Next Article in Journal
Study on the Preparation of Ionic Liquid Doped Chitosan/Cellulose-Based Electroactive Composites
Next Article in Special Issue
Do Mutations Turn p53 into an Oncogene?
Previous Article in Journal
Health Functions and Related Molecular Mechanisms of Tea Components: An Update Review
Previous Article in Special Issue
The Diverse Functions of Mutant 53, Its Family Members and Isoforms in Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Gain-of-Function Mutant p53: All the Roads Lead to Tumorigenesis

1
Department of Molecular Cell Biology, Weizmann Institute of Science, Rehovot 7610001, Israel
2
Department of Biochemistry and Molecular Genetics, Israel Institute for Biological Research, Ness-Ziona 7410001, Israel
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(24), 6197; https://doi.org/10.3390/ijms20246197
Submission received: 30 October 2019 / Revised: 25 November 2019 / Accepted: 5 December 2019 / Published: 8 December 2019
(This article belongs to the Special Issue p53 in Cancer and beyond—40 Years after Its Discovery)

Abstract

:
The p53 protein is mutated in about 50% of human cancers. Aside from losing the tumor-suppressive functions of the wild-type form, mutant p53 proteins often acquire inherent, novel oncogenic functions, a phenomenon termed mutant p53 gain-of-function (GOF). A growing body of evidence suggests that these pro-oncogenic functions of mutant p53 proteins are mediated by affecting the transcription of various genes, as well as by protein–protein interactions with transcription factors and other effectors. In the current review, we discuss the various GOF effects of mutant p53, and how it may serve as a central node in a network of genes and proteins, which, altogether, promote the tumorigenic process. Finally, we discuss mechanisms by which “Mother Nature” tries to abrogate the pro-oncogenic functions of mutant p53. Thus, we suggest that targeting mutant p53, via its reactivation to the wild-type form, may serve as a promising therapeutic strategy for many cancers that harbor mutant p53. Not only will this strategy abrogate mutant p53 GOF, but it will also restore WT p53 tumor-suppressive functions.

1. Introduction

The tumor suppressor p53 functions as the main regulator of several major signaling and cell-fate-decision pathways. Subsequent to various stresses such as DNA damage, oncogene activation or others, p53 undergoes post-translational modifications that lead to its activation, stabilization, and accumulation in the cell. The tumor-suppressor activity of p53 is mainly attributed to its transcriptional regulation of genes that are involved in numerous cellular processes, such as cell cycle arrest, apoptosis, senescence, DNA repair, and differentiation [1,2]. Moreover, accumulating data points to p53 involvement in numerous biological processes, such as forming a barrier to stem cell formation [3,4], metabolism [5,6], regeneration [7], prevention of liver pathologies [8,9], interaction with viruses [10], endocrinology circuits [11] and serving as the guardian of tissue hierarchy [12].
In nearly 50% of human cancers, p53 is found to be mutated, and in many of the remaining cases, where the WT allele is retained, other components in the p53 pathway are defected [13]. Mutations in most of the other tumor suppressors cause the loss of protein production, or the production of truncated or unstable proteins. In contrast, p53 is known to acquire missense mutations that lead to the production of a full-length protein, which tends to accumulate to high levels in tumor cells [14,15]. The most common missense mutations occur in six “hot-spot” amino-acids. Most of these mutations are found in the DNA-binding domain (DBD) of p53 and abrogate its transcriptional activity [16]. Generally, mutations in p53 are divided into two groups [17]. The first group consists of DNA contact mutations, such as p53R248Q and p53R273H, which affect the domains of the protein that are directly involved in DNA binding. The second group consists of conformational mutations, such as p53R175H and p53H179R, which cause either a full or partial distortion of the correct folding of the DBD of the p53 protein.
A unique feature of mutant p53 is its gain-of-function (GOF), which endows it with inherent oncogenic functions that are independent of the loss of WT p53 tumor-suppressor activity [18,19]. One of the first demonstrations of mutant p53 GOF was observed in TP53-null cells that were transfected with the over-expressing vector of mutant p53. Upon injections administered to mice, TP53-null cells formed local tumors that later regressed, while mutant p53 over-expressing cells formed lethal tumors [20]. Furthermore, mutant p53 knock-in mice show novel types of tumors and a significantly higher rate of metastases compared to p53−/− mice, thus demonstrating p53 GOF in vivo [21,22]. Various mechanisms were found to account for this GOF, including increased proliferation, inhibition of apoptosis, resistance to chemotherapy and enhanced inflammation, angiogenesis and invasiveness [18]. In this review, we describe some of the main roads by which GOF mutant p53 leads a cell towards tumorigenesis, and highlighted the broad interactions that mutant p53 has with numerous proteins, as well as its numerous transcriptional targets.

2. Disruption of Cell Cycle Control

WT p53 was shown to be involved in cell cycle regulation, promoting cell cycle arrest or cell death [23]. Thus, it is not surprising that mutant p53 was found to disrupt cell cycle control, leading to enhanced proliferation. Indeed, the notion that GOF p53 accelerates cell proliferation is well-established [24,25]. In studies aimed to understand the mechanism leading to accelerated proliferation, it was shown that tumor-derived p53 mutants interact physically with the master cell cycle regulator nuclear factor Y (NF-Y). These protein complexes can increase DNA synthesis in response to DNA damage through an aberrant upregulation of the expression of the NF-Y cell-cycle target genes, such as cyclin/CDK1 kinase complexes [26]. It is worth mentioning that these genes are clustered with other cell cycle control genes and are annotated as a “proliferation cluster” [27]. In a following study, it was found that mutant p53 interacts with Yes-associated protein (YAP), and together they form a complex with NF-Y, that interacts with the regulatory regions of cyclin A, cyclin B, and CDK1 genes [28]. This was further established in a genome-wide analysis showing that GOF p53 recognizes the promoters of genes encoding cyclin A (CCNA2), which is necessary for origin firing, and CHK1, which is required for preventing a collapse of replication forks, and transcriptionally activates their expression, in a cell-cycle-dependent manner, by localizing on their upstream regulatory sequences [29].
Mutant p53 was also found to trigger the activation of non-coding effectors, such as the circular RNA circPVT1 and miR-497-5p, leading to uncontrolled proliferation through the abnormal enhancement of the expression of cell-cycle-regulated genes. This is regulated through the mutant p53/YAP/TEAD complex via its regulatory region [30]. On the other hand, mutant p53 was shown to suppress the expression of miR-27a, resulting in augmented cell proliferation due to enhanced epidermal growth factor receptor (EGFR) signaling, resulting in the activation of the extracellular signal-regulated kinase (ERK) pathway [31]. In addition, various mutant p53 forms were shown to bind and activate STAT3, leading to increased invasion and tumor growth in colorectal cancer [32]. Aside from affecting signaling pathways, mutant p53 was shown to regulate different chromatin regulators, including the methyltransferases MLL1 and MLL2, and the acetyltransferase MOZ. This regulation was shown to globally affect histone modifications, and to promote the proliferation of cancer cells [33]. Thus, it may be concluded that mutant p53 does not affect only cell signaling and specific gene transcription, but may also underlie global chromatin changes that occur in cancer cells, which facilitate their malignant phenotype.

3. Genomic Instability

Cells harboring mutant p53 exhibit extensive chromosomal aberrations. Mutant p53-expressing pre-tumor thymocytes, but not p53−/−cells, were shown to possess interchromosomal translocations. Moreover, after DNA damage, the G2–M checkpoint was found to be impaired in these p53-mutant cells. The mechanism behind this observation was attributed to the interactions of the p53 mutants with the nuclease Mre11, leading to less binding of the Mre11–Rad50–NBS1 (MRN) complex to DNA double-stranded breaks (DSBs), accounting for impaired Ataxia-telangiectasia mutated (ATM) activation [34]. Notably, genomic instability was also observed in mutant p53 transgenic mice [35,36,37].
Recently, a mechanistic link between mutant p53 expression and a cellular engulfment process, which leads to cell-in-cell (CIC) structures, was described [38]. Cell-in-cell (CIC) structures are mainly seen in aggressive tumors and are related to chromosomal structural abnormalities in human tumor tissue. The study concluded that cell ingestion in the presence of mutant p53 may drive entotic engulfment that would cause abnormal cellular division, and that mutant p53 would allow the abnormal progeny to survive. Interestingly, phosphorylated CHK1 (p-CHK1) was detected in the DNA of the engulfing cells. As mentioned above, CHK1 is upregulated by mutant p53 and prevents the collapse of cells with origins of replication forks. A large-scale genomic study in a set of tumors led to the observation that there is an association between CIC and somatic copy number aberrations. Support for the observation that mutant p53 expression and CIC are linked to genomic instability was provided by the report that chromosomal instability was detected in the Pdx1-Cre; LSL-KrasG12D/+; LSL-Trp53R172H/+ mice, but not in their p53 null counterparts. Moreover, tripolar mitoses in CIC structures were often recorded in the mutant p53 mice and not in the p53 null mice [38].

4. Promoting Dedifferentiation and Acquiring Stemness Properties

The role of WT p53 in differentiation is well documented [39,40,41,42,43]. Moreover, WT p53 has been implicated in the regulation of stem cells self–renewal and differentiation and as a barrier to cancer stem cell (CSC) formation [3]. Thus, it was conceivable to foresee that GOF mutant p53 will contribute to the dedifferentiation processes and the formation of CSCs. Indeed, tumors harboring mutant p53 seem to possess an undifferentiated and aggressive phenotype. p53 mutations were shown to be restricted to poorly differentiated thyroid tumors and thyroid cancer cell lines [44]. Interestingly, examining different areas within a tumor revealed that the mutant p53 was confined to the less differentiated part of the tumor, presenting a positive correlation of mutant p53 and dedifferentiation [44] (for more examples of p53 aberration and dedifferentiation phenotypes, see Table 1 in [3]). Moreover, in breast and lung tumors, stem-cell-like patterns of transcription were found to coincide with the abolishment of WT p53 and the presence of p53 mutations [45]. These data are in line with the observation that WT p53 restrained the reprogramming process of mouse embryonic fibroblasts (MEFs) into induced pluripotent stem cells (iPSCs) [46,47,48,49,50], while GOF mutant p53 enhanced the reprogramming efficacy along with augmented tumorigenicity of the reprogrammed cells [51]. Moreover, accumulation of mutant p53 in progenitor-like cells in the subventricular zone-associated areas in the brain was shown to lead to the initiation of glioma, suggesting improper maturation of neural stem cells due to the presence of mutant p53 [52]. Similarly, mutant p53 bone-marrow mesenchymal stem cells are tumorigenic and induce sarcomas [53]. Interestingly, tumor cell lines derived from these sarcomas exhibited enhanced tumorigenic potential. These cells were smaller in size, had a circular morphology and high proliferation rate [54], characteristics that are assigned to iPSCs and stem cells [55,56]. Moreover, these cells highly expressed cancer- and CSC-related genes and an ESC gene signature. This signature correlates with the one seen in human tumors harboring p53 missense mutation, and is associated with poor patient survival. Furthermore, mutant p53 knockout led to a reduction in the tumor initiation capacity of these tumor-derived cell lines and in the expression of the ESC gene signature, supporting a link between mutant p53, stemness, and tumorigenesis. It is intriguing to speculate that specific genes identified in this ESC signature, can be viewed as a broader signature of CSCs and tumor progression toward a progressive disease, and may be used as biomarkers for disease progression. Recently published, mutant p53 was shown to promote aberrant self-renewal in acute myeloid leukemia (AML). This gain-of-function activity of mutant p53 is mediated by FOXH1, which can bind and regulate the expression of stem cell genes [57].

