ArticlePDF Available

Mutant p53 Disrupts Role of ShcA Protein in Balancing Smad Protein-dependent and -independent Signaling Activity of Transforming Growth Factor-beta (TGF-beta)

Authors:

Abstract and Figures

Biomarkers are lacking for identifying the switch of transforming growth factor-β (TGF-β) from tumor-suppressing to tumor-promoting. Mutated p53 (mp53) has been suggested to switch TGF-β to a tumor promoter. However, we found that mp53 does not always promote the oncogenic role of TGF-β. Here, we show that endogenous mp53 knockdown enhanced cell migration and phosphorylation of ERK in DU145 prostate cancer cells. Furthermore, ectopic expression of mp53 in p53-null PC-3 prostate cancer cells enhanced Smad-dependent signaling but inhibited TGF-β-induced cell migration by down-regulating activated ERK. Reactivation of ERK by the expression of its activator, MEK-1, restored TGF-β-induced cell migration. Because TGF-β is known to activate the MAPK/ERK pathway through direct phosphorylation of the adaptor protein ShcA and MAPK/ERK signaling is pivotal to tumor progression, we investigated whether ShcA contributed to mp53-induced ERK inhibition and the conversion of the role of TGF-β during carcinogenesis. We found that mp53 expression led to a decrease of phosphorylated p52ShcA/ERK levels and an increase of phosphorylated Smad levels in a panel of mp53-expressing cancer cell lines and in mammary glands and tumors from mp53 knock-in mice. By manipulating ShcA levels to regulate ERK and Smad signaling in human untransformed and cancer cell lines, we showed that the role of TGF-β in regulating anchorage-dependent and -independent growth and migration can be shifted between growth suppression and migration promotion. Thus, our results for the first time suggest that mp53 disrupts the role of ShcA in balancing the Smad-dependent and -independent signaling activity of TGF-β and that ShcA/ERK signaling is a major pathway regulating the tumor-promoting activity of TGF-β.
Content may be subject to copyright.
Mutant p53 Disrupts Role of ShcA Protein in Balancing Smad
Protein-dependent and -independent Signaling Activity of
Transforming Growth Factor-
(TGF-
)
*
S
Received for publication, May 27, 2011, and in revised form, October 26, 2011 Published, JBC Papers in Press, October 28, 2011, DOI 10.1074/jbc.M111.265397
Shu Lin
, Lan Yu
, Junhua Yang
, Zhao Liu
, Bijal Karia
‡§
, Alexander J. R. Bishop
द
, James Jackson
,
Guillermina Lozano
, John A. Copland**, Xiaoxin Mu
‡‡
, Beicheng Sun
‡‡
, and Lu-Zhe Sun
‡¶1
From the
Department of Cellular and Structural Biology,
§
Greehey Children’s Cancer Research Institute, and
Cancer Therapy and
Research Center, University of Texas Health Science Center, San Antonio, Texas 78229,
Department of Genetics, The University of
Texas M. D. Anderson Cancer Center, Houston, Texas 77030, **Department of Cancer Biology, Mayo Clinic, Jacksonville, Florida
32224, and
‡‡
Key Laboratory of Living Donor Liver Transplantation, First Affiliated Hospital of Nanjing Medical University,
Nanjing, China 210009
Background: Biomarkers driving TGF-
from tumor-suppressing to tumor-promoting remain elusive.
Results: p53 mutation inhibits TGF-
-induced ShcA/ERK signaling and enhances Smad signaling. Elevated p-p52ShcA levels
shift the role of TGF-
from growth suppression to migration promotion.
Conclusion: Mutant p53 disrupts the role of ShcA in balancing Smad-dependent and -independent signaling activity of TGF-
.
Significance: Elevated p-p52ShcA levels are a promising biomarker for TGF-
as a tumor promoter.
Biomarkers are lacking for identifying the switch of trans-
forming growth factor-
(TGF-
) from tumor-suppressing to
tumor-promoting. Mutated p53 (mp53) has been suggested to
switch TGF-
to a tumor promoter. However, we found that
mp53 does not always promote the oncogenic role of TGF-
.
Here, we show that endogenous mp53 knockdown enhanced
cell migration and phosphorylation of ERK in DU145 prostate
cancer cells. Furthermore, ectopic expression of mp53 in p53-
null PC-3 prostate cancer cells enhanced Smad-dependent sig-
naling but inhibited TGF-
-induced cell migration by down-
regulating activated ERK. Reactivation of ERK by the expression
of its activator, MEK-1, restored TGF-
-induced cell migration.
Because TGF-
is known to activate the MAPK/ERK pathway
through direct phosphorylation of the adaptor protein ShcA and
MAPK/ERK signaling is pivotal to tumor progression, we inves-
tigated whether ShcA contributed to mp53-induced ERK inhi-
bition and the conversion of the role of TGF-
during carcino-
genesis. We found that mp53 expression led to a decrease of
phosphorylated p52ShcA/ERK levels and an increase of phos-
phorylated Smad levels in a panel of mp53-expressing cancer
cell lines and in mammary glands and tumors from mp53
knock-in mice. By manipulating ShcA levels to regulate ERK
and Smad signaling in human untransformed and cancer cell
lines, we showed that the role of TGF-
in regulating anchorage-
dependent and -independent growth and migration can be
shifted between growth suppression and migration promotion.
Thus, our results for the first time suggest that mp53 disrupts
the role of ShcA in balancing the Smad-dependent and -inde-
pendent signaling activity of TGF-
and that ShcA/ERK signal-
ing is a major pathway regulating the tumor-promoting activity
of TGF-
.
Transforming growth factor-
(TGF-
) is a tumor suppres-
sor during early tumor outgrowth. However, carcinogenesis-
mediated elevation of TGF-
production and signaling is often
tumor-promoting at later stages, leading to enhanced tumor
cell migration, invasion, and metastasis (1). Increased TGF-
is
often associated with the loss of the growth-inhibitory activity
of TGF-
and its conversion to promote malignant progression
of cancers (2), making TGF-
a potential therapeutic target.
Indeed, many preclinical studies have shown the efficacy of var-
ious types of TGF-
inhibitors in blocking tumor growth,
angiogenesis, and metastasis in animal models of human and
rodent cancer (3–5). However, biomarkers are lacking for sig-
nifying the complicated molecular alterations mediating the
switch of TGF-
from tumor suppression to tumor promotion
and for identifying appropriate cancer patients for therapy with
TGF-
inhibitors.
As a homodimeric polypeptide in humans and mice, TGF-
signals through cell surface receptors called TGF-
type I (RI)
2
and type II (RII) receptors to regulate multiple cellular func-
tions including cell proliferation, differentiation, migration,
and wound healing (1). RI and RII are transmembrane serine/
threonine kinase receptors that also contain tyrosine kinase
activity (6). Active TGF-
ligands bind first to RII, which then
*This work was supported, in whole or in part, by National Institutes of Health
Grants R01CA075253 (to L.-Z. S.), R01CA079683 (to L.-Z. S.), R01CA104505
(to J. A. C.), and P30CA054174-17, an NCI cancer center support grant (to
the Cancer Therapy and Research Center at the University of Texas Health
Science Center at San Antonio).
S
The on-line version of this article (available at http://www.jbc.org) contains
supplemental Figs. 1 and 2.
1
To whom correspondence should be addressed: Dept. of Cellular and Struc-
tural Biology, University of Texas Health Science Center, 7703 Floyd Curl
Dr., Mail Code 7762, San Antonio, TX 78229-3900. Tel.: 210-567-5746; Fax:
210-567-3803; E-mail: sunl@uthscsa.edu.
2
The abbreviations used are: RI, TGF-
type I receptor; RII, TGF-
type II recep-
tor; mp53, mutant p53; SBE, Smad-binding element; R-Smad, receptor-
regulated class of Smads; DNRII, dominant-negative RII; MTT, 3-(4,5-
dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide; MMTV, murine
mammary tumor virus; p-ERK, phosphorylated ERK; *MEK-1, constitutively
active form of MEK-1.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 286, NO. 51, pp. 44023–44034, December 23, 2011
© 2011 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44023
recruits RI, leading to the phosphorylation and activation of RI.
Active RI directly phosphorylates the receptor-regulated class
of Smads (R-Smad), Smad2/3. Phosphorylated R-Smads in turn
associate with Smad4 and translocate into the nucleus to regu-
late the transcription of TGF-
-responsive target genes (1).
Besides Smad-mediated canonical signaling, TGF-
also sig-
nals through Smad-independent pathways including the Ras/
Raf/ERK pathway (7). The integration of TGF-
-mediated-
Smad-dependent and -independent signaling is believed to
contribute to the key events of TGF-
-induced tumor progres-
sion including ERK signaling-mediated cell migration (8, 9).
TGF-
activates Ras/ERK signaling through the direct phos-
phorylation of the adaptor protein ShcA (10). ShcA belongs to
the family of Shc adaptor proteins, which are substrates of
receptor tyrosine kinases (11). ShcA consists of three isoforms,
p46, p52, and p66. They are derived from two different tran-
scripts, called p66 and p52/p46 mRNAs (12). Compared with
p52ShcA, p46ShcA results from a different in-frame ATG tran-
script and is predominantly expressed in mitochondria with an
elusive role (13). Following the TGF-
engagement, tyrosine-
phosphorylated RI recruits and directly phosphorylates p52/
46ShcA proteins on tyrosine sites, leading to their association
with Grb2 adaptor protein and Sos GTP exchange factor (14).
The ShcA-Grb2-Sos complex activates Ras (15), thereby initi-
ating the sequential activation of c-Raf, MEK, and ERK1/2.
Among the three tyrosine phosphorylation sites at residues
239/240 (Tyr-239/240) and 317 (Tyr-317) in the CH1 domain
of p52/46ShcA, Tyr-317 plays the major role in Ras/ERK sig-
naling activation (11, 16). Clinical studies have shown that a
high amount of Tyr-317-phosphorylated p52/46ShcA alone
with a low amount of p66ShcA serves as an efficient predictor
for identifying aggressive breast tumors with a high risk of
recurrence (17, 18). High levels of TGF-
are also associated
with poor outcome of human cancer (1). However, it is unclear
whether ShcA-mediated activation of ERK signaling contrib-
utes to the switch of TGF-
function during carcinogenesis.
A recently published study has shown that the presence of
mutated p53 (mp53) in the DNA-binding domain in certain
cells together with additional Ras activation can switch TGF-
activity to that of a tumor promoter in the MDA-MB-231 breast
cancer cell line and H1299 non-small cell lung cancer cell line
(19). Somatic mutation-induced inactivation of p53 occurs in
about 50% of human cancers including breast and prostate can-
cer (20). p53 mutation is considered a biomarker of advanced
prostate cancer in which prostate cancer cells lose differenti-
ated phenotypes and transit from androgen-dependent to
androgen-independent growth (21). Rather than losing the
wild-type p53 (WTp53), the tumor with retention of mp53 has
been shown to be more aggressive and associated with poor
outcome in certain cancer types (22). For example, mp53 was
shown to enhance the cell migration and invasion in breast and
lung cancer cell lines (23). However, the role of mp53 in medi-
ating the key steps of cancer progression including cell migra-
tion and invasion is largely cell context-dependent and contro-
versial (24). For example, in human endometrial cancer cells,
the p53 R213Q mutation does not promote cell migration (25).
As shown in another study using the H1299 cell line, p53
R175H negatively regulates cell migration when TGF-
/Smad
signaling is repressed (26). At present, the cellular context for
mp53 to exert its oncogenic activity is underexplored. In addi-
tion, little is known about whether mp53 alone is sufficient to
activate the switch of TGF-
during tumor progression.
To gain a more thorough understanding of the effect of mp53
on tumor progression, particularly with respect to the role of
TGF-
signaling, we investigated whether mp53 contributes to
the conversion of the role of TGF-
in the regulation of cell
growth and migration. Here, we show that mp53 alone is not
sufficient to promote the oncogenic role of TGF-
. We further
demonstrate that mp53 disrupts the role of ShcA in altering the
signaling strength of TGF-
through ERK and Smad pathways
in certain human untransformed and cancer cell lines and mice
with mp53 knock-in. ShcA-mediated ERK signaling appears to
play a more dominant role in conferring the tumor-promoting
activity of TGF-
in the regulation of cell growth and migra-
tion. Our finding provides novel insight into the role of ShcA as
a promising biomarker in driving TGF-
signaling toward
tumor promotion.
