ArticlePDF AvailableLiterature Review

Clinical Overview of MDM2/X-Targeted Therapies

Frontiers
Frontiers in Oncology
Authors:

Abstract and Figures

MDM2 and MDMX are the primary negative regulators of p53, which under normal conditions, maintain low intracellular levels of p53 by targeting it to the proteasome for rapid degradation, and inhibiting its transcriptional activity. Both MDM2 and MDMX function as powerful oncogenes and are commonly over-expressed in some cancers, including sarcoma (~20%), and breast cancer (~15%). In contrast to tumours that are p53 mutant, whereby the current therapeutic strategy is restore the normal active conformation of p53, MDM2 and MDMX represent logical therapeutic targets in cancer for increasing wild type (WT) p53 expression and activities. Recent preclinical studies suggest that there may also be situations that MDM2/X inhibitors could be used in p53 mutant tumours. Since the discovery of nutlin-3a, the first in a class of small molecule MDM2 inhibitors that binds to the hydrophobic cleft in the N-terminus of MDM2, preventing its association with p53, there is now an extensive list of related compounds. In addition, a new class of stapled-peptides that can target both MDM2 and MDMX have also been developed. Importantly, preclinical modeling, which have demonstrated effective in vitro and in vivo killing of WT p53 cancer cells, have now been translated into early clinical trials allowing better assessment of their biological effects and toxicities in patients. In this overview, we will review the current MDM2 and MDMX targeted therapies in development, focusing particularly on compounds that have entered into early phase clinical trials. We will highlight the challenges pertaining to predictive biomarkers for and toxicities associated with these compounds, as well as identify potential combinatorial strategies to enhance its anti-cancer efficacy
This content is subject to copyright.
January 2016 | Volume 6 | Article 71
MINI REVIEW
published: 27 January 2016
doi: 10.3389/fonc.2016.00007
Frontiers in Oncology | www.frontiersin.org
Edited by:
Haining Yang,
University of Hawaii Cancer Center,
USA
Reviewed by:
Takaomi Sanda,
National University of Singapore,
Singapore
Giovanni Gaudino,
University of Hawaii Cancer Center,
USA
*Correspondence:
Elgene Lim
e.lim@garvan.org.au
Andrew Burgess and Kee Ming Chia
contributed equally to this work.
Specialty section:
This article was submitted to
Molecular and Cellular Oncology,
a section of the journal
Frontiers in Oncology
Received: 03December2015
Accepted: 11January2016
Published: 27January2016
Citation:
BurgessA, ChiaKM, HauptS,
ThomasD, HauptY and LimE (2016)
Clinical Overview of
MDM2/X-Targeted Therapies.
Front. Oncol. 6:7.
doi: 10.3389/fonc.2016.00007
Clinical Overview of
MDM2/X-Targeted Therapies
Andrew Burgess1,2† , Kee Ming Chia1† , Sue Haupt3 , David Thomas1,2 , Ygal Haupt3 and
Elgene Lim1,2*
1 The Kinghorn Cancer Centre, Garvan Institute of Medical Research, Sydney, NSW, Australia, 2 Faculty of Medicine,
St. Vincent’s Clinical School, UNSW Australia, Sydney, NSW, Australia, 3 The Sir Peter MacCallum Department of Oncology,
the University of Melbourne, Melbourne, VIC, Australia
MDM2 and MDMX are the primary negative regulators of p53, which under normal con-
ditions maintain low intracellular levels of p53 by targeting it to the proteasome for rapid
degradation and inhibiting its transcriptional activity. Both MDM2 and MDMX function
as powerful oncogenes and are commonly over-expressed in some cancers, including
sarcoma (~20%) and breast cancer (~15%). In contrast to tumors that are p53 mutant,
whereby the current therapeutic strategy restores the normal active conformation of p53,
MDM2 and MDMX represent logical therapeutic targets in cancer for increasing wild-
type (WT) p53 expression and activities. Recent preclinical studies suggest that there
may also be situations that MDM2/X inhibitors could be used in p53 mutant tumors.
Since the discovery of nutlin-3a, the rst in a class of small molecule MDM2 inhibitors
that binds to the hydrophobic cleft in the N-terminus of MDM2, preventing its association
with p53, there is now an extensive list of related compounds. In addition, a new class
of stapled peptides that can target both MDM2 and MDMX have also been developed.
Importantly, preclinical modeling, which has demonstrated effective invitro and invivo
killing of WT p53 cancer cells, has now been translated into early clinical trials allowing
better assessment of their biological effects and toxicities in patients. In this overview, we
will review the current MDM2- and MDMX-targeted therapies in development, focusing
particularly on compounds that have entered into early phase clinical trials. We will
highlight the challenges pertaining to predictive biomarkers for and toxicities associated
with these compounds, as well as identify potential combinatorial strategies to enhance
its anti-cancer efcacy.
Keywords: p53, MDM2, MDMX, cancer therapy, nutlin
INTRODUCTION: RATIONALE FOR TARGETING THE p53
PATHWAY
e tumor suppressor protein p53, nominated “the guardian of the genome,” is mutated in ~50%
of all human cancers. However, the incidence of p53 mutations diers signicantly between cancer
types, ranging from near universal mutation (~96%) in serous ovarian cancer to rare occurrence
(<10%) in thyroid cancer (Figure 1A). is disparity provides therapeutic opportunities for
targeting cancers with p53 wild-type (WT), in a distinct manner from those with p53 mutant can-
cers. Several preclinical studies have demonstrated that reconguration of mutant, to its normal,
FIGURE 1 | Rationale for targeting p53 in cancers. (A) Frequency of alterations are shown with mutation (green), deletion (blue), amplication (red), and
combination of alterations (gray) in p53, MDM2, and MDMX in cancers derived from cBioPortal (5) (http://www.cbioportal.org). Insert shows the mutual exclusivity
observed between MDM2 expression and p53 deletion in sarcomas. (B) Schematic representation of inhibitors in clinical trials (yellow box) or in preclinical studies
(blue box) targeting the p53–MDM2/X axis. Compounds are either small molecules (green circle) or peptide (blue circle).
January 2016 | Volume 6 | Article 72
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
active WT p53 conformation, restores apoptosis and promotes
tumor regression (13). erapeutic targeting of mutant p53,
using small molecule drugs, is in the most advanced state for
PRIMA-1, and its derivative PRIMA-1MET, an approach which
restores the normal, active conformation of p53, which has been
previously explored in depth by Wiman and coworkers (4). In
the current review, we focus on therapies that target MDM2
and MDMX as a means of increase the stability of WT p53 and
the consequences for patients with either WT p53 or mutant
cancer cells.
January 2016 | Volume 6 | Article 73
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
Regulation of p53 Stability by MDM2 and
MDMX
e primary response to a variety of cellular insults and stresses is
to concurrently activate and stabilize p53 within the cell. Activated
p53 then drives a vast transcriptional program that arrests the cell
cycle, promotes repair pathways, and in response to severe stress
initiates apoptosis. erefore, under normal conditions, it is criti-
cal that intracellular levels of p53 are kept low, which is achieved
by the rapid degradation of p53 by the proteasome. is degra-
dation occurs in both ubiquitin-dependent (6) and ubiquitin-
independent mechanisms (7) and can be modulated by various
signaling pathways including sumoylation, phosphorylation,
acetylation, methylation, and glycosylation (8). Of these, ubiqui-
tination is the most important (6, 9) and the E3 ligase MDM2 is
the primary negative regulator of p53 (10, 11), although several
other E3 and E4 ligases of p53 also exist (8, 9). Mechanistically,
engagement of the p53 N-terminal transactivation domain by the
N-terminal of MDM2, facilitates its C-terminal RING nger E3
ligase activity to transfer ubiquitin to multiple lysine residues of
p53, located in central DNA-binding and C-terminal regulatory
regions (8, 9). MDM2 ubiquitination of p53 (either mono- or
poly-ubiquitination) negatively regulates its transcriptional activ-
ity. Mono-ubiquitin triggers nuclear export, while poly-ubiquitin
targets nuclear p53 for degradation by the proteasome (12).
Notably, the C-terminal of MDM2 is also able to bind with the
C-terminal of the highly related protein MDMX (also known as
HDMX and MDM4). Although MDMX does not possess E3 ligase
activity, the MDM2–MDMX heterodimer ubiquitinates p53 with
higher eciency than MDM2 homodimers (13). MDMX, via its
N-terminus, is able to bind p53 and eciently inhibit its tran-
scriptional activity (14). Furthermore, MDM2 is transcriptionally
up regulated by p53 and this negative-feedback loop associated
with cyclical modulation of levels of both proteins, ensures that
p53 levels remain low under normal conditions (15).