5. Modulation of Metabolism

It is well-established that cancer cells rewire a wide variety of metabolic processes to fulfill their energetic demands, as well as their anabolic needs, which require an excess of building blocks [58]. The roles of WT p53 in the regulation of metabolism are well-documented, including suppression of the Warburg effect [5,59,60], as well as controlling the response for glucose [61], lipid [6] and amino-acid starvation [62,63]. Mutant p53 was shown to modulate cancer cell metabolism in several ways. First, mutant p53 was shown to facilitate the cancer cell’s stress response to starvation due to the depletion of metabolites. Several mechanisms were shown by which mutant p53 ameliorates cell starvation, such as inhibition of the starvation response regulator AMP kinase (AMPK) [64], and amelioration of glutamine starvation response, primarily by activating canonical WT p53 targets—p21 in particular [65]. Aside from attenuating starvation response, mutant p53 is also implicated in the accumulation of metabolites, which are essential for tumor cells’ oncogenic properties. For example, in high-grade serous ovarian cancer (HGSOC), mutant p53 was shown to regulate the amount of lipids such as lysophosphatidic acid (LPA). This study shows that mutant p53 leads to LPA accumulation by down-regulating the LPA degrading enzyme, ACP6, in a manner that was sufficient to support HGSOC [66].
The second way by which mutant p53 may modulate cancer cell metabolism is by inducing the mevalonate pathway [67,68], which is implicated in a wide variety of physiological and pathological conditions, including cancer [69]. The targeting of the mevalonate pathway in cancer has been examined in different pre-clinical and clinical studies [69,70]. Intriguingly, it was demonstrated that mutant p53 and the mevalonate pathway exhibit reciprocal interplay, in which mutant p53 upregulates genes pertaining to the mevalonate pathway [67], while the mevalonate pathway was found to be a critical modulator of mutant p53 protein stability, but not the stability of the WT p53 protein [68,71,72,73]. Furthermore, statins, which inhibit a key enzyme in the mevalonate pathway, were shown to attenuate mutant p53 GOF effects, both by downregulating mevalonate pathway-related genes [67], as well as by reducing mutant p53 protein stability [68,71,72,73]. In all, these data indicate that not only it is possible to target mutant p53 harboring tumors by inhibiting metabolic changes induced by mutant p53, but it is also possible to target mutant p53 itself by using agents targeting metabolic pathways, such as statins.
A notable protein that was shown to mediate various metabolic effects of mutant p53 is NRF2. One of the metabolic effects that was shown to be mediated by NRF2 is the regulation of the response to oxidative stress. Although excessive levels of reactive oxygen species (ROS) are toxic to cells, moderately elevated ROS levels are beneficial for cancer cells, as they may activate a variety of signaling pathways that promote cell proliferation and survival [74]. Indeed, different studies indicate that tumor cells exhibit elevated levels of ROS and oxidative stress, which also correlate with tumor cells aggressiveness, as well as with poor prognosis in cancer patients [74,75,76,77,78]. In a previous study that was conducted in our lab, we reported that mutant p53 causes the increase of ROS by diminishing the expression of ROS detoxifying enzymes, via decreasing the induction of NRF2 [79]. A later study corroborated this observation, and demonstrated that the accumulation of ROS in mutant p53 cancer cells can be exploited as a therapeutic strategy, by pharmacologic depletion of glutathione (GSH), which specifically targets mutant-p53-expressing cells [80]. Protein homeostasis is another mutant p53 phenotype that was shown to be mediated by NRF2, through the regulation of proteasome activity [81]. The mutant p53–NRF2 interaction was shown to upregulate the expression of various proteasome subunit genes that lead to increased proteasome activity, which, in turn, lead to inhibition of tumor suppressors, such as KSRP [81]. Aside from NRF2, mutant p53 was shown to increase proteasome activity via upregulating the proteasome activator REGγ [82].
Another noteworthy metabolic effect mediated by mutant p53 is the enhancement of the Warburg effect by various mechanisms, including increased glucose intake via the GLUT1 transporter [83], as well as increasing glycolytic activity and decreasing mitochondrial oxidative phosphorylation [84]. Intriguingly, it was demonstrated that the R72 polymorphic form of mutant p53 endows cancer cells with enhanced oxidative-phosphorylation activity via the regulation of PGC-1α, thus leading to increased invasiveness and metastatic potential [85]. In that regard, it is interesting to note that in our model of highly aggressive, mutant p53 expressing CSC-like tumor lines [54], we observed several metabolic effects. These effects were, at least partially, mutant-p53-dependent, including increased glucose uptake, which correlates with increased GLUT1 membranal translocation, as well as increased mitochondrial mass and oxidative metabolism, as compared to the parental cells from which the tumor lines were derived [86]. The latter study, as well as the study linking mutant p53 forms with increased mitochondrial activity via PGC-1α, are in line with studies that reported increased expression of PGC-1α, as well as mitochondrial activity, in CSCs [87,88]. Taking the aforementioned findings together, it is tempting to speculate that mutant p53 might regulate metabolic changes depending on the cellular context, by suppressing mitochondrial metabolism in non-CSCs to promote the Warburg effect, and conversely, increasing mitochondrial metabolism in CSCs, to enhance their stemness.

6. Tumor Microenvironment

The microenvironment may help to block tumor formation, even when adjacent cells harbor potentially oncogenic genetic aberrations [89]. However, upon progression of the tumorigenic process, the microenvironment may change its phenotype and actively promote tumorigenesis [89]. We have recently published a review [90] summarizing the tumor–stroma crosstalk and the role of mutant p53 in this interaction, and proposed that the microenvironment may be modulated by mutant p53 to promote tumorigenesis. We suggest that mutant p53 may function as a key player in the modulation of the tumor–stroma crosstalk both at the initiation as well as the progression of the tumor–stroma vicious cycle, during which mutant p53 facilitates the adaptation (“re-education”) of the stroma cells to tumorigenic-promoting cells in the stromal compartment, as well as causing a modulation of the signals arriving from the stromal cells in a way that benefits the tumor cells.

7. Mother Nature Knows Best—Retrieving WT Function

The presence of a mutant p53 in various stem cells such as embryonic stem cells (ESCs), iPSCs or adult stem cells may lead to a lethal outcome. Mother Nature evolved various elegant mechanisms to detain these fatal consequences. Indeed, despite malignant properties of mutant p53 in somatic cells, mutant p53 mice develop into viable embryos and mature organisms, and develop aggressive tumors only in adulthood. This phenomenon is intriguing, especially in light of the extreme sensitivity of ESCs to DNA damage and the need for an active WT p53 to protect these cells, suggesting a mechanism to restrain mutant p53 function. A comprehensive study [91] has revealed that WT and mutant p53 ESCs exhibit similar characteristics in doubling time and in vitro differentiation potential, and that despite the difference in p53 status, both ESCs remained pluripotent in vivo and gave rise to benign teratomas upon injection into nude mice. Moreover, mutant p53 ESCs displayed only minor karyotype changes, pointing to an active tumor-suppressive function within these mutant ESCs. Strikingly, analysis of the p53 protein conformation within the mutant ESCs revealed a WT conformation in a ratio of approximately 1:1 with the mutant conformation. This conformational change allowed the activation of DNA damage-responsive genes in these mutant ESCs, in a similar pattern as the one obtained in WT p53 ESCs, providing genomic stability to the mutant ESCs and the birth of mice with no malignant signature.
Another mechanism to maintain at least one WT allele is the loss of heterozygosity (LOH). We have shown that the classical well-known WT p53 LOH is correlated with genome stability of the cells: higher frequencies of LOH were observed in somatic cells compared to adult stem cells, in an age-dependent manner, while LOH was greatly attenuated in iPSCs and even more in ESCs [53]. Interestingly, it was found in single-cell-subclones of iPSCs and MSCs that LOH is a bi-directional process, namely, not just the WT allele may be lost, but events of loss of the mutant allele were also recorded. Strikingly, the majority of the LOH events in ex vivo bone marrow progenitors led to the loss of the mutant allele, suggesting a mechanism that dictates a preference to retain the WT allele and to exclude the mutant allele, assuring genome stability in vivo.
Turrell et al. [92] presented data of yet another mechanism by which the cell may protect itself against the tumorigenic effect of mutant p53. Their data argue against a dominant-negative effect of mutant p53 on the WT form in lung tumors, but rather suggest a dominant-positive function of the WT p53 on the mutant form. Even in the presence of mutant p53, WT p53 retained its transcriptional activity towards anti-proliferative and apoptotic responses [92]. This observation can be explained either by the fact that p53 mutant:WT heterotetramers may retain, in these cells, some WT functions, or that there is a bias towards WT p53 homotetramer formation. Indeed, a subsequent study [93] revealed that WT p53 oligomerization is more efficient than the oligomerization of mutant p53. At equimolar levels of WT p53 and mutant p53, mutant p53 GOF activities are restrained. This may explain the need for the accumulation of mutant p53 in tumors. High levels of mutant p53 contribute not just to the dominant-negative functions of mutant p53, but also to the abolishment of dominant-positive functions of the WT form. Finally, yet importantly, the mechanism by which Mother Nature protects large animals, such as elephants, from cancer, is worth mentioning. It is believed that elephants have a lower-than-expected rate of cancer, when compared to other mammals, due to their multiple copies of TP53, which expanded, coincidently with the evolution of large body size. As a result, elephant cells demonstrate increased apoptotic response following DNA damage, compared to human cells [94,95].

