Вы находитесь на странице: 1из 13

Mutation Research 773 (2017) 113

Contents lists available at ScienceDirect

Mutation Research/Reviews in Mutation Research


journal homepage: www.elsevier.com/locate/reviewsmr
Community address: www.elsevier.com/locate/mutres

Review

When the guardian sleeps: Reactivation of the p53 pathway in cancer


Olaf Merkela,* ,1, Ninon Taylorb , Nicole Prutscha , Philipp B. Staberc , Richard Moriggld,e,
Suzanne D. Turnerf,1, Lukas Kennera,d,g,* ,1
a
Department of Clinical Pathology, Medical University Vienna, Waehringer Guertel 18-20, 1090 Vienna, Austria
b
Department of Internal Medicine III with Hematology, Medical Oncology, Hemostaseology, Infectious Diseases, Rheumatology, Oncologic Center, Laboratory
of Immunological and Molecular Cancer Research Laboratory of Immunological and Molecular Cancer Research, Paracelsus Medical University, Salzburg,
Austria
c
Department of Internal Medicine 1, Division of Hematology and Hemostaseology, Comprehensive Cancer Centre Vienna, Medical University of Vienna, 1090
Vienna, Austria
d
Ludwig Boltzmann Institute for Cancer Research, Waehringerstrasse 13a, 1090 Vienna, Austria
e
Institute of Animal Breeding and Genetics, University of Veterinary Medicine Vienna and Medical University of Vienna, Austria
f
Department of Pathology, University of Cambridge, Lab Block Level 3, Box 231, Addenbrookes Hospital, Cambridge CB20QQ, UK
g
Institute of Laboratory Animal Pathology, University of Veterinary Medicine Vienna, Veterinaerplatz 1, Vienna, Austria

A R T I C L E I N F O A B S T R A C T

Article history:
Received 7 November 2016 The p53 tumor suppressor is inactivated in most cancers, thus suggesting that loss of p53 is a prerequisite
Available online 17 February 2017 for tumor growth. Therefore, its reintroduction through different means bears great clinical potential.
After a brief introduction to current knowledge of p53 and its regulation by the ubiquitin-ligases MDM2/
Keywords: MDMX and post-translational modications, we will discuss small molecules that are able to reactivate
p53 specic, frequently observed mutant forms of p53 and their applicability for clinical purposes. Many
MDM2 malignancies display amplication of MDM genes encoding negative regulators of p53 and therefore
MDMX much effort to date has concentrated on the development of molecules that inhibit MDM2, the most
TP53 mutation
advanced of which are being tested in clinical trials for sarcoma, glioblastoma, bladder cancer and lung
TP53 gain-of-function
adenocarcinoma. These will be discussed as will recent ndings of MDMX inhibitors: these are of special
Gene therapy
importance as it has been shown that cancers that become resistant to MDM2 inhibitors often amplify
MDM4. Finally, we will also touch on gene therapy and vaccination approaches; the former of which aims
to replace mutated TP53 and the latter whose goal is to activate the bodys immune system toward
mutant p53 expressing cells. Besides the obvious importance of MDM2 and MDMX expression for
regulation of p53, other regulatory factors should not be underestimated and are also described. Despite
the beauty of the concept, the past years have shown that many obstacles have to be overcome to bring
p53 reactivation to the clinic on a broad scale, and it is likely that in most cases it will be part of a
combined therapeutic approach. However, improving current p53 targeted molecules and nding the
best therapy partners will clearly impact the future of cancer therapy.
2017 Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Targeting mutant forms of p53 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1. Small molecules targeting mutant p53 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2. Restoration of mutant p53with zinc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3. Proteosomal depletion of mutant p53 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.4. Targeting proteins that modulate p53function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3. Reactivation of WT p53 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

* Corresponding authors at: Department of Clinical Pathology, Medical University Vienna, Waehringer Guertel 1820, 1090 Vienna, Austria
E-mail addresses: olaf.merkel@meduniwien.ac.at (O. Merkel), lukas.kenner@meduniwien.ac.at (L. Kenner).
1
European Research Initiative for ALK- Related Malignancies, www.erialcl.net.

http://dx.doi.org/10.1016/j.mrrev.2017.02.003
1383-5742/ 2017 Published by Elsevier B.V.
2 O. Merkel et al. / Mutation Research 773 (2017) 113

3.1. MDM2and/or MDMX inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6


3.1.1. MDM2inhibitors: peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.2. MDM2inhibitors: small molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1.3. MDMX inhibitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2. Targeting p53upstream regulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4. Vaccination and gene therapy approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1. TP53 vaccination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2. Gene therapy strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.3. TP53 oncolytic viruses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.4. MicroRNA based strategies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5. Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1. Introduction thereby enhancing its activity [9]. As such, both MDM2 and MDM4
are considered oncogenes and their over-expression is seen in
The tumor suppressor protein p53, also called the guardian of cancers such as sarcomas (20%) and breast cancer (15%). Function
the genome, protects the genome by responding to cellular stress and stability of p53 is also regulated by diverse post-translational
and DNA damage; promoting transient cell cycle arrest enabling modications largely in response to stress signals: for example,
DNA repair, an irreversible cell cycle arrest called senescence or cell phosphorylation by the DNA damage-induced kinases ATM/ATR
death depending on the type of stress and the cellular environment and checkpoint control kinases CHK1/CHK2 upon genotoxic stress
[1]. It therefore follows that the p53 protein-coding gene TP53 is at the N-terminal domain (reversed by phosphatases like PPA2),
the most frequently mutated gene in human cancers mostly acetylation at the negatively charged C-terminus and the DNA
through point mutations in exons 510, whereby amino acid binding domain by the histone acetyl transferase P300/CBP and
substitutions lead to disruption of p53 binding to DNA and methylation of the C-terminus by the methyltransferases
subsequent loss of function (LOF). Hence, cells in the process of SET9 and/or PRMT5 (Fig. 1). Conversely, the NAD-dependent
transformation harboring mutant TP53 can develop an unstable histone deacetylase SIRT1 attenuates p53 function by deacetyla-
genome, evade apoptosis and may nally develop into malignan- tion (Fig. 1). It is important to consider that similar to epigenetic
cies. Therefore, it is not surprising that TP53 mutated or deleted marks on histone tails, methylation, acetylation or ubiquitination
tumor cells are resistant to many chemotherapeutic agents since can even affect the same lysine residue within p53, often leading to
the basis of their efcacy is DNA damage leading to activation of opposing biological effects.
intact p53. As such, mutated TP53 is found in over 50% of all human Thus, different strategies are needed for therapeutic approaches
cancers but the incidence of p53 mutatants varies signicantly towards TP53 mutant and TP53 wild-type cancers. In murine
between cancer types, ranging from 95% in uterine carcinomas to models, reconguration of mutant p53 to its active WT function
5% in cervical or thyroid cancer (www.cbioportal.org). However, restores apoptosis and promotes retardation and even regression
there is also growing evidence that some mutant forms of p53 lead of advanced tumors [1012]. In the context of TP53 wild-type, we
not only to loss of wild-type (WT) p53 tumor suppressive function have shown in a PTEN / prostate cancer mouse model that
but also display gain of function (GOF) properties that infer a disrupting STAT3 in the prostate epithelium leads to a defect in
survival advantage during tumor evolution [2,3]. For example, senescence control through the p14(ARF)/TP53 axis and aggressive
mice lacking expression of p53 as a result of gene knock-out show tumor growth ensues. Hence, senescence must be overcome for
a different disease phenotype to mutant p53 knock-in mice cancer development and aggressive tumor growth is promoted by
whereby the presence of mutant p53 enhances tumor metastasis pro-inammatory JAK-STAT pathway activation [13,14].
[2,3]. Consequently, not all mutations detected in p53, in tumor This review focuses on therapeutic methods to reactivate wild-
cells are equal in their consequences to p53 function; as mentioned type or mutant p53 and their outcome in preclinical and clinical
above, some mutations lead to LOF, others to GOF and some act in a studies.
dominant-negative manner over the WT protein when present in
heterozygosity. 2. Targeting mutant forms of p53
Moreover, p53 is highly regulated by an intricate network of
modier proteins enabling or inhibiting p53 functions. These 2.1. Small molecules targeting mutant p53
interdependent signaling pathways provide many, sometimes
unexpected, targets for therapeutic intervention. Central to the Small molecules have been developed that specically target
function of p53 is its transcription factor activity whereby it binds mutant forms of p53 restoring p53 transcriptional activity, thereby
as a tetramer to DNA at the repetitive consensus binding site leading to cell cycle arrest or apoptosis of tumor cells (for overview
sequence (CATGXXXXXCATG) [4,5]. Currently, more than a see Tables 1 and 2). According to the International Agency for
hundred target genes have been described [6] but the most Research on Cancers (IARC) TP53 mutation database (April 2016),
studied include pro-apoptotic BH3 only proteins (e.g. Puma, Bax, the three most often mutated residues of p53 in cancer are the
Noxa), the cell cycle inhibitor p21, death receptors (e.g. Fas) and arginines in positions 175, 248, and 273 (Fig. 2), [15]. In 2002, the
non-coding RNAs like miR-34a [7]. Regulation of p53 activity under low molecular weight compound PRIMA-1 was identied from a
physiological conditions is largely controlled by proteosomal- National Cancer Institute (NCI) substance library screen [16]. A
mediated degradation following ubiquitination at the C-terminus further compound identied by the same group, MIRA-1 was not
by a network of negative regulatory proteins such as murine further followed up due to toxicity issues [17,18]. The selection
double minute 2 (MDM2) and MDMX (encoded by the MDM4 gene) criteria used by the Bykov group to identify these compounds was
[8]. The E3 ubiquitin-ligase MDM2 triggers proteosomal degrada- the ability to preferentially kill mutant p53 as opposed to TP53
tion while MDMX, which has no ubiquitin ligase function (due to a deleted cell lines by restoring transcriptional activity of mutant or
dysfunctional RING domain), forms heterodimers with MDM2, unfolded p53 [16].
O. Merkel et al. / Mutation Research 773 (2017) 113 3

Fig. 1. Schematic representation of the domain structures of p53, MDM2 and MDM4. A p53 domain structure is shown with an N-terminal transactivation domain (TAD) a
proline-rich domain (PR), a DNA binding core domain (DBD), an oligomerization domain (OD) allowing for TP53 tetramerization and a protein stability regulatory domain
(RD) at the C-terminus. Following DNA damage CHK1, CHK2, ATM and ATX kinases phosphorylate p53 at multiple serine and threonine residues leading to p53 stabilization,
whereas the phosphatase PP2A reverses p53 phosphorylation. The NAD-dependent deacetylase SIRT1 attenuates p53 function but p53 acetylation by P300/CBP activates it.
SET and PRMT5 are methyl-(M) transferases that act C-terminally on the oligomerization and regulatory domains of p53. B MDM2 is an ubiquitin (Ub) ligase that acts on the C-
terminus of p53. MDM2 and MDM4 (also called MDMX) have a similar domain structure with a p53 binding domain at the N-terminus, an acidic domain, a zinc nger motif
and nally the RING domain at the C-terminus, which is responsible for ubiquitination (the RING domain is not active in MDMX).