Targeting MDM2 and MDMX
Given the importance of both MDM2 and MDMX in regulating
WT p53, it is unsurprising that they are commonly over-expressed
in some cancers, including sarcoma (~20%) and breast (~15%)
(Figure1A). In this context, they function as powerful oncogenes
and represent logical therapeutic targets for increasing WT p53
expression and activities. e concept of MDM2 targeting was
supported by the discovery of p14ARF (p19ARF in mice), an alternate
reading frame protein produced from the CDKN2A locus (16,
17). P14ARF binds to MDM2, sequestering it in the nucleolus and
preventing it from targeting p53 for degradation (18, 19). More
precisely, the capacity to bind and sequester MDM2 to the nucleus
was assigned to a 22 amino acid fragment from the N-terminus
of p14ARF, revealing a potential method for targeting MDM2 with
small peptide inhibitors (20). e rst successful realization of
this potential came in 2004, when nutlin-3a was discovered by
Vassilev etal. (21). Nutlin-3a potently binds to the hydrophobic
cle in the N-terminus of MDM2, preventing its association with
p53. Importantly, it is highly eective killing of WT p53 cancer
cells, both invit ro and i nvivo in preclinical models, provided vali-
dation for its use. However, its poor bioavailability, high toxicity
(discussed in greater detail below), and its limited eects on
MDMX overexpressing cells (2224) has prevented its translation
to the clinic. Recent interest has switched to compounds that have
better bioavailability and can target both MDM2 and MDMX.
ese new compounds can be broadly segregated according to
their mode of action. e vast majority of preclinical and clinical
small molecule inhibitors work similarly to nutlin-3a, binding to
the N-terminal pocket of MDM2, inhibiting association with p53
(Figure1B). Despite the similarity in the N-terminal p53-binding
domain of MDM2 and MDMX, most of these small molecule
inhibitors bind with signicantly less avidity to MDMX and are
therefore primarily MDM2 specic (12). However, there are now
several new peptide-based inhibitors that are capable of binding
to the N-terminal of both MDM2 and MDMX (Tabl e 1). In addi-
tion, several small molecule inhibitors, which bind specically
to the N-terminus of MDMX, have recently been developed and
are currently undergoing preclinical testing (25, 26). In addition,
there are now a growing number of new MDM2/X inhibitors that
bind outside the N-terminus (Figure1B). ese include small
molecules that inhibit the ubiquitin ligase activity of MDM2
(27); disruptors of MDM2–MDMX heterodimerization (28);
transcriptional inhibitors of both MDM2 (29, 30) and MDMX
(31); MDM2 auto-ubiquitination activators (32, 33); inhibitors
of HSP90 to disrupt MDMX protein folding; and molecules that
directly engage p53 and prevent association with MDM2/X (34).
Cellular Responses to Increased p53
Increased cellular p53 protein levels, resulting from MDM2/X
inhibition, lead to a number of eects that can be simplied into
the broad categories of cell cycle arrest and apoptosis. e decision
between these two pathways is governed by the level and duration
of p53 induction. Lower and cyclical levels of p53 induce arrest,
while sustained levels of elevated p53 expression promotes death
(35). Cell cycle arrest is primarily achieved through transcrip-
tional activation of p53 target genes, primarily p21 and GADD45,
which block the activity of cyclin-dependent kinases (Cdk)
and cause arrest in G1/S (36) and G2 phases, respectively (37).
Interestingly, upregulation of p53 during mitosis does not delay
mitotic progression, but it is an important requirement for arrest-
ing and eliminating aberrant polyploid cells in the subsequent
G1 phase (38, 39). Continued p53 expression occurs when the
damage or stress incurred cannot be repaired or resolved. ese
stresses continue to generate a signaling cascade (e.g., ATM/
ATR, Chk1/2) that leads to the continued stabilization of p53,
and subsequently allows the accumulation of pro-apoptotic p53
targets, including PUMA, Noxa, and Bim within the cell (40, 41).
Once these proteins accumulate to sucient levels, they trigger
apoptosis (42, 43).
MDM2/X INHIBITORS IN CLINICAL
TRIALS
e majority of MDM2-targeted therapies currently in clinical
development are small molecule inhibitors (Tab l e 1 ). ese have
been crystallographically resolved and comprise derivatives
that bind to MDM2 by mimicking Phe19, Trp23, and Leu26,
TABLE 1 | MDM2 and MDMX inhibitors in clinical development.
MDM2 inhibitors in clinical development
Class and
specicity
Nature of
compound
Compound Status p53 NCT identier Company
Small molecule
MDM2 antagonists
Cis-imidazoline RG7112 Phase I in advanced solid and hematological
cancers, and liposarcoma (completed)
n/a NCT00559533
RG7112 with cytarabine Phase I in acute myelogenous leukemia
(completed)
n/a NCT01635296
RG7112 with doxorubicin Phase I in soft tissue sarcoma (completed) n/a NCT01605526 Roche
RO5503781 Phase I in advanced solid cancers (completed) n/a NCT01462175
RO5503781 with
cytarabine
Phase I in acute myelogenous leukemia (active
but not recruiting)
n/a NCT01773408
RO5503781 with
abiraierone
Phase I/II in advanced prostate cancer
(recruiting)
n/a CRUKE/12/032
Spiro-oxindole SAR405838 Phase I in advanced solid cancers (active but
not recruiting)
n/a NCT01636479 Sano-Aventis
SAR405838 with
pimasertib
Phase I in advanced solid cancers (recruiting) n/a NCT01985191
Imidazothiazole DS-3032b Phase I in advanced solid cancers (recruiting) n/a NCT01877382 Daiichi Sankyo
Dihydroisoquinolinone CGM-097 Phase I in advanced solid tumors (recruiting) wtp53 NCT01760525
n/a HDM201 Phase I in advanced solid and hematological
cancers (recruiting)
wtp53 NCT02143635 Novartis
HDM201 with ribociclib Phase Ib/II in liposarcoma (recruiting) wtp53 NCT02343172
Piperidines MK4828 with cytarabine Phase I in acute myelogenous leukemia
(terminated)
n/a NCT01451437 Merck
Piperidinone AMG232 Phase I in advanced solid cancers and multiple
myeloma (recruiting)
n/a NCT01723020 Amgen
AMG 232 with trametinib
and dabrafenib
Phase Ib/IIa in metastatic melanoma
(recruiting)
n/a NCT02110355
Pyrrolidine RG7388 Phase 1 in polycythemia vera and essential
ihrombocythemia (recruiting)
n/a NCT02407080 Pegasys
Stapled peptide
MDM2/X inhibitor
Peptide ALRN-6924 Phase I in advanced solid cancers (recruiting) wtp53 NCT02264613 Aileron
Data extracted from http://www.clinicaltrials.gov, accessed 1st December 2015.
January 2016 | Volume 6 | Article 74
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
which are key residues engaged by p53. ALRN-6924 (Aileron
erapeutics) belongs to a dierent class of therapeutics, which
are stapled peptides designed to disrupt p53 interaction with
both MDM2 and MDMX. A number of these compounds are
also being evaluated clinically in combination with cytotoxics
(doxorubicin and cytarabine), and also molecular-targeted thera-
pies, including ribociclib (CDK4/6 inhibitor), dabrafenib (BRAF
inhibitor), trametinib, and pimasertinib (MEK1/2 inhibitors).
A number of these trials have excluded patients with p53 mutant
tumors; however, the majority have not dened a clear biomarker
for selection criteria, in keeping with the primary end points of
safety and tolerability. It is of interest that a number of these phase
1 trials have yet to be reported even though accrual was started
over 3years ago, which is unusually long in a phase 1 setting.
RG7112 is the most developed in this class of compounds,
and preclinical studies demonstrate strong binding to MDM2,
and eective apoptosis, particularly in MDM2-amplied tumors
(44). One of rst clinical trials reported was in patients with lipo-
sarcoma, a tumor characterized by a high proportion of MDM2
gene amplication and wild-type p53 (45). e primary end point
in this small neoadjuvant study of 20 patients was to assess tumor
biomarkers of p53 pathway activation and cell proliferation. e
results demonstrated an increase in intratumoral p53, p21, and
macrophage-inhibitory cytokine 1 (MIC1, a secreted protein
product of p53) concentrations, an increase in MDM2 mRNA
expression and a small decrease in Ki-67 positive cells in the
treated compared to the pretreated samples. Clinically, the results
were modest, with one partial response and stable disease in 70%
of the cohort. Importantly, there were serious adverse events
(grade 3 or 4) experienced by 40% of the patients, the majority of
which were hematological in nature.