EXPERIMENTAL PROCEDURES
Ethics Statement—All animal experiments were conducted
following appropriate guidelines. They were approved by the
Institutional Animal Care and Use Committee and monitored
by the Department of Laboratory Animal Resources at the Uni-
versity of Texas Health Science Center at San Antonio (proto-
col identification numbers 99142-34-11-A and 05054-34-01-A)
and the M. D. Anderson Institutional Animal Care and Use
Committee (protocol identification number 079906634).
Cell Culture—Human untransformed mammary epithelial
cell line MCF-10A was obtained from the Michigan Cancer
Foundation. These cells were grown in DMEM/F-12 supple-
mented with 5% horse serum, EGF, NaHCO
3
, hydrocortisone,
insulin, Fungizone, CaCl
2
, cholera toxin, and antibiotics.
Human prostate carcinoma cell lines PC-3, DU145, and 22Rv-1
and human breast cancer cell line BT20 was purchased from the
American Type Culture Collection (ATCC, Manassas, VA).
The human breast cancer MCF-7 control cell line and the dom-
inant-negative RII (DNRII)-transfected cell line were provided
by Dr. Michael G. Brattain (27). All these cells were cultured in
McCoy’s 5A medium with 10% fetal bovine serum (FBS) and
other supplements as described previously (28). The human
breast cancer cell line BT474 was obtained from ATCC and
cultured in DMEM (low glucose) with 10% FBS. Cells were
maintained at 37 °C in a 5% CO
2
humidified incubator.
Chemicals—The small RI kinase inhibitor HTS466284
reported previously to be an ATP-competitive inhibitor of RI
kinase (29) was synthesized by the Chemical Synthesis Core of
Vanderbilt University. U0126 is an MEK-1/2 inhibitor from
Calbiochem.
Plasmids and Transfection—p52/46ShcA plasmid was pur-
chased from Origene. p53 R175H-pCMV-Neo-Bam plasmid
was provided by Dr. Harikrishna Nakshatri. p53
R273H-pCDNA3 plasmid was provided by Dr. Zhi-Min Yuan.
Human WTp53 expression plasmid pRc/CMV hp53 was
obtained from Dr. Arnold Levine. Constitutively active MEK-1
plasmid was provided by Dr. Kun-Liang Guan. Stable transfec-
tion of p53 R175H into PC-3 cells was performed by using Lipo-
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44024 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
fectamine 2000 (Invitrogen). Forty-eight hours after the trans-
fection, G418 selection of neomycin-resistant cells was
conducted for 1 week. The sequence of p53 siRNA 1 is 5-CCG
GAC GAU AUU GAA CAA UGG UUC A-3, and the sequence
of p53 siRNA 2 is 5-GCU-UCG AGA UGU UCC GAG AGC
UGA A-3(Invitrogen). The sequence of ShcA siRNA is
5-GAC UAA GGA UCA CCG CUU U-3(Dharmacon, Lafay-
ette, CO). All transient transfections were performed by using
Lipofectamine 2000 according to the manufacturer’s protocol.
The pLKO.1-puro lentiviral RII shRNA and control shRNA
were purchased from Sigma. The sequence of RII shRNA is
5-CCG GCC TGA CTT GTT GCT AGT CAT ACT CGA GTA
TGA CTA GCA ACA AGT CAG GTT TTG-3. The process of
generating MCF-10A cells with stable knockdown of RII was
conducted according to the manufacturer’s protocol.
Immunoblotting Analysis—Immunoblotting analyses were
performed as described previously (30). Primary antibodies
were obtained from the following sources: p-Smad2, p-ERK
(Thr-202/Tyr-204), p-ShcA (Tyr-317), total-ERK, and MEK-
1/2 from Cell Signaling Technology (Danvers, MA); p-Smad3
from Epitomics (Burlingame, CA); total Smad2/3 from BD
Transduction Laboratories; p53 from Santa Cruz Biotechnol-
ogy (Santa Cruz, CA); total ShcA from BD Biosciences; and RII
from Abcam (Cambridge, MA). Relative expression levels of the
indicated genes were quantified with ImageJ software (National
Institutes of Health).
Cell Migration Assay—Cell migration assays were performed
in 24-well Boyden chambers with 8-
m pore polycarbonate
membranes (BD Biosciences). Cells at the indicated number in
serum-free medium were seeded in the upper insert. Complete
medium with or without treatment was added in the lower
chamber. After 18 h, cells that had migrated through the mem-
brane were stained with the Hema 3 Stain 18 kit (Fisher Scien-
tific) according to the manufacturer’s protocol. Migrated cells
were counted under a microscope with 100magnification.
3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium Bro-
mide (MTT) Assay—Cells were plated in a 96-well plate at 2,500
cells/well with or without TGF-
1 treatment. Two hours
before each time point, 50
l of MTT (2 mg/ml in PBS) was
added into each well, and cells were incubated at 37 °C for
another 2 h. DMSO (100
l) was added into each well after the
medium was removed. For dissolving the precipitate, the plate
was gently shaken on a shaker for 10 min. The absorbance was
measured at 595 nm with a microplate reader (BioTek Instru-
ment, Winooski, VT).
Soft Agar Colony Formation Assay—Cells in 1 ml of 0.4% low
melting agarose (Invitrogen) with culture medium were plated
at 6,000 cells/well on top of existing 0.8% agarose in 6-well
plates. The wells were covered with 1 ml of culture medium
containing various treatments and incubated at 37 °C in a 5%
CO
2
incubator for the indicated number of days. Visualized
colonies were counted after staining with p-iodonitrotetrazo-
lium violet (Sigma) overnight.
SBE-Luciferase Reporter Assay—The pSBE4-Luc plasmid
with Smad-responsive promoter and luciferase reporter gene
was used to measure the TGF-
-induced transcriptional activ-
ity (31). Cells at 100,000 cells/well were plated in 12-well plates.
After 24 h, the pSBE4-Luc plasmid (0.4
g) was transiently
co-transfected with a
-galactosidase expression plasmid (0.1
g) into the cells by using Lipofectamine 2000. TGF-
1 was
added to the transfected cells 5 h later. After an additional 20 h
of incubation, cells were lysed, and the activities of luciferase
and
-galactosidase of the cell lysates were assayed as described
previously (32). Luciferase activity was normalized to
-galac-
tosidase activity.
Animal Tissue Protein Extraction—C57BL/6 mice with a
heterozygous p53
R172H
mutation were described previously
(33). Heterozygous p53
R172P
mice were generated in a similar
way (34). They were crossed to C57BL/6 p
un/un
mice (The Jack-
son Laboratory, Bar Harbor, ME) (35). Heterozygous breeding
cohorts of p53
R172P/
p
un/un
and p53
R172H/
p
un/un
were inter-
crossed to produce the p53
R172P/R172P
p
un/un
,p53
R172H/R172H
p
un/un
, and p53
/
p
un/un
mice. MMTV-Wnt1 mice were bred
with p53
/
or p53
R172H/
mice to generate MMTV-Wnt1
mice in p53
/
,p53
R172H/
, and p53
R172H/R172H
backgrounds.
A fraction of tumors in the p53
R172H/
background undergo
loss of heterozygosity (herein referred to as p53
R172H/O
) and
thus are functionally p53 mutants. Loss of heterozygosity anal-
ysis was performed as described previously (36). The tumor
sizes were measured regularly with a caliper in two dimensions.
Tumor volumes (V) were calculated with the equation V(L
W
2
)0.5 where Lis length and Wis width. When tumors
reached a volume of 500 mm
3
, mice were sacrificed, and the
tumors were collected.
The isolated mammary tissues from p53
R172P/R172P
p
un/un
,
p53
R172H/R172H
p
un/un
, and p53
/
p
un/un
mice or breast
tumors from MMTV-Wnt1-p53
/
,-p53
R172H/R172H
and
-p53
R172H/O
mice were snap frozen in liquid nitrogen. Proteins
for immunoblotting analysis were isolated from liquid nitrogen
by grinding tissues using T-Per extraction reagent (Thermo
Fisher Scientific Inc., Rockford, IL) according to the manufac-
turer’s protocol.
Statistical Analysis—Two-tailed Student’s ttests were used
to determine the significant difference between two mean val-
ues from the control and experimental data. All statistical anal-
ysis was performed with GraphPad Prism 3.03 software
(GraphPad Software, La Jolla, CA).
RESULTS
Mutant p53 Inhibits Cell Migration and Down-regulates ERK
Signaling in Prostate Cancer Cell Lines—To investigate
whether mp53 alone can promote tumor migration, we
knocked down mp53 in the human prostate cancer cell line
DU145 containing inactivating endogenous p53 P223L and
V274F mutations in its DNA-binding domain (37). We found
that instead of making cells less migratory mp53 knockdown
significantly enhanced cell migration accompanied by the acti-
vation of ERK via phosphorylation (p-ERK) (Fig. 1, Aand B).
Active ERK signaling is reported to be essential for tumor
metastasis progression including cell migration (9), suggesting
that the mp53 depletion-induced migration was likely caused
by the up-regulated ERK signaling. We further stably intro-
duced mutant p53 R175H with a mutation in its DNA-binding
domain into the p53-null human prostate cancer cell line PC-3
(PC-3/mp53). This resulted in the repression of cell migration
as well as the down-regulation of ERK signaling (Fig. 1, Cand
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44025
D). These data indicate that mp53 in certain prostate cell lines
attenuates cell migration, which is a key step of procancer
metastasis.
Mutant p53 Represses Oncogenic Role of TGF-
and Enhances
TGF-
/Smad Signaling in Prostate Cancer Cell Line
The presence of mp53 has been shown to facilitate the tumor-
promoting activity of TGF-
in some breast cancer models (19).
Because mp53 was found not to enhance cell migration in our
studies with two prostate cancer cells, we explored whether
mp53 showed a different effect on the oncogenic role of TGF-
.
We observed that TGF-
significantly increased migration of
the control PC-3 cells, whereas it suppressed migration in the
PC-3/mp53 cells, suggesting that the activity of TGF-
is
affected by the presence of mp53 (Fig. 2A). Because PC-3 is a
tumorigenic cell line, we next performed a soft agar colony
formation assay to assess the effect of p53 R175H on the growth
of PC-3 cells in an anchorage-independent manner. We found
that the presence of p53 R175H repressed the anchorage-inde-
pendent growth ability of PC-3/mp53 cells compared with cells
without mp53. Interestingly, the colony formation was more
dramatically inhibited by TGF-
, and the treatment with
HTS466284, an RI kinase inhibitor (29), significantly stimu-
lated colony formation in PC-3/mp53 cells (Fig. 2, Band C). To
rule out the possibility that the p53 R175H-reduced cell migra-
tion in PC-3 cells was due to an altered cell growth rate, we
verified that the presence of p53 R175H in PC-3 cells showed
little effect on cell growth but that TGF-
treatment induced a
moderate growth inhibition in a dose-dependent manner (Fig.
2D). Thus, these results indicate that p53 R175H alone is not
sufficient to switch TGF-
to be more tumor-promoting in the
PC-3 cell line. Considering that TGF-
-inhibited anchorage-
dependent and -independent cell growth is mainly due to
Smad-dependent signaling (38, 39), we investigated whether
p53 R175H could alter the activation of TGF-
/Smad signaling
in PC-3/mp53 cells. We found that p53 R175H made PC-3 cells
more sensitive to TGF-
in a dose-dependent manner as evi-
denced by higher levels of TGF-
-induced phosphorylation of
Smad2/3 (Fig. 2E). Additionally, the transcriptional activity of
TGF-
was also increased in PC-3/mp53 cells when compared
with the control cells as detected with transfection of a Smad-
responsive promoter-luciferase reporter plasmid (SBE-Luc)
(Fig. 2F), further indicating that the presence of p53 R175H
enhanced the Smad-dependent signaling. These results
revealed that the addition of p53 R175H to PC-3 cells enhanced
TGF-
/Smad-signaling while inhibiting TGF-
-induced cell
migration.
Mutant p53 Represses Activation of ShcA/ERK Signaling—It
is suggested that Smad-dependent and -independent signaling
pathways work together to drive the key events of TGF-
-in-
duced cell migration and metastasis (8). However, our observa-
tions indicate that Smad-dependent signaling is not responsible
for the loss of TGF-
-induced migration in PC-3/mp53 cells as
reflected by the enhanced TGF-
/Smad signaling. Conse-
quently, we speculated that MAPK/ERK signaling might be
involved in the loss of TGF-
-induced cell migration of PC-3/
mp53 cells. Consistent with our observation above (Fig. 1C), we
found that the basal level of p-ERK was markedly reduced in the
PC-3/mp53 cells when compared with control cells (Fig. 3A).