PRIMA-1 and its derivative PRIMA-1MET, now called APR-246, combination with pegylated doxorubicin and carboplatin
can restore a WT conformation to mutant p53 and in doing so (NCT02098343; www.clinicaltrials.gov).
reinstate transcriptional activity leading to expression of Puma, The activity of APR-246 is not limited to p53 in that it is also able
Noxa and Bax in p53 mutant cells [19,20]. Both compounds are to restore function of the analogous p63 mutant, thereby reversing
converted intracellularly to Michael acceptor methylene quinu- the differentiation block caused by this mutant protein and
clidinone (MQ) that binds covalently to cysteines in mutant p53. extending the use of this exciting compound to mutant p53
Moreover, transfer of MQ-modied mutant p53 into p53 null analogs [27]. TP53 shares high sequence homology with TP63 in its
tumors induces expression of p53 downstream target genes DNA binding domains. Deactivation of p63 as a consequence of
indicating that covalent binding alone of MQ is sufcient [21]. point mutations in the encoding gene leads to severe develop-
In vivo, PRIMA-1 has shown impressive activity in murine mental defects including ectodermal dysplasia, limb defects, and
xenograft models of osteosarcoma [16]. Further cytotoxic and orofacial clefting (EEC syndrome). APR-246 rescues epidermal
apoptotic effects of the derivative APR-246 have been observed in differentiation in skin keratinocytes derived from patients with
diverse murine cancer models such as mutant p53 small cell lung EEC syndrome [28,29].
carcinoma [22], multiple myeloma [23], melanoma [24] and breast Y220C p53 is one of the less frequently observed mutations,
cancer [25]. In 2012, APR-246 was tested in a phase I/IIa clinical although it is estimated that 75,000 new cancer cases per year bear
trial of 22 patients with hematologic malignancies or prostate this mutation, which is comparable to the number of patients
cancer (TP53 mutation status was not a pre-selection criterion). carrying the BCR-ABL translocation. This mutation, in the core
The drug was generally well tolerated with only transient side domain of p53 is not involved in DNA binding, but the Y220C
effects such as fatigue, dizziness, headache, and confusion. Two substitution creates a cavity that destabilizes p53. The small
minor responses were observed in this late stage cohort, one in an molecules PhiKan083 and PhiKan7088 bind to this cavity and
acute myeloid leukemia (AML) patient carrying V173M mutant therefore interact with Y220C p53. They were designed by
p53, the other one in a non-Hodgkin lymphoma patient with a molecular modeling based on the X-ray crystal structure of this
TP53 splice site mutation [26]. Currently, APR-246 is under study in mutant. In vitro, PhiKan7088 permits the correct folding of Y220C
a phase Ib/II clinical trial of high grade serous ovarian cancer in p53 restoring transactivation potential and hence inducing
4 O. Merkel et al. / Mutation Research 773 (2017) 113

Table 1
Overview of selected compounds and peptides that can activate mutant or wt p53 and their mechanism of action. Molecular weight was obtained from the PubChem Library
(National Institute of Health) version November 2016. See Table 2 for 2D structures of compounds.

Compound Derivatives Structure Target Mechanism Caveats MW


PRIMA PRIMAMET quinuclidinone p53 mutant/ restores p53 transactivation intact p19ARF/MDM2 signaling 185.2
(APR-246) WT required
p63 mutant
phiKan083 phiKan5174 carbazole Y220C p53 alters protein stability high in vitro 274.8
phiKan5196 conc. required
CP31398 quinazoline p53 DBR unknown 435.4
mutants
SCH529074 quinazoline p53 DBD restores p53 WT function, inhibits 564.0
MDM2 mediated
degradation
Ellipticine and analogs ellipticine p53 mutants restores p53 transactivation 246.3
RITA p53 WT displaces MDM2 may act upstream of p53 292.4
on N-terminus
JNJ26854165 tryptamine p53 mutants E2F1 induction, blocks cholesterol transport 328.4
(serdemetan)
SAH-p53-8 MDM2 prevents p53 high in vitro peptide
degradation conc. required
sMTide-02a MDM2 prevents p53 cell cycle arrest but no apoptosis peptide
degradation

expression of p21 and Noxa, leading to cell cycle arrest and to p53 [33]. In vivo studies of a transgenic APCmin colon cancer
apoptosis (Fig. 2) [30,31]. Altogether, small molecular weight mouse model demonstrated that CP-31398 suppressed APCmin
compounds constitute a promising approach for targeting specic induced colon cancer development; CP-31398 treatment induced
mutant p53 forms and their further development addresses a great apoptosis of tumor cells with elevated levels of p53 and expression
clinical need. of p21 detected [34]. Furthermore, in a mouse model of UVB-
The quinazoline derivative CP-31398 was selected from a screen induced skin carcinogenesis, a reduction in tumor growth was
of over 100,000 compounds looking for those that could restore observed [35]. More recently, when applied to pancreatic
wild-type conformation of DNA binding domain (DBD) mutants of adenocarcinoma cell lines, CP-31398 induced apoptosis as well
p53, using the conformation-specic antibody PAb1620 as a probe as autophagy, although active concentrations were above 10 mM
[32]. The exact mechanism of action of CP-31398 is unclear, since [36]. New analogs have now been developed that lack the unstable
NMR studies have not been able to show direct binding of CP-31398 styryl linkage in the side chain of CP-31398 and are showing

Table 2
2D structure of compounds that can activate mutant or wt p53. Compound structures were obtained from the PubChem Library (National Institute of Health) version
November 2016.
O. Merkel et al. / Mutation Research 773 (2017) 113 5

R175H p53 mutant. Another compound NSC319726 seems also to be


a zinc-chelator and restores function of the R175H p53 mutant in
both cancer cell lines and tumor xenografts [52].

2.3. Proteosomal depletion of mutant p53

Rationale for this strategy includes selective degradation of


dominant negative mutant forms of p53 or the depletion of
mutant forms that have GOF properties. For this approach,
compounds were screened for degradation of mutant p53 but not
WT p53. Mutant p53 is stabilized by interaction with the heat
Fig. 2. p53 mutations that disrupt DNA binding or destabilize protein folding and shock protein Hsp 70 and Hsp 90 chaperone complex that
substances that can rescue these mutations. Mutations at sites 248 and 273 are the depends on HDAC6 interaction for activation [53]. Several Hsp
most frequently detected in cancer and are involved in direct binding to DNA. Small 90 inhibitors have been developed and include geldanamycin, 17-
molecules that restore DNA binding (collectively denoted by X in the diagram) are
not mutation site-specic and include PRIMA, APR-246, CP31398, Ellipticine
AAG and ganetespib. In murine models, ganetespib showed a 50-
analogs and JNJ26854165. In contrast the arginine 175 mutation is important for fold higher potency than 17-AAG in destabilizing mutant p53
zinc binding to p53, which is required for correct folding. This mutation can be [54]. Ganetespib was tested in a phase II trial for Non-Small Cell
rescued by the novel drug zinc metallochaperone 1 (ZMC-1). The Y220C mutation is Lung Cancer (NSCLC) together with docectaxel and showed
not involved in DNA binding, but destabilizes p53s 3-D structure by forming a
increased progression-free survival although a phase 3 study that
crevice that can be targeted by PhiKan-type small molecules.
closed this year (NCT01798485) could not reproduce these
ndings [55].
promising rst results [37]. Another small molecule that speci- Similar approaches include histone deacetylase inhibitors
cally binds the p53 DBD is SCH529074; this interaction restores (HDAC) such as suberoylanilide hydroxamic acid (SAHA), also
wild-type function and concomitantly inhibits ubiquitination of called Vorinostat or Zolinza. Vorinostat could destabilize the
p53 by HDM2 [38]. complex formed between Hsp90 and mutant p53 and shows
Ellipticine analogs have also been tested for p53-dependent preferential cytotoxicity in mutant p53 cancer cells [56]. Vorino-
cytotoxic activity within the anticancer drug discovery program of stat was approved by the FDA in 2006 for the treatment of
the US National Cancer Institute (NCI) [39]. Ellipticine restores cutaneous T cell lymphoma and is being trialed in other cancers
transactivation inducing expression of p21 and MDM2 in several including myelodysplastic syndrome and Szary syndrome [57].
p53 mutant-expressing cell lines as well as in nude mouse tumor Arsenic trioxide (ATO) is used in combination with all trans
xenografts [40]. Ellipticine was shown to sensitize p53 mutant retinoic acid (ATRA) as a curative therapy for acute promyelocytic
cells to doxorubicin-induced apoptosis in human lymphoma cell leukemia (APL). Like ATRA it also leads to the degradation of the
lines [41]. However, other mechanisms such as DNA intercalation, PML-RARa fusion protein thereby allowing terminal cell differen-
topoisomerase II inhibition and formation of DNA adducts may also tiation [58]. We have shown that p53 mutant cells of chronic
be involved [42]. lymphocytic leukemia patients are selectively killed by ATO [59].
Similar to PRIMA discussed above, RITA (reactivation of tumor This may at least in part be explained by data from Yan and
cell apoptosis) was also derived from the NCI library. RITA is able to colleagues who showed that ATO can reactivate proteasome-
kill p53 WT HCT116 colon cancer cells more effectively than the dependent degradation of mutant p53 in cancer cells partly
isogenic p53 mutant version of this cell line and also shows activity through enhanced expression of the Pirh2 E3 ligase [60].
in murine xenograft models [43,44]. RITA targets the N-terminus of Disulram (DSF) has ubiquitin proteasome-targeting activity
p53 and through conformational change prevents the binding of inhibiting acetaldehyde dehydrogenase and as such is in use for the
negative regulators such as MDM2 or HPV E6 [45,46]. More treatment of chronic alcoholism. Interestingly, DSF was shown to act
recently, Selivanova and colleagues showed in neuroblastoma that synergistically with doxorubicin in the mutant p53 breast cancer cell
RITAs mechanism of action could be extended to both mutant line MDA-MB-231[61]. Recently it has been evaluated as an adjunct
(C135F, M246R) and WT p53. When activated by RITA, to chemotherapy in metastatic NSCLC with remarkable success [62].
p53 inhibited key oncogenes including n-Myc, BCL2, MDM2 and Gambogic acid is derived from the Garcinia hanburyi tree and
MDMX [47]. Recently, a study suggested that RITA-resistant cell induces proteasomal degradation of mutant p53 [63]. Spautin-1, an
lines show increased DNA cross-link repair and are therefore also inhibitor of autophagy also has an unexpected role in mutant
resistant to crosslinking agents like cisplatin, indicating that the p53 degradation. When cancer cells are challenged by starvation,
mechanism of action may be inducing DNA damage upstream of cells start a process of macro-autophagy during which intracellular
p53 rather than a direct effect on p53 [48]. proteins and organelles are degraded through lysosomes, providing
an alternative energy source. Spautin-1 controls levels of p53 by
2.2. Restoration of mutant p53 with zinc regulating the deubiquitinating activity of USP 10 and USP 13 [64].
Under a glucose-free environment with conuent growth con-
p53 contains a single zinc ion close to its DNA binding interface. ditions, spautin-1 inhibits macro-autophagy and induces degrada-
Zn2+ is essential for site-specic DNA binding, activation of tion of several mutant p53 forms by a specic form of autophagy,
downstream targets and correct protein folding. Loss of metal- namely chaperone-mediated autophagy [65]. This is an interesting
lothioneins, which store intracellular zinc, promotes the unfolding new concept that may lead to novel forms of therapy for p53 mutant
of WT p53 and inhibits its transcriptional activity [49]. If p53 bears solid cancers and the recent Nobel prize in medicine awarded for the
the R175H mutation its capacity for zinc binding is abrogated. discovery of autophagy might raise research interest in this
Screening of a small compound library has shown that the small underexplored area.
molecule zinc metallochaperone-1 (ZMC1) was able to restore WT
function specically to this mutant form of p53. The reason for this 2.4. Targeting proteins that modulate p53 function
may be that ZMC1 is a zinc ion chelator, which enables transport of
the ion through the cell membrane [50,51]. ZMC1 may pave the way An alternative approach towards targeting mutant p53 activity
for future clinical molecules that restore function to the common is either to reactivate tumor suppressive pathways inhibited, or to
6 O. Merkel et al. / Mutation Research 773 (2017) 113

