Take a look at the Recent articles

p53-dependent cell cycle checkpoint after DNA damage and its relevance to PARP1

Tadashige Nozaki

Department of Pharmacology, Faculty of Dentistry, Osaka Dental University, Japan

E-mail : bhuvaneswari.bibleraaj@uhsm.nhs.uk

Mitsuko Masutani

Department of Frontier Life Sciences, Graduate School of Biomedical Sciences, Nagasaki University, Japan

Division of Cellular Signaling, Laboratory of Collaborative Research, National Cancer Center Research Institute, Japan

DOI: 10.15761/RRI.1000137

Article
Article Info
Author Info
Figures & Data

Abstract

The poly(ADP-ribose) polymerase (PARP) inhibitors, including 3-aminobenzamide (3-AB), suppress G1 arrest after DNA damage following gamma-irradiation, suggesting that PARP1, a major PARP family protein, is involved in the induction of G1 arrest. Furthermore, p53 stabilization following gamma-irradiation is not inhibited, but the p53-responsive transient increases of WAF1/CIP1/p21 and MDM2 mRNA have been shown to be suppressed by 3-AB. Therefore, it is suggested that PARP1 participates as a downstream mediator of p53 dependent signal-transduction pathway through the modulation of WAF1/CIP1/p21 and MDM2 mRNA expression. In this review, we discuss p53 cell cycle checkpoint after DNA damage, and its relevance to PARP1. Moreover, the role of PARP1 as a sensor of DNA damage will be proposed. Regulation of p53 and PARP1 activities is an attractive and promising target for the development of clinical treatments for particular diseases. Therefore, it is anticipated that the clinical application of drugs that specifically regulate PARP1 activity will develop in the near future.

Key words

PARP1, p53, signal transduction pathway, G1 arrest

p53 and G1 checkpoint in cancer

During the development of cancer, multiple abnormalities occur in the genes that are directly related to the regulation of cell cycle progression [1]. Mutations in the retinoblastoma (RB) protein, a cell cycle regulatory protein with tumor-suppressive functions, have been reported to occur in various types of cancers [2]. Further, cyclin D gets activated by chromosomal translocation and/or amplification in many cancers [3]. Additionally, p16/MTS1, which inhibits the cyclin D dependent kinase activity, was identified as a novel tumor suppressor gene that gets inactivated in melanoma, colon cancer, breast cancer, etc. [4,5]. Mutations in the p53 tumor suppressor gene are involved in approximately 50 % of the human cancers [6,7]. Among 75 % of these cancers, missense mutations with amino acid substitutions were detected [8]. According to Tsuchida et al. p53 mutations occur in approximately 90 % of the cell cultures obtained from human oral squamous cell carcinoma [9].

p53 induces the transcription of p53 target genes that exhibit various functions involved in the regulation of DNA damage, aging, cancer, gene activation, hypoxia stress, etc. As diversities exist in the p53 target genes, in addition to the conventional functions such as apoptosis, cell cycle arrest, DNA repair, and p53 activity suppression, other functions including cellular development, immunity, epigenetic regulation, and undifferentiated cell state maintenance were reported [10]. An important p53 function during cell cycle arrest (following p53 stabilization after the DNA damage) is the induction of cyclin-dependent kinase inhibitor 1 (WAF1/CIP1/p21). WAF1/CIP1/p21 is a transcription regulatory factor that inhibits the cyclin dependent kinase (CDK) and delays or terminates the cell cycle at G1 phase [9,11-13]. The G1 phase arrest checkpoint mechanism regulates the normal cell cycle and is believed to be strongly involved in carcinogenesis subsequent to DNA damage. In cells with p53 mutations or deficiency, the G1 phase arrest does not occur, and progression to the S phase occurs even after DNA damage owing to the abnormal transcriptional regulation by p53 [11,14]. Therefore, it is hypothesized that DNA damage accumulation causes mutated cells to progress into a cancer.

DNA damage after exposure to ionizing radiation

Regarding the resistance to cancer treatment using radiation or chemotherapy, cell cycle arrest in the G1 and G2 phases after DNA damage is considered as an important factor. It is known that the expression levels of cancer genes such as the rat sarcoma oncogene homolog (RAS), myelocytomatosis viral oncogene homolog (MYC), and rapidly accelerated fibrosarcoma (RAF) vary with the radio-sensitivity of cells [15-17], and therefore the cellular response to cell cycle arrest after DNA damage is considered as an important factor and one of mechanisms that determine the radio-sensitivity. Gamma rays produce excited and ionic molecules in cells. Although all intracellular molecules are targeted, DNA damage is broadly classified into direct and indirect effects [18]. The direct effect is caused by direct contact of the ionizing radiation energy with the DNA. The indirect effect is caused by the interaction of free radical species with the DNA. As water molecules (H2O) are the most frequently available intracellular molecules, radicals —including hydroxyl radical (HO·), hydrogen atom (H·), and hydrated electron (e–aq)— are generated from H2O. Therefore, H2O is considered as the molecule that causes the strongest indirect effect. The damage owing to these direct and indirect ionizing radiation effects occurs at the base or sugar–phosphodiester framework that constitutes the DNA. Although the detection of actual damage to the bases present in cells is difficult owing to their instability, the damage to the sugar–phosphodiester framework is detectable as it mainly appears in the form of DNA strand breaks (DSBs). DSBs mainly occur as phosphodiester bond cleavages and, to some extent, as the decomposition of deoxyribose rings.