RG7112 has also been evaluated in a phase 1 trial of patients
with relapsed/refractory leukemia, such as AML, ALL, CML, and
CLL (46). e most common toxicities were gastrointestinal and
hematological in nature, 22% of patients experiencing grade 3
and 4 febrile neutropenia. ere was clinical activity, particularly
in the AML cohort, whereby 5 out of 30 evaluable patients
achieved either a complete or partial response, and another 9
patients had stable disease. ese numbers suggest useful single
agent clinical activity, given the refractory nature of their disease
to other therapies. MDM2 inhibition resulted in p53 stabilization
and transcriptional activation of p53 target genes. Interestingly,
two patients who had p53 mutations (G266E and R181L) also
responded to RG7112 in this trial. e G266E is a gain-of-
function (GOF) p53 mutation that upregulates CXC-chemokine
expression and enhances cell migration (47), while R181L is
capable of inducing MDM2 and instigating a cell cycle arrest,
but not apoptosis (48). Consequently, these mutants (G266E and
R181L) may still be sensitive to MDM2/X inhibitors, and hence
patients with these mutations may benet from these inhibitors.
January 2016 | Volume 6 | Article 75
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
Assessing the eects that MDM2/X inhibitors in the context of
the various GOF p53 mutants will be of signicant importance,
as MDM2/X inhibition has the potential to increase the levels
of GOF p53 mutants. Several GOF mutants have been shown
to increase cell proliferation, metabolism, invasion, and chem-
oresistance in cancer cells (4953). Consequently, inhibition of
MDM2/X could place selective pressure on cancer cells with GOF
p53 mutations, driving the clonal evolution of more aggressive
cancer cells and exacerbating tumor growth and metastasis in
patients. Alternatively, a recent preclinical study demonstrated
that the novel small molecule NSC59984 activates p73, resulting
in an MDM2-dependent degradation of GOF p53 and subse-
quent inhibition of tumor growth (54). Other possible explana-
tions for the varied patient response include multiple clones
being present with the tumor (only some of which are mutant), a
retention of one wild-type allele, certain p53 mutations may still
have functional p53 activity (55). Taken together, it is clear that
much more work needs to be done to clearly identify biomark-
ers to improve patient selection for clinical trials of MDM2/X
inhibitors. Furthermore, understanding the heterogeneity of p53
expression and the specic mutations within a patient’s tumor
prior, during and post treatment will also be of considerable
importance for determining the suitability of treatment with
MDM2/X inhibitors.
e clinical eect of MDM2 inhibitors on p53 reactivation,
range from cytostasis to apoptosis, and a combination strategy
may be more ecacious in certain contexts. Preclincal mod-
eling with nutlin-3a has demonstrated improved anti-cancer
activity in combination with cytotoxic- and molecular-targeted
therapies, in dierent tumor types (45); however, the toxicity
prole of the combination partner is a critical determinant of
the success of such an approach clinically. e high incidence
of hematological toxicities in the clinical trials of RG7112
would suggest that therapies with an overlapping side eect
prole would not be suitable as combination partners (45, 46).
A number of clinical trials combining MDM2 inhibitors with
cytotoxics have completed accrual but have yet to be reported
(Table1).
TOXICITIES
A concern of p53 reactivating therapies is its eect on normal
cells. ese include the stabilization of p53 resulting in increased
apoptosis in these cells. is was reected in the clinical trial
of RG7112 in lipoma, whereby the most common toxicity
was hematological in nature, with a reported 30% of patients
experiencing grade 4 neutropenia, and 15% experiencing some-
times prolonged grade 4 thrombocytopenia (45, 46). Whether
hematologic toxicity correlates with prior exposure to genotoxic
therapies is not known. ere are also reports of an increased
incidence of p53 mutations following prolonged nutlin-3a
exposure (56), and concerns about this eect on the develop-
ment of new cancers (57). Other potential o-target eects on
MDM2 inhibitors include the loss of its ability to ubiquitinate
other proteins, such as the steroid hormone receptors [estrogen
receptor (ER) and androgen receptor (AR)] and Rb, as well as
interference with MDM2’s role in DNA repair and modifying
chromatin structure (58). e clinical relevance of these poten-
tial long-term toxicities have not been reported in the current
early phase trials.
CONCLUSION/PERSPECTIVE
Protein–protein interactions, once considered to be a major
hurdle to p53 therapeutic development, can now be targeted
with a growing number of small molecule inhibitors and stapled
peptides. e strategies to overcome this Achilles heel in many
cancers are increasingly varied, and build upon an understanding
of the crystallographic structure of p53 and its interactions with
its major inhibitors. Most of the major pharmaceutical companies
have one or more lead compounds targeting MDM2/X, and many
of these have only recently progressed from preclinical develop-
ment into early phase clinical trials.
e eect of MDM2/X-targeting therapies range from
cytostasis to apoptosis, and combinatory approaches with other
cytotoxic therapies or therapies that target other major onco-
genic pathways are logical approaches, and may allow for lower
and better tolerated doses of both drugs to be administered. For
example, in p53 mutant tumors, protection of normal cells can
be achieved by triggering p53-dependent cytostatic eects with
short, pulsed exposure to MDM2 inhibitors. is cyclotherapy
can reduce the toxic side eect of chemotherapy in these p53
mutant patients (59). Alternatively, recent preclinical evidence
has demonstrated that inhibition of MDM2 with nutlin-3a pre-
vents repair of DNA damage, providing synthetic lethality with
genotoxic agents, such as cisplatin (60). Importantly, this eect
was independent of p53 status and could provide a rational for
examining MDM2 combination therapy in p53 mutant patients.
Getting the therapeutic index right is critical in patients. It is
not surprising that hematological toxicities have been the most
commonly reported and dose-limiting toxicities in the trials
reported so far (45, 46). Long-term follow-up is also critical to
evaluate for the clinical relevance of the potential eects of an
increase in p53 mutations and other o-target eects of this class
of compounds.
e three major biomarkers that have been used to evaluate
therapeutic responses to MDM2/X inhibitors are p53 status,
MDM2, and MDMX levels. Interestingly, the over expression of
MDM2, MDMX, or mutation of p53 are oen mutually exclusive.
For example, liposarcoma, which is one of the rst tumors in
which MDM2 inhibitors have been evaluated (45), shows highly
signicant tendency toward mutual exclusivity (p-value <0.001)
between overexpression of MDM2 (19%) and p53 mutation
(12%) (Figure1A) (5). Other tumors with similar trends of exclu-
sivity include glioblastoma multiforme, melanoma, bladder, lung
andenocarcioma, prostate, and ER-positive breast cancers. ese
tumors present an obvious starting point for trialing MDM2/X
inhibitors in patients. e high rate of MDM2 overexpression
in prostate and ER-positive breast cancers, and the ability of
MDM2 inhibitors to ubiquitinate steroid hormone receptors, has
led to the evaluation of this class of drugs in combination with
endocrine therapies (CRUKE/12/032). It has also been shown
that estradiol modulates a subset of p53 and ER target genes that
can predict the relapse-free survival of patients with ER-positive
January 2016 | Volume 6 | Article 76
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
breast cancer, and that p53 activation with nutlin in combination
with fulvestrant, a selective ER degrader, led to a greater degree
of apoptosis invitro (61).
Given the risk of mutations in p53 driving resistance
to MDM2/X inhibitors, additional biomarkers need to be
identied to maximize the chances of clinical success. is is
highlighted by evidence that p53 mutation status as currently
measured clinically, may not be an accurate representation of
functional p53 activity (46). In support, the recent discovery
that MDM2 inhibitor sensitivity could be predicted by a panel
of 13 p53 transcriptional target genes (62) was subsequently
shown to be based on a signicant number of miss-classied
p53 mutant cell lines (63). Removal of these lines unfortunately
abolished the predicative power of the gene signature. An
alternative approach would be to select for tumors with MDM2
amplication given the mutual exclusivity of p53 mutations and
MDM2 amplication (64). However, MDM2 and MDMX have
dierent and cooperative inhibitory eects on p53 activity, and
therefore inhibitors of one may not be as eective in the setting
of raised levels of the other protein (23). us, these biomark-
ers, while logical in their choice, unless further improved upon,
may potentially exclude patients who may benet from these
therapies.
AUTHOR CONTRIBUTIONS
All authors contributed to the preparation and writing of the
manuscript.
FUNDING
AB is a CINSW FRL fellow (10/FRL/3-02). KC is an NHMRC
Dora Lush Scholar. DT and YH are NHMRC principal research
fellows (APP1104364 and APP9628426, respectively). EL is an
NBCF/VCA practitioner fellow (PRAC14-002). is work was
supported by the Patricia Helen Guest fellowship and Love Your
Sister foundation. e funders had no role in analysis, decision to
publish, or preparation of the manuscript.
