431_ftp

19
Rev. Med. Virol. 2004; 14: 301–319. Published online in Wiley InterScience (www.interscience.wiley.com). Reviews in Medical Virology DOI: 10.1002/rmv.431 Human tumor suppressor p53 and DNA viruses S. Collot-Teixeira, J. Bass, F. Denis and S. Ranger-Rogez* Department of Virology, Limoges University Teaching Hospital, CHRU Dupuytren, 2 avenue Martin Luther King, 87042 Limoges, France SUMMARY Human tumor suppressor protein p53 plays a major role in the cell cycle, orchestrating a number of important genes involved in cell-cycle control and apoptosis, and seems to be one of the most important molecules protecting cells from malignant transformation. Mutations in the p53 gene are observed in about 50% of primary tumors, inducing defective p53 protein no longer capable of binding DNA and of activating transcription. Certain DNA viruses are thought to act in a similar way and may also contribute to the progression of invasive cancer in infected tissue. One of the most effective strategies employed by these viruses is the inhibition of p53 protein by interaction with viral oncoproteins, implying a direct but also an indirect role of these viruses in the impairment of p53 structure and function. This article provides a summary of current knowledge concerning p53 tumor suppressor protein and reviews the different mechanisms adopted by different DNA viruses in undermining p53 function. Copyright # 2004 John Wiley & Sons, Ltd. Accepted: 23 February 2004 INTRODUCTION Human tumor suppressor protein p53 plays a major role in the cell cycle. It is often described as the ‘guardian of the genome’ [1] because of its key role in cell cycle arrest and in inducing apop- tosis when DNA becomes extensively damaged. P53 gene mutations, catalogued in the Interna- tional Association of Cancer Registries (IARC) database (www.iacr.fr), are an extremely frequent feature of many naturally occurring tumors. They are observed in about 50% of primary tumors [2], and in up to 85% of breast, lung and colon cancers. The mutated p53 allele encodes defective protein that is no longer capable of binding DNA and of activating transcription. Certain viruses, impli- cated in human cancers, could participate in cell neoplasia by encoding viral proteins able to inhibit tumor suppressor protein p53 activity (Table 1). It implies a direct but also an indirect role of these viruses in the impairment of p53 structure and function. This article summarises the different roles of p53 in normal and stressed cells, its interactions, activations and inhibitions, in order to better understand the implications of the different mechanisms used by DNA viruses in abrogating p53 functions in cell transformation. P53 Protein description Human p53 is a phosphoprotein of M r 53000 encoded by a 20 kb gene containing 11 exons [3] and located on chromosome 17q13 [4]. This gene belongs to a highly conserved gene family contain- ing at least two other members; p51/p63 [5,6] and p73 [7]. The proteins share many properties and their intricate interplay serves to optimise and reg- ulate cell proliferation. P53 protein contains 393 amino acids and is composed of several structural and functional domains: an amino-terminal domain required for transcription activation, a proline-rich sequence containing multiple copies R RE EVI IEW Copyright # 2004 John Wiley & Sons, Ltd. *Corresponding author: Dr S. Ranger-Rogez, Department of Virol- ogy, Unit EA3175, CHRU Dupuytren, 2 avenue Martin Luther King, 87042 Limoges, France. E-mail: [email protected] Abbreviations used Akt, phosphatidylinositol 3-OH-kinase; ARF, ADP-ribosylation fac- tor 1; BRCA-1, breast cancer 1 gene; CAK, CDK-activating kinase; CBP, CREB binding protein; CDK, cyclin-dependent kinases; CDKI, CDK inhibitor; DP, dimerisation partner of E2F; E2F, E2 pro- moter binding factor; GADD45, growth arrest and DNA damage inducible gene; HBc, hepatitis B viral core protein; HBx, hepatitis B viral X protein; HDM2, human double minute 2; HIF-1, hypoxia- inducible factor 1 alpha; ING, inhibitor of growth family; NES, nuclear export signal; NQO1, NADPH quinone oxidoreductase 1; PARP-1, poly(ADP-ribose) polymerase-1; PCAF, p300/CREB- binding protein associated factor; PCNA, proliferating cell nuclear antigen; PRb, retinoblastoma protein; RGC, ribosomal gene cluster; SV40, simian virus 40; TAFs, TBP-associated factors; TBP, TATA- binding protein; TRRAP, transformation/transcription domain- associated protein; XPC, xeroderma pigmentosum group C

description

art. p53 dna virus

Transcript of 431_ftp

  • Rev. Med. Virol. 2004; 14: 301319.Published online in Wiley InterScience (www.interscience.wiley.com).

    Reviews in Medical Virology DOI: 10.1002/rmv.431

    Human tumor suppressor p53 and DNA virusesS. Collot-Teixeira, J. Bass, F. Denis and S. Ranger-Rogez*Department of Virology, Limoges University Teaching Hospital, CHRU Dupuytren, 2 avenue Martin LutherKing, 87042 Limoges, France

    SUMMARY

    Human tumor suppressor protein p53 plays a major role in the cell cycle, orchestrating a number of important genesinvolved in cell-cycle control and apoptosis, and seems to be one of the most important molecules protecting cellsfrom malignant transformation. Mutations in the p53 gene are observed in about 50% of primary tumors, inducingdefective p53 protein no longer capable of binding DNA and of activating transcription. Certain DNA viruses arethought to act in a similar way and may also contribute to the progression of invasive cancer in infected tissue. One ofthe most effective strategies employed by these viruses is the inhibition of p53 protein by interaction with viraloncoproteins, implying a direct but also an indirect role of these viruses in the impairment of p53 structure andfunction. This article provides a summary of current knowledge concerning p53 tumor suppressor protein andreviews the different mechanisms adopted by different DNA viruses in undermining p53 function. Copyright # 2004John Wiley & Sons, Ltd.

    Accepted: 23 February 2004

    INTRODUCTIONHuman tumor suppressor protein p53 plays amajor role in the cell cycle. It is often describedas the guardian of the genome [1] because of itskey role in cell cycle arrest and in inducing apop-tosis when DNA becomes extensively damaged.

    P53 gene mutations, catalogued in the Interna-tional Association of Cancer Registries (IARC)database (www.iacr.fr), are an extremely frequentfeature of many naturally occurring tumors. Theyare observed in about 50% of primary tumors [2],and in up to 85% of breast, lung and colon cancers.The mutated p53 allele encodes defective protein

    that is no longer capable of binding DNA and ofactivating transcription. Certain viruses, impli-cated in human cancers, could participate in cellneoplasia by encoding viral proteins able to inhibittumor suppressor protein p53 activity (Table 1). Itimplies a direct but also an indirect role of theseviruses in the impairment of p53 structure andfunction.

    This article summarises the different roles of p53in normal and stressed cells, its interactions,activations and inhibitions, in order to betterunderstand the implications of the differentmechanisms used by DNA viruses in abrogatingp53 functions in cell transformation.

    P53

    Protein descriptionHuman p53 is a phosphoprotein of Mr 53000encoded by a 20 kb gene containing 11 exons [3]and located on chromosome 17q13 [4]. This genebelongs to a highly conserved gene family contain-ing at least two other members; p51/p63 [5,6] andp73 [7]. The proteins share many properties andtheir intricate interplay serves to optimise and reg-ulate cell proliferation. P53 protein contains 393amino acids and is composed of several structuraland functional domains: an amino-terminaldomain required for transcription activation, aproline-rich sequence containing multiple copies

    RR EE V II E W

    Copyright # 2004 John Wiley & Sons, Ltd.

    *Corresponding author: Dr S. Ranger-Rogez, Department of Virol-ogy, Unit EA3175, CHRU Dupuytren, 2 avenue Martin LutherKing, 87042 Limoges, France. E-mail: [email protected]

    Abbreviations usedAkt, phosphatidylinositol 3-OH-kinase; ARF, ADP-ribosylation fac-tor 1; BRCA-1, breast cancer 1 gene; CAK, CDK-activating kinase;CBP, CREB binding protein; CDK, cyclin-dependent kinases;CDKI, CDK inhibitor; DP, dimerisation partner of E2F; E2F, E2 pro-moter binding factor; GADD45, growth arrest and DNA damageinducible gene; HBc, hepatitis B viral core protein; HBx, hepatitis Bviral X protein; HDM2, human double minute 2; HIF-1, hypoxia-inducible factor 1 alpha; ING, inhibitor of growth family; NES,nuclear export signal; NQO1, NADPH quinone oxidoreductase 1;PARP-1, poly(ADP-ribose) polymerase-1; PCAF, p300/CREB-binding protein associated factor; PCNA, proliferating cell nuclearantigen; PRb, retinoblastoma protein; RGC, ribosomal gene cluster;SV40, simian virus 40; TAFs, TBP-associated factors; TBP, TATA-binding protein; TRRAP, transformation/transcription domain-associated protein; XPC, xeroderma pigmentosum group C

  • of the PXXP sequence (where X is any amino acid),a central core DNA-binding domain, a flexiblelinker region, a tetramerisation domain and astrongly basic carboxyl-terminal regulatorydomain [8,9] (Figure 1). Most p53 mutations iden-tified in tumors occur at sites located on the centralcore DNA-binding domain [10]. The binding ofp53 to specific DNA could cause significant shiftsof residues on the DNA-binding interface thattranslated into the beta-sheet protein, and bindingof non-specific DNA, causes weak but qualita-tively similar shifts, corresponding to weakerbinding interactions [11].

    Native state p53 protein contains large unstruc-tured regions in its N- and C-terminal parts. It is amodular protein consisting of defined structuredand unstructured regions. Lack of rigid structurecombined with low overall stability has been pro-posed as an explanation for its physiological inter-action with a multitude of associated proteins andthe regulation of its turnover [12].

    General presentationUnder normal circumstances, p53 is a short-livedprotein maintained at very low or undetectablelevels bound to its major inhibitor HDM2 (humandouble minute 2). In response to damaged DNA,nucleotide depletion, hypoxia, oncogene expres-sion, metabolic changes or viral infections, p53accumulates in the nucleus and is activated as atranscription factor [13]. It then stimulates the

    expression of genes implicated in cell cycle arrest,DNA repair or in apoptosis (Table 2).

    Differential transactivation by the p53 transcrip-tion factor is highly dependent on p53 level andpromoter target sequence [14]. In response todamaged DNA, p53 protein undergoes extensivepost-translational modifications of at least 18 sitesalong its length [15] as a result of an activationprocess involving phosphorylations and acetyla-tions of the protein. During this activation process,six serines and two threonines in the N-terminaldomain are phosphorylated (Figure 1). Onehypothesis postulates that phosphorylations ofresidues on the p53 protein transform its tightlyfolded transactivation domain into a more openconformation that can interact with transcriptionfactors, thus enhancing gene expression [16]. Acet-ylations of p53 C-terminal domain, by PCAF(p300/CREB-binding protein associated factor)and by CBP/p300 [17,18] (Figure 1), are also neces-sary for the activation of the protein. Acetylationscould dramatically stimulate the sequence-specificDNA-binding activity of p53, possibly as a directresult of conformational change [18].

    Activation of p53 occurs in three stages (a) anincrease in p53 protein concentration either byenhanced translation or increased half-life; (b)transformation of the p53 protein from latent toactive conformation; and (c) translocation of p53protein from the cytoplasm to the nucleus. P53protein has been found physically associated

    Table 1. Viruses with malignant potential. List of viral proteins associated with p53 and pRb

    Virus Human associated cancer p53 mutations p53 binding pRb binding

    Adenovirus * * E1B55K/E4orf6 E1ASV40 * * Large T Antigen (LTag) LTagBKV/JCV Brain tumors * LTag LTagCMV Colorectal cancer * IE286 IE286EBV Burkitts lymphoma Frequent BZLF1/EBNA-5 EBNA-5HHV-6 Lymphoproliferative disorders Rare DR7 ?HHV-8 Kaposi sarcoma, Primary Infrequent LANA/v IRF/K8/ORF50 ?

    effusion lymphoma,Multicentric Castlemansdisease

    HSV-1 Oropharyngeal cancer (?) * ICP0 ?HBV Hepatocellular carcinoma Occasional HBx/HBc ?HPV (16/18) Cervical carcinoma Infrequent E6 E7

    *Not described.

    302302 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • with tubulin in cellular microtubules, which whenactive, participate with dynein motor protein totransport p53 into the nucleus after damage toDNA [19].

    The active form of p53 protein is tetrameric incomposition [20] (Figure 1). Biogenesis of p53involves co-translational dimerisation of mono-mers followed by post-translational dimerisationof the resulting dimers [21]. Once in its activatedtetrameric form, p53 protein binds with highaffinity to DNA and activates transcription frompromoters with consensus binding site 50-PuPuPu-C(A/T)(T/A)GpyPyPy-30 [22]. Activated p53 also

    binds, via its amino-terminal activation domain,to TAFII31 (TBP associated factor) [23] and TAFII70proteins (co-activators of the p53 protein) [24]which are sub-units of TFIID. The latter is a multi-protein complex comprising the TATA-bindingprotein TBP and approximately ten associated fac-tors (TAFs) which form a key component of theRNA polymerase II transcription machinery [25].

    Activated p53 protein functions principally as atrigger of gene transcription, controlling genesimplicated in cell cycle control, in DNA repair orin apoptosis (Table 2), as explained above. How-ever, it can also act as a negative regulator of geneswhose promoters lack p53-binding sites. Themechanism by which p53 exerts this inhibition isprobably in part through its physical interactionwith basal components of the transcriptionalmachinery such as TBP [26]. However, p53TBPinteraction alone is insufficient to induce repressionof transcription and the mechanism almost cer-tainly involves interactions between p53 and otherfactors such as TAFs that are also normallyrequired to activate transcription [27]. P53 repressesthe activity of promoters that depend on the pre-sence of a TATA-box. Promoters, whose accuratetranscription is directed by a pyrimidine-rich initia-tor element, are immune to the effects of p53 [28].P53 represses transcription of several genes suchas the IL-6 gene [29] and the pRb gene [30]. In addi-tion, it is also known to repress transcriptionmediated by RNA polymerase I [31], RNA poly-merase II [32] and RNA polymerase III [3335].

    P53 inhibition of gene expression can also takeplace after transcription during the translationphase of mRNA. Data presented by Horton et al.[36] demonstrated that p53 causes a rapid decreasein initiation of translation.

    Role of p53 in cell cycle arrest at theG1/S checkpointThe main mechanism employed by p53 to blockthe cell cycle at the G1/S checkpoint is basedon the inhibition of E2F proteins (E2 promoter bind-ing factor), which are transcription factors impli-cated in activation of genes required for G1/Stransition (Figure 2). E2F consists of heterodimerscontaining a sub-unit encoded by the E2F genefamily and a sub-unit encoded by the DP family.Under normal circumstances, E2F transcriptionfactors are blocked by binding to pRb protein orits relatives, p130 and p107 [37,38] thus inhibiting

    Figure 1. Structure of p53 protein. (A) Domains of p53 with bind-

    ing sites of cellular (below the diagram) and viral (above the dia-

    gram) p53 partners. TAD, transcription activation domain; RPR,

    rich proline region; DBD, DNA binding domain; FLR, flexible

    linker region; TD, tetramerisation domain; RD, regulation

    domain. (B) Structure of p53 monomer and tetramer. (C) Structure

    of the C-terminal domain of p53 tetramer as determined by x-raycrystallography. Derived from Clore et al. Science, 1994 and Jef-frey et al. Science, 1995 and modified. The alpha helix of eachmonomer is shown in dark purple, dark blue, orange and dark

    green. P, phosphorylation; Ac, acetylation

    Human tumor suppressor p53Human tumor suppressor p53 303303

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • E2F transcription activity. At the end of the G1phase, pRb is phosphorylated by several cyclin/CDKs (cyclin dependent kinases) inducing achange in pRb conformation and the release ofan active E2F transcription factor. Free E2F acti-vates genes implicated in DNA synthesis drivingprogression of the cell cycle from G1 to S phases,but the exact mechanism of E2F-mediated activa-

    tion of transcription remains unclear. E2Fs alsoregulate the expression of genes involved in differ-entiation, development, proliferation and apopto-sis [39].

    In normal cells, p53 is maintained at a very lowlevel, coupled with HDM2, and exerts no effect.But in response to damaged DNA, a rapid increasein the amount of the active form of p53 protein

    Table 2. List of p53 transcriptionally targeted genes

    Genes activacted by p53 Function Authors

    p21WAF1/Cip1 CDK inhibitor El Deiry et al., 1993Cell cycle regulation Harper et al., 1993

    GADD45 DNA repair Carrier et al., 199414-3-3 DNA repair Hermecking et al., 1994MDM2 Inhibition of p53 Wu et al., 1993IRF-5 Immune response Mori et al., 2002Cyclin K Cell cycle regulation Mori et al., 2002Cyclin G Cell cycle regulation Okamoto et al., 1994B99/Gtse-1 Cell cycle regulation Collavin et al., 2000MUC2 Defence in lumen Ookawa et al., 2002ICAM-1 Cellcell interaction Gorgoulis et al., 2003ATF3 Response to stress cells Zhang, Gao et al., 2002MC610 Apoptosis Zhu et al., 2000Bax Apoptosis Miyashita et al., 1995IGF-BP3 Apoptosis Buckbinder et al., 1995PIG3 Apoptosis Polyak et al., 1997CD95/FAS Apoptosis Owen-schaub et al., 1995p85 Apoptosis Yin et al., 1998KILLER/DR5 Apoptosis Wu et al., 1997Apaf-1 Apoptosis Rozenfeld-Granot et al., 2002Zac-1 Apoptosis Rozenfeld-Granot et al., 2002mRTVP-1 (mouse gene) Apoptosis Ren et al., 2002PRG3 Apoptosis Ohiro et al., 2002PAG 608 Apoptosis Israeli et al., 1997Caspase-1 Apoptosis Gupta et al., 2001Caspase-6 Apoptosis MacLachlan et al., 2002PERP Apoptosis Attardi et al., 2000p53DINP1 Apoptosis Okamura et al., 2001Pidd (mouse) Apoptosis Lin et al., 2000PUMA Apoptosis Nakano et al., 2001Noxa ? Apoptosis Oda et al., 2000Scotin Apoptosis Bourdon et al., 2002PRG3 Apoptosis Ohiro et al., 2002PAC1 Apoptosis Yin et al., 2003TSAP6 Apoptosis Passer et al., 2003

    304304 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • stimulates transcription of genes like p21WAF1/CIP1

    (Table 2) whose product is a CDKI (CDK inhibitor)[40]. The N-terminal domain of p21 interacts withand inhibits CDK-cyclin [41] halting phosphoryla-tion of pRb protein so that E2F remains inacti-vated, bound to pRb (Figure 2). Consequently,the cell cycle is brought to a halt at the G1 phase,allowing DNA repair to take place. P21 also inhi-bits proliferating cell nuclear antigen (PCNA), anauxiliary factor for DNA polymerases and ",facilitating loading of the polymerases onto DNAtemplates and increasing their activity in bothDNA replication and repair. Inhibition of PCNAby p21 is probably due to its ability to mask ele-ments on PCNA required for the binding of com-ponents of the polymerase assembly [42].Interestingly, PCNA-dependent DNA replicationis inhibited [43,44], whereas PCNA-dependentnucleotide excision-repair is not affected [4547].

    Once the G1/S checkpoint is passed, replicationoccurs normally, and p53 acts as a second regula-tor to avoid re-replication [48]. Failure to exert theS phase control leads to genetic instability [48].

    Role of p53 in cell cycle arrest at theG2/M checkpointP53 is also involved in cell cycle regulation at theG2/M restriction point [49] by a mechanism invol-ving mainly the inhibition of CDK1 (also calledcdc2), a cyclin-dependent kinase required to enterthe mitosis phase [50]. An intact p53 C-terminaldomain is necessary to induce G2 arrest [51].

    In normal cells, the complex CDK1/cyclin B [52]determines G2/M progression [53]. At the end ofthe G2 phase, CDK1/cyclin B complex is phos-phorylated by CAK (CDK-activating kinase),becomes active and allows the cell to progressinto its mitotic phase.

    In damaged cells, p53 inhibits CDK1 through atleast three mechanisms: (1) p53 transcriptionallyactivates three inhibitors of CDK1: GADD45,p21Cip1/Waf1 and 14.3.3 [54] (Figure 3). GADD45was found to directly inhibit the activity of theCDK1/cyclin B complex by physical interactionwith CDK1 but not with cyclin B [55]. P21 is a uni-versal CDK/cyclin inhibitor essential for control-ling the G2 checkpoint [56]. 14.3.3 is needed toprevent mitotic catastrophe after DNA damage.It anchors CDK1 in the cytoplasm where it isunable to phosphorylate substrates required forentry into mitosis [57]. (2) P53 can regulate thecell cycle checkpoint, not only via transcription,but also by interaction with CDK1, binding to thecatalytic sub-unit of CDK1/cyclin B kinase [58]. (3)P53 also negatively regulates CDK1 transcriptionvia the NF-Y transcription factor [59].

    However, p53 is also implicated in the G2/Mcheckpoint by its ability to inhibit DNA topoi-somerase II alpha gene expression [60]. Topoi-somerase II allows the higher order compactionof chromatin, required to form a mitotic chromo-some [61]. Inhibition of topoisomerase II blockscells in G2 because the chromatin is not condensedand decatenated [62].

    Figure 2. Role of p53 in the G1/S cell cycle checkpoint. CdK, cyclin-dependent kinase; E2F, E2 promoter binding factor; DP, dimerisation

    partner of E2F; GADD45, growth arrest and DNA damage inducible gene; pRb, retinoblastoma protein; PCNA, proliferating cell nuclear

    antigen

    Human tumor suppressor p53Human tumor suppressor p53 305305

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • A study carried out by Minemoto et al. [63] alsosuggests a relationship between p53 and ubiquitin-mediated degradation of mitotic cyclins and possi-ble interactions between the G2-DNA damagecheckpoint and the mitotic checkpoint.

    Role of p53 in DNA repairP53 activates genes like GADD45 implicated inDNA repair by stimulating base excision repairvia PCNA [64,65]. GADD45 stimulates DNA exci-sion in vitro and inhibits entry of cells into S phase[66]. GADD45 also recognises an altered chroma-tin state and modulates accessibility to cellularproteins [67] (Figure 2).

    But the exact mechanism of p53 in DNA repairis still unclear. P53 could induce the recruitment ofnucleotide excision repair factors XPC (xerodermapigmentosum group C) and TFIIH to damagedDNA [68]. Therefore, p53 seems to play a signifi-cant role in damage recognition and subsequentassembly of repair machinery during global geno-mic repair, with an in vivo requirement for p53 inregulating the base excision repair response [69].Moreover, evidence was also found of physicaland functional interactions between p53 andDNA polymerase alpha-primase, an enzymeinvolved in DNA repair after damage provokedby stress [70]. P53 could also act as a chromatin

    accessibility factor, mediating UV-induced globalchromatin relaxation prior to global nucleotideexcision repair; a fundamental component ofDNA repair, which plays a direct role in rectifyingDNA damage [71].

    P53 also exhibits 30-to-50 exonuclease activityand it has been suggested that p53 might act asan external proofreader [72]. This activity seemsto be intrinsic to wild-type p53, being associatedwith the central core DNA-binding domain ofthe p53 protein structure [73]. P53, through itsexonuclease activity, could be actively involvedin detecting mismatched nucleotide pairs andin effecting nucleotide excision repair and recom-bination. Further evidence was added in amore recent study demonstrating that the exonu-clease role of p53 primarily consists of excis-ing mismatched nucleotides from DNA, soenhancing the DNA replication fidelity of poly-merase [74].

    This newly discovered DNA repair role of p53,significantly extends the importance of p53 inmaintaining genomic integrity as the guardian ofthe genome. Some authors, however, have recentlyquestioned these conclusions, since p53-associated30-50 exonuclease activity has been detected innuclear and cytoplasmic compartments of cellsand also because wild-type p53 in its non-induced

    Figure 3. Regulation of the G2/M checkpoint by p53. CDK1, cyclin-dependent kinase 1; P, phosphorylation; GADD45, growth arrest and

    DNA damage inducible gene; CAK, CDK-activating kinase

    306306 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • state in the cytoplasm appears to exert 30-50

    exonuclease activity [75]. Therefore the preciserole of p53 in DNA repair still remains unclear.

    Role of p53 in apoptosisApoptosis is a genetically programmed process,which controls many biological events, includingcell homeostasis, tissue morphogenesis, develop-ment of the immune system and aging. It is alsoan important cellular defence against cancer andviral infection. P53 plays a critical role in apoptosis.The mechanisms by which p53 modulates activa-tion of apoptosis remain imperfectly understoodbut seem to be localised in the C-terminal basic reg-ulation domain between residues 364393 and theproline-rich domain between 6491 in the p53 pro-tein [76]. P53 is able to transactivate the transcrip-tion of range of proapoptotic genes including forexample, Bax [77] and CD95/FAS [78] (Table 2).

    P53 appears to activate the apoptotic machineryby triggering translocation of Bax to mitochondria,with a concomitant release of cytochrome c frommitochondria into the cytosol. The cytochrome crelease mechanism induced by p53 was shown todepend on the presence of cytosolic Bax [79]. Inaddition p53 triggered apoptosis in oncogene-expressing fibroblasts seems to be inducedthrough the induction of Noxa and by mitochon-drial Bax translocation. At a cellular level,enforced expression of p53 is sufficient to induceBax translocation and cytochrome c release [80].This then causes activation of several caspasesincluding caspase 8, which might amplify theapoptotic signal by activating caspase 9 and itsdownstream caspases [81]. Cytochrome c bindsto APAF-1 (apoptotic protease activating factor 1)protein and participates in activation of the cas-pase cascade leading to DNA cleavage [82,83]and consequently to cell death. Nakano andcolleagues suggested that PUMA plays a role inmediating p53-induced cell death through thecytochrome c/APAF-1-dependent pathway [84].

    P53 can also directly transactivate the transcrip-tion of caspase-1 [85,86] and caspase-6 [87]. Noneof the members of the caspase family had beenidentified as a direct transcriptional target of p53before caspase-1. The executioner caspase 6 isdirectly regulated by p53, provoking a loweringof the cell death threshold in response to apoptoticsignals. MacLachlan et al. [87] therefore suggestedsuch a model might provide a potential mechan-

    ism for lowering cell-death thresholds and be ofpractical use in chemosensitisation.

    A fraction of p53 seems to translocate directly tothe mitochondria of apoptosing tumor cells. Tar-geting p53 to mitochondria is sufficient to launchapoptosis. P53 translocation to mitochondria couldoccur in vivo in irradiated thymocytes. Further, p53could directly induce permeabilisation of the outermitochondrial membrane by forming complexeswith protective BclXL and Bcl-2 proteins, resultingin cytochrome c release. Tumor-derived transacti-vation-deficient mutants of p53 concomitantly losethe ability to interact with BclXL and to promotecytochrome c release. This implies that mutationsmight represent double-hits by abrogating thetranscriptional and mitochondrial apoptotic activ-ity of p53 [88].

    During apoptosis, p53 can also inhibit the tran-scription of anti-apoptotic genes such as Bcl-2,which protect cells against programmed death[89]. Bcl-2 and Bax, which have opposite effects,belong to the same family. Bcl-2 associates withBax, and it is hypothesised that the response of acell to a death signal is determined by the ratioof Bcl-2 to Bax [90].

    The difference between apoptosis and growtharrest mediated by p53 could result in modulationof the cellular response to p53 by proliferative sig-nals. Studies of p53 signaling upon -irradiation ofprimary mouse lymphocytes demonstrated thatG(0) lymphocytes rapidly went into p53-depen-dent apoptosis whereas stimulated lymphocyteswent into a p53-dependent, p21-mediated growtharrest. The switch in p53 response upon stimula-tion was not the result of a switch in transcrip-tional activation of major p53 target genes butwas mediated by proliferative signals from thecells [91].

    Activation of p53The post-translational modifications seen aboveonly represent a partial picture of p53 functionalregulation. Final p53 activation is a complex me-chanism also affected by a range of proteins thateither enhance or suppress p53 activation throughdifferent pathways in response to cell stress.

    Many cellular proteins are able to augment p53activity under stress conditions (Figure 4). Theprincipal p53 activation pathway appears to bevia HDM2 (main p53 inhibitor) inhibition leadingto an inhibition of p53 degradation. For example,

    Human tumor suppressor p53Human tumor suppressor p53 307307

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • activation of p53 under cell stress conditions dep-ends to a great extent on the presence of c-Abl(Abelson murine leukemia cellular oncogene)which protects p53 from the inhibitory effect ofHDM2 [9295]. P14ARF/p19ARF (also called ARFfor ADP-ribosylation factor 1), with its collaboratorCARF [96] are also strong activators of p53. ARFbinds to HDM2 and promotes its rapid degrada-tion [97]. Other cellular proteins share similarp53 activating mechanisms such as HIF-1 (hypox-ia-inducible factor 1 alpha) [98], ING (inhibitor ofgrowth family) family gene, p47ING3 [99,100].

    There also exist a number of other p53 activationpathways unrelated to HDM2 such as BRCA-1(breast cancer 1 gene) [101104] and PARP-1(poly(ADP-ribose) polymerase-1) [105,106].

    Inhibition of p53HDM2 (often called MDM2: murine double min-ute) probably constitutes the major inhibitor ofp53 (Figure 4). It binds to the transactivationdomain of p53 whilst the protein is localised inthe nucleus (Figure 1). Although the N-terminaldomain of p53 contains the primary docking siteof HDM2, a second HDM2 binding interface with-in the linker region of the core domain of p53 hasbeen described [107].

    HDM2 inhibits the ability of p53 to activate tran-scription of its downstream target genes by two

    known mechanisms: (1) binding directly to thetransactivation domain of p53 and blocking itsinteraction with TBP-associated factors TAFII70and TAFII31 (see details above) [23,24,108]; (2) tar-geting p53 for degradation via the ubiquitin-26Sproteasome pathway [109]. HDM2 functions as aubiquitin ligase for p53 [110] mediating ubiquiti-nation of p53 and allowing it to be recognisedand degraded by the proteasome [111]. HDM2-mediated p53 proteasomal degradation seems tobe regulated by NAD(P)H quinone oxidoreductase1 (NQO1) [112,113] and the HDM2/p53 complex isexported from the nucleus using HDM2 nuclearexport signal (NES). Nuclear export of p53 seemsto be necessary before the protein can be degraded[114]. Regulatory mechanisms exist to controlexport of p53 into the cytoplasm, such as the exis-tence of two NES on the p53 protein itself [115].But in fact, nuclear degradation of p53 could occurduring down-regulation of the p53 response afterDNA damage. And local p53 degradation withinthe nucleus could provide a tighter and faster con-trol route during the down-regulatory phase [116].

    HDM2 also inhibits p300-mediated p53 acetyla-tion by forming a ternary complex with p53 andp300 [117], significantly repressing activation ofp53 transcriptional activity without apparentlyaffecting the level of p53 [118]. Acetylation ofp53 inhibits its ubiquitination by HDM2 [119].

    Figure 4. P53 protein can be positively and negatively regulated by a large number of cellular proteins

    308308 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • P53 transactivates HDM2 transcription afterbeing activated, by recruiting a TRRAP/acetyl-transferase complex to the HDM2 gene to activateits transcription [120] (Figure 1). Interestingly,HDM2 seems to be required for rapid turnoverof p53 [121]. Because HDM2 is transcriptionallyactivated by p53, this degradative pathway maycontribute to the maintenance of low p53 concen-trations in normal cells [111], thus providing anauto-regulatory feedback loop [122].

    Inhibiting the p53-HDM2 interaction withsynthetic molecules is a promising approach foractivating p53 and could lead to p53-mediatedcell-cycle arrest or apoptosis (for review see, [123]).

    Other cellular proteins are also able to inhibitp53 activity such as HDMX [124,125], phosphati-dylinositol 3-OH-kinase (Akt) [126].

    HUMAN TUMOR SUPPRESSOR P53AND DNA VIRUSESSome viruses are associated with human cancers(Table 1), but their real roles in the developmentof cell transformation are still unclear. Forinstance, studies of many viruses revealed theirability to nonspecifically induce cytogeneticdamage to their host cell chromosomes. For exam-ple, the oncogenic adenoviruses, HSV and HCMVhave been found to cause non-random, site-speci-fic chromosomal damage (for review, see [127]). Inthese different cases, if following a break in thechromosome, there is some DNA repair, then thecell may survive, but may harbor specific muta-tions, which could favor cancer development,aggravated by damaged tumor suppressors.

    DNA tumor viruses typically infect resting cellsand induce transition from G0 or G1 into the S-phase of the cell cycle. This phase is permissivefor virus replication, presumably because hostnucleotide pools and proteins involved in DNAsynthesis are necessary for viral replication. Toenter S-phase, cells must traverse the p53-pRbcheckpoint, and induce E2F activity. However,forced entry into S-phase induced by E2F activityprovokes apoptosis, and so these viruses also needto suppress apoptosis to replicate efficiently. AllDNA viruses implicated so far in cell transforma-tion have been shown to encode viral proteins cap-able of inducing inactivation of tumor suppressorprotein p53 functions (Table 1). A similar viralmechanism, based on proteinprotein interaction,is also known to block the functions of pRb, allow-

    ing simultaneously the activation of E2F activityand inhibition of apoptosis.

    Inactivation of p53 is critical for the replicationof DNA tumor viruses and the main mechanismsemployed to achieve this are largely based on thebinding of viral proteins to p53. Binding of viralprotein in most cases reduces the transcriptionalactivation function of p53. In the following para-graphs, the different strategies used by DNAviruses to abrogate p53 functions are summarised.

    Small DNA virusesAlthough adenoviruses are not known to be asso-ciated with tumorigenesis in humans, somehuman adenovirus serotypes can directly inducetumours in rats or hamsters, and all serotypestested can transform cultured rodent cells [128].Two viral genes E1A and E1B were shown to beboth necessary and sufficient for transformation[129]. Transformation of primary rodent cells byadenovirus E1A and E1B oncogenes is a two-stepprocess, where E1A-dependent induction of proli-feration is coupled to E1B-dependent suppressionof apoptosis.

    The adenovirus E1B gene encodes two proteins,the E1B55K protein and the E1B19K. E1B55K pro-tein forms a stable complex with p53 in vitro andin vivo [130,131] and inhibits p53-mediated tran-scription activation [132] by binding to the ami-no-terminal transactivation domain of p53 [131].E1B19K protein does not inhibit p53-mediatedtransactivation but alleviates p53-mediated tran-scriptional repression. P53 may induce apoptosisnot through the activation of transcription butrather through transcriptional repression and theE1B19K protein may overcome p53-mediatedapoptosis by alleviating p53-mediated transcrip-tional repression [133].

    Inhibition of p53 can also depend upon indirectand sometimes more subtle interactions. E1B55Kprotein physically interacts with the P/CAF-p53complex suggesting that the E1B55K protein inhi-bits P/CAF acetylase function on p53 by prevent-ing enzymesubstrate interaction [134]. Withoutacetylation, p53 remains in an inactivated stateand fails to function.

    Adenovirus E4orf6 is also able to interact withp53 within amino acids 318360 of the carboxy-terminal domain of p53 but, unlike E1B55K,E4orf6 blocks p53-stimulated transcription byinterfering with the ability of the amino-terminal

    Human tumor suppressor p53Human tumor suppressor p53 309309

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • domain of p53 to contact TAFII31. E4orf6 proteininteracts at a site on p53 distinct from the domainthat binds to TAFII31 but nevertheless inhibitsp53TAFII31 interaction [135].

    Inhibition of the transcriptional activity of p53 isnot the only consequence of viral proteins bindingto p53; such interaction can also provoke thedegradation of p53. E1B55K and E4orf6 proteinsboth bind independently to p53 and together pro-mote its rapid degradation [136] via a novelmechanism involving a cullin-containing E3 ubi-quitin ligase [137].

    Even though E1A is considered responsible forp53 protein stabilisation [138], E1A protein couldalso repress p53-mediated transactivation, withoutprovoking a significant effect on the DNA-bindingcapacity of p53. Apparently, E1A causes increasedhomo- or hetero-oligomerisation of p53, whichmight result in inactivation of the transcriptionactivation domain of p53, blocking its tetramerisa-tion. Additionally, transfectant cells stably expres-sing E1A lose their ability to arrest the cell cycle atthe G1 stage following DNA damage, suggestingthat E1A can abolish the normal biological functionof p53 [139]. Moreover, p300/CBP and TBP,required for p53 activity, have been identified asfunctional targets of E1A in transcriptional regula-tion, and may therefore provide a pathway bywhich E1A can exert its transforming activity [140].

    The SV40 is a macaque polyomavirus. It has notbeen associated with tumors in its natural host butit is known to induce tumors in new-born rodents.It was iatrogenically introduced into millions ofpeople in North America and Europe between1955 and 1962 through SV40-contaminated poliovaccines [141]. SV40 has also been associatedwith a number of human cancers and there is con-troversy as to whether the virus contributes totumorigenesis. The viral sequence encoding thelarge T antigen (Tag) is always expressed in trans-formed cells, and Tag is sufficient to transformrodent stem cells. Tag is a multifunctional DNA-binding protein that initiates viral DNA replica-tion. Tag is both a transcriptional repressor andactivator, composed of three independent domainsthat contribute to transformation. The first trans-forming region is a J domain (present in membersof DNA J-family of chaperones). The second onegoverns interaction with the pRb and relatives.The third transforming region allows interactionwith p53. All three regions are required for trans-

    formation in most cell types [142144]. SV40 Tagbinds to p53 and inhibits its capacity to activatetranscription. The viral protein could prevent p53from binding to the ribosomal gene cluster (RGC)site [145] on cellular DNA, which contains p53recognition sequences including p53-binding frag-ments, and furthermore wild-type p53 could sti-mulate transcription of reporter genes with RGCin their 50 region [146]. Tag, under certain condi-tions, may also repress p53-dependent transcrip-tion by a mechanism in which the transactivationdomain of p53 is abrogated while DNA binding isunaffected [147]. Simultaneously, p53 has beenshown to stabilise Tag, thus augmenting its onco-genic potential [148]. Tag can also repress p53-mediated transcription by blocking the interactionof p53 with transcription coactivators such asp300. CBP and p300 have been shown to interactwith SV40 large T antigen inducing an alterationof the phosphorylation states of both proteinsand consequently inhibiting their transcriptionalactivities [149]. P300 binds to a carboxyl-terminalfragment of large T antigen which contains itsp53 binding domain, but the implications for p53functions are not yet established [150]. SV40 largeT antigen also represses p53-mediating transcrip-tion by deleteriously affecting the assembly or con-formation of the basal transcription machineryrequired for p53 functions. Large T antigen couldalso interact with both TBP and several TAFs [151].

    Some studies have reported the presence of Tag,and of Tag/p53 complexes, in SV40-associatedhuman tumors [152,153]. More work needs to bedone to accept the hypothesis that SV40 causescancers in humans (for review, see [154]).

    HPVs are small DNA viruses clearly associatedwith the induction of cancer. Clinical and epide-miological studies have implicated HPV infectionin the development of cervical carcinomas. Thehigh-risk HPVs (most notably HPV16 and 18) areassociated with high-grade squamous intraepithe-lial lesions and invasive cervical carcinomas,whereas the low-risk types are found more oftenin low-grade lesions [155]. In the majority of pri-mary cervical carcinoma and cervical tumor celllines studied, the viral genomes of high-risk HPV(16 and 18) types are integrated into the host-genome, allowing active expression of E6 and E7genes. The transforming activities of E6 werefirst revealed when it was shown that efficientimmortalisation of primary human fibroblasts or

    310310 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • keratinocytes requires the combination of E6 andE7 [156]. E6 protein binds to p53 and inhibits tran-scriptional activation, inducing changes within thep53 protein which result both in inhibition of DNAbinding and in dissociation of p53 protein pre-viously bound to the DNA [157].

    It has become increasingly clear that many smallDNA tumor viruses have evolved complexmechanisms to functionally inactivate the p53.But the mechanism employed by HPV E6 proteinsto inactivate p53 functions is different; eventhough Tag, E1B55K and E6 proteins can eachform complexes with p53 that lead to functionalinactivation, these interactions have differentimpacts on the stability of the p53 [158]. SV40Tag and E1B55K are able to provoke an extendedmetabolic half-life of p53 [159,160], caused bysequestration of p53 into stable but nonfunctionalcomplexes. The high-risk HPVs have developed aunique strategy to inactivate the p53 by accelerat-ing its proteolytic degradation. MechanisticallyHPV-16 E6 accelerates p53 degradation by forminga complex with, and reprogramming, the cellularubiquitin-protein ligase E6AP [161,162]. TheE6/E6AP complex acts as a p53-specific ubiqui-tin-protein ligase [163] (for review, see [164]).

    The ability of E6 to induce p53 degradation viaubiquitination would be independent of the sixterminal lysine residues in p53 which had pre-viously been identified as playing an importantrole in effective ubiquitination and degradationof p53 mediated by HDM2. E6 could induce ubi-quitination of p53 using an alternative mechanismto that employed by HDM2 in promoting p53degradation [165]. However, an earlier studyshowed that HPV18 E6 mediated-inhibition oftranscriptional activity of p53 could be indepen-dent of E6 induced degradation [157].

    E6 protein of HPV16 also binds to three regionsof both CBP and p300, resulting in abrogation ofCBP/p300 transcriptional co-activator functions.E6 may also suppress p53 activation via theCBP/p300 complex quite independently from itsknown ability of binding to p53 and removing cel-lular p53 via the proteasomal degradation path-way [166,167].

    HBV produces chronic infections of the liverleading to cirrhosis and hepatocellular carcinoma.The X protein (HBx) of HBV, which is able totransactivate viral and cellular genes by interact-ing directly with cellular transcription factors,

    might be involved in the development of the hepa-tocarcinogenesis. It is capable of neoplasticallytransforming rodent cells [168], and causing HCCin transgenic mice [169]. HBx is a multifunctionalprotein that can interact with p53. However, theexact molecular events involved in liver carcino-genesis remain to be elucidated. HBx binds tothe C-terminus of p53 and partially sequesters itin the cytoplasm, preventing interaction betweenp53 and cellular transcription factors, thereby inhi-biting p53 action on transcriptional machinery[170,171]. A wide range of p53 functions may con-sequently be affected including the down-regula-tion of p53-induced DNA repair mechanismswhich would lead to accumulation of geneticdefects and increasing carcinogenesis [172].

    Moreover HBV core protein (HBc) functions as arepressor of human p53 gene expression by inhi-biting the promoter activity in the human livercell line HepG2, and also acting synergisticallywith HBx protein in this repressive effect [173].

    Recently, transgenic mouse models of HBV-associated hepatocellular carcinoma were used inorder to understand better the role of HBV in liverdisease and to highlight the importance of viralgene products in the cell cycle control checkpointderegulations in the hepatocytes [174].

    Large DNA virusesHerpesviruses are widespread within the world-wide population. Some of them are involved incell transformation in vitro and could also be impli-cated in human cancer development. They seem todevelop similar strategy consisting in p53 functioninactivation but some of them are also capable ofdeveloping many other mechanisms.

    HCMV is an important human pathogen thatcauses congenital birth defects and severe oppor-tunistic diseases. It has also been suspected tohave oncogenic properties. Several regions ofHCMV genome have transforming ability, includ-ing some of the IE genes. The etiological associa-tion between HCMV and human cancer remainsan enigma. The ubiquitous nature of HCMV with-in the population would lead one to expect muchhigher incident rates for each of the malignanciessupposedly linked to the virus. In light of this, ithas been suggested that HCMV may act as a co-etiologic agent in the development of tumorsunder conditions not well understood at this point.HCMV may indirectly contribute to oncogenesis

    Human tumor suppressor p53Human tumor suppressor p53 311311

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • through a hit-and-run mechanism [175]. Essen-tially it would act as a tumor promoter, perhapsby causing genetic instability or by preventingcells from undergoing apoptosis (for review, see[176]). HCMV IE-86 protein can reduce p53 tran-scriptional capacity by direct interaction [177],though without affecting G1 checkpoint function[178]. The oncogenic potential of HCMV was ori-ginally established on the basis of in vitro studieson hamster fibroblasts [179]; recently, HCMVnucleic acids and proteins were found specificallylocalised to neoplastic cells in human colorectalpolyps and adenocarcinomas. Virus infectioncould induce important oncogenic pathways incolon-cancer cells [180].

    HHV-6, a T-lymphotropic virus highly endemicin humans, is the etiologic agent of the commonfebrile childhood exanthem subitum. It has alsobeen suggested to play a role in the developmentof lymphoproliferative disorders and especially inHodgkins disease [181]. The DR7 viral gene is ableto transform NIH3T3 cells and is responsible forthe development of tumours in nude mice. Theprotein DR7 encoded by this gene is also able tobind to p53 and to inhibit the p53-mediated tran-scriptional activation [182].

    EBV is a -herpesvirus, which is the causativeagent of infectious mononucleosis, and whichestablishes a life-long persistent infection in over95% of the human adult population world-wide.It seems to be the most highly transforming mem-ber of the herpesvirus family, efficiently immorta-lising human B-lymphocytes in vitro. It isassociated with the development of several malig-nancies, including Burkitts lymphoma, nasophar-yngeal carcinoma, Hodgkins disease and T-celllymphomas. EBV BZLF1 protein, which initiatesthe switch of latent to lytic EBV infection, interactsdirectly in vitro and in vivo with the carboxy-term-inal portion of p53, and it is possible this may bepart of the latent to lytic viral infection pathway[183]. BZLF1 affects p53 post-translational modifi-cation and is known to inhibit p53 transcriptionalfunction leading some researchers to propose theexistence of an indirect mechanism involving sup-pression of TBP expression [184]. But this virus iscapable of developing a wide range of differentmechanisms in addition to p53 inactivation toinduce cell transformation. For example, the latentmembrane protein 1 (LMP1) plays a central part inthe transforming process by mimicking members

    of the family of tumor necrosis factor (TNF) recep-tors, transmitting growth signals [185].

    HHV-8 is also a human -herpesvirus that playsa role in the development of Kaposis sarcomalesions, primary effusion lymphoma and multi-centric Castlemans disease, which is an atypicallymphoproliferative disorder [186]. Several HHV-8 proteins are known to be oncogenic such asLANA and v-IRF-1.

    V-IRF-1 protein from HHV-8, which acts as aregulator for the expression of several HHV-8genes, interacts and inhibits the transcriptionalco-activator p300 [187], thus provoking indirectinhibition of p53 functions.

    Moreover, K8, an early protein encoded by HHV8also binds to p53 and functions in the recruitment ofp53 to the promyelocytic leukemia protein (PML)bodies which are nuclear sites for both input viralgenome deposition and immediate-early (IE) genetranscription during infection [188].

    Finally, ORF50 protein which is a viral transcrip-tional activator of several early and late genes ofHHV8, also represses the transcriptional activityof p53 and the p53-induced apoptosis [189].

    CONCLUSIONAs seen in this review, the p53 tumor suppressororchestrates a number of important genesinvolved in cell-cycle control and apoptosis. Itseems to be one of the most important moleculesin regulating cell growth and in protecting cellsfrom malignant transformation. When p53 func-tion is impaired, its ability to preserve genomicintegrity is compromised and this may result inan increase in mutations on the p53 gene and con-tribute to progression towards a malignant pheno-type. However, the p53 signaling pathway alsoconnects with tumor suppressors and viral onco-genes known to influence cell cycle machinery.Alterations in components either upstream ordownstream of p53 may therefore be analogousto inactivation of p53 itself, leading to deregulationof cell cycle control, genomic instability and devel-opment of cancer.

    The DNA viruses implicated in the develop-ment of human cancers exhibit a similar mechan-ism aimed at inhibiting both p53 and pRb. The factthat transforming viral proteins form complexeswith two major tumor suppressor proteins sug-gests a remarkable case of convergent evolution.Induction of apoptosis by p53 in response to viral

    312312 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • infection may explain why DNA tumor viruseshave evolved such similar approaches to disablep53 functions [190].

    Researches conducted on interactions betweenp53 and viral oncoproteins provided significantinformation on human cancer development. Thepursuit of the studies on the viral oncogenesshould allow the discovery of new factors impli-cated in the control of the cell cycle, and couldbe helpful in the design of new anti-tumoral drugs.For example, Cassetti et al. have developed Vene-zuelan equine encephalitis (VEE) virus repliconparticle (VRP). The fusion proteins were mutatedat four amino acid positions to inactivate theironcogenic potential, in order to stimulate theimmune system against E6 and E7 [191].

    The question of why viruses need to inactivatep53 has not been formally resolved, although sup-pression of G1/S checkpoint and apoptosis appearlikely to be critical. Small DNA viruses depend onhost cell machinery to replicate their own DNAand, clearly, a protein that blocks S-phase and cap-able of inducing apoptosis could inhibit virusreplication.

    Nevertheless, even if p53 is mutated in morethan 50% of primary human tumors and inacti-vated by interaction with viral oncogenic proteinsor other molecules, this alone is insufficient forthe development of cancer, despite being of greatimportance in the multistep carcinogenic process.Where necessary, other members of the p53 family,p63 and p73, can provide a second control pathwayfor the protection and repair of stressed cells.

    REFERENCES1. Lane DP. Cancer. p53, guardian of the genome.

    Nature 1992; 358: 1516.2. Hollstein M, Sidransky D, Vogelstein B, Harris CC.

    p53 mutations in human cancers. Science 1991; 253:4953.

    3. Lamb P, Crawford L. Characterization of the humanp53 gene. Mol Cell Biol 1986; 6: 13791385.

    4. Isobe M, Emanuel BS, Givol D, Oren M, Croce CM.Localization of gene for human p53 tumour antigento band 17p13. Nature 1986; 320: 8485.

    5. Yang A, Kaghad M, Wang Y, et al. p63, a p53 homo-log at 3q2729, encodes multiple products withtransactivating, death-inducing, and dominant-negative activities. Mol Cell 1998; 2: 305316.

    6. Schmale H, Bamberger C. A novel protein withstrong homology to the tumor suppressor p53.Oncogene 1997; 15: 13631367.

    7. Kaghad M, Bonnet H, Yang A, et al. Monoallelicallyexpressed gene related to p53 at 1p36, a region fre-quently deleted in neuroblastoma and other humancancers. Cell 1997; 90: 809819.

    8. Jayaraman L, Prives C. Covalent and noncovalentmodifiers of the p53 protein. Cell Mol Life Sci 1999;55: 7687.

    9. Somasundaram K. Tumor suppressor p53: regula-tion and function. Front Biosci 2000; 5: D424D437.

    10. Cho Y, Gorina S, Jeffrey PD, Pavletich NP. Crystalstructure of a p53 tumor suppressor-DNA complex:understanding tumorigenic mutations. Science 1994;265: 346355.

    11. Rippin TM, Freund SM, Veprintsev DB, Fersht AR.Recognition of DNA by p53 core domain and loca-tion of intermolecular contacts of cooperative bind-ing. J Mol Biol 2002; 319: 351358.

    12. Bell S, Klein C, Muller L, Hansen S, Buchner J. p53contains large unstructured regions in its nativestate. J Mol Biol 2002; 322: 917927.

    13. Farmer G, Bargonetti J, Zhu H, Friedman P, PrywesR, Prives C. Wild-type p53 activates transcriptionin vitro. Nature 1992; 358: 8386.

    14. Inga A, Storici F, Darden TA, Resnick MA. Differen-tial transactivation by the p53 transcription factor ishighly dependent on p53 level and promoter targetsequence. Mol Cell Biol 2002; 22: 86128625.

    15. Appella E, Anderson CW. Signaling to p53: break-ing the posttranslational modification code. PatholBiol (Paris) 2000; 48: 227245.

    16. Kar S, Sakaguchi K, Shimohigashi Y, et al. Effect ofphosphorylation on the structure and fold of trans-activation domain of p53. J Biol Chem 2002; 277:1557915585.

    17. Sakaguchi K, Herrera JE, Saito S, et al. DNA damageactivates p53 through a phosphorylation-acetyla-tion cascade. Genes Dev 1998; 12: 28312841.

    18. Gu W, Roeder RG. Activation of p53 sequence-spe-cific DNA binding by acetylation of the p53 C-term-inal domain. Cell 1997; 90: 595606.

    19. Giannakakou P, Sackett DL, Ward Y, Webster KR,Blagosklonny MV, Fojo T. p53 is associated with cel-lular microtubules and is transported to the nucleusby dynein. Nat Cell Biol 2000; 2: 709717.

    20. Chene P. The role of tetramerization in p53 func-tion. Oncogene 2001; 20: 26112617.

    21. Nicholls CD, McLure KG, Shields MA, Lee PW. Bio-genesis of p53 involves cotranslational dimerizationof monomers and posttranslational dimerization ofdimers. Implications on the dominant negativeeffect. J Biol Chem 2002; 277: 1293712945.

    22. Tokino T, Thiagalingam S, El-Deiry WS, WaldmanT, Kinzler KW, Vogelstein B. p53 tagged sites fromhuman genomic DNA. Hum Mol Genet 1994; 3:15371542.

    Human tumor suppressor p53Human tumor suppressor p53 313313

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • 23. Lu H, Levine AJ. Human TAFII31 protein is a tran-scriptional coactivator of the p53 protein. Proc NatlAcad Sci USA 1995; 92: 51545158.

    24. Thut CJ, Chen JL, Klemm R, Tjian R. p53 transcrip-tional activation mediated by coactivators TAFII40and TAFII60. Science 1995; 267: 100104.

    25. Burley SK, Roeder RG. Biochemistry and structuralbiology of transcription factor IID (TFIID). Annu RevBiochem 1996; 65: 769799.

    26. Ko LJ, Prives C. p53: puzzle and paradigm. GenesDev 1996; 10: 10541072.

    27. Farmer G, Friedlander P, Colgan J, Manley JL,Prives C. Transcriptional repression by p53involves molecular interactions distinct from thosewith the TATA box binding protein. Nucleic AcidsRes 1996; 24: 42814288.

    28. Mack DH, Vartikar J, Pipas JM, Laimins LA. Specificrepression of TATA-mediated but not initiator-mediated transcription by wild-type p53. Nature1993; 363: 281283.

    29. Santhanam U, Ray A, Sehgal PB. Repression of theinterleukin 6 gene promoter by p53 and the retino-blastoma susceptibility gene product. Proc Natl AcadSci USA 1991; 88: 76057609.

    30. Shiio Y, Yamamoto T, Yamaguchi N. Negative reg-ulation of Rb expression by the p53 gene product.Proc Natl Acad Sci USA 1992; 89: 52065210.

    31. Budde A, Grummt I. p53 represses ribosomal genetranscription. Oncogene 1999; 18: 11191124.

    32. Ljungman M, Zhang F, Chen F, Rainbow AJ, McKayBC. Inhibition of RNA polymerase II as a trigger forthe p53 response. Oncogene 1999; 18: 583592.

    33. Chesnokov I, Chu WM, Botchan MR, Schmid CW.p53 inhibits RNA polymerase III-directed transcrip-tion in a promoter-dependent manner. Mol Cell Biol1996; 16: 70847088.

    34. Cairns CA, White RJ. p53 is a general repressor ofRNA polymerase III transcription. EMBO J 1998;17: 31123123.

    35. Stein T, Crighton D, Warnock LJ, Milner J, White RJ.Several regions of p53 are involved in repression ofRNA polymerase III transcription. Oncogene 2002;21: 55405547.

    36. Horton LE, Bushell M, Barth-Baus D, Tilleray VJ,Clemens MJ, Hensold JO. p53 activation results inrapid dephosphorylation of the eIF4E-binding pro-tein 4E-BP1, inhibition of ribosomal protein S6kinase and inhibition of translation initiation. Onco-gene 2002; 21: 53255334.

    37. Bagchi S, Weinmann R, Raychaudhuri P. The reti-noblastoma protein copurifies with E2F-I, an E1A-regulated inhibitor of the transcription factor E2F.Cell 1991; 65: 10631072.

    38. Qin XQ, Chittenden T, Livingston DM, Kaelin WG,Jr. Identification of a growth suppression domain

    within the retinoblastoma gene product. Genes Dev1992; 6: 953964.

    39. Muller H, Bracken AP, Vernell R, et al. E2Fs regulatethe expression of genes involved in differentiation,development, proliferation, and apoptosis. GenesDev 2001; 15: 267285.

    40. Harper JW, Elledge SJ, Keyomarsi K, et al. Inhibitionof cyclin-dependent kinases by p21. Mol Biol Cell1995; 6: 387400.

    41. Chen J, Jackson PK, Kirschner MW, Dutta A. Sepa-rate domains of p21 involved in the inhibition ofCdk kinase and PCNA. Nature 1995; 374: 386388.

    42. Gulbis JM, Kelman Z, Hurwitz J, ODonnell M,Kuriyan J. Structure of the C-terminal region ofp21(WAF1/CIP1) complexed with human PCNA.Cell 1996; 87: 297306.

    43. Flores-Rozas H, Kelman Z, Dean FB, et al.Cdk-interacting protein 1 directly binds with prolif-erating cell nuclear antigen and inhibits DNA repli-cation catalyzed by the DNA polymerase deltaholoenzyme. Proc Natl Acad Sci USA 1994; 91:86558659.

    44. Waga S, Hannon GJ, Beach D, Stillman B. The p21inhibitor of cyclin-dependent kinases controlsDNA replication by interaction with PCNA. Nature1994; 369: 574578.

    45. Li R, Waga S, Hannon GJ, Beach D, Stillman B. Dif-ferential effects by the p21 CDK inhibitor on PCNA-dependent DNA replication and repair. Nature1994; 371: 534537.

    46. Shivji MK, Grey SJ, Strausfeld UP, Wood RD, BlowJJ. Cip1 inhibits DNA replication but not PCNA-dependent nucleotide excision-repair. Curr Biol1994; 4: 10621068.

    47. Li R, Hannon GJ, Beach D, Stillman B. Subcellulardistribution of p21 and PCNA in normal andrepair-deficient cells following DNA damage. CurrBiol 1996; 6: 189199.

    48. Vaziri C, Saxena S, Jeon Y, et al. A p53-dependentcheckpoint pathway prevents rereplication. Mol Cell2003; 11: 9971008.

    49. Stewart N, Hicks GG, Paraskevas F, Mowat M. Evi-dence for a second cell cycle block at G2/M by p53.Oncogene 1995; 10: 109115.

    50. Azzam EI, de Toledo SM, Pykett MJ, Nagasawa H,Little JB. CDC2 is down-regulated by ionizingradiation in a p53-dependent manner. Cell GrowthDiffer 1997; 8: 11611169.

    51. Nakamura S, Gomyo Y, Roth JA, Mukhopadhyay T.C-terminus of p53 is required for G(2) arrest. Onco-gene 2002; 21: 21022107.

    52. Pines J, Hunter T. Isolation of a human cyclincDNA: evidence for cyclin mRNA and protein reg-ulation in the cell cycle and for interaction withp34cdc2. Cell 1989; 58: 833846.

    314314 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • 53. Draetta G, Beach D. Activation of cdc2 proteinkinase during mitosis in human cells: cell cycle-dependent phosphorylation and subunit rearrange-ment. Cell 1988; 54: 1726.

    54. Taylor WR, Stark GR. Regulation of the G2/M tran-sition by p53. Oncogene 2001; 20: 18031815.

    55. Zhan Q, Antinore MJ, Wang XW, et al. Associationwith Cdc2 and inhibition of Cdc2/Cyclin B1 kinaseactivity by the p53-regulated protein Gadd45. Onco-gene 1999; 18: 28922900.

    56. Bunz F, Dutriaux A, Lengauer C, et al. Requirementfor p53 and p21 to sustain G2 arrest after DNAdamage. Science 1998; 282: 14971501.

    57. Chan TA, Hermeking H, Lengauer C, Kinzler KW,Vogelstein B. 14-3-3Sigma is required to preventmitotic catastrophe after DNA damage. Nature1999; 401: 616620.

    58. Ababneh M, Gotz C, Montenarh M. Downregula-tion of the cdc2/cyclin B protein kinase activity bybinding of p53 to p34(cdc2). Biochem Biophys ResCommun 2001; 283: 507512.

    59. Yun J, Chae HD, Choy HE, et al. p53 negatively reg-ulates cdc2 transcription via the CCAAT-bindingNF-Y transcription factor. J Biol Chem 1999; 274:2967729682.

