ArticlePDF Available

Differential requirement for p19ARF in the p53-dependent arrest induced by DNA damage, microtubule disruption, and ribonucleotide depletion

Authors:

Abstract and Figures

p19ARF has been implicated as a key regulator of p53 stability and activation. While numerous stresses activate the p53 growth arrest pathway, those requiring p19ARF remain to be elucidated. We used p19ARF knockout mouse embryo fibroblasts to show that DNA damage and microtubule disruption require p19ARF to induce p53 responses, whereas ribonucleotide depletion and inhibition of RNA synthesis by low doses of actinomycin D do not. The data provide evidence that the arrest pathway activated by ribonucleotide depletion involves some different signal transducers than those activated by DNA damage or microtubule disruption. We also present biochemical analyses that provide insights into the mechanism by which p53 and p19ARF cooperate in normal cells to induce cell cycle arrest.
Content may be subject to copyright.
Differential requirement for p19ARF in the
p53-dependent arrest induced by DNA damage,
microtubule disruption, and ribonucleotide depletion
Shireen Hussain Khan*
, Janet Moritsugu*, and Geoffrey M. Wahl*
*Gene Expression Laboratory, The Salk Institute, La Jolla, CA 92037; and Department of Biology, University of California at San Diego, La Jolla, CA 92037
Communicated by Tony Hunter, The Salk Institute for Biological Studies, San Diego, CA, December 21, 1999 (received for review July 9, 1999)
p19ARF has been implicated as a key regulator of p53 stability and
activation. While numerous stresses activate the p53 growth arrest
pathway, those requiring p19ARF remain to be elucidated. We
used p19ARF knockout mouse embryo fibroblasts to show that
DNA damage and microtubule disruption require p19ARF to induce
p53 responses, whereas ribonucleotide depletion and inhibition of
RNA synthesis by low doses of actinomycin D do not. The data
provide evidence that the arrest pathway activated by ribonucle-
otide depletion involves some different signal transducers than
those activated by DNA damage or microtubule disruption. We also
present biochemical analyses that provide insights into the mech-
anism by which p53 and p19ARF cooperate in normal cells to induce
cell cycle arrest.
The tumor suppressors p53 and pRb participate in signal
transduction pathways that respond to various environmen-
tal and intracellular challenges, such as hypoxia, ionizing and
ultraviolet radiation, ribonucleotide depletion, microtubule dis-
ruption, and oncogene activation (1– 8). The responses to these
stresses range from induction of a reversible cell cycle arrest
when DNA damage is not incurred to damage-dependent acti-
vation of pathways that lead to premature senescence or apo-
ptosis. The net result of activation of the pathway, therefore, is
to remove cells exposed to stressing conditions from the repli-
cative cycle. During neoplastic progression, there is a strong
selection for mutations that inactivate the p53–pRb pathways,
thus enabling cells to avoid apoptosis or senescence and to
continue dividing under conditions that can increase genetic
instability (9, 10).
Recent experiments have focused attention on the Ink4aARF
locus, which has the potential to regulate both p53 and pRb
function. The p16Ink4a protein (inhibitor of cdk4) ensures that
pRb remains active by preventing the inhibitory phosphoryla-
tions by cyclin DCDK4,6 (11). p19ARF (mouse; p14ARF,
human), partly encoded by exons 2 and 3 of the p16Ink4a gene
in an alternative reading frame, regulates the stability of p53 by
binding to mdm2 (12–15). mdm2 and p53 participate in an
autoregulatory feedback loop in which p53 induces mdm2 ex-
pression (16). mdm2, in turn, mediates p53 degradation and
prevents p53 association with the basal transcriptional machin-
ery (17–19). The binding of mdm2 to p53 is affected by phos-
phorylation of p53 at the N terminus, which compromises mdm2
binding (20, 21). More recently, mdm2 was shown to be regulated
by interaction with p19ARF, which binds to and alters the
subcellular localization of mdm2 so that it has reduced access to
p53 (13, 14, 22, 23).
The available data predict that disruption of the Ink4aARF
locus alone should disable the two tumor suppressors most
commonly mutated in human cancers. Indeed, mice lacking
functional p16Ink4a andor p19ARF develop normally but are
highly susceptible to tumor formation (11, 24). However, the
frequency of mutations at the Ink4a andor ARF locus is not as
high as the frequency of p53 mutations in human cancers,
suggesting that loss of p53 confers a greater advantage during
tumor progression.
The data discussed above raise the possibility that p19ARF
functions in some, but not all, of the p53-responsive pathways
such that loss of p19ARF may not be functionally equivalent to
loss of p53. It has been suggested that oncogene-mediated
activation of p53 requires p19ARF, and occurs in the absence of
DNA damage (25, 26). By contrast, damage induced by ionizing
radiation (IR) involves p53 activation through N-terminal phos-
phorylations and other modifications, but is reported to be
independent of p19ARF (20, 24, 27). However, E1A-expressing
p19ARF
/
mouse embryo fibroblasts (MEFs) were nearly as
radioresistant after exposure to IR as E1A-expressing p53
/
MEFs (25), and p19ARF deficiency prevents induction of pre-
mature senescence in ATM
/
MEFs (28). One interpretation of
the data is that loss of p53 or p19ARF reduces the E1A-induced
radiosensitivity because of the requirement for these proteins in
arrest responses triggered by E1A. Alternatively, the data raise
the possibility that the observed radioresistance and rescue of
premature senescence is due to the requirement for p53, and
possibly p19ARF, in the arrest induced by DNA damage. To
distinguish between these possibilities, we assessed the role of
p19ARF in the p53-mediated arrest responses induced by IR as
well as other stresses that do not induce DNA damage.
Materials and Methods
Cell Culture. Pools of wild-type, p53
/
, p19ARF
/
, p53
/
, and
p16
/
p19ARF
/
double-knockout MEFs were maintained in
DMEM (Cellgro, Waukesha, WI) containing 10% dialyzed fetal
bovine serum, kanamycin sulfate (GIBCO), L-glutamine
(GIBCO), penicillinstreptomycin (GIBCO), and nonessential
amino acids (GIBCO). p19ARF
/
MEFs (passage 6) were
generously provided by Charles Sherr and Martine Roussel;
p16
/
p19ARF
/
double-knockout MEFs (passage 3) were
kindly provided by Manuel Serrano. Cells were maintained at
37°C in an atmosphere containing 7% CO
2
.
Cell Cycle Analysis. Cell cultures were split 1:3 or 1:4 into media
containing 0.05
gml nocodazole, 5 nM actinomycin D, or 100
MN-phosphonacetyl-L-aspartate (PALA), or were exposed to
radiation by using a Gammabeam 150-C irradiator with a
60
Co
source at a dose rate of 2 Gymin. At the indicated time points,
the fibroblasts were pulse labeled with 10
M BrdUrd for 1 h
before fixation in 70% ethanol. Analysis of BrdUrd-labeled cells
was performed as described previously (3). For DNA content
analysis, cells were stained with propidium iodide only. The
samples were analyzed on a Becton Dickinson FACScan and
Abbreviations: IR, ionizing radiation; MEFs, mouse embryo fibroblasts; PALA, N-
phosphonacetyl-L-aspartate.
‡To whom reprint requests should be addressed at: The Salk Institute, Gene Expression
Laboratory, 10010 North Torrey Pines Road, La Jolla, CA 92037. E-mail: wahl@salk.edu.
The publication costs of this article were defrayed in part by page charge payment. This
article must therefore be hereby marked advertisement in accordance with 18 U.S.C.
§1734 solely to indicate this fact.
Article published online before print: Proc. Natl. Acad. Sci. USA, 10.1073pnas.050560997.
Article and publication date are at www.pnas.orgcgidoi10.1073pnas.050560997
3266–3271
PNAS
March 28, 2000
vol. 97
no. 7
quantitated by gating events on dot plots by using CELLQUEST
software. All experiments were repeated a minimum of three
times, 10,000 events were analyzed per sample, and data from
representative experiments are shown.
Western Blot Analysis. Treated cells were lysed in RIPA buffer (150
mM NaCl1% Nonidet P-400.5% sodium deoxycholate0.1%
SDS50 mM TrisHCl, pH 8.0) supplemented with fresh 1 mM
PMSF, 1 mM sodium vanadate, 1 mgml leupeptin, 1 mgml
aprotinin, 1 mgml pepstatin, 1 mgml 1,10-phenanthrolineHCl,
and 160
gml benzamidineHCl. Lysates were sonicated and
protein concentrations were quantitated with a Bio-Rad protein
assay (modified Lowry assay). Fifty or 100
g of protein was
resolved by electrophoresis on a 10% (p21 and p53) or 12%
(p19ARF) polyacrylamide gel and transferred to a nitrocellulose
membrane (Schleicher and Schuell). p53 was detected by using a
mixture of monoclonal antibodies 421 at 1:50 (Ab-1, Oncogene
Science) and 240 at 1:100 (Ab-3, Oncogene Science). p21 was
detected by using the polyclonal antibody C-19 (Santa Cruz Bio-
technology) diluted 1:100; p19ARF was detected by using a poly-
clonal antibody (200-106; Novus Biologicals, Littleton, CO) diluted
1:200; and protein loading was measured with an anti-actin poly-
clonal antibody (A2066, Sigma) diluted 1:50. Membranes were
incubated with secondar y antibodies conjugated to horseradish
peroxidase, and proteins were detected by using ECL chemilumi-
nescence reagents according to the manufacturer’s protocol (New
England Nuclear). Bands were quantitated by computer scan
densitometry (NIH IMAGE), experiments were repeated at least
three times, and representative blots are shown. In all cases, actin
was used as an internal control to enable quantification of relative
band intensities.