REFERENCES
1. Ventura A, Kirsch DG, McLaughlin ME, Tuveson DA, Grimm J, Lintault L,
etal. Restoration of p53 function leads to tumour regression invivo. Nature
(2007) 445:661–5. doi:10.1038/nature05541
2. Martins CP, Brown-Swigart L, Evan GI. Modeling the therapeutic ecacy
of p53 restoration in tumors. Cell (2006) 127:1323–34. doi:10.1016/j.
cell.2006.12.007
3. Xue W, Zender L, Miething C, Dickins RA, Hernando E, Krizhanovsky V, etal.
Senescence and tumour clearance is triggered by p53 restoration in murine
liver carcinomas. Nature (2007) 445:656–60. doi:10.1038/nature05529
4. Bykov VJN, Wiman KG. Mutant p53 reactivation by small molecules
makes its way to the clinic. FEBS Lett (2014) 588:2622–7. doi:10.1016/j.
febslet.2014.04.017
5. Gao J, Aksoy BA, Dogrusoz U, Dresdner G, Gross B, Sumer SO, etal. Integrative
analysis of complex cancer genomics and clinical proles using the cBioPortal.
Sci Signal (2013) 6:l1. doi:10.1126/scisignal.2004088
6. Pant V, Lozano G. Limiting the power of p53 through the ubiquitin
proteasome pathway. Genes Dev (2014) 28:1739–51. doi:10.1101/
gad.247452.114
7. Tsvetkov P, Reuven N, Shaul Y. Ubiquitin-independent p53 proteasomal
degradation. Cell Death Dier (2009) 17:103–8. doi:10.1038/cdd.2009.67
8. Meek DW, Anderson CW. Posttranslational modication of p53: cooperative
integrators of function. Cold Spring Harb Perspect Biol (2009) 1:a000950.
doi:10.1101/cshperspect.a000950
9. Brooks CL, Gu W. p53 regulation by ubiquitin. FEBS Lett (2011) 585:2803–9.
doi:10.1016/j.febslet.2011.05.022
10. Kubbutat MH, Jones SN, Vousden KH. Regulation of p53 stability by Mdm2.
Nature (1997) 387:299–303. doi:10.1038/387299a0
11. Haupt Y, Maya R, Kazaz A, Oren M. Mdm2 promotes the rapid degradation
of p53. Nature (1997) 387:296–9. doi:10.1038/387296a0
12. Li M, Brooks CL, Wu-Baer F, Chen D, Baer R, Gu W. Mono- versus poly-
ubiquitination: dierential control of p53 fate by Mdm2. Science (2003)
302:1972–5. doi:10.1126/science.1091362
13. Leslie PL, Ke H, Zhang Y. e MDM2 RING domain and central acidic domain
play distinct roles in MDM2 protein homodimerization and MDM2-MDMX
protein heterodimerization. J Biol Chem (2015) 290:12941–50. doi:10.1074/
jbc.M115.644435
14. Brooks CL, Gu W. p53 ubiquitination: Mdm2 and beyond. Mol Cell (2006)
21:307–15. doi:10.1016/j.molcel.2006.01.020
15. Shadfan M, Lopez-Pajares V, Yuan Z-M. MDM2 and MDMX: alone and
together in regulation of p53. Transl Cancer Res (2012) 1:88–9.
16. Mao L, Merlo A, Bedi G, Shapiro GI, Edwards CD, Rollins BJ, etal. A novel
p16INK4A transcript. Cancer Res (1995) 55:2995–7.
17. Stone S, Jiang P, Dayananth P, Tavtigian SV, Katcher H, Parry D, et al.
Complex structure and regulation of the P16 (MTS1) locus. Cancer Res (1995)
55:2988–94.
18. Stott FJ, Bates S, James MC, McConnell BB, Starborg M, Brookes S, etal. e
alternative product from the human CDKN2A locus, p14ARF, participates in
a regulatory feedback loop with p53 and MDM2. EMBO J (1998) 17:5001–14.
doi:10.1093/emboj/17.17.5001
19. Honda R, Yasuda H. Association of p19ARF with Mdm2 inhibits ubiquitin
ligase activity of Mdm2 for tumor suppressor p53. EMBO J (1999) 18:22–7.
doi:10.1093/emboj/18.1.22
20. Lohrum MA, Ashcro M, Kubbutat MH, Vousden KH. Contribution of two
independent MDM2-binding domains in p14(ARF) to p53 stabilization. Curr
Biol (2000) 10:539–42. doi:10.1016/S0960-9822(00)00472-3
21. Vassilev LT, Vu BT, Graves B, Carvajal D, Podlaski F, Filipovic Z, et al. In
vivo activation of the p53 pathway by small-molecule antagonists of MDM2.
Science (2004) 303:844–8. doi:10.1126/science.1092472
22. Patton JT, Mayo LD, Singhi AD, Gudkov AV, Stark GR, Jackson MW. Levels
of HdmX expression dictate the sensitivity of normal and transformed cells to
Nutlin-3. Cancer R es (2006) 66:3169–76. doi:10.1158/0008-5472.CAN-05-3832
23. Hu B, Gilkes DM, Farooqi B, Sebti SM, Chen J. MDMX overexpression
prevents p53 activation by the MDM2 inhibitor Nutlin. J Biol Chem (2006)
281:33030–5. doi:10.1074/jbc.C600147200
24. Wade M, Wong ET, Tang M, Stommel JM, Wahl GM. Hdmx modulates
the outcome of p53 activation in human tumor cells. J Biol Chem (2006)
281:33036–44. doi:10.1074/jbc.M605405200
25. Reed D, Shen Y, Shelat AA, Arnold LA, Ferreira AM, Zhu F, etal. Identication
and characterization of the rst small molecule inhibitor of MDMX. J Biol
Chem (2010) 285:10786–96. doi:10.1074/jbc.M109.056747
26. Popowicz GM, Czarna A, Wolf S, Wang K, Wang W, Dömling A, et al.
Structures of low molecular weight inhibitors bound to MDMX and MDM2
reveal new approaches for p53-MDMX/MDM2 antagonist drug discovery.
Cell Cycle (2010) 9:1104–11. doi:10.4161/cc.9.6.10956
27. Herman AG, Hayano M, Poyurovsky MV, Shimada K, Skouta R, Prives C,
et al. Discovery of Mdm2-MdmX E3 ligase inhibitors using a cell-based
ubiquitination assay. Cancer Discov (2011) 1:312–25. doi:10.1158/2159-8290.
CD-11-0104
28. Pellegrino M, Mancini F, Lucà R, Coletti A, Giacchè N, Manni I, etal. Targeting
the MDM2/MDM4 interaction interface as a promising approach for p53
reactivation therapy. Cancer Res (2015) 75:4560–72. doi:10.1158/0008-5472.
CAN-15-0439
29. Huang M, Zhang H, Liu T, Tian D, Gu L, Zhou M. Triptolide inhibits MDM2
and induces apoptosis in acute lymphoblastic leukemia cells through a p53-
independent pathway. Mol Cancer er (2013) 12:184–94. doi:10.1158/1535-
7163.MCT-12-0425
January 2016 | Volume 6 | Article 77
Burgess et al.
P53 Targeted Therapies
Frontiers in Oncology | www.frontiersin.org
30. Qin J-J, Wang W, Voruganti S, Wang H, Zhang W-D, Zhang R. Inhibiting
NFAT1 for breast cancer therapy: new insights into the mechanism of action
of MDM2 inhibitor JapA. Oncotarget (2015) 6:33106–19. doi:10.18632/
oncotarget.5851
31. Wang H, Ma X, Ren S, Buolamwini JK, Yan C. A small-molecule inhibitor
of MDMX activates p53 and induces apoptosis. Mol Cancer er (2011)
10:69–79. doi:10.1158/1535-7163.MCT-10-0581
32. Seo S-K, Hwang C-S, Choe T-B, Hong S-I, Yi JY, Hwang S-G, etal. Selective
inhibition of histone deacetylase 2 induces p53-dependent survivin down-
regulation through MDM2 proteasomal degradation. Oncotarget (2015)
6:26528–40. doi:10.18632/oncotarget.3100
33. Zhang H, Gu L, Liu T, Chi ang K-Y, Zhou M. Inhibition of MDM2 by nilotinib
contributes to cytotoxicity in both Philadelphia-positive and negative acute
lymphoblastic leukemia. PLoS One (2014) 9:e100960. doi:10.1371/journal.
pone.0100960
34. Issaeva N, Bozko P, Enge M, Protopopova M, Verhoef LGGC, Masucci M,
etal. Small molecule RITA binds to p53, blocks p53-HDM-2 interaction and
activates p53 function in tumors. Nat Med (2004) 10:1321–8. doi:10.1038/
nm1146
35. Zhang X-P, Liu F, Wang W. Two-phase dynamics of p53 in the DNA dam-
age response. Proc Natl Acad Sci U S A (2011) 108:8990–5. doi:10.1073/
pnas.1100600108
36. Reisman D, Takahashi P, Polson A, Boggs K. Transcriptional regulation
of the p53 tumor suppressor gene in S-phase of the cell-cycle and the
cellular response to DNA damage. Biochem Res Int (2012) 2012:808934–5.
