMT plays a role in favorable response to CDDP
treatment. However, the demonstration here of partial
methylation in most GCTs provides a possible mecha-
nism for down-regulated expression of MGMT, which is
commonly seen in this tumor. These results, thus, support
the view that the epigenetic alteration in MGMT may be a
factor in the exquisite sensitivity of GCT to CDDP. Such a
model provides opportunities to alter MGMT pathway
and chemosensitize relapsed tumor to CDDP.
Retinoids control gene transcription by activating retinoic
acid receptors (RAR, β and γ) and retinoid X receptors
(RXR, β and γ). Expression of these receptors regulates
organogenesis, organ homeostasis, cell growth, and differ-
entiation and death [28]. It is well established that
changes in expression of RARs play a major role in cancer
development and response of tumor cells to treatment of
all-trans retinoic acid (ATRA). A number of premalignant
lesions and cancers have been shown to exhibit a loss of
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 8 of 12
(page number not for citation purposes)
expression of RARB due to promoter hypermethylation
[28]. In the present study, we found RARB promoter
hypermethylation only in sensitive NSGCTs. The data,
therefore, suggest that RARB down-regulation may favor
response to CDDP treatment. The mechanisms regulated
by RARB responsive genes in GCT require further studies
to understand the role of this gene in chemotherapy
response.
Analysis of promoter methylation by bisulphite sequencing in GCTFigure 5
Analysis of promoter methylation by bisulphite sequencing in GCT. CpG methylation status in 5 independent clones sequenced
is shown for each tumor. Shown on top by vertical lollypops connected to horizontal line are CpGs examined on the 5' pro-
moter region. Site of MSP primer is indicated. Shown below is the CpG sequence information for each tumor. Tumor numbers
are shown on left. Methylation status is indicated in parenthesis next to the tumor. Filled circles indicate methylated CpG sites
and empty circles indicate unmethylated CpGs. NSGCT, nonseminomatous germ cell tumor; SGCT, seminomatous germ cell
tumor.
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 9 of 12
(page number not for citation purposes)
Previous studies indicated that tumor cells exposed to
anticancer agents induce DNA hypermethylation resulting
in the silencing of genes that play role in drug metabolism
and resistance [29]. Human tumor cells exposed to high
concentrations of CDDP in vitro induces alterations in 5-
methyl Cytosine (5-mCyt) [30]. Similarly, in vivo exposure
of bone marrow cells to cytosine arabinoside (araC) alone
or a combination of hydroxyurea, VP-16 and araC also
result in a several-fold increase of 5-mCyt content in
leukemic blasts ([30]. Thus, the exposure of tumor cells to
cytotoxic chemotherapy agents in vitro and in vivo causes
an induction of DNA hypermethylation. In the present
study, we examined whether CDDP treatment in vivo
causes such a hypermethylation in GCT by studying spe-
cific gene promoters. Our results suggest that initial
CDDP treatment in tumors induces promoter hypermeth-
ylation of certain gene promoters such as MGMT, RARB,
and BRCA1 (Fig. 2). This induction of methylation in
these genes hypersensitize the tumor to further treatment,
while tumors that had promoter hypermethylation of
RASSF1A, HIC1, and APC are selected upon further treat-
ment to develop drug resistance (Figs. 1 and 2). Such a
model of CDDP-induced non-random hypermethylation
can predict response to further treatment and allows spe-
cific gene targeted therapeutic approaches for resistant
GCTs. Hypermethylation of RASSF1A, HIC1, and APC
genes provides a plausible mechanism for the propensity
of these tumors to CDDP resistance, and demethylation
could result in restoration of hypersensitivity. Well docu-
mented evidence in certain tumor types suggest that drug
resistance disrupts general mechanisms of chemosensitiv-
ity by targeting mutations and gene amplifications [31].
Here, we demonstrate that epigenetic alterations in spe-
cific genes also play a role in chemosensitivity to CDDP in
GCT. The present data, thus, suggest that specific gene pro-
moter hypermethylation induced by drugs may serve as
prognostic indicator of treatment response in NSGCT. In
view of the biological relevance of DNA methylation,
CDDP-induced hypermethylation shown here in GCTs
will have clinical significance in drug-response
phenotypes.
CDDP-induced promoter hypermethylation in tumor
cells might set in motion a cascade of ectopic gene expres-
sion events that might release tumor from normal home-
ostatic controls. These changes include deamination of 5-
methyl cytosine in CpG causing genetic instability (i.e.,
mutations), transducing epigenetic changes into genetic
alterations, or inactivation of methylated genes. To test
the latter possibility, we tested gene expression in four dif-
ferent clones from two highly resistant cell lines. We could
not reactivate the gene expression by exposure to the
demethylating agent 5-Aza-C or histone deacetylase
inhibitor TSA, implying that a common epigenetic and/or
genetic mechanisms that regulate transcriptional activa-
tion of hypermethylated genes was affected in highly
resistant cells rather than simple methylation changes in
specific gene promoters.
The cytotoxic effectiveness of CDDP against tumor cell is
believed to be mediated through the formation of DNA
adducts, which inhibit DNA replication and transcription
[32,33]. Cisplatin primarily forms intra-strand GpG cross-
links, which are removed by neucleotide excision repair
(NER) [34]. Highly regulated steps involving a number of
proteins coordinate the NER in human cells. One hypoth-
esis to explain the hypersensitivity of GCT to CDDP is that
there is a deficiency in one or more components of this
repair machinery [35]. Recently, it has been shown that
elevated testis-specific high-mobility group (ts-HMG)
DNA-binding proteins may enhance sensitivity to CDDP
[34]. Our results suggest a potential molecular mecha-
nism of CDDP-induced transcriptional inactivation of
genes prone to hypermethylation. The CDDP exposure
may cause genetic damage that might sequester essential
proteins from their designated function such as elements
of DNA repair pathways. The results presented here sup-
port the notion that epigenetic mechanisms play a role in
CDDP-response in a gene specific manner. As cellular
response to CDDP treatment in GCT is believed to be a
complex process, future studies to address this issue need
to examine both epigenetic and genetic alterations.