suppress oncogenic targets activated by it. As seen before, mutant inhibitors of MLL complex formation in p53 GOF mutant cells block
p53 can exert GOF activity by binding and thereby inhibiting the cell proliferation but do not affect p53 WT cells [69].
tumor suppressor functions of p63 and p73. Accordingly, the
disruption of mutant p53 in complex with p63 and p73 appears an 3. Reactivation of WT p53
alternative to the refolding approach of mutant p53. For example,
the small molecule RETRA suppresses mutant p53-bearing cancer 3.1. MDM2 and/or MDMX inhibitors
cells through inhibition of the mutant p53/p73 complex and
induction of p73 expression [66]. In numerous tumors, TP53 is not mutated but inactivated due to
The R280K or R273H DBD mutants of p53 when engineered (over)expression of negative regulatory proteins such as
into mice, disrupt mammary tissue architecture by inducing the MDM2 and MDMX. Deregulation of the autoregulatory feedback
mevalonate/cholesterol synthesis pathway [67]. Statins inhibit loop between p53 and MDM2 proteins neutralizes p53 tumor
the rate limiting enzyme in this pathway 3-hydroxy-3-methyl- suppressor function. Therefore, disruption of p53/MDM2/MDMX
glutaryl coenzyme A (HMG-CoA) reductase and are widely used interactions by small molecules may reactivate the p53 pathway in
for the treatment of hypercholesterinemia and prevention of tumor cells retaining WT p53. The MDM2 protein was isolated in
cardiovascular disease. Hence, statins restore morphology and 1987 as the protein product of extrachromosomal circular DNA
decrease survival of mutant p53 breast cancer cell lines [67]. amplications, also called double minutes, in transformed 3T3
Statins also block phosphorylation of oncogenic MYC, which cell lines [70]. MDM2 regulates p53 by two main mechanisms:
together with their impact on cholesterol metabolism and inducing p53 degradation (by acting as an E3-ubiquitin ligase) and
membrane integrity could help to explain their broader anti- directly suppressing the transactivation domain of p53, hence
cancer action. preventing transcription of downstream targets. In contrast,
Modication of p53 is also conducted by proteins typically MDMX has no ubiquitin ligase function, but shares MDM2s ability
associated with epigenetic histone marks such as those that to inhibit transcription of p53 downstream targets. A prerequisite
mediate lysine methylation and acetylation. For example, the for the ubiquitin ligase function of MDM2 is the formation of
methyltransferases SMYD and SET7 methylate lysines in the C- homo-oligomers, which depends on the carboxy-terminal RING
terminus of WT p53 in teratocarcinoma cell lines compromising its domain; MDMX does not form homo-oligomers. However, hetero-
transcriptional activity. Thus, inhibition of these proteins leads to oligomerization between MDMX and MDM2 is needed for optimal
p53 reactivation and differentiation of the teratocarcinoma cells ubiquitin-ligation through MDM2 [7173]. The p53/
suggestive of a possible therapeutic avenue [68]. In a similar vein, MDM2 interaction brings together the N-terminus of the
mutant GOF p53 has been shown to bind to the transcription factor p53 transactivation domain with the rst 120 amino acids of
ETS2 thereby activating transcription of the methyltransferases the MDM2 N-terminus. The X-ray crystal structure of
MLL1 and MLL2 as well as the acetyl transferase MOZ1. MLL family MDM2 bound to a p53 peptide revealed Phe19, Trp23 and
members are core components of large protein complexes that Leu26 as the key residues creating a locally restricted, hydrophobic
modify chromatin structure and hence gene transcription and in interaction site within p53 (see Fig. 3) [74]. Accordingly, peptides
the case of p53 GOF mutants, induce transcription of genes with the same or similar sequence were screened for their ability to
associated with tumor cell proliferation. Thus, highly specic interfere with MDM/p53 binding. Peptides are polar, large

Fig. 3. Drugs that interact with the MDM2/p53 binding pocket. The mechanism of most MDM2 inhibitors is binding into the hydrophobic p53-binding cleft of MDM2. Critical
for MDM2 binding are the hydrophobic p53 amino acids phenylalanine 19 (F19), tryptophan 23 (W23) and leucine 26 (L26). Drug classes used to displace p53 are peptides (P),
Nutlins (N), Spirooxindoles (S) and Piperidinones (Pi).
O. Merkel et al. / Mutation Research 773 (2017) 113 7

molecular weight compounds; structural mimetics and screens of Another interesting class of MDM2 inhibitors is the spiroox-
compounds blocking MDM2 binding are an area of future indole family, initially developed by the Wang group at the
investigation. University of Michigan [92]. This family of compounds was
identied by structure-guided de novo design in an effort to
3.1.1. MDM2 inhibitors: peptides mimic within an organic molecule the three key hydrophobic
One mechanism to inhibit MDM2 activity is to design peptide- residues in the MDM2 binding pocket of p53: Phe19, Trp23 and
mimics of p53 that block, via steric hindrance the p53 binding site Leu26. This study resulted in the development of a series of MI
of MDM2. WT p53 peptide encompassing the helical compounds, key amongst which is MI-63 which binds MDM2 with
MDM2 binding site has comparatively low binding afnity for a Ki of 3 nM and in turn activates p53, inhibiting cell growth in p53
MDM2. To improve afnity, the amino acid sequence of wild-type cells [93]. Further investigations led to the development
p53 peptides has been optimized using phage display techniques MI-147 showing p53 modulating activity in SJSA-1 osteosarcoma
and in silico structure-guided approaches. Introduction of non- and HCT116 colon carcinoma cell lines [94]. In SJSA-1 osteosarcoma
natural amino acids that allow for additional interpeptidic bonds xenografts, MI-147 alone and in combination with irinotecan
have been introduced into the peptide sequence, a method also showed strong potency and continues to undergo preclinical
called stapling. This has led to enhanced afnity by reducing evaluation. Another member to this class of compounds is MI-888,
conformational exibility, increasing the half live of the peptides which through changes in the stereochemistry reduces the Ki value
and enabling cellular uptake. One of the rst inhibitors of this type to 0.44 nM and has also been shown to have activity in murine
is SAH-p53-8, a stapled p53 peptide made cell permeable by the xenograft models upon oral administration [92]. Finally,
introduction of positively charged amino-acids. In contrast to an SAR405838, a highly potent and selective MDM2 inhibitor was
inactive mutant form, SAH-p53-8F19A, SAH-p53-8 can reactivate identied and shown to halt tumor growth or lead to durable
p53, induce p21 and lead to apoptosis of the SJSA-1 osteosarcoma remissions in murine engraftment models of osteosarcoma, acute
cell line. However, concentrations (1015 mM) needed to exert leukemia, prostate cancer and colon cancer [95]. In dedifferenti-
these effects are relatively high [75]. Another peptide that binds ated liposarcoma, a disease which is characterized by high
MDM2 is sMTide-02A detected using phage-display techniques MDM2 expression and usually wild type p53, SAR405838 gave
and modied (through stapling, removal of proline 12 and a change rise to p53 activation and tumor regression at much lower
of asparagine 8 to alanine) to bind to the p53-binding pocket of concentrations than comparable substances such a Nutlin3a and
MDM2 with high afnity (it has a remarkably low Kd value of 6.76 MI-219 [96]. However, a recent publication has shown develop-
nM) [76,77]. This peptide leads to G1 and G2 arrest in different ment of p53 mutations during SAR405838 treatment in patients
p53 wild-type cell lines without induction of apoptosis [78]. with dedifferentiated liposarcoma, suggesting fast acquisition of
resistance [97].
3.1.2. MDM2 inhibitors: small molecules Another class of MDM2 inhibitors are piperidinones, such as
The small size of the MDM2/p53 interface has made it possible AMG232 that inhibits tumor growth in SJSA-1 osteosarcoma
to design not only peptides but also small organic compounds that xenografted mice [98,99]. Currently, AMG232 is under investiga-
block this interaction (Fig. 3). Compound screens, structure-guided tion in a phase I study of patients with solid tumors or multiple
in silico searches or de novo design, have identied molecules that myeloma (NCT01723020), in a phase Ib study with or without co-
are able to disrupt this interaction by binding the MDM2/ administration of the MEK1/MEK2 inhibitor trametinibin in AML
p53 interface of MDM2. The most promising MDM2/ (NCT02016729) and in a phase Ib/IIa in combination with
p53 disruptors belong to different classes of compounds: nutlins, trametinib and dabrafenib in patients with metastatic melanoma
spirooxindoles and piperidinones. Nutlin-3a was developed in (NCT02110355).
2004 by Hoffmann la Roche with the group of Vassilev et al. [79]
and inhibits the association of MDM2 with p53 via binding to the 3.1.3. MDMX inhibitors
hydrophobic cleft of MDM2 (Fig. 2 B). Nutlin-3a is a potent Resistance to MDM2 antagonists can develop as a consequence
MDM2 inhibitor and shows efcacy in primary chronic lympho- of amplication of MDM4 preventing transactivation of p53 target
cytic leukemia cells that are wild-type for p53 [8082] and in many genes [100102]. Therefore, in such cases parallel inactivation of
tumor cell lines from various malignancies including acute myeloid MDMX may be of benet. Whilst most MDM2 inhibitors also bind
leukemia [83], chemoresistant neuroblastoma [84], anaplastic MDMX they do so with a lower Kd that is not sufcient for total
large cell lymphoma [85,4] or more recently multiple myeloma deactivation. For example, in vitro binding studies have shown that
[86]. However, due to poor bioavailability and high toxicity its use Nutlin-3a binds MDM2 with 500-fold selectivity over MDMX.
in the clinic has been hampered. The nutlin analog RG7112 was rst Hence, there is a need to develop specic MDMX inhibitors. One of
clinically evaluated in MDM2-amplied liposarcoma patients the rst of this class of compounds is SJ-172550, which binds
showing that it could indeed amplify the p53 pathway in patients reversibly to MDMX inhibiting the p53 binding domain and
[87]. In addition, a phase I clinical trial of patients with relapsed inducing apoptosis in retinoblastoma cells known to express high
refractory leukemia was performed recently [88]. Toxicities levels of MDMX [103]. However, the stability of covalent adducts
include gastrointestinal and hematologic side-effects (neutrope- formed between SJ-172550 and thiol groups within the MDMX
nia, thrombocytopenia). Among the AML patients, ve out of p53-binding pocket is dependent on variable factors, which
30 patients reached complete or partial responses and 9 patients complicates its use as a therapeutic agent [5]. A second compound
entered stable disease. Furthermore, two patients with mutant p53 is XI-006, which acts through suppression of MDMX transcription
(R181L and G266E) showed a clinical response. The latest addition as assessed by a MDMX promoter-linked luciferase assay [104]
to this group of compounds is RG-7388 which shows greater effects whereas RO-5963 leads to forced dimerization of MDMX through
in the induction of p53 in the osteosarcoma cell line SJSA-1 and interaction with its RING-domain, inhibiting its ability to form
accordingly is also able to reduce growth of engrafted SJSA-1 cells complexes with p53 and as a consequence was shown to overcome
in mice at lower concentrations than RG-7112 and Nutlin-3a [89]. Nutlin-3a resistance when due to high expression of MDMX [100].
Currently, a number of phase I clinical trials with MDM2 inhibitors A scheme of the mode of action of the MDMX inhibitors described
in combination with other agents are being evaluated for example is given in Fig. 4.
a phase Ib study in AML patients with cytarabine [90] or in Another lead in the design of dual MDM inhibitors is lithocholic
combination with trabectedin in sarcomas [91]. acid, a steroid fatty acid also naturally present in bile acid, which
8 O. Merkel et al. / Mutation Research 773 (2017) 113

Fig. 4. MDMX inhibitors and their mechanism of action. SJ172550 is a compound that binds to the p53 binding pocket of MDMX via disulde bonds. RO-5963 acts by
dimerization of MDMX through binding to the RING domain.

