Skip to main content
F1000Research logoLink to F1000Research
. 2017 Nov 22;6:2041. [Version 1] doi: 10.12688/f1000research.12682.1

P53 at the start of the 21st century: lessons from elephants

Sue Haupt 1,2,a, Ygal Haupt 1,2,3,4
PMCID: PMC5701437  PMID: 29250320

Abstract

Crucial, natural protection against tumour onset in humans is orchestrated by the dynamic protein p53. The best-characterised functions of p53 relate to its cellular stress responses. In this review, we explore emerging insights into p53 activities and their functional consequences. We compare p53 in humans and elephants, in search of salient features of cancer protection.

Keywords: P53, Tumour suppressor protein, cancer

Introduction

Protection from DNA damage defines the primary function of ancestral p53 at its emergence around one billion years ago. This is deduced from germline gametes of the modern-day early metazoan descendent, the sea anemone (reviewed in 1). Conservation of this core business in contemporary p53, which promotes the preservation of DNA integrity in our evolutionarily developed human species, defines its critical function as the ‘guardian of the genome’.

The ascribed gene name, tumour suppressor protein p53 ( TP53), reflects its key role in suppressing malignant transformation in advanced species. Recent analyses reveal that corruption of p53 function (through mutation in approximately 50% of all human cancers 2 and negative regulation in others 3) can wreak havoc across the epigenome 4, the coding and non-coding transcriptome 5, 6, microRNA (miRNA) machinery 7, and the proteasome 8, resulting in altered protein output. Ensuing pathway deregulation impacts on cellular stress responses, including those that facilitate DNA repair or cell termination of the irreparable (reviewed in the recent series edited by Haupt and Blandino 9), autophagy (reviewed in 10), mRNA translation, DNA replication 11, metabolism 12 and immunity (reviewed in 13).

The emergence of cancer in humans is often touted as a largely modern-day affliction that has arisen with extended longevity associated with clinical advances. Remarkably, however, one of the oldest living mammals, the elephant, rarely, if ever, dies of cancer. Highly relevant to this review, p53 appears to be the lynchpin to explain both these scenarios. In humans, cancers are associated with the high prevalence of TP53 gene mutation 2. Elephants, on the other hand, have extended, cancer-free longevity, attributed to their at least 20 paired copies of its TP53 repertoire 14. Dissection of this protection offers fascinating insight into p53 function.

How p53 defends against DNA damage to preserve the genome and fight cancer is still being elucidated, despite more than 35 years of intense molecular and cellular study. Until very recently, the field predominantly focused on the role of p53 as a transcriptional regulator, particularly on its transactivation targets that drive arrest and apoptosis. The capacity of p53 to suppress tumours independent of key mediators of these processes has challenged accepted knowledge 15, 16. The complexity of the p53 response continues to emerge along with new understanding of the contribution of p53 to transcriptional repression 17, 18 and also revisitation of the concept of p53 transactivation-independent function, which was first reported more than 20 years ago 19.

The best-defined p53 responses to DNA damage are either temporary or permanent interruption of cell proliferation. Measured restraint of these potent p53 responses is biologically essential for survival. Despite early suggestions that this feature developed late in evolution, new findings have identified a key orthologue in the fly genome of the major negative regulator of p53 20. The ancient origins of p53 and also the ancestral form of the contemporary regulators MDM2 and MDM4 reflect their fundamental importance for evolutionary fitness.

Tumour suppressor function of p53

The critical role of p53 in preserving genomic integrity is supported by extensive exome sequence data sets (including from the Getz lab 21 and the Tumor Cancer Genome Atlas 22), which identify TP53 as the single most frequently mutated gene in cancer. Furthermore, p53 pathway genes proved to be the most significantly enriched set in the cancer susceptibility loci in the 1000 Genomes Project 23. In parallel, the loss of TP53 24 or germline 25, 26 mutation predisposes mouse models to cancer and its mutation is the major driver of malignancy in the human inherited Li-Fraumeni cancer syndrome. Intriguingly, in stem cells, p53 controls cell differentiation 27, which is inherently distinct from its counterpart role in protecting from DNA corruption in somatic cells. These distinct functional differences have been tentatively attributed to distinct p53 isoform expression (reviewed in 28). These key points condense the message of thousands of individual studies that, in vertebrates, p53 performs a major tumour-suppressive role in somatic cells. In this review, we will focus on the somatic functions of p53.