p53 stabilization after DNA damage

After DNA damage by gamma rays, cell cycle arrest mainly occurs in the G1 and G2 phases. DNA repair is believed to occur prior to the progression of cell cycle into the S or M phase. Kastan et al. suggested the aforementioned fact that the p53 gene product is the key molecule for G1 phase arrest [19]. They investigated the responses of various cells against gamma rays and reported that p53-deficient cells exhibit G2 phase arrest but not G1 phase arrest after gamma-irradiation. Furthermore, in cells treated with ultraviolet (UV) rays, 4-nitro, 4-nitrosoquinoline-1-oxide, or DNA damaging factors—including gamma rays and actinomycin D—an increase in the p53 level was observed owing to the post-translational mechanism that prolongs the p53 half-life [20,21]. Previous reports demonstrated that p53 recognizes the DNA strand broken ends and binds to these strands [22,23]. It is considered that p53 with an extended half-life gets accumulated in cells, activates or suppresses the transcription of gene targets that contain a p53 binding sequence, and induces G1 phase arrest.

Transcriptional response induced by p53 accumulation

p53 forms a tetramer and functions as a transcription factor by binding to its target genes using the p53 consensus binding sequence (a sequence comprising a 10mer sequence composed of RRRCWWGYYY [R: A/G, Y: T/C, W: A/T] is repeated twice in tandem separated by a spacer of 0-13 bp) [24,25]. p53 activates the transcription of genes, such as the growth arrest and DNA-damage-inducible protein (GADD45), mouse double minute 2 homolog (MDM2), WAF1/CIP1/p21, epidermal growth factor receptor (EGFR), and muscle creatine kinase (CK) [26-28]. Furthermore, it is considered that p53 binds to the large T antigen, which is an oncogene product of DNA-type cancer virus SV40 [2,29], E1B of adenovirus [30], and E6 of papilloma virus. These virus-derived oncoproteins inhibit p53 transcriptional activity and cause cell transformation [31]. Moreover, the oncogene product MDM2 was reported to directly bind to p53 [32,33]. In naturally transformed cells, MDM2 is abnormally amplified [34], and this aberrant MDM2 occurs in approximately 30 % and 15 % of human osteosarcoma and breast cancer, respectively. It is believed that MDM2 ubiquitinates, degrades, and binds to p53, and that it participates in a negative feedback regulation of p53 to inhibit its transcription regulatory activity [35,36]. A low molecular weight compound that binds to MDM2/MDMX and suppresses the degradation of wild-type p53 by inhibiting the interaction between MDM2 and p53 has attracted attention as a molecular target drug, and clinical trials have been conducted in this regard [37].

G1 checkpoint after DNA damage

The information regarding the G1 phase arrest mechanism after DNA damage is described in this section. After DNA damage, the stabilization and intracellular accumulation of p53 takes place, therefore, its transcriptional activation ability increases. As a transcription regulation factor, p53 induces the expression of a protein that suppresses the activity of G1 cyclin and CDK complex enzyme [38]. This protein was identified as WAF1/CIP1/p21 and found to be similar to the senescent cell derived inhibitor 1 (SDI1) gene product whose expression increases with cellular aging [39]. Furthermore, it is considered that via the inhibition of RB phosphorylation by the CDK complex and the activity of transcriptional regulatory factors (such as E2F, which is regulated by RB), WAF1/CIP1/p21 hinders the transcription of the gene necessary for G1/S transition, and therefore causes G1 phase arrest. The process from DNA damage to the increase of p53 level is elucidated in the subsequent sentences. It was reported that in cell cultures derived from B lymphocytes of ataxia telangiectasia patients, an increase of p53 level after DNA damage was not observed [19]. Additionally, it was reported that the causative gene product of ataxia is involved in the signaling pathway leading from DNA damage to p53 elevation.

The G1 phase arrest after DNA damage and apoptosis are known to be closely associated. For instance, in B lymphocytes exposed to gamma irradiation, the G1 phase arrest and apoptosis are induced in the presence and absence of growth factor, respectively [38]. Similar to the G1 phase arrest, apoptosis might be accompanied by p53 stabilization [38]. Therefore, the quantitative information of DNA damage might act as the determining factor in the process of cell decision to opt either for G1 phase arrest or apoptosis after gamma irradiation. In other words, there is a possibility that cells opt for G1 phase arrest when DNA damage is considerably low, and apoptosis when damage is high. The important question is, by what sensor is the quantitative information of DNA damage perceived, and through which signal is this quantitative information transmitted to the key molecule, p53?