8. Concluding Remarks

WT p53 functions as a major regulator in many cell-fate-determining processes. It executes its guardian missions through various interactions with other proteins and by transcriptional regulation of different genes. Interestingly, mutations in the WT form not only lead to the abrogation of its normal functions, but also give rise to an active antithetical protein, with its own “social network” of interacting proteins and transcriptional targets that endows it with GOF activities (Table 1). Since the formation of these mutants of p53 is fatal to cells, Mother Nature has evolved mechanisms such as LOH and conformational “correction” of the mutant conformation to secure the cells. This “re-education” of mutant p53, namely, the various mechanisms by which Mother Nature attempts to ameliorate the effects of mutant p53, not just abrogates the GOF of the mutant form, but also reinstates WT p53 tumor-suppressive functions. Based on the wisdom of Mother Nature, the development of new cancer medicines, aimed at the correction of the mutant p53 form [96,97], may provide new horizons in cancer treatment.

Funding

Research in the laboratory of Varda Rotter is supported by a Center of Excellence Grant from the Israel Science Foundation (ISF) and a Center of Excellence Grant from the Flight Attendant Medical Research Institute (FAMRI).

Acknowledgments

Varda Rotter is the incumbent of the Norman and Helen Asher Professorial Chair for Cancer Research at the Weizmann Institute.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AMLAcute myeloid leukemia
CICCell-in-cell
CSCCancer stem cell
DBDDNA binding domain
DSBDouble-strand breaks
ESCEmbryonic stem cell
GOFGain-of-function
HGSOCHigh-grade serous ovarian cancer
iPSCInduced pluripotent stem cell
LOHLoss of heterozygosity
LPAlysophosphatidic acid
LSLLox-Stop-Lox
MEFMouse embryonic fibroblast
MSCMesenchymal stem cell
NMGNucleotide metabolism gene
ROSReactive oxygen species
WTWild-type