Results
p19ARF
/
MEFs Become Arrested After Extended Exposure to PALA.
p19ARF has been proposed to mediate p53 activation in response
to oncogene overexpression in a manner that is independent of
DNA damage (25). Depletion of ribonucleotide pools in cells also
activates p53 without producing DNA damage in human and mouse
fibroblasts (3, 29). Therefore, we determined whether p19ARF is
involved in the activation of p53 after exposure to the UMP
synthesis inhibitor PALA (30). As previous reports showed no
significant differences in the results obtained after ribonucleotide
depletion of asynchronous or G
0
G
1
synchronized MEFs, the
PALA studies reported here were performed on exponentially
growing populations of MEFs (31).
Fig. 1 shows that PALA induces a cell cycle arrest in MEFs
regardless of their p19ARF status. Approximately 20–25% of
untreated wild-type, p53
/
, p53
/
, or p19ARF
/
MEFs were
in S phase (BrdUrd positive) in exponentially growing popula-
tions (Fig. 1 A–D). The fraction of BrdUrd-positive p53
/
,
p53
/
, and p19ARF
/
MEFs increased after 24 h in PALA,
presumably because of depletion of ribonucleotide and deoxy-
ribonucleotide pools, which slows the progression of replication
forks (32, 33). After 48 h of PALA treatment, the fraction of
S-phase cells decreased in wild-type, p53
/
, and p19ARF
/
populations (Fig. 1 A, B, and D), whereas the percentage of
BrdUrd-positive p53
/
MEFs remained elevated, as observed
previously (Fig. 1C) (31). By 72 h, significantly fewer wild-type
and p19ARF
/
MEFs were in S phase compared with untreated
populations or PALA-treated p53
/
MEFs. Although the per-
centage of p53
/
S-phase cells decreased between 48 and 72 h
of PALA treatment, the fraction of cycling cells was as high as
in the untreated population, and significantly higher than
PALA-treated wild-type, p53
/
, or p19ARF
/
populations.
There appears to be a slight increase in the population of cells
in S phase that are BrdUrd negative (below the diagonal line) in
wild-type, p53
/
, and p19ARF
/
MEFs; however, this popu-
lation accounts for only 5% of the total population. As reported
previously, a majority of the PA LA-treated MEFs that were in
S phase at the beginning of the experiment appear to accumulate
in G
2
M, as this population shows a 10% increase relative to the
G
2
M population in untreated cells (31). p16
/
p19ARF
/
double-knockout MEFs also became arrested after 72 h of
PALA treatment, suggesting that the loss of pRb regulation does
not abrogate the arrest response (data not shown). These data
clearly show that the p53-mediated arrest induced by PALA is
intact in MEFs lacking p19ARF function.
The p53-dependent cell cycle arrest induced by PALA re-
quires activation of p53 target genes such as p21 (3, 31). An
analysis of the p21 protein levels in p19ARF
/
MEFs revealed
that p21 was up-regulated 48 h after exposure to PALA com-
pared with untreated cells (Fig. 1E). The fold induction of p21
(normalized with actin protein loading) is approximately 2-fold
in p19ARF
/
and p53
/
MEFs after 48 h of PALA treatment.
This finding is consistent with the previous observation that
p19ARF function is not required for the arrest induced by
ribonucleotide depletion. Induction of p21 was not observed in
p53
/
MEFs treated with PALA for 48 h, which supports
previous reports that p21 expression is activated by p53 in
response to ribonucleotide depletion (data not shown) (3).
To assess whether ribonucleotide depletion was unique in
eliciting a p53-dependent arrest that is independent of p19ARF,
we determined whether the arrest induced by the intercalating
agent actinomycin D required p19ARF function. Recent data
indicate that low doses (20 nM) of actinomycin D did not induce
DNA double-strand breaks but were sufficient to inhibit tran-
scription and activate p53 in normal human fibroblasts, whereas
doses of 200 nM induced DNA damage (34, 35). We found that
p53
/
and p19ARF
/
MEFs become arrested after 24 h of
Fig. 1. p19ARF-null MEFs become arrested during ribonucleotide depletion.
Asynchronous populations of wild-type (WT), p53/, p53/, and p19ARF/
MEFs were untreated or treated with 100
M PALA for 24, 48, or 72 h. Cells
were pulse labeled with BrdUrd and analyzed by flow cytometry. Cells that
incorporated BrdUrd are shown above the diagonal line. The percentages of
BrdUrd-positive cells are shown in the upper right corner of each plot. Rep-
resentative dot plots for untreated or PALA-treated MEFs are shown in A(WT),
B(p53/), C(p53/), and D(p19ARF/). (E) Western blot analysis of p21 and
actin after 48-h PALA treatment of p53/and p19ARF/MEFs. Relative band
intensities were quantified in all Western analyses with actin as an internal
control.
Khan et al. PNAS
March 28, 2000
vol. 97
no. 7
3267
CELL BIOLOGY
exposure to 5 nM actinomycin D, whereas most of the p53
/
MEFs continue cycling (Table 1). Moreover, 5 nM actinomycin
D treatment induced a reversible arrest in wild-type and
p19ARF
/
MEFs; that is, once the drug was removed, the cells
resumed cycling as efficiently as untreated populations (data not
shown). This result provides further evidence that the arrest
induced by 5 nM actinomycin D is unlikely to result from residual
DNA damage. Therefore, two treatments that reduce RNA
synthesis in the absence of detectable DNA damage activate a
cell cycle arrest requiring p53, but not p19ARF.
p19ARF
/
MEFs Show an Increase in Polyploidy After Nocodazole
Treatment. We and others showed previously that the microtu-
bule-depolymerizing agents nocodazole and Colcemid provoke
an arrest in a 4NG
1
-like state in cells with an intact p53–pRb
pathway (5, 36–39). By contrast, mutants with defects in this
pathway undergo rereplication during continuous exposure to
these drugs. As with PALA, the arrest induced by these agents
is largely reversible and is not associated with detectable DNA
damage (39). It was shown previously that p16
/
MEFs also
undergo rereplication after exposure to such drugs (39). How-
ever, these MEFs have nonfunctional p16Ink4a and p19ARF
because of targeted disruption of exons shared by the two
proteins (11). The dual inactivation of the proteins in this mutant
raised the important question of whether this damage-
independent stress required p19ARF to induce a cell cycle
arrest, or whether p16Ink4a deficiency alone allowed rereplica-
tion to occur.
We investigated this possibility by determining whether no-
codazole induced rereplication in MEFs solely lacking p19ARF.
Asynchronous populations of p53
/
, p53
/
, p19ARF
/
, and
p16
/
p19ARF
/
MEFs were treated with 0.05
gml nocoda-
zole for 48 h, after which polyploidy was measured by DNA
content analysis (see Materials and Methods). After 48 h in
nocodazole, 8.7% of p53
/
MEFs were polyploid compared
with 60.3% in p53
/
MEFs, 18.3% in p19ARF
/
MEFs, and
37.9% in p16
/
p19ARF
/
double-knockout MEFs (Fig. 2 and
Table 1). In three independent experiments, the percentage of
polyploid cells in nocodazole-treated p19ARF
/
populations
consistently doubled compared with the untreated control (Fig.
2Cand Table 1). Furthermore, the percentage of cells with
greater than 4NDNA content was consistently higher in the
p16
/
p19ARF
/
double-knockout MEFs (Fig. 2C), suggesting
that the loss of both tumor suppressors is additive in this
response. While it remains to be determined whether the loss of
p16Ink4a alone induces polyploidy after microtubule disruption,
we predict this is likely, as pRb deficiency produces substantial
polyploidy (50%) under the same conditions (36).
