ArticlePDF Available

Decreased DNA Repair Efficiency by Loss or Disruption of p53 Function Preferentially Affects Removal of Cyclobutane Pyrimidine Dimers from Non-transcribed Strand and Slow Repair Sites in Transcribed Strand

Authors:

Abstract

The tumor suppressor protein p53 plays a central role in modulating the cellular responses to DNA damage. Several recent studies, undertaken with the whole genomic DNA or full-length gene segments, have shown that p53 is involved in nucleotide excision repair and it selectively influences the adduct removal from the non-transcribed strand in the genome. In this study, we have analyzed the damage induction at nucleotide resolution by ligase-mediated polymerase chain reaction and compared the repair of ultraviolet radiation-induced cyclobutane pyrimidine dimers within exon 8 of p53 gene in normal and Li-Fraumeni syndrome fibroblasts as well as in normal and human papillomavirus 16 E6 and E7 protein-expressing human mammary epithelial cells. The results demonstrate that (i) loss or disruption of p53 function decreases efficiency of DNA repair, by preferentially affecting the repair of non-transcribed strand and of intrinsically slow repair sites in transcribed strand; (ii) mutant p53 protein affects DNA repair, at least of non-transcribed strand, in a dominant negative manner; and (iii) pRb does not have an effect on the repair of DNA damage within transcribed or non-transcribed strand. The overall data suggest that p53 could regulate excision repair or related events through direct protein-protein interaction.
Decreased DNA Repair Efficiency by Loss or Disruption of p53
Function Preferentially Affects Removal of Cyclobutane Pyrimidine
Dimers from Non-transcribed Strand and Slow Repair Sites in
Transcribed Strand*
(Received for publication, November 17, 1999, and in revised form, January 6, 2000)
Qianzheng Zhu‡, Manzoor A. Wani‡, Mohammed El-mahdy‡, and Altaf A. Wani‡§
From the Department of Radiology, §Biochemistry Program, and James Cancer Hospital and Research Institute, Ohio
State University, Columbus, Ohio 43210
The tumor suppressor protein p53 plays a central role
in modulating the cellular responses to DNA damage.
Several recent studies, undertaken with the whole
genomic DNA or full-length gene segments, have shown
that p53 is involved in nucleotide excision repair and it
selectively influences the adduct removal from the non-
transcribed strand in the genome. In this study, we have
analyzed the damage induction at nucleotide resolution
by ligase-mediated polymerase chain reaction and com-
pared the repair of ultraviolet radiation-induced cy-
clobutane pyrimidine dimers within exon 8 of p53 gene
in normal and Li-Fraumeni syndrome fibroblasts as well
as in normal and human papillomavirus 16 E6 and E7
protein-expressing human mammary epithelial cells.
The results demonstrate that (i) loss or disruption of p53
function decreases efficiency of DNA repair, by prefer-
entially affecting the repair of non-transcribed strand
and of intrinsically slow repair sites in transcribed
strand; (ii) mutant p53 protein affects DNA repair, at
least of non-transcribed strand, in a dominant negative
manner; and (iii) pRb does not have an effect on the
repair of DNA damage within transcribed or non-tran-
scribed strand. The overall data suggest that p53 could
regulate excision repair or related events through di-
rect protein-protein interaction.
Mammalian cells have evolved sophisticated DNA repair
mechanisms to overcome DNA damaging hazards that
threaten the integrity of genome (1–4). The most versatile and
thoroughly studied repair system is the excision repair, of
which two major pathways, nucleotide excision repair (NER)
1
and base excision repair, have been identified (5, 6). NER
removes many types of DNA lesions including cyclobutane
pyrimidine dimers (CPDs) by global genomic repair (GGR) and
transcription-coupled repair (TCR) (5–8). It is now well estab-
lished that NER along genome is heterogeneous, CPDs are
more efficiently removed from transcriptionally active genes,
and TCR is generally faster than GGR (2, 9, 10). GGR acts on
the elimination of lesions from non-transcribed strand (NTS)
and transcriptionally inactive genes, whereas TCR removes
lesions from DNA strand transcribed by RNA polymerase II.
In mammalian cells, a variety of cellular responses following
genotoxic exposure may contribute to prevent DNA lesions
from interfering with essential cellular functions. Of those
cellular responses, activation of p53 pathway is well studied
and documented (11, 12). p53 is a critical protein for maintain-
ing genomic stability and homeostasis. It is believed that p53
activation signals the G
1
arrest to delay the transit from G
1
to
S, thus preventing the effects of DNA lesions on vital cellular
functions. Accumulating evidence indicates that p53 plays a
role in DNA repair, especially in GGR (13–17). Viral proteins
that bind p53, causing p53 inactivation or degradation, inter-
fere with p53-regulated DNA repair (14, 18–20). Conceivably,
p53 could be involved in NER by regulating its downstream
genes, which are either related to or actively participate in
NER. For example, Gadd45 and p21
waf1
, which are among the
many p53-regulated proteins, have been shown to interact with
proliferating cell nuclear antigen (21). Additionally, recent ev-
idence has shown that p48, a UV-damaged DNA-binding pro-
tein is transcriptionally regulated by p53, linking p53 to NER
(22). Besides transcriptional activation, p53 may also be di-
rectly involved in repair or repair-related processes. For exam-
ple, in addition to demonstrated p53 binding to insertion/dele-
tion mismatches (23), p53 has also been found to both
physically and functionally interact with p62, XPD, and XPB,
three components of basal transcription factor IIH (10, 24).
Despite a plethora of experimental data, the definitive role of
p53 in NER and the mechanistic basis of its interaction with
repair machinery has not been fully delineated in eukaryotic
systems. Several observations, either supporting a pronounced
effect of p53 for an efficient repair of NTS alone or its involve-
ment also in the repair of transcribed strand (TS), have been
put forth to identify the role of p53 in NER (13–17, 19, 25, 26).
These observations are primarily based upon the assessment of
an average of the repair events within an entire genome, a
specific gene segment, or in some cases an episomally replicat-
ing plasmid within mammalian host cells. Thus, in the present
study we performed the damage analysis at nucleotide resolu-
tion and systematically compared DNA repair of individual
UV-induced CPD sites in normal, p53 mutant (p53-Mut), and
p53 nullizygous (p53-Null) Li-Fraumeni syndrome (LFS) fibro-
blasts, as well as in normal, human papillomavirus (HPV) 16
E6 and E7 protein-expressing human mammary epithelial cells
(HMEC). Our results show that, compared with the efficiency
* This work was supported by National Institutes of Health NIEHS
Grant ES2388. The costs of publication of this article were defrayed in
part by the payment of page charges. This article must therefore be
hereby marked advertisement in accordance with 18 U.S.C. Section
1734 solely to indicate this fact.
To whom correspondence should be addressed: Molecular Carcino-
genesis Laboratory, 103 Wiseman Hall, 400 W. 12th Ave., Columbus,
OH 43210. Tel.: 614-292-9375; Fax: 614-292-7237; E-mail: wani.2@
osu.edu.