Conclusions
Our studies provide evidence that the RASSF1A and HIC1
inactivation by promoter hypermethylation play a role in
NSGCT resistance and may serve as markers for the iden-
tification of resistant tumors. The epigenetic alterations in
MGMT may be an important factor in conferring the
exquisite sensitivity of GCT to CDDP. Although the
molecular mechanisms of GCT resistance are unclear cur-
rently, our findings of epigenetic alterations in the
RASSF1A, HIC1, MGMT, and RARB genes may serve as
prognostic indicators of CDDP-related treatment
response and provide molecular targets of therapy to
chemo-sensitize the resistant tumors. In view of the bio-
logical relevance of DNA methylation, CDDP-induced
hypermethylation shown here in GCTs will have clinical
significance in drug-response phenotypes and provides
opportunities to modulate pathways controlled by these
genes.
Methods
Tumor specimens and stratification of chemotherapy
resistance and sensitivity
Tumor tissues were identified by retrospective review of
GCT specimens obtained during diagnostic evaluation at
the Memorial Sloan-Kettering Cancer Center, New York,
between 1987 and 1999. Patients were identified based
on known response and resistance to chemotherapy. A
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 10 of 12
(page number not for citation purposes)
total of 70 GCT specimens from 60 patients comprised
the study cohort. The sensitive tumors consisted of 31 tis-
sues obtained from 29 patients that were relapse-free for
more than two years as a result of chemotherapy alone or
in combination with surgery. The resistant panel com-
prised of 39 tumors from 31 patients, with or without ret-
roperitoneal lymph node metastasis, who either did not
respond to one or more cycles of CDDP-based chemo-
therapy or responded and then relapsed, or died of disease
after any number of cycles of treatment. Table 1 summa-
rizes the histologic and phase characteristics of sensitive
and resistant patients. Additionally, 36 unselected GCTs
evaluated at Columbia University Medical Center were
also studied.
Cell lines and drug treatment
Four NSGCT cell lines, an established 833K-E and three
cell lines (169A, 240A and 218A) described by us earlier,
were grown in high-glucose DMEM medium containing
15% fetal bovine serum, L-glutamine, and penicillin-
streptomycin [36]. Sub-cultured cells after 24 hr were
treated with the specific drugs at different concentrations
and time periods. Cells in logarithmic phase were exposed
to CDDP at 0.5 µM and 1.0 µM concentrations for 2 h and
24 hr, at which time drug was removed, and fresh culture
medium was added. The CDDP-treated cells were contin-
ued to grow for 2, 4, and 7 days to clonally expand. We
have also derived CDDP-refractory cells from 833K-E and
240A cell lines by growing for 21 and 16 days, respec-
tively. Two independent clones derived from each of these
cell lines were designated as 833K-E/C10, 833K-E/C13,
240A/C4, and 240A/C10. After further expansion, these
cells were designated as D1-resistant cells for one time
point drug treatment. These D1 cells were further treated
serially with increasing concentrations (1.5 µM to 4.5 µM)
of CDDP and were grown in culture for more than 90
days. The final passage cells were designated as D4 cells
for four time points of drug selection. Untreated and the
CDDP-resistant cell lines were exposed to demethylating
agent 5-Aza-2' deoxycytidine (5-Aza-C) (Sigma) for five
days at a concentration of 2.5 µM and trichostatin (TSA)
at 250 nM for the last 24 hours or a combination of both.
Methylation Specific PCR (MSP) and gene expression
Genomic DNA was treated with sodium bisulphite as pre-
viously described [8]. Placental DNA treated in vitro with
SssI methyltransferase (New England Biolabs, Beverly,
MA) and similarly treated normal lymphocyte DNA were
used as controls for methylated and unmethylated tem-
plates, respectively. The primers used for methylated and
unmethylated-specific PCR have been either described
previously [8] or are available from authors upon request.
PCR products were run on 2% agarose gels and visualized
after ethidium bromide staining.
Gene expression was assessed on total RNA isolated from
four normal testes, a commercially purchased normal tes-
tis RNA (Clontech, Palo Alto, CA) and the cell lines
described above. Reverse transcription was performed
using random primers and the Pro-STAR first strand RT-
PCR kit (Stratagene, La Jolla, CA). A semi-quantitative
analysis of gene expression in replicate experiments was
performed using 26 to 28 cycles of multiplex RT-PCR with
β-actin (ACTB) as a control and gene specific primers
spanning at least 2 exons whenever possible. The gene
primers used have either been described previously [8] or
are available from authors. The PCR products were run on
1.5% agarose gels, visualized by ethidium bromide stain-
ing and quantitated using the Kodak Digital Image Analy-
sis System (Kodak, New Haven, CT).
Bisulphite sequencing
Bisulphite-treated DNA was amplified with primers
designed to amplify both methylated and unmethylated
DNA. Two sets of primers were designed to cover the
entire promoter region of the MGMT gene. The first set of
primers was MGMT-cl-F3 5'-AGGATTTGAGAAAAGTAA-
GAGAG-3' and MGMT-cl-R3 5'-ATT-
TAACAAACTAAAACACAAAACC-3', and the second set of
primers was MGMT-cl-F4 5'-TTTTTTTGTTTTTTTTAG-
GTTTT-3' and MGMT-cl-R4 5'-CAAACACCAACCAT-
AATAACCAA-3'. PCR products were sub-cloned into
pCR2.1-TOPO (Invitrogen) and DNA isolated from 15 to
20 clones for each tumor was sequenced.
Tissue microarray and immunohistochemical analysis
A panel of 36 unselected formalin-fixed, paraffin-embed-
ded tissue specimens from 18 NSGCTs and 18 SGCTs was
used to construct a tissue array (Beecher Instruments, Sil-
ver Spring MD). Representative areas of the biopsy were
chosen to construct a 14 × 8 tissue array. Four micron-
thick sections on the array were immuno-stained
following deparaffinization and antigen retrieval using
citrate buffer at pH6.0. The primary antibody against
MGMT was obtained from NeoMarkers (Fremont, CA).
The antibodies were detected with the Envision plus
(DAKO, Carpenteria, CA) system, using diaminobenzi-
dine as a chromogen. Tumors were considered positive for
MGMT when cells showed brown nuclear staining. Inter-
stitial and intravascular lymphocytes, as well as spermato-
gonia of any residual seminiferous tubules were used as
internal controls.
Statistical analyses
Comparisons for the analyses of sensitive vs. resistant
tumors were done via Fisher's exact test. For the eight cases
that contributed multiple specimens, the specimen with
the greatest number of methylated genes was used. Exact
logistic regression was used for the phase comparisons
with the analyses adjusted for sensitive/resistant status
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 11 of 12
(page number not for citation purposes)
and histology where noted. Here multiple specimens
from the same patient were included. None of the p-val-
ues were adjusted for multiple comparisons due to the
exploratory nature of the analysis.