was found to preferentially target MDMX over MDM2, raising the very limited clinical effect, suggesting effective immune evasion by
possibility that certain steroids may be natural ligands of MDM tumors [110]. Immunomodulatory escape mechanisms include
proteins. However, low solubility and apoptosis inducing activity changes in CD4+FoxP3+ regulatory T cells (Tregs) suppressing the
only at concentrations above 100 mM prevent its use as a adaptive immune response, expression of the co-stimulatory
therapeutic drug [105]. Lipid side chain modications of lith- protein CTLA-4, which induces T-cell anergy and/or differential
ocholic acid and cytochrome P450 action might overcome these regulation of programmed death ligand 1 (PD-L1) and its receptor
limitations in future approaches. PD-1 restraining the tumor killing effect of T-cells [10,111,112].
Ideally, the tumor-associated antigen used for vaccination should
3.2. Targeting p53 upstream regulators have three properties: First, it should be exclusively expressed on
tumor tissues, second, be essential for tumor growth and third,
As previously mentioned, p53 stability and activity is regulated possess immunogenic properties. Unfortunately, p53 is expressed in
by post-translational modications and therefore modulation of both healthy and tumor cells. Nevertheless, it has been shown that
these events also presents therapeutic options. Persistent acetyla- CD8 T cells directed against wild-type p53 can discriminate between
tion of p53 promotes protein stabilization and also activation of p53-presenting tumor cells and normal tissue [113]. The reasons for
downstream targets. Tenovin-6, a small molecular weight sirtuin this remain unclear and may be related to the higher levels of
deacetylase SirT1 enzyme inhibitor blocks p53 deacetylation. p53 expressed in cancers, particular mutant forms of p53 leading to
Hence Tenovin-6 induces apoptosis of acute lymphoblastic elevated levels of processing and presentation on immune cells.
leukemia cells and their stem cells, as well as uveal melanoma Indeed, higher levels of p53 antibodies have been found in patients
cells with mutant p53 [106,107]. When DNA damage induces cell with lung, pancreatic, bladder and breast cancer and in patients
cycle arrest it is no surprise that p53 mutant cells are highly expressing mutant forms of p53, although the levels of antibody
dependent on the G2/M checkpoint, since they are unable to production are obviously insufcient to eradicate the tumors
initiate G1 arrest via p21 and CDK2. Therefore it is an interesting [114,115]. In a study of 21 ovarian cancer patients, p53 peptide-
synthetic lethal strategy to enforce lethal G2/M transition, by pulsed dendritic cells (p53 amino acids 264272) were compared to
keeping the checkpoint kinase CDK1 in its active dephosphory- subcutaneous injection of the same p53 peptide together with the
lated state [108]. This can be achieved by inhibiting the kinases myeloid cytokine GM-CSF. An immune response was detected in 83%
Wee1 (MK-1775) or Chk1 (AZD7762) that usually mediate G2/M and 69% of the patients as tested by ELISPOT-assays in the two arms
arrest by ensuring CDK1 phosphorylation. Accordingly, in a respectively, suggesting successful immunization in both arms.
p53 mutant head and neck squamous cell carcinoma (HNSCC) Administration of IL-2 in both arms generated grade 34 toxicities
engraftment model, the Wee1 kinase inhibitor MK-1775 displays and boosted the presence of Tregs precluding its clinical applicability.
single-agent activity and potentiates the efcacy of cisplatin [109]. Both arms of the study had similar progression-free and overall
survival, with minor toxicities, suggesting subcutaneous injection to
4. Vaccination and gene therapy approaches be as effective as peptide-pulsed approaches [116]. More recently,
the results of a phase Ib trial dendritic cell p53 peptide vaccine for
4.1. TP53 vaccination head and neck cancer were published [117]. Adjuvant
p53 vaccination of 16 patients was safe with a two-year disease-
Tumor vaccination is an exciting concept in the era of free survival of 88%. In 69% of the patients, p53-specic T cells
immunotherapy and tumor neoantigen vaccination strategies increased post-vaccination conrming the rational for the study.
yet cancer vaccination programs have gone through phases of
enthusiasm and disappointment. Recently, further development 4.2. Gene therapy strategies
and a better understanding of the immune system have led to
improved novel strategies. Early phase clinical trials using peptides TP53 gene therapy is based on delivery of a wild-type TP53 gene
that activate dendritic cells or vector-based strategies have had into tumor cells. The rst human study was reported in 1996 by
O. Merkel et al. / Mutation Research 773 (2017) 113 9

Roth et al., in which a retrovirus carrying human TP53 was directly p53 through suppressing MDM2 [133,134]. Of note, miRNA-34a,
injected into NSCLC [118]. However, use of retrovirus remains a initially described as a p53 downstream target, can upregulate
threat due to the potential for oncogene activation as a result of p53 activity through inhibition of multiple p53 negative regu-
retroviral insertion into the host genome [119]. Hence, alternatives lators like SIRT1 [135]. miRNA-34a is an important tumor
that minimize this problem have been explored including suppressor and its expression is silenced by epigenetic alteration
adenoviral vector gene delivery. One such vector is Gendicine, a in various cancers including esophageal, gastric, hepatocellular
recombinant adenoviral human TP53 vector that received approval and colorectal cancers [136]. Two approaches for miRNA-34a
in China in 2003 for use in combination with radiotherapy for the restoration are currently under investigation: Firstly, the use of
treatment of head and neck cancer. Gendicine has also been epigenetic drugs which inhibit DNA methylation and histone
investigated for its efcacy when administered in combination deacetylation and thereby restore silenced miRNA-34a expres-
with fractionated stereotactic radiotherapy for the treatment of sion such as a combination of 5-Aza CDR and Vorinostat in the
hepatocellular carcinoma and showed improved one-year survival setting of pancreatic cancer [137]. Secondly, systemic delivery of
rates [120]. In the USA, adenoviral p53 vectors such as SCH-58500 miRNA-34a mimics such as MRX34 developed by Mirna
(CANJI, Inc.) and Advexin (Introgen Therapeutics) have been Therapeutics which is currently in phase I clinical trial as a
investigated in various clinical trials [121,122]. However, compa- monotherapy. Using this approach, partial responses have been
nies pursuing this avenue have since closed down like Advexin, or reported in hepatocellular carcinoma, renal cell carcinoma and
been bought by Merck like CANJI, Inc. and no active clinical trials acral melanoma patients, although in three patients serious
can currently be found on databases suggesting severe problems adverse events that were possibly immune-related were reported
associated with this strategy. at the ASCO 2016 meeting [138]. Therapeutic approaches
employing miRNA are still hampered by issues of poor cellular
4.3. TP53 oncolytic viruses uptake and will need careful monitoring for potential side effects
due to their multiple targets.
Because of the low transduction efciencies of TP53 via
adenoviral-based vectors, other strategies have been developed. 5. Outlook
This approach uses replicating viruses for selective killing of tumor
cells and is based on the long-known observation that tumors To guard the integrity of a genome one has to restore the
regress following viral infection, which may be caused by normal function of the guardian of the genome, namely p53.
interferon and/or acquired immune responses [123,124]. An There are a plethora of direct or indirect targeting approaches
example of TP53-based oncolytic virotherapy is dl-1520 (Onyx- that all have promising and important clinical implications: from
015) in which the viral E1B-55KDa gene has been deleted. E1B those aiming either directly at, or upstream of p53, e.g. MDM2/
proteins interact with p53, inducing its degradation, therefore MDMX, to immunotherapy and vaccination strategies, to miRNA
enabling host cell survival and subsequent viral replication. Hence, targeting. In the clinical setting it will be important to tailor the
in TP53-decient tumors, viruses lacking the E1B proteins are able strategy to the specic tumor type and its molecular status
to replicate, whereas in normal cells p53 is activated preventing concerning p53, MDM and other biomarkers that may eventually
cellular survival and as a corollary viral replication. As such Onyx- be unraveled.
015 presents a relatively specic therapy whereby virus replication To date, small molecule inhibitors are those holding the greatest
only occurs in tumor cells that lack p53 resulting in cytolysis [125]. promise. Their size, polarity, solubility, pharmacokinetic and
Onyx-015 was investigated in numerous phase I/II clinical trials of pharmacodynamic properties can be improved through sophisti-
malignant gliomas, hepatobiliary tumors and sarcomas [126128]. cated medical chemistry approaches. Furthermore, structural
However, the clinical response to Onyx-015 was mixed, it reached modeling and drug-target protein interaction studies have
phase III in combination with chemotherapy for head and neck revealed a detailed, atomic-level understanding of vulnerable
cancer, but the trial was not completed and Onyx-015 did not nodes of p53 or MDM family members and how key protein
obtain FDA approval. For different reasons, further clinical interphases or pockets can be targeted. p53 cancer research
development was not pursued. A further oncolytic virus, activity in the future will include advanced structure-guided
H101 was developed and licensed for cancer therapy in China. modeling combined with biological read-out systems. Also, in vivo
H101 lacks the viral E1B genes in addition to E3 which encodes the targeting in advanced p53 animal model systems and related early
adenovirus death protein [129]. In phase II/III trials for treatment of phase clinical trials will bring new momentum. Studies of
head and neck carcinomas, it showed synergistic antitumor effects inhibitors of MDM2 are good examples of how basic research
in combination with chemotherapy. Two further oncolytic viruses into regulatory factors of p53 has enabled the development of new
are currently under investigation: H103, which has completed drugs for research and for therapeutic application. Other recently
phase I trials and H102 which is still in the preclinical phase identied non-MDM-targeting negative regulators of p53, includ-
[130,131]. ing PIRH2, COP1 and TRIM24, also appear promising as small
molecule inhibitors [139141]. Assessment of the mechanisms and
4.4. MicroRNA based strategies efcacy of these agents in murine models that overexpress PIRH2,
COP1 or TRIM24 is likely to generate new insights. Novel
MicroRNAs (miRNA) are small noncoding RNAs with a typical p53 regulators, some of them even cancer type-specic, will likely
length of around 20 nucleotides. miRNAs usually bind to the 3- be identied in the near future from ongoing cancer genome DNA
untranslated region of their target mRNAs and can either lead to sequencing efforts and searches for mutual exclusivity with
degradation of the target mRNA or block translational protein p53 aberrations. Finally, the transcriptional p53 targets HAUSP
synthesis. Most miRNAs regulate multiple target genes and play and WIP1 that stabilize MDM2 protein, thus creating a regulatory
an important role in physiological and pathological processes feedback loop, are natural targets for future therapeutic inhibition
including cancer [132]. Recent investigations have highlighted in cancers where one of them is amplied (e.g. invasive breast
the contribution of miRNA in p53 regulation. A group of miRNAs cancer or pancreatic neuroendocrine tumors; TCGA database)
including miR-25, miR-30d and miR-125b have been shown to [142,143]. In regards to gene therapy, future work and improve-
directly downregulate p53 expression whereas others including ment of technology e.g. CRSPR/CAS9 or the use of engineered
miR-192, miR-194, miR-215 and miR-605 indirectly stabilize viruses may solve the problems associated with traditional
10 O. Merkel et al. / Mutation Research 773 (2017) 113