doi:10.1155/2012/808934
37. Fischer M, Quaas M, Steiner L, Engeland K. e p53-p21-DREAM-CDE/
CHR pathway regulates G2/M cell cycle genes. Nucleic Acids Res (2016)
44(1):164–74. doi:10.1093/nar/gkv927
38. Andreassen PR, Lohez OD, Lacroix FB, Margolis RL. Tetraploid state induces
p53-dependent arrest of nontransformed mammalian cells in G1. Mol Biol Cell
(2001) 12:1315–28. doi:10.1091/mbc.12.5.1315
39. ompson SL, Compton DA. Proliferation of aneuploid human cells is limited
by a p53-dependent mechanism. J Cell Biol (2010) 188:369–81. doi:10.1083/
jcb.200905057
40. Kuribayashi K, Finnberg N, Jeers JR, Zambetti GP, El-Deiry WS. e relative
contribution of pro-apoptotic p53-target genes in the triggering of apoptosis
following DNA damage in vitro and in vivo. Cell Cycle (2011) 10:2380–9.
doi:10.4161/cc.10.14.16588
41. Haupt S, Berger M, Goldberg Z, Haupt Y. Apoptosis–the p53 network. J Cell
Sci (2003) 116:4077–85. doi:10.1242/jcs.00739
42. Kracikova M, Akiri G, George A, Sachidanandam R, Aaronson SA. A thresh-
old mechanism mediates p53 cell fate decision between growth arrest and
apoptosis. Cell Death Dier (2013) 20:576–88. doi:10.1038/cdd.2012.155
43. Khoo KH, Hoe KK, Verma CS, Lane DP. Drugging the p53 pathway: under-
standing the route to clinical ecacy. Nat Rev Drug Discov (2014) 13:217–36.
doi:10.1038/nrd4236
44. Tovar C, Graves B, Packman K, Filipovic Z, Higgins B, Xia M, etal. MDM2
small-molecule antagonist RG7112 activates p53 signaling and regresses
human tumors in preclinical cancer models. Cancer Res (2013) 73:2587–97.
doi:10.1158/0008-5472.CAN-12-2807
45. Ray-Coquard I, Blay J-Y, Italiano A, Le Cesne A, Penel N, Zhi J, etal. Eect of
the MDM2 antagonist RG7112 on the P53 pathway in patients with MDM2-
amplied, well-dierentiated or dedierentiated liposarcoma: an exploratory
proof-of-mechanism study. Lancet Oncol (2012) 13:1133–40. doi:10.1016/
S1470-2045(12)70474-6
46. Andree M, Kelly KR, Yee KW, Assouline SE, Strair R, Popplewell L, etal.
Results of the phase 1 trial of RG7112, a small-molecule MDM2 antagonist in
leukemia. Clin Cancer Res (2015). doi:10.1158/1078-0432.CCR-15-0481
47. Yeudall WA, Vaughan CA, Miyazaki H, Ramamoorthy M, Choi M-Y, Chapman
CG, et al. Gain-of-function mutant p53 upregulates CXC chemokines and
enhances cell migration. Carcinogenesis (2012) 33:442–51. doi:10.1093/
carcin/bgr270
48. Ludwig RL, Bates S, Vousden KH. Dierential activation of target cellular
promoters by p53 mutants with impaired apoptotic function. Mol Cell Biol
(1996) 16:4952–60. doi:10.1128/MCB.16.9.4952
49. Dong P, Karaayvaz M, Jia N, Kaneuchi M, Hamada J, Watari H, etal. Mutant
p53 gain-of-function induces epithelial-mesenchymal transition through
modulation of the miR-130b-ZEB1 axis. Oncogene (2013) 32:3286–95.
doi:10.1038/onc.2012.334
50. Zhou G, Wang J, Zhao M, Xie T-X, Tanaka N, Sano D, et al. Gain-of-
function mutant p53 promotes cell growth and cancer cell metabolism via
inhibition of AMPK activation. Mol Cell (2014) 54:960–74. doi:10.1016/j.
molcel.2014.04.024
51. Zhu J, Sammons MA, Donahue G, Dou Z, Vedadi M, Getlik M, etal. Gain-
of-function p53 mutants co-opt chromatin pathways to drive cancer growth.
Nature (2015) 525:206–11. doi:10.1038/nature15251
52. Fiorini C, Cordani M, Padroni C, Blandino G, Di Agostino S, Donadelli M.
Mutant p53 stimulates chemoresistance of pancreatic adenocarcinoma cells
to gemcitabine. Biochim Biophys Acta (2015) 1853:89–100. doi:10.1016/j.
bbamcr.2014.10.003
53. Chee JLY, Saidin S, Lane DP, Leong SM, Noll JE, Neilsen PM, etal. Wild-
type and mutant p53 mediate cisplatin resistance through interaction and
inhibition of active caspase-9. Cell Cycle (2013) 12:278–88. doi:10.4161/
cc.23054
54. Zhang S, Zhou L, Hong B, van den Heuvel APJ, Prabhu VV, Warfel NA,
etal. Small-molecule NSC59984 restores p53 pathway signaling and anti-
tumor eects against colorectal cancer via p73 activation and degradation
of mutant p53. Cancer Res (2015) 75:3842–52. doi:10.1158/0008-5472.
CAN-13-1079
55. Mello SS, Attardi LD. Not all p53 gain-of-function mutants are created equal.
Cell Death Dier (2013) 20:855–7. doi:10.1038/cdd.2013.53
56. Aziz MH, Shen H, Maki CG. Acquisition of p53 mutations in response to
the non-genotoxic p53 activator Nutlin-3. Oncogene (2011) 30:4678–86.
doi:10.1038/onc.2011.185
57. Francoz S, Froment P, Bogaerts S, De Clercq S, Maetens M, Doumont G, etal.
Mdm4 and Mdm2 cooperate to inhibit p53 activity in proliferating and qui-
escent cells invivo. Proc Natl Acad Sci U S A (2006) 103:3232–7. doi:10.1073/
pnas.0508476103
58. Wade M, Li Y-C, Wahl GM. MDM2, MDMX and p53 in oncogenesis and
cancer therapy. Nat Rev Cancer (2013) 13:83–96. doi:10.1038/nrc3430
59. Rao B, Lain S, ompson AM. p53-based cyclotherapy: exploiting the “guard-
ian of the genome” to protect normal cells from cytotoxic therapy. Br J Cancer
(2013) 109:2954–8. doi:10.1038/bjc.2013.702
60. Carrillo AM, Hicks M, Khabele D, Eischen CM. Pharmacologically increasing
Mdm2 inhibits DNA repair and cooperates with genotoxic agents to kill
p53-inactivated ovarian cancer cells. Mol Cancer Res (2015) 13:1197–205.
doi:10.1158/1541-7786.MCR-15-0089
61. Bailey ST, Shin H, Westerling T, Liu XS, Brown M. Estrogen receptor prevents
p53-dependent apoptosis in breast cancer. Proc Natl Acad Sci U S A (2012)
109:18060–5. doi:10.1073/pnas.1018858109
62. Jeay S, Gaulis S, Ferretti S, Bitter H, Ito M, Valat T, etal. A distinct p53 target
gene set predicts for response to the selective p53-HDM2 inhibitor NVP-
CGM097. Elife (2015) 4:12985. doi:10.7554/eLife.06498
63. Sonkin D. Expression signature based on TP53 target genes doesn’t predict
response to TP53-MDM2 inhibitor in wild type TP53 tumors. Elife (2015)
4:e10279. doi:10.7554/eLife.10279
64. Saiki AY, Caenepeel S, Cosgrove E, Su C, Boedigheimer M, Oliner JD.
Identifying the determinants of response to MDM2 inhibition. Oncotarget
(2015) 6:7701–12. doi:10.18632/oncotarget.3116
Conict of Interest Statement:e authors declare that the research was con-
ducted in the absence of any commercial or nancial relationships that could be
construed as a potential conict of interest.
Copyright © 2016 Burgess, Chia, Haupt, omas, Haupt and Lim. is is an open-ac-
cess article distributed under the terms of the Creative Commons Attribution License
(CC BY). e use, distribution or reproduction in other forums is permitted, provided
the original author(s) or licensor are credited and that the original publication in this
journal is cited, in accordance with accepted academic practice. No use, distribution
or reproduction is permitted which does not comply with these terms.
... Both Murine double minute 2 (MDM2, HMD2 in human) and MDMX act as negative regulators of p53, maintaining p53 at a low level by directly binding to its N-terminal and mediating its degradation in normal cells. 215 The primary mechanism of p53 degradation involves ubiquitylation by the E3 ubiquitin ligase MDM2, which leads to proteasomal degradation of p53. MDM2 amplification is frequently observed in several cancer types, especially in tumors that still retain wildtype p53. ...