Authors' contributions
SK carried out the methylation, cloning, sequencing and
gene expression analysis. JMM participated in selection of
tumor specimens, isolation of DNA and RNA. GN partic-
ipated in the analysis of gene expression. JH coordinated
the selection of tumors, tissue culture, isolation of
genomic DNA and RNA. JB performed statistical analysis.
DLB participated in obtaining the follow up on patients.
AMA and MM participated in the preparation of tissue
array and gene expression analysis. VER participated in
histologic diagnosis. GJB was responsible for referring the
patients and clinical information. RSKC and VVVSM have
conceived and coordinated the study. All authors read and
approved the final manuscript.
Acknowledgements
This work was supported by the NIH grant CA75925 (VVVSM), funds from
Lance Armstrong Foundation (VVVSM and JH), Tom Green's Nuts Cancer
Fund of California Community Foundation (VVVSM), and the Herbert Irv-
ing Comprehensive Cancer Center, Columbia University (JMM and
VVVSM).
References
1. Chaganti RS, Houldsworth J: Genetics and biology of adult
human male germ cell tumors. Cancer Res 2000, 60:1475-1482.
2. Motzer RJ, Amsterdam A, Prieto V, Sheinfeld J, Murty VV, Mazumdar
M, Bosl GJ, Chaganti RS, Reuter VE: Teratoma with malignant
transformation: diverse malignant histologies arising in men
with germ cell tumors. J Urol 1998, 159:133-138.
3. Bosl GJ, Motzer RJ: Testicular germ-cell cancer. N Engl J Med
1997, 337:242-253.
4. Masters JR, Koberle B: Curing metastatic cancer: lessons from
testicular germ-cell tumours. Nat Rev Cancer 2003, 3:517-525.
5. Houldsworth J, Xiao H, Murty VV, Chen W, Ray B, Reuter VE, Bosl
GJ, Chaganti RS: Human male germ cell tumor resistance to
cisplatin is linked to TP53 gene mutation. Oncogene 1998,
16:2345-2349.
6. Rao PH, Houldsworth J, Palanisamy N, Murty VV, Reuter VE, Motzer
RJ, Bosl GJ, Chaganti RS: Chromosomal amplification is associ-
ated with cisplatin resistance of human male germ cell
tumors. Cancer Res 1998, 58:4260-4263.
7. Mayer F, Gillis AJ, Dinjens W, Oosterhuis JW, Bokemeyer C, Looi-
jenga LH: Microsatellite instability of germ cell tumors is asso-
ciated with resistance to systemic treatment. Cancer Res 2002,
62:2758-2760.
8. Koul S, Houldsworth J, Mansukhani MM, Donadio A, McKiernan JM,
Reuter VE, Bosl GJ, Chaganti RS, Murty VV: Characteristic pro-
moter hypermethylation signatures in male germ cell
tumors. Mol Cancer 2002, 1:8.
9. Smiraglia DJ, Szymanska J, Kraggerud SM, Lothe RA, Peltomaki P, Plass
C: Distinct epigenetic phenotypes in seminomatous and non-
seminomatous testicular germ cell tumors. Oncogene 2002,
21:3909-3916.
10. Honorio S, Agathanggelou A, Wernert N, Rothe M, Maher ER, Latif
F: Frequent epigenetic inactivation of the RASSF1A tumour
suppressor gene in testicular tumours and distinct methyla-
tion profiles of seminoma and nonseminoma testicular germ
cell tumours. Oncogene 2003, 22:461-466.
11. Pegg AE, Dolan ME, Moschel RC: Structure, function, and inhibi-
tion of O6-alkylguanine-DNA alkyltransferase. Prog Nucleic Acid
Res Mol Biol 1995, 51:167-223.
12. Esteller M, Garcia-Foncillas J, Andion E, Goodman SN, Hidalgo OF,
Vanaclocha V, Baylin SB, Herman JG: Inactivation of the DNA-
repair gene MGMT and the clinical response of gliomas to
alkylating agents. N Engl J Med 2000, 343:1350-1354.
13. Olopade OI, Wei M: FANCF methylation contributes to chem-
oselectivity in ovarian cancer. Cancer Cell 2003, 3:417-420.
14. Taniguchi T, Tischkowitz M, Ameziane N, Hodgson SV, Mathew CG,
Joenje H, Mok SC, D'Andrea AD: Disruption of the Fanconi ane-
mia-BRCA pathway in cisplatin-sensitive ovarian tumors. Nat
Med 2003, 9:568-574.
15. Bedford P, Fichtinger-Schepman AM, Shellard SA, Walker MC, Mas-
ters JR, Hill BT: Differential repair of platinum-DNA adducts in
human bladder and testicular tumor continuous cell lines.
Cancer Res 1988, 48:3019-3024.
16. Koberle B, Grimaldi KA, Sunters A, Hartley JA, Kelland LR, Masters
JR: DNA repair capacity and cisplatin sensitivity of human
testis tumour cells. Int J Cancer 1997, 70:551-555.
17. Lowe SW, Ruley HE, Jacks T, Housman DE: p53-dependent apop-
tosis modulates the cytotoxicity of anticancer agents. Cell
1993, 74:957-967.
18. Kersemaekers AM, Mayer F, Molier M, van Weeren PC, Oosterhuis
JW, Bokemeyer C, Looijenga LH: Role of P53 and MDM2 in treat-
ment response of human germ cell tumors. J Clin Oncol 2002,
20:1551-1561.
19. Zamble DB, Jacks T, Lippard SJ: p53-Dependent and -independ-
ent responses to cisplatin in mouse testicular teratocarci-
noma cells. Proc Natl Acad Sci U S A 1998, 95:6163-6168.
20. Chresta CM, Masters JR, Hickman JA: Hypersensitivity of human
testicular tumors to etoposide-induced apoptosis is associ-
ated with functional p53 and a high Bax:Bcl-2 ratio. Cancer Res
1996, 56:1834-1841.
21. Auersperg N, Edelson MI, Mok SC, Johnson SW, Hamilton TC: The
biology of ovarian cancer. Semin Oncol 1998, 25:281-304.
