Next Article in Journal
Aged Lymphatic Vessels and Mast Cells in Perilymphatic Tissues
Next Article in Special Issue
Control of Nucleotide Metabolism Enables Mutant p53’s Oncogenic Gain-of-Function Activity
Previous Article in Journal
Chemical Profile and Antioxidant, Anti-Inflammatory, Antimutagenic and Antimicrobial Activities of Geopropolis from the Stingless Bee Melipona orbignyi
Previous Article in Special Issue
Essential Roles of E3 Ubiquitin Ligases in p53 Regulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mutant p53 Protein and the Hippo Transducers YAP and TAZ: A Critical Oncogenic Node in Human Cancers

Molecular Chemoprevention and Oncogenomic and Epigenetic Units, Italian National Cancer Institute “Regina Elena”, 00144 Rome, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(5), 961; https://doi.org/10.3390/ijms18050961
Submission received: 2 March 2017 / Revised: 11 April 2017 / Accepted: 24 April 2017 / Published: 3 May 2017
(This article belongs to the Special Issue Emerging Non-Canonical Functions and Regulation of p53)

Abstract

:
p53 protein is a well-known tumor suppressor factor that regulates cellular homeostasis. As it has several and key functions exerted, p53 is known as “the guardian of the genome” and either loss of function or gain of function mutations in the TP53 coding protein sequence are involved in cancer onset and progression. The Hippo pathway is a key regulator of developmental and regenerative physiological processes but if deregulated can induce cell transformation and cancer progression. The p53 and Hippo pathways exert a plethora of fine-tuned functions that can apparently be in contrast with each other. In this review, we propose that the p53 status can affect the Hippo pathway function by switching its outputs from tumor suppressor to oncogenic activities. In detail, we discuss: (a) the oncogenic role of the protein complex mutant p53/YAP; (b) TAZ oncogenic activation mediated by mutant p53; (c) the therapeutic potential of targeting mutant p53 to impair YAP and TAZ oncogenic functions in human cancers.

Graphical Abstract

1. Introduction

The p53 protein is a key factor for the safeguard of different cellular processes such as gene expression, cell cycle regulation, DNA damage repair, ageing, cell migration inhibition, and apoptosis. It was first identified in 1979 as a partner of the T antigen in SV40 viruses [1,2,3,4]. Therefore, p53 was primarily considered as a transforming factor until the 1980s, when it was definitively recognized as a tumor suppressor factor and named “the guardian of the genome”. TP53 gene exhibits a high mutation rate in human tumors (about 50–70%) which results either in loss or gain of function of oncogenic mutant p53 proteins [5,6,7,8].
The Hippo pathway is evolutionarily conserved and tightly controls embryos and adult tissue organ size and growth. It is heavily involved in the control of proliferation, organ size and shape during development, stem cell maintenance, metastasis, tissue regeneration, apoptosis, senescence, and differentiation. It is regulated by many factors such as cell density and polarity, metabolism, senescence, and DNA damage [9,10,11,12]. Hippo crosstalks with other signaling players resembling a network rather than a linear pathway [13,14,15]. Dysregulations in the Hippo pathway play a role in cancer insurgence [9,16,17,18,19]. Recent evidences show that the p53 and Hippo pathways are “functionally” and “physically” connected; they regulate common pathways preserving cellular and tissue homeostasis in healthy conditions. When aberrantly deregulated, the p53 and Hippo pathways promote pathological and pro-tumorigenic onsets. Understanding this complex interaction could be of great interest in dissecting tumor onset and ultimately for the design of new anticancer strategies that could concomitantly target both pathways.

2. p53 and Mutant p53 Protein Functions

p53 is a short-lived protein (6–20 min half-life) that shuttles from cytoplasm to nucleus in response to different stresses that can erode the genome integrity (i.e., DNA damage, activation of oncogenes, hypoxia, viral replication or depletion of ribonucleotides) [3,20]. Moreover, it becomes activated by several post-translational modifications such as phosphorylation, sumoylation, acetylation and prolyl-isomerization mediated by physical interaction with diverse partners such as p300, PCAF, HIPK2, SUMO1, Chk1, Chk2, ATM, ATR, and Pin1 [21,22]. p53 is a modular protein harboring four functional different domains [23]: the N-terminal domain or transactivation domain (NTD or TA) that includes a proline-rich region and a transcriptional activation domain essential for binding to transcription factors and regulators of p53 activity [24,25]; the core domain that allows the binding to DNA (DBD domain) [26,27]; the oligomerization domain (OLD) relevant for the tetramerization of p53 [28] and the C-terminal domain or regulatory domain (RD) which is involved in post-translational modifications (phosphorylation, acetylation, ubiquitination, sumoylation, prolyl-isomerization, methylation, glycosylation, and ADP-ribosylation) which affect p53 activity [23].

2.1. Transcriptional Regulation

The most important biochemical activity of p53 is the ability to regulate gene transcription. Through its TA domain, p53 interacts with TBPs and TAFs inducing the recruitment of the RNA Polymerase II and consequently the transcriptional activation of its targets genes harboring a p53 responsive element in the DNA sequence (p53RE) [29]. p53 transcriptionally controls a few hundred genes involved in the regulation of cell proliferation, DNA damage repair, and apoptosis induction [30,31,32,33].

2.2. Cell Cycle Regulation

The p53 protein exerts an important role in regulating cell cycle checkpoints [34,35]. In particular, p53 induces P21WAF1/CIP1 transcription that inhibits cyclin dependent kinases/cyclin complexes. This results in the activation of the G1 checkpoint [36,37,38]. Moreover, p53 also acts in the late G2 phase activating the DNA damage checkpoint [39]. When activated by stress or exogenously overexpressed, p53 can induce cell cycle arrest in G1 and G2/M phases and, inhibiting PCNA protein, p53 can induce cell replication blockage [3,40]. In response to DNA damage, ATM and ATR kinases phosphorylate p53, Chk2, Chk1, Mdm2, H2AFX, and BRCA1 inducing cell cycle arrest and promoting DNA damage repair [41,42,43,44,45]. G2/M arrest is mediated by p53-dependent 14-3-3σ transcription, which inhibit Cdk1/cyclin B complexes [46]. Moreover, p53 regulates centrosome formation, mitotic spindle assembly, and chromosome segregation activating the spindle checkpoint [47,48] and interacts with the transmembrane protein Gas1 maintaining cells in the G0 phase [3,49].

2.3. DNA Damage Response

One of the main functions of p53 is the maintenance of genome integrity. It has been shown that different types of DNA damage, such as ionizing radiation, chemicals, radiomimetic drugs, ultraviolet radiation, and some pharmacological treatments can induce DNA damage [50]. When the insult is not efficiently repaired, changes in chromatin structure, genetic mutations and oncogenic transformation may occur [45]. These alterations in DNA sequence or structure activate p53, which in turn trans-activates different genes involved in cell cycle arrest, DNA damage repair or apoptosis [51,52]. If TP53 is deleted or mutated, DNA damage will persist and cells replicate transmitting altered genetic heritage; this mechanism could lead to a neoplastic transformation. Moreover, p53 exerts other functions contributing to the maintenance of genome integrity: it exerts a 3′–5′ exonuclease activity [53] that is important in DNA recombination, in DNA repair and DNA replication [54]; p53 prevents chromosomic aberrations inhibiting aneuploidy and polyploidy occurrences; it blocks DNA replication until complete chromosome segregation [55]. p53 can also stimulate DNA damage repair activating GADD45 protein that can bind to PCNA and stimulate the NER system [56]; moreover, p53 can bind to TFIIH factor involved directly in the NER system [57,58].

2.4. Senescence and Aging

The p53 family members are involved in regulating ageing processes [23,59,60]. Mice models overexpressing wild-type p53 protein or p53 heterozygous mutants show reduced susceptibility to tumor formation and reduced lifespan [60,61] associated with premature aging [59,62]. p53 protein is responsible for irreversible cell cycle arrest, a characteristic of senescent cells [63,64]. In fact, P16INK4 and PML, two key factors in senescence, are p53 transcriptional targets [65]. Normal cultured cells, subjected to increased number of cell divisions, display an increased p53 activity and enhanced p21 protein levels [3]. Therefore, following cell cycle arrest the p53 protein can activate two distinct pathways: senescence or apoptosis [62] that both can inhibit cell transformation and cancer onset [66].

2.5. Apoptosis

The most studied p53 function is the induction of apoptosis. The first evidence came from a seminal work of Oren’s lab, in which p53 reintroduction in myeloid leukemia cells-deficient for p53 promoted apoptosis [67,68]. Studies using transgenic and knockout mice demonstrated that endogenous p53 induced apoptosis, and thus counteracted tumorigenesis [69]. p53 protein acts as an apoptosis “facilitator” or “activator” [70] interacting through its DBD domain with anti-apoptotic (Bcl-2, Bcl-xL, Bcl-xW, and Mcl-1) [71] or with pro-apoptotic effectors (Bax, Bak, Bid, Bim, Bik, BMf, Bif-1, Bad, Noxa, and Hrk) inhibiting or activating them, respectively [70,72,73]. Puma protein is directly activated by p53 and can induce apoptosis binding to and inhibiting anti-apoptotic proteins [45,51,70]. p53 transcriptionally activates the pro-apoptotic genes BAX [74], PUMA [51] and FAS-R promoting apoptosis [45]. Moreover, p53 transcribes IGF-BP3 coding Igf-bp3 protein that can bind to the growth factor IGF-1 and inhibits proliferation stimuli transduced by IGF-R receptors; p53 can also interact with WWOX inducing apoptosis synergistically [75]. Without growth factor stimuli, apoptosis pathway is activated [76]. Treatments with transcription or translation inhibitors such as Actinomycin D and Cycloeximide can induce p53-dependent apoptosis [2,77].

2.6. Metabolic Regulation

p53 plays a role in the regulation of cell metabolism. It has been shown that p53 regulates glucose uptake and glycolysis reducing the expression of glucose transporters (GLUT1, GLUT3, and GLUT4), phosphoglycerate mutase (PGM), and fructose-2,6-bisphosphate levels [78,79]; moreover, it induces the transcription of TIGAR and GLS2 target genes [65,80,81] in both tumor and normal cell lines. It is also reported that p53 could induce some branches of the glycolytic pathway [82,83] and could regulate mitochondria metabolism [84,85,86,87,88]. In response to oxidative stress or DNA damage (induced by campothecin or danunorubicin treatments), p53 becomes activated and binds to the GLS2 promoter inducing its transcription. This activation leads to increased glutamine metabolism and reduced ROS production [81]. It was proposed that Gls2 preserves total GSH levels with consequent reduction of oxidative stress and so promoting cell survival after mild genotoxic stress. Moreover, Gls2 exerts a tumor suppressor function representing a link between amino acid metabolism, ROS scavenging, and inhibition of tumor cells proliferation [81]. It is worthwhile to mention that p53 can conversely increase ROS production in response to severe stress conditions and thereby promote apoptosis [81]. Tigar protein encodes for a fructose bisphosphatase that down regulates cellular levels of fructose-2,6,-bisphosphate impairing the glycolytic pathway and leading to increased NADPH formation. High NADPH levels increase the GSH amount with consequent ROS detoxification [80]. p53 also regulates nucleic acid biosynthesis inactivating the glucose-6-phosphate dehydrogenase (G6PD) that is the rate-limiting enzyme of the pentose phosphate pathway (PPP) [89], and the lipid metabolism transcriptionally controlling many targets genes [90]. In detail, p53 can inhibit transcriptionally or by physical interaction some enzymes that promote lipid biosynthesis (SREBP-1 and G6PD) [89,91] or transcriptionally activate enzymes that inhibit lipogenesis promoting fatty acid oxidative metabolism (SIRT1, Aromatase, Acad11, Lipin1, MCD, and DHRS3) or also activate proteins that regulate cholesterol efflux (Caveolin 1) [92,93,94,95,96,97,98,99]. p53 family members p63 and p73 are also involved in cell metabolism supporting p53 tumor suppressor function [100].
p53 metabolic regulation appears crucial in the inhibition of cell transformation and tumorigenesis considering that tumor cells change their metabolism to support their rapid growth. In particular, tumor cells exhibit an increased glucose uptake, glycolysis, lactate production, glutamine consumption, nucleic acid metabolism, and lipid and cholesterol synthesis [90,101,102].

2.7. Cell Migration and Angiogenesis Inhibition

Cell migration is required for the acquisition of migratory and metastatic behavior in cancer cells. p53 protein regulates cellular expansion, cell polarity formation, and cell protrusion formation, which are required for cell motility [103]. The most critical event in tumorigenesis is the conversion from a primary tumor to an invasive and metastatic cancer. This transformation requires actin cytoskeleton remodeling with consequent changing in cell shape, and interactions among cells and extracellular matrix (ECM). Actin rich protrusions favor the interactions between cells and ECM and are regulated by Rho family members Rho, Rac, and Cdc42 [104]. Rho family members regulate epithelial to mesenchymal transition (EMT) [105] and a deregulation of these factors is involved in tumorigenesis and invasiveness [106,107,108,109,110]. p53 protein modulates the transcription of different factors involved in the regulation of cell motility and cell morphology: α-actin, collagen type IIa1 and VIa1, keratin, Msp, PAI1, HGF, VEGF, MMP-1, and fibronectin [111,112,113,114,115]. Moreover, p53 can interact with tubulin, vimentin, and F-actin suggesting a possible involvement in the regulation of microtubules, intermediate filaments, and microfilament formation [103]. p53 inhibits Rac, RhoA, and Cdc42 activation blocking filopodia formation, fibronectin formation, cellular expansion, and polarization essentials in promoting cell migration and invasion [116,117,118,119,120,121]. Moreover, p53 blocks Slug/Snail and EpCAM functions increasing E-cadherin levels, activates the tumor suppressor phosphatase PTEN [122,123], activates RhoGAPs and inhibits RhoGEFs with consequent inactivation of RhoA and ROCK kinase. Rho is also inhibited by Notch signaling activated by p53 [124].
p53 also exerts a key role in the negative regulation of angiogenesis activating a potent inhibitor of this process (TSP-1) and inhibiting the collagen α(II)PH protein involved in the growth of primary human endothelial cells [125,126].
As seen in all the functions previously described, p53 deregulation is primarily involved in cancer progressions instead of early tumorigenesis [127]. Indeed, loss of p53 function contributes to chemoresistance, increased invasiveness, and metastasis formation [128,129,130].

2.8. Mutant p53 and Cancer

The TP53 coding gene is frequently mutated in human cancers (50–70% of cases) [131,132,133,134,135]. Most of these mutations are of missense type [135,136,137,138]. Missense mutations primarily affect the DBD domain (75% of cases) between exons 5–8 [6,139] and are roughly distinct in DNA-contact defective mutants (affecting the region involved in the binding to the DNA (i.e., p53-R275H and p53-R248W)) or in conformational mutants (affecting the tridimensional structure of the protein (i.e., p53-R175H and p53-R179H)) [26,132]. Most of the mutant p53 proteins do not recognize canonical consensus on the DNA sequence of wild-type p53 target genes [26,140] and exhibit an increased half-life (6–24 h) compared to that of wild-type p53 proteins. Heterozygous mutations in TP53 denote increased cancer risk (around 90–95%) frequently in early ages [3]. TP53 mutations are dominant negative and in heterozygous we can observe tetramers composed by native and mutant subunits that inactivate native ones abolishing p53 wild-type functions [6,141,142]. Mutant p53 proteins can also acquire new oncogenic functions (gain of function (GOF) mutations) promoting cell transformation, tumor progression, metastasis, and chemoresistance [5,6,7,143,144,145,146]. Moreover, GOF mutations hamper p53 proteasome-dependent degradation, altering the ordinary p53 turnover and promoting mutant p53 accumulation [141]. GOF of mutant p53 is due to different transcriptional activity caused by the recognition of new DNA secondary structures (i.e., MARs regions) or by the interaction with new protein partners and transcription factors [147,148,149]. GOF of mutant p53 proteins results also in the interaction with well-known tumor suppressor proteins (involved in the inhibition of cell proliferation, in apoptosis induction and differentiation) that are sequestered and inactivated (i.e., p73 and p63) [150,151,152,153,154,155]. It is known that mutant p53 binds with more affinity to p73 and p63 compared to the wild-type form and inhibits the P21WAF1/CIP1 activation [150,151,156]. Knock-in mice with p53 missense mutations elegantly provided the proof that some mutant p53 proteins exert pro-tumorigenic activities [157].