... [206][207][208] MDM2, the most important negative regulator of p53, directly binds to p53 and forms complexes with it, regulating the stability and activity of p53 protein. 215 MDM2 binds to the N-terminus of the transcriptional activation domain of p53, inhibiting the trans-activation of p53 and cell growth, regulating cell cycle and inducing cell apoptosis. [365][366][367][368] Therefore, disrupting the MDM2-p53 interaction, presents a potential approach for treating cancer by restoring the impaired function of p53. ...
... KT-333 is a potent and selective bifunctional small molecule protein degrader targeting the STAT3 protein. 511 Similarly, Wang's team developed a small molecule degradation agent for STAT3, known as SD-36 (215), to overcome the challenges faced in targeting this protein (Table 4). SD-36 effectively degraded STAT3 protein in both xenograft tumor tissue and normal mouse tissue, leading to complete and durable tumor regression in a xenograft tumor model at well-tolerated doses. ...
Article
Full-text available
Undruggable proteins are a class of proteins that are often characterized by large, complex structures or functions that are difficult to interfere with using conventional drug design strategies. Targeting such undruggable targets has been considered also a great opportunity for treatment of human diseases and has attracted substantial efforts in the field of medicine. Therefore, in this review, we focus on the recent development of drug discovery targeting “undruggable” proteins and their application in clinic. To make this review well organized, we discuss the design strategies targeting the undruggable proteins, including covalent regulation, allosteric inhibition, protein–protein/DNA interaction inhibition, targeted proteins regulation, nucleic acid-based approach, immunotherapy and others.
... p53 is closely linked with a variety of transcriptional and non-transcriptional activities that regulate cell proliferation, DNA repair, cell senescence, and cell death [39,40]. As a primary negative regulator of p53, MDM2 targets p53 to the proteasome for rapid degradation and suppresses its transcriptional activity to maintain low intracellular p53 levels under normal circumstances [41]. When TP53 is abnormally activated, MDM2 promotes poly-ubiquitination and leads to proteasomal degradation to downregulate p53, thereby maintaining p53 at a normal level [42]. ...
Article
Full-text available
Prostate cancer (PCa) is a prevalent male malignancy globally. Tripartite motif 47 (TRIM47) has been reported to be associated with PCa. However, how TRIM47 acts on PCa is still incompletely understood. Here, we explored the biological roles of TRIM47 in PCa cells and investigated its potential regulatory mechanism. TRIM47 expression in PCa cells was detected by qRT-PCR and western blot. After TRIM47 silencing, the viability of PCa cells was measured using CCK-8 method. Flow cytometry was employed to estimate cell cycle. Cell apoptotic level was subjected to appraisement with TUNEL assay. Additionally, wound healing- and transwell assays were adopted for evaluation of migration and invasion of PCa cells. Moreover, the Biogrid database and HDOCK SERVER predicated that TRIM47 could interact with mouse double minute 2 (MDM2), which was detected using the Co-immunoprecipitation (co-IP) assay and glutathione S-transferase (GST) pull-down assay. The expression of proteins in MDM2/p53 signaling was detected by western blot analysis. Results indicated that TRIM47 expression was highly expressed in PCa cells. TRIM47 knockdown inhibited PCa proliferation and cell cycle whereas promoted cell apoptosis. Besides, TRIM47 knockdown significantly inhibited the migration and invasion of PCa cells. In addition, TRIM47 was proved to bind to MDM2 and regulated MDM2/p53 expression. Importantly, MDM2 overexpression counteracted the impacts of TRIM47 knockdown on cell viability, cell cycle, apoptosis, migration and invasion by regulating the MDM2/p53 pathway. Collectively, our results suggested that TRIM47 silencing inhibits the malignant biological behaviors of prostate cancer cells by regulating MDM2/p53 signaling, which may provide a novel therapeutic target for PCa treatment.
... However, MDM2 has many roles that are not related to p53. Several recent reviews have comprehensively discussed the function of MDM2 in tumorigenesis, its amplification and expression in diverse human cancers, and the cancer therapeutic agent that specifically target MDM2 protein [23][24][25]. Although the role of MDM2 in malignancy and tumor progression has been the focus of most previous studies, MDM2 has recently received increasing attention for its non-carcinogenic effects in various diseases and conditions, such as cardiovascular disease [26]. ...
Article
Full-text available
Xinnaotongluo liquid has been used to improve the clinical symptoms of patients with myocardial infarction. However, the molecular mechanism of Xinnaotongluo liquid is not completely understood. H9c2 cells exposed to hypoxia/reoxygenation (H/R) was used to simulate damage to cardiomyocytes in myocardial infarction in vitro. The biological indicators of H9c2 cells were measured by cell counting kit-8, enzyme linked immunoabsorbent assay, and western blot assay. In H/R-induced H9c2 cells, a markedly reduced murine double minute 2 (MDM2) was observed. However, the addition of Xinnaotongluo liquid increased MDM2 expression in H/R-induced H9c2 cells. And MDM2 overexpression strengthened the beneficial effects of Xinnaotongluo liquid on H9c2 cells from the perspective of alleviating oxidative damage, cellular inflammation, apoptosis and ferroptosis of H/R-induced H9c2 cells. Moreover, MDM2 overexpression reduced the protein expression of p53 and Six-Transmembrane Epithelial Antigen of Prostate 3 (STEAP3). Whereas, STEAP3 overexpression hindered the function of MDM2-overexpression in H/R-induced H9c2 cells. Our results insinuated that Xinnaotongluo liquid could protect H9c2 cells from H/R-induced damage by regulating MDM2/STEAP3, which provide a potential theoretical basis for further explaining the working mechanism of Xinnaotongluo liquid.
... To fully leverage the potential benefits of combining treatments, it is imperative to conduct a comprehensive assessment that includes both clinical and preclinical studies (Konopleva et al., 2020a). This comprehensive evaluation carefully examines the advantages and disadvantages associated with applying combined therapeutic approaches, with the primary objective being to ensure that the combination of diverse therapeutic approaches does not result in increased toxicity or the occurrence of severe adverse effects that outweigh the predicted positive effects (Burgess et al., 2016). This review further examines the molecular mechanisms of the MDM2 pathways and explores potential combination strategies to enhance their tumorigenic effects. ...
Article
Full-text available
Cancer cells often depend on multiple pathways for their growth and survival, resulting in therapeutic resistance and the limited effectiveness of treatments. Combination therapy has emerged as a favorable approach to enhance treatment efficacy and minimize acquired resistance and harmful side effects. The murine double minute 2 (MDM2) protein regulates cellular proliferation and promotes cancer-related activities by negatively regulating the tumor suppressor protein p53. MDM2 aberrations have been reported in a variety of human cancers, making it an appealing target for cancer therapy. As a result, several small-molecule MDM2 inhibitors have been developed and are currently being investigated in clinical studies. Nevertheless, it has been shown that the inhibition of MDM2 alone is inadequate to achieve long-term suppression of tumor growth, thus prompting the need for further investigation into combination therapeutic strategies. In this review, possible clinical and preclinical MDM2 combination inhibitor regimens are thoroughly analyzed and discussed. It provides a rationale for combining MDM2 inhibitors with other therapeutic approaches in the management of cancer, taking into consideration ongoing clinical trials that evaluate the combination of MDM2 inhibitors. The review explores the current status of MDM2 inhibitors in combination with chemotherapy or targeted therapy, as well as promising approach of combining MDM2 inhibitors with immunotherapy. In addition, it investigates the function of PROTACs as MDM2 degraders in cancer treatment. A comprehensive examination of these combination regimens highlights the potential for advancing MDM2-inhibitor therapy and improving clinical outcomes for cancer patients and establishes the foundation for future research and development in this promising area of study.
... Current research is focused on the development of new drugs targeting WDLPS/DDLPS key driver mutated genes such as CDK4 and MDM2. Two generations of compounds have been created that reactivate TP53 by inhibiting the MDM2-TP53 connection, as MDM2 is a nuclear phosphoprotein that inhibits the TP53 pathway (56). Although preclinical research has shown strong tumor suppressor effects, early clinical trials revealed that MDM2-TP53 inhibitors elicited only partial responses in a minority of patients, with the majority of patients experiencing at least one adverse event (52). ...
Article
Full-text available
Background Common kinds of soft tissue sarcomas (STS) include well-differentiated liposarcoma (WDLPS) and dedifferentiated liposarcoma (DDLPS). In this case, we present a comprehensive clinical profile of a patient who underwent multiple recurrences during the progression from WDLPS to DDLPS. Case presentation A 62-year-old Asian female underwent retroperitoneal resection of a large tumor 11 years ago, the initial pathology revealed a fibrolipoma-like lesion. Over the next six years, the patient underwent three resections for recurrence of abdominal tumors. Postoperative histology shows mature adipose tissue with scattered “adipoblast”-like cells with moderate-to-severe heterogeneous spindle cells, pleomorphic cells, or tumor giant cells. Immunohistochemistry (IHC) demonstrated positive staining for MDM2 and CDK4, confirming that the abdominal tumor was WDLPS and gradually progressing to DDLPS. Post-operative targeted sequencing and IHC confirmed the POC1B::ROS1 fusion gene in DDLPS. Whole-exome sequencing (WES) revealed that WDLPS and DDLPS shared similar somatic mutations and copy number variations (CNVs), whereas DDLPS had more mutated genes and a higher and more concentrated amplification of the chromosome 12q region. Furthermore, somatic mutations in DDLPS were significantly reduced after treatment with CDK4 inhibitors, while CNVs remained elevated. Conclusion Due to the high likelihood of recurrence of liposarcoma, various effective treatments should be taken into consideration even if surgery is the primary treatment for recurrent liposarcoma. To effectively control the course of the disease following surgery, combination targeted therapy may be a viable alternative to chemotherapy and radiotherapy in the treatment of liposarcoma.
Article
Inhibitors of the p53–MDM2 interaction such as RG7388 have been developed to exploit latent tumor suppressive properties in p53 in 50% of tumors in which p53 is wild-type. However, these agents for the most part activate cell cycle arrest rather than death, and high doses in patients elicit on-target dose-limiting neutropenia. Recent work from our group indicates that combination of p53–MDM2 inhibitors with the class-I HDAC inhibitor Entinostat (which itself has dose-limiting toxicity issues) has the potential to significantly augment cell death in p53 wild-type colorectal cancer cells. We investigated whether coencapsulation of RG7388 and Entinostat within polymeric nanoparticles (NPs) could overcome efficacy and toxicity limitations of this drug combination. Combinations of RG7388 and Entinostat across a range of different molar ratios resulted in synergistic increases in cell death when delivered in both free drug and nanoencapsulated formats in all colorectal cell lines tested. Importantly, we also explored the in vivo impact of the drug combination on murine blood leukocytes, showing that the leukopenia induced by the free drugs could be significantly mitigated by nanoencapsulation. Taken together, this study demonstrates that formulating these agents within a single nanoparticle delivery platform may provide clinical utility beyond use as nonencapsulated agents.
Article
Full-text available
Opinion statement The therapeutic approach of pleomorphic liposarcoma (PLPS), a rare high-grade subgroup of soft tissue sarcoma, is commonly extrapolated from the management of other LPS subtypes. Only published retrospective data on PLPS currently serve as a guide for oncologists without clear recommendations or specific guidelines. In the advanced setting, specific systemic therapy such as eribulin and trabectedin showed promising activity in comparison to conventional therapy (doxorubicin- and gemcitabine-based protocols), which currently remains the current standard of care at initial stages of the disease. The better understanding of soft tissue sarcoma (STS) pathophysiology and disease course has led to the development of adapted clinical trial designs for rare STS histotypes with specific treatment approach.
Article
Diffuse large B-cell lymphoma (DLBCL) remains a formidable diagnosis in need of new treatment paradigms. In this work, we elucidated an opportunity for therapeutic synergy in DLBCL by reactivating tumor protein p53 with a stapled peptide, ATSP-7041, thereby priming cells for apoptosis and enhancing their sensitivity to BCL-2 family modulation with a BH3-mimetic, ABT-263 (navitoclax). While this combination was highly effective at activating apoptosis in DLBCL in vitro, it was highly toxic in vivo, resulting in a prohibitively narrow therapeutic window. We, therefore, developed a targeted nanomedicine delivery platform to maintain the therapeutic potency of this combination while minimizing its toxicity via packaging and targeted delivery of a stapled peptide. We developed a CD19-targeted polymersome using block copolymers of poly(ethylene glycol) disulfide linked to poly(propylene sulfide) (PEG-SS-PPS) for ATSP-7041 delivery into DLBCL cells. Intracellular delivery was optimized in vitro and validated in vivo by using an aggressive human DLBCL xenograft model. Targeted delivery of ATSP-7041 unlocked the ability to systemically cotreat with ABT-263, resulting in delayed tumor growth, prolonged survival, and no overt toxicity. This work demonstrates a proof-of-concept for antigen-specific targeting of polymersome nanomedicines, targeted delivery of a stapled peptide in vivo, and synergistic dual intrinsic apoptotic therapy against DLBCL via direct p53 reactivation and BCL-2 family modulation.
Article
Full-text available
Purpose: RG7112 is a small-molecule MDM2 antagonist. MDM2 is a negative regulator of the tumor suppressor p53 and frequently overexpressed in leukemias. Thus, a Phase I study of RG7112 in patients with hematologic malignancies was conducted. Experimental design: Primary study objectives included determination of the dose and safety profile of RG7112. Secondary objectives included evaluation of pharmacokinetics, pharmacodynamics, such as TP53-mutation status and MDM2 expression, and preliminary clinical activity. Patients were divided into 2 cohorts: Stratum A (relapsed/refractory AML (except APL), ALL, and CML) and Stratum B (relapsed/refractory CLL/sCLL). Some Stratum A patients were treated at the MTD to assess clinical activity. Results: RG7112 was administered to 116 patients (96 patients in Stratum A and 20 patients in Stratum B). All patients experienced at least 1 adverse event, and 3 DLTs were reported. PK analysis indicated that twice-daily dosing enhanced daily exposure. Anti-leukemia activity was observed in the 30 patients with AML assessed at the MTD included 5 patients who met IWG criteria for response. Exploratory analysis revealed TP53 mutations in 14% of Stratum A patients and in 40% of Stratum B patients. Two patients with TP53 mutations exhibited clinical activity. p53 target genes were induced only in TP53 wild-type leukemic cells. Baseline expression levels of MDM2 correlated positively with clinical response. Conclusions: RG7112 demonstrated clinical activity against relapsed/refractory AML and CLL/sCLL. MDM2 inhibition resulted in p53 stabilization and transcriptional activation of p53-target genes. We provide proof-of-concept that MDM2 inhibition restores p53 function and generates clinical responses in hematologic malignancies.
Article
Full-text available
Transcription factor NFAT1 has been recently identified as a new regulator of the MDM2 oncogene. Targeting the NFAT1-MDM2 pathway represents a novel approach to cancer therapy. We have recently identified a natural product MDM2 inhibitor, termed JapA. As a specific and potent MDM2 inhibitor, JapA inhibits MDM2 at transcriptional and post-translational levels. However, the molecular mechanism remains to be fully elucidated for its inhibitory effects on MDM2 transcription. Herein, we reported that JapA inhibited NFAT1 and NFAT1-mediated MDM2 transcription, which contributed to the anticancer activity of JapA. Its effects on the expression and activity of NFAT1 were examined in various breast cancer cell lines in vitro and in MCF-7 and MDA-MB-231 xenograft tumors in vivo. The specificity of JapA in targeting NFAT1 and NFAT1-MDM2 pathway and the importance of NFAT1 inhibition in JapA's anticancer activity were demonstrated using NFAT1 overexpression and knockdown cell lines and the pharmacological activators and inhibitors of NFAT1 signaling. Our results indicated that JapA inhibited NFAT1 signaling in breast cancer cells in vitro and in vivo, which plays a pivotal role in its anticancer activity. JapA inhibited the nuclear localization of NFAT1, disrupted the NFAT1-MDM2 P2 promoter complex, and induced NFAT1 proteasomal degradation, resulting in the repression of MDM2 transcription. In conclusion, JapA is a novel NFAT1 inhibitor and the NFAT1 inhibition is responsible for the JapA-induced repression of MDM2 transcription, contributing to its anticancer activity. The results may pave an avenue for validating the NFAT1-MDM2 pathway as a novel molecular target for cancer therapy.
Article
Full-text available
The tumor suppressor p53 functions predominantly as a transcription factor by activating and downregulating gene expression, leading to cell cycle arrest or apoptosis. p53 was shown to indirectly repress transcription of the CCNB2, KIF23 and PLK4 cell cycle genes through the recently discovered p53-p21-DREAM-CDE/CHR pathway. However, it remained unclear whether this pathway is commonly used. Here, we identify genes regulated by p53 through this pathway in a genome-wide computational approach. The bioinformatic analysis is based on genome-wide DREAM complex binding data, p53-depedent mRNA expression data and a genome-wide definition of phylogenetically conserved CHR promoter elements. We find 210 target genes that are expected to be regulated by the p53-p21-DREAM-CDE/CHR pathway. The target gene list was verified by detailed analysis of p53-dependent repression of the cell cycle genes B-MYB (MYBL2), BUB1, CCNA2, CCNB1, CHEK2, MELK, POLD1, RAD18 and RAD54L. Most of the 210 target genes are essential regulators of G2 phase and mitosis. Thus, downregulation of these genes through the p53-p21-DREAM-CDE/CHR pathway appears to be a principal mechanism for G2/M cell cycle arrest by p53.