References

  1. Molchadsky, A.; Rivlin, N.; Brosh, R.; Rotter, V.; Sarig, R. p53 is balancing development, differentiation and de-differentiation to assure cancer prevention. Carcinogenesis 2010, 31, 1501–1508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bieging, K.T.; Mello, S.S.; Attardi, L.D. Unravelling mechanisms of p53-mediated tumour suppression. Nat. Rev. Cancer 2014, 14, 359–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Aloni-Grinstein, R.; Shetzer, Y.; Kaufman, T.; Rotter, V. p53: The barrier to cancer stem cell formation. FEBS Lett. 2014, 588, 2580–2589. [Google Scholar] [CrossRef] [PubMed]
  4. Molchadsky, A.; Rotter, V. p53 and its mutants on the slippery road from stemness to carcinogenesis. Carcinogenesis 2017, 38, 347–358. [Google Scholar] [CrossRef] [Green Version]
  5. Labuschagne, C.F.; Zani, F.; Vousden, K.H. Control of metabolism by p53—Cancer and beyond. Biochim. Biophys. Acta Rev. Cancer 2018, 1870, 32–42. [Google Scholar] [CrossRef]
  6. Goldstein, I.; Rotter, V. Regulation of lipid metabolism by p53—Fighting two villains with one sword. Trends Endocrinol. Metab. 2012, 23, 567–575. [Google Scholar] [CrossRef]
  7. Charni, M.; Aloni-Grinstein, R.; Molchadsky, A.; Rotter, V. p53 on the crossroad between regeneration and cancer. Cell Death Differ. 2017, 24, 8–14. [Google Scholar] [CrossRef] [Green Version]
  8. Charni, M.; Rivlin, N.; Molchadsky, A.; Aloni-Grinstein, R.; Rotter, V. p53 in liver pathologies-taking the good with the bad. J. Mol. Med. (Berl.) 2014, 92, 1229–1234. [Google Scholar] [CrossRef]
  9. Yan, Z.; Miao, X.; Zhang, B.; Xie, J. p53 as a double-edged sword in the progression of non-alcoholic fatty liver disease. Life Sci. 2018, 215, 64–72. [Google Scholar] [CrossRef]
  10. Aloni-Grinstein, R.; Charni-Natan, M.; Solomon, H.; Rotter, V. p53 and the Viral Connection: Back into the Future. Cancers 2018, 10, 178. [Google Scholar] [CrossRef] [Green Version]
  11. Charni-Natan, M.; Aloni-Grinstein, R.; Osher, E.; Rotter, V. Liver and Steroid Hormones—Can a Touch of p53 Make a Difference? Front. Endocrinol. (Lausanne) 2019, 10, 374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Koifman, G.; Aloni-Grinstein, R.; Rotter, V. p53 balances between tissue hierarchy and anarchy. J. Mol. Cell Biol. 2019, 11, 553–563. [Google Scholar] [CrossRef] [PubMed]
  13. Aylon, Y.; Oren, M. New plays in the p53 theater. Curr. Opin. Genet. Dev. 2011, 21, 86–92. [Google Scholar] [CrossRef] [PubMed]
  14. Rotter, V. p53, a transformation-related cellular-encoded protein, can be used as a biochemical marker for the detection of primary mouse tumor cells. Proc. Natl. Acad. Sci. USA 1983, 80, 2613–2617. [Google Scholar] [CrossRef] [Green Version]
  15. Bártek, J.; Bártková, J.; Vojtĕsek, B.; Stasková, Z.; Lukás, J.; Rejthar, A.; Kovarík, J.; Midgley, C.A.; Gannon, J.V.; Lane, D.P. Aberrant expression of the p53 oncoprotein is a common feature of a wide spectrum of human malignancies. Oncogene 1991, 6, 1699–1703. [Google Scholar]
  16. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 mutations in human cancers: Origins, consequences, and clinical use. Cold Spring Harb. Perspect. Biol. 2010, 2, a001008. [Google Scholar] [CrossRef] [Green Version]
  17. Bullock, A.N.; Fersht, A.R. Rescuing the function of mutant p53. Nat. Rev. Cancer 2001, 1, 68–76. [Google Scholar] [CrossRef]
  18. Brosh, R.; Rotter, V. When mutants gain new powers: News from the mutant p53 field. Nat. Rev. Cancer 2009, 9, 701–713. [Google Scholar] [CrossRef]
  19. Muller, P.A.J.; Vousden, K.H. Mutant p53 in cancer: New functions and therapeutic opportunities. Cancer Cell 2014, 25, 304–317. [Google Scholar] [CrossRef] [Green Version]
  20. Wolf, D.; Harris, N.; Goldfinger, N.; Rotter, V. Isolation of a Full-Length Mouse cDNA Clone Coding for an Immunologically Distinct p53 Molecule. Mol. Cell. Biol. 1985, 5, 127–132. [Google Scholar] [CrossRef] [Green Version]
  21. Lang, G.A.; Iwakuma, T.; Suh, Y.-A.; Liu, G.; Rao, V.A.; Parant, J.M.; Valentin-Vega, Y.A.; Terzian, T.; Caldwell, L.C.; Strong, L.C.; et al. Gain of function of a p53 hot spot mutation in a mouse model of Li-Fraumeni syndrome. Cell 2004, 119, 861–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Olive, K.P.; Tuveson, D.A.; Ruhe, Z.C.; Yin, B.; Willis, N.A.; Bronson, R.T.; Crowley, D.; Jacks, T. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 2004, 119, 847–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Levine, A.J.; Oren, M. The first 30 years of p53: Growing ever more complex. Nat. Rev. Cancer 2009, 9, 749–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Oren, M.; Rotter, V. Mutant p53 gain-of-function in cancer. Cold Spring Harb. Perspect. Biol. 2010, 2, a001107. [Google Scholar] [CrossRef]
  25. Freed-Pastor, W.A.; Prives, C. Mutant p53: One name, many proteins. Genes Dev. 2012, 26, 1268–1286. [Google Scholar] [CrossRef] [Green Version]
  26. Di Agostino, S.; Strano, S.; Emiliozzi, V.; Zerbini, V.; Mottolese, M.; Sacchi, A.; Blandino, G.; Piaggio, G. Gain of function of mutant p53: The mutant p53/NF-Y protein complex reveals an aberrant transcriptional mechanism of cell cycle regulation. Cancer Cell 2006, 10, 191–202. [Google Scholar] [CrossRef] [Green Version]
  27. Brosh, R.; Rotter, V. Transcriptional control of the proliferation cluster by the tumor suppressor p53. Mol. Biosyst. 2010, 6, 17–29. [Google Scholar] [CrossRef]
  28. Di Agostino, S.; Sorrentino, G.; Ingallina, E.; Valenti, F.; Ferraiuolo, M.; Bicciato, S.; Piazza, S.; Strano, S.; Del Sal, G.; Blandino, G. YAP enhances the pro-proliferative transcriptional activity of mutant p53 proteins. EMBO Rep. 2016, 17, 188–201. [Google Scholar] [CrossRef]
  29. Singh, S.; Vaughan, C.A.; Frum, R.A.; Grossman, S.R.; Deb, S.; Palit Deb, S. Mutant p53 establishes targetable tumor dependency by promoting unscheduled replication. J. Clin. Investig. 2017, 127, 1839–1855. [Google Scholar] [CrossRef] [Green Version]
  30. Verduci, L.; Ferraiuolo, M.; Sacconi, A.; Ganci, F.; Vitale, J.; Colombo, T.; Paci, P.; Strano, S.; Macino, G.; Rajewsky, N.; et al. The oncogenic role of circPVT1 in head and neck squamous cell carcinoma is mediated through the mutant p53/YAP/TEAD transcription-competent complex. Genome Biol. 2017, 18, 237. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, W.; Cheng, B.; Miao, L.; Mei, Y.; Wu, M. Mutant p53-R273H gains new function in sustained activation of EGFR signaling via suppressing miR-27a expression. Cell Death Dis. 2013, 4, e574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Schulz-Heddergott, R.; Stark, N.; Edmunds, S.J.; Li, J.; Conradi, L.-C.; Bohnenberger, H.; Ceteci, F.; Greten, F.R.; Dobbelstein, M.; Moll, U.M. Therapeutic Ablation of Gain-of-Function Mutant p53 in Colorectal Cancer Inhibits Stat3-Mediated Tumor Growth and Invasion. Cancer Cell 2018, 34, 298–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhu, J.; Sammons, M.A.; Donahue, G.; Dou, Z.; Vedadi, M.; Getlik, M.; Barsyte-Lovejoy, D.; Al-awar, R.; Katona, B.W.; Shilatifard, A.; et al. Gain-of-function p53 mutants co-opt chromatin pathways to drive cancer growth. Nature 2015, 525, 206–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Song, H.; Hollstein, M.; Xu, Y. p53 gain-of-function cancer mutants induce genetic instability by inactivating ATM. Nat. Cell Biol. 2007, 9, 573–580. [Google Scholar] [CrossRef] [PubMed]
  35. Caulin, C.; Nguyen, T.; Lang, G.A.; Goepfert, T.M.; Brinkley, B.R.; Cai, W.-W.; Lozano, G.; Roop, D.R. An inducible mouse model for skin cancer reveals distinct roles for gain- and loss-of-function p53 mutations. J. Clin. Investig. 2007, 117, 1893–1901. [Google Scholar] [CrossRef] [PubMed]
  36. Hingorani, S.R.; Wang, L.; Multani, A.S.; Combs, C.; Deramaudt, T.B.; Hruban, R.H.; Rustgi, A.K.; Chang, S.; Tuveson, D.A. Trp53R172H and KrasG12D cooperate to promote chromosomal instability and widely metastatic pancreatic ductal adenocarcinoma in mice. Cancer Cell 2005, 7, 469–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Murphy, K.L.; Dennis, A.P.; Rosen, J.M. A gain of function p53 mutant promotes both genomic instability and cell survival in a novel p53-null mammary epithelial cell model. FASEB J. 2000, 14, 2291–2302. [Google Scholar] [CrossRef]
  38. Mackay, H.L.; Moore, D.; Hall, C.; Birkbak, N.J.; Jamal-Hanjani, M.; Karim, S.A.; Phatak, V.M.; Piñon, L.; Morton, J.P.; Swanton, C.; et al. Genomic instability in mutant p53 cancer cells upon entotic engulfment. Nat. Commun. 2018, 9, 3070. [Google Scholar] [CrossRef] [Green Version]
  39. Rotter, V.; Aloni-Grinstein, R.; Schwartz, D.; Elkind, N.B.; Simons, A.; Wolkowicz, R.; Lavigne, M.; Beserman, P.; Kapon, A.; Goldfinger, N. Does wild-type p53 play a role in normal cell differentiation? Semin. Cancer Biol. 1994, 5, 229–236. [Google Scholar]
  40. Zhang, J.; Yan, W.; Chen, X. p53 is required for nerve growth factor-mediated differentiation of PC12 cells via regulation of TrkA levels. Cell Death Differ. 2006, 13, 2118–2128. [Google Scholar] [CrossRef] [Green Version]
  41. Molchadsky, A.; Shats, I.; Goldfinger, N.; Pevsner-Fischer, M.; Olson, M.; Rinon, A.; Tzahor, E.; Lozano, G.; Zipori, D.; Sarig, R.; et al. p53 Plays a Role in Mesenchymal Differentiation Programs, in a Cell Fate Dependent Manner. PLoS ONE 2008, 3, e3707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Molchadsky, A.; Ezra, O.; Amendola, P.G.; Krantz, D.; Kogan-Sakin, I.; Buganim, Y.; Rivlin, N.; Goldfinger, N.; Folgiero, V.; Falcioni, R.; et al. p53 is required for brown adipogenic differentiation and has a protective role against diet-induced obesity. Cell Death Differ. 2013, 20, 774–783. [Google Scholar] [CrossRef] [PubMed]