Additionally, TGF-
was unable to further activate ERK in
PC-3/mp53 cells (Fig. 3A). To confirm that this observation was
specifically due to mp53 expression and not an aberrant
selection of a pool of antibiotic-resistant clones, we tran-
siently introduced p53 R175H into PC-3 cells and found that
p-ERK was greatly reduced when compared with PC-3 cells
transfected with a control plasmid or a WTp53 expression
plasmid (Fig. 3B). In parallel, p53 R175H knockdown in
PC-3/mp53 cells with p53 siRNA restored the p-ERK level
(Fig. 3C). We additionally used another p53 mutation in the
DNA-binding domain, p53 R273H, which also enhanced
TGF-
-stimulated phosphorylation of Smad2/3 and inhib-
ited the basal and TGF-
-stimulated p-ERK levels. Consis-
tently, p53 R273H also significantly attenuated TGF-
-in-
duced cell migration of PC-3 cells in comparison with the
control cells (Fig. 3Dand supplemental Fig. 1A). However,
exogenous human WTp53 expression in PC-3 cells showed
no effect on the level of TGF-
-induced cell migration, TGF-
-induced activation of ERK, or Smad signaling (Fig. 3Eand
supplemental Fig. 1B). Because tyrosine-phosphorylated
p52ShcA has been shown to positively mediate TGF-
-acti-
vated Ras/ERK signaling (10), we next examined whether the
mp53-inhibited phosphorylation of ERK was linked to the
down-regulation of p-p52ShcA. Indeed, the expression of
p53 R175H and p53 R273H, but not WTp53, in PC-3 cells
inhibited TGF-
-induced phosphorylation of p52ShcA
FIGURE 1. Mutant p53 inhibits cell migration and ERK signaling in pros-
tate cancer cell lines. A, immunoblotting analysis of the indicated gene
expression in DU145 cells, which were transfected with control or p53 siRNA
for 72 h. B, cell migration was assessed in DU145 cells 48 h after the transfec-
tion. Cells were plated at 40,000 cells/insert and incubated for 18 h as
described under “Experimental Procedures.” Data represent mean S.E.
from three inserts. *, p0.05. C, immunoblotting analysis of the indicated
gene expression in PC-3 cells with stable introduction of p53 R175H expres-
sion plasmid (PC-3/mp53) or empty vector. D, cell migration was assessed in
PC-3/mp53 and control cells plated at 70,000 cells/insert and incubated for
18 h. Data represent mean S.E. from three inserts. *, p0.05. T-ERK, total
ERK. Error bars represent S.E.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44026 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
compared with control cells (Fig. 3F). Thus, these data
revealed that mp53, but not WTp53, has the ability to down-
regulate ShcA-mediated ERK signaling.
Expression of MEK-1 or ShcA Restores Active ERK Level and
TGF-
-induced Cell Migration—To determine whether the
attenuated ERK signaling led to the loss of TGF-
-induced cell
migration in PC-3/mp53 cells, we examined whether TGF-
-
induced cell migration could be rescued by the reactivation of
ERK. As shown in Fig. 4A, ectopic expression of the ERK acti-
vator, a constitutively active form of MEK-1 (*MEK-1), was able
to restore the level of p-ERK in PC-3/mp53 cells. More impor-
tantly, compared with control plasmid-transfected cells, PC-3/
mp53 cells with *MEK-1 expression-rescued p-ERK became
responsive again to TGF-
in cell migration (Fig. 4B). These
results suggest the need for active ERK signaling to mediate the
migration-promoting activity of TGF-
in the presence of
mp53. To further confirm our results, we used DU145 cells to
test whether manipulating the active ERK by elevated ShcA
activation also affected TGF-
-induced cell migration in a dif-
ferent mp53-containing model system. We found that TGF-
slightly induced the phosphorylation of ShcA and did not
induce the phosphorylation of ERK (Fig. 4C). Neither was
TGF-
able to induce DU145 cell migration (Fig. 4D), reca-
pitulating the results with PC-3/mp53 (Fig. 2A). Conversely,
ectopic overexpression of p52/46ShcA in the presence of
TGF-
expression resulted in the clear activation of
p52ShcA and ERK as well as the stimulation of cell migra-
tion. These findings are highly consistent with the findings
from PC-3/mp53 cells with reactivated ERK (Fig. 4B). Our
observations thus far have demonstrated that TGF-
is able
to induce cell migration in the context of mp53 when
p-p52ShcA is elevated and that this correlates with activa-
tion of ERK. To verify that the increase of TGF-
-induced
cell migration after overexpression of p52/46ShcA was in
fact mediated by the activation of ERK, we treated the PC-3
cells with an ERK signaling inhibitor, U0126, and found that
TGF-
-induced cell migration was totally abolished (Fig.
4E). Thus, these findings indicate that TGF-
-induced cell
migration is dependent on its activation of ERK signaling in
prostate cancer cells in which mp53 tends to attenuate TGF-
-induced cancer malignancy.
FIGURE 2. Mutant p53 inhibits oncogenic role of TGF-
and enhances TGF-
/Smad signaling. A, cell migration was assayed with PC-3/mp53 and PC-3
control cells plated at 70,000 cells/insert with or without TGF-
1 (2 ng/ml) for 18 h. Data represent mean S.E. from three inserts. **, p0.01; ***, p0.0001.
B, soft agar colony formation ability was assessed in PC-3/mp53 and PC-3 control cells under the indicated TGF-
1 or HTS466284 (HTS) treatment for 12 days.
Data represent mean S.E. from three wells. *, p0.05; **, p0.01. C, representative pictures of soft agar colonies of PC-3/mp53 and PC-3 control cells were
taken after 12-day treatment with 2 ng/ml TGF-
1or0.1
MHTS466284. D, effect of TGF-
1 on cell proliferation was measured with the MTT assay at day 6 after
incubation with the indicated dose of TGF-
1. Data represent mean S.E. from five wells. A two-tailed ttest was performed to compare the mean of relative
cell number between PC-3/mp53 and PC-3 control cells for all treatments. *, p0.05. E, immunoblotting analysis for the indicated gene expression was
performed in PC-3/mp53 and PC-3 control cells under TGF-
1 treatment for 40 min. F, SBE-luciferase assay was performed using PC-3/mp53 and PC-3 control
cells after TGF-
1 treatment as described under “Experimental Procedures.” Data represent mean S.E. from three independent transfections of relative
luciferase units normalized to
-galactosidase activity. *, p0.05; **, p0.01. T-Smad2/3, total Smad2/3. Error bars represent S.E.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44027
ShcA Alters Role of TGF-
in Cellular Migration and Anchor-
age-dependent and -independent Growth in Transformed Cells
To this point, we have demonstrated that mp53-induced alter-
ation of Smad-dependent and -independent TGF-
signaling is
due to altered ShcA activation. Therefore, we next investigated
whether ShcA could serve as a biomarker in converting the role
of TGF-
from growth suppression to migration promotion in
cancer cells. To this end, we examined the effect of manipulat-
ing the expression level of ShcA in PC-3 cells on growth and
migration. Interestingly, we found that knockdown of ShcA iso-
forms by pan-ShcA siRNA decreased TGF-
-induced phos-
phorylation of ERK and increased TGF-
-induced phosphory-
lation of Smad2/3 (Fig. 5A). Conversely, the ectopic
overexpression of p52/46ShcA raised TGF-
-induced phos-
phorylated ERK levels and reduced TGF-
-induced phospho-
rylated Smad2/3 levels (Fig. 5B). Consistently, we found that
knockdown of p52/46ShcA enhanced, but ectopic overexpres-
sion of p52/46ShcA reduced, TGF-
-inhibited anchorage-de-
pendent (Fig. 5, Cand D) and -independent cell growth (Fig. 5,
Eand F). TGF-
-induced cell migration was significantly
diminished in ShcA-depleted cells but significantly increased in
p52/46ShcA-overexpressing cells (Fig. 5, Gand H). Because
anchorage-independent growth ability is associated with
tumorigenicity in vivo and cell migration is a key step of tumor
progression, our results suggest that ShcA attenuates the tumor
suppressor activity of TGF-
while enhancing its tumor pro-
moter activity.
ShcA Alters Role of TGF-
in Cellular Migration and Cell
Growth in Untransformed Cells—To further investigate the
effect of ShcA on TGF-
-mediated cellular growth and migra-
tion, we introduced pan-ShcA siRNA or p52/46ShcA cDNA
into untransformed MCF-10A cells, which lack complicated
alterations of tumor suppressor genes and oncogenes. By using
this spontaneously immortalized and non-tumorigenic human
mammary epithelial cell line, we were able to examine the tran-
sition of the role of TGF-
in the early stage of transformation.
We essentially observed the same phenotypic changes as in the
transformed PC-3 cells. Specifically, knockdown of ShcA iso-
forms down-regulated TGF-
-induced phosphorylation of
ERK and up-regulated TGF-
-induced phosphorylation of
Smad2/3 (Fig. 6A). In cells with knockdown of ShcA, we found
that TGF-
-inhibited cell growth was significantly enhanced
(Fig. 6C) and that TGF-
-induced cell migration was signifi-
cantly reduced (Fig. 6E) in comparison with the control
siRNA-transfected cells. In contrast, when compared with vehicle-
transfected cells, ectopic overexpression of p52/46ShcA aug-
mented TGF-
-induced phosphorylation of ERK (Fig. 6B) and
TGF-
-induced cell migration (Fig. 6F) but repressed TGF-
-in-
duced phosphorylation of Smad2/3 (Fig. 6B) and significantly
attenuated TGF-
-inhibited cell growth (Fig. 6D). Thus, our data
suggest that ShcA can alter the role of TGF-
in controlling cellu-
lar growth and migration by balancing its signaling between the
ERK and Smad pathways. We next asked whether TGF-
-induced
tyrosine phosphorylation of p52ShcA by RI requires RII. There-
fore, we knocked down RII in MCF-10A cells by using an RII
shRNA and found that the depletion of RII led to the attenuation of
TGF-
-induced phosphorylation of p52ShcA and ERK (Fig. 6G).
Additionally, we have previously shown that ectopic expression of
a DNRII blocked TGF-
signaling and reduced the level of active
ERK in the human MCF-7 breast cancer cell line, which contains a
high level of autocrine TGF-
activity (40). Interestingly, we found
that this DNRII-expressing MCF-7 cell line also has a lower level of
the tyrosine-phosphorylated p52ShcA than the control vector-
transfected cells (Fig. 6H). Thus, our observations suggest that RII
is required for the activation of p52ShcA by TGF-
.