2.8.1. Mutant p53 Oncogenic Targets

Mutant p53 proteins show two critical residues in the N-terminal domain (TA), which are necessary for oncogenic function [158,159]. Different p53 mutants transcriptionally upregulate proto-oncogenes involved in nucleotide metabolism, amino acid and protein synthesis (different tRNA synthetases, ASNS, and EEF1A2), in cell metabolism (IMPDH2, PHGDH, ALDH6A1, and AR), in cell cycle progression (CCNB2, CDC25A, cMyc, and MCL1), in oncogenic transformation, cell migration and invasion (RhoGAPs, RhoGEFs, ANGPT1, ITGA6, Zyxin 2, Fibronectin, and hEGR1), in DNA replication and transcription (POLD2 and RNA pol II E) and in anti-apoptotic signaling (API5, E2F-5, and c-Fos). All these mutant p53-dependend upregulated genes promote tumor transformation and progression in different cancer cell lines [160,161,162,163,164,165]. p53 GOF mutants are also involved in lipid metabolism increasing lipogenesis and cholesterol and fatty acid biosynthesis by the activation of SREBPs proteins [166]. It is still unknown if there is a specific DNA-binding consensus for mutant p53 (despite certain characterization of DNA consensus for wild-type p53), but it is clear that p53 mutants can bind to different transcription factors recruiting large protein complexes containing co-activators (SWI/SNF) or histones modifiers (KMT2A, KMT2D, and KAT6A), which cooperate with mutant p53 to activate or repress target genes [167,168].
Beside metabolic regulation that favors tumor transformation, p53 mutants can reinforce anti-apoptotic signaling downregulating the MSP/MST-1 gene [169], or promote inflammation through the upregulation of interleukin 6 (IL-6) binding to and activating the transcription factor C/EBP β [170]. Moreover, mutant p53 promotes a pro-survival signaling inducing HSP70, EGFR, hsMAD1, PCNA, and NF-κB transcription and activation [171,172,173,174,175,176,177]. As mentioned before, mutant p53 activates target genes by binding to non-canonical partners that, in this specific case, are transcriptional factors that confer specificity to mutant p53 transcriptional activity. On that note, it is known that mutant p53 can interact with NF-Y transcription factor recruiting p300 acetyltransferase on the promoters of NF-Y target genes (CCNA, CCNB1, CDK1, and CDC25C) resulting in increased cell proliferation [178]. Moreover, interacting with Ets-1 transcription factor, mutant p53 activates the MDR1 gene promoting chemoresistance [148]; alternatively, by binding to E2F-1 transcription factor, p53 mutants can activate ID2 and ID4 transcription promoting tumor neo-angiogenesis [179,180]. Binding to the E2F-4 transcription factor, mutant p53 represses BRCA1 and RAD17 genes impairing the DNA repair mechanism and promoting cancer genomic instability [181]. Mutant p53 is also recruited, together with p300, at the promoter of the proteasome activator REGγ, which promotes p53 wild-type, p21, and p16 protein degradation [182] and inhibits the expression of KLF17 promoting cancer progression in invasive breast cancer cells [183]. Increasing evidence highlight the correlation between mutant p53 and increased cell metabolism associated with cancer progression. Freed-Pastor et al., demonstrated that mutant p53 protein cooperates with SREBPs proteins involved in fatty acid and sterol biosynthetic pathways; thereby leading to the upregulation of the mevalonate pathway that promotes survival and invasion of breast cancer cell lines [166]. In pancreas and breast cancer cell lines, mutant p53 can inhibit autophagy by the downregulation of BECN1, DRAM1, ATG12, SESN1/2, and P-AMPK and induce the activation of mTOR signaling promoting survival and cell proliferation [184]. It was also documented a mutant p53-dependent modulation of non-coding microRNAs (miRNAs) that contribute to cancer onset and progression [185,186,187,188].

2.8.2. Mutant p53 Partners

Mutant p53 proteins are abundantly present in tumor specimens and cancer cells due to increased protein stability and nuclear localization [157,189]. They can interact with different partners determining the generation of large multi-protein floating complexes with oncogenic activities. Well-known mutant p53 partners are the p63 and p73 protein family members, which usually function as tumor suppressor genes cooperating with wild-type p53 in senescence, aging, and apoptosis induction [190,191,192]. Various mutant p53 proteins bind to p73 and p63 with high affinity compared to the wild-type form and this binding results in p73 and p63 sequestration and inactivation; thereby impairing senescence and apoptosis, and inducing chemoresistance in different cell lines [150,151,153,155,193,194,195,196,197]. Due to its high binding affinity to p73, mutant p53 disables the inhibitory complex p73/NF-Y promoting the oncogenic expression of PDGFRb in pancreatic cancer metastasis [198]. Similarly, mutant p53 sequesters p63 promoting TGFβ-induced metastasis [199]. Mutant p53 proteins not only transcriptionally activate proteins involved in lipids biogenesis, but also can interact with AMPK, inhibiting its kinase activity. How AMPK inhibition could contribute to increased lipid production, and thus tumor progression is still unclear [200]. It was reported that mutant p53, as an additional GOF activity, can interact with the nuclear matrix and MAR-DNA elements, generating chromatin domains that may enhance or repress transcription, and so perturb nuclear structure and function [147,149,201,202]. Valenti et al., showed that Plk2 protein binds to and phosphorylates mutant p53 leading to a more efficient recruitment of p300 on mutant p53 target genes with consequent enhanced cell proliferation and chemoresistance [203]. Interestingly, VDR physically and functionally interacts with mutant p53 and is able to convert vitamin D into an anti-apoptotic agent [204]. The prolyl-isomerase Pin1 amplifies mutant p53 oncogenic function co-activating a pro-aggressiveness transcriptional program in breast cancer [205]. Recent evidence demonstrate that two Bcl2 family proteins, BAG2 and BAG5, interact with mutant p53 preventing Mdm2 and CHIP-dependent proteasome degradation of mutant p53, this promotes mutant p53 protein accumulation and GOF determining tumor growth, cell migration, and chemoresistance [206,207].

3. The Hippo Pathway

Questions about developmental processes of multicellular organisms led to the discovery of four tumor suppressor genes in Drosophila melanogaster. Loss of function mutation in these genes results in overgrowth and reduced cell death [208,209,210]. These genes encode for the core components (also known as the core kinase cassette) of the Salvador-Warts-Hippo pathway and orthologous were discovered also in mammals. The Hippo pathway is evolutionary conserved and is involved in a plethora of physio- and pathological conditions [211].

3.1. Hippo Pathway Regulation, Regulators and Functions

Mst1/2, the adaptor protein Sav1 (or WW45), Lats1/2 AGC family kinases, and MOBKL1A/B (or Mob1) compose the mammalian Hippo kinase cassette [210,212]. Mst1/2 activate Lats1/2 by phosphorylation; Mob1 is necessary for both Lats1/2 and Mast1/2 activation and Sav1 is necessary for the Mst1/2 kinase activity [213]. This activated complex triggers a phosphorylation cascade that inhibits YAP and its paralog TAZ (or WWTR1) downstream effectors; these proteins are co-transcription factors that lack a DNA binding domain [214,215]. YAP/TAZ phosphorylation at Ser127 and Ser89, respectively, creates a 14-3-3σ binding site and consequent YAP/TAZ cytoplasmic retention and inhibition (Figure 1). Moreover, Lats1/2 can phosphorylate YAP and TAZ at Ser381 and Ser311 respectively, triggering ubiquitin-mediated proteasome degradation [216] (Figure 1). YAP exhibits two splicing variants (YAP1 and YAP2) displaying a different number of WW domains (one WW domain in YAP1 and two tandem WW domains in YAP2), one binding domain for the interaction with transcription factors, highly conserved PDZ-binding motifs, and an activation domain as for its paralog TAZ [215,217,218]. When activated, YAP and TAZ translocate in the nucleus and interact through the WW domain with the PPXY motifs of diverse transcription factors such as RUNX1 and RUNX2 regulating osteoblast differentiation of mesenchymal stem cells (MSCs), chondrocyte hypertrophy, endothelial cell migration and vascular invasion in bone development [218]. They also interact with Smads and TEADs/TEFs transcription factors, regulating cell proliferation, anchorage-independent growth, EMT, embryonic stem cells (ESCs) and induced pluripotent stem cells (iPSCs) [19,219,220]. YAP also interacts with β-catenin, inducing the expression of canonical Wnt target genes (SOXS2 and SNAI2) [221] (Figure 1). TAZ interacts with PPAR-γ, TTF-1, Pax3, and Tbx5 inducing cell proliferation, cell migration, invasion, ESCs pluripotency and EMT [222,223]. Some YAP and TAZ targets are CTGF, ANKRD1, CYR61, Birc2/cIAP1, Birc5, AREG, and cMyc [214,224]. YAP/TAZ can function also as transcriptional co-repressors for the tumor suppressor genes DDIT4 and TRAIL recruiting the NuRD deacetylase histones complex onto the promoters of selected genes [225] (Figure 1).
In contrast with conventional Lats1/2 tumor suppressor functions, it was recently demonstrated that, in different murine syngeneic tumor models (B16, SCC7, and 4T1), Lats1/2 exert an oncogenic function reducing tumor immunogenicity. In detail, Lats1/2 deletion induces interferon I response improving tumor vaccine efficacy; this mechanism opens possibilities in targeting Lats1/2 for cancer immunotherapy and suggests an additional function exerted by the Hippo pathway [226].
The NF2-Frmd6/1-WWC1/2 and Lin-7C/MMP5/Patj/Mpdz tumor suppressor complexes are involved in the mammalian apical-basal cell polarity (ABCP) and induce YAP/TAZ phosphorylation and inhibition [216,219]. AMOTL1/2, PTPN14, and α-catenin proteins maintain epithelial cell polarity and interact with YAP/TAZ sequestering them at the apical membrane and inhibiting their nuclear localization [227,228,229] (Figure 1). Several other proteins involved in ABCP integrity modulate the Hippo pathway [230]. In mechanical stretch conditions, YAP/TAZ activity is increased and is repressed when cells are compressed [10,231]. YAP/TAZ activation is also regulated by cell–cell contact abundance: high cell density induces YAP/TAZ phosphorylation and cytoplasmic retention [219]. It has been shown that GPCRs activate different heterotrimeric G proteins, which can activate or repress YAP/TAZ [232] (Figure 1). In response to stress-induced apoptosis, Mst1/2 are activated by auto-phosphorylation and caspase-dependent cleavage resulting in YAP phosphorylation and inactivation [233].
Organ size is finely regulated during development, and the Hippo signaling pathway plays a key role in this process [234]. YAP/TAZ are expressed during mammalian early embryogenesis and are crucial for inner cell mass (ICM) segregation from the trophectoderm (TE) in blastocyst development [230,234]. The Hippo pathway defines either differentiation or reprogramming to a pluripotent state that is determined by the precise balance of growth factor amounts and cytoskeletal-associated cues. YAP is found in the nucleus of MSCs, ESCs, and iPSCs; its nuclear localization and protein levels decline during MSCs and ESCs differentiation. Concomitantly, the Hippo pathway activity is enhanced and YAP accumulates in the cytoplasm of differentiated cells [224]. Increasing nuclear YAP/TAZ promotes tissue regeneration, rising cell proliferation in the liver, pancreas, salivary glands, kidney, lung, heart, intestine, skin, and nervous system [235].
The Hippo network is also a key player in the regulation of cell metabolism, and conversely is regulated by metabolic factors. It was shown in zebrafish models, that YAP upregulates nucleotide biosynthesis through the induction of glutamine synthetase. This promotes tissue growth during either development or tumorigenesis [236]. Mst1 kinase can inhibit glucose uptake reducing the expression of GLUT1 transporter and lactate production; this highlights the tumor suppressor function mediated by the Hippo core kinase cassette in regulating glucose metabolism [237]. The stearoyl-CoA-desaturase 1 (SCD1) enzyme promotes fatty acid synthesis and it is demonstrated to regulate YAP and TAZ function promoting their activation and nuclear localization that is also dependent by β-catenin activation. This promotes stemness, increased cell proliferation and it is associated with a poor prognosis in lung cancer casuistries [238]. Moreover, the phosphofructokinase (PFK1) enzyme that is involved in glycolytic metabolic pathway induces YAP and TAZ activity [239]; conversely, under glucose starvation conditions, Hippo and AMPK are activated leading to YAP phosphorylation and inhibition [240]. Hippo is regulated by the mevalonate pathway, as discussed below, and by TSC-mTOR pathway with consequent autophagy induction and protein synthesis inhibition [241]. Interestingly, the activated Hippo pathway controls cell metabolism acting as tumor suppressor factor; on the other hand, YAP and TAZ oncogenes induce increased cell metabolism promoting tumorigenesis and cancer progression.

3.2. Hippo Pathway Deregulation in Cancer

Down-regulation of the core kinase cassette components or activators of the Hippo pathway results in uncontrolled tissue overgrowth; moreover, YAP/TAZ overexpression induces cell proliferation, inflammation, acquisition of cancer stem cell features, EMT (forming metastases), inhibition of senescence, suppressed anoikis, reduced apoptosis, and drug resistance [9,16,18,213,216,230,242,243]. Generally, deregulation of the pathway promotes metastatic osteosarcomas, brain tumor development (meningiomas, schwannomas, and acoustic neuromas), hepatocellular carcinoma (HCC), bile duct tumors, and malignant mesothelioma [244]. YAP/TAZ genes are amplified and localized preferentially in the nucleus of several tumors: lung, pancreas, esophagus, gastric, skin, colon, prostate, liver, ovarian and mammary gland carcinomas, medulloblastomas, gliomas, and oral squamous-cell carcinomas [9,16,18]. Finally, the Hippo pathway interacts with many other signaling pathways, creating a complex network in which YAP/TAZ are emerging as a critical node integrating and decoding both oncogenic and tumor suppressor inputs [9,18,245].
Recently it was demonstrated that YAP could regulate non-coding RNAs biogenesis associated with cancer. In detail, at low cell density, nuclear YAP binds to and sequesters a regulatory protein of the miRNA-processing machinery, p72. This leads to widespread miRNA suppression in cells and tumors with concomitant cMyc post-transcriptional induction [246]. In lung cancer, YAP elicits its oncogenic activity sustaining the aberrant expression of the MCM7 gene and its hosted miR-25, 93, and 106 cluster [247]. Moreover, nuclear YAP/TAZ elicits cytoplasm Dicer processing of pre-miRNAs [248]. In liver cancer, YAP/RUNX2 complex binds to the promoter of the tumor suppressor long non-coding (lncRNA) pseudogene MT1DP inhibiting its expression. Down-regulation of MT1DP enhances FoxA1 activity with consequent induction of the oncogenic factor AFP, a classic liver cancer biomarker [249]. It is also reported that YAP up-regulates lncRNA MALAT1 expression at both transcriptional and post-transcriptional level facilitating proliferation and enhancing cell migration [250]. Another two lncRNAs have been investigated in colon and renal cancers. YAP cooperates with β-catenin in colon tumorigenesis, also activating the transcription of the lncRNA RMRP involved in ribosomal RNA processing [251]. The LncARSR and YAP axis forms a feed-forward loop in renal cancer. Forced expression of lncARSR expression enhances tumor renal initiating cells [252]. LncARSR, upregulated in renal tumor specimens is associated with a poor prognosis of renal cell carcinomas (RCCs).

4. TP53 Status Impacts on the Hippo Pathway Activities

4.1. Wild Type p53 Protein and the Hippo Kinase Cassette in Tumor Suppression

It is known that the Hippo pathway and p53 act as tumor suppressors cooperating to induce senescence and apoptosis. In particular, for the Hippo pathway, this is mediated by the canonical function of inhibiting YAP and TAZ oncogenic activation, as previously described, and by additional non-canonical activities in which YAP functions as a tumor suppressor in response to stress conditions.
Bai et al., demonstrate that HepG2 cells treated with diverse chemotherapeutics bear increased YAP nuclear accumulation that contributes to chemosensitivity [253]. In detail, nuclear YAP induces p21, Bax and Caspase 3 expression and inhibits the anti-apoptotic factors Bcl-2 and Bcl-xL. This results in cell cycle arrest and apoptosis induction that is p53-dependent as suggested by the binding of YAP to the p53 promoter [253]. YAP induces p53 transcription that in turn can bind to the promoter of YAP and activate YAP transcription in a positive feedback loop [253] (Figure 2). In this way, YAP and p53 sustain each other to induce apoptosis and chemosensitivity in hepatocellular carcinoma cells. Mitotic spindle checkpoint activation induces Lats2 activation and binding to Mdm2 protein with consequent p53 activation and G1/S arrest induction. In turn, p53 rapidly and selectively up-regulates Lats2 expression defining a positive feedback loop (Figure 2). The Lats2–Mdm2–p53 axis thus constitutes a novel spindle checkpoint pathway critical for the maintenance of proper chromosome number [254]. Lats2 is also demonstrated to explain a key role in the differentiation of ESCs in a p53-dependent manner [255]. In response to oncogenic stress, Lats2 phosphorylates the tumor suppressor protein ASPP1 and drives its translocation into the nucleus. Lats2 and ASPP1 lead p53 to pro-apoptotic promoters inducing death of polyploid cells [256] (Figure 2). Matallanas et al., demonstrated that the tumor suppressor proteins Rassf1-5 can interact to and activate Mst2 and Lats1 inducing cell cycle arrest and Fas-induced apoptosis [257]. In detail, in a Rassf1a overexpressing condition or activation of the death receptor by Fas ligand or in DNA damage conditions, ATM kinases activate Rassf1-5 proteins that in turns activate Lats1; this phosphorylates YAP protein in a different residue inducing nuclear localization and interaction with the p53 family member p73 [257] (Figure 2). YAP can stabilize p73 by preventing nuclear export and degradation, and together with p73 induces the transcription of pro-apoptotic genes such as PUMA and p53AIP1 [257]. YAP functions as the transcriptional regulator of p73-mediated apoptosis, but is negatively regulated by a proto-oncogene Akt. Therefore, Akt phosphorylation of YAP Ser127, sequesters it in the cytoplasm and attenuates p73 apoptosis [258,259]. In response to DNA damage, YAP imparts transcriptional target specificity to p73, favouring its recruitment onto apoptotic, instead of growth suppression genes. The tumor suppressor PML gene is a target of the transcriptional competent complex YAP/p73; this leads to an auto-regulatory feedback loop that stabilizes YAP and maximizes p73’s pro-apoptotic activity [156,260,261,262] (Figure 2). Moreover, under DNA damage conditions, c-Abl kinase directly phosphorylates YAP at Tyr357, increasing protein stability and affinity for p73 [263] (Figure 2). YAP is also involved in the establishment of senescence through the activation of PML and p53 tumor suppressor proteins; in detail, loss of Werner-induced senescence activates ATM, which phosphorylates YAP and triggers the instauration of a transcriptionally active YAP/PML/p53 axis [12,264].