1
The abbreviations used are: NER, nucleotide excision repair; CPD,
cyclobutane pyrimidine dimer; LFS, Li-Fraumeni syndrome; HMEC,
human mammary epithelial cell; LMPCR, ligation-mediated polymer-
ase chain reaction; WT, wild type; Mut, mutant; Null, nullizygous;
HPV, human papillomavirus; NTS, non-transcribed strand; TS, tran-
scribed strand; TCR, transcription-coupled repair; GGR, global genomic
repair.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 275, No. 15, Issue of April 14, pp. 11492–11497, 2000
© 2000 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org11492
of TS repair, a significantly reduced rate of CPD removal was
observable at all the dipyrimidine target sites of exon 8 NTS in
both p53-Null and p53-Mut LFS cell lines. A close comparison
of the extent of damage removal at 24 and 48 h after UV
irradiation indicated that the repair of NTS was more severely
affected in p53-Mut cells than in p53-Null cells. Relatively
slight decrease in the initial rate of repair within TS, discern-
ible at 24 h after treatment, was fully apparent at the inher-
ently slow repair sites, e.g. codons 286 and 294 of p53 gene.
This selective effect of p53 protein on the slow repair site of TS
was much more clearly demonstrable at the same site in HPV
16 E6-expressing HMEC in which p53 is functionally inacti-
vated by E6 protein. Like LFS cells, HPV 16 E6-expressing
HMEC show a reduced rate of repair of CPDs in the NTS than
in the TS. Both normal HMEC, having wild-type p53 and pRb
proteins, and the isogenic HPV 16 E7-expressing HMEC, hav-
ing wild-type p53 and compromised pRb protein expression,
failed to show the deficiency of CPD repair in NTS as seen
clearly due to the absence of p53 within human cells. Thus,
pRb, another important cell cycle regulatory protein, does not
seem to influence the NER process to a meaningful extent.
MATERIALS AND METHODS
Cell Culture and Treatment—The normal human (p53-WT) fibro-
blasts (OSU-2) were established in culture as described earlier (27).
LFS fibroblast strains, MDAH087 (p53-Mut, harboring a codon 248
single-base substitution) and MDAH041 (p53-Null, harboring a codon
184 frameshift mutation, resulting in premature termination of trans-
lation of p53 protein), both 200 population doubling, post-crisis p53
homozygous cell strains, were kindly provided by Dr. Michael Tainsky
(M. D. Anderson Cancer Center, Austin, TX). These fibroblast cells were
grown in DMEM supplemented with 10% fetal calf serum and antibi-
otics at 37 °C in a humidified atmosphere of 5% CO
2
. Normal HMEC
were established in culture according to Stampfer (28). HPV 16 E6
protein-expressing HMEC 76E6 and HPV 16 E7 protein-expressing
HMEC 76E7 were kindly provided by Dr. Vimla Band (Tufts University
School of Medicine, Boston, MA). These cells were grown in DFCI
medium supplemented with required nutrients and growth factors as
described (29). For experiments to assess DNA damage and repair, the
monolayer cells were grown to confluence in 150-mm dishes and then
placed in serum-free medium for 12 h. Under these growth and main-
tenance conditions, the index of cellular DNA replication, as measured
from changes in the specific radioactivity of DNA pre-labeled with
[
3
H]dThd, does not show significant genome duplication during the test
periods of repair analysis (27). The medium was removed, and cells
were washed with prewarmed phosphate-buffered saline and irradiated
with desired doses of UV (254-nm) light at a rate of 0.5 Jm
2
s
1
as
measured by Kettering model 65 radiometer (Cole-Palmer Instrument
Co., Vernon Hill, IL). After irradiation, fresh serum-free medium was
added to cell culture and incubation continued for desired periods.
DNA Isolation and Conversion of CPDs to Ligatable DNA Strand
Breaks—Briefly, after UV exposure, desired incubation periods, and
washes to remove any floating dead cells, the adherent cells were
recovered by trypsinization and immediately lysed for DNA isolation by
a salt precipitation procedure as described (13, 30). The CPDs were
cleaved and converted to single-strand breaks by digestion with T4
endonuclease V (31). The ligation-inhibiting 5-pyrimidine overhang
FIG.1.Repair of CPDs in non-transcribed strand of exon 8 of p53 in genomic DNA. Human fibroblasts were irradiated with an UV dose
of 20 Jm
2
and DNA isolated at indicated times after irradiation. CPDs in non-transcribed strand of genomic DNA were determined by LMPCR
using upper strand specific primers as described under “Materials and Methods.” Maxam-Gilbert-derived sequencing lanes, lanes CTand C, are
shown on the left of each autoradiogram. Formation or remaining amount of CPDs is shown at indicated postirradiation time points. Lanes marked
Con represent LMPCR profiles of DNA sample from cells without UV treatment. A, p53-WT normal human fibroblasts; B, p53-Null LFS fibroblasts;
C, p53-Mut LFS fibroblasts.
NER Modulation by p53 11493
was removed by Escherichia coli photolyase (31) (generous gift from Dr.
Aziz Sancar, University of North Carolina, Chapel Hill, NC). DNA was
then recovered and quantitated as described previously (13). The same
amount of DNA (1–2
g) was used for each reaction set of LMPCR.
LMPCR—LMPCR is an extremely sensitive genomic sequencing
method used for the detection of DNA damage (31, 32). DNA, specifi-
cally cleaved at CPDs, was used to create blunt end DNA fragment by
extension of primer specific to p53 with DNA sequenase 2.0. Blunt-
ended DNA was then ligated to a double-strand linker and followed by
amplification with the longer oligonucleotide of the linker and a second
nested p53-specific primer. After 20–21 cycles of polymerase chain
reaction, DNA fragments were size-fractionated on an 8% urea-polyac-
rylamide sequencing gel, electroblotted onto a nylon filter, and hybrid-
ized to a single-stranded p53-specific
32
P-labeled probe generated by
polymerase chain reaction using a third p53-specific primer. The filter
was used to expose a PhosphorImager screen and the individual band
intensities quantitated upon imaging and processing by Imagequant
software (Molecular Dynamics). The filters were also used to expose
Kodak X-Omat film for autoradiography and documentation by image
scanning. LMPCR analysis of each sample was carried out in duplicate,
and results described are from at least three independent experiments.
RESULTS
Loss or disruption of normal p53 function generally results in
decreased efficiency of repair of CPDs of non-transcribed DNA
strand. To compare repair of CPDs in p53-WT containing nor-
mal human fibroblasts, p53-Mut, and p53-Null LFS fibroblasts,
formation and removal of CPDs induced by UV at a dose of 20
Jm
2
were mapped by LMPCR, in both strands of exon 8 of p53
from genomic DNA. Under such an exposure condition, no
discernible differences in formation of CPDs in genomic DNA
could be seen in these three cell lines as measured by immuno-
slot blot assay (13). However, repair of CPDs by LMPCR could
be seen at almost all the potential dimer sites of NTS, albeit
with a clearly demonstrable variation in the rates of their
repair. For example, repair at many sites in p53-WT normal
human fibroblasts was apparent after 8 h, followed by an
approximately 30–70% repair after 24 h and 60–95% repair
after 48 h (Fig. 1A, Table I). Higher efficiency of removal of
CPDs was seen at 5side of adjacent pyrimidine sites, e.g.
codons 270, 274, and 289. Consistent with earlier observation
by Tornaletti and Pfeifer (32), repair of CPDs at codon 278,
which is one of the frequently mutated in human skin cancers,
was slower than those of surrounding positions. However, re-
pair of CPDs (5-TCTC^C) at codon 289/290 was also found to
be slower. It may be noted that this dimer is located at the 3
end of four adjacent potential pyrimidine sites.