Mutant p53 Positively Correlates with Activation of Smad-
dependent Signaling and Negatively Correlates with Active
ShcA/ERK Signaling in Human Cancer Cell Lines and Trans-
genic Mouse Models—To further generalize our findings, we
tested the correlation among the presence of mp53, activation
of ShcA/ERK signaling, and Smad-dependent signaling in a
FIGURE 3. ShcA-mediated ERK signaling is down-regulated in presence of
mutated p53. A, immunoblotting analysis for the indicated gene expression
was performed in PC-3/mp53 and PC-3 control cells after TGF-
1 treatment
for 40 min. B, immunoblotting analysis for p-ERK and p53 was performed in
PC-3 cells 72 h after transfection of p53 R175H expression plasmid, human
WTp53 expression plasmid, or the empty vector. C, immunoblotting analysis
for the indicated gene expression was performed in PC-3/mp53 cells 72 h
after transfection of p53 siRNA 1, p53 siRNA 2, or control siRNA. D, immuno-
blotting analysis for the indicated gene expression was assessed in PC-3 cells
72 h after transfection of a p53 R273H expression plasmid or empty vector
followed by TGF-
1 treatment for 40 min. E, immunoblotting analysis for the
indicated gene expression was assessed in PC-3 cells 72 h after transfection of
a human WTp53 expression plasmid or empty vector followed by TGF-
1
treatment for 40 min. F, immunoblotting analysis for the indicated gene
expression was conducted in PC-3/mp53 and PC-3 control cells and PC-3 cells
transfected with p53 R273H or WTp53 expression plasmid for 72 h followed
by TGF-
1 treatment for 40 min. T-ERK, total ERK; T-Smad2/3, total Smad2/3;
con; control.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44028 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
panel of human breast and prostate cancer cell lines, each
expressing varying p53 mutations, although all mutations were
within the DNA-binding domain. Human breast cancer cell
lines BT20 with p53 K132Q mutation, BT474 with p53 E285K
mutation (41), and DU145 were used in our study. We also
included human prostate cancer cell line 22Rv1 with p53
Q331R mutation in the dimerization domain (42). mp53 knock-
down in those cell lines increased the phosphorylation of
p52ShcA and ERK, whereas it decreased the phosphorylation of
Smad2 or Smad3 (Fig. 7A). In mammary gland tissues from
WTp53
/
,p53
R172H/R172H
mutant, or p53
R172P/R172P
mutant
(equivalent to human p53 R175H and R175P mutations,
respectively) female mice on a C57BL/6 p
un/un
genetic back-
ground, we found that the total ERK-normalized phosphoryla-
tion of ERK was down-regulated, whereas the total Smad2/3-
normalized phosphorylation of Smad3 was up-regulated in the
tissues bearing the mp53 when compared with WTp53-ex-
pressing tissues (Fig. 7B). We also normalized the p-ERK and
p-Smad3 expression levels with GAPDH protein levels and
obtained a similar outcome (supplemental Fig. 2A). The
p-ShcA expression levels were undetectable, and we speculated
that the normal murine mammary tissues might express low
levels of p-ShcA (data not shown). Furthermore, to evaluate the
correlation among mp53, ShcA/ERK signaling, and Smad sig-
naling in a more clinically relevant setting, we collected mam-
mary tumor tissues with WTp53
/
,p53
R172H/R172H
mutation,
or p53
R172H/O
(p53 R172H heterozygous with loss of heterozy-
gosity) on the background of the MMTV-Wnt1 transgenic
mice, a well established model of breast cancer development
and progression. The presence of mp53 showed negative cor-
relation with the active ShcA/ERK signaling and positive corre-
lation with the active Smad-dependent signaling (Fig. 7Cand
supplemental Fig. 2B). These observations further support the
conclusion that p53 mutation disrupts the role of ShcA in bal-
ancing the Smad-dependent and Smad-independent signaling
activity of TGF-
.
DISCUSSION
The cross-talk between p53 mutation and oncogenic Ras/
ERK signaling has been demonstrated to promote TGF-
-in-
duced cell migration and metastasis in certain breast cancer
models (19). However, it is still unclear whether mp53 alone can
act as a tumor promoter and cause TGF-
signaling to become
oncogenic. Several studies have revealed that the malignancy
gained from mp53 is cell context-dependent and controversial
(24). For example, mp53 can enhance the cell migration and
invasion via up-regulation of the epithelial-mesenchymal tran-
sition factor Slug in human non-small cell lung cancer and
breast cancer cell lines (23). In contrast, p53 R312Q mutation
does not positively regulate migration of human endometrial
cancer cells (25). p53 R175H mutation was shown to inhibit the
migration of the H1299 lung cancer cell line when its RII was
down-regulated and TGF-
/Smad signaling was repressed (26).
Here, we focused on the p53 DNA-binding domain mutations,
FIGURE 4. Expression of MEK-1 or ShcA restores active ERK level and TGF-
-induced cell migration. A, immunoblotting analysis for the indicated gene
expression was assessed in PC-3/mp53 cells 72 h after transfection of an *MEK-1expression plasmid or empty vector. B, cell migration was assayed in
PC-3/mp53 cells 48 h after *MEK-1 transfection. Cells were plated at 70,000 cells/insert with or without TGF-
1 (2 ng/ml) for 18 h. Data represent mean S.E.
from three inserts. *, p0.05. C, 72 h post-transfection with an empty vector or p52/46ShcA expression plasmid, immunoblotting analysis for the indicated
gene expression was performed in DU145 cells with or without 2 ng/ml TGF-
1 treatment for 40 min. D, cell migration was assessed in DU145 cells 48 h after
transfection of p52/46ShcA or empty vector. Cells were plated at 40,000 cells/insert with or without TGF-
1 (2 ng/ml) for 18 h. Data represent mean S.E. from
three inserts. *, p0.05. E, cell migration was assessed in PC-3 cells 48 h after transfection of p52/46ShcA or empty vector. Cells were plated at 100,000
cells/insert and treated with TGF-
1 (2 ng/ml) or U0126 (10
M) for 18 h. Data represent mean S.E. from three inserts. ***, p0.0001. T-ERK, total ERK;
T-Smad2/3, total Smad2/3. Error bars represent S.E.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44029
which have been reported as the majority of p53 mutations in
human breast and prostate cancer cell lines (43). We found that
in DU145 and PC-3 human prostate cancer cells the endoge-
nous or ectopic expression of p53 with a point mutation in its
DNA-binding domain does not enhance the cell migration or
cause TGF-
to be more migration-promoting. Hence, our
results from these models indicate that mp53 alone does not
always function as a promoter of migration in already trans-
formed cells, and its function is likely context-dependent.
Indeed, p53 mutation has been shown to occur relatively late
during multistage oncogenic progression and often follows Ras
mutation-induced ERK signaling (44). It has been shown that
mp53 works together with oncogenic Ras to induce the expres-
sion of several protumorigenesis and prometastasis genes in
gene expression profiling studies (45). Our finding that mp53
repressed the oncogenic role of TGF-
in the model systems we
used suggests that additional signaling activation and genetic
alterations such as ERK signaling activation and attenuation of
TGF-
-induced growth inhibition appear necessary to collab-
oratively confer the tumor-promoting activity of mp53 and
TGF-
during cancer progression. Although the sequestration
of metastasis suppressor p63 by the formation of an mp53-
Smad-p63 ternary complex has been shown to enhance the
oncogenic activity of TGF-
(19), an additional mechanism
involving mp53 but not the formation of the ternary complex
has also been shown to drive TGF-
to become more tumor-
promoting (46). DU145 cells are negative for p63 expression
(47). Thus, the conversion of TGF-
is independent of the ter-
nary complex formation in DU145 cells with restored ShcA/
ERK signaling. The presence of mp53 repressed the basal level
of phosphorylated p52ShcA and also abolished the activation of
p52ShcA by TGF-
, suggesting that mp53 could be the
upstream regulator of p52ShcA. We speculate that the mecha-
nism by which mp53 down-regulates the phosphorylation of
p52ShcA is perhaps through an aberrant or dysregulated pro-
tein interaction. However, further investigation is needed to
elucidate the mechanism.
TGF-
receptors have been shown to possess dual tyrosine
and serine kinase activity and can directly tyrosine phosphoryl-
ate ShcA to activate Ras/MAPK signaling or serine phosphoryl-
ate ShcA with an unclear consequence (10). The same study
suggests that tyrosine phosphorylation of p66Shc, the inhibi-
tory isoform of ShcA, is more dependent on RI in untrans-
formed cells. On the other hand, it is unclear whether RI alone
is sufficient to induce tyrosine phosphorylation of p52ShcA. In
our study, the depletion of RII in the untransformed human
breast cells attenuated the TGF-
-induced activation of the
ShcA/ERK cascade as well as Smad signaling. DNRII has the
intact extracellular and transmembrane domains of the wild-
type RII and is capable of forming a functional complex with
TGF-
and RI. By introducing this DNRII into a human breast
cancer cell line that has autocrine TGF-
activity, we found that
the dysfunctional RII caused the inhibition of ShcA/ERK acti-
vation. Our observations suggest that RII is required for TGF-
to activate the p52ShcA/ERK pathway. It is worth mentioning
that other tyrosine kinase receptors such as epidermal growth
factor receptor and insulin receptor also have the ability to tyro-
sine phosphorylate p52/46ShcA and consequently activate Ras/
FIGURE 5. ShcA alters TGF-
-mediated Smad and ERK signaling and activ-
ity in human cancer cell. PC-3 cells were transfected with a pan-ShcA siRNA
or a p52/46ShcA expression plasmid as shown in Aand B, respectively. Sev-
enty-two hours post-transfection, immunoblotting analysis for the indicated
gene expression was performed using the transfected cells treated with or
without 2 ng/ml TGF-
1 for 40 min. Cand D, TGF-
1 (2 ng/ml)-induced cell
growth inhibition was measured with the MTT assay in cells 72 h after trans-
fection. Data are presented as the percentage of TGF-
-induced inhibition of
cell growth or colony formation (shown in Eand F) relative to untreated cells.
Data represent mean S.E. from four wells. Eand F, soft agar colony forma-
tion ability was assessed in cells 24 h after transfection. Cells were treated with
or without 2 ng/ml TGF-
1 for 6 days. Data represent mean S.E. from three
wells. Gand H, cell migration was assessed in cells 48 h after transfection. Cells
were plated at 70,000 cells/insert with or without TGF-
1 (2 ng/ml) for 18 h.
Data are presented as the -fold change of TGF-
-induced migration relative
to untreated cells. Data represent mean S.E. from three inserts. *, p0.05; **,
p0.01. T-ERK, total ERK; T-Smad2/3, total Smad2/3. Error bars represent S.E.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44030 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
ERK signaling (48). We speculate that RI and other tyrosine
receptors may compete for available ShcA. This competition
may further influence the signaling strength of TGF-
between
Smad and ERK signaling. Currently, how those pathways may
affect the Smad-dependent and -independent signaling of
TGF-
via p52/46ShcA is also under investigation.
The activation of TGF-
/Smad-dependent signaling
requires the serine/threonine kinase activity of RII and RI.
Tyrosine kinases, instead of serine/threonine kinases, have
been reported as the catalytic center of RI in certain structural
features (49). Conceivably, R-Smads and ShcA may compete for
RI kinase, and the balance between the tyrosine phosphoryla-
tion of ShcA and serine phosphorylation of R-Smads could
determine the signaling strength of TGF-
through the Smad-
independent and -dependent pathways, respectively.
It is well known that TGF-
-induced cell growth inhibition
in epithelial cells depends on Smad-dependent signaling (38,
39). Loss of Smad4 during cancer progression has been shown
to contribute to the resistance of TGF-
-inhibited cell growth,
which is considered one mechanism for the TGF-
to be
tumor-promoting (1). However, it is less clear what other
mechanistic biomarkers can indicate the change of TGF-
sig-
naling in favor of tumor progression during tumorigenesis. Our
study showed that knockdown of ShcA isoforms enhanced,
whereas overexpression of p52/46ShcA attenuated, TGF-
-in-
hibited anchorage-dependent and independent-growth of
tumor cells, implicating ShcA as a negative mediator of TGF-
-induced tumor suppression. For the Smad-dependent
TGF-
signaling, TGF-
-mediated cell cycle arrest involves a
Smad-dependent induction of cyclin-dependent kinase inhibi-
tors p15 and p21. On the other hand, the activation of MAPK/
ERK signaling is pivotal for the TGF-
-induced epithelial-mes-
enchymal transition and cell migration, both of which are
considered important steps of prometastatic progression (1).
Thus, it is conceivable that the mechanism by which p52/
46ShcA regulates TGF-
-induced cell migration or growth
inhibition is cell context-dependent. For DU145, PC-3, and
MCF-10A cells, we found that the cells are sensitive to TGF-
as evidenced by the increased p-Smad2/3 and that manipulat-
ing ShcA enhanced TGF-
-induced cell migration accompa-
nied by increased TGF-
/ShcA/ERK signaling and decreased
TGF-
/Smad signaling. Because TGF-
/Smad-dependent
stimulation of p21 expression has been observed in these cells
and p21 was shown to be up-regulated by the Smad3-FoxO
Forkhead transcription factor transcriptional complex (50), it is
likely that the altered cell growth we observed was due to the
manipulation of p52/46ShcA and altered Smad/FOXO/p21
cascade. On the other hand, in cancer cells with Smad4 muta-
tion or aberrant regulation of p21 and/or p15, the alteration of
Smad signaling due to a change in ShcA activity may not result
in TGF-
-inhibited cell growth.