4.2. GOF Mutant p53 Protein, YAP and TAZ in Tumorigenesis

As described above, the cooperation between the wild-type p53 protein and the Hippo components can induce senescence, differentiation and apoptosis contrasting tumor transformation and progression. Interestingly, this scenario can be subverted when cells and tumors harbor mutant p53 protein with GOF activity. Indeed, it was recently documented that mutant p53 and YAP share a common transcriptional program showing a significant overlapping with gene signatures primarily involved in cell cycle regulation. Moreover, different GOF mutant p53 proteins (p53R280K, p53R175H, p53A193T, p53R248L, p53R273H, p53L194F, and p53P309S), but not wild-type p53 protein, can physically interact with YAP [265]. The protein complex YAP/mutant p53 can form with the transcription factor NF-Y a large multi-protein complex that is recruited onto the regulatory regions of CCNA, CCNB, and CDK1 genes. This enhances gene transcription and leads to increased cell proliferation [265] (Figure 2). It was previously demonstrated that statins delocalize YAP in the cytoplasm thereby severely impairing its oncogenic activity [11]. In agreement with these findings, statins induce YAP cytoplasmic retention and consequently reduces the recruitment of mutant p53/NF-Y protein complex on the promoters of cell cycle regulated genes [265]. Since TAZ doesn’t appear to be involved in such a transcriptional network, at least in the reported experimental conditions, YAP may serve as a critical transducer of GOF mutant p53 proteins [265]. A more recent work highlights an oncogenic crosstalk involving mutant p53 proteins (harboring R175H and R273H missense mutations) and YAP/TAZ in glioma and breast cancer stem cells [266]. Mutant p53 proteins induce WASP-interacting protein (WIP) and promote the expression of the CSC-like markers CD133, CD44, YAP, and TAZ [266] (Figure 2). In detail, Akt2 whose activity is sustained by mutant p53/p63 oncogenic complex phosphorylates WIP. WIP phosphorylation induces YAP and TAZ activation and translocation into the nucleus with consequent activation of the YAP/TAZ oncogenic targets [266]. This work indicates that mutant p53 proteins can also cooperate with the TAZ protein [266]. Excess of cholesterol is associated with mammary tumor growth and represents an independent risk factor for breast cancer and for decreased response to endocrine therapies [267]. This is because some cholesterol metabolites induce proliferation of ER-positive breast cancer cells [268]. SREBP proteins are involved in cholesterol (mevalonate pathway), fatty acids, triglyceride, and phospholipid synthesis [269] and are frequently downregulated or overexpressed by tumor suppressor genes or oncogenes, respectively [252,267]. p53 and Lats2 cooperate to transcriptionally repress the expression of SREBPs (Figure 2); instead Lats2 downregulation induces SREBPs activation and overexpression of SREBP’s targets genes [252]. It was reported that Lats2 is crucial for cholesterol synthesis balance: Lats2 dysregulation is observed in some cases of nonalcoholic fatty liver disease (NAFLD). Lats2 restricts SBEBPs activity that is important in liver physiology and pathology. [252] This Lats2 activity is YAP-independent. Lats2 is necessary for p53 activation by cholesterol excess suggesting that the activation of the Lats2-p53 axis exerts a protective role in liver homeostasis [252,267]. On the other hand, when mutated, p53 increases SREBPs protein levels and transcription of their target genes (Figure 2). In this way, mutant p53 GOF increases the mevalonate pathway disrupting cell morphology and driving malignant phenotypes such as invasion [166]. Sorrentino et al., demonstrated that the mevalonate pathway, sustained by mutant p53 SREBPs activation increases YAP/TAZ oncogenic activity inducing cell proliferation and self-renewal in breast cancer cells [11] (Figure 2); moreover, YAP itself induces the transcription of several genes involved in cholesterol metabolism [250]. Mevalonate pathway inhibition, by treatment with statins, is able to reduce the oncogenic transcriptional effects of mutant p53 in a YAP-dependent manner, suggesting that the mevalonate pathway is both crucial as an upstream and downstream regulator of mutant p53 and YAP oncogenic functions [11]. p53 and Hippo crosstalk is also observed by Aylon et al., who demonstrated that Lats knockdown changes p53’s protein interactome and conformation, converting p53 tumor suppressor transcriptional program into ones activated in cancer-associated p53 mutants (Figure 2). This results in increased cell proliferation and migration [252].
p53 mutations, that result in the acquisition of new oncogenic functions, can change the interactome, conformation, and transcriptional program of wild-type p53 transforming the anti-oncogenic cooperation between Hippo and p53 in a pro-tumorigenic crosstalk primarily through the loss of Lats-p53 mutual activation and by the binding of mutant p53 to YAP (Figure 2). YAP exerts dualistic functions in response, to developmental stages, tumor transformation and stress stimuli. Indeed, it is functionally inhibited by the Hippo kinases but also activated by them in response to stress conditions. When activated, YAP promotes PML expression, p73- and p53-dependent senescence, differentiation, and apoptosis [12,260,264,270,271]. In a mutant p53 context, YAP is no more able to induce tumor suppressor responses but is bound to mutant p53, enhancing mutant p53 oncogenic functions with consequent increased cell proliferation, invasion, and chemoresistance in cancer cells (Figure 2).

5. Conclusions and Future Perspectives

Since the crosstalk between the p53 and the Hippo tumor suppressor pathways can either elicit tumor suppressor or oncogenic effects, its therapeutic targeting might hold great potential for the treatment of human cancers. Indeed, p53 in its wild-type or mutant conformation and YAP/TAZ proteins represent suitable targets since the biological outputs of their interaction switch from pro-apoptotic activators (wild-type p53-p73-p63/YAP) to pro-tumorigenic and metastatic inducers (mutant p53/YAP/TAZ). Potentially, compounds that restore mutant p53 conformation to that of wild-type p53 protein might impair various oncogenic pathways promoted by mutant p53, also the oncogenic activity of the mutant p53/YAP protein complex and/or TAZ induction, and could reactivate the Hippo tumor suppressor function. A growing number of small molecules aimed to restore and stabilize the original DBD conformation of p53 have been identified. These include: p53 reactivation and induction of massive apoptosis (PRIMA-1), maleimide-derived molecule MIRA-1/NSC19630, short interfering mutant p53 peptides (SIMPs) (10–15 residues), reactivation of p53 and induction of tumor cell apoptosis (RITA), CP-31398 and many others [272,273,274]. These compounds showed great promise when tested in cancer cell lines, since they promote p53-induced apoptosis in tumor cells with mutant p53 proteins [275,276,277,278]. Despite its great selectivity and no side effects tested in cultured cell lines, PRIMA-1 does not reduce significantly tumor volume in in vivo experiments [277,279,280]. This result could be due to unknown potential secondary effects in in vivo mice models, to the higher time of treatment compared to the in vitro experiments and/or potential toxicity mediated by secondary products of PRIMA-1 metabolic degradation. Moreover, insulin-like receptors and others proto-oncogenes are responsible for resistance mechanisms to PRIMA-1 treatments [281]. Toxicity in actively proliferating cells is demonstrated after MIRA-1/NSC19630 treatments. The drug induces acute toxicity (within 2 h) in normal primary epithelial cells and in cancer cell lines inducing caspase-9-dependent apoptosis [282]. Cell toxicity is independent from p53 levels and mutational status; moreover, treated cells do not show resistance mechanisms in response to MIRA-1 treatments [282].
All these observations highlighted the need to develop new and more effective compounds and strategies, testing also chronic and acute potential toxicity in in vivo treatments. PRIMA-1MET (APR-246, developed by APREA), a compound very similar to PRIMA-1, but much more active at low dosage, restores the pro-apoptotic function of p53 with consequent activation of downstream target genes [275,283,284,285,286,287]. PRIMA-1MET has also successfully completed a Phase I clinical trial, showing a promising efficacy (ClinicalTrials.gov identifier: NCT00900614). Moreover, a small molecule, NSC59984 that restores wild-type p53 signaling in human tumors has been recently reported [288]. As mentioned before, statins can suppress mutant p53-dependent induction of cell proliferation through the inactivation of the YAP protein [11,265]. In a recent study, Parrales et al., demonstrated that statins can also induce degradation of conformational misfolded p53 mutants with minimal effects on wild-type p53. Thus, statins could preferentially suppress mutant p53 inhibiting cancer cell growth [289,290]. Drugs that inhibit YAP and TAZ functions (i.e., Verteporfin, LPA, thrombin, agonist for dopamine receptors, dobutamine, dasatinib, tankyrase inhibitors, and statins) have been reported; it appears that they mostly act indirectly on YAP/TAZ and exert their activity also inhibiting other targets; therefore they are not target-selective [244,291,292,293].
Further studies are necessary to elucidate mutant p53/Hippo crosstalk with the aim of designing specific combined therapies that could reactivate their pro-apoptotic function and disable their aberrant oncogenic interactions. In line with this, the elucidation of how p53 mutants and YAP interact each other, which domains are involved in this interaction, and if it is a direct or indirect interaction could be important to foster the identification of compounds aimed to destroy their cooperation. The targeting of the oncogenic mediators SREBPs and WIP proteins or the activation of Lats1/2 tumor suppression activities could be also very attractive. This might allow activating tumor suppressor pathways and impairing those with oncogenic outputs.

Acknowledgments

The support of the Italian Association for Cancer Research (AIRC) (Grant n.14455), of the Epigenomics Flagship Project (EPIGEN; sub-project 7.6), of the “Associazione Aurora Tomaselli Ricerca e Prevenzione” to Giovanni Blandino and Aboca Società Per Azioni to Sabrina Strano are greatly appreciated. We newly thank AIRC for the FIRC-AIRC Fellowship to Maria Ferraiuolo. (ID. 19371).

Author Contributions

Maria Ferraiuolo provided the structure and the writing of the article together with Lorena Verduci; Giovanni Blandino and Sabrina Strano corrected, integrated and supervised the entirely manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Acad11Acyl-CoA Dehydrogenase Family Member 11
AFPAlfa-fetoprotein
ALDH6A1Aldehyde Dehydrogenase 6
α(II)PHα(II) collagen Prolyl-4-Hydroxylase
AMOTL1/2Angiomotin-like proteins 1 and 2
AMPKAMP-activated Protein Kinase
ANGPT1Angiopoietin 1
ANKRD1Ankyrin Repeat Domain 1
API5Apoptosis Inhibitor 5
ARAldose Reductase
AREGAmphiregulin
ASNSAsparagine Synthetase
ASPP1Apoptosis-Stimulating Protein of p53 1
ATG12Autophagy-related protein 12
ATMAtaxia-Telangiectasia Mutated
ATPAdenosine Triphosphate
ATRATM and Rad3-related protein kinase
BadBcl-2-antagonist of cell death
BAG2BCL2 Associated Athanogene 2
BAG5BCL2 Associated Athanogene 5
BaxBcl-2-associated x protein
Bcl-2B-cell lymphoma 2
Bcl-xLB-cell lymphoma-extra Large
BECN1Beclin-1
Birc2/cIAP1Baculoviral Inhibitor of the apoptosis repeat-containing protein 1
Birc5Baculoviral Inhibitor of the apoptosis repeat-containing protein 5
BRCA1BReast CAncer type 1
CAMTA1Calmodulin-binding Transcription Activator 1
CCNACyclin A
CCNBCyclin B
Cdk1Cyclin dependent kinase 1
CDK1Cyclin Dependent Kinase 1
CHIPCarboxyl terminus of Hsp70-Interacting Protein
Chk1Checkpoint kinase homologue 1
Chk2Checkpoint kinase homologue 2
CK1δ/εCasein Kinase 1 delta/epsilon
cMycV-Myc Avian Myelocytomatosis Viral Oncogene Homolog
CSCsCancer Stem Cells
CTGFConnective Tissue Growth Factor
CYR61Cysteine-Rich angiogenic inducer 61
DDIT4DNA-Damage-Inducible Transcript 4
DRAM1DNA Damage Regulated Autophagy Modulator 1
EEF1A2Elongation factor 1 A-2
FoxA1Forkhead box A1
Frmd6/1FERM domain-containing protein 6 and 1
GADD45Growth Arrest and DNA Damage gene 45
GLS2Phosphate-activated mitochondrial glutaminase
GPCRsG protein-coupled receptors
H2AFXH2A histone Family member X
hEGR1Early Growth Response protein 1
HGFHepatocyte Growth Factor
HIPK2Homeodomain-Interacting Protein Kinase 2
hsMAD1human Mitotic Arrest Deficiency protein 1
ID2Inhibitor of DNA binding 2
ID4Inhibitor of DNA binding 4
IGF-1Insulin-Like Growth Factor-1
IMPDH2Inosine-5′-Monophosphate Dehydrogenase 2
ITGA6Integrin α-6
KATA6Lysine Acetyltransferase 6A
KMT2ALysine Methyltransferase 2A
KMT2DLysine Methyltransferase 2D
Lats1/2Large tumor suppressor 1 and 2
Lin-7CLin-7 homolog C
lncARSRlncRNA Activated in RCC with Sunitinib Resistance, ENST00000424980
LPALysophosphatidic Acid
MALAT1Metastasis-Associated Lung Adenocarcinoma Transcript 1
MARsMatrix Attachment Regions DNA elements
MCDMalonyl-CoA Decarboxylase
MCL1Human Myeloid Cell differentiation protein
Mdm2Mouse double minute 2 homolog
MDR1Multi Drug Reactivity 1
MMP-1Matrix MetalloProteinase-1
MMP5Membrane-associated Palmitoylated protein 5
MOBKL1A/BMps one binder kinase activator-like A and B
MpdzMultiple PDZ domain protein
MspMacrophage stimulator protein
Mst1/2Mammalian Sterile-20 family Serine-Threonine kinases 1 and 2
MT1DPMetallothionein 1D Pseudogene
NERNucleotide Excision Repair
NF2Neurofibromin 2
NF-YNuclear transcription Factor Y
p53AIP1p53 Apoptosis Independent Protein 1
PAI1Plasminogen stimulator Inhibitor 1
P-AMPKPhospho-AMP-activated Protein Kinase
PatjPals1-associated TJ protein
Pax3Paired box 3
PCAFP300/CBP-Associated Factor
PCNAProliferating Cell Nuclear Antigen
PDGFRbPlatelet Derived Growth Factor Receptor b
PHGDH3-Phosphoglycerate Dehydrogenase
Plk2Polo-Like Kinase-2
PMLProMyelocytic Leukemia
POLD2DNA Polymerase Delta subunit 2
PPAR-γPeroxisome Proliferator-Activated Receptor γ
PTENPhosphatase and tensin homolog deleted on chromosome TEN
PTPN14Tyrosine-Protein Phosphatase Non-receptor type 14
Pumap53 upregulated modulator of apoptosis
RassfRas-association domain family
RhoGAPsRho GTPase-Activating Proteins
RhoGEFsRho Guanine nucleotide Exchange Factors
RMRPRNA component of Mitochondrial RNA Processing endoribonuclease
ROCKRho-associated protein Kinase
ROSReactive Oxygen Species
RUNX1Runt-related transcription factor 1
RUNX2Runt-related transcription factor 2
Sav1Salvador 1
SESN1/2Sestrin 1/2
SIRT1Sirtuin 1
SNAI2Snail family transcriptional Inhibitor 2
SOX2Sex determining region Y-box 2
SREBPsSterol Regulatory Element Binding Proteins
TAFsTBP Associated Factors
TAZTafazzin
TBPTATA-box Binding Proteins
Tbx5T-box transcription factor 5
TCATricarboxylic Acid
TEADs/TEFsTEA Domain proteins/Transcription Enhancer Factors
TFIIHTranscription Factor II Human
TIGARTP53-inducible glycolysis and apoptosis regulator
TRAILTNF-Related Apoptosis-inducing Ligand
TSCTuberous Sclerosis Complex 2
TSP-1Thrombospondin-1
TTF-1Thyroid Transcription Factor-1
VEGFVascular Endothelial Growth Factor
WASPWiskott Aldrich Syndrome Protein
WWC1/2WW domain containing protein 1 and 2
WWOXWW domain-containing Oxidoreductase 1
YAPYes-Associated Protein