G2 checkpoint after DNA damage

Unlike the G1 phase arrest mechanism G2 phase arrest is observed in cells with p53 mutation. For a long time, genetic analyses have been performed in budding yeast, and six genes (RAD9, RAD17, RAD24, MEC3, MEC1, and MEC2) have been identified to be essential for G2 phase arrest [40,41]. Moreover, among the aforementioned genes, MEC1 and MEC2 are indispensable in the step in which the completion of DNA replication occurs. The human homologue of budding yeast RAD24 was identified and found to be identical to a mammalian cell factor that promotes the ADP-ribosylation reaction of bacterial mono (ADP-ribose) transferase and is called 14-3-3 protein [42]. Since the 14-3-3 protein binds to the middle T antigen of polyoma virus [43], it is considered to be involved in DNA damage signaling between the cell membrane and the nucleus. However, as previously described, the analysis of the G2 phase arrest mechanism in mammalian cells has not progressed as much as in yeast cells. Consequently, whether the sensor of DNA damage in G2 phase arrest is distinct to that of the G1 phase arrest remains unknown, as well as the mechanisms through which the quantitative information of DNA damage is transmitted.

Poly(ADP-ribose) synthase 1 (PARP1) is involved in the physiological responses to DSBs in the nuclei of highly evolved eukaryotes, such as mammals. PARP1 specifically recognizes DSBs and promptly synthesizes poly (ADP-ribose) chains using β-nicotinamide adenine dinucleotide (β-NAD) as the substrate. It is known that PARP1 constitutively exists in the nucleus of particular cells in the ratio of approximately 1 per 10 kb DNA [44]. PARP1 activity occurs in most eukaryotic cellular nuclei, including slime molds, animals and plants [45,46]. As an exception, PARP1 activity is not observed in mature granulocytes (leukocytes with rod-shaped or segmented nuclei) in mammals [47]. Most PARP1 is present in the chromatin of the nucleus [48,49]. Immunohistological observations have shown that it is found in the periphery of nucleus (heterochromatin region) in some types of cells [47]. Analyses of protein and gene levels indicate that PARP1 exhibits three functional domains that are well conserved throughout various species [50]. PARP1 binds to a DNA nick through its Zn finger that spreads over 30 bases on the periphery of the nick portion [51,52]. Furthermore, the initial PARP1 activation occurs by its binding to the ends of either single- or double-strand DNA breaks [53]. PARP1 catalyzes the poly ADP-ribosylation of proteins containing glutamic and aspartic acid residues. Histone H1 and H2B, high-mobility group (HMG) proteins, DNA polymerase α and β, and topoisomerase I and II are well-known acceptor proteins of poly (ADP-ribose). It was reported that poly ADP-ribosylation inhibits the enzyme activity of acceptor proteins and causes histones to lose affinity toward the DNA. Moreover, when PARP1 loses its affinity toward the DNA, its ability to synthesize poly (ADP-ribose) is suppressed owing to the auto-poly-ADP-ribosylation reaction [54-56].  Studies regarding the physiological functions of PARP1 were performed in the presence of a PARP inhibitor, and the involvement of PARP1 in DNA repair was identified by enhancing cytotoxicity through various DNA damaging treatments such as exposure to gamma rays [57,58]. Since PARP1 specifically recognizes DNA broken ends, it is believed to participate in DNA repair, by removing the damaged DNA portion after breakage. It was reported that DNA repair against alkylating agents is inhibited exclusively in mutant cells with low PARP1 expression or dominant negative mutants that contain a high number of DNA binding sites [59]. Furthermore, in an experiment, the DNA repair was delayed when PARP1 anti-sense RNA was ectopically expressed in cells to inhibit PARP1 function [60]. Moreover, in the cell-free DNA repair system, DNA repair is temporarily interrupted depending on the PARP1 efficiency in the presence of NAD [61]. Additionally, other reports have indicated that PARP1 highly promotes DNA ligase activity on the chromatin DNA [62]. The previously proposed histone-shuttling model indicates that the loss of affinity of poly-ADP-ribosylated histones for DNA causes the structure surrounding DNA broken ends to further loosen, and thereby promotes DNA repair [63]. PARP1 is necessary for DNA replication as approximately 10 Kb DNA replication intermediates were accumulated in cells upon treatment with a PARP1 inhibitor [64]. Additionally, owing to the facts such as an increase in the sister chromosome conversion frequency [65], loss of amplified c-myc in HL-60 cells [66], the loss of external cancer gene in NIH 3T3 cells [67] suggested that PARP1 is involved in DNA recombination. Regarding gene transcription, experiments using PARP1 anti-sense RNA expression plasmid demonstrated that the induction of major histocompatibility complex (MHC) class II gene expression by gamma interferon (γ-IFN) is in turn induced by PARP1 inhibition [68]. Moreover, it was reported that the transcription from the HIV long terminal repeat induced during DNA damage by UV can be suppressed by a PARP inhibitor [69]. Therefore, the involvement of PARP1 in various physiological conditions that cause DSBs is well established. We found that PARP inhibitors, including 3-aminobenzamide (3-AB), suppressed G1 arrest after dNA damage following gamma irradiation, suggesting that PARP1 is critical for the induction of G1 arrest. Furthermore, we found that p53 stabilization was not inhibited by gamma-irradiation, and that the p53-responsive transient increases of WAF1/CIP1/p21 and MDM2 mRNA were suppressed by 3-AB. Therefore, it is suggested that PARP1 participates in p53 dependent G1 arrest signal-transduction pathway through the modulation of WAF1/CIP1/p21 and MDM2 mRNA expression [70,71].

The role of PARP1 as a sensor of DNA damage

As PARP1 is constitutively expressed in the nucleus, it specifically recognizes DSBs and synthesizes the poly (ADP-ribose) strands corresponding to the number of DSBs using NAD as substrate. Hence, it is considered as a promising sensor candidate that monitors DNA damage. A prompt signal for DSBs and an accurate transmission of the vast information regarding DSBs are considered as necessary capabilities of a sensor molecule. After gamma-irradiation, PARP1, present in the ratio of approximately one per several Kb of DNA, rapidly synthesizes poly (ADP-ribose) in a dose dependent manner by consuming NAD. For example, 2 Gy gamma-irradiation produces approximately 1,000 single-strand breaks in a cell. Considering that 3 × 10-18 moles of NAD molecules per cell are consumed within 30 min after 2 Gy gamma irradiation, it was calculated that approximately 105 molecules of poly (ADP-ribose) with an average chain length of 20 molecules are produced per cell through DNA strand breakage. Similarly, approximately 5 × 106 molecules of poly (ADP-ribose) may be synthesized after 100 Gy irradiation. Contrarily, as p53 recognizes DSBs, another theory suggests that p53 might act as the direct signal for DSBs [23]. However, an immunoprecipitation experiment indicated that approximately 104 p53 molecules per cell were present under normal conditions, and thus it is unlikely that p53 could directly and efficiently transmit the vast information regarding ≥ 104 DSBs. Therefore, it is considered that poly (ADP-ribose) might be a more effective signaling candidate to transmit the vast information on DSBs. Due to their capacity to bind DNA broken ends, p53 and PARP1 are considered potential signaling molecules for the detection of DSBs in G1 phase arrest. Several p53 isoforms and family members were identified, and multiple reports have indicated that some of the isoforms interact with p53 and transactivate specific target gene groups [72-74]. Moreover, the possibility that the cooperation between the wild-type p53 and some p53 isoforms may transmit the vast information regarding DSBs needs to be explored. If PARP1 transmits quantitative information regarding DSBs, the decision as to whether the direct poly-ADP-ribosylation or the interaction between PARP1 and the other information transmission factors is selected and the information transmission factor needs to be investigated in order to completely understand the mechanism. Previously, using the western blot method we revealed the presence of intracellular proteins that non-covalently interact with PARP1 [75]. In the future, extensive research will be essential to understand the sensor molecules that sense and transmit the vast information regarding DSBs.

Modification of p53 and PARP1 activities and their application in disease treatments

As mutant p53 stabilizes and accumulates in cancer cells, cancer patients often exhibit p53 antibody positivity; therefore, p53 antibody is used as a tumor marker for diagnosis [76]. Drugs that can convert this mutant p53 into wild-type might be useful as therapeutic agents for cancer [77,78]. Anticancer drugs that activate the p53 pathway without causing DNA damage are being developed [79,80]. The development of protective agents that increase the radiation resistance of normal tissues by the regulation of cell death is currently under progress. As p53 regulators selectively protect normal tissues that exhibit proper p53 function against cell death caused by DNA damage, and do not protect cancer cells that exhibit abnormal p53, it might be applied to overcome the dose limitation of radiation therapy and to reduce the side-effects of anticancer drugs [81,82]. The denaturation and deactivation of p53 is known to occur owing to the dissociation of zinc ions that coordinate with the zinc ion binding site in p53 and the substitution of metal ions other than zinc [83]. Additionally, compounds that target the zinc binding site in p53 have been studied. However, drugs that inhibit p53 function increase the risk of carcinogenesis promotion. Therefore, efforts are being undertaken to prevent p53-dependant cell death by inducing the expression of WAF1/CIP1/p21 and suppressing the expression of the p53-upregulated modulator of apoptosis, which promotes cell death [84]. It is anticipated that these p53 target drug discoveries will encourage the development of novel radiation protection agents. Similarly, because PARP1 participates not only in DNA repair but also in the p53-dependent cell-cycle check-point after DNA damage, we believe that cell death control via the regulation of PARP1 activity might be useful for future applications in cancer therapy. Although PARP inhibitors that block PARP family proteins have started to be employed in clinical settings, we predict the development and clinical application of drugs that will specifically regulate PARP1 activity.

Conflicts of interest

The authors have declared that no conflicts of interest exist.