It is believed that p66ShcA is not involved in the activation
of MAPK signaling (51). In fact, p66ShcA was shown to
FIGURE 6. ShcA alters TGF-
-mediated Smad and ERK signaling and activ-
ity in untransformed human cells. MCF-10A cells were transfected with a
pan-ShcA siRNA or a p52/46ShcA expression plasmid as shown in Aand B,
respectively. Seventy-two hours post-transfection, immunoblotting analysis
for the indicated gene expression was performed using the transfected cells
treated with or without 2 ng/ml TGF-
1 for 40 min. Cand D, TGF-
1 (2 ng/ml)-
induced cell growth inhibition was measured with the MTT assay in cells 72 h
after transfection. Data are presented as the percentage of TGF-
-induced
growth inhibition relative to the untreated cells. Data represent mean S.E.
from four wells. Eand F, cell migration was assayed in cells 48 h after transfec-
tion. Cells were plated at 70,000 cells/insert with or without TGF-
1 (2 ng/ml)
for 18 h. Data are presented as the fold change of TGF-
-induced migration
relative to untreated cells. Data represent mean S.E. from three inserts. *,
p0.05; **, p0.01. G, immunoblotting analysis of the indicated gene
expression was performed in MCF-10A cells with or without RII knockdown
after TGF-
1 treatment for 40 min. H, immunoblotting analysis for the indi-
cated gene expression was performed in MCF-7 cells with or without DNRII
expression. T-ERK, total ERK; T-Smad2/3, total Smad2/3; T-p52ShcA, total
p52ShcA; con, control. Error bars represent S.E.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44031
antagonize ERK activation via negative regulation of c-fos
promoter activity (52) and to be mainly involved in oxidative
stress signaling (53). Additionally, the role of p66ShcA in
tumor metastasis is controversial (54, 55). Therefore, the
reduction of the p-ERK level after ShcA knockdown with the
pan-ShcA siRNA was likely due to the depletion of
p52/46ShcA.
Consistent with a previous study demonstrating that
knockdown of ShcA or a dominant-negative form of ShcA
inhibited TGF-
-induced cell migration and invasion of
murine breast cancer cells bearing activated Neu/ErbB-2
receptor (56), our study showed that ShcA depletion-caused
down-regulation of ERK signaling resulted in the attenua-
tion of TGF-
-induced cell migration in human untrans-
formed and transformed cells. On the other hand, we found
that increased ShcA/ERK signaling augmented TGF-
-in-
duced cell migration. We further showed that in DU145 cells
the restored activation of p52/46ShcA/ERK signaling com-
pletely switched the role of TGF-
from suppression to pro-
motion of cell migration. ERK signaling-induced cell migra-
tion is a key sign of tumor progression (9); therefore, our
results provide a novel concept that the ShcA/ERK pathway
acts as a pivotal driver of the oncogenic role of TGF-
.Inan
unpublished study,
3
we found that knockdown of Smad4 in
MCF-10A cells moderately reduced TGF-
-induced cell
migration in comparison with control siRNA-transfected
cells, indicating that Smad-dependent signaling partially
contributes to TGF-
-induced cell migration. Here, we
showed that the ShcA-mediated increase of TGF-
-induced
migration was in the presence of down-regulated Smad sig-
naling. These observations suggest that ShcA-enhanced ERK
signaling appears to play a dominant role in TGF-
-induced
cell migration. Thus, the down-regulation of p-Smad2/3 is
not likely necessary for the ShcA/ERK signaling-mediated
increase of TGF-
-induced migration.
In our study, we found that mp53 positively correlates with
phosphorylation of R-Smads, whereas it negatively correlates
with phosphorylation of p52ShcA and ERK in a panel of
mp53-expressing breast and prostate cancer cell lines as well
as in mp53 mouse models. We demonstrated that ShcA
tipped the balance of the role of TGF-
in tumor progression
as evidenced by enhanced migration and attenuated prolif-
eration in the presence of TGF-
. The earlier clinical studies
reported that high levels of tyrosine-phosphorylated p52/
46Shc in primary tumors might actually identify patients
with malignant breast tumors (17, 18). Increased levels of
TGF-
are also positively related to poor outcome of human
cancer (1). Our finding that overexpression of ShcA
enhanced the tumor-promoting activity of TGF-
provides
insight in linking these two biomarkers to the disease out-
come. In the preclinical models, we and others have shown
that anti-TGF-
therapies are potent in treating cancer in
which TGF-
functions as a tumor promoter (3–5). How-
ever, the lack of knowledge about when TGF-
is switched
from a tumor suppressor to a tumor promoter is a major
3
S. Lin, L. Yu, J. Yang, and L.-Z. Sun, unpublished data.
FIGURE 7. Mutant p53 correlates with higher level of Smad-dependent
signaling and lower level of ShcA/ERK signaling in human cancer cell
lines and genetically altered mouse models. A, immunoblotting analysis
for the indicated gene expression was performed in BT-20, BT474, 22Rv1, and
DU145 cells 72 h after transfection of p53 siRNA 1 or control siRNA. B, immu-
noblotting analysis for the indicated gene expression was performed in the
tissue lysates from mammary gland of p53
R172H/R172H
p
un/un
(p53 R172H; n2)
and p53
R172P/R172P
p
un/un
(p53 R172P; n1) mice and the littermate p53
/
p
un/un
(WTp53; n5). The scatter plot figures under the immunoblots are the
total ERK-normalized p-ERK levels and the total Smad2/3-normalized
p-Smad3 levels quantified with ImageJ software. C, immunoblotting analysis
for the indicated gene expression was performed in the tissue lysates from
mammary tumors of MMTV-Wnt1-WTp53
/
(WTp53; n8), MMTV-Wnt1-
p53
R172H/R172H
(p53 R172H/H; n4), and MMTV-Wnt1-p53
R172H/O
(p53
R172H/O; n4) mice. The scatter plot figures under the immunoblots are the
total p52ShcA-normalized p-p52ShcA levels, total ERK-normalized p-ERK lev-
els, and the total Smad2/3-normalized p-Smad3 levels for each sample quan-
tified with ImageJ software. *, p0.05. T-ERK, total ERK; T-Smad2/3, total
Smad2/3; T-p52ShcA, total p52ShcA.
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44032 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
obstacle for the utility of TGF-
antagonists for treating
metastatic carcinomas. Thus, our study indicates that the
TGF-
-dependent increase of tyrosine-phosphorylated p52/
46ShcA may be a promising biomarker for an effective anti-
TGF-
cancer therapy. Biomarkers identified by this type of
research may help cancer patients benefit from anti-TGF-
therapy in blocking tumor progression in the future.
Acknowledgments—We thank Dr. Andrew Hinck (University of Texas
Health Science Center at San Antonio, UTHSCSA) for the recombi-
nant TGF-
1, Dr. Michael G. Brattain (University of Nebraska,
Omaha) for the MCF-7 cell lines, Dr. Bert Vogelstein (Johns Hopkins
Oncology Center, Baltimore, Maryland) for the pSBE4-Luc plasmid,
Dr. Harikrishna Nakshatri (Indiana University, Indianapolis, IN) for
the p53 R175H plasmid, Dr. Zhi-Min Yuan (UTHSCSA) for the p53
R273H plasmid, Dr. Arnold Levine (Institute for Advanced Study,
Princeton, NJ) for the human WTp53 plasmid and Dr. Kun-Liang
Guan (University of California, San Diego) for the constitutively
active MEK1 plasmid.
REFERENCES
1. Massagué, J. (2008) Cell 134, 215–230
2. Sun, L. (2004) Front. Biosci. 9, 1925–1935
3. Bandyopadhyay, A., Agyin, J. K., Wang, L., Tang, Y., Lei, X., Story, B. M.,
Cornell, J. E., Pollock, B. H., Mundy, G. R., and Sun, L. Z. (2006) Cancer Res.
66, 6714–6721
4. Muraoka, R. S., Dumont, N., Ritter, C. A., Dugger, T. C., Brantley, D. M.,
Chen, J., Easterly, E., Roebuck, L. R., Ryan, S., Gotwals, P. J., Koteliansky,
V., and Arteaga, C. L. (2002) J. Clin. Investig. 109, 1551–1559
5. Yang, Y. A., Dukhanina, O., Tang, B., Mamura, M., Letterio, J. J.,
MacGregor, J., Patel, S. C., Khozin, S., Liu, Z. Y., Green, J., Anver, M. R.,
Merlino, G., and Wakefield, L. M. (2002) J. Clin. Investig. 109,
1607–1615
6. Manning, G., Whyte, D. B., Martinez, R., Hunter, T., and Sudarsanam, S.
(2002) Science 298, 1912–1934
7. Mulder, K. M. (2000) Cytokine Growth Factor Rev. 11, 23–35
8. Davies, M., Robinson, M., Smith, E., Huntley, S., Prime, S., and Paterson, I.
(2005) J. Cell. Biochem. 95, 918–931
9. Reddy, K. B., Nabha, S. M., and Atanaskova, N. (2003) Cancer Metastasis
Rev. 22, 395–403
10. Lee, M. K., Pardoux, C., Hall, M. C., Lee, P. S., Warburton, D., Qing, J.,
Smith, S. M., and Derynck, R. (2007) EMBO J. 26, 3957–3967
11. Ravichandran, K. S. (2001) Oncogene 20, 6322–6330
12. Luzi, L., Confalonieri, S., Di Fiore, P. P., and Pelicci, P. G. (2000) Current
Opin. Genet. Dev. 10, 668–674
13. Ventura, A., Maccarana, M., Raker, V. A., and Pelicci, P. G. (2004) J. Biol.
Chem. 279, 2299–2306
14. van der Geer, P., Wiley, S., Gish, G. D., and Pawson, T. (1996) Curr. Biol. 6,
1435–1444
15. Lotti, L. V., Lanfrancone, L., Migliaccio, E., Zompetta, C., Pelicci, G., Sal-
cini, A. E., Falini, B., Pelicci, P. G., and Torrisi, M. R. (1996) Mol. Cell. Biol.
16, 1946–1954
16. Ursini-Siegel, J., and Muller, W. J. (2008) Cell Cycle 7, 1936–1943
17. Davol, P. A., Bagdasaryan, R., Elfenbein, G. J., Maizel, A. L., and Frackelton,
A. R., Jr. (2003) Cancer Res. 63, 6772–6783
18. Frackelton, A. R., Jr., Lu, L., Davol, P. A., Bagdasaryan, R., Hafer, L. J., and
Sgroi, D. C. (2006) Breast Cancer Res. 8, R73
19. Adorno, M., Cordenonsi, M., Montagner, M., Dupont, S., Wong, C.,
Hann, B., Solari, A., Bobisse, S., Rondina, M. B., Guzzardo, V., Parenti,
A. R., Rosato, A., Bicciato, S., Balmain, A., and Piccolo, S. (2009) Cell 137,
87–98
20. Levine, A. J. (1997) Cell 88, 323–331
21. Navone, N. M., Troncoso, P., Pisters, L. L., Goodrow, T. L., Palmer, J. L.,
Nichols, W. W., von Eschenbach, A. C., and Conti, C. J. (1993) J. Natl.
Cancer Inst. 85, 1657–1669
22. Soussi, T., and Béroud, C. (2001) Nat. Rev. Cancer 1, 233–240
23. Wang, S. P., Wang, W. L., Chang, Y. L., Wu, C. T., Chao, Y. C., Kao, S. H.,
Yuan, A., Lin, C. W., Yang, S. C., Chan, W. K., Li, K. C., Hong, T. M., and
Yang, P. C. (2009) Nat. Cell Biol. 11, 694–704
24. Oren, M., and Rotter, V. (2010) Cold Spring Harb. Perspect. Biol. 2,
a001107
25. Dong, P., Tada, M., Hamada, J., Nakamura, A., Moriuchi, T., and Sakuragi,
N. (2007) Clin. Exp. Metastasis 24, 471–483
26. Kalo, E., Buganim, Y., Shapira, K. E., Besserglick, H., Goldfinger, N., Weisz,
L., Stambolsky, P., Henis, Y. I., and Rotter, V. (2007) Mol. Cell. Biol. 27,
8228–8242
27. Ko, Y., Koli, K. M., Banerji, S. S., Li, W., Zborowska, E., Willson, J. K.,
Brattain, M. G., and Arteaga, C. L. (1998) Int. J. Oncol. 12, 87–94
28. Bandyopadhyay, A., Wang, L., López-Casillas, F., Mendoza, V., Yeh, I. T.,
and Sun, L. (2005) Prostate 63, 81–90
29. Singh, J., Chuaqui, C. E., Boriack-Sjodin, P. A., Lee, W. C., Pontz, T.,
Corbley, M. J., Cheung, H. K., Arduini, R. M., Mead, J. N., Newman, M. N.,
Papadatos, J. L., Bowes, S., Josiah, S., and Ling, L. E. (2003) Bioorg. Med.