  43. Lin, T.; Chao, C.; Saito, S.; Mazur, S.J.; Murphy, M.E.; Appella, E.; Xu, Y. p53 induces differentiation of mouse embryonic stem cells by suppressing Nanog expression. Nat. Cell Biol. 2005, 7, 165–171. [Google Scholar] [CrossRef] [PubMed]
  44. Fagin, J.A.; Matsuo, K.; Karmakar, A.; Chen, D.L.; Tang, S.H.; Koeffler, H.P. High prevalence of mutations of the p53 gene in poorly differentiated human thyroid carcinomas. J. Clin. Investig. 1993, 91, 179–184. [Google Scholar] [CrossRef] [PubMed]
  45. Mizuno, H.; Spike, B.T.; Wahl, G.M.; Levine, A.J. Inactivation of p53 in breast cancers correlates with stem cell transcriptional signatures. Proc. Natl. Acad. Sci. USA 2010, 107, 22745–22750. [Google Scholar] [CrossRef] [Green Version]
  46. Hong, H.; Takahashi, K.; Ichisaka, T.; Aoi, T.; Kanagawa, O.; Nakagawa, M.; Okita, K.; Yamanaka, S. Suppression of induced pluripotent stem cell generation by the p53-p21 pathway. Nature 2009, 460, 1132–1135. [Google Scholar] [CrossRef]
  47. Kawamura, T.; Suzuki, J.; Wang, Y.V.; Menendez, S.; Morera, L.B.; Raya, A.; Wahl, G.M.; Izpisúa Belmonte, J.C. Linking the p53 tumour suppressor pathway to somatic cell reprogramming. Nature 2009, 460, 1140–1144. [Google Scholar] [CrossRef] [Green Version]
  48. Li, H.; Collado, M.; Villasante, A.; Strati, K.; Ortega, S.; Cañamero, M.; Blasco, M.A.; Serrano, M. The Ink4/Arf locus is a barrier for iPS cell reprogramming. Nature 2009, 460, 1136–1139. [Google Scholar] [CrossRef] [Green Version]
  49. Marión, R.M.; Strati, K.; Li, H.; Murga, M.; Blanco, R.; Ortega, S.; Fernandez-Capetillo, O.; Serrano, M.; Blasco, M.A. A p53-mediated DNA damage response limits reprogramming to ensure iPS cell genomic integrity. Nature 2009, 460, 1149–1153. [Google Scholar] [CrossRef] [Green Version]
  50. Utikal, J.; Polo, J.M.; Stadtfeld, M.; Maherali, N.; Kulalert, W.; Walsh, R.M.; Khalil, A.; Rheinwald, J.G.; Hochedlinger, K. Immortalization eliminates a roadblock during cellular reprogramming into iPS cells. Nature 2009, 460, 1145–1148. [Google Scholar] [CrossRef]
  51. Sarig, R.; Rivlin, N.; Brosh, R.; Bornstein, C.; Kamer, I.; Ezra, O.; Molchadsky, A.; Goldfinger, N.; Brenner, O.; Rotter, V. Mutant p53 facilitates somatic cell reprogramming and augments the malignant potential of reprogrammed cells. J. Exp. Med. 2010, 207, 2127–2140. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.; Yang, J.; Zheng, H.; Tomasek, G.J.; Zhang, P.; McKeever, P.E.; Lee, E.Y.-H.P.; Zhu, Y. Expression of mutant p53 proteins implicates a lineage relationship between neural stem cells and malignant astrocytic glioma in a murine model. Cancer Cell 2009, 15, 514–526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Shetzer, Y.; Kagan, S.; Koifman, G.; Sarig, R.; Kogan-Sakin, I.; Charni, M.; Kaufman, T.; Zapatka, M.; Molchadsky, A.; Rivlin, N.; et al. The onset of p53 loss of heterozygosity is differentially induced in various stem cell types and may involve the loss of either allele. Cell Death Differ. 2014, 21, 1419–1431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Koifman, G.; Shetzer, Y.; Eizenberger, S.; Solomon, H.; Rotkopf, R.; Molchadsky, A.; Lonetto, G.; Goldfinger, N.; Rotter, V. A mutant p53-dependent embryonic stem cell gene signature is associated with augmented tumorigenesis of stem cells. Cancer Res. 2018, 78, 5833–5847. [Google Scholar] [CrossRef] [Green Version]
  55. Smith, Z.D.; Sindhu, C.; Meissner, A. Molecular features of cellular reprogramming and development. Nat. Rev. Mol. Cell Biol. 2016, 17, 139–154. [Google Scholar] [CrossRef] [Green Version]
  56. Courtot, A.-M.; Magniez, A.; Oudrhiri, N.; Féraud, O.; Bacci, J.; Gobbo, E.; Proust, S.; Turhan, A.G.; Bennaceur-Griscelli, A. Morphological analysis of human induced pluripotent stem cells during induced differentiation and reverse programming. Biores. Open Access 2014, 3, 206–216. [Google Scholar] [CrossRef]
  57. Loizou, E.; Banito, A.; Livshits, G.; Ho, Y.-J.; Koche, R.P.; Sánchez-Rivera, F.J.; Mayle, A.; Chen, C.-C.; Kinalis, S.; Bagger, F.O.; et al. A Gain-of-Function p53-Mutant Oncogene Promotes Cell Fate Plasticity and Myeloid Leukemia through the Pluripotency Factor FOXH1. Cancer Discov. 2019, 9, 962–979. [Google Scholar] [CrossRef] [Green Version]
  58. Hanahan, D.; Weinberg, R. a Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  59. Humpton, T.J.; Vousden, K.H. Regulation of Cellular Metabolism and Hypoxia by p53. Cold Spring Harb. Perspect. Med. 2016, 6, a026146. [Google Scholar] [CrossRef] [Green Version]
  60. Gomes, A.S.; Ramos, H.; Soares, J.; Saraiva, L. p53 and glucose metabolism: An orchestra to be directed in cancer therapy. Pharmacol. Res. 2018, 131, 75–86. [Google Scholar] [CrossRef]
  61. Goldstein, I.; Yizhak, K.; Madar, S.; Goldfinger, N.; Ruppin, E.; Rotter, V. p53 promotes the expression of gluconeogenesis-related genes and enhances hepatic glucose production. Cancer Metab. 2013, 1, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Maddocks, O.D.K.; Berkers, C.R.; Mason, S.M.; Zheng, L.; Blyth, K.; Gottlieb, E.; Vousden, K.H. Serine starvation induces stress and p53-dependent metabolic remodelling in cancer cells. Nature 2012, 493, 542–546. [Google Scholar] [CrossRef] [PubMed]
  63. Prokesch, A.; Graef, F.A.; Madl, T.; Kahlhofer, J.; Heidenreich, S.; Schumann, A.; Moyschewitz, E.; Pristoynik, P.; Blaschitz, A.; Knauer, M.; et al. Liver p53 is stabilized upon starvation and required for amino acid catabolism and gluconeogenesis. FASEB J. 2017, 31, 732–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhou, G.; Wang, J.; Zhao, M.; Xie, T.-X.; Tanaka, N.; Sano, D.; Patel, A.A.; Ward, A.M.; Sandulache, V.C.; Jasser, S.A.; et al. Gain-of-Function Mutant p53 Promotes Cell Growth and Cancer Cell Metabolism via Inhibition of AMPK Activation. Mol. Cell 2014, 54, 960–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Tran, T.Q.; Lowman, X.H.; Reid, M.A.; Mendez-Dorantes, C.; Pan, M.; Yang, Y.; Kong, M. Tumor-associated mutant p53 promotes cancer cell survival upon glutamine deprivation through p21 induction. Oncogene 2017, 36, 1991–2001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Chryplewicz, A.; Tienda, S.M.; Nahotko, D.A.; Peters, P.N.; Lengyel, E.; Eckert, M.A. Mutant p53 regulates LPA signaling through lysophosphatidic acid phosphatase type 6. Sci. Rep. 2019, 9, 5195. [Google Scholar] [CrossRef]
  67. Freed-Pastor, W.A.A.; Mizuno, H.; Zhao, X.; Langerød, A.; Moon, S.-H.; Rodriguez-Barrueco, R.; Barsotti, A.; Chicas, A.; Li, W.; Polotskaia, A.; et al. Mutant p53 Disrupts Mammary Tissue Architecture via the Mevalonate Pathway. Cell 2012, 148, 244–258. [Google Scholar] [CrossRef] [Green Version]
  68. Parrales, A.; Thoenen, E.; Iwakuma, T. The interplay between mutant p53 and the mevalonate pathway. Cell Death Differ. 2018, 25, 460–470. [Google Scholar] [CrossRef]
  69. Mullen, P.J.; Yu, R.; Longo, J.; Archer, M.C.; Penn, L.Z. The interplay between cell signalling and the mevalonate pathway in cancer. Nat. Rev. Cancer 2016, 16, 718–731. [Google Scholar] [CrossRef]
  70. Iannelli, F.; Lombardi, R.; Milone, M.R.; Pucci, B.; De Rienzo, S.; Budillon, A.; Bruzzese, F. Targeting Mevalonate Pathway in Cancer Treatment: Repurposing of Statins. Recent Pat. Anticancer. Drug Discov. 2018, 13, 184–200. [Google Scholar] [CrossRef]
  71. Parrales, A.; Ranjan, A.; Iyer, S.V.; Padhye, S.; Weir, S.J.; Roy, A.; Iwakuma, T. DNAJA1 controls the fate of misfolded mutant p53 through the mevalonate pathway. Nat. Cell Biol. 2016, 18, 1233–1243. [Google Scholar] [CrossRef] [PubMed]
  72. Freed-Pastor, W.; Prives, C. Targeting mutant p53 through the mevalonate pathway. Nat. Cell Biol. 2016, 18, 1122–1124. [Google Scholar] [CrossRef] [PubMed]
  73. Ingallina, E.; Sorrentino, G.; Bertolio, R.; Lisek, K.; Zannini, A.; Azzolin, L.; Severino, L.U.; Scaini, D.; Mano, M.; Mantovani, F.; et al. Mechanical cues control mutant p53 stability through a mevalonate-RhoA axis. Nat. Cell Biol. 2018, 20. [Google Scholar] [CrossRef] [PubMed]
  74. Liou, G.-Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Gorrini, C.; Harris, I.S.; Mak, T.W. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 2013, 12, 931–947. [Google Scholar] [CrossRef]
  76. Moloney, J.N.; Cotter, T.G. ROS signalling in the biology of cancer. Semin. Cell Dev. Biol. 2018, 80, 50–64. [Google Scholar] [CrossRef]
  77. Kumar, B.; Koul, S.; Khandrika, L.; Meacham, R.B.; Koul, H.K. Oxidative Stress Is Inherent in Prostate Cancer Cells and Is Required for Aggressive Phenotype. Cancer Res. 2008, 68, 1777–1785. [Google Scholar] [CrossRef] [Green Version]
  78. Leone, A.; Roca, M.S.; Ciardiello, C.; Costantini, S.; Budillon, A. Oxidative Stress Gene Expression Profile Correlates with Cancer Patient Poor Prognosis: Identification of Crucial Pathways Might Select Novel Therapeutic Approaches. Oxid. Med. Cell. Longev. 2017, 2017, 2597581. [Google Scholar] [CrossRef] [Green Version]
  79. Kalo, E.; Kogan-Sakin, I.; Solomon, H.; Bar-Nathan, E.; Shay, M.; Shetzer, Y.; Dekel, E.; Goldfinger, N.; Buganim, Y.; Stambolsky, P.; et al. Mutant p53 R273H attenuates the expression of phase 2 detoxifying enzymes and promotes the survival of cells with high levels of reactive oxygen species. J. Cell Sci. 2012, 125, 5578–5586. [Google Scholar] [CrossRef] [Green Version]
  80. Liu, D.S.; Duong, C.P.; Haupt, S.; Montgomery, K.G.; House, C.M.; Azar, W.J.; Pearson, H.B.; Fisher, O.M.; Read, M.; Guerra, G.R.; et al. Inhibiting the system xC−/glutathione axis selectively targets cancers with mutant-p53 accumulation. Nat. Commun. 2017, 8, 14844. [Google Scholar] [CrossRef] [Green Version]