References

  1. Marx J (1994) How cells cycle toward cancer. Science 263: 319-321. [Crossref] 
  2. Lane DP, Crawford LV (1979) T antigen is bound to a host protein in SV40-transformed cells. Nature 278: 261-263. [Crossref] 
  3. Sherr CJ (1993) Mammalian G1 cyclins. Cell 73: 1059-1065. [Crossref] 
  4. Kamb A, Gruis NA, Weaver-Feldhaus J, Liu Q, Harshman K, et al. (1994) A cell cycle regulator potentially involved in genesis of many tumor types. Science 264: 436-440.
  5. Serrano M, Hannon GJ, Beach D (1993) A new regulatory motif in cell-cycle control causing specific inhibition of cyclin D/CDK4. Nature 366: 704-707. [Crossref] 
  6. Hollstein M, Sidransky D, Vogelstein B, Harris CC (1991) p53 mutations in human cancers. Science 253: 49-53. [Crossref] 
  7. Levine AJ, Momand J, Finlay CA (1991) The p53 tumour suppressor gene. Nature 351: 453-456. [Crossref] 
  8. Kato S, Han SY, Liu W, Otsuka K, Shibata H, et al. (2003) Understanding the function-structure and function-mutation relationships of p53 tumor suppressor protein by high-resolution missense mutation analysis. Proc Natl Acad Sci USA 100: 8424-8429.
  9. Sakai E, Tsuchida N (1992) Most human squamous cell carcinomas in the oral cavity contain mutated p53 tumor-suppressor genes. Oncogene 7: 927-933.
  10. Marcel V, Nguyen Van Long F, Diaz JJ (2018) 40 years of research put p53 in translation. Cancers (Basel) 10. [Crossref] 
  11. el-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R, et al. (1993) WAF1, a potential mediator of p53 tumor suppression. Cell 75: 817-825. [Crossref] 
  12. Harper JW, Adami GR, Wei N, Keyomarsi K, Elledge SJ (1993) The p21 Cdk-interacting protein Cip1 is a potent inhibitor of G1 cyclin-dependent kinases. Cell 75: 805-816. [Crossref] 
  13. Xiong Y, Hannon GJ, Zhang H, Casso D, Kobayashi R, et al. (1993) p21 is a universal inhibitor of cyclin kinases. Nature 366: 701-704. [Crossref] 
  14. Dulic V, Kaufmann WK, Wilson SJ, Tisty TD, Lees E, et al. (1994) p53-dependent inhibition of cyclin-dependent kinase activities in human fibroblasts during radiation-induced G1 arrest. Cell 76: 1013-1023.
  15. Sawey MJ, Hood AT, Burns FJ, Garte SJ (1987) Activation of c-myc and c-K-ras oncogenes in primary rat tumors induced by ionizing radiation. Mol Cell Biol 7: 932-935. [Crossref] 
  16. Chang EH, Pirollo KF, Zou ZQ, Cheung HY, Lawler EL, et al. (1987) Oncogenes in radioresistant, noncancerous skin fibroblasts from a cancer-prone family. Science 237: 1036-1039. [Crossref] 
  17. Kasid U, Pfeifer A, Weichselbaum RR, Dritschilo A (1987) The raf oncogene is associated with a radiation-resistant human laryngeal cancer. Science 237: 1039-1041.
  18. Friedberg EC (1985) DNA Repair. New York: W. H. Freeman and company. 54-59.
  19. Kastan MB, Zhan Q, el-Deiry WS, Carrier F, Jacks T, et al. (1992) A mammalian cell cycle checkpoint pathway utilizing p53 and GADD45 is defective in ataxia-telangiectasia. Cell 71: 587-597. [Crossref] 
  20. Tishler RB, Calderwood SK, Coleman CN, Price BD (1993) Increases in sequence specific DNA binding by p53 following treatment with chemotherapeutic and DNA damaging agents. Cancer Res 53: 2212-2216. [Crossref] 
  21. Maltzman W, Czyzyk L (1984) UV irradiation stimulates levels of p53 cellular tumor antigen in nontransformed mouse cells. Mol Cell Biol 4: 1689-1694. [Crossref] 
  22. Oberosler P, Hloch P, Ramsperger U, Stahl H (1993) p53-catalyzed annealing of complementary single-stranded nucleic acids. EMBO J 12: 2389-2396.
  23. Bakalkin G, Yakovleva T, Selivanova G, Magnusson KP, Szekely L, et al. (1994) p53 binds single-stranded DNA ends and catalyzes DNA renaturation and strand transfer. Proc Natl Acad Sci U S A 91: 413-417. [Crossref] 
  24. Jordan JJ, Menendez D, Inga A, Noureddine M, Bell DA, et al. (2008) Noncanonical DNA motifs as transactivation targets by wild type and mutant p53. PLoS Genet 4: e1000104. [Crossref] 
  25. Wang B, Xiao Z, Ko HL, Ren EC (2010) The p53 response element and transcriptional repression. Cell Cycle 9: 870-879. [Crossref] 
  26. Price BD, Park SJ (1994) DNA damage increases the levels of MDM2 messenger RNA in wtp53 human cells. Cancer Res 54: 896-899. [Crossref] 
  27. Deb SP, Munoz RM, Brown DR, Subler MA, Deb S (1994) Wild-type human p53 activates the human epidermal growth factor receptor promoter. Oncogene 9: 1341-1349.
  28. Zambetti GP, Bargonetti J, Walker K, Prives C,Levine AJ (1992) Wild-type p53 mediates positive regulation of gene expression through a specific DNA sequence element. Genes Dev 6: 1143-1152.
  29. Linzer DI, Levine AJ (1979) Characterization of a 54K Dalton cellular SV40 tumor antigen present in SV40-transformed cells and uninfected embryonal carcinoma cells. Cell 17: 43-52.
  30. Sarnow P, Ho YS, Williams J, Levine AJ (1982) Adenovirus E1b-58kd tumor antigen and SV40 large tumor antigen are physically associated with the same 54 kd cellular protein in transformed cells. Cell 28: 387-394. [Crossref] 
  31. Werness BA, Levine AJ, Howley PM (1990) Association of human papillomavirus types 16 and 18 E6 proteins with p53. Science 248: 76-79. [Crossref] 
  32. Oliner JD, Pietenpol JA, Thiagalingam S, Gyuris J, Kinzler KW, et al. (1993) Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature 362: 857-860. [Crossref] 
  33. Olson DC, Marechal V, Momand J, Chen J, Romocki C, et al. (1993) Identification and characterization of multiple mdm-2 proteins and mdm-2-p53 protein complexes. Oncogene 8: 2353-2360. [Crossref] 
  34. Oliner JD, Kinzler KW, Meltzer PS, George DL, Vogelstein B (1992) Amplification of a gene encoding a p53-associated protein in human sarcomas. Nature 358: 80-83. [Crossref] 
  35. Chen CY, Oliner JD, Zhan Q, Fornace AJ Jr, Vogelstein B, et al. (1994) Interactions between p53 and MDM2 in a mammalian cell cycle checkpoint pathway. Proc Natl Acad Sci U S A 91: 2684-2688. [Crossref] 
  36. Fuchs SY, Adler V, Buschmann T, Wu X, Ronai Z (1998) Mdm2 association with p53 targets its ubiquitination. Oncogene 17: 2543-2547. [Crossref] 
  37. Burgess A, Chia KM, Haupt S, Thomas D, Haupt Y, et al. (2016) Clinical Overview of MDM2/X-Targeted Therapies. Front Oncol 6: 7. [Crossref] 
  38. el-Deiry WS, Harper JW, O'Connor PM, Velculescu VE, Canman CE, et al. (1994) WAF1/CIP1 is induced in p53-mediated G1 arrest and apoptosis. Cancer Res 54: 1169-1174. [Crossref] 
  39. Noda A, Ning Y, Venable SF, Pereira-Smith OM, Smith JR (1994) Cloning of senescent cell-derived inhibitors of DNA synthesis using an expression screen. Exp Cell Res 211: 90-98. [Crossref] 
  40. Weinert TA, Hartwell LH (1990) Characterization of RAD9 of Saccharomyces cerevisiae and evidence that its function acts posttranslationally in cell cycle arrest after DNA damage. Mol Cell Biol 10: 6554-6564. [Crossref] 
  41. Murray AW (1992) Creative blocks: cell-cycle checkpoints and feedback controls. Nature 359: 599-604. [Crossref] 
  42. Ford JC, al-Khodairy F, Fotou E, Sheldrick KS, Griffiths DJ, et al. (1994) 14-3-3 protein homologs required for the DNA damage checkpoint in fission yeast. Science 265: 533-535. [Crossref] 
  43. Pallas DC, Fu H, Haehnel LC, Weller W, Collier RJ, et al. (1994) Association of polyomavirus middle tumor antigen with 14-3-3 proteins. Science 265: 535-537. [Crossref] 
  44. 2021 Copyright OAT. All rights reserv
  45. Althaus FR, Richter C (1987) ADP-ribosylation of proteins. Enzymology and biological significance. Mol Biol Biochem Biophys 37: 1-237. [Crossref]
  46. Hilz H, Stone P (1976) Poly(ADP-ribose) and ADP-ribosylation of proteins. Rev Physiol Biochem Pharmacol 76: 1-58, 177. [Crossref] 
  47. Hayaishi O, Ueda K (1977) Poly(ADP-ribose) and ADP-ribosylation of proteins. Annu Rev Biochem 46: 95-116. [Crossref] 
  48. Hayaishi O, Ueda K, Kawaichi, M, Ogata N, Oka J, et al. (1979) Poly(ADP-ribose) and ADP-ribosylation of proteins. In: Russel TR, Brew K, Faber H, Schultz J, eds. From gene to protein: Information transfer in normal and abnormal cells. New York: Academic Press 545-566.
  49. Hilz H, Kittler M (1968) On the localization of poly ADPribose synthetase in the nucleus. Hoppe Seylers Z Physiol Chem 349: 1793-1796. [Crossref]