Chem. Lett. 13, 4355–4359
30. Lei, X., Bandyopadhyay, A., Le, T., and Sun, L. (2002) Oncogene 21,
7514–7523
31. Zawel, L., Dai, J. L., Buckhaults, P., Zhou, S., Kinzler, K. W., Vogelstein, B.,
and Kern, S. E. (1998) Mol. Cell 1, 611–617
32. Chen, C., Wang, X. F., and Sun, L. (1997) J. Biol. Chem. 272, 12862–12867
33. Lang, G. A., Iwakuma, T., Suh, Y. A., Liu, G., Rao, V. A., Parant, J. M.,
Valentin-Vega, Y. A., Terzian, T., Caldwell, L. C., Strong, L. C., El-Naggar,
A. K., and Lozano, G. (2004) Cell 119, 861–872
34. Liu, G., Parant, J. M., Lang, G., Chau, P., Chavez-Reyes, A., El-Naggar,
A. K., Multani, A., Chang, S., and Lozano, G. (2004) Nat. Genet. 36, 63–68
35. Aubrecht, J., Secretan, M. B., Bishop, A. J., and Schiestl, R. H. (1999) Car-
cinogenesis 20, 2229–2236
36. Post, S. M., Quintás-Cardama, A., Terzian, T., Smith, C., Eischen, C. M.,
and Lozano, G. (2010) Oncogene 29, 1260–1269
37. Isaacs, W. B., Carter, B. S., and Ewing, C. M. (1991) Cancer Res. 51,
4716–4720
38. Liu, X., Sun, Y., Constantinescu, S. N., Karam, E., Weinberg, R. A., and
Lodish, H. F. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 10669–10674
39. Ramachandra, M., Atencio, I., Rahman, A., Vaillancourt, M., Zou, A.,
Avanzini, J., Wills, K., Bookstein, R., and Shabram, P. (2002) Cancer Res.
62, 6045–6051
40. Lei, X., Yang, J., Nichols, R. W., and Sun, L. Z. (2007) Exp. Cell Res. 313,
1687–1695
41. Bartek, J., Iggo, R., Gannon, J., and Lane, D. P. (1990) Oncogene 5, 893–899
42. van Bokhoven, A., Varella-Garcia, M., Korch, C., Johannes, W. U., Smith,
E. E., Miller, H. L., Nordeen, S. K., Miller, G. J., and Lucia, M. S. (2003)
Prostate 57, 205–225
43. Berglind, H., Pawitan, Y., Kato, S., Ishioka, C., and Soussi, T. (2008) Cancer
Biol. Ther. 7, 699–708
44. Vogelstein, B., and Kinzler, K. W. (1993) Trends Genet. 9, 138–141
45. Buganim, Y., Solomon, H., Rais, Y., Kistner, D., Nachmany, I., Brait, M.,
Madar, S., Goldstein, I., Kalo, E., Adam, N., Gordin, M., Rivlin, N., Kogan,
I., Brosh, R., Sefadia-Elad, G., Goldfinger, N., Sidransky, D., Kloog, Y., and
Rotter, V. (2010) Cancer Res. 70, 2274–2284
46. Muller, P. A., Caswell, P. T., Doyle, B., Iwanicki, M. P., Tan, E. H., Karim, S.,
Lukashchuk, N., Gillespie, D. A., Ludwig, R. L., Gosselin, P., Cromer, A.,
Brugge, J. S., Sansom, O. J., Norman, J. C., and Vousden, K. H. (2009) Cell
139, 1327–1341
47. Signoretti, S., Waltregny, D., Dilks, J., Isaac, B., Lin, D., Garraway, L., Yang,
A., Montironi, R., McKeon, F., and Loda, M. (2000) Am. J. Pathol.157,
1769–1775
48. Okada, S., Yamauchi, K., and Pessin, J. E. (1995) J. Biol. Chem. 270,
20737–20741
49. Huse, M., Chen, Y. G., Massagué, J., and Kuriyan, J. (1999) Cell 96,
425–436
50. Seoane, J., Le, H. V., Shen, L., Anderson, S. A., and Massagué, J. (2004) Cell
117, 211–223
51. Jackson, J. G., Yoneda, T., Clark, G. M., and Yee, D. (2000) Clin. Cancer
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
DECEMBER 23, 2011VOLUME 286 •NUMBER 51 JOURNAL OF BIOLOGICAL CHEMISTRY 44033
Res. 6, 1135–1139
52. Migliaccio, E., Mele, S., Salcini, A. E., Pelicci, G., Lai, K. M., Superti-Furga,
G., Pawson, T., Di Fiore, P. P., Lanfrancone, L., and Pelicci, P. G. (1997)
EMBO J. 16, 706–716
53. Migliaccio, E., Giorgio, M., Mele, S., Pelicci, G., Reboldi, P., Pandolfi, P. P.,
Lanfrancone, L., and Pelicci, P. G. (1999) Nature 402, 309–313
54. Ma, Z., Liu, Z., Wu, R. F., and Terada, L. S. (2010) Oncogene 29,
5559–5567
55. Rajendran, M., Thomes, P., Zhang, L., Veeramani, S., and Lin, M. F. (2010)
Cancer Metastasis Rev. 29, 207–222
56. Northey, J. J., Chmielecki, J., Ngan, E., Russo, C., Annis, M. G., Muller,
W. J., and Siegel, P. M. (2008) Mol. Cell. Biol. 28, 3162–3176
Mutant p53 and ShcA Alter Oncogenic Role of TGF-
44034 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 286 •NUMBER 51• DECEMBER 23, 2011
... This was apparent in the level of Smad3 activation and nuclear import in response to TGF-β and the activation of TGF-β/Smad target gene expression. These findings are consistent with results on the control of TGF-β signaling by a mutant p53 in human prostate carcinoma cell lines, showing that Smad2/3 activation is decreased upon overexpression of ShcA and enhanced when ShcA levels are reduced [60]. The increased Smad3 activation when ShcA expression is down-regulated did not result from an increase in cell surface TGF-β receptors or from generation of active TGF-β by the cells, but from enhanced Smad3 recruitment to the activated TβRI. ...
... Moreover, mammary carcinoma cells that express a mutant ShcA lacking a functional phosphoTyr-binding PTB domain show increased expression of mesenchymal fibronectin and α5β1 integrin [75], arguing that with impaired ShcA function the carcinoma cells might transition toward a mesenchymal phenotype. Distinct functions of p66SchA, with some antagonizing those of p52SchA [1,2,60] may explain the complex roles of SchA in controlling epithelial plasticity of carcinomas. Accordingly, analyses of breast cancer cell lines suggest differential regulation of p52ShcA and p66ShcA expression [76]. ...
... Although in our nontransformed epithelial cells the ShcA expression level defines the sensitivity to EMT through modulation of TGF-β signaling, crosstalk of ShcA with oncogenic signaling may confer a more complex role of ShcA in the epithelial plasticity of cancer cells. Indeed, ShcA cooperates with Neu/ErbB2 signaling in the control of cell motility and invasion in transformed epithelial cells [77], through effects on focal adhesion turnover [5,77], and increased p52/46ShcA levels enhance migration of prostate carcinoma cells [60]. Additionally, p66ShcA overexpression promotes EMT in ErbB2-driven breast cancer cells, through up-regulated activation of the c-Met receptor by its ligand hepatocyte growth factor (HGF) [76]. ...
Article
Full-text available
Epithelial–mesenchymal transition (EMT) is a normal cell differentiation event during development and contributes pathologically to carcinoma and fibrosis progression. EMT often associates with increased transforming growth factor-β (TGF-β) signaling, and TGF-β drives EMT, in part through Smad-mediated reprogramming of gene expression. TGF-β also activates the Erk MAPK pathway through recruitment and Tyr phosphorylation of the adaptor protein ShcA by the activated TGF-β type I receptor. We found that ShcA protects the epithelial integrity of nontransformed cells against EMT by repressing TGF-β-induced, Smad-mediated gene expression. p52ShcA competed with Smad3 for TGF-β receptor binding, and down-regulation of ShcA expression enhanced autocrine TGF-β/Smad signaling and target gene expression, whereas increased p52ShcA expression resulted in decreased Smad3 binding to the TGF-β receptor, decreased Smad3 activation, and increased Erk MAPK and Akt signaling. Furthermore, p52ShcA sequestered TGF-β receptor complexes to caveolin-associated membrane compartments, and reducing ShcA expression enhanced the receptor localization in clathrin-associated membrane compartments that enable Smad activation. Consequently, silencing ShcA expression induced EMT, with increased cell migration, invasion, and dissemination, and increased stem cell generation and mammosphere formation, dependent upon autocrine TGF-β signaling. These findings position ShcA as a determinant of the epithelial phenotype by repressing TGF-β-induced Smad activation through differential partitioning of receptor complexes at the cell surface.
... Moreover, TGF-β1 supplement significantly restored Erk1/2 and SMAD3 phosphorylation levels individually or simultaneously which is inhibited by LAE supplement in three subtypes of breast cancer cells (Fig. 3d). Mutant p53 mediates the TGF-β1 signaling pathway via Erk1/2 as well as SMAD3 signals, and the p53 in all of the three subtypes of breast cancer cells are mutant [30][31][32][33]. We examined the mutant p53 protein level in three subtypes of breast cancer cells with LAE supplement. ...
... Interestingly, although LAE treatment significantly reduced the phosphorylation levels of Erk1/2 and SMAD3, the expression of TGF-β1 protein which regulated them both did not change significantly in MDA-MB-231 cells. This might be explained by the significantly decreased expression of mutant p53 in MDA-MB-231 cells because of the essential role of mutant p53 in mediating the TGF-β1 signaling pathway [30]. Taken together, our findings strongly suggested that LAE inhibit EMT by primarily targeting the TGF-β1 pathway, and the underlying mechanism of LAE-suppressed cell migration is differently within the three cell lines. ...
Article
Full-text available
Background Patients with estrogen receptor negative (ER⁻) breast cancer have poor prognosis due to high rates of metastasis. However, there is no effective treatment and drugs for ER⁻ breast cancer metastasis. Our purpose of this study was to evaluate the effect of lotus leaf alcohol extract (LAE) on the cell migration and metastasis of ER⁻ breast cancer. Methods The anti-migratory effect of LAE were analyzed in ER⁻ breast cancer cells including SK-BR-3, MDA-MB-231 and HCC1806 cell lines. Cell viability assay, wound-healing assay, RNA-sequence analysis and immunoblotting assay were used to evaluate the cytotoxicity and anti-migratory effect of LAE. To further investigate the inhibitory effect of LAE on metastasis in vivo, subcutaneous xenograft and intravenous injection nude mice models were established. Lung and liver tissues were analyzed by the hematoxylin and eosin staining and immunoblotting assay. Results We found that lotus LAE, not nuciferine, inhibited cell migration significantly in SK-BR-3, MDA-MB-231 and HCC1806 breast cancer cells, and did not affect viability of breast cancer cells. The anti-migratory effect of LAE was dependent on TGF-β1 signaling, while independent of Wnt signaling and autophagy influx. Intracellular H2O2 was involved in the TGF-β1-related inhibition of cell migration. LAE inhibited significantly the breast cancer cells metastasis in mice models. RNA-sequence analysis showed that extracellular matrix signaling pathways are associated with LAE-suppressed cell migration. Conclusions Our findings demonstrated that lotus leaf alcohol extract inhibits the cell migration and metastasis of ER⁻ breast cancer, at least in part, via TGF-β1/Erk1/2 and TGF-β1/SMAD3 signaling pathways, which provides a potential therapeutic strategy for ER⁻ breast cancer.
... Mutant p53 mediates the TGF-β1 signaling pathway via Erk1/2 as well as SMAD3 signals, and the p53 in all of the three subtypes of breast cancer cells are mutant [30][31][32][33]. We examined the mutant p53 protein level in three subtypes of breast cancer cells with LAE supplement. ...
... Interestingly, although LAE treatment signi cantly reduced the phosphorylation levels of Erk1/2 and SMAD3, the expression of TGF-β1 protein which regulated them both did not change signi cantly in MDA-MB-231 cells. This might be explained by the signi cantly decreased expression of mutant p53 in MDA-MB-231 cells because of the essential role of mutant p53 in mediating the TGF-β1 signaling pathway [30]. Taken together, our ndings strongly suggested that LAE inhibit EMT by primarily targeting the TGF-β1 pathway, and the underlying mechanism of LAE-suppressed cell migration is differently within the three cell lines. ...
Preprint
Full-text available
Background: Patients with estrogen receptor negative (ER-) breast cancer have poor prognosis due to high rates of metastasis. However, there is no effective treatment and drugs for ER⁻ breast cancer metastasis. Our purpose of this study was to evaluate the effect of lotus leaf alcohol extract (LAE) on the cell migration and metastasis of ER- breast cancer. Methods: The anti-migratory effect of LAE were analyzed in ER- breast cancer cells including SK-BR-3, MDA-MB-231 and HCC1806 cell lines. Cell viability assay, wound-healing assay, RNA-sequence analysis and immunoblotting assay were used to evaluate the cytotoxicity and anti-migratory effect of LAE. To further investigate the inhibitory effect of LAE on metastasis in vivo, subcutaneous xenograft and intravenous injection nude mice models were established. Lung and liver tissues were analyzed by the Hematoxylin & eosin staining and immunoblotting assay. Results: We found that lotus leaf alcohol extract (LAE), not nuciferine, inhibited cell migration significantly in SK-BR-3, MDA-MB-231 and HCC1806 breast cancer cells, and did not affect viability of breast cancer cells. The anti-migratory effect of LAE was dependent on TGF-β1 signaling, while independent of Wnt signaling and autophagy influx. Intracellular H2O2 was involved in the TGF-β1-related inhibition of cell migration. LAE inhibited significantly the breast cancer cells metastasis in mice models. RNA-sequence analysis showed that extracellular matrix signaling pathways are associated with LAE-suppressed cell migration. Conclusions: Our findings demonstrated that lotus leaf alcohol extract inhibits the cell migration and metastasis of ER- breast cancer, at least in part, via TGF-β1/Erk1/2 and TGF-β1/SMAD3 signaling pathways, which provides a potential therapeutic strategy for ER- breast cancer.
... Mutant p53 mediates the TGF-β1 signaling pathway via Erk1/2 as well as SMAD3 signals, and the p53 in all of the three subtypes of breast cancer cells are mutant [30][31][32][33]. We next examined the mutant p53 protein level in three subtypes of breast cancer cells with LAE supplement. ...
... Interestingly, although LAE treatment signi cantly reduced the phosphorylation levels of Erk1/2 and SMAD3, the expression of TGF-β1 protein which regulated them both did not change signi cantly in MDA-MB-231 cells. This might be explained by the signi cantly decreased expression of mutant p53 in MDA-MB-231 cells because of the essential role of mutant p53 in mediating the TGF-β1 signaling pathway [30]. Taken together, our ndings strongly suggested that LAE inhibits EMT by primarily targeting the TGF-β1 pathway, and the underlying mechanism of LAE-suppressed cell migration is differential within the three cell lines. ...
Preprint
Full-text available
Background: Patients with estrogen receptor negative (ER-) breast cancer have poor prognosis due to high rates of metastasis. However, there is no effective treatment and drugs for ER⁻ breast cancer metastasis. Our purpose of this study was to evaluate the effect of lotus leaf alcohol extract (LAE) on the cell migration and metastasis of ER- breast cancer. Methods: The anti-migratory effect of LAE were analyzed in ER- breast cancer cells including SK-BR-3, MDA-MB-231 and HCC1806 cell lines. Cell viability assay, wound-healing assay, RNA-sequence analysis and immunoblotting assay were used to evaluate the cytotoxicity and anti-migratory effect of LAE. To further investigate the inhibitory effect of LAE on metastasis in vivo, subcutaneous xenograft and intravenous injection nude mice models were established. Lung and liver tissues were analyzed by the Hematoxylin & eosin staining and immunoblotting assay. Results: We found that lotus leaf alcohol extract (LAE), not nuciferine, inhibited cell migration significantly in SK-BR-3, MDA-MB-231 and HCC1806 breast cancer cells, and did not affect viability of breast cancer cells. The anti-migratory effect of LAE was dependent on TGF-β1 signaling, while independent of Wnt signaling and autophagy influx. Intracellular H2O2 was involved in the TGF-β1-related inhibition of cell migration. LAE inhibited significantly the breast cancer cells metastasis in mice models. RNA-sequence analysis showed that extracellular matrix signaling pathways are associated with LAE-suppressed cell migration. Conclusions: Our findings demonstrated that lotus leaf alcohol extract inhibits the cell migration and metastasis of ER- breast cancer, at least in part, via TGF-β1/Erk1/2 and TGF-β1/SMAD3 signaling pathways, which provides a potential therapeutic strategy for ER- breast cancer.
... Since mutant p53 mediates the TGF-β1 signaling pathway via Erk1/2 as well as SMAD3 signals, and the p53 in all of the three subtypes of breast cancer cells are mutant [32][33][34][35]. To address this issue, we examined the mutant p53 protein level in three subtypes of breast cancer cells with LAE supplement. ...
... Interestingly, although LAE treatment significantly reduced the phosphorylation levels of Erk1/2 and SMAD3, the expression of TGF-β1 protein which regulated them both did not change significantly in MDA-MB-231 cells. This might be explained by the significantly decreased expression of mutant p53 in MDA-MB-231 cells because the essential effects of mutant p53 in mediating the TGF-β1 signaling pathway [32]. Taken together, our findings strongly suggested that LAE inhibit EMT by primarily targeting the TGF-β1 pathway, and the underlying mechanism of LAE-suppressed cell migration is differently dependent on cell lines. ...
Preprint
Full-text available
Background: Patients with estrogen receptor negative (ER-) breast cancer have poor prognosis because of their high rates of metastasis. However, there is no effective treatment and drugs for ER⁻ breast cancer metastasis. Our purpose of this study was to evaluate the effect and mechanism of lotus leaf alcohol extract (LAE) on the cell migration and metastasis of ER- breast cancer. Methods: The anti-migratory effect and mechanism of LAE were analysed in ER- breast cancer cells including SK-BR-3, MDA-MB-231 and HCC1806. Cell viability assay, wound-healing assay, RNA-sequence analysis and immunoblotting assay were applied in examining the cytotoxicity, anti-migratory effect and its possible pathways of LAE. To further investigate the inhibitory effect of LAE on metastasis in vivo, subcutaneous xenograft nude mice model and intravenous injection nude mice model were established. Lung and liver tissues were analysed by the Hematoxylin & eosin staining and immunoblotting assay. Results: We found that lotus leaf alcohol extract (LAE), not nuciferine, inhibited cell migration significantly in SK-BR-3, MDA-MB-231 and HCC1806 breast cancer cells, and did not change viability of breast cancer cells. The anti-migratory effect of LAE was dependent on TGF-β1 signaling, while independent of Wnt signaling and autophagy influx. Intracellular H2O2 participated in the TGF-β1-related inhibition of cell migration. LAE inhibited significantly the breast cancer cells metastasis in mice models. RNA-sequence analysis showed that extracellular matrix signaling pathways are associated with LAE-suppressed cell migration. Conclusions: Our findings demonstrated that lotus leaf alcohol extract inhibits the cell migration and metastasis of ER- breast cancer via TGF-β1/Erk1/2 and TGF-β1/SMAD3 signaling, which provides a potential therapeutic strategy for ER- breast cancer.
... Although a variety of types of cells can transform into myofibroblasts, fibroblasts are thought to be the dominant position of myofibroblasts transition [9]. In multiple fibrogenic factors related to the pathogenesis of EF, transforming growth factor (TGF)-β has been suggested to be a major source for myofibroblast accumulation via regulating Smad2/3 and PI3K /Akt signaling pathways [10,11]. Clearly, EF remains a huge therapeutic challenge and there is an urgent need to develop effective anti-fibrotic drugs. ...
... Moreover, and in contrast to PDAC cells utilized in our previous study 18 , NKC cells harbor wild-type expression of p53. Mutation of this master tumor-suppressor gene has been described to synergize with Ras activation in inducing the TGFβ switch by driving non-canonical/ Smad-independent TGFβ signaling 58,59 . Hence, our findings underscore the strong context-dependency of TGFβ-mediated transcription and function in PDAC progression. ...
Article
Full-text available
Given its aggressive tumor biology and its exceptional therapy resistance, pancreatic ductal adenocarcinoma (PDAC) remains a major challenge in cancer medicine and is characterized by a 5-year survival rate of <8%. At the cellular level, PDAC is largely driven by the activation of signaling pathways that eventually converge in altered, tumor-promoting transcription programs. In this study, we sought to determine the interplay between transforming growth factor β (TGFβ) signaling and activation of the inflammatory transcription factor nuclear factor of activated T cells (NFATc1) in the regulation of transcriptional programs throughout PDAC progression. Genome-wide transcriptome analysis and functional studies performed in primary PDAC cells and transgenic mice linked nuclear NFATc1 expression with pro-proliferative and anti-apoptotic gene signatures. Consistently, NFATc1 depletion resulted in downregulation of target genes associated with poor PDAC outcome and delayed pancreatic carcinogenesis in vivo. In contrast to previous reports and consistent with a concept of retained tumor suppressive TGFβ activity, even in established PDAC, TGFβ treatment reduced PDAC cell proliferation and promoted apoptosis even in the presence of oncogenic NFATc1. However, combined TGFβ treatment and NFATc1 depletion resulted in a tremendous abrogation of tumor-promoting gene signatures and functions. Chromatin studies implied that TGFβ-dependent regulators compete with NFATc1 for the transcriptional control of jointly regulated target genes associated with an unfavorable PDAC prognosis. Together, our findings suggest opposing consequences of TGFβ and NFATc1 activity in the regulation of pro-tumorigenic transcription programs in PDAC and emphasize the strong context-dependency of key transcription programs in the progression of this devastating disease.
... There is an exciting piece of evidence suggesting that mutant p53 enhances TGF-β/Smad signaling in prostate cancer cells. Mutant p53 expressing prostate cancer cells were more sensitive to TGF-β dose-dependently as evidenced by notably raised levels of phosphorylated levels of Smad2/3 (8). ...
Article
Research over the decades has gradually and sequentially shown that both intratumor heterogeneity and multifocality make prostate cancer difficult to target. Different challenges associated with generation of risk-stratification tools that correlate genomic landscape with clinical outcomes severely influence clinical efficacy of therapeutic strategies. Androgen receptor mediated signaling has gained great appreciation and rewiring of AR induced signaling cascade in absence of androgen, structural variants of AR have provided near complete resolution of genomic landscape and underlying mechanisms of prostate cancer. In this review we have attempted to provide an overview of most recent advancements in our knowledge related to different signaling cascades including TGF, SHH, Notch, JAK-STAT in prostate cancer progression and development.
... The final or net effect of TGF-b is determined by the sum of the signals activated. In this sense signaling through the Smad pathway or the Ras-ERK pathway has been shown to be dependent on p53 [28]. For example, in the context of embryo development, TGF-b-induced cytostasis has been linked to its capacity to bind p53. ...
Article
Full-text available
Background/aims: Defective tissue repair underlies renal tissue degeneration during chronic kidney disease (CKD) progression. Unbalanced presence of TGF-β opposes effective cell proliferation and differentiation processes, necessary to replace damaged epithelia. TGF-β also retains arrested cells in a fibrotic phenotype responsible for irreversible scarring. In order to identify prospective molecular targets to prevent the effect of TGF-β during CKD, we studied the signaling pathways responsible for the antiproliferative effect of this cytokine. Methods: Tubule epithelial HK2 and MDCK cells were treated with TGF-β (or not as control) to study cell proliferation (by MTT), cell signaling (by Western blot), cell cycle (by flow cytometry) and apoptosis (DNA fragmentation). Results: TGF-β fully activates the ALK-5 receptor pathway, whereas it has no effect on the ALK-1 and MAPK pathways in both HK2 and MDCK cells. Interestingly, TGF-β exerts an antiproliferative effect only on MDCK cells, through a cytostatic effect in G0/G1. Inhibition of the ALK-5 pathway with SB431542 prevents the cytostatic effect of TGF-β on MDCK cells. Conclusion: Activation of the ALK-5 pathway is not sufficient for the antiproliferative effect of TGF-β. The presence of undetermined permissive conditions or absence of undetermined inhibitory conditions seems to be necessary for this effect. The ALK-5 pathway appears to provide targets to modulate fibrosis, but further research is necessary to identify critical circumstances allowing or inhibiting its role at modulating tubule epithelial cell proliferation and tubule regeneration in the context of CKD progression.
... Treatment with combination TGFb-dasatinib did not have a significant effect on the non-canonical TGFb pathway intracellular mediators ERK, AKT, or p38 (S3 Figure). In addition, treatment with TGFb-1 did not result in an increase in Shc activation in A549 cells, as previously reported for other cell types [42,43], ...