22. Smith-Sorensen B, Lind GE, Skotheim RI, Fossa SD, Fodstad O, Sten-
wig AE, Jakobsen KS, Lothe RA: Frequent promoter hypermeth-
ylation of the O6-Methylguanine-DNA Methyltransferase
(MGMT) gene in testicular cancer. Oncogene 2002,
21:8878-8884.
23. Pfeifer GP, Yoon JH, Liu L, Tommasi S, Wilczynski SP, Dammann R:
Methylation of the RASSF1A gene in human cancers. Biol
Chem 2002, 383:907-914.
24. Chen WY, Zeng X, Carter MG, Morrell CN, Chiu Yen RW, Esteller
M, Watkins DN, Herman JG, Mankowski JL, Baylin SB: Hetero-
zygous disruption of Hic1 predisposes mice to a gender-
dependent spectrum of malignant tumors. Nat Genet 2003,
33:197-202.
25. Esteller M, Corn PG, Baylin SB, Herman JG: A gene hypermethyl-
ation profile of human cancer. Cancer Res 2001, 61:3225-3229.
26. Esteller M, Gaidano G, Goodman SN, Zagonel V, Capello D, Botto B,
Rossi D, Gloghini A, Vitolo U, Carbone A, Baylin SB, Herman JG:
Hypermethylation of the DNA repair gene O(6)-methylgua-
nine DNA methyltransferase and survival of patients with
diffuse large B-cell lymphoma. J Natl Cancer Inst 2002, 94:26-32.
27. Christmann M, Pick M, Lage H, Schadendorf D, Kaina B: Acquired
resistance of melanoma cells to the antineoplastic agent
fotemustine is caused by reactivation of the DNA repair
gene MGMT. Int J Cancer 2001, 92:123-129.
28. Sun SY, Lotan R: Retinoids and their receptors in cancer devel-
opment and chemoprevention. Crit Rev Oncol Hematol 2002,
41:41-55.
29. Nyce JW: Drug-induced DNA hypermethylation: a potential
mediator of acquired drug resistance during cancer
chemotherapy. Mutat Res 1997, 386:153-161.
30. Nyce J: Drug-induced DNA hypermethylation and drug
resistance in human tumors. Cancer Res 1989, 49:5829-5836.
31. Gambacorti-Passerini CB, Gunby RH, Piazza R, Galietta A, Rostagno
R, Scapozza L: Molecular mechanisms of resistance to imatinib
in Philadelphia-chromosome-positive leukaemias. Lancet
Oncol 2003, 4:75-85.
32. Zamble DB, Lippard SJ: Cisplatin and DNA repair in cancer
chemotherapy. Trends Biochem Sci 1995, 20:435-439.
33. Huang JC, Zamble DB, Reardon JT, Lippard SJ, Sancar A: HMG-
domain proteins specifically inhibit the repair of the major
DNA adduct of the anticancer drug cisplatin by human exci-
sion nuclease. Proc Natl Acad Sci U S A 1994, 91:10394-10398.
Publish with Bio Med Central and every
scientist can read your work free of charge
"BioMed Central will be the most significant development for
disseminating the results of biomedical researc h in our lifetime."
Sir Paul Nurse, Cancer Research UK
Your research papers will be:
available free of charge to the entire biomedical community
peer reviewed and published immediately upon acceptance
cited in PubMed and archived on PubMed Central
yours — you keep the copyright
Submit your manuscript here:
http://www.biomedcentral.com/info/publishing_adv.asp
BioMedcentral
Molecular Cancer 2004, 3 http://www.molecular-cancer.com/content/3/1/16
Page 12 of 12
(page number not for citation purposes)
34. Zamble DB, Mikata Y, Eng CH, Sandman KE, Lippard SJ: Testis-spe-
cific HMG-domain protein alters the responses of cells to
cisplatin. J Inorg Biochem 2002, 91:451-462.
35. Koberle B, Masters JR, Hartley JA, Wood RD: Defective repair of
cisplatin-induced DNA damage caused by reduced XPA pro-
tein in testicular germ cell tumours. Curr Biol 1999, 9:273-276.
36. Bala S, Oliver H, Renault B, Montgomery K, Dutta S, Rao P, Houlds-
worth J, Kucherlapati R, Wang X, Chaganti RS, Murty VV: Genetic
analysis of the APAF1 gene in male germ cell tumors. Genes
Chromosomes Cancer 2000, 28:258-268.
... The current body of work is limited to few differentially methylated genes. RASSF1A hypermethylation has been proposed as a risk factor in a couple of studies, followed by a meta-analysis with results confirmed in a cohort of 32 patients [27][28][29]. RASSF1A promoter hypermethylation was identifiable in the peripheral blood of TGCT patients, both in the seminoma and non-seminoma groups. Although stratification by tumor subtype was not possible because of the limited number of patients, the meta-analysis provided additional evidence. ...
Article
Full-text available
Testicular germ cell tumors (TGCTs) are the most common type of testicular cancer, with a particularly high incidence in the 15–45-year age category. Although highly treatable, resistance to therapy sometimes occurs, with devastating consequences for the patients. Additionally, the young age at diagnosis and the treatment itself pose a great threat to patients’ fertility. Despite extensive research concerning genetic and environmental risk factors, little is known about TGCT etiology. However, epigenetics has recently come into the spotlight as a major factor in TGCT initiation, progression, and even resistance to treatment. As such, recent studies have been focusing on epigenetic mechanisms, which have revealed their potential in the development of novel, non-invasive biomarkers. As the most studied epigenetic mechanism, DNA methylation was the first revelation in this particular field, and it continues to be a main target of investigations as research into its association with TGCT has contributed to a better understanding of this type of cancer and constantly reveals novel aspects that can be exploited through clinical applications. In addition to biomarker development, DNA methylation holds potential for developing novel treatments based on DNA methyltransferase inhibitors (DNMTis) and may even be of interest for fertility management in cancer survivors. This manuscript is structured as a literature review, which comprehensively explores the pivotal role of DNA methylation in the pathogenesis, progression, and treatment resistance of TGCTs.
... Therefore, the epigenetic status might also influence cisplatin sensitivity and clinical outcomes for TGCTs. This is supported by investigations of Koul and co-workers, who observed different promoter hypermethylation of individual genes in cisplatin-sensitive versus resistant germ cell tumors [90]. ...