  81. Walerych, D.; Lisek, K.; Sommaggio, R.; Piazza, S.; Ciani, Y.; Dalla, E.; Rajkowska, K.; Gaweda-Walerych, K.; Ingallina, E.; Tonelli, C.; et al. Proteasome machinery is instrumental in a common gain-of-function program of the p53 missense mutants in cancer. Nat. Cell Biol. 2016, 18, 897–909. [Google Scholar] [CrossRef] [PubMed]
  82. Ali, A.; Wang, Z.; Fu, J.; Ji, L.; Liu, J.; Li, L.; Wang, H.; Chen, J.; Caulin, C.; Myers, J.N.; et al. Differential regulation of the REGγ-proteasome pathway by p53/TGF-β signalling and mutant p53 in cancer cells. Nat. Commun. 2013, 4, 2667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhang, C.; Liu, J.; Liang, Y.; Wu, R.; Zhao, Y.; Hong, X.; Lin, M.; Yu, H.; Liu, L.; Levine, A.J.; et al. Tumour-associated mutant p53 drives the Warburg effect. Nat. Commun. 2013, 4, 2935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hernández-Reséndiz, I.; Gallardo-Pérez, J.C.; López-Macay, A.; Robledo-Cadena, D.X.; García-Villa, E.; Gariglio, P.; Saavedra, E.; Moreno-Sánchez, R.; Rodríguez-Enríquez, S. Mutant p53 R248Q downregulates oxidative phosphorylation and upregulates glycolysis under normoxia and hypoxia in human cervix cancer cells. J. Cell. Physiol. 2018, 1–13. [Google Scholar]
  85. Basu, S.; Gnanapradeepan, K.; Barnoud, T.; Kung, C.-P.; Tavecchio, M.; Scott, J.; Watters, A.; Chen, Q.; Kossenkov, A.V.; Murphy, M.E. Mutant p53 controls tumor metabolism and metastasis by regulating PGC-1α. Genes Dev. 2018, 32, 230–243. [Google Scholar] [CrossRef]
  86. Lonetto, G.; Koifman, G.; Silberman, A.; Attery, A.; Solomon, H.; Levin-Zaidman, S.; Goldfinger, N.; Porat, Z.; Erez, A.; Rotter, V. Mutant p53-dependent mitochondrial metabolic alterations in a mesenchymal stem cell-based model of progressive malignancy. Cell Death Differ. 2018, 26, 1566–1581. [Google Scholar] [CrossRef] [Green Version]
  87. Viale, A.; Pettazzoni, P.; Lyssiotis, C.A.; Ying, H.; Sánchez, N.; Marchesini, M.; Carugo, A.; Green, T.; Seth, S.; Giuliani, V.; et al. Oncogene ablation-resistant pancreatic cancer cells depend on mitochondrial function. Nature 2014, 514, 628–632. [Google Scholar] [CrossRef] [Green Version]
  88. Sancho, P.; Burgos-Ramos, E.; Tavera, A.; Bou Kheir, T.; Jagust, P.; Schoenhals, M.; Barneda, D.; Sellers, K.; Campos-Olivas, R.; Graña, O.; et al. MYC/PGC-1α Balance Determines the Metabolic Phenotype and Plasticity of Pancreatic Cancer Stem Cells. Cell Metab. 2015, 22, 590–605. [Google Scholar] [CrossRef] [Green Version]
  89. Bissell, M.J.; Hines, W.C. Why don’t we get more cancer? A proposed role of the microenvironment in restraining cancer progression. Nat. Med. 2011, 17, 320–329. [Google Scholar] [CrossRef] [Green Version]
  90. Stein, Y.; Aloni-Grinstein, R.; Rotter, V. Mutant p53—A potential player in shaping the tumor–stroma crosstalk. J. Mol. Cell Biol. 2019, 11, 600–604. [Google Scholar] [CrossRef]
  91. Rivlin, N.; Katz, S.; Doody, M.; Sheffer, M.; Horesh, S.; Molchadsky, A.; Koifman, G.; Shetzer, Y.; Goldfinger, N.; Rotter, V.; et al. Rescue of embryonic stem cells from cellular transformation by proteomic stabilization of mutant p53 and conversion into WT conformation. Proc. Natl. Acad. Sci. USA 2014, 111, 7006–7011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Turrell, F.K.; Kerr, E.M.; Gao, M.; Thorpe, H.; Doherty, G.J.; Cridge, J.; Shorthouse, D.; Speed, A.; Samarajiwa, S.; Hall, B.A.; et al. Lung tumors with distinct p53 mutations respond similarly to p53 targeted therapy but exhibit genotype-specific statin sensitivity. Genes Dev. 2017, 31, 1339–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Walerych, D.; Pruszko, M.; Zyla, L.; Wezyk, M.; Gaweda-Walerych, K.; Zylicz, A. Wild-type p53 oligomerizes more efficiently than p53 hot-spot mutants and overcomes mutant p53 gain-of-function via a “dominant-positive” mechanism. Oncotarget 2018, 9, 32063–32080. [Google Scholar] [CrossRef] [PubMed]
  94. Abegglen, L.M.; Caulin, A.F.; Chan, A.; Lee, K.; Robinson, R.; Campbell, M.S.; Kiso, W.K.; Schmitt, D.L.; Waddell, P.J.; Bhaskara, S.; et al. Potential Mechanisms for Cancer Resistance in Elephants and Comparative Cellular Response to DNA Damage in Humans. JAMA 2015, 314, 1850. [Google Scholar] [CrossRef]
  95. Sulak, M.; Fong, L.; Mika, K.; Chigurupati, S.; Yon, L.; Mongan, N.P.; Emes, R.D.; Lynch, V.J. TP53 copy number expansion is associated with the evolution of increased body size and an enhanced DNA damage response in elephants. Elife 2016, 5, e11994. [Google Scholar] [CrossRef]
  96. Bykov, V.J.N.; Eriksson, S.E.; Bianchi, J.; Wiman, K.G. Targeting mutant p53 for efficient cancer therapy. Nat. Rev. Cancer 2018, 18, 89–102. [Google Scholar] [CrossRef]
  97. Tal, P.; Eizenberger, S.; Cohen, E.; Goldfinger, N.; Pietrokovski, S.; Oren, M.; Rotter, V. Cancer therapeutic approach based on conformational stabilization of mutant p53 protein by small peptides. Oncotarget 2016, 7, 11817–11837. [Google Scholar] [CrossRef] [Green Version]
  98. Liu, K.; Lin, F.-T.; Graves, J.D.; Lee, Y.-J.; Lin, W.-C. Mutant p53 perturbs DNA replication checkpoint control through TopBP1 and Treslin. Proc. Natl. Acad. Sci. USA 2017, 114, E3766–E3775. [Google Scholar] [CrossRef] [Green Version]
  99. Polotskaia, A.; Xiao, G.; Reynoso, K.; Martin, C.; Qiu, W.-G.; Hendrickson, R.C.; Bargonetti, J. Proteome-wide analysis of mutant p53 targets in breast cancer identifies new levels of gain-of-function that influence PARP, PCNA, and MCM4. Proc. Natl. Acad. Sci. USA 2015, 112, E1220–E1229. [Google Scholar] [CrossRef] [Green Version]
  100. Kollareddy, M.; Dimitrova, E.; Vallabhaneni, K.C.; Chan, A.; Le, T.; Chauhan, K.M.; Carrero, Z.I.; Ramakrishnan, G.; Watabe, K.; Haupt, Y.; et al. Regulation of nucleotide metabolism by mutant p53 contributes to its gain-of-function activities. Nat. Commun. 2015, 6, 7389. [Google Scholar] [CrossRef] [Green Version]
  101. Vaughan, C.A.; Singh, S.; Windle, B.; Yeudall, W.A.; Frum, R.; Grossman, S.R.; Deb, S.P.; Deb, S. Gain-of-Function Activity of Mutant p53 in Lung Cancer through Up-Regulation of Receptor Protein Tyrosine Kinase Axl. Genes Cancer 2012, 3, 491–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Masciarelli, S.; Fontemaggi, G.; Di Agostino, S.; Donzelli, S.; Carcarino, E.; Strano, S.; Blandino, G. Gain-of-function mutant p53 downregulates miR-223 contributing to chemoresistance of cultured tumor cells. Oncogene 2014, 33, 1601–1608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Irwin, M.S.; Kondo, K.; Marin, M.C.; Cheng, L.S.; Hahn, W.C.; Kaelin, W.G. Chemosensitivity linked to p73 function. Cancer Cell 2003, 3, 403–410. [Google Scholar] [CrossRef] [Green Version]
  104. Di Como, C.J.; Gaiddon, C.; Prives, C. p73 Function Is Inhibited by Tumor-Derived p53 Mutants in Mammalian Cells. Mol. Cell. Biol. 1999, 19, 1438–1449. [Google Scholar] [CrossRef] [Green Version]
  105. Stambolsky, P.; Tabach, Y.; Fontemaggi, G.; Weisz, L.; Maor-Aloni, R.; Siegfried, Z.; Sigfried, Z.; Shiff, I.; Kogan, I.; Shay, M.; et al. Modulation of the vitamin D3 response by cancer-associated mutant p53. Cancer Cell 2010, 17, 273–285. [Google Scholar] [CrossRef] [Green Version]
  106. Do, P.M.; Varanasi, L.; Fan, S.; Li, C.; Kubacka, I.; Newman, V.; Chauhan, K.; Daniels, S.R.; Boccetta, M.; Garrett, M.R.; et al. Mutant p53 cooperates with ETS2 to promote etoposide resistance. Genes Dev. 2012, 26, 830–845. [Google Scholar] [CrossRef] [Green Version]
  107. Lisek, K.; Campaner, E.; Ciani, Y.; Walerych, D.; Del Sal, G. Mutant p53 tunes the NRF2-dependent antioxidant response to support survival of cancer cells. Oncotarget 2018, 9, 20508–20523. [Google Scholar] [CrossRef] [Green Version]
  108. Krishnan, S.R.; Nair, B.C.; Sareddy, G.R.; Roy, S.S.; Natarajan, M.; Suzuki, T.; Peng, Y.; Raj, G.; Vadlamudi, R.K. Novel role of PELP1 in regulating chemotherapy response in mutant p53-expressing triple negative breast cancer cells. Breast Cancer Res. Treat. 2015, 150, 487–499. [Google Scholar] [CrossRef] [Green Version]
  109. Chee, J.L.Y.; Saidin, S.; Lane, D.P.; Leong, S.M.; Noll, J.E.; Neilsen, P.M.; Phua, Y.T.; Gabra, H.; Lim, T.M. Wild-type and mutant p53 mediate cisplatin resistance through interaction and inhibition of active caspase-9. Cell Cycle 2013, 12, 278–288. [Google Scholar] [CrossRef] [Green Version]
  110. Kolukula, V.K.; Sahu, G.; Wellstein, A.; Rodriguez, O.C.; Preet, A.; Iacobazzi, V.; D’Orazi, G.; Albanese, C.; Palmieri, F.; Avantaggiati, M.L. SLC25A1, or CIC, is a novel transcriptional target of mutant p53 and a negative tumor prognostic marker. Oncotarget 2014, 5, 1212–1225. [Google Scholar] [CrossRef] [Green Version]
  111. Frank, A.K.; Pietsch, E.C.; Dumont, P.; Tao, J.; Murphy, M.E. Wild-type and mutant p53 proteins interact with mitochondrial caspase-3. Cancer Biol. Ther. 2011, 11, 740–745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Alam, S.K.; Yadav, V.K.; Bajaj, S.; Datta, A.; Dutta, S.K.; Bhattacharyya, M.; Bhattacharya, S.; Debnath, S.; Roy, S.; Boardman, L.A.; et al. DNA damage-induced ephrin-B2 reverse signaling promotes chemoresistance and drives EMT in colorectal carcinoma harboring mutant p53. Cell Death Differ. 2016, 23, 707–722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Sampath, J.; Sun, D.; Kidd, V.J.; Grenet, J.; Gandhi, A.; Shapiro, L.H.; Wang, Q.; Zambetti, G.P.; Schuetz, J.D. Mutant p53 cooperates with ETS and selectively up-regulates human MDR1 not MRP1. J. Biol. Chem. 2001, 276, 39359–39367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Ali, A.; Shah, A.S.; Ahmad, A. Gain-of-function of mutant p53: Mutant p53 enhances cancer progression by inhibiting KLF17 expression in invasive breast carcinoma cells. Cancer Lett. 2014, 354, 87–96. [Google Scholar] [CrossRef]