In p53-Null cells, repair of CPDs at most of the dipyrimidine
sites of NTS was slower compared with that of the same CPDs
at same position in p53-WT cells. Moreover, slow repair sites
FIG.2.Repair of CPDs in transcribed strand of exon 8 of p53
genomic DNA in p53-WT, p53-Null, and p53-Mut LFS fibroblasts.
The LMPCR processing of the sample was as described under “Materi-
als and Methods.” The composite is a representative autoradiogram of
data from several independent experiments.
TABLE I
Repair of CPDs of exon 8 of p53 genomic DNA in normal and LFS fibroblasts
DNA Codon Sequence (5–3)
CPD remaining
a
after
p53-WT p53-Mut p53-Null
24 h 48 h 24 h 48 h 24 h 48 h
Non-transcribed %
270 TTT 22 6 65 60 58 38
TTT 4912 73606044
274/275 GTTT 6848 80566042
276/277 CTG 22 11 100 53 50 22
277/278(M) TCCT 36 34 100 53 65 40
278(M) CCT 32 18 100 97 65 40
289 CTC 10 9 100 80 34 20
CTC 17 10 100 80 54 22
289/290 CTCC 41 15 100 84 60 30
Transcribed 285/286(M) CCTC 44 23 62 31 40 40
286(M) TTC 4130 55315232
291 CTT 3018 40204722
292 TTT 10 8 34 8 30 14
293 CCC 32 17 47 20 34 8
294(M) CTC 3620 75504220
a
Repair rates were measured at mutation hotspots (M) and at various surrounding sites. Repair at each position is described as average
percentage derived from time versus repair plots of three independent experiments. Percentage of CPDs remaining was calculated from band
intensities at 24 and 48 h in reference to the band intensity at 0 h and normalized for any intensity observed at same sites in control lanes.
NER Modulation by p5311494
were more prominently affected (Fig. 1, Aand B, Table I). For
example, dimer C^CT, at codon 278, was 68% repaired at 24 h
and 82% repaired at 48 h in p53-WT cells, whereas in p53-Null
cells, there was only 35% and 60% repair observed at these
sites within 24 and 48 h, respectively. Repair of CPDs (5-
TCTC^C) at codon 289/290 was also drastically affected by the
loss of p53 function (Fig. 1B). These observations were further
confirmed by comparison of normal HMEC with HPV 16 E6
protein-expressing HMEC. In the case of 76E6 HMEC, in
which p53 protein is degraded by E6 protein-mediated ubiq-
uitin proteolysis pathway (33), the overall p53 modulation of
DNA repair events appeared exactly like that of p53-Null fi-
broblast cells described above (redundant data not shown).
Among the cell lines tested, repair of CPDs in NTS was most
dramatically affected in p53-Mut fibroblasts (Fig. 1Cand Table
I). In this cell line, repair of CPDs at all dipyrimidine sites was
significantly slow. Approximately 80–100% of CPDs remained
after 24 h, and 50–100% of CPDs remained 48 h after UV
irradiation. These results, in conjunction with the data from
p53-Null cells, indicate that mutant p53 protein affects DNA
repair in a dominant negative manner. This would seem to
suggest that p53 protein regulates DNA repair, at least of the
non-transcribed strand, by direct protein-protein interaction.
Such a dominant effect could be exerted by interaction with the
proteins of NER assembly or by altered transcription of pro-
teins essential for optimal assembling and targeting of the
damage recognition complex.
At the position of one potential dipyrimidine site, (5-
C^CTCACC) at codon 295, an abnormal signal was distinctly and
reproducibly detected in p53-Null and p53-Mut cells, but not in
p53-WT cells (Fig. 1, A–C). Accordingly, a distinct LMPCR gen-
erated band could be seen in the control unirradiated sample
lane. We surmise that this band could not arise due to nonspecific
or enzymatic cleavage as the signal gradually decreased between
FIG.3.Repair of CPDs in transcribed
strand of exon 8 of p53 genomic DNA in
normal, HPV 16 E6 and E7 protein-ex-
pressing HMEC. A, normal HMEC; B,
HPV 16 E6 protein-expressing HMEC,
76E6; C, HPV 16 E7 protein-expressing
HMEC, 76E7. The LMPCR processing of the
sample was as described under “Materials
and Methods.”
NER Modulation by p53 11495
4- and 48-h time intervals following irradiation. According to the
nature of LMPCR, an assay specialized for detecting nicks in
individual DNA strand, this signal could only represent a specific
DNA strand break at this site; due to its occurrence within DNA
topoisomerase consensus sequence, it could be the result of an
arrested cleavable complex. Surprisingly, such an abnormal
DNA break signal was also found in HPV 16 E6-expressing
HMEC, but not in normal or in HPV 16 E7-expressing HMEC.
This further confirmed that the DNA strand break at this codon
site is specific and only appears in cells that are rendered defi-
cient for normal p53 function. Interestingly, exposure of cells to
UV seemed to induce the repair of this DNA strand break, as was
evident from time-dependent gradual disappearance of the band
at this site.
Loss or Disruption of Normal p53 Function Affects the Repair
of CPDs at Slow Repair Sites of Transcribed DNA Strand—To
examine the effects of loss of normal p53 function on the repair
of CPDs within TS, repair of CPDs in TS of exon 8 of p53 gene
was mapped by LMPCR (Fig. 2). As expected, removal of CPDs
from TS was generally faster than from NTS, albeit with a
clearly discernible site-specific variation in the removal of
CPDs at individual dipyrimidine sites. Consistent with earlier
observation (32), repair of CPDs in p53-WT cells at codons 286
and 294 was seen to be slower than that of surrounding posi-
tions. A p53-dependent decreased repair of CPDs was found at
several sites of exon 8 in both p53-Null and p53-Mut LFS cells
at 24 h after UV irradiation. The intrinsically slow repair
dipyrimidine sites, e.g. at codons 286 and 294, were preferen-
tially affected by the absence of normal wild type function than
surrounding CPD sites (Fig. 2 and Table I). The intrinsically
slow repair is suggested to be the basis for mutational predis-
position of these p53 gene sites in human cancers (32). An
absence of p53 function would be expected to further exacer-
bate the cellular instability through decreased repair of exog-
enously or endogenously induced DNA damage.