References

  1. Kress, M.; May, E.; Cassingena, R.; May, P. Simian virus 40-transformed cells express new species of proteins precipitable by anti-simian virus 40 tumor serum. J. Virol. 1979, 31, 472–483. [Google Scholar] [PubMed]
  2. Lane, D.P.; Crawford, L.V. T antigen is bound to a host protein in SV40-transformed cells. Nature 1979, 278, 261–263. [Google Scholar] [CrossRef] [PubMed]
  3. Levine, A.J. p53, the cellular gatekeeper for growth and division. Cell 1997, 88, 323–331. [Google Scholar] [CrossRef]
  4. Linzer, D.I.; Maltzman, W.; Levine, A.J. The SV40 a gene product is required for the production of a 54,000 MW cellular tumor antigen. Virology 1979, 98, 308–318. [Google Scholar] [CrossRef]
  5. Dittmer, D.; Pati, S.; Zambetti, G.; Chu, S.; Teresky, A.K.; Moore, M.; Finlay, C.; Levine, A.J. Gain of function mutations in p53. Nat. Genet. 1993, 4, 42–46. [Google Scholar] [CrossRef] [PubMed]
  6. Iwakuma, T.; Lozano, G.; Flores, E.R. Li-Fraumeni syndrome: A p53 family affair. Cell Cycle 2005, 4, 865–867. [Google Scholar] [CrossRef] [PubMed]
  7. Prives, C.; Hall, P.A. The p53 pathway. J. Pathol. 1999, 187, 112–126. [Google Scholar] [CrossRef]
  8. Strano, S.; Rossi, M.; Fontemaggi, G.; Munarriz, E.; Soddu, S.; Sacchi, A.; Blandino, G. From p63 to p53 across p73. FEBS Lett. 2001, 490, 163–170. [Google Scholar] [CrossRef]
  9. Harvey, K.F.; Zhang, X.; Thomas, D.M. The Hippo pathway and human cancer. Nat. Rev. Cancer 2013, 13, 246–257. [Google Scholar] [CrossRef] [PubMed]
  10. Piccolo, S.; Dupont, S.; Cordenonsi, M. The biology of YAP/TAZ: Hippo signaling and beyond. Physiol. Rev. 2014, 94, 1287–1312. [Google Scholar] [CrossRef] [PubMed]
  11. Sorrentino, G.; Ruggeri, N.; Specchia, V.; Cordenonsi, M.; Mano, M.; Dupont, S.; Manfrin, A.; Ingallina, E.; Sommaggio, R.; Piazza, S.; et al. Metabolic control of YAP and TAZ by the mevalonate pathway. Nat. Cell Biol. 2014, 16, 357–366. [Google Scholar] [CrossRef] [PubMed]
  12. Strano, S.; Fausti, F.; Di Agostino, S.; Sudol, M.; Blandino, G. PML surfs into Hippo tumor suppressor pathway. Front. Oncol. 2013, 3, 36. [Google Scholar] [CrossRef] [PubMed]
  13. Baena-Lopez, L.A.; Rodriguez, I.; Baonza, A. The tumor suppressor genes dachsous and fat modulate different signalling pathways by regulating dally and dally-like. Proc. Natl. Acad. Sci. USA 2008, 105, 9645–9650. [Google Scholar] [CrossRef] [PubMed]
  14. Cho, E.; Feng, Y.; Rauskolb, C.; Maitra, S.; Fehon, R.; Irvine, K.D. Delineation of a Fat tumor suppressor pathway. Nat. Genet. 2006, 38, 1142–1150. [Google Scholar] [CrossRef] [PubMed]
  15. Karpowicz, P.; Perez, J.; Perrimon, N. The Hippo tumor suppressor pathway regulates intestinal stem cell regeneration. Development 2010, 137, 4135–4145. [Google Scholar] [CrossRef] [PubMed]
  16. Pan, D. The Hippo signaling pathway in development and cancer. Dev. Cell 2010, 19, 491–505. [Google Scholar] [CrossRef] [PubMed]
  17. Park, H.W.; Guan, K.L. Regulation of the Hippo pathway and implications for anticancer drug development. Trends Pharmacol. Sci. 2013, 34, 581–589. [Google Scholar] [CrossRef] [PubMed]
  18. Zanconato, F.; Cordenonsi, M.; Piccolo, S. YAP/TAZ at the roots of cancer. Cancer Cell 2016, 29, 783–803. [Google Scholar] [CrossRef] [PubMed]
  19. Zhao, B.; Ye, X.; Yu, J.; Li, L.; Li, W.; Li, S.; Yu, J.; Lin, J.D.; Wang, C.Y.; Chinnaiyan, A.M.; et al. TEAD mediates YAP-dependent gene induction and growth control. Genes Dev. 2008, 22, 1962–1971. [Google Scholar] [CrossRef] [PubMed]
  20. Linzer, D.I.; Levine, A.J. Characterization of a 54K dalton cellular SV40 tumor antigen present in SV40-transformed cells and uninfected embryonal carcinoma cells. Cell 1979, 17, 43–52. [Google Scholar] [CrossRef]
  21. Brooks, C.L.; Gu, W. Ubiquitination, phosphorylation and acetylation: The molecular basis for p53 regulation. Curr. Opin. Cell Biol. 2003, 15, 164–171. [Google Scholar] [CrossRef]
  22. Xu, Y. Regulation of p53 responses by post-translational modifications. Cell Death Differ. 2003, 10, 400–403. [Google Scholar] [CrossRef] [PubMed]
  23. Romer, L.; Klein, C.; Dehner, A.; Kessler, H.; Buchner, J. p53—A natural cancer killer: Structural insights and therapeutic concepts. Angew. Chem. 2006, 45, 6440–6460. [Google Scholar] [CrossRef] [PubMed]
  24. Unger, T.; Nau, M.M.; Segal, S.; Minna, J.D. p53: A transdominant regulator of transcription whose function is ablated by mutations occurring in human cancer. EMBO J. 1992, 11, 1383–1390. [Google Scholar] [PubMed]
  25. Zhang, Y.; Xiong, Y. A p53 amino-terminal nuclear export signal inhibited by DNA damage-induced phosphorylation. Science 2001, 292, 1910–1915. [Google Scholar] [CrossRef] [PubMed]
  26. Cho, Y.; Gorina, S.; Jeffrey, P.D.; Pavletich, N.P. Crystal structure of a p53 tumor suppressor-DNA complex: Understanding tumorigenic mutations. Science 1994, 265, 346–355. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, X.W.; Harris, C.C. Tp53 tumour suppressor gene: Clues to molecular carcinogenesis and cancer therapy. Cancer Surv. 1996, 28, 169–196. [Google Scholar] [PubMed]
  28. Stommel, J.M.; Marchenko, N.D.; Jimenez, G.S.; Moll, U.M.; Hope, T.J.; Wahl, G.M. A leucine-rich nuclear export signal in the p53 tetramerization domain: Regulation of subcellular localization and p53 activity by NES masking. EMBO J. 1999, 18, 1660–1672. [Google Scholar] [CrossRef] [PubMed]
  29. Thut, C.J.; Chen, J.L.; Klemm, R.; Tjian, R. p53 transcriptional activation mediated by coactivators TAFII40 and TAFII60. Science 1995, 267, 100–104. [Google Scholar] [CrossRef] [PubMed]
  30. Bourdon, J.C.; Deguin-Chambon, V.; Lelong, J.C.; Dessen, P.; May, P.; Debuire, B.; May, E. Further characterisation of the p53 responsive element—Identification of new candidate genes for trans-activation by p53. Oncogene 1997, 14, 85–94. [Google Scholar] [CrossRef] [PubMed]
  31. El-Deiry, W.S.; Kern, S.E.; Pietenpol, J.A.; Kinzler, K.W.; Vogelstein, B. Definition of a consensus binding site for p53. Nat. Genet. 1992, 1, 45–49. [Google Scholar] [CrossRef] [PubMed]
  32. Mack, D.H.; Vartikar, J.; Pipas, J.M.; Laimins, L.A. Specific repression of TATA-mediated but not initiator-mediated transcription by wild-type p53. Nature 1993, 363, 281–283. [Google Scholar] [CrossRef] [PubMed]
  33. Seto, E.; Usheva, A.; Zambetti, G.P.; Momand, J.; Horikoshi, N.; Weinmann, R.; Levine, A.J.; Shenk, T. Wild-type p53 binds to the TATA-binding protein and represses transcription. Proc. Natl. Acad. Sci. USA 1992, 89, 12028–12032. [Google Scholar] [CrossRef] [PubMed]
  34. Adimoolam, S.; Ford, J.M. p53 and regulation of DNA damage recognition during nucleotide excision repair. Dna Repair 2003, 2, 947–954. [Google Scholar] [CrossRef]
  35. Zhou, J.X.; Niehans, G.A.; Shar, A.; Rubins, J.B.; Frizelle, S.P.; Kratzke, R.A. Mechanisms of G1 checkpoint loss in resected early stage non-small cell lung cancer. Lung Cancer 2001, 32, 27–38. [Google Scholar] [CrossRef]
  36. El-Deiry, W.S.; Tokino, T.; Velculescu, V.E.; Levy, D.B.; Parsons, R.; Trent, J.M.; Lin, D.; Mercer, W.E.; Kinzler, K.W.; Vogelstein, B. WAF1, a potential mediator of p53 tumor suppression. Cell 1993, 75, 817–825. [Google Scholar] [CrossRef]
  37. Harper, J.W.; Adami, G.R.; Wei, N.; Keyomarsi, K.; Elledge, S.J. The p21 Cdk-interacting protein Cip1 is a potent inhibitor of G1 cyclin-dependent kinases. Cell 1993, 75, 805–816. [Google Scholar] [CrossRef]
  38. Xiong, Y.; Hannon, G.J.; Zhang, H.; Casso, D.; Kobayashi, R.; Beach, D. p21 is a universal inhibitor of cyclin kinases. Nature 1993, 366, 701–704. [Google Scholar] [CrossRef] [PubMed]
  39. Jung-Hynes, B.; Ahmad, N. SIRT1 controls circadian clock circuitry and promotes cell survival: A connection with age-related neoplasms. FASEB J. 2009, 23, 2803–2809. [Google Scholar] [CrossRef] [PubMed]
  40. Pines, J. Cell cycle. p21 inhibits cyclin shock. Nature 1994, 369, 520–521. [Google Scholar] [CrossRef] [PubMed]
  41. Ahn, J.Y.; Schwarz, J.K.; Piwnica-Worms, H.; Canman, C.E. Threonine 68 phosphorylation by ataxia telangiectasia mutated is required for efficient activation of Chk2 in response to ionizing radiation. Cancer Res. 2000, 60, 5934–5936. [Google Scholar] [PubMed]
  42. Cortez, D.; Wang, Y.; Qin, J.; Elledge, S.J. Requirement of ATM-dependent phosphorylation of brca1 in the DNA damage response to double-strand breaks. Science 1999, 286, 1162–1166. [Google Scholar] [CrossRef] [PubMed]
  43. Fernandez-Capetillo, O.; Chen, H.T.; Celeste, A.; Ward, I.; Romanienko, P.J.; Morales, J.C.; Naka, K.; Xia, Z.; Camerini-Otero, R.D.; Motoyama, N.; et al. DNA damage-induced G2-M checkpoint activation by histone H2AX and 53BP1. Nat. Cell Biol. 2002, 4, 993–997. [Google Scholar] [CrossRef] [PubMed]
  44. Matsuoka, S.; Huang, M.; Elledge, S.J. Linkage of ATM to cell cycle regulation by the Chk2 protein kinase. Science 1998, 282, 1893–1897. [Google Scholar] [CrossRef] [PubMed]
  45. Roos, W.P.; Kaina, B. DNA damage-induced cell death by apoptosis. Trends Mol. Med. 2006, 12, 440–450. [Google Scholar] [CrossRef] [PubMed]
  46. Hermeking, H.; Lengauer, C.; Polyak, K.; He, T.C.; Zhang, L.; Thiagalingam, S.; Kinzler, K.W.; Vogelstein, B. 14-3-3 sigma is a p53-regulated inhibitor of G2/M progression. Mol. Cell 1997, 1, 3–11. [Google Scholar] [CrossRef]
  47. Cross, S.M.; Sanchez, C.A.; Morgan, C.A.; Schimke, M.K.; Ramel, S.; Idzerda, R.L.; Raskind, W.H.; Reid, B.J. A p53-dependent mouse spindle checkpoint. Science 1995, 267, 1353–1356. [Google Scholar] [CrossRef] [PubMed]
  48. Fukasawa, K.; Choi, T.; Kuriyama, R.; Rulong, S.; Vande Woude, G.F. Abnormal centrosome amplification in the absence of p53. Science 1996, 271, 1744–1747. [Google Scholar] [CrossRef] [PubMed]
  49. Del Sal, G.; Ruaro, E.M.; Utrera, R.; Cole, C.N.; Levine, A.J.; Schneider, C. Gas1-induced growth suppression requires a transactivation-independent p53 function. Mol. Cell. Biol. 1995, 15, 7152–7160. [Google Scholar] [CrossRef] [PubMed]
  50. Lakin, N.D.; Jackson, S.P. Regulation of p53 in response to DNA damage. Oncogene 1999, 18, 7644–7655. [Google Scholar] [CrossRef] [PubMed]
  51. Nakano, K.; Vousden, K.H. PUMA, a novel proapoptotic gene, is induced by p53. Mol. Cell 2001, 7, 683–694. [Google Scholar] [CrossRef]
  52. Oda, K.; Arakawa, H.; Tanaka, T.; Matsuda, K.; Tanikawa, C.; Mori, T.; Nishimori, H.; Tamai, K.; Tokino, T.; Nakamura, Y.; et al. p53AIP1, a potential mediator of p53-dependent apoptosis, and its regulation by Ser-46-phosphorylated p53. Cell 2000, 102, 849–862. [Google Scholar] [CrossRef]
  53. Mummenbrauer, T.; Janus, F.; Muller, B.; Wiesmuller, L.; Deppert, W.; Grosse, F. p53 Protein exhibits 3′-to-5′ exonuclease activity. Cell 1996, 85, 1089–1099. [Google Scholar] [CrossRef]
  54. Morris, S.M. A role for p53 in the frequency and mechanism of mutation. Mutat. Res. 2002, 511, 45–62. [Google Scholar] [CrossRef]
  55. Hainaut, P.; Wiman, K.G. p53, Cell Cycle Arrest and Apoptosis. In 25 Years Of P53 Research; Springer: Berlin, Germany, 2005; pp. 141–163. [Google Scholar]
  56. Smith, M.L.; Chen, I.T.; Zhan, Q.; Bae, I.; Chen, C.Y.; Gilmer, T.M.; Kastan, M.B.; O’Connor, P.M.; Fornace, A.J., Jr. Interaction of the p53-regulated protein GADD45 with proliferating cell nuclear antigen. Science 1994, 266, 1376–1380. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, Z.; Buratowski, S.; Svejstrup, J.Q.; Feaver, W.J.; Wu, X.; Kornberg, R.D.; Donahue, T.F.; Friedberg, E.C. The yeast TFB1 and SSL1 genes, which encode subunits of transcription factor IIH, are required for nucleotide excision repair and RNA polymerase II transcription. Mol. Cell. Biol. 1995, 15, 2288–2293. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, Z.; Svejstrup, J.Q.; Feaver, W.J.; Wu, X.; Kornberg, R.D.; Friedberg, E.C. Transcription factor b (TFIIH) is required during nucleotide-excision repair in yeast. Nature 1994, 368, 74–76. [Google Scholar] [CrossRef] [PubMed]
  59. Flores, E.R.; Lozano, G. The p53 family grows old. Genes Dev. 2012, 26, 1997–2000. [Google Scholar] [CrossRef] [PubMed]
  60. Tyner, S.D.; Venkatachalam, S.; Choi, J.; Jones, S.; Ghebranious, N.; Igelmann, H.; Lu, X.; Soron, G.; Cooper, B.; Brayton, C.; et al. p53 mutant mice that display early ageing-associated phenotypes. Nature 2002, 415, 45–53. [Google Scholar] [CrossRef] [PubMed]
  61. Donehower, L.A. Does p53 affect organismal aging? J. Cell. Physiol. 2002, 192, 23–33. [Google Scholar] [CrossRef] [PubMed]
  62. Campisi, J. Cancer and ageing: Rival demons? Nat. Rev. Cancer 2003, 3, 339–349. [Google Scholar] [CrossRef] [PubMed]
  63. Di Leonardo, A.; Linke, S.P.; Clarkin, K.; Wahl, G.M. DNA damage triggers a prolonged p53-dependent G1 arrest and long-term induction of Cip1 in normal human fibroblasts. Genes Dev. 1994, 8, 2540–2551. [Google Scholar] [CrossRef] [PubMed]
  64. Linke, S.P.; Harris, M.P.; Neugebauer, S.E.; Clarkin, K.C.; Shepard, H.M.; Maneval, D.C.; Wahl, G.M. p53-mediated accumulation of hypophosphorylated pRb after the G1 restriction point fails to halt cell cycle progression. Oncogene 1997, 15, 337–345. [Google Scholar] [CrossRef] [PubMed]
  65. Li, T.; Kon, N.; Jiang, L.; Tan, M.; Ludwig, T.; Zhao, Y.; Baer, R.; Gu, W. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell 2012, 149, 1269–1283. [Google Scholar] [CrossRef] [PubMed]
  66. Mathon, N.F.; Lloyd, A.C. Cell senescence and cancer. Nat. Rev. Cancer 2001, 1, 203–213. [Google Scholar] [CrossRef] [PubMed]
  67. 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]
  68. Yonish-Rouach, E.; Resnitzky, D.; Lotem, J.; Sachs, L.; Kimchi, A.; Oren, M. Wild-type p53 induces apoptosis of myeloid leukaemic cells that is inhibited by interleukin-6. Nature 1991, 352, 345–347. [Google Scholar] [CrossRef] [PubMed]