approaches [144]. Despite the dwindling number of gene therapy [12] A. Ventura, D.G. Kirsch, M.E. McLaughlin, D.A. Tuveson, J. Grimm, L. Lintault,
trials following the dramatic setbacks caused by the development et al., Restoration of p53 function leads to tumour regression in vivo, Nature
445 (7128) (2007) 661665, doi:http://dx.doi.org/10.1038/nature05541.
of deadly leukemia in 9 out of 9 pediatric patients after viral gene [13] J. Pencik, H.T. Pham, J. Schmoellerl, T. Javaheri, M. Schlederer, Z. Culig, et al.,
therapy for X-linked severe combined immunodeciency [145], JAK-STAT signaling in cancer: from cytokines to non-coding genome,
site-specic, targeted integration of virally delivered genes into the Cytokine 87 (2016) 2636, doi:http://dx.doi.org/10.1016/j.cyto.2016.06.017.
[14] J. Pencik, M. Schlederer, W. Gruber, C. Unger, S.M. Walker, A. Chalaris, et al.,
genome of patient cells is an active area of research. Newly STAT3 regulated ARF expression suppresses prostate cancer metastasis, Nat.
engineered viruses derived from the lentivirus HIV and the adeno- Commun. 6 (2015) 7736, doi:http://dx.doi.org/10.1038/ncomms8736.
associated virus (AAV) which belongs to the parvovirus family are [15] L. Bouaoun, D. Sonkin, M. Ardin, M. Hollstein, G. Byrnes, J. Zavadil, M. Olivier,
TP53 variations in human cancers: new lessons from the IARC TP53 database
potential alternatives [146]. Finally, the advent of CRISPR/ and genomics data, Hum. Mutat. 37 (9) (2016) 865876, doi:http://dx.doi.
CAS9 technology allowing for gene insertion at dened sites in org/10.1002/humu.23035.
the genome may overcome the oncogenicity observed with classic [16] V.J. Bykov, N. Issaeva, A. Shilov, M. Hultcrantz, E. Pugacheva, P. Chumakov,
et al., Restoration of the tumor suppressor function to mutant p53 by a low-
viral approaches [147].
molecular-weight compound, Nat. Med. 8 (3) (2002) 282288, doi:http://dx.
It is remarkable how technological and scientic developments doi.org/10.1038/nm0302-282.
in cancer research are mirrored in the evolution of drugs targeting [17] C. Bou-Hanna, A. Jarry, L. Lode, I. Schmitz, K. Schulze-Osthoff, S. Kury, et al.,
a single pathway: that of the p53 tumor suppressor. In the future Acute cytotoxicity of MIRA-1/NSC19630, a mutant p53-reactivating small
molecule, against human normal and cancer cells via a caspase-9-dependent
we expect that basic research into p53 and its direct and indirect apoptosis, Cancer Lett. 359 (2) (2015) 211217, doi:http://dx.doi.org/10.1016/
regulators will continue to generate insights that inspire academia j.canlet.2015.01.014.
and pharma companies to develop new, improved therapeutics for [18] V.J. Bykov, N. Issaeva, N. Zache, A. Shilov, M. Hultcrantz, J. Bergman, et al.,
Reactivation of mutant p53 and induction of apoptosis in human tumor cells
cancer patients. by maleimide analogs, J. Biol. Chem. 280 (34) (2005) 3038430391, doi:
http://dx.doi.org/10.1074/jbc.M501664200.
Acknowledgements [19] J. Shen, H. Vakifahmetoglu, H. Stridh, B. Zhivotovsky, K.G. Wiman, PRIMA-
1MET induces mitochondrial apoptosis through activation of caspase-2,
Oncogene 27 (51) (2008) 65716580, doi:http://dx.doi.org/10.1038/
We thank Brigitte Hantusch and Gertrude Krainz for proof- onc.2008.249.
reading the manuscript. This study was supported by an EU Marie [20] T. Wang, K. Lee, A. Rehman, S.S. Daoud, PRIMA-1 induces apoptosis by
inhibiting JNK signaling but promoting the activation of Bax, Biochem.
Curie Actions Innovative Training Network grant (ALKATRAS
Biophys. Res. Commun. 352 (1) (2007) 203212, doi:http://dx.doi.org/
675712). We are also grateful to collaborators of the European 10.1016/j.bbrc.2006.11.006.
Research Initiative of ALK-related malignancies (www.erialcl.net). [21] J.M. Lambert, P. Gorzov, D.B. Veprintsev, M. Soderqvist, D. Segerback, J.
Bergman, et al., PRIMA-1 reactivates mutant p53 by covalent binding to the
We thank the Jubilumsfond der sterreichischen Nationalbank
core domain, Cancer Cell 15 (5) (2009) 376388, doi:http://dx.doi.org/
(Grant No. 14856, to OM) and the Fonds der Stadt Wien fur 10.1016/j.ccr.2009.03.003.
innovative interdisziplinre Krebsforschung (The role of Brg1 in [22] R. Zandi, G. Selivanova, C.L. Christensen, T.A. Gerds, B.M. Willumsen, H.S.
Anaplastic Large Cell Lymphoma, to OM) as well as the Austrian Poulsen, PRIMA-1Met/APR-246 induces apoptosis and tumor growth delay in
small cell lung cancer expressing mutant p53, Clin. Cancer Res. 17 (9) (2011)
Science Funds (FWF; Grant Nos P 26011 and P 29251, to LK). SDT is a 28302841, doi:http://dx.doi.org/10.1158/1078-0432.ccr-10-3168.
Senior Lecturer supported by funding from Bloodwise. RM was [23] M.N. Saha, H. Jiang, Y. Yang, D. Reece, H. Chang, PRIMA-1Met/APR-
supported by grants SFB F47 (SFB F4707-B20) from the Austrian 246 displays high antitumor activity in multiple myeloma by induction of
p73 and Noxa, Mol. Cancer Ther. 12 (11) (2013) 23312341, doi:http://dx.doi.
Science Fund. org/10.1158/1535-7163.mct-12-1166.
[24] W. Bao, M. Chen, X. Zhao, R. Kumar, C. Spinnler, M. Thullberg, et al., PRIMA-
1Met/APR-246 induces wild-type p53-dependent suppression of malignant
References melanoma tumor growth in 3D culture and in vivo, ABBV Cell Cycle 10 (2)
(2011) 301307, doi:http://dx.doi.org/10.4161/cc.10.2.14538.
[1] B. Vogelstein, D. Lane, A.J. Levine, Surng the p53 network, Nature 408 (6810) [25] Y. Liang, C. Besch-Williford, S.M. Hyder, PRIMA-1 inhibits growth of breast
(2000) 307310. cancer cells by re-activating mutant p53 protein, Int. J. Oncol. 35 (5) (2009)
[2] G. Liu, T.J. McDonnell, R. Montes de Oca Luna, M. Kapoor, B. Mims, A.K. El- 10151023.
Naggar, G. Lozano, High metastatic potential in mice inheriting a targeted [26] S. Lehmann, V.J. Bykov, D. Ali, O. Andren, H. Cherif, U. Tidefelt, et al., Targeting
p53 missense mutation, Proc. Natl. Acad. Sci. U. S. A. 97 (8) (2000) 41744179. p53 in vivo: a rst-in-human study with p53-targeting compound APR-
[3] S. Xiong, H. Tu, M. Kollareddy, V. Pant, Q. Li, Y. Zhang, et al., 246 in refractory hematologic malignancies and prostate cancer, J. Clin.
Pla2g16 phospholipase mediates gain-of-function activities of mutant p53, Oncol. 30 (29) (2012) 36333639, doi:http://dx.doi.org/10.1200/
Proc. Natl. Acad. Sci. U. S. A. 111 (30) (2014) 1114511150, doi:http://dx.doi. jco.2011.40.7783.
org/10.1073/pnas.1404139111. [27] N. Rokaeus, J. Shen, I. Eckhardt, V.J. Bykov, K.G. Wiman, M.T. Wilhelm, PRIMA-
[4] Y.X. Cui, A. Kerby, F.K. McDuff, H. Ye, S.D. Turner, NPM-ALK inhibits the 1(MET)/APR-246 targets mutant forms of p53 family members p63 and p73,
p53 tumor suppressor pathway in an MDM2 and JNK-dependent manner, Oncogene 29 (49) (2010) 64426451, doi:http://dx.doi.org/10.1038/
Blood 113 (21) (2009) 52175227. onc.2010.382.
[5] F. Nikulenkov, C. Spinnler, H. Li, C. Tonelli, Y. Shi, M. Turunen, et al., Insights [28] R. Shalom-Feuerstein, L. Serror, E. Aberdam, F.J. Muller, H. van Bokhoven, K.G.
into p53 transcriptional function via genome-wide chromatin occupancy and Wiman, et al., Impaired epithelial differentiation of induced pluripotent stem
gene expression analysis, Cell Death Differ. 19 (12) (2012) 19922002, doi: cells from ectodermal dysplasia-related patients is rescued by the small
http://dx.doi.org/10.1038/cdd.2012.89. compound APR-246/PRIMA-1MET, Proc. Natl. Acad. Sci. U. S. A. 110 (6) (2013)
[6] T. Riley, E. Sontag, P. Chen, A. Levine, Transcriptional control of human p53- 21522156, doi:http://dx.doi.org/10.1073/pnas.1201753109.
regulated genes, Nat. Rev. Mol. Cell Biol. 9 (5) (2008) 402412, doi:http://dx. [29] J. Shen, E.H. van den Bogaard, E.N. Kouwenhoven, V.J. Bykov, T. Rinne, Q.
doi.org/10.1038/nrm2395. Zhang, et al., APR-246/PRIMA-1(MET) rescues epidermal differentiation in
[7] D. Asslaber, J.D. Pinon, I. Seyfried, P. Desch, M. Stocher, I. Tinhofer, et al., skin keratinocytes derived from EEC syndrome patients with p63 mutations,
microRNA-34a expression correlates with MDM2 SNP309 polymorphism Proc. Natl. Acad. Sci. U. S. A. 110 (6) (2013) 21572162, doi:http://dx.doi.org/
and treatment-free survival in chronic lymphocytic leukemia, Blood 115 (21) 10.1073/pnas.1201993110.
(2010) 41914197, doi:http://dx.doi.org/10.1182/blood-2009-07-234823. [30] F.M. Boeckler, A.C. Joerger, G. Jaggi, T.J. Rutherford, D.B. Veprintsev, A.R.
[8] U.M. Moll, O. Petrenko, The MDM2-p53 interaction, Mol. Cancer Res. 1 (14) Fersht, Targeted rescue of a destabilized mutant of p53 by an in silico
(2003) 10011008. screened drug, Proc. Natl. Acad. Sci. U. S. A. 105 (30) (2008) 1036010365, doi:
[9] L.K. Linares, A. Hengstermann, A. Ciechanover, S. Muller, M. Scheffner, HdmX http://dx.doi.org/10.1073/pnas.0805326105.
stimulates Hdm2-mediated ubiquitination and degradation of p53, Proc. [31] X. Liu, R. Wilcken, A.C. Joerger, I.S. Chuckowree, J. Amin, J. Spencer, A.R. Fersht,
Natl. Acad. Sci. U. S. A. 100 (21) (2003) 1200912014, doi:http://dx.doi.org/ Small molecule induced reactivation of mutant p53 in cancer cells, Nucleic
10.1073/pnas.2030930100. Acids Res. 41 (12) (2013) 60346044, doi:http://dx.doi.org/10.1093/nar/
[10] W. Ma, B.M. Gilligan, J. Yuan, T. Li, Current status and perspectives in gkt305.
translational biomarker research for PD-1/PD-L1 immune checkpoint [32] B.A. Foster, H.A. Coffey, M.J. Morin, F. Rastinejad, Pharmacological rescue of
blockade therapy, J. Hematol. Oncol. 9 (1) (2016) 47, doi:http://dx.doi.org/ mutant p53 conformation and function, Science 286 (5449) (1999) 25072510.
10.1186/s13045-016-0277-y. [33] T.M. Rippin, V.J. Bykov, S.M. Freund, G. Selivanova, K.G. Wiman, A.R. Fersht,
[11] C.P. Martins, L. Brown-Swigart, G.I. Evan, Modeling the therapeutic efcacy of Characterization of the p53-rescue drug CP-31398 in vitro and in living cells,
p53 restoration in tumors, Cell 127 (7) (2006) 13231334, doi:http://dx.doi. Oncogene 21 (14) (2002) 21192129, doi:http://dx.doi.org/10.1038/sj.
org/10.1016/j.cell.2006.12.007. onc.1205362.
O. Merkel et al. / Mutation Research 773 (2017) 113 11