P53 transactivation function

In response to a range of cellular stresses, p53 transactivates multiple target genes to regulate a range of outcomes, including cell growth arrest, apoptosis, DNA damage repair, oncogene activation, telomere shortening and metabolic disturbance (reviewed in 29). While p53 transcriptional activity is widely regarded as its core function, intriguing recent findings have exposed greater complexity. Consistent with conventional understanding, transactivation-incompetent p53 mutants fail to suppress tumour development 15. This provoked an intense search for decisive p53 targets. Although thousands of putative p53 transcriptional targets have been reported, only a couple of hundred were identified at high confidence 17, 30 and these appear independent of cell type and treatment 31. More specifically and unexpectedly, a synthetic p53 mutant rendered incapable of transactivating its key known mediators of growth arrest and apoptosis is still able to suppress cancer development 15. Consistently, ablation of prime transcriptional p53 targets (specifically: p21, Puma and Noxa) in the cell-cycle inhibitory pathways failed to completely recapitulate p53 loss 16. Intense study is under way, in many labs, to define the critical p53 targets that execute its downstream effects (see ‘P53 gene repressor function’, ‘DNA damage response’ (DDR), ‘Emerging p53 growth suppression mechanism: ferroptosis-induced cell death’ sections below). To comprehensively evaluate the significance of these studies, a key distinction must be drawn between the prevention of cancer development that was assessed and intervention to treat a developed cancer that remains to be tested (particularly pertinent to the very elegant in vivo experiments from the labs of Gerard Evan 32, Scott Lowe 33 and Tyler Jacks 34).

P53 gene repressor function

A mechanism of p53 transcriptional repression has recently been delineated which has led to a fundamental revision of our understanding of p53 activity, particularly the induction of cell-cycle arrest. This new concept supersedes the former dogma that p53 represses gene transcription through direct engagement of particular response elements 35. P53 transcription inhibition is now attributed to an indirect p53 action where p53 transactivates the cyclin-dependent kinase (CDK) inhibitor p21 (CDKN1A/p21), causing interference with phosphorylation of RB-like pocket protein homologs RBL1 (p107) and RBL2 (p130). These hypo-phosphorylated RB-like proteins then cause stabilization of the multi-protein repressor ‘DREAM’ complex (which is composed of dimerization proteins [DPs], RB-like proteins, E2F4 and MUVB). DREAM is a transcriptional repressor complex that engages E2F or CHR promoter sites. A link between p53 and DREAM is a new concept that is now referred to as the ‘p53-p21-DREAM pathway’ of gene repression. The repertoire of genes repressed by this complex are largely cell cycle–associated and include those in DNA repair (as discussed below). It is the recruitment of this complex to the promoters of target genes that causes transcriptional repression (reviewed in 17). This is a fundamental change in thinking and is relevant to the multitude of targets repressed when p53 is activated.

P53 responses to stress

DNA damage response

A decisive role for p53 in the DDR has been recognized (reviewed in 36). Indeed, in response to disruption of the genome, critical directives by p53 seal cellular fate. P53 is capable of activating molecular processes to either initiate temporary arrest and repair or induce permanent arrest or death. Despite the many thousands of studies defining the role of p53 in these mechanisms, what directs these choices still awaits comprehensive elucidation (see ‘Emerging functions: p53 regulation of the epigenome’ section below 37).

This brief survey of DDR p53 targets reveals extensive intervention across multiple processes. P53 participation in a range of DDR was recently reviewed 36; however, we will limit this discussion to the high-confidence targets. P53 regulates DDR facilitators, through direct engagement of facilitating proteins during repair but also by transactivating key targets. First, p53 contributes to detection of DNA damage by promoting chromatin relaxation, which it achieves by engaging and consequently reducing the activity of two DNA helicases: XPB and XPD. This relaxation is further stimulated by p53 recruitment of p300 histone acetylase (HAT) to mediate histone H3 subunit acetylation at damage sites (reviewed in 36). Second, halting cell division to enable repair is understood to be a key activity of p53 target CDKN1A, which is one of the most high-confidence targets identified in multiple screens (reviewed in 17).

Third, in response to single-stranded DNA damage caused by ultraviolet radiation, nucleotide excision repair (NER) (reviewed in 36) is provoked by the recruitment, to the break site, of two high-confidence transactivation targets of p53 that are NER pathway components: damage-specific DNA-binding protein 2 (DDB2/XPE) and XPC 17. RRM2B promotes DNA repair by feeding precursor deoxyribonucleoside diphosphates (dNTPs), which it catalytically converts from ribonucleoside diphosphates 38. Identification of RRM2B as one of the top two transactivation targets of p53, together with CDKN1A 17, predicts the significance of p53 in directing arrest and DNA repair.

An additional p53 target of indirect p53-p21-DREAM repression that has been linked to NER is the mismatch repair (MMR) core component M2H2 17. Suggestion that p53 and M2H2 engage physically 36 queries the existence of additional levels of regulation at the level of transcription. In addition, p53 has been linked to other DDR pathways, including homologous recombination (HR), MMR and base excision repair.

RAD51 is another target of p53 involved in HR to regulate double-strand break repair. Interestingly, while reported as a high-confidence target of repression by the p53-p21-DREAM pathway 17, it once again appears that this is dependent on the isoform of p53 present. Specifically, the p53 isoform Delta133p53 is reported to upregulate RAD51 39. Furthermore, direct interaction between p53 and RAD51 was identified to control HR, suggesting that this engagement prevents aberrant recombination events (reviewed in 36).

Fourth, in response to DNA damage, repair is promoted through p53 transactivation of proliferating cell nuclear antigen (PCNA) 40, which is a definitive component of the DNA replication fork, and acts as a co-factor of DNA POL Delta. This augments the essential cell cycle–regulated function of PCNA 41. P53 mutation eliminates this DDR 40. A fascinating finding is that PCNA has response elements for both p53 and DREAM 17. This predicts an inbuilt ‘rheostat’ to properly meter-out the DDR, where p53 could initially activate a DDR target such as PCNA, simultaneously with CDKN1A, and as a secondary containment event, the p21-DREAM repression complex would override. Similarly, DNA polymerase H ( POLH) appears to be differentially regulated by p53 direct and indirect activity 17. This interesting concept of regulation needs testing, but oscillating levels of p53 products is an established concept (as evidenced in the p53-MDM2 feedback loop 42, 43).

At face value, if we assume that it is vital to interrupt progression through the cell cycle to facilitate repair, it is surprising that p21 has not been identified in germline models to be critical for TS. However, perhaps it is actually the repair functions of p53 that are vital—with redundancy in arrest induction and tolerance of imprecise DREAM dampening of repair gene repression? In this context, the involvement of p53 in both G 1 and G 2 arrest (reviewed in 44) is pertinent.

Emerging p53 growth suppression mechanism: ferroptosis-induced cell death

A critical role for p53 in triggering cell death through iron-mediated ferroptosis has taken centre stage recently. In non-colorectal cancer (CRC) cells, p53 was reported to inhibit the transcription of solute carrier cystine-glutamate antiporter SLC7A11, which in turn drives ferroptosis 45. In contrast, a surprising new study reports that CRC cells specifically are protected from ferroptosis through wild-type (wt) p53 engagement in a transcriptional-independent manner 46. In these cells, p53 physically binds and promotes the relocation of the ferroptosis promoter DPP4 (CD26) into the nucleus where it is inactive. This is reported as a p53 transcription-independent function in these cells. Therapeutic opportunities for targeting this antiporter system in the absence of functional p53 show promise across a range of cells 46, 47 and constitute an active area of research. These studies highlight the importance of understanding p53 activity in context. Further clarification is warranted regarding the breadth of p53 transcriptional-dependent and -independent activities across healthy and disease contexts to execute ferroptosis and how this meshes with p53-induced arrest and apoptosis.

Emerging functions: p53 regulation of the epigenome

A role for p53 in epigenomic regulation of the extensive non-coding elements of the genome is newly emerging. Epigenetics refers to DNA and histone modifications, plus chromatin remodelling. Importantly, promoter methylation is linked to transcription repression while methylation in gene bodies is associated with transcription activation (reviewed in 48).

P53 is involved in transcriptional silencing of repetitive, short interspersed nuclear elements (SINEs) and non-coding RNAs. In keeping with this, p53 has also attracted the title of ‘guardian of repeats’. These are DNA elements attributed to ancient viral invasion of the genome. As a fateful safety precaution, a pathway of type I interferon (IFN)-mediated self-destructive cell death is triggered if these elements are transcribed in normal cells (where formation of double-stranded RNA appears key to activating the response). It is unsurprising then that p53 mutation, DNA hypomethylation and breakdown of regulated IFN function frequently accompany tumorigenesis 49.

The exact mechanism of p53’s involvement in this silencing is awaiting elucidation. Pertinently, links between p53 and DNA methyl transferases (DNMTs) have been reported. Cell-cycle interruption is attributed to p53 recruitment of DMNT1 and consequent methylation of promoters of genes that promote cell growth (for example, through targeting the inhibitor of p53 growth arrest and the cell division cycle (cdc) 25C (CDC25C) tyrosine phosphatase and also by downregulating the anti-apoptotic gene Survivin ( BIRC5) 50. Of therapeutic relevance, DNMT1 inhibition induced by 5-aza-2′-deoxycytidine induces DNA hypomethylation exclusively in a wt p53 context resulting in a protective G 2/M checkpoint arrest, while cells (both primary and non-transformed cells) lacking p53 do not stop dividing and undergo extreme chromosomal abnormalities, then apoptose 51. In contrast, DMNT3a also interacts with p53, but in this instance, it counters growth inhibition through the methylation of genes involved in arrest (for example, CDKN1A/p21). This is in keeping with elevated DMNT3a levels in cancers 52.

In an attempt to identify the chromatin-related factors that discriminate the capacity of p53 to instigate either arrest or apoptosis, Shelley Berger’s group undertook an RNA interference analysis of the chromatin mediators involved in p53-dependent transcription of its key respective targets: CDKN1A/p21 and BBC3/puma 37. This work was rationalized upon the evident overlap between p53 and chromatin regulatory pathways. In this study, DNMTs did not stand out as key regulators of transcription; however, other chromatin mediators were distinguished as either positive or negative regulators of p53 activity under basal or stress conditions. Furthermore, target specificity was clear among these regulators. These rich lists are composed of both known and new candidates and their biological significance beyond a single cancer cell line awaits unmasking ( 37 and references within).

Another fascinating study by the Berger group identified that mutant p53 can drive cancer by instigating epigenomic disruption. Transcription of chromatin regulators was found to be subject to mutant p53 but not affected by wt p53. Mutant p53 is able to co-localize with ETS2 and Pol II at the transcriptional start sites of a number of vital chromatin regulatory genes, methyltransferases (notably MLL1 and MLL2), and also of the acetyltransferase (MOZ). The result is genome-wide increased histone methylation and acetylation associated with cancer growth.

Consistently, analyses of large, publicly available genome data sets demonstrate the trend that gain-of-function p53 mutation correlates with elevated expression levels of MLL1, MLL2 and MOZ. This work defines a rational new application for the burgeoning field of epigenetic drug regulators 4 and predicts the relevance of stratification according to p53 status.

P53 regulation beyond MDM2

New aspects regarding the regulation of p53 are also emerging. While elevated levels of p53 in response to stress are largely attributed to its post-translational modifications to protect from the proteasome, additional levels of transcriptional control are also evident. At the post-transcriptional level, the well-established role of MDM2 as the major regulator of p53, in partnership with MDM4, is being elaborated to define therapeutic relevance (reviewed in 53) and the oncogenic role of these regulators in mutant p53 cancers is also being exposed (for example, 54).

At the level of mRNA processing, the contribution of mRNA splicing is reflected by the generation of discretely spliced p53 isoforms in stem cells, compared with differentiated tissues 27. Abnormal p53 isoforms that appear in cancers reflect the subversion of mRNA splicing associated with malignancy (reviewed in 55).

In addition, a host of miRNAs are being identified that also regulate p53 levels (reviewed in 56) and again their deregulation is a risk for cancer. Beyond this level of control, regulation during translation also occurs, and an alternative site for ribosome attachment to TP53 RNA, termed the internal ribosomal entry sites which are engaged at different phases of the cell cycle 57.

P53 and elephants

The multiple copies of TP53 in cancer-resistant elephants cast a fascinating perspective on its tumour-suppressive function ( Table 1). These additional copies of TP53 are referred to as TP53 retrogenes ( TP53RTGs) and while a number of these are transcribed, they do not appear to be directly transcriptionally active. Intriguingly, elephants have an exceptionally well-developed response to DNA damage and their dermal fibroblasts trigger apoptotic cell death when exposed to low doses of stimuli but do not respond to Nutlin 3a that relieves p53 from MDM2 suppression. The dual functions identified for TP53RTGs are, first, to repress p53 signalling in the absence of activating stimuli and, second, to promote greater sensitivity to DNA damage. The proposed mechanism for these functions stems from the finding that a TP53RTG is transcriptionally incompetent but can dimerize with p53 in the absence of stress, effectively protecting it from MDM2 and the proteasome. This ‘pool of protected p53’ can then respond rapidly to DNA damage. The suggestion that these extra copies encode TP53 isoforms that are not subject to MDM2 regulation predicts a fascinating biological perspective and has potential ramifications for consideration for human cancer therapy and longevity 14.

Table 1. Comparison of humans and elephants and parameters of relevance to cancer-free survival.

Species Humans Elephants
Average survival ~70 years 58 60–70 years 59
Number of pairs of
p53 copies 14
1 1
Number of pairs of
p53 transgenes 14
0 19
Cancer incidence 1:4 (US humans
by 85 years 60)
~0 14

These studies open many additional questions related to elephant p53 regulation during its cancer-free development by analogy to the suggestion that distinct p53 isoforms drive individual functions in stem cell differentiation 27. The speculation that orthologous isoforms exist between elephants and humans is intriguing 14, particularly with recent findings indicating that p53 isoforms moderate the DDR in humans 39. These studies also raise many other exciting questions such as the status of other p53 family members in elephants, their interplay with p53 and the nature of their regulation. The study of elephants suggests that the full extent of the tumour-suppressive capacity of p53 is yet to be tapped in humans and predicts that vital cancer resistance possibilities await discovery and adoption.

Editorial Note on the Review Process

F1000 Faculty Reviews are commissioned from members of the prestigious F1000 Faculty and are edited as a service to readers. In order to make these reviews as comprehensive and accessible as possible, the referees provide input before publication and only the final, revised version is published. The referees who approved the final version are listed with their names and affiliations but without their reports on earlier versions (any comments will already have been addressed in the published version).

The referees who approved this article are:

  • Jerson L Silva, Institute of Medical Biochemistry, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil

  • Ashish Lal, Regulatory RNAs and Cancer Section, Genetics Branch, Center for Cancer Research, National Cancer Institute, National Institutes of Health, Bethesda, USA

  • Xinbin Chen, Comparative Oncology Laboratory, School of Veterinary Medicine, School of Medicine, University of California, Davis, USA

Funding Statement

The Haupt lab acknowledges funding from the following sources: National Health and Medical Research Council (1123057), Cancer Council Victoria (1085154), National Breast Cancer Foundation (IN-16-042), the Peter MacCallum Foundation, and the MD Anderson Sister Institute Network Fund (SINF).

The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

[version 1; referees: 3 approved]

References

  • 1. Belyi VA, Ak P, Markert E, et al. : The origins and evolution of the p53 family of genes. Cold Spring Harb Perspect Biol. 2010;2(6):a001198. 10.1101/cshperspect.a001198 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Olivier M, Hollstein M, Hainaut P: TP53 mutations in human cancers: origins, consequences, and clinical use. Cold Spring Harb Perspect Biol. 2010;2(1):a001008. 10.1101/cshperspect.a001008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Wade M, Li YC, Wahl GM: MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nat Rev Cancer. 2013;13(2):83–96. 10.1038/nrc3430 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Zhu J, Sammons MA, Donahue G, et al. : Gain-of-function p53 mutants co-opt chromatin pathways to drive cancer growth. Nature. 2015;525(7568):206–11. 10.1038/nature15251 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 5. Di Agostino S, Strano S, Emiliozzi V, et al. : 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(3):191–202. 10.1016/j.ccr.2006.08.013 [DOI] [PubMed] [Google Scholar]
  • 6. Strano S, Dell'Orso S, Di Agostino S, et al. : Mutant p53: an oncogenic transcription factor. Oncogene. 2007;26(15):2212–9. 10.1038/sj.onc.1210296 [DOI] [PubMed] [Google Scholar]
  • 7. Gurtner A, Falcone E, Garibaldi F, et al. : Dysregulation of microRNA biogenesis in cancer: the impact of mutant p53 on Drosha complex activity. J Exp Clin Cancer Res. 2016;35:45. 10.1186/s13046-016-0319-x [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 8. Walerych D, Lisek K, Sommaggio R, et al. : Proteasome machinery is instrumental in a common gain-of-function program of the p53 missense mutants in cancer. Nat Cell Biol. 2016;18(8):897–909. 10.1038/ncb3380 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 9. Haupt Y, Blandino G: Editorial: Human Tumor-Derived p53 Mutants: A Growing Family of Oncoproteins. Front Oncol. 2016;6:170. 10.3389/fonc.2016.00170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Cordani M, Butera G, Pacchiana R, et al. : Molecular interplay between mutant p53 proteins and autophagy in cancer cells. Biochim Biophys Acta. 2017;1867(1):19–28. 10.1016/j.bbcan.2016.11.003 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 11. Qiu WG, Polotskaia A, Xiao G, et al. : Identification, validation, and targeting of the mutant p53-PARP-MCM chromatin axis in triple negative breast cancer. NPJ Breast Cancer. 2017;3: pii: 1. 10.1038/s41523-016-0001-7 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 12. Zhou G, Myers JN: Mutant p53 exerts oncogenic functions by modulating cancer cell metabolism. Mol Cell Oncol. 2014;1(3):e963441. 10.4161/23723548.2014.963441 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Cui Y, Guo G: Immunomodulatory Function of the Tumor Suppressor p53 in Host Immune Response and the Tumor Microenvironment. Int J Mol Sci. 2016;17(11): pii: E1942. 10.3390/ijms17111942 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 14. Sulak M, Fong L, Mika K, et al. : TP53 copy number expansion is associated with the evolution of increased body size and an enhanced DNA damage response in elephants. eLife. 2016;5: pii: e11994. 10.7554/eLife.11994 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 15. Jiang D, Brady CA, Johnson TM, et al. : Full p53 transcriptional activation potential is dispensable for tumor suppression in diverse lineages. Proc Natl Acad Sci U S A. 2011;108(41):17123–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Valente LJ, Gray DH, Michalak EM, et al. : p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 2013;3(5):1339–45. 10.1016/j.celrep.2013.04.012 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 17. Fischer M: Census and evaluation of p53 target genes. Oncogene. 2017;36(28):3943–56. 10.1038/onc.2016.502 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 18. Fischer M, Quaas M, Nickel A, et al. : Indirect p53-dependent transcriptional repression of Survivin, CDC25C, and PLK1 genes requires the cyclin-dependent kinase inhibitor p21/CDKN1A and CDE/CHR promoter sites binding the DREAM complex. Oncotarget. 2015;6(39):41402–17. 10.18632/oncotarget.6356 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 19. Haupt Y, Rowan S, Shaulian E, et al. : Induction of apoptosis in HeLa cells by trans-activation-deficient p53. Genes Dev. 1995;9(17):2170–83. 10.1101/gad.9.17.2170 [DOI] [PubMed] [Google Scholar]
  • 20. Lu WJ, Amatruda JF, Abrams JM: p53 ancestry: gazing through an evolutionary lens. Nat Rev Cancer. 2009;9(10):758–62. 10.1038/nrc2732 [DOI] [PubMed] [Google Scholar]
  • 21. Lawrence MS, Stojanov P, Mermel CH, et al. : Discovery and saturation analysis of cancer genes across 21 tumour types. Nature. 2014;505(7484):495–501. 10.1038/nature12912 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 22. Kandoth C, McLellan MD, Vandin F, et al. : Mutational landscape and significance across 12 major cancer types. Nature. 2013;502(7471):333–9. 10.1038/nature12634 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 23. Stracquadanio G, Wang X, Wallace MD, et al. : The importance of p53 pathway genetics in inherited and somatic cancer genomes. Nat Rev Cancer. 2016;16(4):251–65. 10.1038/nrc.2016.15 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 24. Jackson JG, Lozano G: The mutant p53 mouse as a pre-clinical model. Oncogene. 2013;32(37):4325–30. 10.1038/onc.2012.610 [DOI] [PubMed] [Google Scholar]
  • 25. Lang GA, Iwakuma T, Suh YA, et al. : Gain of function of a p53 hot spot mutation in a mouse model of Li-Fraumeni syndrome. Cell. 2004;119(6):861–72. 10.1016/j.cell.2004.11.006 [DOI] [PubMed] [Google Scholar]
  • 26. Olive KP, Tuveson DA, Ruhe ZC, et al. : Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell. 2004;119(6):847–60. 10.1016/j.cell.2004.11.004 [DOI] [PubMed] [Google Scholar]
  • 27. Ungewitter E, Scrable H: Delta40p53 controls the switch from pluripotency to differentiation by regulating IGF signaling in ESCs. Genes Dev. 2010;24(21):2408–19. 10.1101/gad.1987810 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Levine AJ, Puzio-Kuter AM, Chan CS, et al. : The Role of the p53 Protein in Stem-Cell Biology and Epigenetic Regulation. Cold Spring Harb Perspect Med. 2016;6(9): pii: a026153. 10.1101/cshperspect.a026153 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 29. Kruiswijk F, Labuschagne CF, Vousden KH: p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat Rev Mol Cell Biol. 2015;16(7):393–405. 10.1038/nrm4007 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 30. Allen MA, Andrysik Z, Dengler VL, et al. : Global analysis of p53-regulated transcription identifies its direct targets and unexpected regulatory mechanisms. eLife. 2014;3:e02200. 10.7554/eLife.02200 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Verfaillie A, Svetlichnyy D, Imrichova H, et al. : Multiplex enhancer-reporter assays uncover unsophisticated TP53 enhancer logic. Genome Res. 2016;26(7):882–95. 10.1101/gr.204149.116 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 32. Martins CP, Brown-Swigart L, Evan GI: Modeling the therapeutic efficacy of p53 restoration in tumors. Cell. 2006;127(7):1323–34. 10.1016/j.cell.2006.12.007 [DOI] [PubMed] [Google Scholar]
  • 33. Xue W, Zender L, Miething C, et al. : Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature. 2007;445(7128):656–60. 10.1038/nature05529 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Ventura A, Kirsch DG, McLaughlin ME, et al. : Restoration of p53 function leads to tumour regression in vivo. Nature. 2007;445(7128):661–5. 10.1038/nature05541 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 35. Wang B, Xiao Z, Ren EC: Redefining the p53 response element. Proc Natl Acad Sci U S A. 2009;106(34):14373–8. 10.1073/pnas.0903284106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Williams AB, Schumacher B: p53 in the DNA-Damage-Repair Process. Cold Spring Harb Perspect Med. 2016;6(5): pii: a026070. 10.1101/cshperspect.a026070 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 37. Sammons MA, Zhu J, Berger SL: A Chromatin-Focused siRNA Screen for Regulators of p53-Dependent Transcription. G3 (Bethesda). 2016;6(8):2671–8. 10.1534/g3.116.031534 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38. Tanaka H, Arakawa H, Yamaguchi T, et al. : A ribonucleotide reductase gene involved in a p53-dependent cell-cycle checkpoint for DNA damage. Nature. 2000;404(6773):42–9. 10.1038/35003506 [DOI] [PubMed] [Google Scholar]