Since the effect of loss of p53 function on TCR has not been
fully resolved, we extensively mapped repair of CPDs in TS of
exon 8 of p53 gene in normal, isogenic HPV 16 E6 and E7
protein-expressing HMECs. It may be noted that HPV 16 E7
protein selectively activates ubiquitin proteolysis pathway
causing degradation of pRb protein, while it stabilizes p53
protein (29). The data shows that the repair of CPD was faster
in HMEC than fibroblast, and dimers at most of the sites were
quantitatively removed within 48 h after UV treatment. Fur-
thermore, unlike normal fibroblasts, slow repair of CPDs
within sites like codon 286 and 294 was not very obvious in
either the normal or HPV 16 E7 protein-expressing HMEC
(Fig. 3, Aand C). Thus, p53-expressing cells appear to have
normal NER despite the absence of a functional pRb protein.
On the other hand, a clearly visible slower repair at the same
sites was found in p53-compromised HPV 16 E6 protein-ex-
pressing HMEC (Fig. 3Band Table II). This observation fur-
ther confirmed that loss of functional p53 affects the removal of
CPDs from TS by TCR and preferentially affects removal of
CPDs at slow repair sites.
DISCUSSION
The biochemical mechanisms of NER involve damage recog-
nition and open complex formation by factors such as XPA,
RPA, XPC, and transcription factor IIH, dual incision of the
damaged DNA strand by endonucleases XPF-ERCC1 and XPG,
repairsynthesismediated by a proliferating cell nuclearantigen-
dependent DNA polymerase and ligation of newly synthesized
DNA strand. The precise reaction mechanisms of NER have
recently been established to a significant extent (for review, see
Refs. 1–6). Nonetheless, how NER is regulated or connected to
other cellular functions still remains to be explored. Several
investigators have examined the involvement of p53 in regula-
tion of NER. Different systems and approaches have been used
for assessment of DNA repair. Immunoassay was mostly used
for direct assessment of GGR, whereas endonuclease-sensitive
site assay was used for examination of strand-specific repair. It
is becoming clear that functional p53 is required for efficient
GGR (13–17). However, due to different views being supported
by various TCR studies (14–17, 19, 26), precise nature of p53
participation in TCR still remains unclear. In this study, we
provide the first detailed analysis of effects of loss of functional
p53 on the removal of CPDs in both DNA strands at nucleotide
resolution. First, the results confirmed that functional p53 is
required for efficient GGR (13, 15, 16, 19). Furthermore, the
results show that expression of mutant p53 protein more sig-
nificantly affects removal of CPDs from non-transcribed strand
than loss of p53 protein, indicating that mutant p53 protein
affects GGR in a dominant negative manner. The results also
show that p53 is involved in TCR and repair of CPDs at slow
repair sites is the first target to sustain meaningful effects due
to the loss of p53 function.
In a recent review, Mckay et al. (17) have strongly argued
that p53 plays a definitive role in TCR. The difference of ob-
servations by various laboratories may be the result of different
assays used to detect strand-specific repair. Using the normal
endonuclease-sensitive site assay, we too were unable to dem-
onstrate any detectable differences in TCR between normal
and LFS fibroblasts as well as between normal and HPV 16 E6
protein-expressing HMEC for both UV- and benzo[a]pyrene
diol epoxide-induced DNA damage.
2
Repair differences be-
tween cells did not become pronounced and meaningful until
full mapping of repair of CPDs in normal, HPV 16 E6 and E7
protein-expressing HMEC was conducted within the same gene
2
Q. Zhu, M. A. Wani, M. El-mahdy, and A. A. Wani, unpublished
results.
TABLE II
Repair of CPDs of exon 8 of p53 genomic DNA in normal, HPV 16 E6 and E7 protein-expressing HMEC
DNA Codon Sequence (5–3)
CPD remaining after
HMEC 76E6
a
76E7
a
24 h 48 h 24 h 48 h 24 h 48 h
Transcribed %
285/286(M) TTC^C 5 1 30 18 10 2
286(M) TT^C 14 2 40 20 8 0
T^TC 26 6 50 16 20 0
287 CT^C 16 0 24 6 12 2
291 CT^T 10 1 13 0 7 0
292 TT^T 16 2 35 10 16 3
293 C^CC 8 0 40 30 14 2
294(M) CT^C 10 0 40 22 12 1
a
76E6 and 76E7 represent HPV 16 E6 and E7 protein-expressing HMEC, respectively. Extent of repair at each site was quantitated at indicated
times as described in Table I.
NER Modulation by p5311496
segment by LMPCR. The reasons are that (i) LMPCR is a much
more powerful assay in demonstrating variations of repair
along particular sites and stretches of specific gene sequences,
(ii) fully discernible differences were visualized mainly during
the initial stages of repair, i.e. before 24-h time points, and (iii)
slow repair sites were more prominently subjected to the influ-
ence exerted by the loss of p53 function.
Several investigations suggest that p53-regulated gene prod-
ucts may participate or be associated with NER. Using host
reactivation assay, it has been shown that UV- and heat shock-
inducible NER is p53-dependent (14, 17). More recently, it has
been shown that the expression of p53-downstream genes, e.g.
p48 gene, was dependent on p53 and involved in GGR (22). It
should be noted that p48 has been suggested to have a role in
the repair of DNA in chromatin and damage recognition. If
effects of loss of function in p53-Null LFS fibroblasts and HPV
16 E6 protein-expressing HMEC on NER reflect requirements
of p53-activated downstream genes in NER, dominant negative
effects of mutant p53 protein on DNA repair of non-transcribed
strand may reflect direct participation of p53 in NER. It has
been shown that p53 binds to three components of basal tran-
scription factor: p62, XPD, and XPB (10). Furthermore, XPB
and XPD cells are deficient in repair of non-transcribed DNA
and inefficiently repair the transcribed strand including se-
quences near the transcription start site (34). However, contact
with these factors may not be the only mechanism by which p53
directly participates in NER. Given that mutant p53 protein
negatively regulates NER by binding XPB and XPD, TCR
should be the first target. However, dominant negative effects
of mutant p53 protein on DNA repair of non-transcribed strand
were more distinctly observed. In fact, wild-type rather than
mutant p53 protein has been shown to inhibit DNA helicase
activity of XPD and XPB (35). In search for the components of
NER complex that could be interacting in vivo with p53, we
have found that recognition of UV-induced damage links p53
pathway to NER and that HHR23A is involved in regulating
transcriptional activity of p53.
3
Interestingly, it has been
shown that XPC protein, complexed with HHR23A and
HHR23B, also plays a role in damage recognition and chroma-
tin unfolding (37). It seems that p53 regulates the early steps of
damage recognition of NER or chromatin unfolding during
NER processes through both transcription activation and pro-
tein-protein interaction. However, no experimental evidence
shows that p53 protein recognizes or binds UV-induced CPDs.
In eukaryotic cells, genomic DNA is wrapped around histone
octamers forming nucleosomes, which are the repeating units
of chromatin. Proteins involved in cellular processes, such as
DNA replication, transcription, and DNA repair, require access
to DNA within chromatin structural hierarchy. Heterogeneity
of NER may partially reflect the accessibility of damaged DNA
to NER components. In support of this, very fast repair has
been seen in both DNA strands near the transcription initia-
tion site (38). It has also been shown that several sequence
positions in 5-flanking region of the tRNA
val
gene, which lacks
TCR, were also repaired more efficiently than the gene itself
(38, 39). In the case of a run of potential CPD sites in the
non-transcribed DNA strand, if each DNA repair event in GGR
is considered as an independent event, as suggested by its
biochemical mechanisms, there is no reason for higher effi-
ciency of CPD removal at 5end. This is only possible if TCR or
transcription somehow helped damaged DNA to become more
accessible at 5end. Thus, besides transcription coupling, ac-
cessibility of damaged DNA contributes a very important pa-
rameter to the heterogeneity of NER. In this regard, p53 may
regulate NER through modulating accessibility of damaged
DNA rather than damage recognition. One would expect that
CBP/p300, a p53 coactivator that has been shown to have
acetyltransferase activity (36), could also be involved in p53-
regulated DNA repair. Investigation of such DNA repair par-
ticipating principles should become an area of active interest in
the near future.
Acknowledgments—We are grateful to Dr. Aziz Sancar for providing
photolyase enzyme and John Croyle for assistance with high resolution
image scanning.
REFERENCES
1. Hanawalt, P. C. (1998) Mutat. Res. Fundam. Mol. Mech. Mutagen. 400,
117–125
2. Petit, C., and Sancar, A. (1999) Biochimie 81, 15–25
3. Hoeijmakers, J. H. J., and Bootsma, D. (1994) Nature 371, 654–655
4. Lindahl, T., Karran, P., and Wood, R. D. (1997) Curr. Opin. Genet. Dev. 7,
158–169
5. Wood, R. D. (1997) J. Biol. Chem. 272, 23465–23468
6. Mullenders, L. H. (1998) Mutat. Res. DNA Repair 409, 59–64
7. Drapkin, R., Sancar, A., and Reinberg, D. (1994) Cell 77, 9–12
8. Hanawalt, P. C., Donahue, B. A., and Sweder, K. S. (1994) Curr. Biol. 4,
518–521
9. Tornaletti, S., and Hanawalt, P. C. (1999) Biochimie 81, 139–146
10. Frit, P., Bergmann, E., and Egly, J.-M. (1999) Biochimie 81, 27–38
11. Oren, M., and Rotter, V. (1999) Cell. Mol. Life Sci. 55, 9–11
12. Agarwal, M. L., Taylor, W. R., Chernov, M. V., Chernova, O. B., and Stark,
G. R. (1998) J. Biol. Chem. 273, 1–4
13. Wani, M. A., Zhu, Q. Z., El-mahdy, M., and Wani, A. A. (1999) Carcinogenesis
20, 765–772
14. Smith, M. L., Chen, I.-T., Zhan, Q., O’Connor, P. M., and Fornace, A. J., Jr.
(1995) Oncogene 10, 1053–1059
15. Ford, J. M., and Hanawalt, P. C. (1995) Proc. Natl. Acad. Sci. U. S. A. 92,
8876–8880
16. Ford, J. M., and Hanawalt, P. C. (1997) J. Biol. Chem. 272, 28073–28080
17. McKay, B. C., Francis, M. A., and Rainbow, A. J. (1997) Carcinogenesis 18,
245–249
18. Ford, J. M., Baron, E. L., and Hanawalt, P. C. (1998) Cancer Res. 58, 599–603
19. Prost, S., Ford, J. M., Taylor, C., Doig, J., and Harrison, D. J. (1998) J. Biol.
Chem. 273, 33327–33332
20. Jia, L., Wang, X. W., and Harris, C. C. (1999) Int. J. Cancer 80, 875–879
21. Smith, M. L., Chen, I. T., Zhan, Q., Bae, I., Chen, C.-Y., Gilmer, T. M., Kastan,
M. B., O’Connor, P. M., and Fornace, A. J., Jr. (1994) Science 266,
1376–1380
22. Hwang, B. J., Ford, J. M., Hanawalt, P. C., and Chu, G. (1999) Proc. Natl.
Acad. Sci. U. S. A. 96, 424– 428
23. Lee, S., Elenbass, B., Levine, A. J., and Griffith, J. (1995) Cell 81, 1013–1020
24. Wang, X. W., Forrester, K., Yeh, H., Feitelson, M. A., Gu, J.-R., and Harris,
C. C. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 2230–2234
25. McKay, B. C., Ljungman, M., and Rainbow, A. J. (1999) Carcinogenesis 20,
1389–1396
26. Mirzayans, R., Enns, L., Dietrich, K., Barley, R. D. C., and Paterson, M. C.
(1996) Carcinogenesis 17, 691–698
27. Venkatachalam, S., Denissenko, M. F., and Wani, A. A. (1995) Carcinogenesis
16, 2029–2036
28. Stampfer, M. R. (1985) J. Tissue Culture Methods 9, 107–115
29. Wazer, D. E., Liu, X.-L., Chu, Q., Gao, Q., and Band, V. (1995) Proc. Natl. Acad.
Sci. U. S. A. 92, 3687–3691
30. Miller, S. A., Dykes, D. D., and Polesky, H. F. (1988) Nucleic Acids Res. 16,
12–15
31. Tornaletti, S., and Pfeifer, G. (1996) in Technologies for Detection of DNA
Damage and Mutations (Pfeifer, G. P., ed) pp. 199–210, Plenum Press, New
York
32. Tornaletti, S., and Pfeifer, G. P. (1994) Science 263, 1436–1438
33. Scheffner, M., Werness, J. M., Huibregtse, J. M., Levine, A. J., and Howley,
P. M. (1990) Cell 63, 1129–1136
34. Tu, Y., Bates, S., and Pfeifer, G. P. (1997) J. Biol. Chem. 272, 20747–20755
35. Wang, X. W., Yeh, H., Schaeffer, L., Roy, R., Moncollin, V., Egly, J.-M., Wang,
Z., Friedberg, E. C., Evans, M. K., Taffe, B. G., Bohr, V. A., Weeda, G.,
Hoeijmakers, J. H. J., Forrester, K., and Harris, C. C. (1995) Nat. Genet. 10,
188–195
36. Chakravarti, D., Ogryzko, V., Kao, H.-Y., Nash, A., Chen, H., Nakatani, Y.,
and Evans, R. M. (1999) Cell 96, 393–403
37. Baxter, B. K., and Smerdon, M. J. (1998) J. Biol. Chem. 273, 17517–17524
38. Tu, Y., Tornaletti, S., and Pfeifer, G. P. (1996) EMBO J. 15, 675–683
39. Dammann, R., and Pfeifer, G. P. (1997) Mol. Cell. Biol. 17, 219–229
3
Q. Z. Zhu, M. El-mahdy, M. A. Wani, and A. A. Wani, submitted for
publication.
NER Modulation by p53 11497
... In addition, it was recently found that p53 is able to reduce the frequency of chemically induced point mutations (5). Disruption of p53 function in some systems resulted in a deficiency in the rate and extent of nucleotide excision repair (NER) (5)(6)(7). ...
... It was suggested that p53 might facilitate the initial stages of DNA repair in a transcription-independent manner, probably through protein-protein interactions (7). The ability to interact with gapped DNA might provide the basis for a possible role of p53 as a recruitment factor in NER, especially when taking into consideration the ability of p53 to bind to NER helicases XPB, XPD and CSB (14). ...
Article
Full-text available
In addition to binding DNA in a sequence-specific manner, the p53 tumour suppressor protein can interact with damaged DNA. In order to understand which structural features in DNA the C-teminal domain recognises we have studied the interaction of p53 protein with different types of DNA oligonucleotides imitating damaged DNA. Here we show that one unpaired nucleotide within double-stranded (ds)DNA is sufficient for recognition by the p53 C-terminus, either as a protruding end or as an internal gap in dsDNA. C-terminal interaction with DNA ends facilitated core domain binding to DNA, whereas interaction with gaps prevented core domain–DNA complexing, implying that p53 might adopt distinct conformations upon binding to different DNA lesions. These observations suggest that both single-strand and double-strand breaks can serve as a target for p53 C-terminal recognition in vivo and indicate that p53 might recruit different repair factors to the sites of damaged DNA depending on the type of lesion.
... It recognizes DNA adducts and activates apoptosis [62]. All of them are regulated by p53, which can trigger cell cycle arrest and DNA repair or apoptosis ( Fig. 1) [63][64][65][66]. ...
Article
Although platinum-based chemotherapy remains the standard treatment for advanced NSCLC patients, clinical outcomes are poor and most patients develop high-grade toxicities. Genetic factors, such as single nucleotide polymorphisms (SNPs) involved in platinum pharmacodynamics, metabolism and mechanism of action, may account for inter-individual differences shown in effectiveness and toxicity. Polymorphisms in genes involved in DNA repair and others such as PI3K/PTEN/AKT and TGF-β pathways have been demonstrated to be associated with response, survival and toxicity in advanced NSCLC patients treated with platinum-based chemotherapy. Other cellular processes, like DNA methylation and proliferation have been connected with clinical outcome for platinum-based chemotherapy regimens through folate metabolism and cytokine signaling.
... Consistent results have generally confirmed that p53 is important in the GG-NER pathway, so the controversy mostly surrounds its involvement in TC-NER. Clouding the issue even further, three later studies from Altaf Wani's group provided additional evidence that p53 is primarily involved in GGNER, contradicting the results of Mirzayans et al. (1996) and agreeing with the earlier results from the Hanawalt group ( Zhu et al. 2000;Wani et al. 2000;. Can these contradictions be reconciled? ...
Article
The cells in the human body are continuously challenged by a variety of genotoxic attacks. Erroneous repair of the DNA can lead to mutations and chromosomal aberrations that can alter the functions of tumor suppressor genes or oncogenes, thus causing cancer development. As a central tumor suppressor, p53 guards the genome by orchestrating a variety of DNA-damage-response (DDR) mechanisms. Already early in metazoan evolution, p53 started controlling the apoptotic demise of genomically compromised cells. p53 plays a prominent role as a facilitator of DNA repair by halting the cell cycle to allow time for the repair machineries to restore genome stability. In addition, p53 took on diverse roles to also directly impact the activity of various DNA-repair systems. It thus appears as if p53 is multitasking in providing protection from cancer development by maintaining genome stability.
... More recent results have shown that cells with a homozygous mutation in the p53 gene and fibroblasts expressing the HPV E6 gene were deficient in global nucleotide excision repair (NER), but not in transcription-coupled repair, i.e. the repair of the transcribed strand in active genes (11)(12)(13). Other studies reported an impairment of both NER sub-pathways in various cell lines (14)(15)(16)(17)(18). Thus, it was suggested that p53 could play a role in DNA repair (19) and a direct involvement in the base excision repair (BER) process has recently been demonstrated (20,21). ...
Article
Full-text available
The tumour suppressor gene TP53 plays an important role in the regulation of DNA repair, and particularly of nucleotide excision repair. The influence of p53 status on the efficiency of the principal steps of this repair pathway was investigated after UV-C irradiation in the human ovarian carcinoma cell line IGROV-1 (expressing wild-type p53) and in the derived clone IGROV-1/Pt1 (with p53 mutations at codons 270 and 282). Clonogenic survival after UV-C irradiation showed that IGROV-1/Pt1 cells were ∼2-fold more resistant to DNA damage than parental cells. Modulation of p53 protein levels, cell cycle arrest and apoptosis were induced in UV-irradiated IGROV-1 cells, but not in the p53-mutant cell line. Exposure to UV or cisplatin induced down-regulation of p53-replication protein A (RPA) interaction in parental, but not in IGROV-1/Pt1 cells. However, persistent binding of p53 to RPA did not affect the early steps of DNA repair. In fact, both UV-induced DNA incision and the recruitment of proliferating cell nuclear antigen (PCNA) to DNA repair sites occurred to a comparable extent in p53-wild type and -mutant cell lines, although PCNA remained associated with chromatin for a longer period of time in IGROV-1/Pt1 cells. Global genome repair, as detected by immunoblot analysis of cyclobutane pyrimidine dimers, was not significantly different in the two cell lines at 3 h after UV irradiation. In contrast, lesion removal at 24 h was markedly reduced in IGROV-1/Pt1 cells, being ∼25% of the initial amount of damage, as compared with ∼50% repair in parental cells. These results indicate that the presence of mutant p53 protein and its persistent interaction with RPA do not affect the early steps of nucleotide excision repair in IGROV-1/Pt1 cells. Thus, repair defects in p53-mutant ovarian carcinoma cells may be attributed to late events, possibly related to a reduced removal/recycling of PCNA at repair sites. Chemicals/CAS: DNA, Neoplasm; DNA-Binding Proteins; Proliferating Cell Nuclear Antigen; Replication Protein A; RPA1 protein, human; Tumor Suppressor Protein p53
... The exact mechanism involved in the downregulation of DNA repair gene expression is currently unknown; however, p53 suppression might play a role in this process (Smith et al., 2000). p53 controls several aspects of the cell cycle that allow for DNA repair, and it also has been implicated directly in the regulation of DNA repair genes (Zhu et al., 2000). A previous report showed a decrease in p53 protein levels in immortalized keratinocytes (HaCaT) in response to AsIII exposure (Hamadeh et al., 1999) and hypermethylation of a p53 promoter region was observed in A549 cells exposed to AsIII (Mass and Wang, 1997). ...
Article
Human exposure to arsenic, a ubiquitous and toxic environmen- tal pollutant, is associated with an increased incidence of skin cancer. However, the mechanism(s) associated with AsIII-medi- ated toxicity and carcinogenesis at low levels of exposure remains elusive. Aberrations in cell proliferation, oxidative damage, and DNA-repair fidelity have been implicated in sodium arsenite (AsIII)-mediated carcinogenicity and toxicity, but these events have been examined in isolation in the majority of biological models of arsenic exposure. We hypothesized that the simulta- neous interaction of these effects may be important in arsenic- mediated neoplasia in the skin. To evaluate this, normal human epidermal keratinocytes (NHEK) were exposed to nontoxic doses (0.005-5 M) of AsIII and monitored for several physiological endpoints at the times when cells were harvested for gene expres- sion measurements (1-24 h). Two-fluor cDNA microarray analy- ses indicated that AsIII treatment decreased the expression of genes associated with DNA repair (e.g., p53 and Damage-specific DNA-binding protein 2) and increased the expression of genes indicative of the cellular response to oxidative stress (e.g., Super- oxide dismutase 1, NAD(P)H quinone oxidoreductase, and Serine/ threonine kinase 25). AsIII also modulated the expression of cer- tain transcripts associated with increased cell proliferation (e.g., Cyclin G1, Protein kinase C delta), oncogenes, and genes associ- ated with cellular transformation (e.g., Gro-1 and V-yes). These observations correlated with measurements of cell proliferation and mitotic measurements as AsIII treatment resulted in a dose- dependent increase in cellular mitoses at 24 h and an increase in cell proliferation at 48 h of exposure. Data in this manuscript demonstrates that AsIII exposure simultaneously modulates DNA repair, cell proliferation, and redox-related gene expression in nontransformed, normal NHEK. It is anticipated that data in this report will serve as a foundation for furthering our knowledge of AsIII-regulated gene expression in skin and other tissues and contribute to a better understanding of arsenic toxicity and carci- nogenesis.
Article
Xeroderma pigmentosum group C protein (XPC) acts as a DNA damage recognition factor for bulky adducts and as an initiator of global genome nucleotide excision repair (GG-NER). Novel insights have shown that the role of XPC is not limited to NER, but is also implicated in DNA damage response (DDR), as well as in cell fate decisions upon stress. Moreover, XPC has a proteolytic role through its interaction with p53 and casp-2S. XPC is also able to determine cellular outcomes through its interaction with downstream proteins, such as p21, ARF, and p16. XPC interactions with effector proteins may drive cells to various fates such as apoptosis, senescence, or tumorigenesis. In this review, we explore XPC’s involvement in different molecular pathways in the cell and suggest that XPC can be considered not only as a genomic caretaker and gatekeeper but also as a tumor suppressor and cellular-fate decision maker. These findings envisage that resistance to cell death, induced by DNA-damaging therapeutics, in highly prevalent P53-deficent tumors might be overcome through new therapeutic approaches that aim to activate XPC in these tumors. Moreover, this review encourages care providers to consider XPC status in cancer patients before chemotherapy in order to improve the chances of successful treatment and enhance patients’ survival.
Article
Exposure of skin cells to UV radiation results in DNA damage, which if inadequately repaired, may cause mutations. UV-induced DNA damage and reactive oxygen and nitrogen species also cause local and systemic suppression of the adaptive immune system. Together, these changes underpin the development of skin tumours. The hormone derived from vitamin D, calcitriol (1,25-dihydroxyvitamin D3) and other related compounds, working via the vitamin D receptor and at least in part through endoplasmic reticulum protein 57 (ERp57), reduce cyclobutane pyrimidine dimers and oxidative DNA damage in keratinocytes and other skin cell types after UV. Calcitriol and related compounds enhance DNA repair in keratinocytes, in part through decreased reactive oxygen species, increased p53 expression and/or activation, increased repair proteins and increased energy availability in the cell when calcitriol is present after UV exposure. There is mitochondrial damage in keratinocytes after UV. In the presence of calcitriol, but not vehicle, glycolysis is increased after UV, along with increased energy-conserving autophagy and changes consistent with enhanced mitophagy. Reduced DNA damage and reduced ROS/RNS should help reduce UV-induced immune suppression. Reduced UV immune suppression is observed after topical treatment with calcitriol and related compounds in hairless mice. These protective effects of calcitriol and related compounds presumably contribute to the observed reduction in skin tumour formation in mice after chronic exposure to UV followed by topical post-irradiation treatment with calcitriol and some, though not all, related compounds.
Chapter
Exposure of skin cells to UV radiation results in DNA damage, which if inadequately repaired, causes mutations. UV-induced DNA damage and reactive oxygen and nitrogen species also cause local and systemic suppression of the adaptive immune system. Together these changes facilitate the development of skin tumors. The vitamin D hormone, 1,25-dihydroxyvitamin D3, working via the vitamin D receptor and at least in part, through endoplasmic reticulum protein 57, reduces oxidative DNA damage by decreasing reactive oxygen and nitrogen species. This is facilitated by increased metallothionein expression and increased nuclear factor-erythroid-2-related factor 2 transcriptional activity. 1,25-Dihydroxyvitamin D3 also enhances DNA repair, through decreased reactive oxygen and nitrogen species, increased p53 expression and/or activation, increased energy availability in the cell, increased phosphatase and tensin homolog deleted on chromosome 10, and increased repair proteins. Reduced DNA damage and reduced reactive oxygen and nitrogen species reduce UV-induced immune suppression. Several related metabolites and vitamin D-like compounds also reduce DNA damage and immune suppression after UV. These protective effects of 1,25-dihydroxyvitamin D and some related compounds result in less skin tumor formation after chronic exposure to UV.
Article
Full-text available
Dermal fibroblast strains cultured from affected members of a cancer-prone family with Li-Fraumeni syndrome (LFS) harbor a point mutation in one allele of the p53 tumor suppressor gene, resulting in loss of normal p53 function, In this study we have examined the ability of these p53-deficient strains to carry out the long-patch mode of excision repair, mediated by DNA polymerases delta and epsilon, after exposure to Co-60 gamma radiation or far ultraviolet (UV) (chiefly 254 mm) light, Repair was monitored by incubation of the irradiated cultures in the presence of aphidicolin (ape) or 1-beta-D-arabinofuranosylcytosine (araC), each a specific inhibitor of long-patch repair, followed by measurement of drug-induced DNA strand breaks (reflecting non-ligated strand incision events) by alkaline sucrose velocity sedimentation, The LFS strains displayed deficient repair capacity in response to both gamma rays and UV light. The repair anomaly in UV-irradiated LFS cultures was manifested not only in the overall genome, but also in the transcriptionally active, preferentially repaired c-myc gene, Using autoradiography we also assessed unscheduled DNA synthesis (UDS) after UV irradiation and found this conventional measure of repair replication to be deficient in LFS strains. Moreover, both ape and araC decreased the level of UV-induced UDS by similar to 75% in normal cells, but each had only a marginal effect on LFS cells, We further demonstrated that the LFS strains are impaired in the recovery of both RNA and replicative DNA syntheses after UV treatment, two molecular anomalies of the DNA repair deficiency disorders xeroderma pigmentosum and Cockayne's syndrome, Together these results imply a critical role for wild-type p53 protein in DNA polymerase delta/epsilon-mediated excision repair, both the mechanism operating on the entire genome and that acting on expressed genes.
Chapter
UV irradiation of DNA produces two major types of DNA lesions:the cyclobutane pyrimidine dimers (CPDs) and the pyrimidine (6−4) pyrimidone photoproducts [(6−4) photoproducts]. Cyclobutane dimers are formed between the 5,6 bonds of two adjacent pyrimidines, 5′-TpT, 5′-TpC, 5′-CpT, or 5′-CpC. The (6−4) photoproducts have a covalent bond between positions 6 and 4 of two adjacent pyrimidines, and are detected most frequently at 5′-TpC and 5′-CpC sequences. The (6−4) photoproducts are formed at a rate of approximately 30% of that of CPDs and this ratio appears to be DNA sequence-dependent (Mitchell and Nairn, 1989). Both photoproducts are mutagenic in Escherichia coli and in mammalian cells (Brash, 1988).
Article
Ultraviolet (UV) light-induced DNA damage is repaired by the nucleotide excision repair pathway, which can be subdivided into transcription-coupled repair (TCR) and global genome repair (GGR). Treatment of cells with a priming dose of UV light appears to stimulate both GGR and TCR, suggesting that these processes are inducible. GGR appears to be disrupted in p53-deficient fibroblasts, whereas the effect of p53 disruption on TCR remains somewhat controversial. Normal recovery of mRNA synthesis following UV irradiation is thought to depend on TCR. We have found that the recovery of mRNA synthesis following exposure to UV light is severely attenuated in p53-deficient human fibroblasts. Therefore, p53 disruption may lead to a defect in TCR or a post-repair process required for the recovery of mRNA synthesis. Several different functions of p53 have been proposed which could contribute to these cellular processes. We suggest that p53 could participate in GGR and the recovery of mRNA synthesis following UV exposure through the regulation of steady-state levels of one or more p53-regulated gene products important for these processes. Furthermore, we suggest that the role of p53 in the recovery of mRNA synthesis is important for resistance to UV-induced apoptosis.
Article
Human hepatitis B virus (HBV) is a major risk factor of human hepatocellular carcinoma. Both in vivo and in vitro studies have shown that HBV X protein (HBx) can bind to the p53 tumor-suppressor protein and interfere with the role that p53 plays in the cellular response to DNA damage. Our previous work has shown that HBx protein inhibits p53 sequence-specific transcriptional activation, p53-mediated apoptosis and p53 binding to the TFIIH transcription nucleotide excision repair (NER) factors, including XPB and XPD. To investigate whether HBx interferes with the NER pathway, we utilized cell-proliferation and colony-formation assays to determine if cells expressing HBx are more sensitive to UVC-induced DNA damage. NER was also measured by a plasmid host cell re-activation assay using a vector containing a luciferase reporter gene. UV-irradiated plasmids were transfected into a human RKO colon carcinoma cell line that contains wild-type (wt) p53 as well as its derivatives, either mutant p53-143(ala) (RKO-143(ala)) or human papillomavirus E6 (RKO-E6, a wt p53 protein that is rapidly degraded and non-functional). We found that cells expressing HBx are more sensitive to UVC induced killing. Moreover, expression of HBx resulted in a reduction of NER efficiency in RKO cells to 52 +/- 2% (compared with control), RKO-143(ala) cells to 46 +/- 3% and RKO-E6 cells to 60 +/- 3%. Similar results were also obtained with a HepG2 hepatoblastoma cell line carrying wt p53. In addition, we found that HBx bound directly to either XPB or XPD DNA helicase in vitro. Thus, our data indicate that HBx may interfere with the NER pathway through both p53-dependent and p53-independent mechanisms. Because HBx binds to TFIIH-associated proteins, we propose that HBx may interfere with the NER pathway also through binding to and altering the activities of helicases necessary for NER and, thereby, increase the mutation rate induced by chemical carcinogens, such as aflatoxin B-1, during human liver carcinogenesis. Int. J. Cancer 80:875-879, 1999, Published 1999 Wiley-Liss, Inc.dagger
Article
Techniques for the isolation and culture of human mammary epithelial cells are described. The isolation procedure consists of dissection followed by partial enzymatic digestion with collagenase and hyal-uronidase and subsequent filtration to separate the epithelial clusters from the digested stromal elements. Culture procedures utilizing two different growth media are presented. A serum-free medium, MCDB170, permits long-term growth (45 to 60 population doublings) of a pure epithelial population; a less defined medium, MM, yields fewer population doublings but increased expression of some mammary-specific properties.
Article
Deleterious alterations in cellular DNA result from endogenous sources of damage, as well as from external radiations and genotoxic chemicals in the environment. Although it is often difficult to ascertain the relative contributions to biological endpoints from endogenous vs. environmental sources of genomic instability, such determinations are highly relevant to risk estimates based upon perceived toxic levels of environmental agents. Of particular concern are the DNA lesions caused by reactive oxygen species that are generated both as a byproduct of oxidative metabolism and as a consequence of exposure to ionizing radiation and some other toxicants. We need to better understand the sequence of biochemical events that occurs between the initial formation of a DNA lesion and the biological outcome. These events may include transcription, replication, and cell cycle regulation, as well as DNA repair. Heterogeneity in the intragenomic distribution of lesions and their repair must also be taken into count. Expressed genes are unusually susceptible to alteration by some agents, and preferential repair of some lesions is targeted to transcribed DNA strands. An arrested RNA polymerase at a lesion may block access of repair enzymes, and it may also serve as a signal for upregulation of repair enzymes, cell cycle arrest and/or apoptosis. Our current understanding of the role of transcription in lesion processing and biological outcomes will be summarized, with particular emphasis upon the information gained from characterization of human genetic diseases expressing defects in the processing of damaged DNA. In some cases, the clinical features of these diseases might be understood in terms of deficiencies in the repair of lesions that arrest transcription.
Article
The E6 protein encoded by the oncogenic human papillomavirus types 16 and 18 is one of two viral products expressed in HPV-associated cancers. E6 is an oncoprotein which cooperates with E7 to immortalize primary human keratinocytes. Insight into the mechanism by which E6 functions in oncogenesis is provided by the observation that the E6 protein encoded by HPV-16 and HPV-18 can complex the wild-type p53 protein in vitro. Wild-type p53 gene has tumor suppressor properties, and is a target for several of the oncoproteins encoded by DNA tumor viruses. In this study we demonstrate that the E6 proteins of the oncogenic HPVs that bind p53 stimulate the degradation of p53. The E6-promoted degradation of p53 is ATP dependent and involves the ubiquitin-dependent protease system. Selective degradation of cellular proteins such as p53 with negative regulatory functions provides a novel mechanism of action for dominant-acting oncoproteins.
Article
Multiple mammary epithelial cell (MEC) types are observed both in mammary ducts in vivo and in primary cultures in vitro; however, the oncogenic potential of different cell types remains unknown. Here, we used human papilloma virus 16 E6 and E7 oncogenes, which target p53 and Rb tumor suppressor proteins, respectively, to immortalize MECs present in early or late passages of human mammary tissue-derived cultures or in milk. One MEC subtype was exclusively immortalized by E6; such cells predominated in late-passage cultures but were rare at early passages and apparently absent in milk. Surprisingly, a second cell type, present only in early-passage tissue-derived cultures, was fully immortalized by E7 alone. A third cell type, observed in tissue-derived cultures and in milk, showed a substantial extension of life span with E7 but eventually senesced. Finally, both E6 and E7 were required to fully immortalize milk-derived MECs and a large proportion of MECs in early-passage tissue-derived cultures, suggesting the presence of another discrete subpopulation. Identification of MECs with distinct susceptibilities to p53- and Rb-targeting human papillomavirus oncogenes raises the possibility that these cells may serve as precursors for different forms of breast cancer.