[34] C.V. Rao, M.V. Swamy, J.M. Patlolla, L. Kopelovich, Suppression of familial [54] E.M. Alexandrova, A.R. Yallowitz, D. Li, S. Xu, R. Schulz, D.A. Proia, et al.,
adenomatous polyposis by CP-31398, a TP53 modulator, in APCmin/+ mice, Improving survival by exploiting tumour dependence on stabilized mutant
Cancer Res. 68 (18) (2008) 76707675, doi:http://dx.doi.org/10.1158/0008- p53 for treatment, Nature 523 (7560) (2015) 352356, doi:http://dx.doi.org/
5472.can-08-1610. 10.1038/nature14430.
[35] X. Tang, Y. Zhu, L. Han, A.L. Kim, L. Kopelovich, D.R. Bickers, M. Athar, CP- [55] S. Ramalingam, G. Goss, R. Rosell, G. Schmid-Bindert, B. Zaric, Z. Andric, et al.,
31398 restores mutant p53 tumor suppressor function and inhibits UVB- A randomized phase II study of ganetespib, a heat shock protein 90 inhibitor,
induced skin carcinogenesis in mice, J. Clin. Invest. 117 (12) (2007) 3753 in combination with docetaxel in second-line therapy of advanced non-small
3764, doi:http://dx.doi.org/10.1172/JCI32481. cell lung cancer (GALAXY-1), Ann. Oncol. 26 (8) (2015) 17411748, doi:http://
[36] C. Fiorini, M. Menegazzi, C. Padroni, I. Dando, E. Dalla Pozza, A. Gregorelli, dx.doi.org/10.1093/annonc/mdv220.
et al., Autophagy induced by p53-reactivating molecules protects pancreatic [56] D. Li, N.D. Marchenko, U.M. Moll, SAHA shows preferential cytotoxicity in
cancer cells from apoptosis, Apoptosis 18 (3) (2013) 337346, doi:http://dx. mutant p53 cancer cells by destabilizing mutant p53 through inhibition of
doi.org/10.1007/s10495-012-0790-6. the HDAC6-Hsp90 chaperone axis, Cell Death Differ. 18 (12) (2011) 1904
[37] H.S. Sutherland, I.Y. Hwang, E.S. Marshall, B.S. Lindsay, W.A. Denny, C. 1913, doi:http://dx.doi.org/10.1038/cdd.2011.71.
Gilchrist, et al., Therapeutic reactivation of mutant p53 protein by [57] L. Ding, Z. Zhang, G. Liang, Z. Yao, H. Wu, B. Wang, et al., SAHA triggered MET
quinazoline derivatives, Invest. New Drugs 30 (5) (2012) 20352045, doi: activation contributes to SAHA tolerance in solid cancer cells, Cancer Lett 356
http://dx.doi.org/10.1007/s10637-011-9744-z. (2 Pt B) (2015) 828836, doi:http://dx.doi.org/10.1016/j.canlet.2014.10.034.
[38] M. Demma, E. Maxwell, R. Ramos, L. Liang, C. Li, D. Hesk, et al., SCH529074, a [58] V. Lallemand-Breitenbach, H. de The, Retinoic acid plus arsenic trioxide, the
small molecule activator of mutant p53, which binds p53 DNA binding ultimate panacea for acute promyelocytic leukemia? Blood 122 (12) (2013)
domain (DBD), restores growth-suppressive function to mutant p53 and 20082010, doi:http://dx.doi.org/10.1182/blood-2013-06-505115.
interrupts HDM2-mediated ubiquitination of wild type p53, J. Biol. Chem. [59] O. Merkel, C. Heyder, D. Asslaber, F. Hamacher, I. Tinhofer, C. Holler, et al.,
285 (14) (2010) 1019810212, doi:http://dx.doi.org/10.1074/jbc. Arsenic trioxide induces apoptosis preferentially in B-CLL cells of patients
M109.083469. with unfavourable prognostic factors including del17p13, J. Mol. Med. (Berl.)
[39] L.M. Shi, T.G. Myers, Y. Fan, P.M. O'Connor, K.D. Paull, S.H. Friend, J.N. 86 (5) (2008) 541552, doi:http://dx.doi.org/10.1007/s00109-008-0314-6.
Weinstein, Mining the National Cancer Institute Anticancer Drug Discovery [60] W. Yan, Y.S. Jung, Y. Zhang, X. Chen, Arsenic trioxide reactivates proteasome-
Database: cluster analysis of ellipticine analogs with p53-inverse and central dependent degradation of mutant p53 protein in cancer cells in part via
nervous system-selective patterns of activity, Mol. Pharmacol. 53 (2) (1998) enhanced expression of Pirh2 E3 ligase, PLoS One 9 (8) (2014) e103497, doi:
241251. http://dx.doi.org/10.1371/journal.pone.0103497.
[40] Y. Peng, C. Li, L. Chen, S. Sebti, J. Chen, Rescue of mutant p53 transcription [61] T.J. Robinson, M. Pai, J.C. Liu, F. Vizeacoumar, T. Sun, S.E. Egan, et al., High-
function by ellipticine, Oncogene 22 (29) (2003) 44784487, doi:http://dx. throughput screen identies disulram as a potential therapeutic for triple-
doi.org/10.1038/sj.onc.1206777. negative breast cancer cells: interaction with IQ motif-containing factors,
[41] F. Wang, J. Liu, D. Robbins, K. Morris, A. Sit, Y.Y. Liu, Y. Zhao, Mutant ABBV Cell Cycle 12 (18) (2013) 30133024, doi:http://dx.doi.org/10.4161/
p53 exhibits trivial effects on mitochondrial functions which can be cc.26063.
reactivated by ellipticine in lymphoma cells, Apoptosis 16 (3) (2011) 301310, [62] H. Nechushtan, Y. Hamamreh, S. Nidal, M. Gotfried, A. Baron, Y.I. Shalev, et al.,
doi:http://dx.doi.org/10.1007/s10495-010-0559-8. A phase IIb trial assessing the addition of disulram to chemotherapy for the
[42] J. Poljakova, T. Eckschlager, J. Hrabeta, J. Hrebackova, S. Smutny, E. Frei, et al., treatment of metastatic non-small cell lung cancer, Oncologist 20 (4) (2015)
The mechanism of cytotoxicity and DNA adduct formation by the anticancer 366367, doi:http://dx.doi.org/10.1634/theoncologist.2014-0424.
drug ellipticine in human neuroblastoma cells, Biochem. Pharmacol. 77 (9) [63] J. Wang, Q. Zhao, Q. Qi, H.Y. Gu, J.J. Rong, R. Mu, et al., Gambogic acid-induced
(2009) 14661479, doi:http://dx.doi.org/10.1016/j.bcp.2009.01.021. degradation of mutant p53 is mediated by proteasome and related to CHIP, J.
[43] N. Issaeva, P. Bozko, M. Enge, M. Protopopova, L.G. Verhoef, M. Masucci, et al., Cell. Biochem. 112 (2) (2011) 509519, doi:http://dx.doi.org/10.1002/
Small molecule RITA binds to p53, blocks p53-HDM-2 interaction and jcb.22941.
activates p53 function in tumors, Nat. Med. 10 (12) (2004) 13211328, doi: [64] J. Liu, H. Xia, M. Kim, L. Xu, Y. Li, L. Zhang, et al., Beclin1 controls the levels of
http://dx.doi.org/10.1038/nm1146. p53 by regulating the deubiquitination activity of USP10 and USP13, Cell 147
[44] J. Yang, A. Ahmed, E. Poon, N. Perusinghe, A. de Haven Brandon, G. Box, et al., (1) (2011) 223234, doi:http://dx.doi.org/10.1016/j.cell.2011.08.037.
Small-molecule activation of p53 blocks hypoxia-inducible factor 1alpha and [65] H. Vakifahmetoglu-Norberg, M. Kim, H.G. Xia, M.P. Iwanicki, D. Ofengeim, J.L.
vascular endothelial growth factor expression in vivo and leads to tumor cell Coloff, et al., Chaperone-mediated autophagy degrades mutant p53, Genes
apoptosis in normoxia and hypoxia, Mol. Cell. Biol. 29 (8) (2009) 22432253, Dev. 27 (15) (2013) 17181730, doi:http://dx.doi.org/10.1101/gad.220897.113.
doi:http://dx.doi.org/10.1128/mcb.00959-08. [66] J.E. Kravchenko, G.V. Ilyinskaya, P.G. Komarov, L.S. Agapova, D.V. Kochetkov, E.
[45] M. Enge, W. Bao, E. Hedstrom, S.P. Jackson, A. Moumen, G. Selivanova, MDM2- Strom, et al., Small-molecule RETRA suppresses mutant p53-bearing cancer
dependent downregulation of p21 and hnRNP K provides a switch between cells through a p73-dependent salvage pathway, Proc. Natl. Acad. Sci. U. S. A.
apoptosis and growth arrest induced by pharmacologically activated p53, 105 (17) (2008) 63026307, doi:http://dx.doi.org/10.1073/pnas.0802091105.
Cancer Cell 15 (3) (2009) 171183, doi:http://dx.doi.org/10.1016/j. [67] W.A. Freed-Pastor, H. Mizuno, X. Zhao, A. Langerod, S.H. Moon, R. Rodriguez-
ccr.2009.01.019. Barrueco, et al., Mutant p53 disrupts mammary tissue architecture via the
[46] C.Y. Zhao, L. Szekely, W. Bao, G. Selivanova, Rescue of p53 function by small- mevalonate pathway, Cell 148 (12) (2012) 244258, doi:http://dx.doi.org/
molecule RITA in cervical carcinoma by blocking E6-mediated degradation, 10.1016/j.cell.2011.12.017.
Cancer Res. 70 (8) (2010) 33723381, doi:http://dx.doi.org/10.1158/0008- [68] J. Zhu, Z. Dou, M.A. Sammons, A.J. Levine, S.L. Berger, Lysine methylation
5472.CAN-09-2787. represses p53 activity in teratocarcinoma cancer cells, Proc. Natl. Acad. Sci. U.
[47] M. Burmakin, Y. Shi, E. Hedstrom, P. Kogner, G. Selivanova, Dual targeting of S. A. 113 (35) (2016) 98229827, doi:http://dx.doi.org/10.1073/
wild-type and mutant p53 by small molecule RITA results in the inhibition of pnas.1610387113.
N-Myc and key survival oncogenes and kills neuroblastoma cells in vivo and [69] J. Zhu, M.A. Sammons, G. Donahue, Z. Dou, M. Vedadi, M. Getlik, et al., Gain-
in vitro, Clin. Cancer Res. 19 (18) (2013) 50925103, doi:http://dx.doi.org/ of-function p53 mutants co-opt chromatin pathways to drive cancer growth,
10.1158/1078-0432.CCR-12-2211. Nature 525 (7568) (2015) 206211, doi:http://dx.doi.org/10.1038/
[48] M. Wanzel, J.B. Vischedyk, M.P. Gittler, N. Gremke, J.R. Seiz, M. Hefter, et al., nature15251.
CRISPR-Cas9-based target validation for p53-reactivating model compounds, [70] L. Cahilly-Snyder, T. Yang-Feng, U. Francke, D.L. George, Molecular analysis
Nat. Chem. Biol. 12 (1) (2016) 2228, doi:http://dx.doi.org/10.1038/ and chromosomal mapping of amplied genes isolated from a transformed
nchembio.1965. mouse 3T3 cell line, Somat. Cell Mol. Genet. 13 (3) (1987) 235244.
[49] R. Puca, L. Nardinocchi, G. Bossi, A. Sacchi, G. Rechavi, D. Givol, G. D'Orazi, [71] S. Francoz, P. Froment, S. Bogaerts, S. De Clercq, M. Maetens, G. Doumont,
Restoring wtp53 activity in HIPK2 depleted MCF7 cells by modulating et al., Mdm4 and Mdm2 cooperate to inhibit p53 activity in proliferating and
metallothionein and zinc, Exp. Cell Res. 315 (1) (2009) 6775, doi:http://dx. quiescent cells in vivo, Proc. Natl. Acad. Sci. U. S. A. 103 (9) (2006) 32323237,
doi.org/10.1016/j.yexcr.2008.10.018. doi:http://dx.doi.org/10.1073/pnas.0508476103.
[50] A.R. Blanden, X. Yu, A.J. Wolfe, J.A. Gilleran, D.J. Augeri, R.S. O'Dell, et al., [72] J. Momand, G.P. Zambetti, D.C. Olson, D. George, A.J. Levine, The mdm-
Synthetic metallochaperone ZMC1 rescues mutant p53 conformation by 2 oncogene product forms a complex with the p53 protein and inhibits p53-
transporting zinc into cells as an ionophore, Mol. Pharmacol. 87 (5) (2015) mediated transactivation, Cell 69 (7) (1992) 12371245.
825831, doi:http://dx.doi.org/10.1124/mol.114.097550. [73] X. Wu, J.H. Bayle, D. Olson, A.J. Levine, The p53-mdm-2 autoregulatory
[51] X. Yu, A.R. Blanden, S. Narayanan, L. Jayakumar, D. Lubin, D. Augeri, feedback loop, Genes Dev. 7 (7A) (1993) 11261132.
et al., Small molecule restoration of wildtype structure and function of [74] P.H. Kussie, S. Gorina, V. Marechal, B. Elenbaas, J. Moreau, A.J. Levine, N.P.
mutant p53 using a novel zinc-metallochaperone based mechanism, Pavletich, Structure of the MDM2 oncoprotein bound to the p53 tumor
Oncotarget 5 (19) (2014) 88798892, doi:http://dx.doi.org/10.18632/ suppressor transactivation domain, Science 274 (5289) (1996) 948953.
oncotarget.2432. [75] F. Bernal, A.F. Tyler, S.J. Korsmeyer, L.D. Walensky, G.L. Verdine, Reactivation
[52] X. Yu, A. Vazquez, A.J. Levine, D.R. Carpizo, Allele-specic p53 mutant of the p53 tumor suppressor pathway by a stapled p53 peptide, J. Am. Chem.
reactivation, Cancer Cell 21 (5) (2012) 614625, doi:http://dx.doi.org/ Soc. 129 (9) (2007) 24562457, doi:http://dx.doi.org/10.1021/ja0693587.
10.1016/j.ccr.2012.03.042. [76] M. Pazgier, M. Liu, G. Zou, W. Yuan, C. Li, J. Li, et al., Structural basis for high-
[53] D. Li, N.D. Marchenko, R. Schulz, V. Fischer, T. Velasco-Hernandez, F. Talos, U. afnity peptide inhibition of p53 interactions with MDM2 and MDMX, Proc.
M. Moll, Functional inactivation of endogenous MDM2 and CHIP by Natl. Acad. Sci. U. S. A. 106 (12) (2009) 46654670, doi:http://dx.doi.org/
HSP90 causes aberrant stabilization of mutant p53 in human cancer cells, 10.1073/pnas.0900947106.
Mol. Cancer Res. 9 (5) (2011) 577588, doi:http://dx.doi.org/10.1158/1541- [77] J. Phan, Z. Li, A. Kasprzak, B. Li, S. Sebti, W. Guida, et al., Structure-based
7786.MCR-10-0534. design of high afnity peptides inhibiting the interaction of p53 with
12 O. Merkel et al. / Mutation Research 773 (2017) 113

MDM2 and MDMX, J. Biol. Chem. 285 (3) (2010) 21742183, doi:http://dx.doi. [99] D. Sun, Z. Li, Y. Rew, M. Gribble, M.D. Bartberger, H.P. Beck, et al., Discovery of
org/10.1074/jbc.M109.073056. AMG 232, a potent, selective, and orally bioavailable MDM2-p53 inhibitor in
[78] C.J. Brown, S.T. Quah, J. Jong, A.M. Goh, P.C. Chiam, K.H. Khoo, et al., Stapled clinical development, J. Med. Chem. 57 (4) (2014) 14541472, doi:http://dx.
peptides with improved potency and specicity that activate p53, ACS Chem. doi.org/10.1021/jm401753e.
Biol. 8 (3) (2013) 506512, doi:http://dx.doi.org/10.1021/cb3005148. [100] B. Graves, T. Thompson, M. Xia, C. Janson, C. Lukacs, D. Deo, et al., Activation of
[79] L.T. Vassilev, B.T. Vu, B. Graves, D. Carvajal, F. Podlaski, Z. Filipovic, et al., In the p53 pathway by small-molecule-induced MDM2 and MDMX
vivo activation of the p53 pathway by small-molecule antagonists of MDM2, dimerization, Proc. Natl. Acad. Sci. U. S. A. 109 (29) (2012) 1178811793, doi:
Science 303 (5659) (2004) 844848, doi:http://dx.doi.org/10.1126/ http://dx.doi.org/10.1073/pnas.1203789109.
science.1092472. [101] M. Wade, L.W. Rodewald, J.M. Espinosa, G.M. Wahl, BH3 activation blocks
[80] L. Coll-Mulet, D. Iglesias-Serret, A.F. Santidrian, A.M. Cosialls, M. de Frias, E. Hdmx suppression of apoptosis and cooperates with Nutlin to induce cell
Castano, et al., MDM2 antagonists activate p53 and synergize with genotoxic death, ABBV Cell Cycle 7 (13) (2008) 19731982.
drugs in B-cell chronic lymphocytic leukemia cells, Blood 107 (10) (2006) [102] M. Wade, E.T. Wong, M. Tang, J.M. Stommel, G.M. Wahl, Hdmx modulates the
41094114. outcome of p53 activation in human tumor cells, J. Biol. Chem. 281 (44)
[81] K. Kojima, M. Konopleva, T. McQueen, S. O'Brien, W. Plunkett, M. Andreeff, (2006) 3303633044, doi:http://dx.doi.org/10.1074/jbc.M605405200.
Mdm2 inhibitor Nutlin-3a induces p53-mediated apoptosis by transcription- [103] D. Reed, Y. Shen, A.A. Shelat, L.A. Arnold, A.M. Ferreira, F. Zhu, et al.,
dependent and transcription-independent mechanisms and may overcome Identication and characterization of the rst small molecule inhibitor of
Mdm2 and Atm-mediated resistance to udarabine in chronic lymphocytic MDMX, J. Biol. Chem. 285 (14) (2010) 1078610796, doi:http://dx.doi.org/
leukemia, Blood (2006). 10.1074/jbc.M109.056747.
[82] P. Secchiero, E. Barbarotto, M. Tiribelli, C. Zerbinati, M.G. di Iasio, A. Gonelli, [104] H. Wang, X. Ma, S. Ren, J.K. Buolamwini, C. Yan, A small-molecule inhibitor of
et al., Functional integrity of the p53-mediated apoptotic pathway induced MDMX activates p53 and induces apoptosis, Mol. Cancer Ther. 10 (1) (2011)
by the nongenotoxic agent nutlin-3 in B-cell chronic lymphocytic leukemia 6979, doi:http://dx.doi.org/10.1158/1535-7163.MCT-10-0581.
(B-CLL), Blood 107 (10) (2006) 41224129. [105] S.M. Vogel, M.R. Bauer, A.C. Joerger, R. Wilcken, T. Brandt, D.B. Veprintsev,
[83] K. Kojima, M. Konopleva, I.J. Samudio, M. Shikami, M. Cabreira-Hansen, T. et al., Lithocholic acid is an endogenous inhibitor of MDM4 and MDM2, Proc.
McQueen, et al., MDM2 antagonists induce p53-dependent apoptosis in Natl. Acad. Sci. U. S. A. 109 (42) (2012) 1690616910, doi:http://dx.doi.org/
AML: implications for leukemia therapy, Blood 106 (9) (2005) 31503159. 10.1073/pnas.1215060109.
[84] T. Van Maerken, L. Ferdinande, J. Taildeman, I. Lambertz, N. Yigit, L. [106] W. Dai, J. Zhou, B. Jin, J. Pan, Class III-specic HDAC inhibitor Tenovin-
Vercruysse, et al., Antitumor activity of the selective MDM2 antagonist 6 induces apoptosis, suppresses migration and eliminates cancer stem cells
nutlin-3 against chemoresistant neuroblastoma with wild-type p53, J. Natl. in uveal melanoma, Sci. Rep. 6 (2016) 22622, doi:http://dx.doi.org/10.1038/
Cancer Inst. 101 (22) (2009) 15621574, doi:http://dx.doi.org/10.1093/jnci/ srep22622.
djp355. [107] Y. Jin, Q. Cao, C. Chen, X. Du, B. Jin, J. Pan, Tenovin-6-mediated inhibition of
[85] Y. Cui, H. Zhang, J. Meadors, R. Poon, M. Guimond, C.L. Mackall, Harnessing SIRT1/2 induces apoptosis in acute lymphoblastic leukemia (ALL) cells and
the physiology of lymphopenia to support adoptive immunotherapy in eliminates ALL stem/progenitor cells, BMC Cancer 15 (2015) 226, doi:http://
lymphoreplete hosts, Blood 114 (18) (2009) 38313840, doi:http://dx.doi. dx.doi.org/10.1186/s12885-015-1282-1.
org/10.1182/blood-2009-03-212134. [108] J.E. Bauman, C.H. Chung, CHK it out! Blocking WEE kinase routs TP53 mutant
[86] S. Surget, D. Chiron, P. Gomez-Bougie, G. Descamps, E. Menoret, R. Bataille, cancer, Clin. Cancer Res. 20 (16) (2014) 41734175, doi:http://dx.doi.org/
et al., Cell death via DR5, but not DR4, is regulated by p53 in myeloma cells, 10.1158/1078-0432.CCR-14-0720.
Cancer Res. 72 (17) (2012) 45624573, doi:http://dx.doi.org/10.1158/0008- [109] R. Moser, C. Xu, M. Kao, J. Annis, L.A. Lerma, C.M. Schaupp, et al., Functional
5472.Can-12-0487. kinomics identies candidate therapeutic targets in head and neck cancer,
[87] I. Ray-Coquard, J.Y. Blay, A. Italiano, A. Le Cesne, N. Penel, J. Zhi, et al., Effect of Clin. Cancer Res. 20 (16) (2014) 42744288, doi:http://dx.doi.org/10.1158/
the MDM2 antagonist RG7112 on the P53 pathway in patients with MDM2- 1078-0432.CCR-13-2858.
amplied, well-differentiated or dedifferentiated liposarcoma: an [110] B. Goldman, L. DeFrancesco, The cancer vaccine roller coaster, Nat. Biotechnol.
exploratory proof-of-mechanism study, Lancet Oncol. 13 (11) (2012) 1133 27 (2) (2009) 129139, doi:http://dx.doi.org/10.1038/nbt0209-129.
1140, doi:http://dx.doi.org/10.1016/S1470-2045(12)70474-6. [111] L. Chen, X. Han, Anti-PD-1/PD-L1 therapy of human cancer: past, present, and
[88] M. Andreeff, K.R. Kelly, K. Yee, S. Assouline, R. Strair, L. Popplewell, et al., future, J. Clin. Invest. 125 (9) (2015) 33843391, doi:http://dx.doi.org/10.1172/
Results of the phase I trial of RG7112, a small-Molecule MDM2 antagonist in JCI80011.
leukemia, Clin. Cancer Res. 22 (4) (2016) 868876, doi:http://dx.doi.org/ [112] K.C. Ohaegbulam, A. Assal, E. Lazar-Molnar, Y. Yao, X. Zang, Human cancer
10.1158/1078-0432.CCR-15-0481. immunotherapy with antibodies to the PD-1 and PD-L1 pathway, Trends Mol.
[89] Q. Ding, Z. Zhang, J.J. Liu, N. Jiang, J. Zhang, T.M. Ross, et al., Discovery of Med. 21 (1) (2015) 2433, doi:http://dx.doi.org/10.1016/j.
RG7388, a potent and selective p53-MDM2 inhibitor in clinical development, molmed.2014.10.009.
J. Med. Chem. (2013), doi:http://dx.doi.org/10.1021/jm400487c. [113] H.W. Nijman, A. Lambeck, S.H. van der Burg, A.G. van der Zee, T. Daemen,
[90] G. Martinelli, S. Assouline, M. Kasner, N. Vey, K.R. Kelly, M.W. Drummond, Immunologic aspect of ovarian cancer and p53 as tumor antigen, J. Transl.
et al., Phase 1b study of the MDM2 antagonist RG7112 In combination with 2 Med. 3 (2005) 34, doi:http://dx.doi.org/10.1186/1479-5876-3-34.
Doses/Schedules of cytarabine, Blood 122 (21) (2013) 498498. [114] K. Angelopoulou, E.P. Diamandis, D.J. Sutherland, J.A. Kellen, P.S. Bunting,
[91] A. Obrador-Hevia, E. Martinez-Font, I. Felipe-Abrio, S. Calabuig-Farinas, M. Prevalence of serum antibodies against the p53 tumor suppressor gene
Serra-Sitjar, J.A. Lopez-Guerrero, et al., RG7112, a small-molecule inhibitor of protein in various cancers, Int. J. Cancer 58 (4) (1994) 480487.
MDM2, enhances trabectedin response in soft tissue sarcomas, Cancer Invest. [115] O. Met, E. Balslev, H. Flyger, I.M. Svane, High immunogenic potential of
33 (9) (2015) 440450, doi:http://dx.doi.org/10.3109/ p53 mRNA-transfected dendritic cells in patients with primary breast cancer,
07357907.2015.1064534. Breast Cancer Res. Treat. 125 (2) (2011) 395406, doi:http://dx.doi.org/
[92] Y. Zhao, L. Liu, W. Sun, J. Lu, D. McEachern, X. Li, et al., Diastereomeric 10.1007/s10549-010-0844-9.
spirooxindoles as highly potent and efcacious MDM2 inhibitors, J. Am. [116] O.E. Rahma, E. Ashtar, M. Czystowska, M.E. Szajnik, E. Wieckowski, S.
Chem. Soc. 135 (19) (2013) 72237234, doi:http://dx.doi.org/10.1021/ Bernstein, et al., A gynecologic oncology group phase II trial of two
ja3125417. p53 peptide vaccine approaches: subcutaneous injection and intravenous
[93] K. Ding, Y. Lu, Z. Nikolovska-Coleska, G. Wang, S. Qiu, S. Shangary, et al., pulsed dendritic cells in high recurrence risk ovarian cancer patients, Cancer
Structure-based design of spiro-oxindoles as potent, specic small-molecule Immunology Immunotherapy 61 (3) (2012) 373384, doi:http://dx.doi.org/
inhibitors of the MDM2-p53 interaction, J. Med. Chem. 49 (12) (2006) 3432 10.1007/s00262-011-1100-9.
3435, doi:http://dx.doi.org/10.1021/jm051122a. [117] P.J. Schuler, M. Harasymczuk, C. Visus, A. Deleo, S. Trivedi, Y. Lei, et al., Phase I
[94] S. Yu, D. Qin, S. Shangary, J. Chen, G. Wang, K. Ding, et al., Potent and orally dendritic cell p53 peptide vaccine for head and neck cancer, Clin. Cancer Res.
active small-molecule inhibitors of the MDM2-p53 interaction, J. Med. Chem. 20 (9) (2014) 24332444, doi:http://dx.doi.org/10.1158/1078-0432.CCR-13-
52 (24) (2009) 79707973, doi:http://dx.doi.org/10.1021/jm901400z. 2617.
[95] S. Wang, W. Sun, Y. Zhao, D. McEachern, I. Meaux, C. Barriere, et al., [118] J.A. Roth, D. Nguyen, D.D. Lawrence, B.L. Kemp, C.H. Carrasco, D.Z. Ferson,
SAR405838: an optimized inhibitor of MDM2-p53 interaction that induces et al., Retrovirus-mediated wild-type p53 gene transfer to tumors of patients
complete and durable tumor regression, Cancer Res. 74 (20) (2014) 5855 with lung cancer, Nat. Med. 2 (9) (1996) 985991.
5865, doi:http://dx.doi.org/10.1158/0008-5472.can-14-0799. [119] J.M. Bishop, Retroviruses and cancer genes, Adv. Cancer Res. 37 (1982) 132.
[96] K.L. Bill, J. Garnett, I. Meaux, X. Ma, C.J. Creighton, S. Bolshakov, et al., [120] Z.X. Yang, D. Wang, G. Wang, Q.H. Zhang, J.M. Liu, P. Peng, X.H. Liu, Clinical
SAR405838: a novel and potent inhibitor of the MDM2:p53 axis for the study of recombinant adenovirus-p53 combined with fractionated
treatment of dedifferentiated liposarcoma, Clin. Cancer Res. 22 (5) (2016) stereotactic radiotherapy for hepatocellular carcinoma, J. Cancer Res. Clin.
11501160, doi:http://dx.doi.org/10.1158/1078-0432.ccr-15-1522. Oncol. 136 (4) (2010) 625630, doi:http://dx.doi.org/10.1007/s00432-009-
[97] J. Jung, J.S. Lee, M.A. Dickson, G.K. Schwartz, A. Le Cesne, A. Varga, et al., 0701-6.
TP53 mutations emerge with HDM2 inhibitor SAR405838 treatment in de- [121] M. Schuler, R. Herrmann, J.L. De Greve, A.K. Stewart, U. Gatzemeier, D.J.
differentiated liposarcoma, Nat. Commun. 7 (2016) 12609, doi:http://dx.doi. Stewart, et al., Adenovirus-mediated wild-type p53 gene transfer in patients
org/10.1038/ncomms12609. receiving chemotherapy for advanced non-small-cell lung cancer: results of a
[98] J. Canon, T. Osgood, S.H. Olson, A.Y. Saiki, R. Robertson, D. Yu, et al., The multicenter phase II study, J. Clin. Oncol. 19 (6) (2001) 17501758.
MDM2 inhibitor AMG 232 demonstrates robust antitumor efcacy and [122] H. Shimada, H. Matsubara, T. Shiratori, T. Shimizu, S. Miyazaki, S. Okazumi,
potentiates the activity of p53-Inducing cytotoxic agents, Mol. Cancer Ther. et al., Phase I/II adenoviral p53 gene therapy for chemoradiation resistant
14 (3) (2015) 649658, doi:http://dx.doi.org/10.1158/1535-7163.mct-14- advanced esophageal squamous cell carcinoma, Cancer Sci. 97 (6) (2006)
0710. 554561, doi:http://dx.doi.org/10.1111/j.1349-7006.2006.00206.x.
O. Merkel et al. / Mutation Research 773 (2017) 113 13

[123] A.Z. Bluming, J.L. Ziegler, Regression of Burkitt's lymphoma in association [136] H. Hermeking, The miR-34 family in cancer and apoptosis, Cell Death Differ.
with measles infection, Lancet 2 (7715) (1971) 105106. 17 (2) (2010) 193199, doi:http://dx.doi.org/10.1038/cdd.2009.56.
[124] R.M. Hansen, J.A. Libnoch, Remission of chronic lymphocytic leukemia after [137] D. Nalls, S.N. Tang, M. Rodova, R.K. Srivastava, S. Shankar, Targeting epigenetic
smallpox vaccination, Arch. Intern. Med. 138 (7) (1978) 11371138. regulation of miR-34a for treatment of pancreatic cancer by inhibition of
[125] J.R. Bischoff, D.H. Kirn, A. Williams, C. Heise, S. Horn, M. Muna, et al., An pancreatic cancer stem cells, PLoS One 6 (8) (2011) e24099, doi:http://dx.doi.
adenovirus mutant that replicates selectively in p53-decient human tumor org/10.1371/journal.pone.0024099.
cells, Science 274 (5286) (1996) 373376. [138] M. Beg Shaalan, First-in-human trial of microRNA cancer therapy with
[126] E.A. Chiocca, K.M. Abbed, S. Tatter, D.N. Louis, F.H. Hochberg, F. Barker, et al., A MRX34, a liposomal miR-34 mimic: phase Ia expansion in patients with
phase I open-label, dose-escalation, multi-institutional trial of injection with advanced solid tumors, J. Clin. Oncol. (2016) 34.
an E1B-Attenuated adenovirus, ONYX-015, into the peritumoral region of [139] K. Allton, A.K. Jain, H.M. Herz, W.W. Tsai, S.Y. Jung, J. Qin, et al., Trim24 targets
recurrent malignant gliomas, in the adjuvant setting, Mol. Ther. 10 (5) (2004) endogenous p53 for degradation, Proc. Natl. Acad. Sci. U. S. A. 106 (28) (2009)
958966, doi:http://dx.doi.org/10.1016/j.ymthe.2004.07.021. 1161211616, doi:http://dx.doi.org/10.1073/pnas.0813177106.
[127] E. Galanis, S.H. Okuno, A.G. Nascimento, B.D. Lewis, R.A. Lee, A.M. Oliveira, [140] D. Dornan, I. Wertz, H. Shimizu, D. Arnott, G.D. Frantz, P. Dowd, et al., The
et al., Phase I-II trial of ONYX-015 in combination with MAP chemotherapy in ubiquitin ligase COP1 is a critical negative regulator of p53, Nature 429
patients with advanced sarcomas, Gene Ther. 12 (5) (2005) 437445, doi: (6987) (2004) 8692, doi:http://dx.doi.org/10.1038/nature02514.
http://dx.doi.org/10.1038/sj.gt.3302436. [141] R.P. Leng, Y. Lin, W. Ma, H. Wu, B. Lemmers, S. Chung, et al., Pirh2, a p53-
[128] D. Makower, A. Rozenblit, H. Kaufman, M. Edelman, M.E. Lane, J. Zwiebel, induced ubiquitin-protein ligase, promotes p53 degradation, Cell 112 (6)
et al., Phase II clinical trial of intralesional administration of the oncolytic (2003) 779791.
adenovirus ONYX-015 in patients with hepatobiliary tumors with correlative [142] W. Hu, Z. Feng, I. Modica, D.S. Klimstra, L. Song, P.J. Allen, et al., Gene
p53 studies, Clin. Cancer Res. 9 (2) (2003) 693702. amplications in well-Differentiated pancreatic neuroendocrine tumors
[129] W. Yu, H. Fang, Clinical trials with oncolytic adenovirus in China, Curr. Cancer inactivate the p53 pathway, Genes Cancer 1 (4) (2010) 360368, doi:http://
Drug Targets 7 (2) (2007) 141148. dx.doi.org/10.1177/1947601910371979.
[130] M. Farnebo, V.J. Bykov, K.G. Wiman, The p53 tumor suppressor: a master [143] Y. Jiao, C. Shi, B.H. Edil, R.F. de Wilde, D.S. Klimstra, A. Maitra, et al., DAXX/
regulator of diverse cellular processes and therapeutic target in cancer, ATRX, MEN1, and mTOR pathway genes are frequently altered in pancreatic
Biochem. Biophys. Res. Commun. 396 (1) (2010) 8589, doi:http://dx.doi.org/ neuroendocrine tumors, Science 331 (6021) (2011) 11991203, doi:http://dx.
10.1016/j.bbrc.2010.02.152. doi.org/10.1126/science.1200609.
[131] Y.S. Guan, Q. He, Q. Zou, Status quo of p53 in the treatment of tumors, [144] L. Vannucci, M. Lai, F. Chiuppesi, L. Ceccherini-Nelli, M. Pistello, Viral vectors:
Anticancer Drugs (2016), doi:http://dx.doi.org/10.1097/ a look back and ahead on gene transfer technology, New Microbiol. 36 (1)
CAD.0000000000000397. (2013) 122.
[132] D.P. Bartel, MicroRNAs: target recognition and regulatory functions, Cell 136 [145] S. Hacein-Bey-Abina, C. Von Kalle, M. Schmidt, M.P. McCormack, N. Wulffraat,
(2) (2009) 215233, doi:http://dx.doi.org/10.1016/j.cell.2009.01.002. P. Leboulch, et al., LMO2-associated clonal T cell proliferation in two patients
[133] G. Deng, G. Sui, Noncoding RNA in oncogenesis: a new era of identifying key after gene therapy for SCID-X1, Science 302 (5644) (2003) 415419, doi:
players, Int. J. Mol. Sci. 14 (9) (2013) 1831918349, doi:http://dx.doi.org/ http://dx.doi.org/10.1126/science.1088547.
10.3390/ijms140918319. [146] I. Gil-Farina, M. Schmidt, Interaction of vectors and parental viruses with the
[134] F. Pichiorri, S.S. Suh, A. Rocci, L. De Luca, C. Taccioli, R. Santhanam, et al., host genome, Curr Opin Virol 21 (2016) 3540, doi:http://dx.doi.org/10.1016/
Downregulation of p53-inducible microRNAs 192, 194, and 215 impairs the j.coviro.2016.07.004.
p53/MDM2 autoregulatory loop in multiple myeloma development, Cancer [147] L. Cong, F.A. Ran, D. Cox, S. Lin, R. Barretto, N. Habib, et al., Multiplex genome
Cell 18 (4) (2010) 367381, doi:http://dx.doi.org/10.1016/j.ccr.2010.09.005. engineering using CRISPR/Cas systems, Science 339 (6121) (2013) 819823,
[135] S. Hunten, H. Siemens, M. Kaller, H. Hermeking, The p53/microRNA network doi:http://dx.doi.org/10.1126/science.1231143.
in cancer: experimental and bioinformatics approaches, Adv. Exp. Med. Biol.
774 (2013) 77101, doi:http://dx.doi.org/10.1007/978-94-007-5590-1_5.

Вам также может понравиться