Article
Full-text available
Restoration of wild-type p53 tumor suppressor function has emerged as an attractive anticancer strategy. Therapeutics targeting the two p53 negative regulators, MDM2 and MDM4, have been developed, but most agents selectively target the ability of only one of these molecules to interact with p53, leaving the other free to operate. Therefore, we developed a method that targets the activity of MDM2 and MDM4 simultaneously based on recent studies indicating that formation of MDM2/MDM4 heterodimer complexes are required for efficient inactivation of p53 function. Using computational and mutagenesis analyses of the heterodimer binding interface, we identified a peptide that mimics the MDM4 C-terminus, competes with endogenous MDM4 for MDM2 binding, and activates p53 function. This peptide induces p53-dependent apoptosis in vitro and reduces tumor growth in vivo. Interestingly, interfering with the MDM2/MDM4 heterodimer specifically activates a p53-dependent oxidative stress response. Consistently, distinct subcellular pools of MDM2/MDM4 complexes were differentially sensitive to the peptide; nuclear MDM2/MDM4 complexes were particularly highly susceptible to the peptide-displacement activity. Taken together, these data identify the MDM2/MDM4 interaction interface as a valuable molecular target for therapeutic reactivation of p53 oncosuppressive function.
Article
Full-text available
The oncoprotein murine double minute 2 (MDM2) is an E3 ligase that plays a prominent role in p53 suppression by promoting its polyubiquitination and proteasomal degradation. In its active form, MDM2 forms homodimers as well as heterodimers with the homologous protein MDMX, both of which are thought to occur through their respective C-terminal RING (Really Interesting New Gene) domains. In this study, using multiple MDM2 mutants, we show evidence suggesting that MDM2 homo- and heterodimerization occur through distinct mechanisms, as MDM2 RING domain mutations that inhibit MDM2 interaction with MDMX do not affect MDM2 interaction with wild-type (WT) MDM2. Intriguingly, deletion of a portion of the MDM2 central acidic domain selectively inhibits interaction with MDM2 while leaving intact the ability of MDM2 to interact with MDMX and to ubiquitinate p53. Further analysis of an MDM2 C-terminal deletion mutant reveals that the C-terminal residues of MDM2 are required for both MDM2 and MDMX interaction. Collectively, our results suggest a model whereby MDM2-MDMX heterodimerization requires the extreme C-terminus and proper RING domain structure of MDM2, whereas MDM2 homodimerization requires the extreme C-terminus and the central acidic domain of MDM2, suggesting that MDM2 homo- and heterodimers utilize distinct MDM2 domains. Our study is the first to report mutations capable of separating MDM2 homo- and heterodimerization. Copyright © 2015, The American Society for Biochemistry and Molecular Biology.
Article
Full-text available
Previous reports have provided evidence that p53 mutation is a strong negative predictor of response to MDM2 inhibitors. However, this correlation is not absolute, as many p53Mutant cell lines have been reported to respond to MDM2 inhibition, while many p53WT cell lines have been shown not to respond. To better understand the nature of these exceptions, we screened a panel of 260 cell lines and noted similar discrepancies. However, upon extensive curation of this panel, these apparent exceptions could be eliminated, revealing a perfect correlation between p53 mutational status and MDM2 inhibitor responsiveness. It has been suggested that the MDM2-amplified subset of p53WT tumors might be particularly sensitive to MDM2 inhibition. To facilitate clinical testing of this hypothesis, we identified a rationally derived copy number cutoff for assignment of functionally relevant MDM2 amplification. Applying this cutoff resulted in a pan-cancer MDM2 amplification rate far lower than previously published.
Article
p53, a critical tumor suppressor, is activated by various cellular stresses to prevent and repair damages that can lead to tumor development. In response to these stresses, p53 activation can cause very serious cellular effects including permanent cell cycle arrest and cell death. p53 must therefore be very tightly regulated to avoid unnecessary pathological effects. The homologs MDM2 and MDMX have been shown to be the major, essential negative regulators of p53. In normal cells, MDM2 and MDMX suppress p53 activity, but in the event of cellular stress, they themselves must be inhibited so that p53 may respond to the stress. MDM2 and MDMX are known to bind together, and play multifaceted, non-redundant roles in modulating p53 protein activity. Recently, evidence has emerged showing that MDM2 and MDMX most effectively inhibit p53 as a complex, and possibly play non-redundant roles because they must function as one to control p53. In this review, we give an overview of MDM2 and MDMX and discuss a few ways in which they are modified so that p53 may be activated. Lastly, we discuss the non-redundant roles of MDM2 and MDMX and how it is important to investigate the effect on the complex as a whole when investigating either protein.
Article
The tumor-suppressor p53 prevents cancer development via initiating cell-cycle arrest, cell death, repair, or antiangiogenesis processes. Over 50% of human cancers harbor cancer-causing mutant p53. p53 mutations not only abrogate its tumor-suppressor function, but also endow mutant p53 with a gain of function (GOF), creating a proto-oncogene that contributes to tumorigenesis, tumor progression, and chemo- or radiotherapy resistance. Thus, targeting mutant p53 to restore a wild-type p53 signaling pathway provides an attractive strategy for cancer therapy. We demonstrate that small-molecule NSC59984 not only restores wild-type p53 signaling, but also depletes mutant p53 GOF. NSC59984 induces mutant p53 protein degradation via MDM2 and the ubiquitin-proteasome pathway. NSC59984 restores wild-type p53 signaling via p73 activation, specifically in mutant p53-expressing colorectal cancer cells. At therapeutic doses, NSC59984 induces p73-dependent cell death in cancer cells with minimal genotoxicity and without evident toxicity toward normal cells. NSC59984 synergizes with CPT11 to induce cell death in mutant p53-expressing colorectal cancer cells and inhibits mutant p53-associated colon tumor xenograft growth in a p73-dependent manner in vivo. We hypothesize that specific targeting of mutant p53 may be essential for anticancer strategies that involve the stimulation of p73 in order to efficiently restore tumor suppression. Taken together, our data identify NSC59984 as a promising lead compound for anticancer therapy that acts by targeting GOF-mutant p53 and stimulates p73 to restore the p53 pathway signaling. Cancer Res; 75(18); 1-11. ©2015 AACR.
Article
TP53 (which encodes p53 protein) is the most frequently mutated gene among all human cancers. Prevalent p53 missense mutations abrogate its tumour suppressive function and lead to a 'gain-of-function' (GOF) that promotes cancer. Here we show that p53 GOF mutants bind to and upregulate chromatin regulatory genes, including the methyltransferases MLL1 (also known as KMT2A), MLL2 (also known as KMT2D), and acetyltransferase MOZ (also known as KAT6A or MYST3), resulting in genome-wide increases of histone methylation and acetylation. Analysis of The Cancer Genome Atlas shows specific upregulation of MLL1, MLL2, and MOZ in p53 GOF patient-derived tumours, but not in wild-type p53 or p53 null tumours. Cancer cell proliferation is markedly lowered by genetic knockdown of MLL1 or by pharmacological inhibition of the MLL1 methyltransferase complex. Our study reveals a novel chromatin mechanism underlying the progression of tumours with GOF p53, and suggests new possibilities for designing combinatorial chromatin-based therapies for treating individual cancers driven by prevalent GOF p53 mutations.
Article
The Mdm2 oncogene is a negative regulator of the p53 tumor suppressor and recently identified inhibitor of DNA break repair. Nutlin-3 is a small molecule inhibitor of Mdm2/p53 interaction that can induce apoptosis in cancer cells through activation of p53. While this is promising therapy for those cancers with wild-type p53, half of all human cancers have inactivated p53. Here, we reveal a previously unappreciated effect of Nutlin is inhibition of DNA break repair, stemming from its ability to increase Mdm2 protein levels. The Nutlin-induced increase in Mdm2 inhibited DNA double-strand break (DSB) repair and prolonged DNA damage response signaling independent of p53. Mechanistically, this effect of Nutlin required Mdm2 and acted through Nbs1 of the Mre11/Rad50/Nbs1 DNA repair complex. In ovarian cancer cells where greater than 90% have inactivated p53, Nutlin combined with the genotoxic agents, cisplatin or etoposide, had a cooperative lethal effect resulting in increased DNA damage and apoptosis. Therefore, these data demonstrate an unexpected consequence of pharmacologically increasing Mdm2 levels that when utilized in combination with genotoxic agents induces synthetic lethality in ovarian cancer cells, and likely other malignant cell types, that have inactivated p53. Data reveal a therapeutically beneficial effect of pharmacologically increasing Mdm2 levels combined with chemotherapeutic agents for malignancies that have lost functional p53. Copyright © 2015, American Association for Cancer Research.