  • 39. Gong L, Gong H, Pan X, et al. : p53 isoform Δ113p53/Δ133p53 promotes DNA double-strand break repair to protect cell from death and senescence in response to DNA damage. Cell Res. 2015;25(3):351–69. 10.1038/cr.2015.22 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 40. Morris GF, Bischoff JR, Mathews MB: Transcriptional activation of the human proliferating-cell nuclear antigen promoter by p53. Proc Natl Acad Sci U S A. 1996;93(2):895–9. 10.1073/pnas.93.2.895 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Moldovan GL, Pfander B, Jentsch S: PCNA, the maestro of the replication fork. Cell. 2007;129(4):665–79. 10.1016/j.cell.2007.05.003 [DOI] [PubMed] [Google Scholar]
  • 42. Lahav G: Oscillations by the p53-Mdm2 feedback loop. Adv Exp Med Biol. 2008;641:28–38. 10.1007/978-0-387-09794-7_2 [DOI] [PubMed] [Google Scholar]
  • 43. Stewart-Ornstein J, Lahav G: p53 dynamics in response to DNA damage vary across cell lines and are shaped by efficiency of DNA repair and activity of the kinase ATM. Sci Signal. 2017;10(476): pii: eaah6671. 10.1126/scisignal.aah6671 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Taylor WR, Stark GR: Regulation of the G2/M transition by p53. Oncogene. 2001;20(15):1803–15. 10.1038/sj.onc.1204252 [DOI] [PubMed] [Google Scholar]
  • 45. Jiang L, Kon N, Li T, et al. : Ferroptosis as a p53-mediated activity during tumour suppression. Nature. 2015;520(7545):57–62. 10.1038/nature14344 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 46. Xie Y, Zhu S, Song X, et al. : The Tumor Suppressor p53 Limits Ferroptosis by Blocking DPP4 Activity. Cell Rep. 2017;20(7):1692–704. 10.1016/j.celrep.2017.07.055 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 47. Liu DS, Duong CP, Haupt S, et al. : Inhibiting the system x C -/glutathione axis selectively targets cancers with mutant-p53 accumulation. Nat Commun. 2017;8:14844. 10.1038/ncomms14844 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. Medvedeva YA, Lennartsson A, Ehsani R, et al. : EpiFactors: a comprehensive database of human epigenetic factors and complexes. Database (Oxford). 2015;2015:bav067. 10.1093/database/bav067 [DOI] [PMC free article] [PubMed] [Google Scholar]; F1000 Recommendation
  • 49. Leonova KI, Brodsky L, Lipchick B, et al. : p53 cooperates with DNA methylation and a suicidal interferon response to maintain epigenetic silencing of repeats and noncoding RNAs. Proc Natl Acad Sci U S A. 2013;110(1):E89–98. 10.1073/pnas.1216922110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Estève PO, Chin HG, Pradhan S: Human maintenance DNA (cytosine-5)-methyltransferase and p53 modulate expression of p53-repressed promoters. Proc Natl Acad Sci U S A. 2005;102(4):1000–5. 10.1073/pnas.0407729102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Nieto M, Samper E, Fraga MF, et al. : The absence of p53 is critical for the induction of apoptosis by 5-aza-2'-deoxycytidine. Oncogene. 2004;23(3):735–43. 10.1038/sj.onc.1207175 [DOI] [PubMed] [Google Scholar]
  • 52. Wang YA, Kamarova Y, Shen KC, et al. : DNA methyltransferase-3a interacts with p53 and represses p53-mediated gene expression. Cancer Biol Ther. 2005;4(10):1138–43. 10.4161/cbt.4.10.2073 [DOI] [PubMed] [Google Scholar]
  • 53. Burgess A, Chia KM, Haupt S, et al. : Clinical Overview of MDM2/X-Targeted Therapies. Front Oncol. 2016;6:7. 10.3389/fonc.2016.00007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54. Miranda PJ, Buckley D, Raghu D, et al. : MDM4 is a rational target for treating breast cancers with mutant p53. J Pathol. 2017;241(5):661–70. 10.1002/path.4877 [DOI] [PubMed] [Google Scholar]
  • 55. Bourdon JC: p53 isoforms change p53 paradigm. Mol Cell Oncol. 2014;1(4):e969136. 10.4161/23723548.2014.969136 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56. Liu J, Zhang C, Zhao Y, et al. : MicroRNA Control of p53. J Cell Biochem. 2017;118(1):7–14. 10.1002/jcb.25609 [DOI] [PubMed] [Google Scholar]; F1000 Recommendation
  • 57. Sharathchandra A, Katoch A, Das S: IRES mediated translational regulation of p53 isoforms. Wiley Interdiscip Rev RNA. 2014;5(1):131–9. 10.1002/wrna.1202 [DOI] [PubMed] [Google Scholar]
  • 58. Nations U: World Population Propsects. (United Nations, New York).2015. Reference Source [Google Scholar]
  • 59. Nowak R: Walker’s Mammals of the World.(Johns Hopkins University Press).1999. Reference Source [Google Scholar]
  • 60. Prevention, C. f. D. C. a. (US).2017. [Google Scholar]

Articles from F1000Research are provided here courtesy of F1000 Research Ltd

RESOURCES