  69. Lowe, S.W.; Schmitt, E.M.; Smith, S.W.; Osborne, B.A.; Jacks, T. p53 is required for radiation-induced apoptosis in mouse thymocytes. Nature 1993, 362, 847–849. [Google Scholar] [CrossRef] [PubMed]
  70. Yee, K.S.; Vousden, K.H. Complicating the complexity of p53. Carcinogenesis 2005, 26, 1317–1322. [Google Scholar] [CrossRef] [PubMed]
  71. Mihara, M.; Erster, S.; Zaika, A.; Petrenko, O.; Chittenden, T.; Pancoska, P.; Moll, U.M. p53 has a direct apoptogenic role at the mitochondria. Mol. Cell 2003, 11, 577–590. [Google Scholar] [CrossRef]
  72. Chipuk, J.E.; Kuwana, T.; Bouchier-Hayes, L.; Droin, N.M.; Newmeyer, D.D.; Schuler, M.; Green, D.R. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science 2004, 303, 1010–1014. [Google Scholar] [CrossRef] [PubMed]
  73. Leu, J.I.; Dumont, P.; Hafey, M.; Murphy, M.E.; George, D.L. Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcl1 complex. Nat. Cell Biol. 2004, 6, 443–450. [Google Scholar] [CrossRef] [PubMed]
  74. Miyashita, T.; Reed, J.C. Tumor suppressor p53 is a direct transcriptional activator of the human bax gene. Cell 1995, 80, 293–299. [Google Scholar] [PubMed]
  75. Chang, N.S.; Doherty, J.; Ensign, A.; Schultz, L.; Hsu, L.J.; Hong, Q. WOX1 is essential for tumor necrosis factor-, UV light-, staurosporine-, and p53-mediated cell death, and its tyrosine 33-phosphorylated form binds and stabilizes serine 46-phosphorylated p53. J. Biol. Chem. 2005, 280, 43100–43108. [Google Scholar] [CrossRef] [PubMed]
  76. Buckbinder, L.; Talbott, R.; Velasco-Miguel, S.; Takenaka, I.; Faha, B.; Seizinger, B.R.; Kley, N. Induction of the growth inhibitor IGF-binding protein 3 by p53. Nature 1995, 377, 646–649. [Google Scholar] [CrossRef] [PubMed]
  77. Caelles, C.; Helmberg, A.; Karin, M. p53-dependent apoptosis in the absence of transcriptional activation of p53-target genes. Nature 1994, 370, 220–223. [Google Scholar] [CrossRef] [PubMed]
  78. Kondoh, H.; Lleonart, M.E.; Gil, J.; Wang, J.; Degan, P.; Peters, G.; Martinez, D.; Carnero, A.; Beach, D. Glycolytic enzymes can modulate cellular life span. Cancer Res. 2005, 65, 177–185. [Google Scholar] [PubMed]
  79. Schwartzenberg-Bar-Yoseph, F.; Armoni, M.; Karnieli, E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 2004, 64, 2627–2633. [Google Scholar] [CrossRef] [PubMed]
  80. Bensaad, K.; Tsuruta, A.; Selak, M.A.; Vidal, M.N.; Nakano, K.; Bartrons, R.; Gottlieb, E.; Vousden, K.H. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 2006, 126, 107–120. [Google Scholar] [CrossRef] [PubMed]
  81. Suzuki, S.; Tanaka, T.; Poyurovsky, M.V.; Nagano, H.; Mayama, T.; Ohkubo, S.; Lokshin, M.; Hosokawa, H.; Nakayama, T.; Suzuki, Y.; et al. Phosphate-activated glutaminase (GLS2), a p53-inducible regulator of glutamine metabolism and reactive oxygen species. Proc. Natl. Acad. Sci. USA 2010, 107, 7461–7466. [Google Scholar] [CrossRef] [PubMed]
  82. Mathupala, S.P.; Ko, Y.H.; Pedersen, P.L. Hexokinase II: Cancer’s double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene 2006, 25, 4777–4786. [Google Scholar] [CrossRef] [PubMed]
  83. Ruiz-Lozano, P.; Hixon, M.L.; Wagner, M.W.; Flores, A.I.; Ikawa, S.; Baldwin, A.S., Jr.; Chien, K.R.; Gualberto, A. p53 is a transcriptional activator of the muscle-specific phosphoglycerate mutase gene and contributes in vivo to the control of its cardiac expression. Cell Growth Differ. 1999, 10, 295–306. [Google Scholar] [PubMed]
  84. Bourdon, A.; Minai, L.; Serre, V.; Jais, J.P.; Sarzi, E.; Aubert, S.; Chretien, D.; de Lonlay, P.; Paquis-Flucklinger, V.; Arakawa, H.; et al. Mutation of RRM2B, encoding p53-controlled ribonucleotide reductase (p53R2), causes severe mitochondrial DNA depletion. Nat. Genet. 2007, 39, 776–780. [Google Scholar] [CrossRef] [PubMed]
  85. Kulawiec, M.; Ayyasamy, V.; Singh, K.K. p53 regulates mtDNA copy number and mitocheckpoint pathway. J. Carcinog. 2009, 8, 8. [Google Scholar] [PubMed]
  86. Lebedeva, M.A.; Eaton, J.S.; Shadel, G.S. Loss of p53 causes mitochondrial DNA depletion and altered mitochondrial reactive oxygen species homeostasis. Biochim. Biophys. Acta 2009, 1787, 328–334. [Google Scholar] [CrossRef] [PubMed]
  87. Matoba, S.; Kang, J.G.; Patino, W.D.; Wragg, A.; Boehm, M.; Gavrilova, O.; Hurley, P.J.; Bunz, F.; Hwang, P.M. p53 regulates mitochondrial respiration. Science 2006, 312, 1650–1653. [Google Scholar] [CrossRef] [PubMed]
  88. Okamura, S.; Ng, C.C.; Koyama, K.; Takei, Y.; Arakawa, H.; Monden, M.; Nakamura, Y. Identification of seven genes regulated by wild-type p53 in a colon cancer cell line carrying a well-controlled wild-type p53 expression system. Oncol. Res. 1999, 11, 281–285. [Google Scholar] [PubMed]
  89. Jiang, P.; Du, W.; Wang, X.; Mancuso, A.; Gao, X.; Wu, M.; Yang, X. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat. Cell Biol. 2011, 13, 310–316. [Google Scholar] [CrossRef] [PubMed]
  90. Parrales, A.; Iwakuma, T. p53 as a regulator of lipid metabolism in cancer. Int. J. Mol. Sci. 2016, 17, 2074. [Google Scholar] [CrossRef] [PubMed]
  91. Yahagi, N.; Shimano, H.; Matsuzaka, T.; Najima, Y.; Sekiya, M.; Nakagawa, Y.; Ide, T.; Tomita, S.; Okazaki, H.; Tamura, Y.; et al. p53 activation in adipocytes of obese mice. J. Biol. Chem. 2003, 278, 25395–25400. [Google Scholar] [CrossRef] [PubMed]
  92. Assaily, W.; Rubinger, D.A.; Wheaton, K.; Lin, Y.; Ma, W.; Xuan, W.; Brown-Endres, L.; Tsuchihara, K.; Mak, T.W.; Benchimol, S. ROS-mediated p53 induction of Lpin1 regulates fatty acid oxidation in response to nutritional stress. Mol. Cell 2011, 44, 491–501. [Google Scholar] [CrossRef] [PubMed]
  93. Bist, A.; Fielding, C.J.; Fielding, P.E. p53 regulates caveolin gene transcription, cell cholesterol, and growth by a novel mechanism. Biochemistry 2000, 39, 1966–1972. [Google Scholar] [CrossRef] [PubMed]
  94. Deisenroth, C.; Itahana, Y.; Tollini, L.; Jin, A.; Zhang, Y. p53-Inducible DHRS3 is an endoplasmic reticulum protein associated with lipid droplet accumulation. J. Biol. Chem. 2011, 286, 28343–28356. [Google Scholar] [CrossRef] [PubMed]
  95. Jiang, D.; LaGory, E.L.; Kenzelmann Broz, D.; Bieging, K.T.; Brady, C.A.; Link, N.; Abrams, J.M.; Giaccia, A.J.; Attardi, L.D. Analysis of p53 transactivation domain mutants reveals Acad11 as a metabolic target important for p53 pro-survival function. Cell Rep. 2015, 10, 1096–1109. [Google Scholar] [CrossRef] [PubMed]
  96. Kirschner, R.D.; Rother, K.; Muller, G.A.; Engeland, K. The retinal dehydrogenase/reductase retSDR1/DHRS3 gene is activated by p53 and p63 but not by mutants derived from tumors or EEC/ADULT malformation syndromes. Cell Cycle 2010, 9, 2177–2188. [Google Scholar] [CrossRef] [PubMed]
  97. Liu, Y.; He, Y.; Jin, A.; Tikunov, A.P.; Zhou, L.; Tollini, L.A.; Leslie, P.; Kim, T.H.; Li, L.O.; Coleman, R.A.; et al. Ribosomal protein-Mdm2-p53 pathway coordinates nutrient stress with lipid metabolism by regulating MCD and promoting fatty acid oxidation. Proc. Natl. Acad. Sci. USA 2014, 111, E2414–E2422. [Google Scholar] [CrossRef] [PubMed]
  98. Nemoto, S.; Fergusson, M.M.; Finkel, T. Nutrient availability regulates SIRT1 through a forkhead-dependent pathway. Science 2004, 306, 2105–2108. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, X.; Zhao, X.; Gao, X.; Mei, Y.; Wu, M. A new role of p53 in regulating lipid metabolism. J. Mol. Cell Biol. 2013, 5, 147–150. [Google Scholar] [CrossRef] [PubMed]
  100. Napoli, M.; Flores, E.R. The p53 family orchestrates the regulation of metabolism: Physiological regulation and implications for cancer therapy. Br. J. Cancer 2017, 116, 149–155. [Google Scholar] [CrossRef] [PubMed]
  101. De Berardinis, R.J.; Mancuso, A.; Daikhin, E.; Nissim, I.; Yudkoff, M.; Wehrli, S.; Thompson, C.B. Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc. Natl. Acad. Sci. USA 2007, 104, 19345–19350. [Google Scholar] [CrossRef] [PubMed]
  102. Vousden, K.H.; Ryan, K.M. p53 and metabolism. Nat. Rev. Cancer 2009, 9, 691–700. [Google Scholar] [CrossRef] [PubMed]
  103. Roger, L.; Gadea, G.; Roux, P. Control of cell migration: A tumour suppressor function for p53? Biol. Cell 2006, 98, 141–152. [Google Scholar] [CrossRef] [PubMed]
  104. Sahai, E.; Marshall, C.J. Differing modes of tumour cell invasion have distinct requirements for Rho/ROCK signalling and extracellular proteolysis. Nat. Cell Biol. 2003, 5, 711–719. [Google Scholar] [CrossRef] [PubMed]
  105. Lozano, E.; Betson, M.; Braga, V.M. Tumor progression: Small GTPases and loss of cell-cell adhesion. Bioessays 2003, 25, 452–463. [Google Scholar] [CrossRef] [PubMed]
  106. Clark, E.A.; Golub, T.R.; Lander, E.S.; Hynes, R.O. Genomic analysis of metastasis reveals an essential role for RhoC. Nature 2000, 406, 532–535. [Google Scholar] [CrossRef] [PubMed]
  107. Fritz, G.; Just, I.; Kaina, B. Rho GTPases are over-expressed in human tumors. Int. J. Cancer 1999, 81, 682–687. [Google Scholar] [CrossRef]
  108. Horiuchi, A.; Imai, T.; Wang, C.; Ohira, S.; Feng, Y.; Nikaido, T.; Konishi, I. Up-regulation of small GTPases, RhoA and RhoC, is associated with tumor progression in ovarian carcinoma. Lab. Investig. 2003, 83, 861–870. [Google Scholar] [CrossRef] [PubMed]
  109. Itoh, K.; Yoshioka, K.; Akedo, H.; Uehata, M.; Ishizaki, T.; Narumiya, S. An essential part for Rho-associated kinase in the transcellular invasion of tumor cells. Nat. Med. 1999, 5, 221–225. [Google Scholar] [CrossRef] [PubMed]
  110. Jaffe, A.B.; Hall, A. Rho GTPases in transformation and metastasis. Adv. Cancer Res. 2002, 84, 57–80. [Google Scholar] [PubMed]
  111. Comer, K.A.; Dennis, P.A.; Armstrong, L.; Catino, J.J.; Kastan, M.B.; Kumar, C.C. Human smooth muscle α-actin gene is a transcriptional target of the p53 tumor suppressor protein. Oncogene 1998, 16, 1299–1308. [Google Scholar] [CrossRef] [PubMed]
  112. Iotsova, V.; Stehelin, D. Down-regulation of fibronectin gene expression by the p53 tumor suppressor protein. Cell Growth Differ. 1996, 7, 629–634. [Google Scholar] [PubMed]
  113. Mukhopadhyay, D.; Tsiokas, L.; Sukhatme, V.P. Wild-type p53 and v-Src exert opposing influences on human vascular endothelial growth factor gene expression. Cancer Res. 1995, 55, 6161–6165. [Google Scholar] [PubMed]
  114. Sun, Y.; Sun, Y.; Wenger, L.; Rutter, J.L.; Brinckerhoff, C.E.; Cheung, H.S. Human metalloproteinase-1 (collagenase-1) is a tumor suppressor protein p53 target gene. Ann. N. Y. Acad. Sci. 1999, 878, 638–641. [Google Scholar] [CrossRef] [PubMed]
  115. Zhao, R.; Gish, K.; Murphy, M.; Yin, Y.; Notterman, D.; Hoffman, W.H.; Tom, E.; Mack, D.H.; Levine, A.J. The transcriptional program following p53 activation. Cold Spring Harb. Symp. Quant. Biol. 2000, 65, 475–482. [Google Scholar] [CrossRef] [PubMed]
  116. Etienne-Manneville, S. Cdc42—The centre of polarity. J. Cell Sci. 2004, 117, 1291–1300. [Google Scholar] [CrossRef] [PubMed]
  117. Etienne-Manneville, S.; Hall, A. Integrin-mediated activation of Cdc42 controls cell polarity in migrating astrocytes through PKCzeta. Cell 2001, 106, 489–498. [Google Scholar] [CrossRef]
  118. Etienne-Manneville, S.; Hall, A. Cdc42 regulates GSK-3β and adenomatous polyposis coli to control cell polarity. Nature 2003, 421, 753–756. [Google Scholar] [CrossRef] [PubMed]
  119. Gadea, G.; Lapasset, L.; Gauthier-Rouviere, C.; Roux, P. Regulation of Cdc42-mediated morphological effects: A novel function for p53. EMBO J. 2002, 21, 2373–2382. [Google Scholar] [CrossRef] [PubMed]
  120. Gadea, G.; Roger, L.; Anguille, C.; de Toledo, M.; Gire, V.; Roux, P. TNFα induces sequential activation of Cdc42- and p38/p53-dependent pathways that antagonistically regulate filopodia formation. J. Cell Sci. 2004, 117, 6355–6364. [Google Scholar] [CrossRef] [PubMed]
  121. Guo, F.; Zheng, Y. Rho family GTPases cooperate with p53 deletion to promote primary mouse embryonic fibroblast cell invasion. Oncogene 2004, 23, 5577–5585. [Google Scholar] [CrossRef] [PubMed]
  122. Shiota, M.; Izumi, H.; Onitsuka, T.; Miyamoto, N.; Kashiwagi, E.; Kidani, A.; Hirano, G.; Takahashi, M.; Naito, S.; Kohno, K. Twist and p53 reciprocally regulate target genes via direct interaction. Oncogene 2008, 27, 5543–5553. [Google Scholar] [CrossRef] [PubMed]
  123. Smit, M.A.; Peeper, D.S. Deregulating EMT and senescence: Double impact by a single twist. Cancer Cell 2008, 14, 5–7. [Google Scholar] [CrossRef] [PubMed]
  124. Muller, P.A.; Vousden, K.H.; Norman, J.C. p53 and its mutants in tumor cell migration and invasion. J. Cell Biol. 2011, 192, 209–218. [Google Scholar] [CrossRef] [PubMed]
  125. Dameron, K.M.; Volpert, O.V.; Tainsky, M.A.; Bouck, N. The p53 tumor suppressor gene inhibits angiogenesis by stimulating the production of thrombospondin. Cold Spring Harb. Symp. Quant. Biol. 1994, 59, 483–489. [Google Scholar] [CrossRef] [PubMed]
  126. Teodoro, J.G.; Parker, A.E.; Zhu, X.; Green, M.R. p53-mediated inhibition of angiogenesis through up-regulation of a collagen prolyl hydroxylase. Science 2006, 313, 968–971. [Google Scholar] [CrossRef] [PubMed]
  127. Kemp, C.J.; Donehower, L.A.; Bradley, A.; Balmain, A. Reduction of p53 gene dosage does not increase initiation or promotion but enhances malignant progression of chemically induced skin tumors. Cell 1993, 74, 813–822. [Google Scholar] [CrossRef]
  128. Schmitt, C.A.; Fridman, J.S.; Yang, M.; Baranov, E.; Hoffman, R.M.; Lowe, S.W. Dissecting p53 tumor suppressor functions in vivo. Cancer Cell 2002, 1, 289–298. [Google Scholar] [CrossRef]
  129. Schmitt, C.A.; McCurrach, M.E.; de Stanchina, E.; Wallace-Brodeur, R.R.; Lowe, S.W. INK4a/ARF mutations accelerate lymphomagenesis and promote chemoresistance by disabling p53. Genes Dev. 1999, 13, 2670–2677. [Google Scholar] [CrossRef] [PubMed]
  130. Thompson, T.C.; Park, S.H.; Timme, T.L.; Ren, C.; Eastham, J.A.; Donehower, L.A.; Bradley, A.; Kadmon, D.; Yang, G. Loss of p53 function leads to metastasis in ras+myc-initiated mouse prostate cancer. Oncogene 1995, 10, 869–879. [Google Scholar] [PubMed]
  131. Hainaut, P.; Hollstein, M. p53 and human cancer: The first ten thousand mutations. Adv. Cancer Res. 2000, 77, 81–137. [Google Scholar] [PubMed]
  132. Hollstein, M.; Rice, K.; Greenblatt, M.S.; Soussi, T.; Fuchs, R.; Sorlie, T.; Hovig, E.; Smith-Sorensen, B.; Montesano, R.; Harris, C.C. Database of p53 gene somatic mutations in human tumors and cell lines. Nucleic Acids Res. 1994, 22, 3551–3555. [Google Scholar] [PubMed]
  133. Hollstein, M.; Sidransky, D.; Vogelstein, B.; Harris, C.C. p53 mutations in human cancers. Science 1991, 253, 49–53. [Google Scholar] [CrossRef] [PubMed]
  134. Lutzker, S.G.; Levine, A.J. A functionally inactive p53 protein in teratocarcinoma cells is activated by either DNA damage or cellular differentiation. Nat. Med. 1996, 2, 804–810. [Google Scholar] [CrossRef] [PubMed]
  135. Oliver, R.T.; Shamash, J.; Berney, D.M. p53 and MDM2 in germ cell cancer treatment response. J. Clin. Oncol. 2002, 20, 3928–3929. [Google Scholar] [CrossRef] [PubMed]
  136. Hainaut, P.; Soussi, T.; Shomer, B.; Hollstein, M.; Greenblatt, M.; Hovig, E.; Harris, C.C.; Montesano, R. Database of p53 gene somatic mutations in human tumors and cell lines: Updated compilation and future prospects. Nucleic Acids Res. 1997, 25, 151–157. [Google Scholar] [CrossRef] [PubMed]
  137. Hollstein, M.; Soussi, T.; Thomas, G.; von Brevern, M.C. Bartsch, P53 gene alterations in human tumors: Perspectives for cancer control. Recent Results Cancer Res. 1997, 143, 369–389. [Google Scholar] [PubMed]
  138. Hussain, S.P.; Harris, C.C. Molecular epidemiology of human cancer: Contribution of mutation spectra studies of tumor suppressor genes. Cancer Res. 1998, 58, 4023–4037. [Google Scholar] [CrossRef]
  139. Walker, D.R.; Bond, J.P.; Tarone, R.E.; Harris, C.C.; Makalowski, W.; Boguski, M.S.; Greenblatt, M.S. Evolutionary conservation and somatic mutation hotspot maps of p53: Correlation with p53 protein structural and functional features. Oncogene 1999, 18, 211–218. [Google Scholar] [CrossRef] [PubMed]
  140. Joerger, A.C.; Ang, H.C.; Veprintsev, D.B.; Blair, C.M.; Fersht, A.R. Structures of p53 cancer mutants and mechanism of rescue by second-site suppressor mutations. J. Biol. Chem. 2005, 280, 16030–16037. [Google Scholar] [CrossRef] [PubMed]
  141. Shimamura, A.; Fisher, D.E. p53 in life and death. Clin. Cancer Res. 1996, 2, 435–440. [Google Scholar] [PubMed]
  142. Sigal, A.; Rotter, V. Oncogenic mutations of the p53 tumor suppressor: The demons of the guardian of the genome. Cancer Res. 2000, 60, 6788–6793. [Google Scholar] [PubMed]
  143. Blandino, G.; Levine, A.J.; Oren, M. Mutant p53 gain of function: Differential effects of different p53 mutants on resistance of cultured cells to chemotherapy. Oncogene 1999, 18, 477–485. [Google Scholar] [CrossRef] [PubMed]
  144. Oren, M.; Rotter, V. Mutant p53 gain-of-function in cancer. Cold Spring Harb. Perspect. Biol. 2010, 2, a001107. [Google Scholar] [CrossRef] [PubMed]
  145. Walerych, D.; Lisek, K.; del Sal, G. Mutant p53: One, no one, and one hundred thousand. Front. Oncol. 2015, 5, 289. [Google Scholar] [CrossRef] [PubMed]
  146. Zambetti, G.P.; Levine, A.J. A comparison of the biological activities of wild-type and mutant p53. FASEB J. 1993, 7, 855–865. [Google Scholar] [PubMed]
  147. Gohler, T.; Jager, S.; Warnecke, G.; Yasuda, H.; Kim, E.; Deppert, W. Mutant p53 proteins bind DNA in a DNA structure-selective mode. Nucleic Acids Res. 2005, 33, 1087–1100. [Google Scholar] [CrossRef] [PubMed]
  148. 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]
  149. Will, K.; Warnecke, G.; Wiesmuller, L.; Deppert, W. Specific interaction of mutant p53 with regions of matrix attachment region DNA elements (MARs) with a high potential for base-unpairing. Proc. Natl. Acad. Sci. USA 1998, 95, 13681–13686. [Google Scholar] [CrossRef] [PubMed]
  150. 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] [PubMed]
  151. 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] [PubMed]
  152. 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]
  153. Marin, M.C.; Jost, C.A.; Brooks, L.A.; Irwin, M.S.; O’Nions, J.; Tidy, J.A.; James, N.; McGregor, J.M.; Harwood, C.A.; Yulug, I.G.; et al. A common polymorphism acts as an intragenic modifier of mutant p53 behaviour. Nat. Genet. 2000, 25, 47–54. [Google Scholar] [CrossRef] [PubMed]
  154. 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]
  155. Strano, S.; Munarriz, E.; Rossi, M.; Cristofanelli, B.; Shaul, Y.; Castagnoli, L.; Levine, A.J.; Sacchi, A.; Cesareni, G.; Oren, M.; et al. Physical and functional interaction between p53 mutants and different isoforms of p73. J. Biol. Chem. 2000, 275, 29503–29512. [Google Scholar] [CrossRef] [PubMed]
  156. Strano, S.; Munarriz, E.; Rossi, M.; Castagnoli, L.; Shaul, Y.; Sacchi, A.; Oren, M.; Sudol, M.; Cesareni, G.; Blandino, G. Physical interaction with Yes-associated protein enhances p73 transcriptional activity. J. Biol. Chem. 2001, 276, 15164–15173. [Google Scholar] [CrossRef] [PubMed]
  157. Soussi, T.; Lozano, G. p53 mutation heterogeneity in cancer. Biochem. Biophys. Res. Commun. 2005, 331, 834–842. [Google Scholar] [CrossRef] [PubMed]
  158. Lin, S.L.; Chang, D.; Ying, S.Y. Asymmetry of intronic pre-miRNA structures in functional RISC assembly. Gene 2005, 356, 32–38. [Google Scholar] [CrossRef] [PubMed]
  159. Matas, D.; Sigal, A.; Stambolsky, P.; Milyavsky, M.; Weisz, L.; Schwartz, D.; Goldfinger, N.; Rotter, V. Integrity of the N-terminal transcription domain of p53 is required for mutant p53 interference with drug-induced apoptosis. EMBO J. 2001, 20, 4163–4172. [Google Scholar] [CrossRef] [PubMed]
  160. Frazier, M.W.; He, X.; Wang, J.; Gu, Z.; Cleveland, J.L.; Zambetti, G.P. Activation of c-myc gene expression by tumor-derived p53 mutants requires a discrete C-terminal domain. Mol. Cell. Biol. 1998, 18, 3735–3743. [Google Scholar] [CrossRef] [PubMed]
  161. 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] [PubMed]
  162. Mizuarai, S.; Yamanaka, K.; Kotani, H. Mutant p53 induces the GEF-H1 oncogene, a guanine nucleotide exchange factor-H1 for RhoA, resulting in accelerated cell proliferation in tumor cells. Cancer Res. 2006, 66, 6319–6326. [Google Scholar] [CrossRef] [PubMed]
  163. O’Farrell, T.J.; Ghosh, P.; Dobashi, N.; Sasaki, C.Y.; Longo, D.L. Comparison of the effect of mutant and wild-type p53 on global gene expression. Cancer Res. 2004, 64, 8199–8207. [Google Scholar] [CrossRef] [PubMed]
  164. Scian, M.J.; Stagliano, K.E.; Ellis, M.A.; Hassan, S.; Bowman, M.; Miles, M.F.; Deb, S.P.; Deb, S. Modulation of gene expression by tumor-derived p53 mutants. Cancer Res. 2004, 64, 7447–7454. [Google Scholar] [CrossRef] [PubMed]
  165. Weisz, L.; Zalcenstein, A.; Stambolsky, P.; Cohen, Y.; Goldfinger, N.; Oren, M.; Rotter, V. Transactivation of the EGR1 gene contributes to mutant p53 gain of function. Cancer Res. 2004, 64, 8318–8327. [Google Scholar] [CrossRef] [PubMed]
  166. Freed-Pastor, W.A.; Mizuno, H.; Zhao, X.; Langerod, 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] [PubMed]
  167. Pfister, N.T.; Fomin, V.; Regunath, K.; Zhou, J.Y.; Zhou, W.; Silwal-Pandit, L.; Freed-Pastor, W.A.; Laptenko, O.; Neo, S.P.; Bargonetti, J.; et al. Mutant p53 cooperates with the SWI/SNF chromatin remodeling complex to regulate VEGFR2 in breast cancer cells. Genes Dev. 2015, 29, 1298–1315. [Google Scholar] [CrossRef] [PubMed]
  168. 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]
  169. Zalcenstein, A.; Weisz, L.; Stambolsky, P.; Bar, J.; Rotter, V.; Oren, M. Repression of the Msp/Mst-1 gene contributes to the antiapoptotic gain of function of mutant p53. Oncogene 2006, 25, 359–369. [Google Scholar] [CrossRef] [PubMed]
  170. Margulies, L.; Sehgal, P.B. Modulation of the human interleukin-6 promoter (IL-6) and transcription factor C/EBP β (NF-IL6) activity by p53 species. J. Biol. Chem. 1993, 268, 15096–15100. [Google Scholar] [PubMed]
  171. Cooks, T.; Pateras, I.S.; Tarcic, O.; Solomon, H.; Schetter, A.J.; Wilder, S.; Lozano, G.; Pikarsky, E.; Forshew, T.; Rosenfeld, N.; et al. Mutant p53 prolongs NF-κB activation and promotes chronic inflammation and inflammation-associated colorectal cancer. Cancer Cell 2013, 23, 634–646. [Google Scholar] [CrossRef] [PubMed]
  172. Deb, S.; Jackson, C.T.; Subler, M.A.; Martin, D.W. Modulation of cellular and viral promoters by mutant human p53 proteins found in tumor cells. J. Virol. 1992, 66, 6164–6170. [Google Scholar] [PubMed]
  173. Iwanaga, Y.; Jeang, K.T. Expression of mitotic spindle checkpoint protein hsMAD1 correlates with cellular proliferation and is activated by a gain-of-function p53 mutant. Cancer Res. 2002, 62, 2618–2624. [Google Scholar] [PubMed]
  174. Ludes-Meyers, J.H.; Subler, M.A.; Shivakumar, C.V.; Munoz, R.M.; Jiang, P.; Bigger, J.E.; Brown, D.R.; Deb, S.P.; Deb, S. Transcriptional activation of the human epidermal growth factor receptor promoter by human p53. Mol. Cell. Biol. 1996, 16, 6009–6019. [Google Scholar] [CrossRef] [PubMed]
  175. Scian, M.J.; Stagliano, K.E.; Anderson, M.A.; Hassan, S.; Bowman, M.; Miles, M.F.; Deb, S.P.; Deb, S. Tumor-derived p53 mutants induce NF-κB2 gene expression. Mol. Cell. Biol. 2005, 25, 10097–10110. [Google Scholar] [CrossRef] [PubMed]
  176. Tsutsumi-Ishii, Y.; Tadokoro, K.; Hanaoka, F.; Tsuchida, N. Response of heat shock element within the human HSP70 promoter to mutated p53 genes. Cell Growth Differ. 1995, 6, 1–8. [Google Scholar] [PubMed]
  177. 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]
  178. 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] [PubMed]
  179. Fontemaggi, G.; Dell’Orso, S.; Muti, P.; Blandino, G.; Strano, S. Id2 gene is a transcriptional target of the protein complex mutant p53/E2F1. Cell Cycle 2010, 9, 2464–2466. [Google Scholar] [CrossRef] [PubMed]
  180. Fontemaggi, G.; Dell’Orso, S.; Trisciuoglio, D.; Shay, T.; Melucci, E.; Fazi, F.; Terrenato, I.; Mottolese, M.; Muti, P.; Domany, E.; et al. The execution of the transcriptional axis mutant p53, E2F1 and ID4 promotes tumor neo-angiogenesis. Nat. Struct. Mol. Biol. 2009, 16, 1086–1093. [Google Scholar] [CrossRef] [PubMed]
  181. 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, 5547–5566. [Google Scholar] [CrossRef] [PubMed]
  182. 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 REGgamma-proteasome pathway by p53/TGF-β signalling and mutant p53 in cancer cells. Nat. Commun. 2013, 4, 2667. [Google Scholar] [CrossRef] [PubMed]
  183. 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] [PubMed]
  184. Cordani, M.; Oppici, E.; Dando, I.; Butturini, E.; Dalla Pozza, E.; Nadal-Serrano, M.; Oliver, J.; Roca, P.; Mariotto, S.; Cellini, B.; et al. Mutant p53 proteins counteract autophagic mechanism sensitizing cancer cells to mTOR inhibition. Mol. Oncol. 2016, 10, 1008–1029. [Google Scholar] [CrossRef] [PubMed]
  185. 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]
  186. Donzelli, S.; Strano, S.; Blandino, G. microRNAs: Short non-coding bullets of gain of function mutant p53 proteins. Oncoscience 2014, 1, 427–433. [Google Scholar] [CrossRef] [PubMed]
  187. Ganci, F.; Sacconi, A.; Bossel Ben-Moshe, N.; Manciocco, V.; Sperduti, I.; Strigari, L.; Covello, R.; Benevolo, M.; Pescarmona, E.; Domany, E.; et al. Expression of Tp53 mutation-associated microRNAs predicts clinical outcome in head and neck squamous cell carcinoma patients. Ann. Oncol. 2013, 24, 3082–3088. [Google Scholar] [CrossRef] [PubMed]
  188. 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]
  189. Strano, S.; Dell’Orso, S.; Di Agostino, S.; Fontemaggi, G.; Sacchi, A.; Blandino, G. Mutant p53: An oncogenic transcription factor. Oncogene 2007, 26, 2212–2219. [Google Scholar] [CrossRef] [PubMed]
  190. Gonfloni, S.; Caputo, V.; Iannizzotto, V. P63 in health and cancer. Int. J. Dev. Biol. 2015, 59, 87–93. [Google Scholar] [CrossRef] [PubMed]
  191. Strano, S.; Dell’Orso, S.; Mongiovi, A.M.; Monti, O.; Lapi, E.; Di Agostino, S.; Fontemaggi, G.; Blandino, G. Mutant p53 proteins: Between loss and gain of function. Head Neck 2007, 29, 488–496. [Google Scholar] [CrossRef] [PubMed]
  192. Yoon, M.K.; Ha, J.H.; Lee, M.S.; Chi, S.W. Structure and apoptotic function of p73. BMB Rep. 2015, 48, 81–90. [Google Scholar] [CrossRef] [PubMed]
  193. Ferraiuolo, M.; Di Agostino, S.; Blandino, G.; Strano, S. Oncogenic intra-p53 family member interactions in human cancers. Front. Oncol. 2016, 6, 77. [Google Scholar] [CrossRef] [PubMed]
  194. Gualberto, A.; Aldape, K.; Kozakiewicz, K.; Tlsty, T.D. An oncogenic form of p53 confers a dominant, gain-of-function phenotype that disrupts spindle checkpoint control. Proc. Natl. Acad. Sci. USA 1998, 95, 5166–5171. [Google Scholar] [CrossRef] [PubMed]
  195. Irwin, M.S.; Kondo, K.; Marin, M.C.; Cheng, L.S.; Hahn, W.C.; Kaelin, W.G., Jr. Chemosensitivity linked to p73 function. Cancer Cell 2003, 3, 403–410. [Google Scholar] [CrossRef]
  196. Strano, S.; Blandino, G. p73-mediated chemosensitivity: A preferential target of oncogenic mutant p53. Cell Cycle 2003, 2, 348–349. [Google Scholar] [CrossRef] [PubMed]
  197. Strano, S.; Fontemaggi, G.; Costanzo, A.; Rizzo, M.G.; Monti, O.; Baccarini, A.; Del Sal, G.; Levrero, M.; Sacchi, A.; Oren, M.; et al. Physical interaction with human tumor-derived p53 mutants inhibits p63 activities. J. Biol. Chem. 2002, 277, 18817–18826. [Google Scholar] [CrossRef] [PubMed]
  198. Weissmueller, S.; Manchado, E.; Saborowski, M.; Morris, J.P.T.; 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] [PubMed]
  199. Adorno, M.; Cordenonsi, M.; Montagner, M.; Dupont, S.; Wong, C.; Hann, B.; Solari, A.; Bobisse, S.; Rondina, M.B.; Guzzardo, V.; et al. A Mutant-p53/Smad complex opposes p63 to empower TGFβ-induced metastasis. Cell 2009, 137, 87–98. [Google Scholar] [CrossRef] [PubMed]
  200. 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]
  201. Kim, E.; Deppert, W. The versatile interactions of p53 with DNA: When flexibility serves specificity. Cell Death Differ. 2006, 13, 885–889. [Google Scholar] [CrossRef] [PubMed]
  202. Koga, H.; Deppert, W. Identification of genomic DNA sequences bound by mutant p53 protein (Gly245-->Ser) in vivo. Oncogene 2000, 19, 4178–4183. [Google Scholar] [CrossRef] [PubMed]
  203. Valenti, F.; Fausti, F.; Biagioni, F.; Shay, T.; Fontemaggi, G.; Domany, E.; Yaffe, M.B.; Strano, S.; Blandino, G.; Di Agostino, S. Mutant p53 oncogenic functions are sustained by Plk2 kinase through an autoregulatory feedback loop. Cell Cycle 2011, 10, 4330–4340. [Google Scholar] [CrossRef] [PubMed]
  204. Stambolsky, P.; Tabach, Y.; Fontemaggi, G.; Weisz, L.; Maor-Aloni, R.; Siegfried, Z.; Shiff, I.; Kogan, I.; Shay, M.; Kalo, E.; et al. Modulation of the vitamin D3 response by cancer-associated mutant p53. Cancer Cell 2010, 17, 273–285. [Google Scholar] [CrossRef] [PubMed]
  205. 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] [PubMed]
  206. Yue, X.; Zhao, Y.; Huang, G.; Li, J.; Zhu, J.; Feng, Z.; Hu, W. A novel mutant p53 binding partner BAG5 stabilizes mutant p53 and promotes mutant p53 GOFs in tumorigenesis. Cell Discov. 2016, 2, 16039. [Google Scholar] [CrossRef] [PubMed]
  207. Yue, X.; Zhao, Y.; Liu, J.; Zhang, C.; Yu, H.; Wang, J.; Zheng, T.; Liu, L.; Li, J.; Feng, Z.; et al. BAG2 promotes tumorigenesis through enhancing mutant p53 protein levels and function. eLife 2015, 4, e08401. [Google Scholar] [CrossRef] [PubMed]
  208. Lai, Z.C.; Wei, X.; Shimizu, T.; Ramos, E.; Rohrbaugh, M.; Nikolaidis, N.; Ho, L.L.; Li, Y. Control of cell proliferation and apoptosis by mob as tumor suppressor, mats. Cell 2005, 120, 675–685. [Google Scholar] [CrossRef] [PubMed]
  209. Staley, B.K.; Irvine, K.D. Hippo signaling in Drosophila: Recent advances and insights. Dev. Dyn. 2012, 241, 3–15. [Google Scholar] [CrossRef] [PubMed]
  210. Tapon, N.; Harvey, K.F.; Bell, D.W.; Wahrer, D.C.; Schiripo, T.A.; Haber, D.; Hariharan, I.K. salvador Promotes both cell cycle exit and apoptosis in Drosophila and is mutated in human cancer cell lines. Cell 2002, 110, 467–478. [Google Scholar] [CrossRef]
  211. Ferraiuolo, M.S.S.; Blandino, G. The Hippo Pathway. In Encyclopedia of Cell Biology; Stahl, P., Bradshaw, R., Eds.; Academic Press: Waltham, MA, USA, 2016; Volume 3, pp. 99–106. [Google Scholar]
  212. Yabuta, N.; Okada, N.; Ito, A.; Hosomi, T.; Nishihara, S.; Sasayama, Y.; Fujimori, A.; Okuzaki, D.; Zhao, H.; Ikawa, M.; et al. Lats2 is an essential mitotic regulator required for the coordination of cell division. J. Biol. Chem. 2007, 282, 19259–19271. [Google Scholar] [CrossRef] [PubMed]
  213. Yabuta, N.; Fujii, T.; Copeland, N.G.; Gilbert, D.J.; Jenkins, N.A.; Nishiguchi, H.; Endo, Y.; Toji, S.; Tanaka, H.; Nishimune, Y.; et al. Structure, expression, and chromosome mapping of LATS2, a mammalian homologue of the Drosophila tumor suppressor gene lats/warts. Genomics 2000, 63, 263–270. [Google Scholar] [CrossRef] [PubMed]
  214. Dong, J.; Feldmann, G.; Huang, J.; Wu, S.; Zhang, N.; Comerford, S.A.; Gayyed, M.F.; Anders, R.A.; Maitra, A.; Pan, D. Elucidation of a universal size-control mechanism in Drosophila and mammals. Cell 2007, 130, 1120–1133. [Google Scholar] [CrossRef] [PubMed]
  215. Sudol, M. Yes-associated protein (YAP65) is a proline-rich phosphoprotein that binds to the SH3 domain of the Yes proto-oncogene product. Oncogene 1994, 9, 2145–2152. [Google Scholar] [PubMed]
  216. Zhao, B.; Li, L.; Wang, L.; Wang, C.Y.; Yu, J.; Guan, K.L. Cell detachment activates the Hippo pathway via cytoskeleton reorganization to induce anoikis. Genes Dev. 2012, 26, 54–68. [Google Scholar] [CrossRef] [PubMed]
  217. Sudol, M.; Bork, P.; Einbond, A.; Kastury, K.; Druck, T.; Negrini, M.; Huebner, K.; Lehman, D. Characterization of the mammalian YAP (Yes-associated protein) gene and its role in defining a novel protein module, the WW domain. J. Biol. Chem. 1995, 270, 14733–14741. [Google Scholar] [CrossRef] [PubMed]
  218. Yagi, R.; Chen, L.F.; Shigesada, K.; Murakami, Y.; Ito, Y. A WW domain-containing yes-associated protein (YAP) is a novel transcriptional co-activator. EMBO J. 1999, 18, 2551–2562. [Google Scholar] [CrossRef] [PubMed]
  219. Varelas, X.; Samavarchi-Tehrani, P.; Narimatsu, M.; Weiss, A.; Cockburn, K.; Larsen, B.G.; Rossant, J.; Wrana, J.L. The Crumbs complex couples cell density sensing to Hippo-dependent control of the TGF-β-SMAD pathway. Dev. Cell 2010, 19, 831–844. [Google Scholar] [CrossRef] [PubMed]
  220. Zhao, B.; Lei, Q.Y.; Guan, K.L. The Hippo-YAP pathway: New connections between regulation of organ size and cancer. Curr. Opin. Cell Biol. 2008, 20, 638–646. [Google Scholar] [CrossRef] [PubMed]
  221. Heallen, T.; Zhang, M.; Wang, J.; Bonilla-Claudio, M.; Klysik, E.; Johnson, R.L.; Martin, J.F. Hippo pathway inhibits Wnt signaling to restrain cardiomyocyte proliferation and heart size. Science 2011, 332, 458–461. [Google Scholar] [CrossRef] [PubMed]
  222. Hong, J.H.; Hwang, E.S.; McManus, M.T.; Amsterdam, A.; Tian, Y.; Kalmukova, R.; Mueller, E.; Benjamin, T.; Spiegelman, B.M.; Sharp, P.A.; et al. TAZ, a transcriptional modulator of mesenchymal stem cell differentiation. Science 2005, 309, 1074–108. [Google Scholar] [CrossRef] [PubMed]
  223. Yang, N.; Morrison, C.D.; Liu, P.; Miecznikowski, J.; Bshara, W.; Han, S.; Zhu, Q.; Omilian, A.R.; Li, X.; Zhang, J. TAZ induces growth factor-independent proliferation through activation of EGFR ligand amphiregulin. Cell Cycle 2012, 11, 2922–2930. [Google Scholar] [CrossRef] [PubMed]
  224. Lian, I.; Kim, J.; Okazawa, H.; Zhao, J.; Zhao, B.; Yu, J.; Chinnaiyan, A.; Israel, M.A.; Goldstein, L.S.; Abujarour, R.; et al. The role of YAP transcription coactivator in regulating stem cell self-renewal and differentiation. Genes Dev. 2010, 24, 1106–1118. [Google Scholar] [CrossRef] [PubMed]
  225. Kim, M.; Kim, T.; Johnson, R.L.; Lim, D.S. Transcriptional co-repressor function of the Hippo pathway transducers YAP and TAZ. Cell Rep. 2015, 11, 270–282. [Google Scholar] [CrossRef] [PubMed]
  226. Moroishi, T.; Hayashi, T.; Pan, W.W.; Fujita, Y.; Holt, M.V.; Qin, J.; Carson, D.A.; Guan, K.L. The Hippo pathway kinases LATS1/2 suppress cancer immunity. Cell 2016, 167, 1525–1539 e17. [Google Scholar] [CrossRef] [PubMed]
  227. Liu-Chittenden, Y.; Huang, B.; Shim, J.S.; Chen, Q.; Lee, S.J.; Anders, R.A.; Liu, J.O.; Pan, D. Genetic and pharmacological disruption of the TEAD-YAP complex suppresses the oncogenic activity of YAP. Genes Dev. 2012, 26, 1300–1305. [Google Scholar] [CrossRef] [PubMed]
  228. Schlegelmilch, K.; Mohseni, M.; Kirak, O.; Pruszak, J.; Rodriguez, J.R.; Zhou, D.; Kreger, B.T.; Vasioukhin, V.; Avruch, J.; Brummelkamp, T.R.; et al. Yap1 acts downstream of α-catenin to control epidermal proliferation. Cell 2011, 144, 782–795. [Google Scholar] [CrossRef] [PubMed]
  229. Zhao, B.; Li, L.; Lu, Q.; Wang, L.H.; Liu, C.Y.; Lei, Q.; Guan, K.L. Angiomotin is a novel Hippo pathway component that inhibits YAP oncoprotein. Genes Dev. 2011, 25, 51–63. [Google Scholar] [CrossRef] [PubMed]
  230. Yu, F.X.; Guan, K.L. The Hippo pathway: Regulators and regulations. Genes Dev. 2013, 27, 355–371. [Google Scholar] [CrossRef] [PubMed]
  231. Dupont, S.; Morsut, L.; Aragona, M.; Enzo, E.; Giulitti, S.; Cordenonsi, M.; Zanconato, F.; Le Digabel, J.; Forcato, M.; Bicciato, S.; et al. Role of YAP/TAZ in mechanotransduction. Nature 2011, 474, 179–183. [Google Scholar] [CrossRef] [PubMed]
  232. Yu, F.X.; Zhao, B.; Panupinthu, N.; Jewell, J.L.; Lian, I.; Wang, L.H.; Zhao, J.; Yuan, H.; Tumaneng, K.; Li, H.; et al. Regulation of the Hippo-YAP pathway by G-protein-coupled receptor signaling. Cell 2012, 150, 780–791. [Google Scholar] [CrossRef] [PubMed]
  233. Zhou, D.; Conrad, C.; Xia, F.; Park, J.S.; Payer, B.; Yin, Y.; Lauwers, G.Y.; Thasler, W.; Lee, J.T.; Avruch, J.; et al. Mst1 and Mst2 maintain hepatocyte quiescence and suppress hepatocellular carcinoma development through inactivation of the Yap1 oncogene. Cancer Cell 2009, 16, 425–438. [Google Scholar] [CrossRef] [PubMed]
  234. Camargo, F.D.; Gokhale, S.; Johnnidis, J.B.; Fu, D.; Bell, G.W.; Jaenisch, R.; Brummelkamp, T.R. YAP1 increases organ size and expands undifferentiated progenitor cells. Curr. Biol. 2007, 17, 2054–2060. [Google Scholar] [CrossRef] [PubMed]
  235. Varelas, X. The Hippo pathway effectors TAZ and YAP in development, homeostasis and disease. Development 2014, 141, 1614–1626. [Google Scholar] [CrossRef] [PubMed]
  236. Cox, A.G.; Hwang, K.L.; Brown, K.K.; Evason, K.J.; Beltz, S.; Tsomides, A.; O’Connor, K.; Galli, G.G.; Yimlamai, D.; Chhangawala, S.; et al. Yap reprograms glutamine metabolism to increase nucleotide biosynthesis and enable liver growth. Nat. Cell Biol. 2016, 18, 886–896. [Google Scholar] [CrossRef] [PubMed]
  237. Valis, K.; Talacko, P.; Grobarova, V.; Cerny, J.; Novak, P. Shikonin regulates C-MYC and GLUT1 expression through the MST1-YAP1-TEAD1 axis. Exp. Cell Res. 2016, 349, 273–281. [Google Scholar] [CrossRef] [PubMed]
  238. Noto, A.; de Vitis, C.; Pisanu, M.E.; Roscilli, G.; Ricci, G.; Catizone, A.; Sorrentino, G.; Chianese, G.; Taglialatela-Scafati, O.; Trisciuoglio, D.; et al. Stearoyl-CoA-desaturase 1 regulates lung cancer stemness via stabilization and nuclear localization of YAP/TAZ. Oncogene 2017. [Google Scholar] [CrossRef] [PubMed]
  239. Enzo, E.; Santinon, G.; Pocaterra, A.; Aragona, M.; Bresolin, S.; Forcato, M.; Grifoni, D.; Pession, A.; Zanconato, F.; Guzzo, G.; et al. Aerobic glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 2015, 34, 1349–1370. [Google Scholar] [CrossRef] [PubMed]
  240. Wang, W.; Xiao, Z.D.; Li, X.; Aziz, K.E.; Gan, B.; Johnson, R.L.; Chen, J. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat. Cell Biol. 2015, 17, 490–499. [Google Scholar] [CrossRef] [PubMed]
  241. Santinon, G.; Pocaterra, A.; Dupont, S. Control of YAP/TAZ Activity by Metabolic and Nutrient-Sensing Pathways. Trends Cell Biol. 2016, 26, 289–299. [Google Scholar] [CrossRef] [PubMed]
  242. Vigneron, A.M.; Ludwig, R.L.; Vousden, K.H. Cytoplasmic ASPP1 inhibits apoptosis through the control of YAP. Genes Dev. 2010, 24, 2430–2439. [Google Scholar] [CrossRef] [PubMed]
  243. Xie, Q.; Chen, J.; Feng, H.; Peng, S.; Adams, U.; Bai, Y.; Huang, L.; Li, J.; Huang, J.; Meng, S.; et al. YAP/TEAD-mediated transcription controls cellular senescence. Cancer Res. 2013, 73, 3615–3624. [Google Scholar] [CrossRef] [PubMed]
  244. Johnson, R.; Halder, G. The two faces of Hippo: Targeting the Hippo pathway for regenerative medicine and cancer treatment. Nat. Rev. Drug Discov. 2014, 13, 63–79. [Google Scholar] [CrossRef] [PubMed]
  245. Fallahi, E.; O’Driscoll, N.A.; Matallanas, D. The MST/Hippo pathway and cell death: A non-canonical affair. Genes 2016, 7, 28. [Google Scholar] [CrossRef] [PubMed]
  246. Mori, M.; Triboulet, R.; Mohseni, M.; Schlegelmilch, K.; Shrestha, K.; Camargo, F.D.; Gregory, R.I. Hippo signaling regulates microprocessor and links cell-density-dependent miRNA biogenesis to cancer. Cell 2014, 156, 893–906. [Google Scholar] [CrossRef] [PubMed]
  247. Lo Sardo, F.; Forcato, M.; Sacconi, A.; Capaci, V.; Zanconato, F.; Di Agostino, S.; Del Sal, G.; Pandolfi, P.P.; Strano, S.; Bicciato, S.; et al. MCM7 and its hosted miR-25, 93 and 106b cluster elicit YAP/TAZ oncogenic activity in lung cancer. Carcinogenesis 2017, 38, 64–75. [Google Scholar] [CrossRef] [PubMed]
  248. Chaulk, S.G.; Lattanzi, V.J.; Hiemer, S.E.; Fahlman, R.P.; Varelas, X. The Hippo pathway effectors TAZ/YAP regulate dicer expression and microRNA biogenesis through Let-7. J. Biol. Chem. 2014, 289, 1886–1891. [Google Scholar] [CrossRef] [PubMed]
  249. Yu, W.; Qiao, Y.; Tang, X.; Ma, L.; Wang, Y.; Zhang, X.; Weng, W.; Pan, Q.; Yu, Y.; Sun, F.; et al. Tumor suppressor long non-coding RNA, MT1DP is negatively regulated by YAP and Runx2 to inhibit FoxA1 in liver cancer cells. Cell. Signal. 2014, 26, 2961–2968. [Google Scholar] [CrossRef] [PubMed]
  250. Wang, Z.; Wu, Y.; Wang, H.; Zhang, Y.; Mei, L.; Fang, X.; Zhang, X.; Zhang, F.; Chen, H.; Liu, Y.; et al. Interplay of mevalonate and Hippo pathways regulates RHAMM transcription via YAP to modulate breast cancer cell motility. Proc. Natl. Acad. Sci. USA 2014, 111, E89–E98. [Google Scholar] [CrossRef] [PubMed]
  251. Park, J.; Jeong, S. Wnt activated β-catenin and YAP proteins enhance the expression of non-coding RNA component of RNase MRP in colon cancer cells. Oncotarget 2015, 6, 34658–34668. [Google Scholar] [PubMed]
  252. Aylon, Y.; Gershoni, A.; Rotkopf, R.; Biton, I.E.; Porat, Z.; Koh, A.P.; Sun, X.; Lee, Y.; Fiel, M.I.; Hoshida, Y.; et al. The LATS2 tumor suppressor inhibits SREBP and suppresses hepatic cholesterol accumulation. Genes Dev. 2016, 30, 786–797. [Google Scholar] [CrossRef] [PubMed]
  253. Bai, N.; Zhang, C.; Liang, N.; Zhang, Z.; Chang, A.; Yin, J.; Li, Z.; Li, N.; Tan, X.; Luo, N.; et al. Yes-associated protein (YAP) increases chemosensitivity of hepatocellular carcinoma cells by modulation of p53. Cancer Biol. Ther. 2013, 14, 511–520. [Google Scholar] [CrossRef] [PubMed]
  254. Aylon, Y.; Michael, D.; Shmueli, A.; Yabuta, N.; Nojima, H.; Oren, M. A positive feedback loop between the p53 and Lats2 tumor suppressors prevents tetraploidization. Genes Dev. 2006, 20, 2687–2700. [Google Scholar] [CrossRef] [PubMed]
  255. Aylon, Y.; Sarver, A.; Tovy, A.; Ainbinder, E.; Oren, M. Lats2 is critical for the pluripotency and proper differentiation of stem cells. Cell Death Differ. 2014, 21, 624–633. [Google Scholar] [CrossRef] [PubMed]
  256. Aylon, Y.; Ofir-Rosenfeld, Y.; Yabuta, N.; Lapi, E.; Nojima, H.; Lu, X.; Oren, M. The Lats2 tumor suppressor augments p53-mediated apoptosis by promoting the nuclear proapoptotic function of ASPP1. Genes Dev. 2010, 24, 2420–2429. [Google Scholar] [CrossRef] [PubMed]
  257. Matallanas, D.; Romano, D.; Yee, K.; Meissl, K.; Kucerova, L.; Piazzolla, D.; Baccarini, M.; Vass, J.K.; Kolch, W.; O’Neill, E. RASSF1A elicits apoptosis through an MST2 pathway directing proapoptotic transcription by the p73 tumor suppressor protein. Mol. Cell 2007, 27, 962–975. [Google Scholar] [CrossRef] [PubMed]
  258. Basu, S.; Totty, N.F.; Irwin, M.S.; Sudol, M.; Downward, J. Akt phosphorylates the Yes-associated protein, YAP, to induce interaction with 14-3-3 and attenuation of p73-mediated apoptosis. Mol. Cell 2003, 11, 11–23. [Google Scholar] [CrossRef]
  259. Downward, J.; Basu, S. YAP and p73: A complex affair. Mol. Cell. Biol. 2008, 32, 749–750. [Google Scholar] [CrossRef] [PubMed]
  260. Lapi, E.; Di Agostino, S.; Donzelli, S.; Gal, H.; Domany, E.; Rechavi, G.; Pandolfi, P.P.; Givol, D.; Strano, S.; Lu, X.; et al. PML, YAP, and p73 are components of a proapoptotic autoregulatory feedback loop. Mol. Cell 2008, 32, 803–814. [Google Scholar] [CrossRef] [PubMed]
  261. Levy, D.; Adamovich, Y.; Reuven, N.; Shaul, Y. The Yes-associated protein 1 stabilizes p73 by preventing Itch-mediated ubiquitination of p73. Cell Death Differ. 2007, 14, 743–751. [Google Scholar] [CrossRef] [PubMed]
  262. Strano, S.; Monti, O.; Pediconi, N.; Baccarini, A.; Fontemaggi, G.; Lapi, E.; Mantovani, F.; Damalas, A.; Citro, G.; Sacchi, A.; et al. The transcriptional coactivator Yes-associated protein drives p73 gene-target specificity in response to DNA Damage. Mol. Cell 2005, 18, 447–459. [Google Scholar] [CrossRef] [PubMed]
  263. Levy, D.; Adamovich, Y.; Reuven, N.; Shaul, Y. Yap1 phosphorylation by c-Abl is a critical step in selective activation of proapoptotic genes in response to DNA damage. Mol. Cell 2008, 29, 350–361. [Google Scholar] [CrossRef] [PubMed]
  264. Fausti, F.; Di Agostino, S.; Cioce, M.; Bielli, P.; Sette, C.; Pandolfi, P.; Oren, M.; Sudol, M.; Strano, S.; Blandino, G. ATM kinase enables the functional axis of YAP, PML and p53 to ameliorate loss of Werner protein-mediated oncogenic senescence. Cell Death Differ. 2013, 20, 1498–1509. [Google Scholar] [CrossRef] [PubMed]
  265. 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] [PubMed]
  266. Escoll, M.; Gargini, R.; Cuadrado, A.; Anton, I.M.; Wandosell, F. Mutant p53 oncogenic functions in cancer stem cells are regulated by WIP through YAP/TAZ. Oncogene 2017. [Google Scholar] [CrossRef] [PubMed]
  267. Aylon, Y.; Oren, M. The Hippo pathway, p53 and cholesterol. Cell Cycle 2016, 15, 2248–2255. [Google Scholar] [CrossRef] [PubMed]
  268. Wu, Q.; Ishikawa, T.; Sirianni, R.; Tang, H.; McDonald, J.G.; Yuhanna, I.S.; Thompson, B.; Girard, L.; Mineo, C.; Brekken, R.A.; et al. 27-Hydroxycholesterol promotes cell-autonomous, ER-positive breast cancer growth. Cell Rep. 2013, 5, 637–645. [Google Scholar] [CrossRef] [PubMed]
  269. Horton, J.D.; Goldstein, J.L.; Brown, M.S. SREBPs: Activators of the complete program of cholesterol and fatty acid synthesis in the liver. J. Clin. Investig. 2002, 109, 1125–1131. [Google Scholar] [CrossRef] [PubMed]
  270. Guo, A.; Salomoni, P.; Luo, J.; Shih, A.; Zhong, S.; Gu, W.; Pandolfi, P.P. The function of PML in p53-dependent apoptosis. Nat. Cell Biol. 2000, 2, 730–736. [Google Scholar] [PubMed]
  271. Pearson, M.; Pelicci, P.G. PML interaction with p53 and its role in apoptosis and replicative senescence. Oncogene 2001, 20, 7250–7256. [Google Scholar] [CrossRef] [PubMed]
  272. Girardini, J.E.; del Sal, G. Improving pharmacological rescue of p53 function: RITA targets mutant p53. Cell Cycle 2010, 9, 2062. [Google Scholar] [CrossRef] [PubMed]
  273. Parrales, A.; Iwakuma, T. Targeting Oncogenic Mutant p53 for Cancer Therapy. Front. Oncol. 2015, 5, 288. [Google Scholar] [CrossRef] [PubMed]
  274. Rao, C.V.; Patlolla, J.M.; Qian, L.; Zhang, Y.; Brewer, M.; Mohammed, A.; Desai, D.; Amin, S.; Lightfoot, S.; Kopelovich, L. Chemopreventive effects of the p53-modulating agents CP-31398 and Prima-1 in tobacco carcinogen-induced lung tumorigenesis in A/J mice. Neoplasia 2013, 15, 1018–1027. [Google Scholar] [CrossRef] [PubMed]
  275. Bykov, V.J.; Wiman, K.G. Mutant p53 reactivation by small molecules makes its way to the clinic. FEBS Lett. 2014, 588, 2622–2627. [Google Scholar] [CrossRef] [PubMed]
  276. Di Agostino, S.; Cortese, G.; Monti, O.; Dell’Orso, S.; Sacchi, A.; Eisenstein, M.; Citro, G.; Strano, S.; Blandino, G. The disruption of the protein complex mutantp53/p73 increases selectively the response of tumor cells to anticancer drugs. Cell Cycle 2008, 7, 3440–3447. [Google Scholar] [CrossRef] [PubMed]
  277. Lambert, J.M.; Gorzov, P.; Veprintsev, D.B.; Soderqvist, M.; Segerback, D.; Bergman, J.; Fersht, A.R.; Hainaut, P.; Wiman, K.G.; Bykov, V.J. PRIMA-1 reactivates mutant p53 by covalent binding to the core domain. Cancer Cell 2009, 15, 376–388. [Google Scholar] [CrossRef] [PubMed]
  278. Saha, M.N.; Chen, Y.; Chen, M.H.; Chen, G.; Chang, H. Small molecule MIRA-1 induces in vitro and in vivo anti-myeloma activity and synergizes with current anti-myeloma agents. Br. J. Cancer 2014, 110, 2224–2231. [Google Scholar] [CrossRef] [PubMed]
  279. Watanabe, F.T.; Chade, D.C.; Reis, S.T.; Piantino, C.; Dall’ Oglio, M.F.; Srougi, M.; Leite, K.R. Curcumin, but not Prima-1, decreased tumor cell proliferation in the syngeneic murine orthotopic bladder tumor model. Clinics 2011, 66, 2121–2124. [Google Scholar] [CrossRef] [PubMed]
  280. Zandi, R.; Selivanova, G.; Christensen, C.L.; Gerds, T.A.; Willumsen, B.M.; Poulsen, H.S. PRIMA-1Met/APR-246 induces apoptosis and tumor growth delay in small cell lung cancer expressing mutant p53. Clin. Cancer Res. 2011, 17, 2830–2841. [Google Scholar] [CrossRef] [PubMed]
  281. Lewis, E.J. PRIMA-1 as a cancer therapy restoring mutant p53: A review. Biosciencehorizons 2015, 8, 1–5. [Google Scholar] [CrossRef]
  282. Bou-Hanna, C.; Jarry, A.; Lode, L.; Schmitz, I.; Schulze-Osthoff, K.; Kury, S.; Bezieau, S.; Mosnier, J.F.; Laboisse, C.L. Acute cytotoxicity of MIRA-1/NSC19630, a mutant p53-reactivating small molecule, against human normal and cancer cells via a caspase-9-dependent apoptosis. Cancer Lett. 2015, 359, 211–217. [Google Scholar] [CrossRef] [PubMed]
  283. Piantino, C.B.; Reis, S.T.; Viana, N.I.; Silva, I.A.; Morais, D.R.; Antunes, A.A.; Dip, N.; Srougi, M.; Leite, K.R. Prima-1 induces apoptosis in bladder cancer cell lines by activating p53. Clinics 2013, 68, 297–303. [Google Scholar] [CrossRef]
  284. Saha, M.N.; Jiang, H.; Yang, Y.; Reece, D.; Chang, H. PRIMA-1Met/APR-246 displays high antitumor activity in multiple myeloma by induction of p73 and Noxa. Mol. Cancer Ther. 2013, 12, 2331–2341. [Google Scholar] [CrossRef] [PubMed]
  285. Shalom-Feuerstein, R.; Serror, L.; Aberdam, E.; Muller, F.J.; van Bokhoven, H.; Wiman, K.G.; Zhou, H.; Aberdam, D.; Petit, I. Impaired epithelial differentiation of induced pluripotent stem cells from ectodermal dysplasia-related patients is rescued by the small compound APR-246/PRIMA-1MET. Proc. Natl. Acad. Sci. USA 2013, 110, 2152–2156. [Google Scholar] [CrossRef] [PubMed]
  286. Sobhani, M.; Abdi, J.; Manujendra, S.N.; Chen, C.; Chang, H. PRIMA-1Met induces apoptosis in Waldenstrom’s Macroglobulinemia cells independent of p53. Cancer Biol. Ther. 2015, 16, 799–806. [Google Scholar] [CrossRef] [PubMed]
  287. Zache, N.; Lambert, J.M.; Wiman, K.G.; Bykov, V.J. PRIMA-1MET inhibits growth of mouse tumors carrying mutant p53. Cell. Oncol. 2008, 30, 411–418. [Google Scholar] [PubMed]
  288. Zhang, S.; Zhou, L.; Hong, B.; van den Heuvel, A.P.; Prabhu, V.V.; Warfel, N.A.; Kline, C.L.; Dicker, D.T.; Kopelovich, L.; El-Deiry, W.S. Small-molecule NSC59984 restores p53 pathway signaling and antitumor effects against colorectal cancer via p73 activation and degradation of mutant p53. Cancer Res. 2015, 75, 3842–3852. [Google Scholar] [CrossRef] [PubMed]
  289. Freed-Pastor, W.; Prives, C. Targeting mutant p53 through the mevalonate pathway. Nat. Cell Biol. 2016, 18, 1122–1124. [Google Scholar] [CrossRef] [PubMed]
  290. 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]
  291. Oku, Y.; Nishiya, N.; Shito, T.; Yamamoto, R.; Yamamoto, Y.; Oyama, C.; Uehara, Y. Small molecules inhibiting the nuclear localization of YAP/TAZ for chemotherapeutics and chemosensitizers against breast cancers. FEBS Open Bio. 2015, 5, 542–549. [Google Scholar] [CrossRef] [PubMed]
  292. Sun, H.; Ying, M.Y. Small molecule drug Verteporfin inhibits TAZ/YAP-driven signaling and tumorigenicity of breast cancer cells. Cancer Res. 2015, 75. [Google Scholar] [CrossRef]
  293. Zanconato, F.; Battilana, G.; Cordenonsi, M.; Piccolo, S. YAP/TAZ as therapeutic targets in cancer. Curr. Opin. Pharmacol. 2016, 29, 26–33. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The Hippo pathway is activated by different stimuli and this activation induces Yes-Associated Protein (YAP) and Tafazzin (TAZ) phosphorylation and cytoplasmic retention. Moreover, the activated core kinase cassette (Mst1/2-Mob1-Sav1-Lats1/2) can contribute with CDK1δ/ε in inducing YAP and TAZ protein degradation. YAP and TAZ oncogenes, when activated and not phosphorylated, can translocate into the nucleus and bind to different transcription factors inducing osteoblast differentiation, cell proliferation, survival, mevalonate pathway activation, increased cell metabolism, anchorage-independent growth, EMT, ESCs and iPSCs pluripotency, cell migration, cell invasion, and tissue regeneration. Moreover, YAP and TAZ can function as co-repressors binding to the NuRD complex and inhibiting transcription of target genes. Arrows and T-bars indicate activating and inhibiting functions respectively. The symbol “X” in red color represents a transcriptional blockage.
Figure 1. The Hippo pathway is activated by different stimuli and this activation induces Yes-Associated Protein (YAP) and Tafazzin (TAZ) phosphorylation and cytoplasmic retention. Moreover, the activated core kinase cassette (Mst1/2-Mob1-Sav1-Lats1/2) can contribute with CDK1δ/ε in inducing YAP and TAZ protein degradation. YAP and TAZ oncogenes, when activated and not phosphorylated, can translocate into the nucleus and bind to different transcription factors inducing osteoblast differentiation, cell proliferation, survival, mevalonate pathway activation, increased cell metabolism, anchorage-independent growth, EMT, ESCs and iPSCs pluripotency, cell migration, cell invasion, and tissue regeneration. Moreover, YAP and TAZ can function as co-repressors binding to the NuRD complex and inhibiting transcription of target genes. Arrows and T-bars indicate activating and inhibiting functions respectively. The symbol “X” in red color represents a transcriptional blockage.
Ijms 18 00961 g001
Figure 2. In a wild-type p53 context, DNA damage, oncogenic stress, senescence or apoptosis induction can trigger Lats1/2 activation and p53 and YAP mutual induction leading to a tumor suppressor response that is sustained by the p53/YAP/PML/p73 axis. Cancers harboring mutant p53 show an oncogenic cooperation between YAP/TAZ and mutant p53 that reinforce tumor growth, chemoresistance and migration in cancer cells. Restoring wild-type p53 functions or destroying oncogenic mutant p53/YAP/TAZ crosstalk could be good strategies for recovering the tumor suppressor program promoted by YAP and p53. Arrows and T-bars indicate activating and inhibiting functions respectively. The symbol “X” in red color represents a downregulation in protein levels.
Figure 2. In a wild-type p53 context, DNA damage, oncogenic stress, senescence or apoptosis induction can trigger Lats1/2 activation and p53 and YAP mutual induction leading to a tumor suppressor response that is sustained by the p53/YAP/PML/p73 axis. Cancers harboring mutant p53 show an oncogenic cooperation between YAP/TAZ and mutant p53 that reinforce tumor growth, chemoresistance and migration in cancer cells. Restoring wild-type p53 functions or destroying oncogenic mutant p53/YAP/TAZ crosstalk could be good strategies for recovering the tumor suppressor program promoted by YAP and p53. Arrows and T-bars indicate activating and inhibiting functions respectively. The symbol “X” in red color represents a downregulation in protein levels.
Ijms 18 00961 g002

Share and Cite

MDPI and ACS Style

Ferraiuolo, M.; Verduci, L.; Blandino, G.; Strano, S. Mutant p53 Protein and the Hippo Transducers YAP and TAZ: A Critical Oncogenic Node in Human Cancers. Int. J. Mol. Sci. 2017, 18, 961. https://doi.org/10.3390/ijms18050961

AMA Style

Ferraiuolo M, Verduci L, Blandino G, Strano S. Mutant p53 Protein and the Hippo Transducers YAP and TAZ: A Critical Oncogenic Node in Human Cancers. International Journal of Molecular Sciences. 2017; 18(5):961. https://doi.org/10.3390/ijms18050961

Chicago/Turabian Style

Ferraiuolo, Maria, Lorena Verduci, Giovanni Blandino, and Sabrina Strano. 2017. "Mutant p53 Protein and the Hippo Transducers YAP and TAZ: A Critical Oncogenic Node in Human Cancers" International Journal of Molecular Sciences 18, no. 5: 961. https://doi.org/10.3390/ijms18050961

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