Article
Full-text available
The clinical utility of the chemotherapeutic agent cisplatin is restricted by cancer drug resistance, which is either intrinsic to the tumor or acquired during therapy. Epigenetics is increasingly recognized as a factor contributing to cisplatin resistance and hence influences drug efficacy and clinical outcomes. In particular, epigenetics regulates gene expression without changing the DNA sequence. Common types of epigenetic modifications linked to chemoresistance are DNA methylation, histone modification, and non-coding RNAs. This review provides an overview of the current findings of various epigenetic modifications related to cisplatin efficacy in cell lines in vitro and in clinical tumor samples. Furthermore, it discusses whether epigenetic alterations might be used as predictors of the platinum agent response in order to prevent avoidable side effects in patients with resistant malignancies. In addition, epigenetic targeting therapies are described as a possible strategy to render cancer cells more susceptible to platinum drugs.
... Moreover, we also explore the relationship between HIC1 expression and the anti-cancer drug sensitivity of cancer patients. Koul S et al. have reported that the promoter hypermethylation of HIC1 was involved in the resistance of Male germ cell tumor (GCT) to cisplatin (49). Our findings showed that HIC1 is closely related to the sensitivity of multiple anti-cancer drugs, especially small molecule inhibitors, including MEK inhibitors trametinib and PD−98059, indicating HIC1 plays a critical role in predicting the sensitivity of anti-cancer drugs. ...
Article
Full-text available
Background Hypermethylated in Cancer 1 (HIC1) was originally confirmed as a tumor suppressor and has been found to be hypermethylated in human cancers. Although growing evidence has supported the critical roles of HIC1 in cancer initiation and development, its roles in tumor immune microenvironment and immunotherapy are still unclear, and no comprehensive pan-cancer analysis of HIC1 has been conducted. Methods HIC1 expression in pan-cancer, and differential HIC1 expression between tumor and normal samples were investigated. Immunohistochemistry (IHC) was employed to validate HIC1 expression in different cancers by our clinical cohorts, including lung cancer, sarcoma (SARC), breast cancer, and kidney renal clear cell carcinoma (KIRC). The prognostic value of HIC1 was illustrated by Kaplan-Meier curves and univariate Cox analysis, followed by the genetic alteration analysis of HIC1 in pan-cancer. Gene Set Enrichment Analysis (GSEA) was conducted to illustrate the signaling pathways and biological functions of HIC1. The correlations between HIC1 and tumor mutation burden (TMB), microsatellite instability (MSI), and the immunotherapy efficacy of PD-1/PD-L1 inhibitors were analyzed by Spearman correlation analysis. Drug sensitivity analysis of HIC1 was performed by extracting data from the CellMiner™ database. Results HIC1 expression was abnormally expressed in most cancers, and remarkable associations between HIC1 expression and prognostic outcomes of patients in pan-cancer were detected. HIC1 was significantly correlated with T cells, macrophages, and mast cell infiltration in different cancers. Moreover, GSEA revealed that HIC1 was significantly involved in immune-related biological functions and signaling pathways. There was a close relationship of HIC1 with TMB and MSI in different cancers. Furthermore, the most exciting finding was that HIC1 expression was significantly correlated with the response to PD-1/PD-L1 inhibitors in cancer treatment. We also found that HIC1 was significantly correlated with the sensitivity of several anti-cancer drugs, such as axitinib, batracylin, and nelarabine. Finally, our clinical cohorts further validated the expression pattern of HIC1 in cancers. Conclusions Our investigation provided an integrative understanding of the clinicopathological significance and functional roles of HIC1 in pan-cancer. Our findings suggested that HIC1 can function as a potential biomarker for predicting the prognosis, immunotherapy efficacy, and drug sensitivity with immunological activity in cancers.
... Hypermethylation of specific gene promoters, such as RASSF1A, HIC1, CALCA, MLH1, and MGMT, has been associated with cisplatin resistance [98,99] (Figure 1BI), whereas RARB promoter hypermethylation is associated with tumors sensitive to cisplatin (Figure 1(AI)) ( Table 2). None of these genes has been validated as a DNA methylation-based biomarker for cisplatin response. ...
Article
Full-text available
Testicular germ-cell tumors (TGCT) have been widely recognized for their outstanding survival rates, commonly attributed to their high sensitivity to cisplatin-based therapies. Despite this, a subset of patients develops cisplatin resistance, for whom additional therapeutic options are unsuccessful, and ~20% of them will die from disease progression at an early age. Several efforts have been made trying to find the molecular bases of cisplatin resistance. However, this phenomenon is still not fully understood, which has limited the development of efficient biomarkers and precision medicine approaches as an alternative that could improve the clinical outcomes of these patients. With the aim of providing an integrative landscape, we review the most recent genomic and epigenomic features attributed to chemoresponse in TGCT patients, highlighting how we can seek to combat cisplatin resistance through the same mechanisms by which TGCTs are particularly hypersensitive to therapy. In this regard, we explore ongoing treatment directions for resistant TGCT and novel targets to guide future clinical trials. Through our exploration of recent findings, we conclude that epidrugs are promising treatments that could help to restore cisplatin sensitivity in resistant tumors, shedding light on potential avenues for better prognosis for the benefit of the patients.
... BRCA1 (19.8%), and HIC1 (19.6%) (Koul et al., 2002). The hypermethylation of the RASSF1A and HIC1 was associated with cisplatin resistance (Koul et al., 2004). More recently, MGMT and CALCA promotor methylations were observed in NSGCTs and assigned as a poor prognosis. ...
Article
Testicular germ cell tumors (TGCTs) are a common malignancy occurring in young adult men. The various genetic risk factors have been suggested to contribute to TGCT pathogenesis, however, they have a distinct mutational profile with a low rate of somatic point mutations, more frequent chromosomal gains, and aneuploidy. The most frequently mutated oncogenes in human cancers are RAS oncogenes, while their impact on testicular carcinogenesis and refractory disease is still poorly understood. In this mini-review, we summarize current knowledge on genetic alternations of RAS signaling-associated genes (the single nucleotide polymorphisms and point mutations) in this particular cancer type and highlight their link to chemotherapy resistance mechanisms. We also mention the impact of epigenetic changes on TGCT progression. Lastly, we propose a model for RAS-dependent signaling networks, regulation, cross-talks, and outcomes in TGCTs.
... Another genomic trait of seminomas versus non-seminomas is the low level of DNA methylation. 112,113 If we trace the methylation status of TGCT precursor cells (PGCs), we will find that at the earliest stage when an ESC transforms into a gonocyte (early PGC), it loses its DNA methylation pattern. In the transition from early to late PGC and during further specialization of PGC into spermatogonia, cells will start de novo DNA methylation to re-establish the parental imprinting pattern. ...
Article
Full-text available
Objective: Testicular germ cell tumors (TGCT) are the most common solid malignancy in adolescent and young men, with a rising incidence over the past 20 years. Overall, TGCTs are second in terms of the average life years lost per person dying of cancer, and clinical therapeutics without adverse long-term side effects are lacking. Platinum-based regimens for TGCTs have heterogeneous outcomes even within the same histotype that frequently leads to under- and over-treatment. Understanding of molecular differences that lead to diverse outcomes of TGCT patients may improve current treatment approaches. Seminoma is the most common subtype of TGCTs, which can either be pure or present in combination with other histotypes. Methods: Here we conducted a computational study of 64 pure seminoma samples from The Cancer Genome Atlas, applied consensus clustering approach to their transcriptomic data and revealed two clinically relevant seminoma subtypes: seminoma subtype 1 and 2. Results: Our analysis identified significant differences in pluripotency stage, activity of double stranded DNA breaks repair mechanisms, rates of loss of heterozygosity, and expression of lncRNA responsible for cisplatin resistance between the subtypes. Seminoma subtype 1 is characterized by higher pluripotency state, while subtype 2 showed attributes of reprogramming into non-seminomatous TGCT. The seminoma subtypes we identified may provide a molecular underpinning for variable responses to chemotherapy and radiation. Conclusion: Translating our findings into clinical care may help improve risk stratification of seminoma, decrease overtreatment rates, and increase long-term quality of life for TGCT survivors.
... The mismatch repair pathway (MMR) is a vital tool to keep genome stability, and a deficiency in this system has been shown to cause CDDP resistance in cells, associated with poor prognosis in some tumors [26,133]. Several previous studies have shown that promoter hypermethylation of genes involved in this pathway, such as MLH1 and hMSH2, is associated with the acquisition of resistance to platinum therapy [72,92,105,110,121,124]. The NRF2/KEAP1 pathway plays a key role in the chemoresistance process of different tumor types and is capable of inhibiting apoptosis, promoting cell proliferation, and chemoresistance [134]. ...
Article
Full-text available
Simple Summary Platinum-based agents are one of the most widely used chemotherapy drugs for various types of cancer. However, one of the main challenges in the application of platinum drugs is resistance, which is currently being widely investigated. Epigenetic DNA methylation-based biomarkers are promising to aid in the selection of patients, helping to foresee their platinum therapy response in advance. These biomarkers enable minimally invasive patient sample collection, short analysis, and good sensitivity. Hence, improved methodologies for the detection and quantification of DNA methylation biomarkers will facilitate their use in the choice of an optimal treatment strategy. Abstract Platinum-based chemotherapy is routinely used for the treatment of several cancers. Despite all the advances made in cancer research regarding this therapy and its mechanisms of action, tumor resistance remains a major concern, limiting its effectiveness. DNA methylation-based biomarkers may assist in the selection of patients that may benefit (or not) from this type of treatment and provide new targets to circumvent platinum chemoresistance, namely, through demethylating agents. We performed a systematic search of studies on biomarkers that might be predictive of platinum-based chemotherapy resistance, including in vitro and in vivo pre-clinical models and clinical studies using patient samples. DNA methylation biomarkers predictive of response to platinum remain mostly unexplored but seem promising in assisting clinicians in the generation of more personalized follow-up and treatment strategies. Improved methodologies for their detection and quantification, including non-invasively in liquid biopsies, are additional attractive features that can bring these biomarkers into clinical practice, fostering precision medicine.
... Different methylation profiles of several specific gene promoters have been reported in cisplatin-sensitive and -resistant GCTs. Promoter hypermethylation of the RASSF1A and HIC1 genes was observed in resistant non-seminomas, while sensitive non-seminomas showed hypermethylation of MGMT and RARB [108]. Martinelli et al. showed the association of CALCA hypermethylation with non-seminomas (90.5%, 19/21; p < 0.026) and cisplatin-refractory disease (47.4%, 09/19; p = 0.005). ...
Article
Full-text available
Testicular germ cell tumors (GCTs) are highly curable malignancies. Excellent survival rates in patients with metastatic disease can be attributed to the exceptional sensitivity of GCTs to cisplatin-based chemotherapy. This hypersensitivity is probably related to alterations in the DNA repair of cisplatin-induced DNA damage, and an excessive apoptotic response. However, chemotherapy fails due to the development of cisplatin resistance in a proportion of patients. The molecular basis of this resistance appears to be multifactorial. Tracking the mechanisms of cisplatin resistance in GCTs, multiple molecules have been identified as potential therapeutic targets. A variety of therapeutic agents have been evaluated in preclinical and clinical studies. These include different chemotherapeutics, targeted therapies, such as tyrosine kinase inhibitors, mTOR inhibitors, PARP inhibitors, CDK inhibitors, and anti-CD30 therapy, as well as immune-checkpoint inhibitors, epigenetic therapy, and others. These therapeutics have been used as single agents or in combination with cisplatin. Some of them have shown promising in vitro activity in overcoming cisplatin resistance, but have not been effective in clinical trials in refractory GCT patients. This review provides a summary of current knowledge about the molecular mechanisms of cisplatin sensitivity and resistance in GCTs and outlines possible therapeutic approaches that seek to overcome this chemoresistance.
... The promoter methylation status of several candidate genes has been assessed in various individual studies on TGCTs (reviewed in [103]), both in patients' tissue and plasma cohorts and cell lines [104][105][106][107][108][109][110][111][112][113][114][115][116][117][118][119]. Some focused on genes involved in stages of embryonic development (such as CRIPTO [105] and OCT3/4 [116]), others on genes having a tumor-suppressor role in a variety of neoplasms (such as RASFF1A [107,120]), others coding for DNA repair proteins (such as MGMT [109]) and others explored cancer-testis antigens (CTAs, such as PRAME [119]). ...
Article
The 5th edition of the WHO Classification of Tumours of the Urinary and Male Genital Systems contains relevant revisions and introduces a group of molecularly defined renal tumour subtypes. Herein we present the World Health Organization (WHO) 2022 perspectives on papillary and chromophobe renal cell carcinoma with emphasis on their evolving classification, differential diagnosis, and emerging entities. The WHO 2022 classification eliminated the type 1/2 papillary renal cell carcinoma (pRCC) subcategorization, given the recognition of frequent mixed tumour phenotypes and the existence of entities with different molecular background within the type 2 pRCC category. Additionally, emerging entities such as biphasic squamoid alveolar RCC, biphasic hyalinizing psammomatous RCC, papillary renal neoplasm with reverse polarity, and Warthin‐like pRCC are included as part of the pRCC spectrum, while additional morphological and molecular data are being gathered. In addition to oncocytomas and chromophobe renal cell carcinoma (chRCC), a category of “other oncocytic tumours” with oncocytoma/chRCC‐like features has been introduced, including emerging entities, most with TSC/mTOR pathway alterations (eosinophilic vacuolated tumour and so‐called “low‐grade” oncocytic tumour), deserving additional research. Eosinophilic solid and cystic RCC was accepted as a new and independent tumour entity. Finally, a highly reproducible and clinically relevant universal grading system for chRCC is still missing and is another niche of ongoing investigation. This review discusses these developments and highlights emerging morphological and molecular data relevant for the classification of renal cell carcinoma.
Article
Full-text available
We are in an era where the potential exists for deriving comprehensive profiles of DNA alterations characterizing each form of human cancer. Such profiles would provide invaluable insight into mechanisms underly- ing the evolution of each tumor type and will provide molecular markers, which could radically improve cancer detection. To date, no one type of DNA change has been defined which accomplishes this purpose. Herein, by using a candidate gene approach, we show that one category of DNA alteration, aberrant methylation of gene promoter regions, can enor- mously contribute to the above goals. We have now analyzed a series of promoter hypermethylation changes in 12 genes (p16INK4a, p15INK4b, p14ARF, p73, APC,5 BRCA1, hMLH1, GSTP1, MGMT, CDH1, TIMP3, and DAPK), each rigorously characterized for association with abnormal gene silencing in cancer, in DNA from over 600 primary tumor samples rep- resenting 15 major tumor types. The genes play known important roles in processes encompassing tumor suppression, cell cycle regulation, apopto- sis, DNA repair, and metastastic potential. A unique profile of promoter hypermethylation exists for each human cancer in which some gene changes are shared and others are cancer-type specific. The hypermethy- lation of the genes occurs independently to the extent that a panel of three to four markers defines an abnormality in 70 -90% of each cancer type. Our results provide an unusual view of the pervasiveness of DNA alter- ations, in this case an epigenetic change, in human cancer and a powerful set of markers to outline the disruption of critical pathways in tumori- genesis and for derivation of sensitive molecular detection strategies for virtually every human tumor type.
Article
Full-text available
The formation and removal of four platinum-DNA adducts were immunochemically quantitated in cultured cells derived from a human bladder carcinoma cell line (RT112) and from two lines derived from germ cell tumors of the testis (833K and SUSA), following exposure in vitro to 16.7 microM (5 micrograms/ml) cisplatin. RT112 cells were least sensitive to the drug and were proficient in the repair of all four adducts, whereas SUSA cells, which were 5-fold more sensitive, were deficient in the repair of DNA-DNA intrastrand cross-links in the sequences pApG and pGpG. Despite expressing a similar sensitivity to SUSA cells, 833K cells were proficient in the repair of all four adducts, although less so than the RT112 bladder tumor cells. In addition, SUSA cells were unable to repair DNA-DNA interstrand cross-links whereas 50-85% of these lesions were removed in RT112 and 833K cells 24 h following drug exposure. It is possible that the inability of SuSa cells to repair platinated DNA may account for their hypersensitivity to cisplatin.
Article
The DNA-repair enzyme O6-methylguanine-DNA methyltransferase (MGMT) inhibits the killing of tumor cells by alkylating agents. MGMT activity is controlled by a promoter; methylation of the promoter silences the gene in cancer, and the cells no longer produce MGMT. We examined gliomas to determine whether methylation of the MGMT promoter is related to the responsiveness of the tumor to alkylating agents. METHODS: We analyzed the MGMT promoter in tumor DNA by a methylation-specific polymerase-chain-reaction assay. The gliomas were obtained from patients who had been treated with carmustine (1,3-bis(2-chloroethyl)-1-nitrosourea, or BCNU). The molecular data were correlated with the clinical outcome. RESULTS: The MGMT promoter was methylated in gliomas from 19 of 47 patients (40 percent). This finding was associated with regression of the tumor and prolonged overall and disease-free survival. It was an independent and stronger prognostic factor than age, stage, tumor grade, or performance status. CONCLUSIONS: Methylation of the MGMT promoter in gliomas is a useful predictor of the responsiveness of the tumors to alkylating agents.
Article
The most frequent DNA adduct made by the anticancer drug cisplatin, the 1,2-intrastrand d(GpG) cross-link, as well as the minor 1,3-intrastrand d(GpTpG) adduct, were both repaired by an in vitro human excision repair system. Fragments of 27-29 nt containing the platinum damage were excised. The high mobility group (HMG)-domain proteins HMG1 and human mitochondrial transcription factor specifically inhibited repair of the 1,2-intrastrand cross-link by the human excision nuclease. These results suggest that the types and levels of HMG-domain proteins in a given tumor may influence the responsiveness of that cancer to cisplatin chemotherapy and they provide a rational basis for the synthesis of new platinum anticancer drug candidates.
Article
Metastatic testicular germ cell tumours can be cured using cisplatin-based combination chemotherapy. To investigate the role of DNA repair in cisplatin sensitivity, we measured the formation and removal of cisplatin adducts in the whole genome and in specific genomic regions in 3 testis and 3 bladder cancer cell lines. Following a 1 hr exposure to 50 μM cisplatin, the mean level of DNA platination was lower in the testis tumour cell lines. During a 72 hr post-treatment incubation period, the 3 bladder cell lines removed platinum from the overall genome, whereas 2 of the testis tumour cell lines showed relatively little reduction of DNA platination. The third testis tumour cell line, SuSa, showed an intermediate capacity to remove cisplatin. Cisplatin-induced damage and repair in selected regions of the actively transcribed N-ras gene and the inactive CD3δ gene were measured using quantitative PCR. The data were in agreement with those obtained with atomic absorption spectroscopy for the whole genome, showing that the bladder lines were repair-proficient: 2 of the testis tumour cell lines showed no repair, and the third testis line, SuSa, showed an intermediate level of repair in these 2 genes. Our findings confirm that reduced capacity to repair cisplatin-damaged DNA may contribute to the hypersensitivity of testis tumour cells to DNA-damaging agents. Int. J. Cancer 70:551–555. © 1997 Wiley-Liss Inc.
Article
Acquired resistance to antineoplastic agents is a frequent obstacle in tumor therapy. Malignant melanoma cells are particularly well known for their unresponsiveness to chemotherapy; only about 30% of tumors exhibit a transient clinical response to treatment. In our study, we investigated the molecular mechanism of acquired resistance of melanoma cells (MeWo) to the chloroethylating drug fotemustine. Determination of O6-methylguanine-DNA methyltransferase (MGMT) activity showed that MeWo cells that acquired resistance to fotemustine upon repeated treatment with the drug display high MGMT activity, whereas the parental cell line had no detectable MGMT. The resistant cell lines exhibit cross-resistance to other O6-alkylating agents, such as N-methyl-N′-nitro-N-nitrosoguanidine. Acquired resistance to fotemustine was alleviated by treatment with the MGMT inhibitor O6-benzylguanine demonstrating that reactivation of MGMT is the main underlying cause of acquired alkylating drug resistance. As compared with control cells, both MGMT mRNA and MGMT protein were expressed at a high level in fotemustine resistant cells. Southern blot analysis proved that the MGMT gene was not amplified. There was also only an insignificant difference in the CpG methylation pattern of the MGMT promoter whereas a clear hypermethylation in the body of the gene was observed in fotemustine resistant cells. The conclusion that hypermethylation is responsible for reactivation of the MGMT gene gained support by the finding that MGMT activity significantly declined and cells reverted (partially) to the parental sensitive phenotype upon treatment with 5-azacytidine. This is the first report of acquired resistance to a chloroethylating antineoplastic drug of melanoma cells due to gene hypermethylation. © 2001 Wiley-Liss, Inc.
Article
Thesis (Ph. D.)--Massachusetts Institute of Technology, Dept. of Biology, 1994. Includes bibliographical references (leaves 226-254).
Article
Drug-induced DNA hypermethylation was observed to constitute one component of the response of human tumor cells to toxic concentrations of commonly used cancer chemotherapy agents. In both human lung adenocarcinoma cells (HTB-54) and human rhabdomyosarcoma cells (CCl-136), pulse exposures to the topoisomerase II inhibitors etoposide and nalidixic acid; to the antibiotic doxorubicin; to the microtubule inhibitors vincristine, vinblastine, and colchicine; to the DNA cross-linking agent cisplatinum; to hydroxyurea; and to the antimetabolites 1-beta-D-arabinofuranosylcytosine, 5-fluorouracil, 5-fluorodeoxyuridine, and methotrexate were associated with profound drug-induced DNA hypermethylation. Exposure of human T-lymphocytes (MOLT-4) to toxic pulse doses of 3'-azidodideoxythymidine was associated with similar drug-induced DNA hypermethylation. In every case, drug-induced DNA hypermethylation was observed only when the degree of DNA synthesis inhibition caused by the drug exceeded 90% and when drug levels or duration of exposure was sufficient to kill 90-100% of exposed cells. Drug-induced DNA hypermethylation was shown not to represent a tissue culture phenomenon, since it occurred in vivo during high-dose 1-beta-D-arabinofuranosylcytosine and hydroxyurea treatments in two leukemic patients. Drug-induced alterations in DNA methylation were frequently biphasic, with hypomethylation occurring at drug concentrations which produced mild DNA synthesis inhibition and which killed less than 50% of exposed cells. Exposure to the alkylating agents 1,3-bis(2-chloroethyl)-1-nitrosourea and cyclophosphamide and to the antimetabolites 5-azadeoxycytidine and 6-thioguanine was associated with DNA hypomethylation at all studied concentrations in HTB-54 cells. Drug-induced DNA hypermethylation could be blocked by preexposure to hypomethylating agents administered at nontoxic to mildly toxic concentrations. Drug-induced DNA hypermethylation may be capable of creating drug-resistant phenotypes by inactivating genes the products of which are required for drug cytotoxicity. Perhaps paradoxically, drug-induced DNA hypermethylation may also produce a second class of drug-resistant tumor cells, characterized by overexpression of particular gene products, by potentiating the process of gene amplification.
Article
This chapter focuses on the structure and function of alkyltransferase protein. O⁶-alkylguanine-DNA alkyltransferase is a remarkable protein that alone can, in a single step, remove adducts from DNA that are formed at the O⁶-position of guanine and the O⁴-position of thymine and can, thus, restore the original DNA. Production of such adducts is a major contributor to the toxic, mutagenic, and carcinogenic effects of alkylating agents. Alkyltransferase activity has been detected in many species, including microorganisms, insects, fish, and mammals. The unique activity of alkyltransferase suggests that it is an ideal target for biochemical modulation. The efficient repair of toxic lesions formed at the O⁶-position of guanine without additional enzymes or cofactors provides a less complex target to modulate than other DNA repair proteins. Furthermore, the high degree of correlation that exists between alkyltransferase activity and sensitivity to nitrosoureas indicates that elimination of this protein may reverse resistance in many cases. Two methods have been used to overcome alkylnitrosourea resistance by inactivation of alkyltransferase. One uses methylating agents that indirectly decrease alkyltransferase levels by introducing O⁶-methylguanine residues in DNA that are then repaired by the alkyltransferase. The second method uses direct alkyltransferase inactivators such as O⁶-methylguanine.
Article
The most frequent DNA adduct made by the anticancer drug cisplatin, the 1,2-intrastrand d(GpG) cross-link, as well as the minor 1,3-intrastrand d(GpTpG) adduct, were both repaired by an in vitro human excision repair system. Fragments of 27-29 nt containing the platinum damage were excised. The high mobility group (HMG)-domain proteins HMG1 and human mitochondrial transcription factor specifically inhibited repair of the 1,2-intrastrand cross-link by the human excision nuclease. These results suggest that the types and levels of HMG-domain proteins in a given tumor may influence the responsiveness of that cancer to cisplatin chemotherapy and they provide a rational basis for the synthesis of new platinum anticancer drug candidates.