  115. Donzelli, S.; Fontemaggi, G.; Fazi, F.; Di Agostino, S.; Padula, F.; Biagioni, F.; Muti, P.; Strano, S.; Blandino, G. MicroRNA-128-2 targets the transcriptional repressor E2F5 enhancing mutant p53 gain of function. Cell Death Differ. 2012, 19, 1038–1048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Chin, K.V.; Ueda, K.; Pastan, I.; Gottesman, M.M. Modulation of activity of the promoter of the human MDR1 gene by Ras and p53. Science 1992, 255, 459–462. [Google Scholar] [CrossRef] [PubMed]
  117. Buganim, Y.; Kalo, E.; Brosh, R.; Besserglick, H.; Nachmany, I.; Rais, Y.; Stambolsky, P.; Tang, X.; Milyavsky, M.; Shats, I.; et al. Mutant p53 protects cells from 12-O-tetradecanoylphorbol-13-acetate-induced death by attenuating activating transcription factor 3 induction. Cancer Res. 2006, 66, 10750–10759. [Google Scholar] [CrossRef] [Green Version]
  118. Pugacheva, E.N.; Ivanov, A.V.; Kravchenko, J.E.; Kopnin, B.P.; Levine, A.J.; Chumakov, P.M. Novel gain of function activity of p53 mutants: Activation of the dUTPase gene expression leading to resistance to 5-fluorouracil. Oncogene 2002, 21, 4595–4600. [Google Scholar] [CrossRef] [Green Version]
  119. Zalcenstein, A.; Stambolsky, P.; Weisz, L.; Müller, M.; Wallach, D.; Goncharov, T.M.; Krammer, P.H.; Rotter, V.; Oren, M. Mutant p53 gain of function: Repression of CD95(Fas/APO-1) gene expression by tumor-associated p53 mutants. Oncogene 2003, 22, 5667–5676. [Google Scholar] [CrossRef] [Green Version]
  120. Scian, M.J.; Stagliano, K.E.R.; Anderson, M.A.E.; Hassan, S.; Bowman, M.; Miles, M.F.; Deb, S.P.; Deb, S. Tumor-derived p53 mutants induce NF-kappaB2 gene expression. Mol. Cell. Biol. 2005, 25, 10097–10110. [Google Scholar] [CrossRef] [Green Version]
  121. Huang, X.; Zhang, Y.; Tang, Y.; Butler, N.; Kim, J.; Guessous, F.; Schiff, D.; Mandell, J.; Abounader, R. A novel PTEN/mutant p53/c-Myc/Bcl-XL axis mediates context-dependent oncogenic effects of PTEN with implications for cancer prognosis and therapy. Neoplasia 2013, 15, 952–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Weisz, L.; Damalas, A.; Liontos, M.; Karakaidos, P.; Fontemaggi, G.; Maor-Aloni, R.; Kalis, M.; Levrero, M.; Strano, S.; Gorgoulis, V.G.; et al. Mutant p53 enhances nuclear factor κB activation by tumor necrosis factor α in cancer cells. Cancer Res. 2007, 67, 2396–2401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Di Minin, G.; Bellazzo, A.; Dal Ferro, M.; Chiaruttini, G.; Nuzzo, S.; Bicciato, S.; Piazza, S.; Rami, D.; Bulla, R.; Sommaggio, R.; et al. Mutant p53 Reprograms TNF Signaling in Cancer Cells through Interaction with the Tumor Suppressor DAB2IP. Mol. Cell 2014, 56, 617–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ubertini, V.; Norelli, G.; D’Arcangelo, D.; Gurtner, A.; Cesareo, E.; Baldari, S.; Gentileschi, M.P.; Piaggio, G.; Nisticò, P.; Soddu, S.; et al. Mutant p53 gains new function in promoting inflammatory signals by repression of the secreted interleukin-1 receptor antagonist. Oncogene 2015, 34, 2493–2504. [Google Scholar] [CrossRef]
  125. Solomon, H.; Buganim, Y.; Kogan-Sakin, I.; Pomeraniec, L.; Assia, Y.; Madar, S.; Goldstein, I.; Brosh, R.; Kalo, E.; Beatus, T.; et al. Various p53 mutant proteins differently regulate the Ras circuit to induce a cancer-related gene signature. J. Cell Sci. 2012, 125, 3144–3152. [Google Scholar] [CrossRef] [Green Version]
  126. Ham, S.W.; Jeon, H.-Y.; Jin, X.; Kim, E.-J.; Kim, J.-K.; Shin, Y.J.; Lee, Y.; Kim, S.H.; Lee, S.Y.; Seo, S.; et al. TP53 gain-of-function mutation promotes inflammation in glioblastoma. Cell Death Differ. 2019, 26, 409–425. [Google Scholar] [CrossRef] [Green Version]
  127. Addadi, Y.; Moskovits, N.; Granot, D.; Lozano, G.; Carmi, Y.; Apte, R.N.; Neeman, M.; Oren, M. p53 status in stromal fibroblasts modulates tumor growth in an SDF1-dependent manner. Cancer Res. 2010, 70, 9650–9658. [Google Scholar] [CrossRef] [Green Version]
  128. Buganim, Y.; Solomon, H.; Rais, Y.; Kistner, D.; Nachmany, I.; Brait, M.; Madar, S.; Goldstein, I.; Kalo, E.; Adam, N.; et al. p53 Regulates the Ras Circuit to Inhibit the Expression of a Cancer-Related Gene Signature by Various Molecular Pathways. Cancer Res. 2010, 70, 2274–2284. [Google Scholar] [CrossRef] [Green Version]
  129. Yeudall, W.A.; Vaughan, C.A.; Miyazaki, H.; Ramamoorthy, M.; Choi, M.-Y.; Chapman, C.G.; Wang, H.; Black, E.; Bulysheva, A.A.; Deb, S.P.S.; et al. Gain-of-function mutant p53 upregulates CXC chemokines and enhances cell migration. Carcinogenesis 2012, 33, 442–451. [Google Scholar] [CrossRef]
  130. Kalo, E.; Buganim, Y.; Shapira, K.E.; Besserglick, H.; Goldfinger, N.; Weisz, L.; Stambolsky, P.; Henis, Y.I.; Rotter, V. Mutant p53 attenuates the SMAD-dependent transforming growth factor beta1 (TGF-beta1) signaling pathway by repressing the expression of TGF-beta receptor type II. Mol. Cell. Biol. 2007, 27, 8228–8242. [Google Scholar] [CrossRef] [Green Version]
  131. Madar, S.; Harel, E.; Goldstein, I.; Stein, Y.; Kogan-Sakin, I.; Kamer, I.; Solomon, H.; Dekel, E.; Tal, P.; Goldfinger, N.; et al. Mutant p53 attenuates the anti-tumorigenic activity of fibroblasts-secreted interferon beta. PLoS ONE 2013, 8, e61353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Cooks, T.; Pateras, I.S.; Jenkins, L.M.; Patel, K.M.; Robles, A.I.; Morris, J.; Forshew, T.; Appella, E.; Gorgoulis, V.G.; Harris, C.C. Mutant p53 cancers reprogram macrophages to tumor supporting macrophages via exosomal miR-1246. Nat. Commun. 2018, 9, 771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Zhao, Y.; Li, Y.; Sheng, J.; Wu, F.; Li, K.; Huang, R.; Wang, X.; Jiao, T.; Guan, X.; Lu, Y.; et al. P53-R273H mutation enhances colorectal cancer stemness through regulating specific lncRNAs. J. Exp. Clin. Cancer Res. 2019, 38, 379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Solomon, H.; Dinowitz, N.; Pateras, I.S.; Cooks, T.; Shetzer, Y.; Molchadsky, A.; Charni, M.; Rabani, S.; Koifman, G.; Tarcic, O.; et al. Mutant p53 gain of function underlies high expression levels of colorectal cancer stem cells markers. Oncogene 2018, 37, 1669–1684. [Google Scholar] [CrossRef] [Green Version]
  135. Butera, G.; Pacchiana, R.; Mullappilly, N.; Margiotta, M.; Bruno, S.; Conti, P.; Riganti, C.; Donadelli, M. Mutant p53 prevents GAPDH nuclear translocation in pancreatic cancer cells favoring glycolysis and 2-deoxyglucose sensitivity. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1914–1923. [Google Scholar] [CrossRef]
  136. Valenti, F.; Ganci, F.; Fontemaggi, G.; Sacconi, A.; Strano, S.; Blandino, G.; Di Agostino, S. Gain of function mutant p53 proteins cooperate with E2F4 to transcriptionally downregulate RAD17 and BRCA1 gene expression. Oncotarget 2015, 6. [Google Scholar] [CrossRef] [Green Version]
  137. Subramanian, M.; Francis, P.; Bilke, S.; Li, X.L.; Hara, T.; Lu, X.; Jones, M.F.; Walker, R.L.; Zhu, Y.; Pineda, M.; et al. A mutant p53/let-7i-axis-regulated gene network drives cell migration, invasion and metastasis. Oncogene 2015, 34, 1094–1104. [Google Scholar] [CrossRef] [Green Version]
  138. Gaiddon, C.; Lokshin, M.; Ahn, J.; Zhang, T.; Prives, C. A subset of tumor-derived mutant forms of p53 down-regulate p63 and p73 through a direct interaction with the p53 core domain. Mol. Cell. Biol. 2001, 21, 1874–1887. [Google Scholar] [CrossRef] [Green Version]
  139. Ji, L.; Xu, J.; Liu, J.; Amjad, A.; Zhang, K.; Liu, Q.; Zhou, L.; Xiao, J.; Li, X. Mutant p53 promotes tumor cell malignancy by both positive and negative regulation of the transforming growth factor β (TGF-β) pathway. J. Biol. Chem. 2015, 290, 11729–11740. [Google Scholar] [CrossRef] [Green Version]
  140. Zhao, Y.; Zhang, C.; Yue, X.; Li, X.; Liu, J.; Yu, H.; Belyi, V.A.; Yang, Q.; Feng, Z.; Hu, W. Pontin, a new mutant p53-binding protein, promotes gain-of-function of mutant p53. Cell Death Differ. 2015, 22, 1824–1836. [Google Scholar] [CrossRef] [Green Version]
  141. Muller, P.A.J.; Trinidad, A.G.; Timpson, P.; Morton, J.P.; Zanivan, S.; Van Den Berghe, P.V.E.; Nixon, C.; Karim, S.A.; Caswell, P.T.; Noll, J.E.; et al. Mutant p53 enhances MET trafficking and signalling to drive cell scattering and invasion. Oncogene 2012, 32, 1252–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Muller, P.A.J.; Caswell, P.T.; Doyle, B.; Iwanicki, M.P.; Tan, E.H.; Karim, S.; Lukashchuk, N.; Gillespie, D.A.; Ludwig, R.L.; Gosselin, P.; et al. Mutant p53 Drives Invasion by Promoting Integrin Recycling. Cell 2009, 139, 1327–1341. [Google Scholar] [CrossRef] [PubMed]
  143. Wei, S.; Wang, H.; Lu, C.; Malmut, S.; Zhang, J.; Ren, S.; Yu, G.; Wang, W.; Tang, D.D.; Yan, C. The activating transcription factor 3 protein suppresses the oncogenic function of mutant p53 proteins. J. Biol. Chem. 2014, 289, 8947–8959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Stindt, M.H.; Muller, P.A.J.; Ludwig, R.L.; Kehrloesser, S.; Dötsch, V.; Vousden, K.H. Functional interplay between MDM2, p63/p73 and mutant p53. Oncogene 2015, 34, 4300–4310. [Google Scholar] [CrossRef]
  145. Vogiatzi, F.; Brandt, D.T.; Schneikert, J.; Fuchs, J.; Grikscheit, K.; Wanzel, M.; Pavlakis, E.; Charles, J.P.; Timofeev, O.; Nist, A.; et al. Mutant p53 promotes tumor progression and metastasis by the endoplasmic reticulum UDPase ENTPD5. Proc. Natl. Acad. Sci. USA 2016, 113, E8433–E8442. [Google Scholar] [CrossRef] [Green Version]