The polyploid population was significantly higher in p53
/
MEFs compared with either p19ARF
/
or p16
/
p19ARF
/
MEFs, indicating that the loss of p19ARF is not equivalent to
loss of p53 in the nocodazole-induced arrest response. Similarly,
a reduced level of polyploidy was observed in p21
/
MEFs
compared with MEFs lacking p53 (39). This observation may be
due to a p53-dependent activation of additional pathways that
may compensate for the loss of p16, p19ARF, or p21, such that
pRb is eventually restored to the antiproliferative form to
promote cell cycle arrest.
p19ARF
/
MEFs Exhibit a Defective DNA-Damage Response. The
p53-dependent cell cycle arrest induced by ionizing radiation
requires prolonged induction of p21 (7, 40). Previous reports
indicated that p53 and p21 induction occurs in p19ARF
/
Table 1. ARF dependence of p53-mediated stress responses
MEFs
%S
/% S untreated % BrdUrd-positive Polyploidy
T48 Noc5 Gy 6 Gy 7 Gy 20 Gy Untreated T72 PALA T24 Act-D
p53
/
39.6 0.5 36.9 1.8 34.4 5.2 25.9 4.5 22.8 2.4 12.0 3.1 5.1 0.1 6.5 3.4
p53
/
101.0 1.4 100.8 3.2 105.2 8.3 83.5 0.7 24.4 3.8 27.7 3.3 20.8 5.4 60.1 2.2
p19ARF
/
75.0 9.9 85.0 2.8 81.5 0.7 64.5 0.7 19.7 1.2 6.9 2.6 4.6 0.5 26.1 4.9
Asynchronous populations of p53/, p53/, and p19ARF/MEFs were irradiated with 5, 6, 7, or 20 Gy, pulse-labeled with BrdUrd, and fixed 24 h later. The
values represent the percentage of BrdUrd-positive cells in
-irradiated cultures (% S
) relative to untreated. Cells were also untreated or treated with 100
M
PALA for 72 h or 5 nM actinomycin D for 24 h, and the percentage of BrdUrd-positive cells is shown. The polyploid fraction was assessed by treating cells with
0.05
g/ml nocodazole for 48 h and quantitating the percentage of cells with 4NDNA content. The values represent the mean of at least two independent
experiments and standard deviations are shown.
Fig. 2. Increased polyploidy in p19ARF-null fibroblasts after nocodazole
treatment. Asynchronous cultures of p53/, p53/, p19ARF/, and
p16/p19ARF/MEFs were treated with or without 0.05
gml nocodazole
for 48 h, fixed, and stained with propidium iodide for FACS analysis. Values in
the upper right corner of each plot represent the percentage of cells with 4N
DNA content. Histogram plots of untreated or nocodazole-treated MEFs are
shown in A(p53/), B(p53/), C(p19ARF/), and D(p16/p19ARF/).
3268
www.pnas.org Khan et al.
MEFs 4 h after IR, but protein analyses at later time points were
not carried out and the stringency of the cell cycle arrest was not
shown (24). The increased viability of E1A-expressing
p19ARF
/
MEFs after IR invites the speculation that these
cells are compromised in their ability to undergo cell cycle arrest
or apoptosis in response to IR (25). Hence, we assessed whether
the p19ARF
/
MEFs exhibit an altered cell cycle response
after IR.
The cell cycle responses of asynchronous p53
/
, p53
/
, and
p19ARF
/
MEFs are shown in Fig. 3 and Table 1. In the
experiment shown, the percentage of BrdUrd-positive cells in
irradiated populations relative to untreated cells was approxi-
mately 37% for p53
/
MEFs, 103% in p53
/
MEFs, and 87%
in p19ARF
/
MEFs 24 h after exposure to 6 Gy of
-radiation
(Fig. 3 A, B, and C, respectively). This observation suggests that
loss of p19ARF results in a defective DNA-damage arrest
pathway even though these cells express wild-type p53 and p21.
The experiments were repeated a minimum of five times at
passages 8–10 with various radiation dosages (Table 1). Similar
results were observed in irradiated populations of p19ARF
/
MEFs that were synchronized in G
0
G
1
by serum deprivation
(data not shown). We also found that the percentage of irradi-
ated cells in S phase relative to untreated was 70% in
p16
/
p19ARF
/
MEFs after 24 h (data not shown), suggesting
that the loss of pRb regulation does not augment the defective
DNA-damage arrest created by loss of p19ARF alone.
The data show that p19ARF
/
MEFs continue cycling after
DNA damage, although not to the same extent as p53
/
MEFs.
Interestingly, the lack of cell cycle arrest in p19ARF
/
MEFs is
similar to that reported for p21
/
MEFs, in which 80% of the
population continued to cycle 24 h after DNA damage (31). This
observation implies that factors in addition to p19ARF and p21
contribute to the p53-dependent DNA-damage-induced arrest in
wild-type cells.
As the response to ionizing radiation involves rapid p53-
dependent induction of p21, it is conceivable that p53 activation
and p21 induction could be compromised in p19ARF-null cells.
We investigated this possibility by measuring p53 and p21 protein
levels as a function of time after IR. As reported previously, p53
induction is biphasic in wild-type cells, consisting of a rapid
increase, followed by a decrease, and then a second increase to
a level that remained above the uninduced control up to 96 h
after IR (7, 8). The p53 levels in p19ARF
/
MEFs showed a
similar initial increase in abundance 2–10 h after irradiation, but
then diminished to undetectable levels by 24 h and remained
undetectable 48 h after IR (Fig. 4A). In contrast, the p53 levels
were elevated between 1.9- and 3.5-fold at each of the time points
analyzed after IR in wild-type MEFs (Fig. 4A). A possible
explanation for the apparent p53 instability is that the absence
of p19ARF allows a higher proportion of p53 to be bound by
mdm2, which increases p53 turnover.
Consistent with the p53 expression levels, p21 protein levels
started to increase4hafterirradiation in wild-type MEFs and
remained elevated up to 48 h (Fig. 4A). While the p21 levels are
detectable in p19ARF
/
MEFs at all times tested, the induced
levels are lower than observed in wild-type MEFs. Wild-type
MEFs show a 1.9- to 5.0-fold increase in p21 between 2 and 48 h
after exposure to IR, whereas the maximum induction of p21 in
p19ARF
/
MEFs during the same time course is 2-fold.
Furthermore, 2448 h after IR the levels of p21 are not
increased relative to the untreated control in p19ARF
/
MEFs
(Fig. 4A). p53
/
MEFs showed no detectable levels of p21 at
several time points after exposure to IR; hence the induction of
p21 in this stress response is largely, if not entirely, dependent on
transactivation by p53 (data not shown) (40).
We next determined whether p19ARF is induced by DNA
damage by measuring the p19ARF protein levels at several
intervals after IR. As expected, p19ARF was not present in the
p19ARF-null MEFs, and it is present at high levels in p53
/
MEFs (Fig. 4B) (25, 26). p19ARF levels increased significantly
(2- to 8-fold induction) 2–10 h after
-irradiation in wild-type
MEFs (Fig. 4B). Although the level of p19ARF appears to
Fig. 3. Defective DNA-damage response in p19ARF-null MEFs. Asynchronous
cultures of p53/, p53/, and p19ARF/MEFs were exposed to 6 Gy of
-radiation. After 24 h, cells were pulse labeled with BrdUrd and analyzed by
FACS. Cells above the diagonal line are BrdUrd positive. The values shown in
the upper right represent the percentage of irradiated cells in S phase relative
to untreated (% S irradiated% S untreated). Representative plots are shown
for p53/(A), p53/(B), and p19ARF/(C).
Fig. 4. Involvement of p19ARF in the sustained induction of p21 after IR.
Asynchronous populations of wild-type (WT) and p19ARF/MEFs were
treated with 6 Gy of
-radiation and harvested after 2, 4, 6, 8, 10, 24, or 48 h.
For p53 and p21 analysis, 50
g of protein was resolved on an SDS10%
polyacrylamide gel. For analysis of p19ARF, 100
g of protein was resolved on
a 12% polyacrylamide gel. (A) p53, p21, and actin protein levels in WT and
p19ARF/MEFs 2 h (T2), 4 h (T4), 6 h (T6), 8 h (T8), 10 h (T10), 24 h (T24), and
48 h (T48) after IR compared with untreated (Unt). (B) p19ARF and actin
protein levels in WT MEFs at the indicated times after IR compared with
untreated (Unt). p53/and p19ARF/MEFs are also shown as positive and
negative controls for p19ARF protein detection.
Khan et al. PNAS
March 28, 2000
vol. 97
no. 7
3269
CELL BIOLOGY
decrease 2448 h after IR, the protein levels were still 1.5- to
1.7-fold elevated relative to untreated controls at the latest time
points measured (Fig. 4B). These results show that p19ARF
protein levels increase after exposure to
-radiation.
Discussion
The studies reported here reveal that p19ARF contributes to the
ability of MEFs to undergo p53-dependent cell cycle arrest
induced by IR and by microtubule disruption. The lack of a
damage-induced arrest in p19ARF
/
MEFs correlates with
decreased basal p53 levels and an inability to sustain the
induction of p53. This, in turn, appears to affect the induction of
the CDK inhibitor p21, a key downstream target gene in the
damage response. By contrast, the arrest induced by ribonucle-
otide depletion and actinomycin D treatment is intact in
p19ARF-deficient fibroblasts. The data are consistent with the
model that mdm2 is a key negative regulator of p53 function and
that p19ARF contributes to a subset of p53-dependent arrest
pathways to counteract the inhibitory effects of mdm2.
The requirement for p19ARF in the p53-dependent arrest
induced by DNA damage is consistent with current models that
p19ARF is a regulator of mdm2–p53 interactions (12–15, 22).
p19ARF appears to reduce the ability of mdm2 to interact with
p53 by changing mdm2 subcellular localization (22, 23). Our
observation that p53 levels are reduced in p19ARF-null MEFs
is consistent with the increased p53–mdm2 associations that
should occur in cells lacking p19ARF. It is conceivable that the
reduced p53 levels may contribute to the defective arrest induced
by IR in the p19ARF-null MEFs. However, the arrest induced
by IR appears to be normal in cell lines expressing low amounts
of a transfected wild-type p53 gene and high quantities of
dominant-negative p53 encoded by the mutated endogenous
alleles (e.g., see ref. 41). This observation leads us to favor the
interpretation that it is the altered p53 and p21 induction kinetics
in the p19ARF-null MEFs that prevents them from mounting a
durable arrest response. This view is compatible with the ob-
servation of long-term induction of p53 and p21 in irradiated
fibroblasts that undergo a senescent-like arrest (7, 42). The
difference in the response kinetics at later times suggests that
p19ARF andor its induction by DNA damage may contribute
to the sustained activation of p53 and its target genes. While the
mechanism for p19ARF induction remains to be elucidated, it is
tempting to speculate that the transcriptional regulator E2F-1
may be involved, as p19ARF expression can be transactivated by
E2F-1 (43), and recent studies show that IR increases E2F-1
abundance by a posttranslational mechanism (44). We also note
that the rapid induction of p53 and p21 in p19ARF-null MEFs
appears to be equivalent to that of wild-type MEFs. We suggest
that kinases such as ATM, ATR, and others that modify the N
terminus of p53 to limit interactions with mdm2 may be involved
in establishing the early stages of the damage response (20, 21,
45–47).
The involvement of p19ARF in the DNA-damage-inducible
arrest is compatible with studies analyzing the cell cycle re-
sponses and radiosensitivity of MEFs with various genetic
deficiencies in the p53 pathway. p21-null MEFs continue to cycle
to approximately the same extent as p19ARF-null MEFs after
IR, and both are more radioresistant than are wild-type MEFs
(31, 48). By contrast, ATM-deficient MEFs are more radiosen-
sitive than wild-type MEFs, and they senesce prematurely (49,
50). These phenotypes are consistent with the generation and
accumulation of DNA damage when cells cycle in the absence of
the ATM kinase (51). Importantly, premature senescence is
rescued in MEFs doubly deficient in ATM and p19ARF, p53, or
p21 (28, 52, 53). Thus, p19ARF deficiency results in defects
similar to those observed in cells lacking two key members of the
damage-response pathway, p53 and p21. However, it is impor-
tant to note that neither p19ARF, p21, nor p53 deficiency
rescues the radiosensitivity of ATM-deficient mice (28, 53, 54).
This observation is most likely explained by the fact that
intestinal epithelia and other cell types in ATM-deficient mice
appear to be exquisitely sensitive to damage-induced apoptosis
that occurs independent of p53 function (54). Hence, the survival
of ATM-deficient mice after radiation is influenced by factors
outside of the p53 signaling pathway, whereas the isolated cells
of specific tissues appear to be profoundly dependent on p53,
p21, and p19ARF.
Previous studies showed that p19ARF cooperates with p53 to
mediate cell cycle arrest or programmed cell death in response to
activated oncogenes such as E1A, Myc, Ras, and E2F-1 (25, 26, 43,
55). E1A overexpression did not lead to phosphorylation of p53 at
serine-15, which is one residue that is modified after DNA damage
(20, 25, 27). This finding led to the proposal that overexpression of
E1A activates p53 independent of DNA damage. However, phos-
phorylation at serine-15 is just one of many p53 modifications that
occur in response to DNA damage; recent data indicate the
importance of serine-20 phosphorylation in the DNA-damage
response (21, 47). Our data linking p19ARF to the damage-induced
arrest raise the possibility that activated oncogenes may induce
DNA damage, which could then lead to senescence in fibroblasts
with an intact p53 pathway. Indeed, this model would be c ompatible
with numerous reports showing that activated c-myc, ras, or mos can
induce chromosome breakage in cycling cells (refs. 56 –58; O. Vafa
and G.M.W., unpublished results).
Our data demonstrate that stresses such as ribonucleotide de-
pletion or RNA synthesis inhibition clearly function independently
of p19ARF. PALA treatment is associated with increased p53
abundance and activation of p21 transcription, but it remains to be
determined whether induction of mdm2 is involved (3). It is possible
that p19ARF is not required for the arrest induced by stresses that
reduce transcription because such conditions may preferentially
reduce the levels of the unstable mdm2 transcript. Consistent with
this proposal, previous studies showed that low doses of actinomy-
cin D and UV doses that reduce transcription also reduce the
abundance of mdm2 mRNA (44, 59). Similarly, we found reduced
levels of mdm2 protein in PALA-treated cells, but not in those
treated with nocodazole (S.H.K. and G.M.W., unpublished data).
Hence, inhibition of mdm2 expression by agents such as actinomy-
cin D, and possibly PALA, serves as a p19ARF-independent
mechanism to promote the stability of p53.
These and other data also emphasize the complexity of the
biochemical and cell cycle responses to diverse stresses that
activate the p53 pathway. The different responses of cells with
defects in this pathway indicate that additional genes remain to
be elucidated. Thus, as p21
/
and p53
/
MEFs exhibit com-
parable defects in PALA-induced arrest, we surmise that p21 is
a key downstream modulator of this response. By contrast, since
p21
/
and p19ARF
/
MEFs retain a partial ability to undergo
arrest after IR and show less rereplication after nocodazole
treatment than do p53
/
MEFs (Figs. 1, 2, and 3; refs. 31 and
39), we infer that additional upstream effectors and downstream
targets remain to be defined. Additional studies to elucidate the
mechanisms by which p19ARF is activated and the kinetics of
p53, mdm2, and p19ARF associations will enhance our under-
standing of the cellular response to DNA damage and damage-
independent stresses.
We thank Drs. Martine Roussel and Charles Sherr for generously
providing p19ARF
/
MEFs and Dr. Manuel Serrano for kindly pro-
viding p16
/
p19ARF
/
MEFs; Dr. Gretchen Jimenez and Ms. Jayne
Stommel for insightful discussions and critical review of this manuscript;
and Ms. Michelle Beeche for isolating the wild-type and p53
/
MEFs.
These studies were supported by grants to G.M.W. from the National
Institutes of Health (CA61449) and the G. Harold and Leila Y. Mathers
Charitable Foundation and by a National Institutes of Health Predoc-
toral Training Grant to S.H.K. (1T32CA64041).
3270
www.pnas.org Khan et al.
1. Huang, L. C., Clarkin, K. C. & Wahl, G. M. (1996) Proc. Natl. Acad. Sci. USA
93, 4827–4832.
2. Hermeking, H. & Eick, D. (1994) Science 265, 2091–2093.
3. Linke, S. P., Clarkin, K. C., Di Leonardo, A., Tsou, A. & Wahl, G. M. (1996)
Genes Dev. 10, 934–947.
4. Maltzman, W. & Czyzyk, L. (1984) Mol. Cell. Biol. 4, 1689–1694.
5. Cross, S. M., Sanchez, C. A., Morgan, C. A., Schimke, M. K., Ramel, S., Idzerda,
R. L., Raskind, W. H. & Reid, B. J. (1995) Science 267, 1353–1356.
6. Graeber, T. G., Peterson, J. F., Tsai, M., Monica, K., Fornace, A. J., Jr., &
Giaccia, A. J. (1994) Mol. Cell. Biol. 14, 6264– 6277.
7. Di Leonardo, A., Linke, S. P., Clarkin, K. & Wahl, G. M. (1994) Genes Dev.
8, 2540–2551.
8. Kastan, M. B., Onyekwere, O., Sidransky, D., Vogelstein, B. & Craig, R. W.
(1991) Cancer Res. 51, 6304– 6311.
9. Paulson, T. G., Almasan, A., Brody, L. L. & Wahl, G. M. (1998) Mol. Cell. Biol.
18, 3089–3100.
10. Chernova, O. B., Chernov, M. V., Agarwal, M. L., Taylor, W. R. & Stark, G. R.
(1995) Trends Biochem. Sci. 20, 431–434.
11. Serrano, M., Lee, H., Chin, L., Cordon-Cardo, C., Beach, D. & DePinho, R. A.
(1996) Cell 85, 27–37.
12. Stott, F. J., Bates, S., James, M. C., McConnell, B. B., Starborg, M., Brookes,
S., Palmero, I., Ryan, K., Hara, E., Vousden, K. H. & Peters, G. (1998) EMBO
J. 17, 5001–5014.
13. Zhang, Y., Xiong, Y. & Yarbrough, W. G. (1998) Cell 92, 725–734.
14. Pomerantz, J., Schreiber-Agus, N., Liegeois, N. J., Silverman, A., Alland, L.,
Chin, L., Potes, J., Chen, K., Orlow, I., Lee, H. W., et al. (1998) Cell 92, 713–723.
15. Kamijo, T., Weber, J. D., Zambetti, G., Zindy, F., Roussel, M. F. & Sherr, C. J.
(1998) Proc. Natl. Acad. Sci. USA 95, 8292–8297.
16. Barak, Y., Juven, T., Haffner, R. & Oren, M. (1993) EMBO J. 12, 461–468.
17. Kubbutat, M. H., Jones, S. N. & Vousden, K. H. (1997) Nature (London) 387,
299–303.
18. Haupt, Y., Maya, R., Kazaz, A. & Oren, M. (1997) Nature (London) 387,
296–299.
19. Thut, C. J., Goodrich, J. A. & Tjian, R. (1997) Genes Dev. 11, 1974 –1986.
20. Shieh, S. Y., Ikeda, M., Taya, Y. & Prives, C. (1997) Cell 91, 325–334.
21. Unger, T., Juven-Gershon, T., Moallem, E., Berger, M., Vogt Sionov, R.,
Lozano, G., Oren, M. & Haupt, Y. (1999) EMBO J. 18, 1805–1814.
22. Tao, W. & Lev ine, A. J. (1999) Proc. Natl. Acad. Sci . USA 96, 6937–6941.
23. Weber, J. D., Taylor, L. J., Roussel, M. F., Sherr, C. J. & Bar-Sagi, D. (1999)
Nat. Cell Biol. 1, 20 –26.
24. Kamijo, T., Zindy, F., Roussel, M. F., Quelle, D. E., Downing, J. R., Ashmun,
R. A., Grosveld, G. & Sherr, C. J. (1997) Cell 91, 649– 659.
25. de Stanchina, E., McCurrach, M. E., Zindy, F., Shieh, S. Y., Ferbeyre, G.,
Samuelson, A. V., Prives, C., Roussel, M. F., Sherr, C. J. & Lowe, S. W. (1998)
Genes Dev. 12, 2434–2442.
26. Zindy, F., Eischen, C. M., Randle, D. H., Kamijo, T., Cleveland, J. L., Sherr,
C. J. & Roussel, M. F. (1998) Genes Dev. 12, 2424–2433.
27. Siliciano, J. D., Canman, C. E., Taya, Y., Sakaguchi, K., Appella, E. & Kastan,
M. B. (1997) Genes Dev. 11, 3471–3481.
28. Kamijo, T., van de Kamp, E., Chong, M. J., Zindy, F., Diehl, J. A., Sherr, C. J.
& McKinnon, P. J. (1999) Cancer Res. 59, 2464–2469.
29. Agarwal, M. L., Agarwal, A., Taylor, W. R., Chernova, O., Sharma, Y. & Stark,
G. R. (1998) Proc. Natl. Acad. Sci. USA 95, 14775–14780.
30. Collins, K. D. & Stark, G. R. (1971) J. Biol. Chem. 246, 6599–6605.
31. Deng, C., Zhang, P., Harper, J. W., Elledge, S. J. & Leder, P. (1995) Cell 82,
675–684.
32. Chernova, O. B., Chernov, M. V., Ishizaka, Y., Agarwal, M. L. & Stark, G. R.
(1998) Mol. Cell. Biol. 18, 536 –545.
33. Moyer, J. D., Smith, P. A., Levy, E. J. & Handschumacher, R. E. (1982) Cancer
Res. 42, 4525–4531.
34. Chang, D., Chen, F., Zhang, F., McKay, B. C. & Ljungman, M. (1999) Cell
Growth Differ. 10, 155–162.
35. Ljungman, M., Zhang, F., Chen, F., Rainbow, A. J. & McKay, B. C. (1999)
Oncogene 18, 583–592.
36. Di Leonardo, A., Khan, S. H., Linke, S. P., Greco, V., Seidita, G. & Wahl, G. M.
(1997) Cancer Res. 57, 1013–1019.
37. Minn, A. J., Boise, L. H. & Thompson, C. B. (1996) Genes Dev. 10, 2621–2631.
38. Lanni, J. S. & Jacks, T. (1998) Mol. Cell Biol. 18, 1055–1064.
39. Khan, S. H. & Wahl, G. M. (1998) Cancer Res. 58, 396– 401.
40. el-Deir y, W. S., Harper, J. W., O’Connor, P. M., Velculescu, V. E., Canman,
C. E., Jackman, J., Pietenpol, J. A., Burrell, M., Hill, D. E., Wang, Y., et al.
(1994) Cancer Res. 54, 1169–1174.
41. Yin, Y., Tainsky, M. A., Bischoff, F. Z., Strong, L. C. & Wahl, G. M. (1992)
Cell 70, 937–948.
42. Robles, S. J. & Adami, G. R. (1998) Oncogene 16, 1113–1123.
43. Bates, S., Phillips, A. C., Clark, P. A., Stott, F., Peters, G., Ludwig, R. L. &
Vousden, K. H. (1998) Nature (London) 395, 124–125.
44. Blattner, C., Sparks, A. & Lane, D. (1999) Mol. Cell. Biol. 19, 3704 –3713.
45. Canman, C. E., Lim, D. S., Cimprich, K. A., Taya, Y., Tamai, K., Sak aguchi,
K., Appella, E., Kastan, M. B. & Siliciano, J. D. (1998) Science 281, 1677–1679.
46. Banin, S., Moyal, L., Shieh, S., Taya, Y., Anderson, C. W., Chessa, L.,
Smorodinsky, N. I., Prives, C., Reiss, Y., Shiloh, Y. & Ziv, Y. (1998) Science 281,
1674–1677.
47. Shieh, S. Y., Taya, Y. & Prives, C. (1999) EMBO J. 18, 1815–1823.
48. Brugarolas, J., Chandrasekaran, C., Gordon, J. I., Beach, D., Jacks, T. &
Hannon, G. J. (1995) Nature (London) 377, 552–557.
49. Barlow, C., Hirotsune, S., Paylor, R., Liyanage, M., Eckhaus, M., Collins, F.,
Shiloh, Y., Crawley, J. N., Ried, T., Tagle, D. & Wynshaw-Boris, A. (1996) Cell
86, 159–171.
50. Xu, Y., Ashley, T., Brainerd, E. E., Bronson, R. T., Meyn, M. S. & Baltimore,
D. (1996) Genes Dev. 10, 2411–2422.
51. Elson, A., Wang, Y., Daugherty, C. J., Morton, C. C., Zhou, F., Campos-Torres,
J. & Leder, P. (1996) Proc. Natl. Acad. Sci. USA 93, 13084–13089.
52. Westphal, C. H., Schmaltz, C., Rowan, S., Elson, A., Fisher, D. E. & Leder, P.
(1997) Cancer Res. 57, 1664–1667.
53. Wang, Y. A., Elson, A. & Leder, P. (1997) Proc. Natl. Acad. Sci . USA 94,
14590–14595.
54. Westphal, C. H., Hoyes, K. P., Canman, C. E., Huang, X., Kastan, M. B.,
Hendry, J. H. & Leder, P. (1998) Cancer Res. 58, 5637–5639.
55. Palmero, I., Pantoja, C. & Serrano, M. (1998) Nature (London) 395, 125–126.
56. Felsher, D. W. & Bishop, J. M. (1999) Proc. Natl. Acad. Sci. USA 96, 3940 –3944.
57. Denko, N., Stringer, J., Wani, M. & Stambrook, P. (1995) Somatic Cell Mol.
Genet. 21, 241–253.
58. Fukasawa, K. & Vande Woude, G. F. (1997) Mol. Cell. Biol. 17, 506 –518.
59. Wu, L. & Lev ine, A. J. (1997) Mol. Med. 3, 441– 451.
Khan et al. PNAS
March 28, 2000
vol. 97
no. 7
3271
CELL BIOLOGY
... The primary function of ARF appears to be in mediating p53-dependent oncogenic checkpoints, as its involvement in other p53-dependent checkpoints such as the response to UV or IR is debatable (Kamijo et al., 1997), (Khan et al., 2000). ...
... Active p53 instigates transcription of genes involved in mediating cell cycle arrest, primarily p21^'^^ and thus maintains the arrest initiated by Cdc25a and cyclin D l degradation. The finding that E2F1 and ARF can also be induced in response to ATM activation could indicate that this pathway may have a role in amplifying the p53 response , (Khan et al., 2000). A delayed induction of pioccurring after the p53 damage response, has also been observed (Pavey et al., 1999), (Robles and Adami, 1998). ...
... ; Rb-/-MEFs (Camero et al., 2000). The most commonly held view is that DNA damage response checkpoints in MEFs do not require p i9^^^ function and thus differ from culture-induced senescence in this respect (Kamijo et al., 1997) but see also (Khan et al., 2000). However, other findings suggest similarities between DNA damage responses and senescence of cultured MEFs: loss of proteins required for DNA repair (e.g.ATM) accelerate senescence presumably due to increased levels of DNA damage. ...
Thesis
Replicative limits have been thought to constrain the extended proliferation of primary cells. After dividing a set number of times, cells irreversibly withdraw from the cell cycle and adopt a characteristic phenotype in a process defined as cellular senescence. The certainty that all cells will ultimately senescence has resulted in a finite replicative capacity being used as one of the defining features of primary cells. In this thesis I show that primary rat Schwann cells can proliferate indefinitely in culture without acquiring immortalizing mutations. These results demonstrate that senescence can no longer be considered an unavoidable barrier to primary cell culture. The only known mechanism of limiting replicative lifespan is telomere shortening, which occurs upon division of telomerase negative cells. Other uncharacterised cell-intrinsic mechanisms were thought to regulate lifespan in cells that normally express telomerase yet senesce in culture. I show that senescence can be induced in primary Schwann cells, which express telomerase, by altering the conditions in which they are cultured. These results provide the first demonstration that extrinsic as well as intrinsic factors can regulate replicative lifespan and has fuelled speculation that senescence in telomerase positive cells may also be induced by external conditions rather than an intrinsic cell division timer. The cyclin-dependent kinase inhibitor p16INK4A progressively upregulated in most primary cell types as they divide in culture. This induction was thought to be a response to an intrinsic timer that operated to limit replicative capacity. I demonstrate that p16INKA induction in rat Schwann cells can be uncoupled from a senescent arrest. Moreover, the rate and level of p16INKA induction is dependent on the conditions in which the cells are cultured. These results demonstrate that p16INKA4A induction in primary Schwann cells is a response to extrinsic signals rather than an intrinsic cell division timer.
... As shown in Figure 3E, p53 mRNA levels were decreased by 8.92% in BSA-GNP-treated cells, yet increased by 1.21fold (p < 0.05) in CTAB-GNP treated cells. The increase of P53 can be contributed to microtubule disruption [25]. ...
... Increased levels of PCNA can cause cell cycle arrest in G 0 /G 1 through the inactivation of CDK4/6. Moreover, increased levels of p53 and PCNA can contribute to microtubule damage [25,29]. fold in response to mispositioned spindles. ...
... Increased levels of PCNA can cause cell cycle arrest in G0/G1 through the inactivation of CDK4/6. Moreover, increased levels of p53 and PCNA can contribute to microtubule damage [25,29]. ...
Article
Full-text available
Gold nanoparticles (GNPs) are usually wrapped with biocompatible polymers in biomedical field, however, the effect of biocompatible polymers of gold nanoparticles on cellular responses are still not fully understood. In this study, GNPs with/without polymer wrapping were used as model probes for the investigation of cytotoxicity and cell cycle progression. Our results show that the bovine serum albumin (BSA) coated GNPs (BSA-GNPs) had been transported into lysosomes after endocytosis. The lysosomal accumulation had then led to increased binding between kinesin 5 and microtubules, enhanced microtubule stabilization, and eventually induced G2/M arrest through the regulation of cadherin 1. In contrast, the bare GNPs experienced lysosomal escape, resulting in microtubule damage and G0/G1 arrest through the regulation of proliferating cell nuclear antigen. Overall, our findings showed that both naked and BSA wrapped gold nanoparticles had cytotoxicity, however, they affected cell proliferation via different pathways. This will greatly help us to regulate cell responses for different biomedical applications.
... Several studies in the last two decades have shown that depletion of intracellular pyrimidine nucleotides induces activation or upregulation of p53 (Lane, 1992;Huang, 1996;Wahl, 1997;Herrmann, 1997;Huang, 1999;Foxa, 1999;Abcouwer, 1999;Hail, 2012). While it is well established that p53 is activated by DNA damage, possibly triggered by instructions from a sensory "checkpoint" protein which may be p19ARF (Khan, 2000), evidence also shows that ribonucleotide depletion can activate p53 in the absence of DNA damage (Linke, 1996) and without p19ARF involvement (Ries, 2000). Moreover, evidence suggests that, in the case of ribonucleotide depletion, the trigger for p53 activation may be repression of Mdm2. ...
... Moreover, evidence suggests that, in the case of ribonucleotide depletion, the trigger for p53 activation may be repression of Mdm2. Significantly, Mdm2 was shown to be repressed by PALAinduced pyrimidine nucleotide depletion (Khan, 2000). VRAC-released pyrimidine nucleotides autocrinally activate ERK. ...
Article
Full-text available
Several types of cytoproliferative diseases, including cancers and autoimmune diseases, are associated with elevated cellular levels of pyrimidine nucleotides. This review assembles literature evidence supporting a thesis that elevated pyrimidine levels drive cytoproliferation largely through DNA-synthesis-independent mechanisms, which circumstance, it is proposed, inherently limits clinical efficacy of DNA-synthesis-targeting cytostatic drugs. Five hypothetical mechanisms,with supporting literature evidence, are presented: (1) Pyrimidine nucleotides released from ion channels auto/paracrinally and persistently activate the EGFR, PKC and ERK signaling pathway, upregulating cytokines and Cdk2; (2) One effect ofERK activation is reversal of allosteric control of de novo pyrimidine biosynthesis; thus pyrimidines’ auto/paracrine ERKactivation may reflect positive autoregulation; (3) Elevated intracellular pyrimidine nucleotides preferentially upregulate tumorigenic genes; (4) Elevated pyrimidines promote aberrant glycosylation of transcription factors; and (5) These diverse pyrimidine effects contribute, by different paths yet collectively, to repression of p53 or of its function. Also reviewed are alternate pyrimidine biosynthesis pathways relevant to disease. Based on these hypotheses, a clinical strategy for cytoproliferative disease is proposed that co-targets three cellular processes seen as “linchpins” of pyrimidines’ auto/paracrine effects, including targeting alternate pyrimidine biosynthesis pathways.
... Liu pointed out that DDR signaling and DNA repair were impaired in Mcph1-∆BR1 primary MEFs [21]. In MEFs, DNA damage caused by exposure to γ-radiation can increase the expression of p19ARF, thereby causing cell cycle arrest [41]. Although we have not further investigated the direct relationship between DDR and p19ARF expression in Mcph1 knockout cells, it is reasonable to speculate that the knockout of Mcph1 may increase p19ARF by causing DNA damage and then lead to cell cycle arrest and senescence. ...
Article
Full-text available
MCPH1 has been identified as the causal gene for primary microcephaly type 1, a neurodevelopmental disorder characterized by reduced brain size and delayed growth. As a multifunction protein, MCPH1 has been reported to repress the expression of TERT and interact with transcriptional regulator E2F1. However, it remains unclear whether MCPH1 regulates brain development through its transcriptional regulation function. This study showed that the knockout of Mcph1 in mice leads to delayed growth as early as the embryo stage E11.5. Transcriptome analysis (RNA-seq) revealed that the deletion of Mcph1 resulted in changes in the expression levels of a limited number of genes. Although the expression of some of E2F1 targets, such as Satb2 and Cdkn1c, was affected, the differentially expressed genes (DEGs) were not significantly enriched as E2F1 target genes. Further investigations showed that primary and immortalized Mcph1 knockout mouse embryonic fibroblasts (MEFs) exhibited cell cycle arrest and cellular senescence phenotype. Interestingly, the upregulation of p19ARF was detected in Mcph1 knockout MEFs, and silencing p19Arf restored the cell cycle and growth arrest to wild-type levels. Our findings suggested it is unlikely that MCPH1 regulates neurodevelopment through E2F1-mediated transcriptional regulation, and p19ARF-dependent cell cycle arrest and cellular senescence may contribute to the developmental abnormalities observed in primary microcephaly.
... It has been suggested that ARF attenuates rRNA transcription [27]. There are reports where it suggests ARF has a role in DNA damage response by inducing oncogenes via p53 pathway [35][36][37]. Other functions include activation of autophagy [38,39]. ...
Article
Full-text available
Owing to its poor prognosis, the World Health Organization (WHO) lists lung cancer on top of the list when it comes to growing mortality rates and incidence. Usually, there are two types of lung cancer, small-cell lung cancer (SCLC) and non-small-cell lung cancer (NSCLC), which also includes adenocarcinoma, squamous cell carcinoma and large cell carcinomas. ARF, also known in humans as p14ARF and in the mouse as p19ARF, is a nucleolar protein and a member of INK4, a family of cyclin-independent kinase inhibitors (CKI). These genes are clustered on chromosome number 9p21 within the locus of CDKN2A. NSCLC has reported the role of p14ARF as a potential target. p14ARF has a basic mechanism to inhibit mouse double minute 2 protein that exhibits inhibitory action on p53, a phosphoprotein tumour suppressor, thus playing a role in various tumour-related activities such as growth inhibition, DNA damage, autophagy, apoptosis, cell cycle arrest and others. Extensive cancer research is ongoing and updated reports regarding the role of ARF in lung cancer are available. This article summarizes the available lung cancer ARF data, its molecular mechanisms and its associated signalling pathways. Attempts have been made to show how p14ARF functions in different types of lung cancer providing a thought to look upon ARF as a new target for treating the debilitating condition of lung cancer.
... The induction of p i 9^'^'' in response to oncogenic stress and certain types of DNA damage is thought to promote p53 activation (Quelle et al, 1995b;Kamijo et al, 1997;Stott, 1998;Damalas et al, 2001;Khan et al, 2000). Overexpression of pl9^*^^ in mouse fibroblasts is associated with p53 stabilisation and transactivation of p21'^''^', resulting in a G1 and/or G2/M cell cycle arrest (Quelle et al, 1995b;Kamijo et al, 1997;Stott, 1998). ...
Thesis
Oncogenic activation of the Ras gene has been implicated in many human tumours. However, despite the ability of Ras to transform immortal cell lines, activated Ras is growth inhibitory in primary cells. We have previously shown that activation of Ras/Raf signalling in primary Schwann cells results in a proliferative arrest due to the induction of the cyclin dependent kinase inhibitor (CDKI) p21Cip1. In this thesis I examine the mechanisms involved in p21Cip1 induction and the roles of other CDKIs in the Ras/Raf induced cell-cycle arrest. I show that Raf activation is also associated with the induction of the CDKI pl5INK4b, however, in contrast to other primary cell types, p16INK4a levels decrease and the induction of p19ARF is not associated with p53 stabilisation. In certain cell types, the Ras induced proliferative arrest is associated with differentiation. I therefore investigated whether the Ras/Raf induced proliferative arrest in Schwann cells was associated with an induction of differentiation. Surprisingly in vitro I found that Raf/MAPK signalling blocks Schwann cell differentiation, as measured by the downregulation of differentiation markers. In addition, Raf is able to induce Schwann cell de-differentiation. In vivo, differentiated Schwann cells are found in the peripheral nervous system in association with axons and are capable of de-differentiating and proliferating in response to nerve damage throughout life. To investigate the effect of Ras/Raf activation on Schwann cell-neuron interactions I have set up a Schwann cell-dorsal root ganglion (DRG) co-culture system and I have generated transgenic mice expressing an inducible Raf protein in myelinating Schwann cells. Using the Schwann cell-DRG system I show that activation of Raf does not prevent the recognition or association of Schwann cells with axons. However, using time-lapse microscopy I have found that Raf activation results in subtle changes in the dynamics of Schwann cell-axon interactions.
... Superfluous BER proteins not bound in a complex are targeted by two different E3 ubiquitin ligases called Mule/ARF-BP1 and CHIP: firstly, Mule adds a monoubiquitin to 'prime' the protein for subsequent ubiquitination and, secondly, CHIP extends this ubiquitin chain, thus labelling the target for proteasomal degradation [134,136]. Interestingly, the activity of Mule can be inhibited by the tumour suppressor ARF, which accumulates in response to DNA damage [137,138] by a transcriptional regulation through a PARP1-SIRT1-E2F1 axis [132]. Inhibition of Mule activity by ARF binding results in accumulation of active Pol β -XRCC1 -Lig III complexes able undertake DNA repair. ...
Article
Full-text available
Oxidative DNA damage constitutes a major threat to genetic integrity, and has thus been implicated in the pathogenesis of a wide variety of diseases, including cancer and neurodegeneration. 7,8-dihydro-8oxo-deoxyGuanine (8-oxo-G) is one of the best characterised oxidative DNA lesions, and it can give rise to point mutations due to its miscoding potential that instructs most DNA polymerases (Pols) to preferentially insert Adenine (A) opposite 8-oxo-G instead of the correct Cytosine (C). If uncorrected, A:8-oxo-G mispairs can give rise to C:G→A:T transversion mutations. Cells have evolved a variety of pathways to mitigate the mutational potential of 8-oxo-G that include i) mechanisms to avoid incorporation of oxidized nucleotides into DNA through nucleotide pool sanitisation enzymes (by MTH1, MTH2, MTH3 and NUDT5), ii) base excision repair (BER) of 8-oxo-G in DNA (involving MUTYH, OGG1, Pol λ, and other components of the BER machinery), and iii) faithful bypass of 8-oxo-G lesions during replication (using a switch between replicative Pols and Pol λ). In the following, the fate of 8-oxo-G in mammalian cells is reviewed in detail. The differential origins of 8-oxo-G in DNA and its consequences for genetic stability will be covered. This will be followed by a thorough discussion of the different mechanisms in place to cope with 8-oxo-G with an emphasis on Pol λ-mediated correct bypass of 8-oxo-G during MUTYH-initiated BER as well as replication across 8-oxo-G. Furthermore, the multitude of mechanisms in place to regulate key proteins involved in 8-oxo-G repair will be reviewed. Novel functions of 8-oxo-G as an epigenetic-like regulator and insights into the repair of 8-oxo-G within the cellular context will be touched upon. Finally, a discussion will outline the relevance of 8-oxo-G and the proteins involved in dealing with 8-oxo-G to human diseases with a special emphasis on cancer.
Article
Melanoma represents an aggressive malignant tumor, encapsulating frequent loss of differentiation markers, with familial melanoma constituting a relatively commonly encountered entity, in direct relationship with cyclin-dependent kinase inhibitor 2A (CDKN2A). The present study aims to identify the association between the immunohistochemical p14–p16 profile, the molecular CDKN2A findings and clinically diagnosed familial or multiple primary melanomas (MPM). We conducted a 5-year retrospective cross-sectional study, on patients diagnosed with familial or MPM. P14 and p16 immunohistochemical staining has been applied on the selected surgical specimens simultaneously with the performance of fluorescence in situ hybridization (FISH) CDKN2A testing. 13 out of the 23 included cases displayed p14 and/or p16 immunohistochemical absence and the main positive relationships were encountered between CDKN2A homozygous deletion and p14 ± p16 negative immunoreactions. Cases with exclusive p16 absent reaction (n = 7) were more frequently associated with the presence of distant metastases (85.71%), while samples with exclusive p14 immunohistochemical loss exhibited more favorable histopathological prognostic markers. The average percentage of p16-stained nuclei in the superficial dermis and the deep dermis were equal (29.54% for each), therefore infirming its potential predictive and/or prognostic utility. The present study is the first of its type to approach the clinical, evolutionary and immunophenotypic correlations between p14–p16 immunohistochemical testing, CDKN2A molecular biology pattern, familial melanoma and spontaneous MPM in a cohort of Romanian patients. This analysis highlighted the value of singular p16 immunohistochemical absence as a predictor for aggressive biological behavior and unfavorable prognosis in familial melanoma and/or MPM, in comparison with the exclusive loss of p14, indifferent to the histopathological subtype. The present study emphasizes the utility of immunohistochemistry as a less expensive method of complementing the current testing arsenal and could represent the starting point for the elaboration of tailored diagnostic and therapeutic algorithms, based on the discovered p14–p16-CDKN2A significant correlation.
Article
Fibroblasts are typically quiescent and tumor suppressive in adult mammals but secrete factors upon activation in benign and malignant diseases such as wound healing and tumor growth1. As residents of the tumor microenvironment (TME), fibroblasts are subjected to the same stresses that tumor cells experience including nutrient deprivation that activates the integrated stress response (ISR) pathway. This eventually leads to either apoptosis due to intense or prolonged stress or autophagy to promote cellular survival in the absence of essential nutrients2–5. Studies have shown that nutrient deprivation also activates the tumor suppressor p53 in fibroblasts. However, little is known about the role of other tumor suppressors such as the INK4a locus of the CDKN2A gene, the second most commonly mutated tumor suppressor gene in human cancers6–8. The INK4a locus encodes two tumor suppressor genes including the p19 Alternative reading frame (p19Arf) that activates p53 by sequestering the ubiquitin ligase MDM2. While P19Arf regulation has been well characterized in cancer cells, its role in the TME has not been investigated. Endothelial cells in the TME are also subjected to stress such as shear stress that is altered upon changes in blood flow as a consequence of tumor vascularization as well as activities such as aerobic exercise. Exercise is commonly prescribed to cancer patients to enhance quality of life however, the effect of exercise during chemotherapy has not been extensively investigated. My work shows that the growth of transplanted tumor cells in the flank of mice is significantly upregulated in p19Arf-/- mice as compared to wild-type littermate controls indicating a role for P19Arf in the TME. My studies show that primary murine adult lung fibroblasts (ALFs) induce P19Arf expression upon nutrient deprivation and that prolonged leucine deprivation triggers apoptosis in wild-type ALFs. However, p19Arf-/- ALFs demonstrate enhanced proliferation, migration and survival in response to long-term leucine deprivation due in part to upregulation of the ISR pathway and increased autophagic flux. My data also suggests that loss of p19Arf in fibroblasts promotes survival during nutrient deprivation through increased proliferation and autophagy. My studies investigating the effect of acute aerobic exercise on endothelial cell activation and subsequent tumor vascular normalization show that a single round of acute exercise enhances chemotherapeutic efficacy when administered after exercise in melanoma xenograft tumor models independent of tumor vascular normalization. My data suggests that acute exercise may provide an opportunity to enhance chemotherapeutic efficacy.
Article
Full-text available
The INK4a-ARF locus encodes two proteins, p16INK4a and p19ARF, that restrain cell growth by affecting the functions of the retinoblastoma protein and p53, respectively. Disruption of this locus by deletions or point mutations is a common event in human cancer, perhaps second only to the loss of p53. Using insect cells infected with baculovirus vectors and NIH 3T3 fibroblasts infected with ARF retrovirus, we determined that mouse p19ARF can interact directly with p53, as well as with the p53 regulator mdm2. ARF can bind p53-DNA complexes, and it depends upon functional p53 to transcriptionally induce mdm2 and the cyclin-dependent kinase inhibitor p21Cip1, and to arrest cell proliferation. Binding of p19ARF to p53 requires the ARF N-terminal domain (amino acids 1–62) that is necessary and sufficient to induce cell cycle arrest. Overexpression of p19ARF in wild type or ARF-null mouse embryo fibroblasts increases the half-life of p53 from 15 to ≈75 min, correlating with an increased p53-dependent transcriptional response and growth arrest. Surprisingly, when overexpressed at supra-physiologic levels after introduction into ARF-null NIH 3T3 cells or mouse embryo fibroblasts, the p53 protein is handicapped in inducing this checkpoint response. In this setting, reintroduction of p19ARF restores p53’s ability to induce p21Cip1 and mdm2, implying that, in addition to stabilizing p53, ARF modulates p53-dependent function through an additional mechanism.
Article
Full-text available
The occurrence of DNA double strand breaks induces cell cycle arrest in mortal and immortal human cells. In normal, mortal fibroblasts this block to proliferation is permanent. It depends on the growth regulator p53 and a protein p53 induces, the cyclin dependent kinase inhibitor, p21. We show here that following DNA damage in mortal fibroblasts, the induction of p21 and p53 is to a large degree shortlived. By 8 days after a brief exposure to DNA strand breaking agents, bleomycin or actinomycin D, p53 protein is at baseline levels, while the p53 transactivation level is only slightly above its baseline. By this time the concentration of p21 protein, which goes up as high as 100-fold shortly after treatment, is down to just 2-4-fold over baseline levels. Following the drop in p21 concentration a large increase in the expression level of the tumor suppressor gene p16INK4a is observed. This scenario, where a transient increase in p21 is followed by a delayed induction of p16INK4a, also happens with the permanent arrest that occurs with cellular senescence. In fact, these cells treated with agents that cause DNA double strand breaks share a number of additional markers with senescent cells. Our findings indicate that these cells are very similar to senescent cells and that they have additional factor(s) beside p21 and p53 that maintain cell cycle arrest.
Article
The Ink4/Arf locus encodes two tumour-suppressor proteins, p16Ink4a and p19Arf, that govern the antiproliferative functions of the retinoblastoma and p53 proteins, respectively. Here we show that Arf binds to the product of the Mdm2 gene and sequesters it into the nucleolus, thereby preventing negative-feedback regulation of p53 by Mdm2 and leading to the activation of p53 in the nucleoplasm. Arf and Mdm2 co-localize in the nucleolus in response to activation of the oncoprotein Myc and as mouse fibroblasts undergo replicative senescence. These topological interactions of Arf and Mdm2 point towards a new mechanism for p53 activation.
Article
Most human cancers show perturbation of growth regulation mediated by the tumour-suppressor proteins retinoblastoma (RB) and p53 (ref. 1), indicating that loss of both pathways is necessary for tumour development. Loss of RB function leads to abnormal proliferation related to the deregulation of the E2F transcription factors, but also results in the activation of p53, which suppresses cell growth. Here we show that E2F-1 directly activates expression of the human tumour-suppressor protein p14ARF (the mouse homologue is called p19ARF), which binds to the MDM2-p53 complex and prevents p53 degradation2,5. These results complete a pathway linking abnormal proliferative signals, such as loss of RB, with the activation of a p53 response, through E2F-1 and p14ARF. They suggest that E2F-1, a protein inherently activated by cell-cycle progression, is part of a fail-safe mechanism to protect against aberrant cell growth.
Article
p21CIP11WAF1 is a CDK Inhibitor regulated by the tumor suppressor p53 and is hypothesized to mediate G1 arrest. p53 has been suggested to derive anti-oncogenic properties from this relationship. To test these notions, we created mice lacking p21CIP1/WAF1. They develop normally and (unlike p53−/− mice) have not developed spontaneous malignancies during 7 months of observation. Nonetheless, p21−/− embryonic fibroblasts are significantly deficient in their ability to arrest in G1 In response to DNA damage and nucleotide pool perturbation. p21−/− cells also exhibit a significant growth alteration in vitro, achieving a saturation density as high as that observed In p53−/− cells. In contrast, other aspects of p53 function, such as thymocytic apoptosis and the mitotic spindle checkpoint, appear normal. These results establish the role of p21CIP1/WAF1 in the G1 checkpoint, but suggest that the antiapoptotic and the anti-oncogenic effects of p53 are more complex.
Article
The INK4a-ARF locus encodes two unrelated proteins that both function in tumor suppression. p16INK4 binds to and inhibits the activity of CDK4 and CDK6, and ARF arrests the cell cycle in a p53-dependent manner. We show here that ARF binds to MDM2 and promotes the rapid degradation of MDM2. This interaction is mediated by the exon 1beta-encoded N-terminal domain of ARF and a C-terminal region of MDM2. ARF-promoted MDM2 degradation is associated with MDM2 modification and concurrent p53 stabilization and accumulation. The functional consequence of ARF-regulated p53 levels via MDM2 proteolysis is evidenced by the ability of ectopically expressed ARF to restore a p53-imposed G1 cell cycle arrest that is otherwise abrogated by MDM2. Thus, deletion of the ARF-INK4a locus simultaneously impairs both the INK4a-cyclin D/CDK4-RB and the ARF-MDM2-p53 pathways.
Article
The INK4a gene encodes two distinct growth inhibitors--the cyclin-dependent kinase inhibitor p16Ink4a, which is a component of the Rb pathway, and the tumor suppressor p19Arf, which has been functionally linked to p53. Here we show that p19Arf potently suppresses oncogenic transformation in primary cells and that this function is abrogated when p53 is neutralized by viral oncoproteins and dominant-negative mutants but not by the p53 antagonist MDM2. This finding, coupled with the observations that p19Arf and MDM2 physically interact and that p19Rrf blocks MDM2-induced p53 degradation and transactivational silencing, suggests that p19Arf functions mechanistically to prevent MDM2's neutralization of p53. Together, our findings ascribe INK4a's potent tumor suppressor activity to the cooperative actions of its two protein products and their relation to the two central growth control pathways, Rb and p53.
Article
Induced expression of a mutant human Ha-ras oncogene in NIH3T3 cells leads to the rapid production of multicentric chromosomes, acentric chromosome fragments, double minute chromosomes, increased heteroploidy, and increased capacity to undergo gene amplification. In this study we have used fluorescent-in-situ hybridization (FISH) to demonstrate that induction of the Ha-ras oncogene also leads to disruption of the mitotic machinery, resulting in aberrant mitoses and abnormal daughter cells. Cells induced to express an oncogenic Ha-ras transgene accumulate chromosomes that lag outside of the rest of the chromosomal architecture, chromosomes that form bridges between daughter nuclei at anaphase, and that form micronuclei. Many of these mitotic aberrations contain structurally abnormal chromosomes. These ras-induced changes were suppressed by the introduction of a gene encoding the dominant negative effector of ras, raf 301. Expression of raf301 in cells induced to express Ha-ras reduced the level of growth in soft agar, chromosome aberrations, mitotic aberrations, and frequency of gene amplification. These data provide evidence for an association between Ha-ras induced transformation, chromosome aberrations and gene amplification. Furthermore they offer insight into how the cell responds to the formation of aberrant chromosomes, and how disrupting chromosomal architecture could lead to further imbalances in the distribution of genetic material between daughter cells.
Article
Loss of cell cycle control and acquisition of chromosomal rearrangements such as gene amplification often occur during tumor progression, suggesting that they may be correlated. We show here that the wild-type p53 allele is lost when fibroblasts from patients with the Li-Fraumeni syndrome (LFS) are passaged in vitro. Normal and LFS cells containing wild-type p53 arrested in G1 when challenged with the uridine biosynthesis inhibitor PALA and did not undergo PALA-selected gene amplification. The converse occurred in cells lacking wild-type p53 expression. Expression of wild-type p53 in transformants of immortal and tumor cells containing mutant p53 alleles restored G1 control and reduced the frequency of gene amplification to undetectable levels. These studies reveal that p53 contributes to a metabolically regulated G1 check-point, and they provide a model for understanding how abnormal cell cycle progression leads to the genetic rearrangements involved in tumor progression.
Article
The inhibition of replicative DNA synthesis that follows DNA damage may be critical for avoiding genetic lesions that could contribute to cellular transformation. Exposure of ML-1 myeloblastic leukemia cells to nonlethal doses of the DNA damaging agents, gamma-irradiation or actinomycin D, causes a transient inhibition of replicative DNA synthesis via both G1 and G2 arrests. Levels of p53 protein in ML-1 cells and in proliferating normal bone marrow myeloid progenitor cells increase and decrease in temporal association with the G1 arrest. In contrast, the S-phase arrest of ML-1 cells caused by exposure to the anti-metabolite, cytosine arabinoside, which does not directly damage DNA, is not associated with a significant change in p53 protein levels. Caffeine treatment blocks both the G1 arrest and the induction of p53 protein after gamma-irradiation, thus suggesting that blocking the induction of p53 protein may contribute to the previously observed effects of caffeine on cell cycle changes after DNA damage. Unlike ML-1 cells and normal bone marrow myeloid progenitor cells, hematopoietic cells that either lack p53 gene expression or overexpress a mutant form of the p53 gene do not exhibit a G1 arrest after gamma-irradiation; however, the G2 arrest is unaffected by the status of the p53 gene. These results suggest a role for the wild-type p53 protein in the inhibition of DNA synthesis that follows DNA damage and thus suggest a new mechanism for how the loss of wild-type p53 might contribute to tumorigenesis.