  50. Uchida K, Morita T, Sato T, Ogura T, Yamashita R, et al. (1987) Nucleotide sequence of a full-length cDNA for human fibroblast poly(ADP-ribose) polymerase. Biochem Biophys Res Commun 148: 617-622.
  51. Kameshita I, Matsuda Z, Taniguchi T, Shizuta Y (1984) Poly (ADP-Ribose) synthetase. Separation and identification of three proteolytic fragments as the substrate-binding domain, the DNA-binding domain, and the automodification domain. J Biol Chem 259: 4770-4776. [Crossref] 
  52. Ikejima M, Noguchi S, Yamashita R, Ogura T, Sugimura T, et al. (1990) The zinc fingers of human poly(ADP-ribose) polymerase are differentially required for the recognition of DNA breaks and nicks and the consequent enzyme activation. J Biol Chem 265: 21907-21913.
  53. Gradwohl G, Menissier-de Murcia J, Molinete M, Simonin F, Kohen M, et al. (1990) The second zinc-finger domain of poly(ADP-ribose) polymerase determines specificity for signal-stranded in DNA. Proc Natl Acad Sci USA 87: 2990-2994.
  54. Benjamin RC, Gill DM (1980) Poly(ADP-ribose) synthesis in vitro programmed by damaged DNA. A comparison of DNA molecules containing different types of strand breaks. J Biol Chem 255: 10502-10508. [Crossref] 
  55. Jump DP, Smulson M (1980) Purification and Characterization of the major nonhistone protein acceptor for poly(adenosine diphosphate ribose) in HeLa cell nuclei. Biochemistry 19: 1024-1030.
  56. Yoshihara K, Hashida T, Yoshihara H, Tanaka Y, Ohgushi H (1977) Enzyme-bound early product of purified poly(ADP-ribose) polymerase. Biochem Biophys Res Commun 78: 1281-1288. [Crossref] 
  57. Ogata N, Kawaichi M, Ueda K, Hayaishi O (1980) Poly(ADP-ribsyl)ation of 110,000 Dalton protein in human lymphocytes treated with N-methyl-N'-nitro-N-nitrosoguanidine. Biochem Int 1: 229-236.
  58. Nduka N, Skidmore CJ, Shall S (1980) The enhancement of cytotoxicity of N-methyl-N-nitrosourea and of gamma-radiation by inhibitors of poly(ADP-ribose) polymerase. Eur J Biochem 105: 525-530. [Crossref] 
  59. Hirai T, Shirai H, Fujimori H, Okayasu R, Sasai K, et al. (2012) Radiosensitization effect of poly(ADP-ribose) polymerase inhibition in cells exposed to low and high liner energy transfer radiation. Cancer Sci 103: 1045-1050.
  60. Molinete M, Vermeulen W, Burkel A, Menissier-de Murcia J, Kupper JH, et al. (1993) Overproduction of the poly(ADP-ribose) polymerase DNA-binding domain blocks alkylation-induced DNA repair synthesis in mammalian cells. EMBO J 12: 2109-2117.
  61. Ding R, Pommier Y, Kang VH, Smulson M (1992) Depletion of poly(ADP-ribose) polymerase by antisense RNA expression results in a delay in DNA strand break rejoining. J Biol Chem 267: 12804-12812. [Crossref] 
  62. Satoh MS, Lindahl T (1992) Role of poly(ADP-ribose) formation in DNA repair. Nature 356: 356-358. [Crossref] 
  63. Creissen D, Shall S (1982) Regulation of DNA ligase activity by poly(ADP-ribose). Nature 296: 271-272. [Crossref] 
  64. Realini CA, Althaus FR (1992) Histone shuttling by poly(ADP-ribosylation). J Biol Chem 267: 18858-18865. [Crossref] 
  65. Lönn U, Lönn S (1985) Accumulation of 10-kilobase DNA replication intermediates in cells treated with 3-aminobenzamide. Proc Natl Acad Sci U S A 82: 104-108. [Crossref]
  66. Oikawa A, Tohda H, Kanai M, Miwa M, Sugimura T (1980) Inhibitors of poly(adenosine diphosphate ribose) polymerase induce sister chromatid exchanges. Biochem Biophys Res Commun 97: 1311-1316.
  67. Shima H, Nakayasu M, Aonuma S, Sugimura T, Nagao M (1989) Loss of the MYC gene amplified in human HL-60 cells after treatment with inhibitors of poly(ADP-ribose) polymerase or with dimethyl sulfoxide. Proc Natl Acad Sci USA 86: 7442-7445.
  68. Nakayasu M, Shima H, Aonuma S, Nakagama H, Nagao M, et al. (1988) Deletion of transfected oncogenes from NIH 3T3 transformants by inhibitors of poly(ADP-ribose) polymerase. Proc Natl Acad Sci U S A 85: 9066-9070. [Crossref] 
  69. Qu Z, Fujimoto S, Taniguchi T (1994) Enhancement of interferon-γ-induced major histocompatibility complex class II gene expression by expressing an antisense RNA of poly(ADP-ribose) synthetase. J Biol Chem 269: 5543-5547.