  146. Coffill, C.R.; Muller, P.A.J.J.; Oh, H.K.; Neo, S.P.; Hogue, K.A.; Cheok, C.F.; Vousden, K.H.; Lane, D.P.; Blackstock, W.P.; Gunaratne, J. Mutant p53 interactome identifies nardilysin as a p53R273H-specific binding partner that promotes invasion. EMBO Rep. 2012, 13, 638–644. [Google Scholar] [CrossRef] [Green Version]
  147. Girardini, J.E.; Napoli, M.; Piazza, S.; Rustighi, A.; Marotta, C.; Radaelli, E.; Capaci, V.; Jordan, L.; Quinlan, P.; Thompson, A.; et al. A Pin1/Mutant p53 Axis Promotes Aggressiveness in Breast Cancer. Cancer Cell 2011, 20, 79–91. [Google Scholar] [CrossRef]
  148. Wang, H.; Bao, W.; Jiang, F.; Che, Q.; Chen, Z.; Wang, F.; Tong, H.; Dai, C.; He, X.; Liao, Y.; et al. Mutant p53 (p53-R248Q) functions as an oncogene in promoting endometrial cancer by up-regulating REGγ. Cancer Lett. 2015, 360, 269–279. [Google Scholar] [CrossRef]
  149. Arjonen, A.; Kaukonen, R.; Mattila, E.; Rouhi, P.; Högnäs, G.; Sihto, H.; Miller, B.W.; Morton, J.P.; Bucher, E.; Taimen, P.; et al. Mutant p53-associated myosin-X upregulation promotes breast cancer invasion and metastasis. J. Clin. Investig. 2014, 124, 1069–1082. [Google Scholar] [CrossRef] [Green Version]
  150. Weissmueller, S.; Manchado, E.; Saborowski, M.; Morris, J.P.; Wagenblast, E.; Davis, C.A.; Moon, S.-H.; Pfister, N.T.; Tschaharganeh, D.F.; Kitzing, T.; et al. Mutant p53 drives pancreatic cancer metastasis through cell-autonomous PDGF receptor β signaling. Cell 2014, 157, 382–394. [Google Scholar] [CrossRef] [Green Version]
  151. Neilsen, P.M.; Noll, J.E.; Mattiske, S.; Bracken, C.P.; Gregory, P.A.; Schulz, R.B.; Lim, S.P.; Kumar, R.; Suetani, R.J.; Goodall, G.J.; et al. Mutant p53 drives invasion in breast tumors through up-regulation of miR-155. Oncogene 2013, 32, 2992–3000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Tanaka, N.; Zhao, M.; Tang, L.; Patel, A.A.; Xi, Q.; Van, H.T.; Takahashi, H.; Osman, A.A.; Zhang, J.; Wang, J.; et al. Gain-of-function mutant p53 promotes the oncogenic potential of head and neck squamous cell carcinoma cells by targeting the transcription factors FOXO3a and FOXM1. Oncogene 2018, 37, 1279–1292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Xu, Y.; Zhang, Q.; Miao, C.; Dongol, S.; Li, Y.; Jin, C.; Dong, R.; Li, Y.; Yang, X.; Kong, B. CCNG1 (Cyclin G1) regulation by mutant-P53 via induction of Notch3 expression promotes high-grade serous ovarian cancer (HGSOC) tumorigenesis and progression. Cancer Med. 2019, 8, 351–362. [Google Scholar] [CrossRef] [PubMed]
  154. Kogan-Sakin, I.; Tabach, Y.; Buganim, Y.; Molchadsky, A.; Solomon, H.; Madar, S.; Kamer, I.; Stambolsky, P.; Shelly, A.; Goldfinger, N.; et al. Mutant p53(R175H) upregulates Twist1 expression and promotes epithelial-mesenchymal transition in immortalized prostate cells. Cell Death Differ. 2011, 18, 271–281. [Google Scholar] [CrossRef] [PubMed]
  155. Lv, T.; Wu, X.; Sun, L.; Hu, Q.; Wan, Y.; Wang, L.; Zhao, Z.; Tu, X.; Xiao, Z.-X.J. p53-R273H upregulates neuropilin-2 to promote cell mobility and tumor metastasis. Cell Death Dis. 2017, 8, e2995. [Google Scholar] [CrossRef] [PubMed]
  156. Muller, P.A.J.; Trinidad, A.G.; Caswell, P.T.; Norman, J.C.; Vousden, K.H. Mutant p53 regulates Dicer through p63-dependent and -independent mechanisms to promote an invasive phenotype. J. Biol. Chem. 2014, 289, 122–132. [Google Scholar] [CrossRef] [Green Version]
  157. Iwanicki, M.P.; Chen, H.-Y.; Iavarone, C.; Zervantonakis, I.K.; Muranen, T.; Novak, M.; Ince, T.A.; Drapkin, R.; Brugge, J.S. Mutant p53 regulates ovarian cancer transformed phenotypes through autocrine matrix deposition. JCI Insight 2016, 1, 1–20. [Google Scholar] [CrossRef]
Table 1. “Social network” of mutant p53 1.
Table 1. “Social network” of mutant p53 1.
Cancer Hallmarks Associated with GOFMutant p53 Interacting Molecules/Gene TargetsRef.
Proliferation/Abrogation of Growth SuppressorsProtein interactions (*indirect)NF-Y (R175H, R273H), p300 (R175H, R273H, R280K), TopBP1 (V143A, R175H, R248W, R249S, R273H), MCM4 and PCNA (R273H), YAP (R175H, H193L, L194F, R273H) TEAD* (R175H, H193L), EGFR * (R273H), ETS2 (R175H, R248Q, R248W, R273H), STAT3 (R175H, R248W, R248Q, R273H, R282W)[25,26,28,30,31,32,33,82,98,99]
Transcriptional targets (*indirect)REGγ (R175H, R248W, R273H, R282W), nucleotide metabolism genes (NMGs) (R175H, R249S, R273L, R273H, R280K), Axl receptor tyrosine kinase (R175H, R267P, R273C, R273H, D281G), cyclin A+cyclin B2+CDK1 (R175H, H193L, R248L, R273H, R280K), circPVT1 (R175H, H193L), miR-27a (R273H), MLL1, MLL2 and MOZ chromatin regulators (R248Q, R248W, R249S, R273H)[28,30,31,33,82,100,101]
Resistance to Cell Death/ChemoresistanceProtein interactions (*indirect)NF-Y (R273H, R249S), ZEB1 (R273H), p300 (R273H), p73 (R175H, R248W, R273H, R273C), VDR (R175H, R280K), ETS1 (D281G), ETS2 (R175H, R248W), NRF2 (R175H, R280K), PELP1 (R273H,280K), Caspase 3 (R175H, R273H) Caspase 9 (D42Y, R175H, R337H), FOXO1 (R175H, D281G)[79,102,103,104,105,106,107,108,109,110,111,112,113]
Transcriptional targets (*indirect)EFNB2 (R273H), SLC25A1 (R175H, G245A, R273H, R280K, D281G), miR-223 (Negative regulator, R175H, R273H, R280K), STMN1 (R273H, R280K), REGγ (R282W, R175H, R248W, R273H), KLF17 (R175H, R280K, R273H, R282W), NF-κB2 (R175H, R273H, D281G), Bcl-xL (R273H), miR-128-2 (R175H), MDR1 (V143A, R175H, R248W, R273H, D281G), ATF3 (Negative regulator, R175H, H179G, R248W, D281G), MEF2D (Negative regulator, R175H), dUTPase (R175H, R248W), CD95/Fas (R175 H, R248W, R273H)[102,112,113,114,115,116,117,118,119,120,121]
Tumor Microenvironment/inflammationProtein interactions (*indirect)NF-κB (R175H, R248Q, R273H), DAB2IP (R175H, M237I, R248W, R273H, R280K), c-MAFF (R273H), BTG2 (Negative regulator, R175H, H179R)[122,123,124,125]
Transcriptional targets (*indirect)sIL-1Rα (Negative regulator, R175H, R273H), CCL2 (R248L), SDF1/CXCL12 (murine R172H, R175H, R273H), CXC chemokines (R175H, H179L, R248Q, R273H, D281G) MMPs and IL-1β (R248Q, R273H), TGFβ-R2 (Negative regulator, R175H, H179R, R248W), SOCS1 (R175H, R248Q)[124,125,126,127,128,129,130,131]
Exosome releasemiR-1246 (V157F, R175H, R248W, R249S, R273H)[132]
Self-Renewal/StemnessTranscriptional targetsFOXH1 (mechanism unknown, murine R172H), lnc273–31 and lnc273–34 (R273H), embryonic stem cell gene signature (murine R172H), CD44 (R248W, R273H) Lgr5 (R172H, R273H) ALDH1A1 (R175H R248W, R273H)[45,54,57,133,134]
MetabolismProtein interactions (*indirect)AMPKα (R175H, P151S, R175H, G245C, and R282W), RhoA * (mechanism unknown, R175H, L194F, M237I, R280K), SREBP proteins (R273H), GAPDH* (R273H), PGC-1α (P72-mut-R175H and R273H), NRF2 (R175H, M237I, R248Q, R249S, R248W, R273H, C277F, R280K)[64,66,67,79,80,81,83,85,107,135]
Transcriptional targets (*indirect)SLC25A1 (R175H, G281D, G245A), SLC7A11 (Negative regulator, R175H, G266E, R273H, C277F), HMOX-1 and NQO-1 (Negative regulators, R273H), TXN and PSMC-1 (R175H, R280K), mevalonate pathway enzymes (R273H), mitochondrial metabolism genes (murine R172H), ACP6 (Negative regulator, R175H, R249S and R273H), CDKN1A (p21) (R248Q, R273H), Nucleotide metabolism genes (NMGs) (R175H, R249S, R273L, R273H, R280K), Proteasome subunit genes (R175H, M237I, R249S, R273H, R280K)[65,66,67,79,80,81,86,100,107,110]
Genomic InstabilityProtein interactions (*indirect)E2F4 (R175H), Mre11 (R248W, R273H)[34,136]
Transcriptional targets (*indirect)BRCA1 (R175H), RAD17 (R175H)[136]
Invasion and MetastasisProtein interactions (*indirect)p63 (V143A, R175H, H179Y, Y220C, I254R, R273H, R280K), p73 (V143A, R175P, R175H, H179Y, Y220C, R248W, I254R, R273H, R280K), PTEN (R273H), SMAD3 (C135Y, V143A, R175H, R248W, R273H, R282W), NF-Y (R175H, R273H), SREBPs (R273H), Pontin (R175H, R248Q, R273H) KLF17 (R280K, R282W), RCP* and MET* (R175H, R248W, R273H) Integrins* (R175H, R273H), ATF3 (V143A, R175H, R249S, R273H), MDM2 (R175H, I254R, R273H), SP1 (R248W, R280K), NRD1 (R273H), Pin1 (murine R172H, R175H, R280K), STAT3 (R175H, R248W, R248Q, R273H, R282W)[32,67,112,114,121,137,138,139,140,141,142,143,144,145,146,147]
Transcriptional targets (*indirect)REGγ (R248Q), let-7i (S241F, R248W, P273H, R280K), KLF17 (R280K, R282W), MYO10 (R175H, R273H, R280K), STMN1* (R175H), PDGFR-β (R175H, Y220C, C242R, T155P, R248W, R273H, R280K), Nucleotide metabolism genes (NMGs) (R175H, R249S, R273L, R273H, R280K), DICER1 (R273H, R280K), EFNB2 (R175H, R248W, R273H, R280K), MYC* (R273H), ZNF652* and miR-155* (R248Q, R249S, R282W), FOXM1* (R175H, C238F, G245D, R248W), Notch3 and CCNG1* (R248Q), Twist1* (epigenetic, R175H, R273H, P223L/V274F), DLX2 and NRP2* (R273H), endoplasmic reticulum UDPase ENTPD5 (C176S, R248W, R273H, R280K)[100,112,114,121,137,140,145,148,149,150,151,152,153,154,155,156,157]
1 Genes/proteins that are part of mutant p53 network are marked in bold. Shown in parentheses are the mutant p53 forms that were implicated in the interaction or transcriptional regulation of the indicated proteins or genes, respectively.

Share and Cite

MDPI and ACS Style

Stein, Y.; Rotter, V.; Aloni-Grinstein, R. Gain-of-Function Mutant p53: All the Roads Lead to Tumorigenesis. Int. J. Mol. Sci. 2019, 20, 6197. https://doi.org/10.3390/ijms20246197

AMA Style

Stein Y, Rotter V, Aloni-Grinstein R. Gain-of-Function Mutant p53: All the Roads Lead to Tumorigenesis. International Journal of Molecular Sciences. 2019; 20(24):6197. https://doi.org/10.3390/ijms20246197

Chicago/Turabian Style

Stein, Yan, Varda Rotter, and Ronit Aloni-Grinstein. 2019. "Gain-of-Function Mutant p53: All the Roads Lead to Tumorigenesis" International Journal of Molecular Sciences 20, no. 24: 6197. https://doi.org/10.3390/ijms20246197

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop