Next Article in Journal
Machine Learning Characterization of Ictal and Interictal States in EEG Aimed at Automated Seizure Detection
Previous Article in Journal
Heterocycle-Based Multicomponent Reactions in Drug Discovery: From Hit Finding to Rational Design
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

P63 and P73 Activation in Cancers with p53 Mutation

1
School of Medicine, I-Shou University, No. 8, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
2
Department of Medical Laboratory Science, I-Shou University, No. 8, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
3
Department of Biological Science and Technology, I-Shou University, No. 8, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
4
Department of Otolaryngology-Head and Neck Surgery, E-DA Hospital, No. 1, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
5
Department of Physical Therapy, I-Shou University, No. 8, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
6
School of Chinese Medicine for Post-Baccalaureate, I-Shou University, No. 8, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
7
Department of Pathology, E-Da Hospital, I-Shou University, No. 1, Yida Rd., Jiaosu Village Yanchao District, Kaohsiung 82445, Taiwan
*
Authors to whom correspondence should be addressed.
Biomedicines 2022, 10(7), 1490; https://doi.org/10.3390/biomedicines10071490
Submission received: 23 May 2022 / Revised: 17 June 2022 / Accepted: 21 June 2022 / Published: 23 June 2022
(This article belongs to the Special Issue Advanced Research in Tumor Suppressor)

Abstract

:
The members of the p53 family comprise p53, p63, and p73, and full-length isoforms of the p53 family have a tumor suppressor function. However, p53, but not p63 or p73, has a high mutation rate in cancers causing it to lose its tumor suppressor function. The top and second-most prevalent p53 mutations are missense and nonsense mutations, respectively. In this review, we discuss possible drug therapies for nonsense mutation and a missense mutation in p53. p63 and p73 activators may be able to replace mutant p53 and act as anti-cancer drugs. Herein, these p63 and p73 activators are summarized and how to improve these activator responses, particularly focusing on p53 gain-of-function mutants, is discussed.

1. Introduction to the p53 Family

The p53 family has three members, p53, p63, and p73 [1,2,3]. TA (transactivation) isoforms of p53 family members are tumor suppressor genes [4,5]. p53 has a high frequency of mutation in cancers causing loss of its tumor suppression function [6,7]; however, p63 and p73 are rarely mutated in cancers [8,9,10]. In this review, we briefly introduce each of the members with an emphasis on the most common mutations of p53 making it nonfunctional. Further, we summarize p63 and p73 activators that can replace them to obtain a similar tumor suppressor function in the p53 family. Some p53 mutants can obtain oncogenic function as a gain of function similar to an oncogene [11,12]. We also discuss how to improve p63 or p73 activator drug response in p53 gain-of-function mutation cancer cells.

1.1. p53

p53 was the second tumor suppressor gene identified, although p53 was actually discovered in 1979, before the first tumor suppressor gene Rb, which was cloned in 1986 [13]. p53 has a function in apoptosis, cell cycle arrest, autophagy, metabolism, DNA repair, translational control, and feedback mechanisms [14,15]. p53 knockout mice are prone to the spontaneous variety of tumors by 6 months of age (~age 34 in humans) [16]. p53 has an average ~50% mutation rate in cancers [6,7,17]. According to the Catalogue of Somatic Mutations in Cancer (COSMIC) database [18], there are 177,561 unique clinical samples with 50,215 unique samples having p53 mutations (Figure 1A). The top p53 mutation type is missense mutation accounting for 62.74% of mutations, and the second most prevalent mutation of p53 is a nonsense mutation which accounts for 10.72% of mutations. Eight missense mutations R175H, G245S, R248Q, R248W, R249S, R273H, R273S, and R282W called hotspot mutations account for ~28% of all p53 mutations identified in cancers [19]. p53 germline mutation can cause Li-Fraumeni syndrome (LFS) which is a hereditary syndrome with a relatively early age of cancer diagnosis usually before the age of 36 [20]. This syndrome is characterized by the early onset of various types of cancer such as soft-tissue, breast, brain, leukemia, lymphoma, gastrointestinal, head and neck, kidney, larynx, lung, skin, ovary, pancreas, prostate, testis, thyroid, and adrenocortical cancers [21,22]. Some congenital p53 mutations of the LFS are similar to those acquired by p53 hotspot mutations in cancers such as G245S [23,24], but some congenital p53 mutations only appear in LFS as germ-line specific mutations such as R337H [25,26]. In this review, we focus on drug therapy of nonsense and missense mutations of p53. The missense mutation is divided into two types, loss-of-function mutations, and gain-of-function mutations. We focus particularly on p63 and p73 reactivation with drug treatment to replace part of the p53 tumor-suppressor function.

1.2. p63

p63 was cloned in 1998 [27]. The major C-terminal p63 isoforms are p63α, while dominant-negative ΔNp63 was the predominant N-terminal isoform in most tissues from the p73-High/p63-High group [28]. p63 is a rare rarely mutated in cancers [8]. According to the COSMIC database [18], there are 55,869 unique clinical samples with 2160 unique samples having p63 mutations (Figure 1B). The nonsense mutation rate of p63 is only 2.97%, and the missense mutation rate of p63 is 32.34%. Although p63 is rarely mutated in cancers, distinct p63 germline mutation can cause several different types of abnormal development issues. Ectrodactyly, ectodermal dysplasia, and cleft lip/palate (EEC) syndrome are mainly characterized by severe ectrodactyly and limb defects with a p63 missense mutation in the middle of the DNA binding domain [29]. The distinguishing features of ankyloblepharon-ectodermal defects-cleft lip/palate (AEC) syndrome are ankyloblepharon, congenital erythroderma, skin fragility, atrophy, palmoplantar hyperkeratosis, and extensive skin erosions with p63 missense mutation in the C-terminal the sterile-α-motif (SAM) domain and transcriptional inhibitory (TI) domains [30,31]. Different p63 mutations can also cause isolated split hand/foot malformation (SHFM) nonsyndromic diseases with a missense mutation of K193E and K194E and nonsense mutants of Q634X and E639X on TAp63α [32]. p63 knockout mice have been found to fail to form the stratified epidermis, limbs, teeth, mammary glands, and postnatal lethality due to dehydration [33,34].

1.3. p73

p73 was cloned in 1997 [35]. Compared to p63, p73 also commonly expresses α isoforms [28] and the expression of TAp73 is higher than ∆Np73 [36]. p73 is rarely mutated in cancers [9,10]. According to the COSMIC database [18], there are 42,580 unique clinical samples with 879 unique samples with p73 mutations (Figure 1C). The nonsense mutation rate of p73 is only 2.28%, and the missense mutation rate of p73 is 27.33%. There are no reports about p73 germline mutation in relation to any type of genetic disorder or syndrome. But hydrocephalus, hippocampal dysgenesis, and pheromonal defects can be found in p73 knockout mice [37].

2. Types of p53 Mutations in Cancers

2.1. Nonsense Mutation

p53 nonsense mutations comprise ~10% of all p53 mutants (Figure 1A); the actual p53 nonsense mutation rate is higher than the average ~5% [38]. There are three pre-stop DNA codons, TAA, TAG, and TGA. Nonsense mutation leads to the generation of premature termination codons (PTC), which leads to nonsense-mediated mRNA decay (NMD), resulting in the inability to express full-length proteins and extremely low expression levels of truncated proteins [39]. Two mechanisms are known to regenerate full-length proteins, one is to inhibit NMD, and the other is for PTC readthrough. It is known that aminoglycoside drugs, such as G418 and gentamicin, can inhibit NMD and promote p53 PTC readthrough [40,41]. But these drugs are highly toxic and cannot be used in clinical practice. 2,6-Diaminopurine (DAP), can inhibit the activity of putative ribosomal RNA methyltransferase 1 (FTSJ1) to increase the capacity of tRNATrp to recognize the UGA stop codon to promote p53 PTC readthrough, but this drug does not have the ability to inhibit NMD [42]. Furthermore, DAP is only effective for nonsense mutations of TGA but not TAA or TAG [42], and this greatly reduces the available targets. In addition, some phthalimide derivatives and antimalarial drug quinines can promote the p53 PTC read-through ability of G418 to increase the proportion of full-length p53 and to reduce the expression of the truncated protein, but these drugs alone have no effect on PTC read-through [43,44]. Non-aminoglycoside drugs, such as Ataluren (PTC124), also increase the read-through ability of PTC without the ability to inhibit NMD, and PTC124 has been used in clinical phase II or III trials to treat genetic diseases with specific nonsense mutations [45,46]. PTC124 can also promote p53 PTC readthrough [47]. A recent study has shown that CC-885 and CC-90009 can inhibit NMD; of note, the effective concentration of CC-885 for treatment of p53 nonsense mutation with TAA is only one-tenth of that of CC-90009 [48].

2.2. Loss-of-Function Mutants

p53 missense mutations contain both loss of function and gain of function. p53 mutation is mainly located at the N-terminal transactivation domain or middle DNA binding domain [2]. In addition, several point mutants still have normal DNA binding function [49,50,51,52]; most p53 mutations within the N-terminal transactivation domain or the DNA binding domain lose their transactivation function or DNA binding function causing loss of their tumor suppressor functions such as cell cycle checkpoint controls and apoptosis [53,54,55]. These p53 mutations can associate with p63 and p73, whereas wild-type p53 cannot [56,57]. Therefore, loss-of-function p53 mutants act in the same way as the ∆N isofroms of the p53 family having a dominant-negative effect to repress the functions of normal TA isoforms of p53 family members [57,58].

2.3. Gain-of-Function Mutations

Some p53 mutants can obtain some oncogenic functions such as cell migration, invasion, and metastasis to enhance tumorigenesis [59], and these p53 mutants are called gain-of-function mutants. The acquisition of p53 gain-of-function is via three mechanisms [60,61]. First, mutant p53 can directly bind to the novel binding site with a p53 non-canonical sequence to activate several oncogenic genes [51]. Second, mutant p53 can act as a co-activator to bind to other transcription factors to activate some oncogenic genes [62]. Third, mutant p53 can bind to other tumor suppressive-type transcription factors to cause loss of transcription ability [63]. Some p53 mutants can become aggregated in several types of cancer, such as breast, lung ovary, colorectal, and head and neck cancers [64,65,66,67]. It is known that these aggregations of mutant p53 can sequester other tumor suppressor genes as a third mechanism to cause p53 gain of function [68,69,70].

3. Factors Influencing p53 Mutant Gain of Function

Several factors have been reported to influence p53 mutant gain of function (Figure 2A). The heat shock protein 70 (HSP70) has been reported to enhance mutant p53 aggregation [71], and heat shock protein 90 (HSP90) can repress mutant p53 aggregation [71,72]. SIRT1 is an NAD+ dependent histone deacetylase that has been reported to deacetylate HSF1 to enhance HSF1 transcriptional activity to increase HSP70 expression [73]. NAMPT can enhance SIRT1 activity by increasing the amount of NAD+ [74]. p53 gain-of-function mutant can induce MYC [62,75], and MYC can enhance NAPMT [76]. Wild-type p53 can induce 14-3-3σ expression [77,78], and 14-3-3σ can promote MYC poly-ubiquitination and degradation [79]. Another article also reported that wild-type p53 can bind to G-quadruplexes on MYC promoter to repress MYC expression [80]. Both wild-type and non-gain-of-function p53 mutants can activate lincRNA-p21 through binding to G-quadruplexes on lincRNA-p21 promoter [81], and lincRNA-p21 can repress STAT3 [82]. STAT3 can bind to the HSP70 promoter to active HSP70 expression [83,84]. The relationships between factors that influence p53 mutant gain of function are summarized in Figure 2A.
MircoRNA and long-noncoding RNA are also key regulators of mutant p53. miR-150 can repress p53 [85,86,87]. An oncogenic type long-noncoding RNA LINC00460 can act as a sponge to repress miR-150 targeting to enhance mutant p53 expression [88]. Another tumor-suppressor type transcription factor, AP2, frequently interacts synergistically with p53 to activate downstream genes such as p21, CD82, and NEU4 [89,90,91,92]. AP2 can repress another oncogenic-type long-noncoding RNA, LINC00511 expression [93]. LINC00511 also can act as a sponge to repress miR-150 targeting [94]. So AP2 may also repress mutant p53 expression through the LINC00511-miR-150 axis (Figure 2B). AP2 has also been reported to decrease the amount of p53 [95].

4. p63 Activation Drugs

Because TAp63 isoforms have limited expression in organs [28], only a few reports have looked at TAp63 activation for anti-cancer purposes; they are summarized in Table 1. Bliotoxin was able to upregulate the levels of DAPK1 to induce TAp63 but not p53 or TAp73 expression to induce apoptosis in paclitaxel pretreated paclitaxel-resistant CaOV-3 and SK-OV-3 ovarian cancer cells [96]. Lovastatin was able to induce TP63 transactivation through phosphorylation of the AMPK-p38MAPK-TAp63 cascade to cause hypopharyngeal carcinoma FaDu cell death [97]. TAp63 can active PUMA (p53 upregulated modulator of apoptosis) [98,99], and interferon-α can induce TP63 and PUMA expression in hepatocyte derived cellular carcinoma cell line HuH7 cells [100].
miR-130b mimics can activate TAp63 and repress ΔNp63 to decrease cell viability in ovarian cancer Ovcar1-8 cells [101]. miR-124 mimics can activate TAp63 and repress ΔNp63 to inhibit cell growth in LoVo and SW480 colorectal cancer cells [102]. miR-140 can directly target p63 3′-UTR sequences to repress p63, and the miR-140 inhibitor can activate TAp63 in HGC-27 and BGC-823 gastric cancer cells to induce cell apoptosis [103].
Table 1. Table summarizing p63 activation drugs used to treat cancer cell lines. If there is no COSMIC ID or ATCC ID for a certain cell line from the p53 database [104], the PubMed ID is provided for information about the p53 status within indicated cell line. CDS, coding sequence. *, premature stop codons.
Table 1. Table summarizing p63 activation drugs used to treat cancer cell lines. If there is no COSMIC ID or ATCC ID for a certain cell line from the p53 database [104], the PubMed ID is provided for information about the p53 status within indicated cell line. CDS, coding sequence. *, premature stop codons.
p63
Activation Drugs
Cell LineCOSMIC IDATCC IDPubMed IDTissuep53 StatusZygosityCDS
Mutation
Protein ChangeRef.
gliotoxinCaov-3 906825HTB-75-OvaryMUTHomozygousc.406C > Tp.Q136*[96]
SK-OV-3905959HTB-77-OvaryMUTHomozygousc.267delp.S90Pfs*33[96]
LovastatinFaDu906863HTB-43-Head and neckMUTHeterozygousc.743G > Tp.R248L[97]
interferon-αHuH7907071--LiverMUTHomozygousc.659A > Gp.Y220C[100]
miR-130b mimicsOvcar-8905991--OvaryMUTHomozygousc.376_396delp.Y126_K132del[101]
miR-214
mimics
LoVo907790--Large
intestine
Wild typeHomozygous--[102]
SW480-CCL-228-Large
intestine
MUTHeterozygousc.818G > A & 925C > Tp.R273H & P309S[102]
miR-140
inhibitor
HGC-27907055--StomachMUTHeterozygousc.455dupp.P153Afs*28[103]
BGC-823--9999992Stomachwild typeHomozygous--[103]

5. p73 Activation Drugs

Over 20 research papers have covered TAp73 activation for anti-cancer purposes. This research is summarized in Table 2. RETRA was the first p73 activator identified. It was found to block the interaction between p73 and mutant p53 in A431 and SW480 p53 mutant cells to relieve p73 transactivation function [105]. NSC59984 can induce mutant p53 protein degradation and activate p73 in SW480 and DLD1 cells [106]. Prodigiosin can disrupt mutant p53 and p73 interaction [107], and it can not only activate TAp73 but can also repress the dominant-negative isoform ΔNp73 in SW480 and DLD1 cells [107,108]. Metformin can activate TAp73 through the LKB1-AMPK axis in HCT116 cells [109], and it also represses dominant-negative isoform ΔNp63 under low glucose conditions in FaDu, H596 and H292 cells [110]. Bortezomib can promote TAp73 activation to reduce colon cancer cell viability in both p53−/− HCT116 and HT29 cells [111]. Protoporphyrin can disrupt p73/MDM2 complexes to restore p73 transcriptional activity in p53−/− HCT116 cells [112]. Nutlin 3 is an MDM2 antagonist [113], and it also disrupts p73-MDM2 binding to enhance p73 function in IEC-6 and Caco2 cells [114]. Diallyl disulfide can upregulate TAp73 and downregulate ΔNp73 expression in HeLa cells to enhance carbon ion beam–induced apoptosis [115]. Abrus agglutinin can inhibit p73 and Snail interaction to cause p73 activation in FaDu cells [116]. Cinobufagin can repress AURKA serine/threonine kinase to decrease phosphorylation levels of p53 (S315 and S392) [117], but it can promote phosphorylation levels of p73 (Y99) to induce the apoptosis in Huh-7 cells [117]. Panobinostat caused p73 and p21 upregulation in TP53−/− Saos-2 osteosarcoma cells [118]. Thymoquinone can induce p73 expression to cause cell apoptosis in a p53-deficient acute lymphoblastic leukemia (ALL) Jurkat cell line [119,120]. Thymoquinone can inhibit Itch, the E3-ubiquitin ligase of p73 leading to the upregulation of tumor suppressor p73 in Jurkat cells, MDA-MB-468 cells, and HL60 cells [121,122]. Extracts of Piper betle leaf (PBL) can promote p73 expression to induce cell cycle arrest in Hep3B cells [123]. MEK1 inhibitor (PD98059 or PD184352) can induce TAp73 and reduce dominant-negative ΔNp73 in NB4 and K562 cell lines [124]. Etoposide can induce p73 expression in HOC313 and Ca9-22 cells [125], and it can also induce p73 expression in both HCT116 p53−/− and H1299 p53−/− cells [126].
miRNA-1180 can directly target p73 3′-UTR sequences to repress p73, and miR-1180 inhibitor can induce p73 in SK-NEP-1 cells [127]. miRNA-193a-5p can directly target p73 3′-UTR sequences to repress p73 [128], miRNA-193a-5p inhibitor can induce p73 in JHU-029 and MG63 cells [128,129]. miR-647 can directly target p73 3′-UTR sequences to repress p73, and miR-647 inhibitor can induce p73 in MGC-803 cell lines [130]. miR-323 can repress the expression of p73 in PC-3 cell lines, and miR-323 inhibitor can arrest the PC-3 cell cycle and cause apoptosis with increased expression of p73 downstream gene p21 [131]. miR106b mimic can directly target Itch, the E3-ubiquitin ligase of p73, and miR106b mimic can repress Itch to accumulate p73 in K562 cells [132].

6. Influence of Interactions between p53 Family Members on p63 or p73 Activators and the Importance of Combination Treatment Strategy

Several factors can influence p63 or p73 activation in p53 mutant cells. First, a high level of dominant-negative ∆Np63 or ∆Np73 expression in cells can block the TAp63 or TAp73 activity [57,58]. Second, gain-of-function mutant p53 can become aggregated and co-aggregate with TAp63 or TAp73 [68,133]. miR-130b and miR-124 mimics not only induce TA isoform p63 in p53 mutant cells, but also repress the relative proportion of ΔN isoform p63 to resolve the first issue [101,102]. Prodigiosin, metformin, diallyl disulfide, and MEK1 inhibitors not only induce TA isoform p73 in p53 mutant cells, but also repress the relative proportion of ΔN isoform p63 or p73 [107,110,115,124,134]. Recently we found that p73 activators RETEA and NSC59984 have a poor response in aggregative p53 mutant HNSCC cells compared to non-aggregative p53 mutant cells. Furthermore, using an NAMPT inhibitor to block p53 aggregation can enhance the anti-cancer effect of p73 activator RETEA and NSC59984 in p53 gain-of-function mutant HNSCC cells [135]. This combination treatment strategy comprised of a p73 activator and a p53 aggregation inhibitor was able to resolve the second issue to active p73 in p53 gain-of-function mutants. Because there are fewer p63 activators than p73 activators (Table 1 and Table 2), whether this combination treatment strategy with a p63 activator and p53 aggregation inhibitor is workable or not in p53 gain-of-function mutants still needs further investigation. However, it is likely that co-treatment with a p53 aggregation inhibitor can improve p63 and p73 drug response in p53 gain-of-function mutants.

7. Viral Proteins and the p53 Family

Some cells contain wild-type p53, but several of these cell lines may be infected by certain viruses to express viral protein(s) that interact with p53 family members to cause their loss of function. Hela cells were infected by oncogenic HPV type 16 (HPV16) with E6 oncoproteins to induce ubiquitin-dependent proteolytic degradation of wild-type p53 [136,137]. HPV18 E6 can interact with TAp63β but not the other p63 isoforms to induce degradation of the TAp63β [138]. HPV18 E6 can also interact with TAp73α or TAp73β to reduce its transcriptional activity but has no influence on p73 stability. Diallyl disulfide can still induce p73 to mediate the cell apoptotic program in Hela cells [115]. Other oncogenic viral proteins like the LANA protein of the Kaposi sarcoma virus, the BZLF1 protein of the Epstein-Barr virus, and the papain-like protease of the nonstructural protein 3 of SARS-CoV also can cause p53 degradation [139,140,141]. In addition, the HbX protein of the Hepatitis virus B and the NS2 proteins of hepatitis virus C can interact with p53 to reduce its transcriptional activity [142,143]. But whether all these oncogenic viral proteins can influence p63 and p73 or not still remains to be addressed. This means that p63 and p73 activators may also not be such powerful anti-cancer agents in p53 mutant cells infected by certain viruses. Interestingly, HPV18 E7 can inhibit the interaction of p53 and MDM2 to stabilize p53 to increase the transcriptional activation function of p53 [144]. The NS5 proteins of the Zika virus can interact with p53 to prolong the half-life of p53 to cause cell death [145]. Therefore, some viral protein(s) can also promote wild-type p53 function. Whether p63 and p73 can be activated by these viral protein(s) or not is also unknown.

8. p53 Isoforms in p53 Mutant Cancer Cells

p53, like p63 and p73, has p53 isoforms [146,147]. Two dominate-negative isoforms ∆133p53 and ∆160p53 are frequently expressed in p53 mutant cancer cells [148,149]. Wild-type p53 does not interact with p73 [56,57], but Δ133p53α, Δ133p53β, and Δ133p53γ isoforms can interact with p73 to alter p73 downstream genes RAD51, RAD52, and LIG4 [150]. One study reported that treatment of 5-lipooxygenase (5-LOX) peptide inhibitor, YWCS, can induce p73, Δ133p53, and Δ160p53 expression to prevent neurotoxicity in pre-treatment of Aβ25–35 in SH-SY5Y cells [151], so Δ133p53 and Δ160p53 play a neuroprotective role to offset the cell apoptotic role of p73. Wild-type p53 does not interact with p63 [56,57], but the Δ133p53α isoform can interact with p63 to repress the anti-proliferative activities of TAp63β [152]. Δ133p53β, but not Δ133p53α or Δ160p53β, can aggregate to deactivate its dominate-negative effect [153]. If free-form Δ133p53β is released from aggregates, Δ133p53β can interact with TAp63 or ∆Np63 to repress transactivation or to enhance an oncogenic effect separately [153]. The p53 isoforms likely also influence p63 and p73 activator response in p53 mutant cells; this it also needs to be clarified.

9. Discussion and Concluding Remarks

There are also several p53 activators that can directly bind to mutant p53 to convert it into a wild-type like structure [154,155,156]. These drugs directly reactivate and change p53 activity. Such drugs were not emphasized in this review. For example, PRIMA-1 and its methylated form PRIMA-1Met (APR-246) can directly bind to mutant p53 to refold it as a wild-type p53-like conformation [157,158,159], because mutant p53 can associate with p63 and p73, whereas wild-type p53 cannot [56,57]. Therefore, PRIMA-1Met has also been reported to activate p63 and p73 through relieving them from mutant p53 that had been converted into wild-type like structure [160,161,162].
In conclusion, this review focused on summarizing the drugs that can activate p63 or p73 in p53 mutant cancer cells. Besides synthetic chemicals and purified natural products, synthesized small biomolecules, like microRNA mimics or inhibitors have also been used to activate p63 and p73 in p53 mutant cancer cells (Table 1 and Table 2). It has also been pointed out in this review that viral proteins and p53 isoforms may also influence the drug response of p63 and p73 activators in p53 mutant cancers.

Author Contributions

B.-H.C. wrote the manuscript. Y.-C.H. (Yun-Chien Hsu), F.-Y.Y. and Y.-R.L. searched database and draw tables. R.-Y.L. (Rui-Yu Lu) and S.-J.Y. draw figures. M.-H.W., Y.-Z.T., Y.-C.L., Z.-Y.B., R.-Y.L. (Ruo-Yu Liao) and W.-H.K. searched literatures. J.-F.S., Y.-C.S., Y.-C.H. (Yi-Chiang Hsu), C.-T.H., C.-F.L. and C.-C.C. conducted review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the I-Shou University and E-Da Medical Consortium Legal Person Research Cooperation Project [ISU-110-IUC-08 to B.-H.C., C.-F.L., Y.-C.H. and C.-C.C.] and [ISU-111-IUC-07 to B.-H.C., J.-F.S., C.-F.L. and Y.-C.S.]. This work was also partially supported by the Medical Student Research and Development Scholarship Program [EDAHS110001 to B.-H.C., M.-H.W. and C.-C.C.].

Acknowledgments

We would like to acknowledge the lab maintenance given by the Basic Medical Core Laboratory, I-Shou University College of Medicine. We thank the p53 family lab in Taiwan past members (https://p53familylabtw.weebly.com; accessed date (12 May 2022) Wei-Chen Wu (Department of Physical Therapy, I-Shou University; date of visit 30 September 2021–31 March 2022) for assistance. We give special thanks to Reiji Kannagi (Institute of Biomedical Sciences, Academia Sinica) and Jang-Yi Chen (Department of Biology and Anatomy, National Defense Medical Center) for materials support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scoumanne, A.; Harms, K.L.; Chen, X. Structural basis for gene activation by p53 family members. Cancer Biol. Ther. 2005, 4, 1178–1185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Harms, K.L.; Chen, X. The functional domains in p53 family proteins exhibit both common and distinct properties. Cell Death Differ. 2006, 13, 890–897. [Google Scholar] [CrossRef] [PubMed]
  3. Zawacka-Pankau, J.E. The Role of p53 Family in Cancer. Cancers 2022, 14, 823. [Google Scholar] [CrossRef] [PubMed]
  4. Levrero, M.; De Laurenzi, V.; Costanzo, A.; Gong, J.; Wang, J.Y.; Melino, G. The p53/p63/p73 family of transcription factors: Overlapping and distinct functions. J. Cell Sci. 2000, 113 Pt 10, 1661–1670. [Google Scholar] [CrossRef]
  5. Flores, E.R.; Sengupta, S.; Miller, J.B.; Newman, J.J.; Bronson, R.; Crowley, D.; Yang, A.; McKeon, F.; Jacks, T. Tumor predisposition in mice mutant for p63 and p73: Evidence for broader tumor suppressor functions for the p53 family. Cancer Cell 2005, 7, 363–373. [Google Scholar] [CrossRef] [Green Version]
  6. Hollstein, M.; Sidransky, D.; Vogelstein, B.; Harris, C.C. p53 mutations in human cancers. Science 1991, 253, 49–53. [Google Scholar] [CrossRef] [Green Version]
  7. Olivier, M.; Hollstein, M.; Hainaut, P. TP53 mutations in human cancers: Origins, consequences, and clinical use. Cold Spring Harb. Perspect. Biol. 2010, 2, a001008. [Google Scholar] [CrossRef] [Green Version]
  8. Hagiwara, K.; McMenamin, M.G.; Miura, K.; Harris, C.C. Mutational analysis of the p63/p73L/p51/p40/CUSP/KET gene in human cancer cell lines using intronic primers. Cancer Res. 1999, 59, 4165–4169. [Google Scholar]
  9. Schwartz, D.I.; Lindor, N.M.; Walsh-Vockley, C.; Roche, P.C.; Mai, M.; Smith, D.I.; Liu, W.; Couch, F.J. p73 mutations are not detected in sporadic and hereditary breast cancer. Breast Cancer Res. Treat. 1999, 58, 25–29. [Google Scholar] [CrossRef]
  10. Han, S.; Semba, S.; Abe, T.; Makino, N.; Furukawa, T.; Fukushige, S.; Takahashi, H.; Sakurada, A.; Sato, M.; Shiiba, K.; et al. Infrequent somatic mutations of the p73 gene in various human cancers. Eur. J. Surg. Oncol. 1999, 25, 194–198. [Google Scholar] [CrossRef]
  11. Lu, X.; Liu, D.P.; Xu, Y. The gain of function of p53 cancer mutant in promoting mammary tumorigenesis. Oncogene 2013, 32, 2900–2906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Pitolli, C.; Wang, Y.; Mancini, M.; Shi, Y.; Melino, G.; Amelio, I. Do Mutations Turn p53 into an Oncogene? Int. J. Mol. Sci. 2019, 20, 6241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Soussi, T. The history of p53. A perfect example of the drawbacks of scientific paradigms. EMBO Rep. 2010, 11, 822–826. [Google Scholar] [CrossRef] [PubMed]
  14. Fischer, M. Census and evaluation of p53 target genes. Oncogene 2017, 36, 3943–3956. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Cai, B.H.; Chao, C.F.; Huang, H.C.; Lee, H.Y.; Kannagi, R.; Chen, J.Y. Roles of p53 Family Structure and Function in Non-Canonical Response Element Binding and Activation. Int. J. Mol. Sci. 2019, 20, 3681. [Google Scholar] [CrossRef] [Green Version]
  16. Donehower, L.A.; Harvey, M.; Slagle, B.L.; McArthur, M.J.; Montgomery, C.A.; Butel, J.S.; Bradley, A. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 1992, 356, 215–221. [Google Scholar] [CrossRef]
  17. Nigro, J.M.; Baker, S.J.; Preisinger, A.C.; Jessup, J.M.; Hostetter, R.; Cleary, K.; Bigner, S.H.; Davidson, N.; Baylin, S.; Devilee, P. Mutations in the p53 gene occur in diverse human tumour types. Nature 1989, 342, 705–708. [Google Scholar] [CrossRef]
  18. Tate, J.G.; Bamford, S.; Jubb, H.C.; Sondka, Z.; Beare, D.M.; Bindal, N.; Boutselakis, H.; Cole, C.G.; Creatore, C.; Dawson, E.; et al. COSMIC: The Catalogue of Somatic Mutations In Cancer. Nucleic Acids Res. 2019, 47, D941–D947. [Google Scholar] [CrossRef] [Green Version]
  19. Baugh, E.H.; Ke, H.; Levine, A.J.; Bonneau, R.A.; Chan, C.S. Why are there hotspot mutations in the TP53 gene in human cancers? Cell Death Differ. 2018, 25, 154–160. [Google Scholar] [CrossRef]
  20. Malkin, D. Li-fraumeni syndrome. Genes Cancer 2011, 2, 475–484. [Google Scholar] [CrossRef]
  21. Adam, M.P.; Ardinger, H.H.; Pagon, R.A.; Wallace, S.E.; Bean, L.J.H.; Gripp, K.W.; Mirzaa, G.M.; Amemiya, A. GeneReviews; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  22. Hisada, M.; Garber, J.E.; Fung, C.Y.; Fraumeni, J.F.; Li, F.P. Multiple primary cancers in families with Li-Fraumeni syndrome. J. Natl. Cancer Inst. 1998, 90, 606–611. [Google Scholar] [CrossRef] [PubMed]
  23. Lepre, M.G.; Omar, S.I.; Grasso, G.; Morbiducci, U.; Deriu, M.A.; Tuszynski, J.A. Insights into the Effect of the G245S Single Point Mutation on the Structure of p53 and the Binding of the Protein to DNA. Molecules 2017, 22, 1358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Meneghetti, B.V.; Wilson, R.; Dias, C.K.; Cadore, N.A.; Klamt, F.; Zaha, A.; Ferreira, H.B.; Monteiro, K.M. p53 mutants G245S and R337H associated with the Li-Fraumeni syndrome regulate distinct metabolic pathways. Biochimie 2022, 198, 141–154. [Google Scholar] [CrossRef] [PubMed]
  25. Ribeiro, R.C.; Sandrini, F.; Figueiredo, B.; Zambetti, G.P.; Michalkiewicz, E.; Lafferty, A.R.; DeLacerda, L.; Rabin, M.; Cadwell, C.; Sampaio, G.; et al. An inherited p53 mutation that contributes in a tissue-specific manner to pediatric adrenal cortical carcinoma. Proc. Natl. Acad. Sci. USA 2001, 98, 9330–9335. [Google Scholar] [CrossRef] [Green Version]
  26. Achatz, M.I.; Olivier, M.; Le Calvez, F.; Martel-Planche, G.; Lopes, A.; Rossi, B.M.; Ashton-Prolla, P.; Giugliani, R.; Palmero, E.I.; Vargas, F.R.; et al. The TP53 mutation, R337H, is associated with Li-Fraumeni and Li-Fraumeni-like syndromes in Brazilian families. Cancer Lett. 2007, 245, 96–102. [Google Scholar] [CrossRef]
  27. Yang, A.; Kaghad, M.; Wang, Y.; Gillett, E.; Fleming, M.D.; Dötsch, V.; Andrews, N.C.; Caput, D.; McKeon, F. p63, a p53 homolog at 3q27-29, encodes multiple products with transactivating, death-inducing, and dominant-negative activities. Mol. Cell 1998, 2, 305–316. [Google Scholar] [CrossRef]
  28. Marshall, C.B.; Beeler, J.S.; Lehmann, B.D.; Gonzalez-Ericsson, P.; Sanchez, V.; Sanders, M.E.; Boyd, K.L.; Pietenpol, J.A. Tissue-specific expression of p73 and p63 isoforms in human tissues. Cell Death Dis. 2021, 12, 745. [Google Scholar] [CrossRef]
  29. Celli, J.; Duijf, P.; Hamel, B.C.; Bamshad, M.; Kramer, B.; Smits, A.P.; Newbury-Ecob, R.; Hennekam, R.C.; Van Buggenhout, G.; van Haeringen, A.; et al. Heterozygous germline mutations in the p53 homolog p63 are the cause of EEC syndrome. Cell 1999, 99, 143–153. [Google Scholar] [CrossRef] [Green Version]
  30. McGrath, J.A.; Duijf, P.H.; Doetsch, V.; Irvine, A.D.; de Waal, R.; Vanmolkot, K.R.; Wessagowit, V.; Kelly, A.; Atherton, D.J.; Griffiths, W.A.; et al. Hay-Wells syndrome is caused by heterozygous missense mutations in the SAM domain of p63. Hum. Mol. Genet. 2001, 10, 221–229. [Google Scholar] [CrossRef] [Green Version]
  31. Julapalli, M.R.; Scher, R.K.; Sybert, V.P.; Siegfried, E.C.; Bree, A.F. Dermatologic findings of ankyloblepharon-ectodermal defects-cleft lip/palate (AEC) syndrome. Am. J. Med. Genet. A 2009, 149A, 1900–1906. [Google Scholar] [CrossRef]
  32. van Bokhoven, H.; Hamel, B.C.; Bamshad, M.; Sangiorgi, E.; Gurrieri, F.; Duijf, P.H.; Vanmolkot, K.R.; van Beusekom, E.; van Beersum, S.E.; Celli, J.; et al. p63 Gene mutations in eec syndrome, limb-mammary syndrome, and isolated split hand-split foot malformation suggest a genotype-phenotype correlation. Am. J. Hum. Genet. 2001, 69, 481–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Mills, A.A.; Zheng, B.; Wang, X.J.; Vogel, H.; Roop, D.R.; Bradley, A. p63 is a p53 homologue required for limb and epidermal morphogenesis. Nature 1999, 398, 708–713. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, A.; Schweitzer, R.; Sun, D.; Kaghad, M.; Walker, N.; Bronson, R.T.; Tabin, C.; Sharpe, A.; Caput, D.; Crum, C.; et al. p63 is essential for regenerative proliferation in limb, craniofacial and epithelial development. Nature 1999, 398, 714–718. [Google Scholar] [CrossRef] [PubMed]
  35. Kaghad, M.; Bonnet, H.; Yang, A.; Creancier, L.; Biscan, J.C.; Valent, A.; Minty, A.; Chalon, P.; Lelias, J.M.; Dumont, X.; et al. Monoallelically expressed gene related to p53 at 1p36, a region frequently deleted in neuroblastoma and other human cancers. Cell 1997, 90, 809–819. [Google Scholar] [CrossRef] [Green Version]
  36. Conforti, F.; Yang, A.L.; Agostini, M.; Rufini, A.; Tucci, P.; Nicklison-Chirou, M.V.; Grespi, F.; Velletri, T.; Knight, R.A.; Melino, G.; et al. Relative expression of TAp73 and ΔNp73 isoforms. Aging 2012, 4, 202–205. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, A.; Walker, N.; Bronson, R.; Kaghad, M.; Oosterwegel, M.; Bonnin, J.; Vagner, C.; Bonnet, H.; Dikkes, P.; Sharpe, A.; et al. p73-deficient mice have neurological, pheromonal and inflammatory defects but lack spontaneous tumours. Nature 2000, 404, 99–103. [Google Scholar] [CrossRef]
  38. Lynch, M. Rate, molecular spectrum, and consequences of human mutation. Proc. Natl. Acad. Sci. USA 2010, 107, 961–968. [Google Scholar] [CrossRef] [Green Version]
  39. Hug, N.; Longman, D.; Cáceres, J.F. Mechanism and regulation of the nonsense-mediated decay pathway. Nucleic Acids Res. 2016, 44, 1483–1495. [Google Scholar] [CrossRef] [Green Version]
  40. Floquet, C.; Deforges, J.; Rousset, J.P.; Bidou, L. Rescue of non-sense mutated p53 tumor suppressor gene by aminoglycosides. Nucleic Acids Res. 2011, 39, 3350–3362. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, M.; Heldin, A.; Palomar-Siles, M.; Öhlin, S.; Bykov, V.J.N.; Wiman, K.G. Synergistic Rescue of Nonsense Mutant Tumor Suppressor p53 by Combination Treatment with Aminoglycosides and Mdm2 Inhibitors. Front. Oncol. 2017, 7, 323. [Google Scholar] [CrossRef] [Green Version]
  42. Trzaska, C.; Amand, S.; Bailly, C.; Leroy, C.; Marchand, V.; Duvernois-Berthet, E.; Saliou, J.M.; Benhabiles, H.; Werkmeister, E.; Chassat, T.; et al. 2,6-Diaminopurine as a highly potent corrector of UGA nonsense mutations. Nat. Commun. 2020, 11, 1509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ferguson, M.W.; Gerak, C.A.N.; Chow, C.C.T.; Rastelli, E.J.; Elmore, K.E.; Stahl, F.; Hosseini-Farahabadi, S.; Baradaran-Heravi, A.; Coltart, D.M.; Roberge, M. The antimalarial drug mefloquine enhances TP53 premature termination codon readthrough by aminoglycoside G418. PLoS ONE 2019, 14, e0216423. [Google Scholar] [CrossRef] [PubMed]
  44. Baradaran-Heravi, A.; Balgi, A.D.; Zimmerman, C.; Choi, K.; Shidmoossavee, F.S.; Tan, J.S.; Bergeaud, C.; Krause, A.; Flibotte, S.; Shimizu, Y.; et al. Novel small molecules potentiate premature termination codon readthrough by aminoglycosides. Nucleic Acids Res. 2016, 44, 6583–6598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kerem, E.; Hirawat, S.; Armoni, S.; Yaakov, Y.; Shoseyov, D.; Cohen, M.; Nissim-Rafinia, M.; Blau, H.; Rivlin, J.; Aviram, M.; et al. Effectiveness of PTC124 treatment of cystic fibrosis caused by nonsense mutations: A prospective phase II trial. Lancet 2008, 372, 719–727. [Google Scholar] [CrossRef]
  46. Kerem, E.; Konstan, M.W.; De Boeck, K.; Accurso, F.J.; Sermet-Gaudelus, I.; Wilschanski, M.; Elborn, J.S.; Melotti, P.; Bronsveld, I.; Fajac, I.; et al. Ataluren for the treatment of nonsense-mutation cystic fibrosis: A randomised, double-blind, placebo-controlled phase 3 trial. Lancet Respir. Med. 2014, 2, 539–547. [Google Scholar] [CrossRef] [Green Version]
  47. Roy, B.; Friesen, W.J.; Tomizawa, Y.; Leszyk, J.D.; Zhuo, J.; Johnson, B.; Dakka, J.; Trotta, C.R.; Xue, X.; Mutyam, V.; et al. Ataluren stimulates ribosomal selection of near-cognate tRNAs to promote nonsense suppression. Proc. Natl. Acad. Sci. USA 2016, 113, 12508–12513. [Google Scholar] [CrossRef] [Green Version]
  48. Baradaran-Heravi, A.; Balgi, A.D.; Hosseini-Farahabadi, S.; Choi, K.; Has, C.; Roberge, M. Effect of small molecule eRF3 degraders on premature termination codon readthrough. Nucleic Acids Res. 2021, 49, 3692–3708. [Google Scholar] [CrossRef]
  49. Nakamura, Y.; Futamura, M.; Kamino, H.; Yoshida, K.; Arakawa, H. Identification of p53-46F as a super p53 with an enhanced ability to induce p53-dependent apoptosis. Cancer Sci. 2006, 97, 633–641. [Google Scholar] [CrossRef]
  50. Emamzadah, S.; Tropia, L.; Vincenti, I.; Falquet, B.; Halazonetis, T.D. Reversal of the DNA-binding-induced loop L1 conformational switch in an engineered human p53 protein. J. Mol. Biol. 2014, 426, 936–944. [Google Scholar] [CrossRef] [Green Version]
  51. 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] [Green Version]
  52. Nichols, N.M.; Matthews, K.S. Human p53 phosphorylation mimic, S392E, increases nonspecific DNA affinity and thermal stability. Biochemistry 2002, 41, 170–178. [Google Scholar] [CrossRef] [PubMed]
  53. Blagosklonny, M.V. Loss of function and p53 protein stabilization. Oncogene 1997, 15, 1889–1893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. 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]
  55. Vousden, K.H.; Lu, X. Live or let die: The cell’s response to p53. Nat. Rev. Cancer 2002, 2, 594–604. [Google Scholar] [CrossRef] [Green Version]
  56. Davison, T.S.; Vagner, C.; Kaghad, M.; Ayed, A.; Caput, D.; Arrowsmith, C.H. p73 and p63 are homotetramers capable of weak heterotypic interactions with each other but not with p53. J. Biol. Chem. 1999, 274, 18709–18714. [Google Scholar] [CrossRef] [Green Version]
  57. Gaiddon, C.; Lokshin, M.; Ahn, J.; Zhang, T.; Prives, C. A subset of tumor-derived mutant forms of p53 down-regulate p63 and p73 through a direct interaction with the p53 core domain. Mol. Cell Biol. 2001, 21, 1874–1887. [Google Scholar] [CrossRef] [Green Version]
  58. Chan, W.M.; Siu, W.Y.; Lau, A.; Poon, R.Y. How many mutant p53 molecules are needed to inactivate a tetramer? Mol. Cell Biol. 2004, 24, 3536–3551. [Google Scholar] [CrossRef] [Green Version]
  59. 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]
  60. Donzelli, S.; Biagioni, F.; Fausti, F.; Strano, S.; Fontemaggi, G.; Blandino, G. Oncogenomic Approaches in Exploring Gain of Function of Mutant p53. Curr. Genom. 2008, 9, 200–207. [Google Scholar] [CrossRef]
  61. Liebl, M.C.; Hofmann, T.G. The Role of p53 Signaling in Colorectal Cancer. Cancers 2021, 13, 2125. [Google Scholar] [CrossRef]
  62. Liao, P.; Zeng, S.X.; Zhou, X.; Chen, T.; Zhou, F.; Cao, B.; Jung, J.H.; Del Sal, G.; Luo, S.; Lu, H. Mutant p53 Gains Its Function via c-Myc Activation upon CDK4 Phosphorylation at Serine 249 and Consequent PIN1 Binding. Mol. Cell 2017, 68, 1134–1146.e1136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Stindt, M.H.; Muller, P.A.; Ludwig, R.L.; Kehrloesser, S.; Dötsch, V.; Vousden, K.H. Functional interplay between MDM2, p63/p73 and mutant p53. Oncogene 2015, 34, 4300–4310. [Google Scholar] [CrossRef] [PubMed]
  64. Yang-Hartwich, Y.; Soteras, M.G.; Lin, Z.P.; Holmberg, J.; Sumi, N.; Craveiro, V.; Liang, M.; Romanoff, E.; Bingham, J.; Garofalo, F.; et al. p53 protein aggregation promotes platinum resistance in ovarian cancer. Oncogene 2015, 34, 3605–3616. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, S.; An, S.S. Role of p53 isoforms and aggregations in cancer. Medicine 2016, 95, e3993. [Google Scholar] [CrossRef]
  66. De Smet, F.; Saiz Rubio, M.; Hompes, D.; Naus, E.; De Baets, G.; Langenberg, T.; Hipp, M.S.; Houben, B.; Claes, F.; Charbonneau, S.; et al. Nuclear inclusion bodies of mutant and wild-type p53 in cancer: A hallmark of p53 inactivation and proteostasis remodelling by p53 aggregation. J. Pathol. 2017, 242, 24–38. [Google Scholar] [CrossRef]
  67. Kanapathipillai, M. Treating p53 Mutant Aggregation-Associated Cancer. Cancers 2018, 10, 154. [Google Scholar] [CrossRef] [Green Version]
  68. Xu, J.; Reumers, J.; Couceiro, J.R.; De Smet, F.; Gallardo, R.; Rudyak, S.; Cornelis, A.; Rozenski, J.; Zwolinska, A.; Marine, J.C.; et al. Gain of function of mutant p53 by coaggregation with multiple tumor suppressors. Nat. Chem. Biol. 2011, 7, 285–295. [Google Scholar] [CrossRef] [Green Version]
  69. de Oliveira, G.A.P.; Petronilho, E.C.; Pedrote, M.M.; Marques, M.A.; Vieira, T.C.R.G.; Cino, E.A.; Silva, J.L. The Status of p53 Oligomeric and Aggregation States in Cancer. Biomolecules 2020, 10, 548. [Google Scholar] [CrossRef] [Green Version]
  70. Pedrote, M.M.; Motta, M.F.; Ferretti, G.D.S.; Norberto, D.R.; Spohr, T.C.L.S.; Lima, F.R.S.; Gratton, E.; Silva, J.L.; de Oliveira, G.A.P. Oncogenic Gain of Function in Glioblastoma Is Linked to Mutant p53 Amyloid Oligomers. iScience 2020, 23, 100820. [Google Scholar] [CrossRef]
  71. Boysen, M.; Kityk, R.; Mayer, M.P. Hsp70- and Hsp90-Mediated Regulation of the Conformation of p53 DNA Binding Domain and p53 Cancer Variants. Mol. Cell 2019, 74, 831–843.e4. [Google Scholar] [CrossRef]
  72. Wu, H.; Dyson, H.J. Aggregation of zinc-free p53 is inhibited by Hsp90 but not other chaperones. Protein Sci. 2019, 28, 2020–2023. [Google Scholar] [CrossRef] [PubMed]
  73. Westerheide, S.D.; Anckar, J.; Stevens, S.M.; Sistonen, L.; Morimoto, R.I. Stress-inducible regulation of heat shock factor 1 by the deacetylase SIRT1. Science 2009, 323, 1063–1066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Zhang, T.; Berrocal, J.G.; Frizzell, K.M.; Gamble, M.J.; DuMond, M.E.; Krishnakumar, R.; Yang, T.; Sauve, A.A.; Kraus, W.L. Enzymes in the NAD+ salvage pathway regulate SIRT1 activity at target gene promoters. J. Biol. Chem. 2009, 284, 20408–20417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ganci, F.; Pulito, C.; Valsoni, S.; Sacconi, A.; Turco, C.; Vahabi, M.; Manciocco, V.; Mazza, E.M.C.; Meens, J.; Karamboulas, C.; et al. PI3K Inhibitors Curtail MYC-Dependent Mutant p53 Gain-of-Function in Head and Neck Squamous Cell Carcinoma. Clin. Cancer Res. 2020, 26, 2956–2971. [Google Scholar] [CrossRef] [Green Version]
  76. Menssen, A.; Hydbring, P.; Kapelle, K.; Vervoorts, J.; Diebold, J.; Lüscher, B.; Larsson, L.G.; Hermeking, H. The c-MYC oncoprotein, the NAMPT enzyme, the SIRT1-inhibitor DBC1, and the SIRT1 deacetylase form a positive feedback loop. Proc. Natl. Acad. Sci. USA 2012, 109, E187–E196. [Google Scholar] [CrossRef] [Green Version]
  77. 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]
  78. Cai, B.H.; Chen, J.Y.; Lu, M.H.; Chang, L.T.; Lin, H.C.; Chang, Y.M.; Chao, C.F. Functional four-base A/T gap core sequence CATTAG of P53 response elements specifically bound tetrameric P53 differently than two-base A/T gap core sequence CATG bound both dimeric and tetrameric P53. Nucleic Acids Res. 2009, 37, 1984–1990. [Google Scholar] [CrossRef]
  79. Phan, L.; Chou, P.C.; Velazquez-Torres, G.; Samudio, I.; Parreno, K.; Huang, Y.; Tseng, C.; Vu, T.; Gully, C.; Su, C.H.; et al. The cell cycle regulator 14-3-3σ opposes and reverses cancer metabolic reprogramming. Nat. Commun. 2015, 6, 7530. [Google Scholar] [CrossRef] [Green Version]
  80. Petr, M.; Helma, R.; Polášková, A.; Krejčí, A.; Dvořáková, Z.; Kejnovská, I.; Navrátilová, L.; Adámik, M.; Vorlíčková, M.; Brázdová, M. Wild-type p53 binds to MYC promoter G-quadruplex. Biosci. Rep. 2016, 36, e00397. [Google Scholar] [CrossRef] [Green Version]
  81. He, Y.H.; Yeh, M.H.; Chen, H.F.; Wang, T.S.; Wong, R.H.; Wei, Y.L.; Huynh, T.K.; Hu, D.W.; Cheng, F.J.; Chen, J.Y.; et al. ERα determines the chemo-resistant function of mutant p53 involving the switch between lincRNA-p21 and DDB2 expressions. Mol. Ther. Nucleic Acids 2021, 25, 536–553. [Google Scholar] [CrossRef]
  82. Jin, S.; Yang, X.; Li, J.; Yang, W.; Ma, H.; Zhang, Z. p53-targeted lincRNA-p21 acts as a tumor suppressor by inhibiting JAK2/STAT3 signaling pathways in head and neck squamous cell carcinoma. Mol. Cancer 2019, 18, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Madamanchi, N.R.; Li, S.; Patterson, C.; Runge, M.S. Reactive oxygen species regulate heat-shock protein 70 via the JAK/STAT pathway. Arter. Thromb. Vasc. Biol. 2001, 21, 321–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Jego, G.; Hermetet, F.; Girodon, F.; Garrido, C. Chaperoning STAT3/5 by Heat Shock Proteins: Interest of Their Targeting in Cancer Therapy. Cancers 2019, 12, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zhang, N.; Wei, X.; Xu, L. miR-150 promotes the proliferation of lung cancer cells by targeting P53. FEBS Lett. 2013, 587, 2346–2351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wang, D.T.; Ma, Z.L.; Li, Y.L.; Wang, Y.Q.; Zhao, B.T.; Wei, J.L.; Qi, X.; Zhao, X.T.; Jin, Y.X. miR-150, p53 protein and relevant miRNAs consist of a regulatory network in NSCLC tumorigenesis. Oncol. Rep. 2013, 30, 492–498. [Google Scholar] [CrossRef] [Green Version]
  87. Liu, F.; Di Wang, X. miR-150-5p represses TP53 tumor suppressor gene to promote proliferation of colon adenocarcinoma. Sci. Rep. 2019, 9, 6740. [Google Scholar] [CrossRef]
  88. Meng, X.; Sun, W.; Yu, J.; Zhou, Y.; Gu, Y.; Han, J.; Zhou, L.; Jiang, X.; Wang, C. LINC00460-miR-149-5p/miR-150-5p-Mutant p53 Feedback Loop Promotes Oxaliplatin Resistance in Colorectal Cancer. Mol. Ther. Nucleic Acids 2020, 22, 1004–1015. [Google Scholar] [CrossRef]
  89. McPherson, L.A.; Loktev, A.V.; Weigel, R.J. Tumor suppressor activity of AP2alpha mediated through a direct interaction with p53. J. Biol. Chem. 2002, 277, 45028–45033. [Google Scholar] [CrossRef] [Green Version]
  90. Marreiros, A.; Dudgeon, K.; Dao, V.; Grimm, M.O.; Czolij, R.; Crossley, M.; Jackson, P. KAI1 promoter activity is dependent on p53, junB and AP2: Evidence for a possible mechanism underlying loss of KAI1 expression in cancer cells. Oncogene 2005, 24, 637–649. [Google Scholar] [CrossRef] [Green Version]
  91. Tsai, Y.C.; Weissman, A.M. Dissecting the diverse functions of the metastasis suppressor CD82/KAI1. FEBS Lett. 2011, 585, 3166–3173. [Google Scholar] [CrossRef] [Green Version]
  92. Cai, B.H.; Wu, P.H.; Chou, C.K.; Huang, H.C.; Chao, C.C.; Chung, H.Y.; Lee, H.Y.; Chen, J.Y.; Kannagi, R. Synergistic activation of the NEU4 promoter by p73 and AP2 in colon cancer cells. Sci. Rep. 2019, 9, 950. [Google Scholar] [CrossRef] [PubMed]
  93. Quan, X.; Zhao, M.; Yang, X.; Zhu, Y.; Tian, X. AP2γ mediated downregulation of lncRNA LINC00511 as a ceRNA suppresses trophoblast invasion by regulating miR-29b-3p/Cyr61 axis. Biomed. Pharmacother. 2019, 120, 109269. [Google Scholar] [CrossRef] [PubMed]
  94. Shi, G.; Cheng, Y.; Zhang, Y.; Guo, R.; Li, S.; Hong, X. Long non-coding RNA LINC00511/miR-150/MMP13 axis promotes breast cancer proliferation, migration and invasion. Biochim. Biophys. Acta Mol. Basis Dis. 2021, 1867, 165957. [Google Scholar] [CrossRef] [PubMed]
  95. Stabach, P.R.; Thiyagarajan, M.M.; Woodfield, G.W.; Weigel, R.J. AP2alpha alters the transcriptional activity and stability of p53. Oncogene 2006, 25, 2148–2159. [Google Scholar] [CrossRef] [Green Version]
  96. Park, G.B.; Jeong, J.Y.; Kim, D. Gliotoxin Enhances Autophagic Cell Death via the DAPK1-TAp63 Signaling Pathway in Paclitaxel-Resistant Ovarian Cancer Cells. Mar. Drugs 2019, 17, 412. [Google Scholar] [CrossRef] [Green Version]
  97. Yen, C.S.; Chen, J.C.; Chang, Y.F.; Hsu, Y.F.; Chiu, P.T.; Shiue, C.; Chuang, Y.F.; Ou, G.; Hsu, M.J. Lovastatin causes FaDu hypopharyngeal carcinoma cell death via AMPK-p63-survivin signaling cascade. Sci. Rep. 2016, 6, 25082. [Google Scholar] [CrossRef] [Green Version]
  98. Pyati, U.J.; Gjini, E.; Carbonneau, S.; Lee, J.S.; Guo, F.; Jette, C.A.; Kelsell, D.P.; Look, A.T. p63 mediates an apoptotic response to pharmacological and disease-related ER stress in the developing epidermis. Dev. Cell 2011, 21, 492–505. [Google Scholar] [CrossRef] [Green Version]
  99. Cai, B.H.; Chao, C.F.; Lin, H.C.; Huang, H.Y.; Kannagi, R.; Chen, J.Y. A/T gap tolerance in the core sequence and flanking sequence requirements of non-canonical p53 response elements. J. Biochem. 2016, 159, 563–572. [Google Scholar] [CrossRef]
  100. Maeda, S.; Wada, H.; Naito, Y.; Nagano, H.; Simmons, S.; Kagawa, Y.; Naito, A.; Kikuta, J.; Ishii, T.; Tomimaru, Y.; et al. Interferon-α acts on the S/G2/M phases to induce apoptosis in the G1 phase of an IFNAR2-expressing hepatocellular carcinoma cell line. J. Biol. Chem. 2014, 289, 23786–23795. [Google Scholar] [CrossRef] [Green Version]
  101. Gunaratne, P.H.; Pan, Y.; Rao, A.K.; Lin, C.; Hernandez-Herrera, A.; Liang, K.; Rait, A.S.; Venkatanarayan, A.; Benham, A.L.; Rubab, F.; et al. Activating p53 family member TAp63: A novel therapeutic strategy for targeting p53-altered tumors. Cancer 2019, 125, 2409–2422. [Google Scholar] [CrossRef] [Green Version]
  102. Liu, K.; Yao, H.; Lei, S.; Xiong, L.; Qi, H.; Qian, K.; Liu, J.; Wang, P.; Zhao, H. The miR-124-p63 feedback loop modulates colorectal cancer growth. Oncotarget 2017, 8, 29101–29115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Lu, C.; Yang, Y.; Ma, S. A functional Variant (Rs35592567) in TP63 at 3q28 is Associated with Gastric Cancer Risk via Modifying its Regulation by MicroRNA-140. Cell Physiol. Biochem. 2018, 47, 235–244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. de Andrade, K.C.; Lee, E.E.; Tookmanian, E.M.; Kesserwan, C.A.; Manfredi, J.J.; Hatton, J.N.; Loukissas, J.K.; Zavadil, J.; Zhou, L.; Olivier, M.; et al. The TP53 Database: Transition from the International Agency for Research on Cancer to the US National Cancer Institute. Cell Death Differ. 2022, 29, 1071–1073. [Google Scholar] [CrossRef] [PubMed]
  105. Kravchenko, J.E.; Ilyinskaya, G.V.; Komarov, P.G.; Agapova, L.S.; Kochetkov, D.V.; Strom, E.; Frolova, E.I.; Kovriga, I.; Gudkov, A.V.; Feinstein, E.; et al. Small-molecule RETRA suppresses mutant p53-bearing cancer cells through a p73-dependent salvage pathway. Proc. Natl. Acad. Sci. USA 2008, 105, 6302–6307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. 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] [Green Version]
  107. Hong, B.; Prabhu, V.V.; Zhang, S.; van den Heuvel, A.P.; Dicker, D.T.; Kopelovich, L.; El-Deiry, W.S. Prodigiosin rescues deficient p53 signaling and antitumor effects via upregulating p73 and disrupting its interaction with mutant p53. Cancer Res. 2014, 74, 1153–1165. [Google Scholar] [CrossRef] [Green Version]
  108. Prabhu, V.V.; Hong, B.; Allen, J.E.; Zhang, S.; Lulla, A.R.; Dicker, D.T.; El-Deiry, W.S. Small-Molecule Prodigiosin Restores p53 Tumor Suppressor Activity in Chemoresistant Colorectal Cancer Stem Cells via c-Jun-Mediated ΔNp73 Inhibition and p73 Activation. Cancer Res. 2016, 76, 1989–1999. [Google Scholar] [CrossRef] [Green Version]
  109. Adamovich, Y.; Adler, J.; Meltser, V.; Reuven, N.; Shaul, Y. AMPK couples p73 with p53 in cell fate decision. Cell Death Differ. 2014, 21, 1451–1459. [Google Scholar] [CrossRef] [Green Version]
  110. Yi, Y.; Chen, D.; Ao, J.; Sun, S.; Wu, M.; Li, X.; Bergholz, J.; Zhang, Y.; Xiao, Z.X. Metformin Promotes AMP-activated Protein Kinase-independent Suppression of ΔNp63α Protein Expression and Inhibits Cancer Cell Viability. J. Biol. Chem. 2017, 292, 5253–5261. [Google Scholar] [CrossRef] [Green Version]
  111. Dabiri, Y.; Kalman, S.; Gürth, C.M.; Kim, J.Y.; Mayer, V.; Cheng, X. The essential role of TAp73 in bortezomib-induced apoptosis in p53-deficient colorectal cancer cells. Sci. Rep. 2017, 7, 5423. [Google Scholar] [CrossRef]
  112. Sznarkowska, A.; Maleńczyk, K.; Kadziński, L.; Bielawski, K.P.; Banecki, B.; Zawacka-Pankau, J. Targeting of p53 and its homolog p73 by protoporphyrin IX. FEBS Lett. 2011, 585, 255–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Lau, L.M.; Nugent, J.K.; Zhao, X.; Irwin, M.S. HDM2 antagonist Nutlin-3 disrupts p73-HDM2 binding and enhances p73 function. Oncogene 2008, 27, 997–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Ray, R.M.; Bhattacharya, S.; Johnson, L.R. Mdm2 inhibition induces apoptosis in p53 deficient human colon cancer cells by activating p73- and E2F1-mediated expression of PUMA and Siva-1. Apoptosis 2011, 16, 35–44. [Google Scholar] [CrossRef] [PubMed]
  115. Di, C.; Sun, C.; Li, H.; Si, J.; Zhang, H.; Han, L.; Zhao, Q.; Liu, Y.; Liu, B.; Miao, G.; et al. Diallyl disulfide enhances carbon ion beams-induced apoptotic cell death in cervical cancer cells through regulating Tap73/ΔNp73. Cell Cycle 2015, 14, 3725–3733. [Google Scholar] [CrossRef] [Green Version]
  116. Sinha, N.; Meher, B.R.; Naik, P.P.; Panda, P.K.; Mukhapadhyay, S.; Maiti, T.K.; Bhutia, S.K. p73 induction by Abrus agglutinin facilitates Snail ubiquitination to inhibit epithelial to mesenchymal transition in oral cancer. Phytomedicine 2019, 55, 179–190. [Google Scholar] [CrossRef]
  117. Zhao, L.; Fu, L.; Xu, Z.; Fan, R.; Xu, R.; Fu, R.; Zou, S.; Wang, C.; Zhang, Y.; Wang, J.; et al. The anticancer effects of cinobufagin on hepatocellular carcinoma Huh-7 cells are associated with activation of the p73 signaling pathway. Mol. Med. Rep. 2019, 19, 4119–4128. [Google Scholar] [CrossRef] [Green Version]
  118. Wirries, A.; Jabari, S.; Jansen, E.P.; Roth, S.; Figueroa-Juárez, E.; Wissniowski, T.T.; Neureiter, D.; Klieser, E.; Lechler, P.; Ruchholtz, S.; et al. Panobinostat mediated cell death: A novel therapeutic approach for osteosarcoma. Oncotarget 2018, 9, 32997–33010. [Google Scholar] [CrossRef] [Green Version]
  119. Alhosin, M.; Abusnina, A.; Achour, M.; Sharif, T.; Muller, C.; Peluso, J.; Chataigneau, T.; Lugnier, C.; Schini-Kerth, V.B.; Bronner, C.; et al. Induction of apoptosis by thymoquinone in lymphoblastic leukemia Jurkat cells is mediated by a p73-dependent pathway which targets the epigenetic integrator UHRF1. Biochem. Pharmacol. 2010, 79, 1251–1260. [Google Scholar] [CrossRef] [Green Version]
  120. Abusnina, A.; Alhosin, M.; Keravis, T.; Muller, C.D.; Fuhrmann, G.; Bronner, C.; Lugnier, C. Down-regulation of cyclic nucleotide phosphodiesterase PDE1A is the key event of p73 and UHRF1 deregulation in thymoquinone-induced acute lymphoblastic leukemia cell apoptosis. Cell. Signal. 2011, 23, 152–160. [Google Scholar] [CrossRef]
  121. Omran, Z.; Dalhat, M.H.; Abdullah, O.; Kaleem, M.; Hosawi, S.; Al-Abbasi, F.A.; Wu, W.; Choudhry, H.; Alhosin, M. Targeting Post-Translational Modifications of the p73 Protein: A Promising Therapeutic Strategy for Tumors. Cancers 2021, 13, 1916. [Google Scholar] [CrossRef]
  122. Alghamdi, R.M.; Hassan, M.A.; Kaleem, M.; Kayali, A.; Halwani, M.A.; Zamzami, M.A.; Choudhry, H.; Alhosin, M. Targeting Itch/p73 pathway by thymoquinone as a novel therapeutic strategy for cancers with p53 mutation. Eur. J. Cell Sci. 2020, 2, 20–26. [Google Scholar] [CrossRef]
  123. Wu, P.F.; Tseng, H.C.; Chyau, C.C.; Chen, J.H.; Chou, F.P. Piper betle leaf extracts induced human hepatocellular carcinoma Hep3B cell death via MAPKs regulating the p73 pathway in vitro and in vivo. Food Funct. 2014, 5, 3320–3328. [Google Scholar] [CrossRef] [PubMed]
  124. Lunghi, P.; Costanzo, A.; Levrero, M.; Bonati, A. Treatment with arsenic trioxide (ATO) and MEK1 inhibitor activates the p73-p53AIP1 apoptotic pathway in leukemia cells. Blood 2004, 104, 519–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Liu, J.; Uematsu, H.; Tsuchida, N.; Ikeda, M.A. Essential role of caspase-8 in p53/p73-dependent apoptosis induced by etoposide in head and neck carcinoma cells. Mol. Cancer 2011, 10, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Uboveja, A.; Satija, Y.K.; Siraj, F.; Sharma, I.; Saluja, D. p73-NAV3 axis plays a critical role in suppression of colon cancer metastasis. Oncogenesis 2020, 9, 12. [Google Scholar] [CrossRef] [PubMed]
  127. Jiang, X.; Li, H. MiR-1180-5p regulates apoptosis of Wilms’ tumor by targeting. OncoTargets Ther. 2018, 11, 823–831. [Google Scholar] [CrossRef] [Green Version]
  128. Ory, B.; Ramsey, M.R.; Wilson, C.; Vadysirisack, D.D.; Forster, N.; Rocco, J.W.; Rothenberg, S.M.; Ellisen, L.W. A microRNA-dependent program controls p53-independent survival and chemosensitivity in human and murine squamous cell carcinoma. J. Clin. Investig. 2011, 121, 809–820. [Google Scholar] [CrossRef] [Green Version]
  129. Jacques, C.; Calleja, L.R.; Baud’huin, M.; Quillard, T.; Heymann, D.; Lamoureux, F.; Ory, B. miRNA-193a-5p repression of p73 controls Cisplatin chemoresistance in primary bone tumors. Oncotarget 2016, 7, 54503–54514. [Google Scholar] [CrossRef] [Green Version]
  130. Zhang, X.; Zhang, M.; Wang, G.; Tian, Y.; He, X. Tumor promoter role of miR-647 in gastric cancer via repression of TP73. Mol. Med. Rep. 2018, 18, 3744–3750. [Google Scholar] [CrossRef] [Green Version]
  131. Gao, Q.; Zheng, J. microRNA-323 upregulation promotes prostate cancer growth and docetaxel resistance by repressing p73. Biomed. Pharmacother. 2018, 97, 528–534. [Google Scholar] [CrossRef]
  132. Sampath, D.; Calin, G.A.; Puduvalli, V.K.; Gopisetty, G.; Taccioli, C.; Liu, C.G.; Ewald, B.; Liu, C.; Keating, M.J.; Plunkett, W. Specific activation of microRNA106b enables the p73 apoptotic response in chronic lymphocytic leukemia by targeting the ubiquitin ligase Itch for degradation. Blood 2009, 113, 3744–3753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Kehrloesser, S.; Osterburg, C.; Tuppi, M.; Schäfer, B.; Vousden, K.H.; Dötsch, V. Intrinsic aggregation propensity of the p63 and p73 TI domains correlates with p53R175H interaction and suggests further significance of aggregation events in the p53 family. Cell Death Differ. 2016, 23, 1952–1960. [Google Scholar] [CrossRef] [PubMed]
  134. Yi, Y.; Zhang, W.; Yi, J.; Xiao, Z.X. Role of p53 Family Proteins in Metformin Anti-Cancer Activities. J. Cancer 2019, 10, 2434–2442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Cai, B.H.; Bai, Z.Y.; Lien, C.F.; Yu, S.J.; Lu, R.Y.; Wu, M.H.; Wu, W.C.; Chen, C.C.; Hsu, Y.C. NAMPT Inhibitor and P73 Activator Represses P53 R175H Mutated HNSCC Cell Proliferation in a Synergistic Manner. Biomolecules 2022, 12, 438. [Google Scholar] [CrossRef] [PubMed]
  136. Crook, T.; Tidy, J.A.; Vousden, K.H. Degradation of p53 can be targeted by HPV E6 sequences distinct from those required for p53 binding and trans-activation. Cell 1991, 67, 547–556. [Google Scholar] [CrossRef]
  137. Huibregtse, J.M.; Scheffner, M.; Howley, P.M. Cloning and expression of the cDNA for E6-AP, a protein that mediates the interaction of the human papillomavirus E6 oncoprotein with p53. Mol. Cell Biol. 1993, 13, 775–784. [Google Scholar] [CrossRef]
  138. Ben Khalifa, Y.; Teissier, S.; Tan, M.K.; Phan, Q.T.; Daynac, M.; Wong, W.Q.; Thierry, F. The human papillomavirus E6 oncogene represses a cell adhesion pathway and disrupts focal adhesion through degradation of TAp63β upon transformation. PLoS Pathog. 2011, 7, e1002256. [Google Scholar] [CrossRef] [Green Version]
  139. Friborg, J.; Kong, W.; Hottiger, M.O.; Nabel, G.J. p53 inhibition by the LANA protein of KSHV protects against cell death. Nature 1999, 402, 889–894. [Google Scholar] [CrossRef]
  140. Mauser, A.; Saito, S.; Appella, E.; Anderson, C.W.; Seaman, W.T.; Kenney, S. The Epstein-Barr virus immediate-early protein BZLF1 regulates p53 function through multiple mechanisms. J. Virol. 2002, 76, 12503–12512. [Google Scholar] [CrossRef] [Green Version]
  141. Ma-Lauer, Y.; Carbajo-Lozoya, J.; Hein, M.Y.; Müller, M.A.; Deng, W.; Lei, J.; Meyer, B.; Kusov, Y.; von Brunn, B.; Bairad, D.R.; et al. p53 down-regulates SARS coronavirus replication and is targeted by the SARS-unique domain and PLpro via E3 ubiquitin ligase RCHY1. Proc. Natl. Acad. Sci. USA 2016, 113, E5192–E5201. [Google Scholar] [CrossRef]
  142. Elmore, L.W.; Hancock, A.R.; Chang, S.F.; Wang, X.W.; Chang, S.; Callahan, C.P.; Geller, D.A.; Will, H.; Harris, C.C. Hepatitis B virus X protein and p53 tumor suppressor interactions in the modulation of apoptosis. Proc. Natl. Acad. Sci. USA 1997, 94, 14707–14712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Bittar, C.; Shrivastava, S.; Bhanja Chowdhury, J.; Rahal, P.; Ray, R.B. Hepatitis C virus NS2 protein inhibits DNA damage pathway by sequestering p53 to the cytoplasm. PLoS ONE 2013, 8, e62581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Seavey, S.E.; Holubar, M.; Saucedo, L.J.; Perry, M.E. The E7 oncoprotein of human papillomavirus type 16 stabilizes p53 through a mechanism independent of p19(ARF). J. Virol. 1999, 73, 7590–7598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Li, P.; Jiang, H.; Peng, H.; Zeng, W.; Zhong, Y.; He, M.; Xie, L.; Chen, J.; Guo, D.; Wu, J.; et al. Non-Structural Protein 5 of Zika Virus Interacts with p53 in Human Neural Progenitor Cells and Induces p53-Mediated Apoptosis. Virol. Sin. 2021, 36, 1411–1420. [Google Scholar] [CrossRef] [PubMed]
  146. Bourdon, J.C.; Fernandes, K.; Murray-Zmijewski, F.; Liu, G.; Diot, A.; Xirodimas, D.P.; Saville, M.K.; Lane, D.P. p53 isoforms can regulate p53 transcriptional activity. Genes Dev. 2005, 19, 2122–2137. [Google Scholar] [CrossRef] [Green Version]
  147. Vieler, M.; Sanyal, S. p53 Isoforms and Their Implications in Cancer. Cancers 2018, 10, 288. [Google Scholar] [CrossRef] [Green Version]
  148. Hofstetter, G.; Berger, A.; Schuster, E.; Wolf, A.; Hager, G.; Vergote, I.; Cadron, I.; Sehouli, J.; Braicu, E.I.; Mahner, S.; et al. Δ133p53 is an independent prognostic marker in p53 mutant advanced serous ovarian cancer. Br. J. Cancer 2011, 105, 1593–1599. [Google Scholar] [CrossRef] [Green Version]
  149. Candeias, M.M.; Hagiwara, M.; Matsuda, M. Cancer-specific mutations in p53 induce the translation of Δ160p53 promoting tumorigenesis. EMBO Rep. 2016, 17, 1542–1551. [Google Scholar] [CrossRef]
  150. Gong, H.; Zhang, Y.; Jiang, K.; Ye, S.; Chen, S.; Zhang, Q.; Peng, J.; Chen, J. p73 coordinates with Δ133p53 to promote DNA double-strand break repair. Cell Death Differ. 2018, 25, 1063–1079. [Google Scholar] [CrossRef]
  151. Shekhar, S.; Dey, S. Induction of p73, Δ133p53, Δ160p53, pAKT lead to neuroprotection via DNA repair by 5-LOX inhibition. Mol. Biol. Rep. 2020, 47, 269–274. [Google Scholar] [CrossRef]
  152. Marcel, V.; Petit, I.; Murray-Zmijewski, F.; Goullet de Rugy, T.; Fernandes, K.; Meuray, V.; Diot, A.; Lane, D.P.; Aberdam, D.; Bourdon, J.C. Diverse p63 and p73 isoforms regulate Δ133p53 expression through modulation of the internal TP53 promoter activity. Cell Death Differ. 2012, 19, 816–826. [Google Scholar] [CrossRef] [PubMed]
  153. Arsic, N.; Slatter, T.; Gadea, G.; Villain, E.; Fournet, A.; Kazantseva, M.; Allemand, F.; Sibille, N.; Seveno, M.; de Rossi, S.; et al. Δ133p53β isoform pro-invasive activity is regulated through an aggregation-dependent mechanism in cancer cells. Nat. Commun. 2021, 12, 5463. [Google Scholar] [CrossRef] [PubMed]
  154. Ramos, H.; Soares, M.I.L.; Silva, J.; Raimundo, L.; Calheiros, J.; Gomes, C.; Reis, F.; Monteiro, F.A.; Nunes, C.; Reis, S.; et al. A selective p53 activator and anticancer agent to improve colorectal cancer therapy. Cell Rep. 2021, 35, 108982. [Google Scholar] [CrossRef]
  155. Lindemann, A.; Patel, A.A.; Silver, N.L.; Tang, L.; Liu, Z.; Wang, L.; Tanaka, N.; Rao, X.; Takahashi, H.; Maduka, N.K.; et al. COTI-2, A Novel Thiosemicarbazone Derivative, Exhibits Antitumor Activity in HNSCC through p53-dependent and -independent Mechanisms. Clin. Cancer Res. 2019, 25, 5650–5662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Abrams, S.L.; Duda, P.; Akula, S.M.; Steelman, L.S.; Follo, M.L.; Cocco, L.; Ratti, S.; Martelli, A.M.; Montalto, G.; Emma, M.R.; et al. Effects of the Mutant TP53 Reactivator APR-246 on Therapeutic Sensitivity of Pancreatic Cancer Cells in the Presence and Absence of WT-TP53. Cells 2022, 11, 794. [Google Scholar] [CrossRef]
  157. Bykov, V.J.; Issaeva, N.; Shilov, A.; Hultcrantz, M.; Pugacheva, E.; Chumakov, P.; Bergman, J.; Wiman, K.G.; Selivanova, G. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat. Med. 2002, 8, 282–288. [Google Scholar] [CrossRef]
  158. Lambert, J.M.; Gorzov, P.; Veprintsev, D.B.; Söderqvist, M.; Segerbäck, 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] [Green Version]
  159. Menichini, P.; Monti, P.; Speciale, A.; Cutrona, G.; Matis, S.; Fais, F.; Taiana, E.; Neri, A.; Bomben, R.; Gentile, M.; et al. Antitumor Effects of PRIMA-1 and PRIMA-1 Met (APR246) in Hematological Malignancies: Still a Mutant P53-Dependent Affair? Cells 2021, 10, 98. [Google Scholar] [CrossRef]
  160. Rökaeus, N.; Shen, J.; Eckhardt, I.; Bykov, V.J.; Wiman, K.G.; Wilhelm, M.T. PRIMA-1(MET)/APR-246 targets mutant forms of p53 family members p63 and p73. Oncogene 2010, 29, 6442–6451. [Google Scholar] [CrossRef] [Green Version]
  161. 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] [Green Version]
  162. Kobayashi, T.; Makino, T.; Yamashita, K.; Saito, T.; Tanaka, K.; Takahashi, T.; Kurokawa, Y.; Yamasaki, M.; Nakajima, K.; Morii, E.; et al. APR-246 induces apoptosis and enhances chemo-sensitivity via activation of ROS and TAp73-Noxa signal in oesophageal squamous cell cancer with TP53 missense mutation. Br. J. Cancer 2021, 125, 1523–1532. [Google Scholar] [CrossRef] [PubMed]
Figure 1. p53 has a high nonsense and missense mutation rate compared to p63 and p73 in cancer. (A) p53 has, respectively, a 10.72% and 62.74% nonsense mutation rate and missense mutation rate in all cancer mutation samples from the COSMIC database (https://cancer.sanger.ac.uk/cosmic; accessed date (29 April 2022) [18]. (B) p63 has, respectively, 2.97% and 32.34% nonsense mutation rate and missense mutation rate in all cancer mutation samples. (C) p73 has, respectively, 2.28% and 27.33%, nonsense mutation rate and missense mutation rate in all cancer mutation samples.
Figure 1. p53 has a high nonsense and missense mutation rate compared to p63 and p73 in cancer. (A) p53 has, respectively, a 10.72% and 62.74% nonsense mutation rate and missense mutation rate in all cancer mutation samples from the COSMIC database (https://cancer.sanger.ac.uk/cosmic; accessed date (29 April 2022) [18]. (B) p63 has, respectively, 2.97% and 32.34% nonsense mutation rate and missense mutation rate in all cancer mutation samples. (C) p73 has, respectively, 2.28% and 27.33%, nonsense mutation rate and missense mutation rate in all cancer mutation samples.
Biomedicines 10 01490 g001
Figure 2. Key upstream molecules that influence p53 gain of function. (A) MYC and lincRNA-p21 are likely the key switch molecules that determine the gain-of-function or non-gain-of-function p53 signal to regulate the HSP70 expression to influence p53 mutant aggregation. Gain-of-function p53 mutant may induce the MYC-NAMPT- SIRT1-HSP70 axis to induce p53 auto-aggregation, and non-gain of function p53 mutants still can maintain non-aggregated p53 through the lincRNA-p21-STAT3-HSP70 axis. Wild-type p53 likely keeps its non-aggregated status through repression of MYC and upregulation of lincRNA-p21. (B) Oncogenic-type long-noncoding RNA LINC00511 and LINC00460 can act as a sponge to repress miR-150 to promote mutant p53 accumulation. A tumor-suppressor type transcription factor, AP2, can repress LINC00511 expression to decrease the amount of mutant p53.
Figure 2. Key upstream molecules that influence p53 gain of function. (A) MYC and lincRNA-p21 are likely the key switch molecules that determine the gain-of-function or non-gain-of-function p53 signal to regulate the HSP70 expression to influence p53 mutant aggregation. Gain-of-function p53 mutant may induce the MYC-NAMPT- SIRT1-HSP70 axis to induce p53 auto-aggregation, and non-gain of function p53 mutants still can maintain non-aggregated p53 through the lincRNA-p21-STAT3-HSP70 axis. Wild-type p53 likely keeps its non-aggregated status through repression of MYC and upregulation of lincRNA-p21. (B) Oncogenic-type long-noncoding RNA LINC00511 and LINC00460 can act as a sponge to repress miR-150 to promote mutant p53 accumulation. A tumor-suppressor type transcription factor, AP2, can repress LINC00511 expression to decrease the amount of mutant p53.
Biomedicines 10 01490 g002
Table 2. Table summarizing p73 activation drugs used to treat cancer cell lines. If there is no COSMIC ID or ATCC ID of a certain cell line from the p53 database [104], the PubMed ID is provided for information about the p53 status within indicated cell line. CDS, coding sequence. *, premature stop codons.
Table 2. Table summarizing p73 activation drugs used to treat cancer cell lines. If there is no COSMIC ID or ATCC ID of a certain cell line from the p53 database [104], the PubMed ID is provided for information about the p53 status within indicated cell line. CDS, coding sequence. *, premature stop codons.
p73
Activation Drugs
Cell LineCOSMIC IDATCC IDPubMed IDTissuep53 StatusZygosityCDS
Mutation
Protein ChangeRef.
RETRAA431910925CRL-1555-SkinMUTHomozygousc.818G > Ap.R273H[105]
SW480-CCL-228-Large intestineMUTHeterozygousc.818G > A & 925C > Tp.R273H & P309S[105]
NSC59984SW480-CCL-228-Large intestineMUTHeterozygousc.818G > A & 925C > Tp.R273H & P309S[106]
DLD1-CCL-221-Large intestineMUTHeterozygousc.722C > Tp.S241F[106]
ProdigiosinSW480-CCL-228-Large intestineMUTHeterozygousc.818G > A & c.925C > Tp.R273H &p.P309S[107,108]
DLD1-CCL-221-Large intestineMUTHeterozygousc.722C > Tp.S241F[107,108]
MetforminHCT-116905936CCL-247-Large intestinewild typeHomozygous--[109]
FaDu906863HTB-43-Head and neckMUTHeterozygousc.743G > Tp.R248L[110]
H596908459HTB-178-LungMUTHomozygousc.733G > Tp.G245C[110]
H292753604--Lungwild typeHeterozygous--[110]
BortezomibHT29905939HTB-38-ColonMUTHomozygousc.818G > Ap.R273H[111]
Nutlin 3IEC-6--25230151Small intestine; Epitheliumwild typeHomozygous--[114]
Caco2--16418264Large intestineMUTHomozygousc.610G > Tp.E204*[114]
Diallyl disulfideHeLa1298134--Cervixwild typeHomozygous--[115]
Abrus agglutininFaDu906863HTB-43-Head and neckMUTHeterozygousc.743G > Tp.R248L[116]
CinobufaginHuh-7907071 -LiverMUTHomozygousc.659A > Gp.Y220C[117]
ThymoquinoneMDA-MB-468908123HTB-132-BreastMUTHomozygousc.818G > Ap.R273H[121,122]
HL60905938CCL-240-Haematopoietic and lymphoidNULLHomozygousc.1_1182delNo
expression
[121,122]
Extracts of Piper betle leaf (PBL)Hep3B-HB-8064-LiverNULLHomozygousgross deletionNo
expression
[123]
MEK1 inhibitorNB-41323913--Haematopoietic and lymphoidMUTHomozygousc.743G > Ap.R248Q[124]
K562905940CCL-243-Haematopoietic and lymphoidMUTHomozygousc.406dupp.Q136Pfs*13[124]
EtoposideHOC313--1570156MouthMUTUnknownc.853G > Ap.E285K[125]
Ca9-22753538--MouthMUTHomozygousc.742C > Tp.R248W[125]
miR-1180
inhibitor
SK-NEP-1909730HTB-48-KidneyMUTHomozygousc.733G > Ap.G245S[127]
miRNA-193a-5p inhibitorJHU-0291298156--Head and neckMUTHeterozygousc.323delp.G108Vfs*15[128,129]
MG63908131--Bonewild typeHomozygous--[128,129]
miR-647 inhibitorMGC-803--9999992Stomachwild typeHomozygous--[130]
miR-323 inhibitorPC-3905934CRL-1435-ProstateMUTHomozygousc.414delp.K139Rfs*31[131]
miR106b mimicK562905940CCL-243-Haematopoietic and lymphoidMUTHomozygousc.406dupp.Q136Pfs*13[132]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cai, B.-H.; Hsu, Y.-C.; Yeh, F.-Y.; Lin, Y.-R.; Lu, R.-Y.; Yu, S.-J.; Shaw, J.-F.; Wu, M.-H.; Tsai, Y.-Z.; Lin, Y.-C.; et al. P63 and P73 Activation in Cancers with p53 Mutation. Biomedicines 2022, 10, 1490. https://doi.org/10.3390/biomedicines10071490

AMA Style

Cai B-H, Hsu Y-C, Yeh F-Y, Lin Y-R, Lu R-Y, Yu S-J, Shaw J-F, Wu M-H, Tsai Y-Z, Lin Y-C, et al. P63 and P73 Activation in Cancers with p53 Mutation. Biomedicines. 2022; 10(7):1490. https://doi.org/10.3390/biomedicines10071490

Chicago/Turabian Style

Cai, Bi-He, Yun-Chien Hsu, Fang-Yu Yeh, Yu-Rou Lin, Rui-Yu Lu, Si-Jie Yu, Jei-Fu Shaw, Ming-Han Wu, Yi-Zhen Tsai, Ying-Chen Lin, and et al. 2022. "P63 and P73 Activation in Cancers with p53 Mutation" Biomedicines 10, no. 7: 1490. https://doi.org/10.3390/biomedicines10071490

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