  70. Yamagoe S, Kohda T, Oishi M (1991) Poly(ADP-ribose) polymerase inhibitors suppress UV-induced human immunodeficiency virus type 1 gene expression at the posttranscriptional level. Mol Cell Biol 11: 3522-3527.
  71. Masutani M, Nozaki T, Wakabayashi K, Sugimura T (1995) Role of poly(ADP-ribose) polymerase in cell-cycle checkpoint mechanisms following gamma-irradiation. Biochimie 77: 462-465.
  72. Wieler S, Gagné JP, Vaziri H, Poirier GG, Benchimol S (2003) Poly(ADP-ribose) polymerase-1 is a positive regulator of the p53-mediated G1 arrest response following ionizing radiation. J Biol Chem 278: 18914-18921.
  73. Bourdon JC, Fernandes K, Murray-Zmijewski F, Liu G, Diot A, et al. (2005) p53 isoforms can regulate p53 transcriptional activity. Genes Dev 19: 2122-2137. [Crossref] 
  74. Rohaly G, Chemnitz J, Dehde S, Nunez AM, Heukeshoven J, et al. (2005) A novel human p53 isoform is an essential element of the ATR-intra-S phase checkpoint. Cell 122: 21-32. [Crossref] 
  75. Ferraiuolo M, Di Agostino S, Blandino G, Strano S (2016) Oncogenic Intra-p53 Family Member Interactions in Human Cancers. Front Oncol 6: 77. [Crossref] 
  76. Nozaki T, Masutani M, Akagawa T, Sugimura T, Esumi H (1994) Non-covalent interaction between poly(ADP-ribose) and cellular proteins: an application of a poly(ADP-ribose)-western blotting method to detect poly(ADP-ribose) binding on protein-blotted filter. Biochem Biophys Res Commun 198: 45-51.
  77. Shimada H, Ochiai T, Nomura F, Japan p53 Antibody Research Group (2003) Titration of serum p53 antibodies in 1,085 patients with various types of malignant tumors: a multiinstitutional analysis by the Japan p53 Antibody Research Group. Cancer 97: 682-689.
  78. Bykov VJ, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E, et al. (2002) Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat Med 8: 282-288. [Crossref] 
  79. Lehmann S, Bykov VJ, Ali D, Andrén O, Cherif H, et al. (2012) Targeting p53 in vivo: a first-in-human study with p53-targeting compound APR-246 in refractory hematologic malignancies and prostate cancer. J Clin Oncol 30: 3633-3639.
  80. Sasaki M, Kawahara K, Nishio M, Mimori K, Kogo R, et al. (2011) Regulation of the MDM2-P53 pathway and tumor growth by PICT1 via nucleolar RPL11. Nat Med 17: 944-951. [Crossref] 
  81. Xu H, Di Antonio M, McKinney S, Mathew V, et al. (2017) CX-5461 is a DNA G-quadruplex stabilizer with selective lethality in BRCA1/2 deficient tumours. Nat Commun 8: 14432. [Crossref] 
  82. Kirsch DG, Santiago PM, di Tomaso E, Sullivan JM, Hou WS, et al. (2010) p53 controls radiation-induced gastrointestinal syndrome in mice independent of apoptosis. Science 327(5965): 593-596. Erratum in: Science 2011 334: 761.
  83. Wang B, Tanaka K, Morita A, Ninomiya Y, Maruyama K, et al. (2013) Sodium orthovanadate (vanadate), a potent mitigator of radiation-induced damage to the hematopoietic system in mice. J Radiat Res 54: 620-629.
  84. Méplan C, Richard MJ, Hainaut P (2000) Metalloregulation of the tumor suppressor protein p53: zinc mediates the renaturation of p53 after exposure to metal chelators in vitro and in intact cells. Oncogene 19: 5227-5236.
  85. Morita A, Takahashi I, Sasatani M, et al. (2018) A chemical modulator of p53 transactivation that acts as a radioprotective agonist. Mol Cancer Ther 17: 432-442. [Crossref]

Editorial Information

Editor-in-Chief

Article Type

Review Article

Publication history

Received date: June 21, 2018
Accepted date: July 09, 2018
Published date: July 11, 2018

Copyright

© 2018 Nozaki T. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Citation

Nozaki T, Masutani M (2017) p53-dependent cell cycle checkpoint after DNA damage and its relevance to PARP1. Res Rev Insights 2: DOI: 10.15761/RRI.1000137

Corresponding author

Tadashige Nozaki

Department of Pharmacology, Faculty of Dentistry, Osaka Dental University, Japan

E-mail : bhuvaneswari.bibleraaj@uhsm.nhs.uk

No Figures