Article
Full-text available
Lung cancer is the second most common cancer and the leading cause of cancer-related deaths. Despite recent advances in the development of targeted therapies, patients with advanced disease remain incurable, mostly because metastatic non-small cell lung carcinomas (NSCLC) eventually become resistant to tyrosine kinase inhibitors (TKIs). Kinase inhibitors have the potential for target promiscuity because the kinase super family is the largest family of druggable genes that binds to a common substrate (ATP). As a result, TKIs often developed for a specific purpose have been found to act on other targets. Drug affinity chromatography has been used to show that dasatinib interacts with the TGFβ type I receptor (TβR-I), a serine-threonine kinase. To determine the potential biological relevance of this association, we studied the combined effects of dasatinib and TGFβ on lung cancer cell lines. We found that dasatinib treatment alone had very little effect; however, when NSCLC cell lines were treated with a combination of TGFβ and dasatinib, apoptosis was induced. Combined TGFβ-1 + dasatinib treatment had no effect on the activity of Smad2 or other non-canonical TGFβ intracellular mediators. Interestingly, combined TGFβ and dasatinib treatment resulted in a transient increase in p-Smad3 (seen after 3 hours). In addition, when NSCLC cells were treated with this combination, the pro-apoptotic protein BIM was up-regulated. Knockdown of the expression of Smad3 using Smad3 siRNA also resulted in a decrease in BIM protein, suggesting that TGFβ-1 + dasatinib-induced apoptosis is mediated by Smad3 regulation of BIM. Dasatinib is only effective in killing EGFR mutant cells, which is shown in only 10% of NSCLCs. Therefore, the observation that wild-type EGFR lung cancers can be manipulated to render them sensitive to killing by dasatinib could have important implications for devising innovative and potentially more efficacious treatment strategies for this disease.
Article
Full-text available
While transforming growth factor β (TGFβ) type III receptor (RIII) is known to increase TGFβ1binding to its type II receptor (RII), the significance of this phenomenon is not known. We used human breast cancer MCF-7 cells to study the role of RIII in regulating autocrine TGFβ1activity because they express very little RIII and no detectable autocrine TGFβ activity. A tetracycline-repressible RIII expression vector was stably transfected into this cell line. Expression of RIII increased TGFβ1 binding to TGFβ type I receptor (RI) as well as RII. Treatment with tetracycline suppressed RIII expression and abolished TGFβ1 binding to RI and RII. Growth of RIII-transfected cells was reduced by 40% when plated at low density on plastic. This reduction was reversed by tetracycline treatment and was partially reversed by treatment with a TGFβ1neutralizing antibody. The activity of a TGFβ-responsive promoter construct when transiently transfected was more than 3-fold higher in the RIII-transfected cells than in the control cells. Treating the cells with tetracycline or the TGFβ1 neutralizing antibody also significantly attenuated the increased promoter activity. These results suggest that expression of RIII restored autocrine TGFβ1 activity in MCF-7 cells. The RIII-transfected cells were also much less clonogenic in soft agarose than the control cells indicating a reversion of progression. Thus, RIII may be essential for an optimal level of the autocrine TGFβ activity in some cells, especially in the transformed cells with reduced RII expression.
Article
Full-text available
Shc proteins are targets of activated tyrosine kinases and are implicated in the transmission of activation signals to Ras. The p46shc and p52shc isoforms share a C-terminal SH2 domain, a proline- and glycine-rich region (collagen homologous region 1; CH1) and a N-terminal PTB domain. We have isolated cDNAs encoding for a third Shc isoform, p66shc. The predicted amino acid sequence of p66shc overlaps that of p52shc and contains a unique N-terminal region which is also rich in glycines and prolines (CH2). p52shc/p46shc is found in every cell type with invariant reciprocal relationship, whereas p66shc expression varies from cell type to cell type. p66shc differs from p52shc/p46shc in its inability to transform mouse fibroblasts in vitro. Like p52shc/p46shc, p66shc is tyrosine-phosphorylated upon epidermal growth factor (EGF) stimulation, binds to activated EGF receptors (EGFRs) and forms stable complexes with Grb2. However, unlike p52shc/p46shc it does not increase EGF activation of MAP kinases, but inhibits fos promoter activation. The isolated CH2 domain retains the inhibitory effect of p66shc on the fos promoter. p52shc/p46shc and p66shc, therefore, appear to exert different effects on the EGFR-MAP kinase and other signalling pathways that control fos promoter activity. Regulation of p66shc expression might, therefore, influence the cellular response to growth factors.
Article
Full-text available
p66Shc, a 66kDa proto-oncogene Src homologous-collagen homologue (Shc) adaptor protein, is classically known in mediating receptor tyrosine kinase signaling and recently identified as a sensor to oxidative stress-induced apoptosis and as a longevity protein in mammals. The expression of p66Shc is decreased in mice and increased in human fibroblasts upon aging and in aging-related diseases, including prostate cancer. p66Shc protein level correlates with the proliferation of several carcinoma cells and can be regulated by steroid hormones. Recent advances point that p66Shc protein plays a role in mediating cross-talk between steroid hormones and redox signals by serving as a common convergence point in signaling pathways on cell proliferation and apoptosis. This article first reviews the unique function of p66Shc protein in regulating oxidative stress-induced apoptosis. Subsequently, we discuss its novel role in androgen-regulated prostate cancer cell proliferation and metastasis and the mechanism by which it mediates androgen action via the redox signaling pathway. The data together indicate that p66Shc might be a useful biomarker for the prognosis of prostate cancer and serve as an effective target for its cancer treatment. Keywordsp66Shc-Apoptosis-Redox-Cell proliferation-Metastasis-Prostate cancer
Article
Full-text available
In this study, we focus on the analysis of a previously identified cancer-related gene signature (CGS) that underlies the cross talk between the p53 tumor suppressor and Ras oncogene. CGS consists of a large number of known Ras downstream target genes that were synergistically upregulated by wild-type p53 loss and oncogenic H-Ras(G12V) expression. Here we show that CGS expression strongly correlates with malignancy. In an attempt to elucidate the molecular mechanisms underling the cooperation between p53 loss and oncogenic H-Ras(G12V), we identified distinguished pathways that may account for the regulation of the expression of the CGS. By knocking-down p53 or by expressing mutant p53, we revealed that p53 exerts its negative effect by at least two mechanisms mediated by its targets B-cell translocation gene 2 (BTG2) and activating transcription factor 3 (ATF3). Whereas BTG2 binds H-Ras(G12V) and represses its activity by reducing its GTP loading state, which in turn causes a reduction in CGS expression, ATF3 binds directly to the CGS promoters following p53 stabilization and represses their expression. This study further elucidates the molecular loop between p53 and Ras in the transformation process.
Article
Full-text available
The effect of p53-dependent cell-cycle arrest and senescence on Emu-myc-induced B-cell lymphoma development remains controversial. To address this question, we crossed Emu-myc mice with the p53(515C) mutant mouse, encoding the mutant p53R172P protein that retains the ability to activate the cell-cycle inhibitor and senescence activator p21. Importantly, this mutant lacks the ability to activate p53-dependent apoptotic genes. Hence, Emu-myc mice that harbor two p53(515C) alleles are completely defective for p53-dependent apoptosis. Both Emu-myc::p53(515C/515C) and Emu-myc::p53(515C/+) mice survive significantly longer than Emu-myc::p53(+/-) mice, indicating the importance of the p53-dependent non-apoptotic pathways in B-cell lymphomagenesis. In addition, the p53(515C) allele is deleted in several Emu-myc::p53(515C/+) lymphomas, further emphasizing the functionality of p53R172P in tumor inhibition. Lymphomas from both Emu-myc::p53(515C/515C) and Emu-myc::p53(515C/+) mice retain the ability to upregulate p21, resulting in cellular senescence. Senescence-associated beta-galactosidase (SA beta-gal) activity was observed in lymphomas from Emu-myc::p53(+/+), Emu-myc::p53(515C/515C) and Emu-myc::p53(515C /+) mice but not in lymphomas isolated from Emu-myc::p53(+/-) mice. Thus, in the absence of p53-dependent apoptosis, the ability of p53R172P to induce senescence leads to a significant delay in B-cell lymphoma development.
Article
Full-text available
Gene mutations in invertebrates have been identified that extend life span and enhance resistance to environmental stresses such as ultraviolet light or reactive oxygen species. In mammals, the mechanisms that regulate stress response are poorly understood and no genes are known to increase individual life span. Here we report that targeted mutation of the mouse p66shc gene induces stress resistance and prolongs life span. p66shc is a splice variant of p52shc/p46shc (ref. 2), a cytoplasmic signal transducer involved in the transmission of mitogenic signals from activated receptors to Ras. We show that: (1) p66shc is serine phosphorylated upon treatment with hydrogen peroxide (H2O2) or irradiation with ultraviolet light; (2) ablation of p66shc enhances cellular resistance to apoptosis induced by H2O2 or ultraviolet light; (3) a serine-phosphorylation defective mutant of p66shc cannot restore the normal stress response in p66shc-/- cells; (4) the p53 and p21 stress response is impaired in p66shc-/- cells; (5) p66shc-/- mice have increased resistance to paraquat and a 30% increase in life span. We propose that p66shc is part of a signal transduction pathway that regulates stress apoptotic responses and life span in mammals.
Article
The Shc protein family is characterized by the (CH2)–PTB–CH1–SH2 modularity. Its complexity increased during evolution from one locus in Drosophila (dShc), to at least three loci in mammals (shc, rai and sli). The three mammalian loci encode, because of alternative initiation codon usage and splicing pattern, at least six Shc-like proteins. Genetic and biological evidence indicates that the mammalian Shc isoforms regulate functions as diverse as growth (p52/p46Shc), apoptosis (p66Shc) and life-span (p66Shc). Available structure–function data and analysis of sequence similarities of Shc-like genes and proteins suggest complex diversification of Shc functions during evolution. Notably, Ras activation, the best-characterized Shc activity, appears to be a recent evolutionary acquisition.
Article
Normal tissue cells survive and proliferate only while anchored to solid substrate. Conversely, transformed cells both survive and proliferate following detachment, having lost attachment context through unclear mechanisms. p66(Shc) is a focal adhesion-associated protein that reports cell attachment through a RhoA-dependent mechanosensory test. We find that human small cell lung cancer (SCLC) cells and mouse Lewis lung carcinoma (LLC), which display aggressive metastatic behavior, lack both p66(Shc) and retinoblastoma (pRB) and bypass anoikis. Re-expression of p66(Shc) in these cells restores anoikis and provides striking protection from metastasis by LLC cells in vivo. Notably, knockdown of p66(Shc) in normal epithelial cells leads to unrestrained Ras activation, preventing anoikis through downstream suppression of RhoA but blocking proliferation in a pRB-dependent manner, thus mimicking oncogenic Ras. Conversely, LLC and SCLC cells display constitutive Ras activation necessary to bypass anoikis, which is reversed by re-expression of p66(Shc). p66(Shc) therefore coordinates Ras-dependent control of proliferation and anchorage sensation, which can be defeated in the evolution of highly metastatic tumors by combined loss of both p66(Shc) and pRB.
Article
In its wild-type form, p53 is a major tumor suppressor whose function is critical for protection against cancer. Many human tumors carry missense mutations in the TP53 gene, encoding p53. Typically, the affected tumor cells accumulate excessive amounts of the mutant p53 protein. Various lines of evidence indicate that, in addition to abrogating the tumor suppressor functions of wild-type p53, the common types of cancer-associated p53 mutations also endow the mutant protein with new activities that can contribute actively to various stages of tumor progression and to increased resistance to anticancer treatments. Collectively, these activities are referred to as mutant p53 gain-of-function. This article addresses the biological manifestations of mutant p53 gain-of-function, the underlying molecular mechanisms, and their possible clinical implications.
Article
p53 is a tumor suppressor protein whose function is frequently lost in cancers through missense mutations within the Tp53 gene. This results in the expression of point-mutated p53 proteins that have both lost wild-type tumor suppressor activity and show gain of functions that contribute to transformation and metastasis. Here, we show that mutant p53 expression can promote invasion, loss of directionality of migration, and metastatic behavior. These activities of p53 reflect enhanced integrin and epidermal growth factor receptor (EGFR) trafficking, which depends on Rab-coupling protein (RCP) and results in constitutive activation of EGFR/integrin signaling. We provide evidence that mutant p53 promotes cell invasion via the inhibition of TAp63, and simultaneous loss of p53 and TAp63 recapitulates the phenotype of mutant p53 in cells. These findings open the possibility that blocking alpha5/beta1-integrin and/or the EGF receptor will have therapeutic benefit in mutant p53-expressing cancers.