    60. Wang Q, Zambetti GP, Suttle DP. Inhibition of DNAtopoisomerase II alpha gene expression by the p53tumor suppressor. Mol Cell Biol 1997; 17: 389397.

    61. Isaacs RJ, Davies SL, Sandri MI, Redwood C, WellsNJ, Hickson ID. Physiological regulation of eukar-yotic topoisomerase II. Biochim Biophys Acta 1998;1400: 121137.

    62. Downes CS, Clarke DJ, Mullinger AM, Gimenez-Abian JF, Creighton AM, Johnson RT. A topoisome-rase II-dependent G2 cycle checkpoint in mamma-lian cells. Nature 1994; 372: 467470.

    63. Minemoto Y, Uchida S, Ohtsubo M, et al. Loss of p53induces M-phase retardation following G2 DNAdamage checkpoint abrogation. Arch Biochem Bio-phys 2003; 412: 1319.

    64. Kazantsev A, Sancar A. Does the p53 up-regulatedGadd45 protein have a role in excision repair?Science 1995; 270: 10031004; author reply 10051006.

    65. Kearsey JM, Shivji MK, Hall PA, Wood RD. Doesthe p53 up-regulated Gadd45 protein have a rolein excision repair? Science 1995; 270: 10041005;author reply 10051006.

    66. Smith ML, Chen IT, Zhan Q, et al. Interaction of thep53-regulated protein Gadd45 with proliferatingcell nuclear antigen. Science 1994; 266: 13761380.

    67. Carrier F, Georgel PT, Pourquier P, et al. Gadd45, ap53-responsive stress protein, modifies DNA acces-sibility on damaged chromatin. Mol Cell Biol 1999;19: 16731685.

    68. Wang Q, Zhu Q, Wani MA, Wani G, Chen J, WaniAA. Tumor suppressor p53 dependent recruitmentof nucleotide excision repair factors XPC and TFIIHto DNA damage. DNA Repair (Amst) 2003; 2:483499.

    69. Soe K, Hartmann H, Schlott B, Stevnsner T, GrosseF. The tumor suppressor protein p53 stimulates theformation of the human topoisomerase I doublecleavage complex in vitro. Oncogene 2002; 21:66146623.

    70. Melle C, Nasheuer HP. Physical and functionalinteractions of the tumor suppressor protein p53and DNA polymerase alpha-primase. Nucleic AcidsRes 2002; 30: 14931499.

    71. Rubbi CP, Milner J. p53 is a chromatin accessibilityfactor for nucleotide excision repair of DNAdamage. EMBO J 2003; 22: 975986.

    72. Mummenbrauer T, Janus F, Muller B, Wiesmuller L,Deppert W, Grosse F. p53 Protein exhibits 30-to-50

    exonuclease activity. Cell 1996; 85: 10891099.73. Janus F, Albrechtsen N, Knippschild U, Wiesmuller

    L, Grosse F, Deppert W. Different regulation of thep53 core domain activities 30-to-50 exonuclease andsequence-specific DNA binding. Mol Cell Biol 1999;19: 21552168.

    74. Ballal K, Zhang W, Mukhopadyay T, Huang P. Sup-pression of mismatched mutation by p53: amechanism for guarding genomic integrity. J MolMed 2002; 80: 2532.

    75. Lilling G, Elena N, Sidi Y, Bakhanashvili M. p53-associated 30> 50 exonuclease activity in nuclearand cytoplasmic compartments of cells. Oncogene2003; 22: 233245.

    76. Zhu J, Zhang S, Jiang J, Chen X. Definition of the p53functional domains necessary for inducing apopto-sis. J Biol Chem 2000; 275: 3992739934.

    77. Miyashita T, Reed JC. Tumor suppressor p53 is adirect transcriptional activator of the human Baxgene. Cell 1995; 80: 293299.

    78. Owen-Schaub LB, Zhang W, Cusack JC, et al. Wild-type human p53 and a temperature-sensitivemutant induce Fas/APO-1 expression. Mol Cell Biol1995; 15: 30323040.

    79. Schuler M, Bossy-Wetzel E, Goldstein JC, FitzgeraldP, Green DR. p53 induces apoptosis by caspase acti-vation through mitochondrial cytochrome c release.J Biol Chem 2000; 275: 73377342.

    80. Schuler M, Maurer U, Goldstein JC, et al. p53 trig-gers apoptosis in oncogene-expressing fibroblastsby the induction of Noxa and mitochondrial Baxtranslocation. Cell Death Differ 2003; 10: 451460.

    81. Gao CF, Ren S, Zhang L, et al. Caspase-dependentcytosolic release of cytochrome c and membranetranslocation of Bax in p53-induced apoptosis. ExpCell Res 2001; 265: 145151.

    Human tumor suppressor p53Human tumor suppressor p53 315315

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • 82. Goping IS, Gross A, Lavoie JN, et al. Regulated tar-geting of BAX to mitochondria. J Cell Biol 1998; 143:207215.

    83. Eskes R, Antonsson B, Osen-Sand A, et al. Bax-induced cytochrome C release from mitochondriais independent of the permeability transition porebut highly dependent on Mg2 ions. J Cell Biol1998; 143: 217224.

    84. Nakano K, Vousden KH. PUMA, a novel proapopto-tic gene, is induced by p53. Mol Cell 2001; 7: 683694.

    85. Gupta S, Radha V, Furukawa Y, Swarup G. Directtranscriptional activation of human caspase-1 bytumor suppressor p53. J Biol Chem 2001; 276:1058510588.

    86. Gupta S, Radha V, Sudhakar C, Swarup G. Anuclear protein tyrosine phosphatase activates p53and induces caspase-1-dependent apoptosis. FEBSLett 2002; 532: 6166.

    87. MacLachlan TK, El-Deiry WS. Apoptotic thresholdis lowered by p53 transactivation of caspase-6. ProcNatl Acad Sci USA 2002; 99: 94929497.

    88. Mihara M, Erster S, Zaika A, et al. p53 has a directapoptogenic role at the mitochondria. Mol Cell2003; 11: 577590.

    89. Reed JC. Bcl-2 and the regulation of programmedcell death. J Cell Biol 1994; 124: 16.

    90. Oltvai ZN, Milliman CL, Korsmeyer SJ. Bcl-2 het-erodimerizes in vivo with a conserved homolog,Bax, that accelerates programmed cell death. Cell1993; 74: 609619.

    91. Heinrichs S, Deppert W. Apoptosis or growtharrest: modulation of the cellular response to p53by proliferative signals. Oncogene 2003; 22: 555571.

    92. Lu W, Pochampally R, Chen L, Traidej M, Wang Y,Chen J. Nuclear exclusion of p53 in a subset oftumors requires MDM2 function. Oncogene 2000;19: 232240.

    93. Sionov RV, Coen S, Goldberg Z, et al. c-Abl regu-lates p53 levels under normal and stress conditionsby preventing its nuclear export and ubiquitination.Mol Cell Biol 2001; 21: 58695878.

    94. Goldberg Z, Vogt Sionov R, Berger M, et al. Tyrosinephosphorylation of Mdm2 by c-Abl: implicationsfor p53 regulation. EMBO J 2002; 21: 37153727.

    95. Nie Y, Li HH, Bula CM, Liu X. Stimulation of p53DNA binding by c-Abl requires the p53 C terminusand tetramerization. Mol Cell Biol 2000; 20: 741748.

    96. Hasan MK, Yaguchi T, Sugihara T, et al. CARF is anovel protein that cooperates with mouse p19ARF(human p14ARF) in activating p53. J Biol Chem2002; 277: 3776537770.

    97. Zhang Y, Xiong Y, Yarbrough WG. ARF promotesMDM2 degradation and stabilizes p53: ARF-INK4alocus deletion impairs both the Rb and p53 tumorsuppression pathways. Cell 1998; 92: 725734.

    98. Chen D, Li M, Luo J, Gu W. Direct interactionsbetween HIF-1 alpha and Mdm2 modulate p53function. J Biol Chem 2003; 278: 1359513598.

    99. Nagashima M, Shiseki M, Pedeux RM, et al. A novelPHD-finger motif protein, p47ING3, modulatesp53-mediated transcription, cell cycle control, andapoptosis. Oncogene 2003; 22: 343350.

    100. Leung KM, Po LS, Tsang FC, et al. The candidatetumor suppressor ING1b can stabilize p53 by dis-rupting the regulation of p53 by MDM2. CancerRes 2002; 62: 48904893.

    101. Scully R, Chen J, Ochs RL, et al. Dynamic changes ofBRCA1 subnuclear location and phosphorylationstate are initiated by DNA damage. Cell 1997; 90:425435.

    102. Somasundaram K, MacLachlan TK, Burns TF, et al.BRCA1 signals ARF-dependent stabilization andcoactivation of p53. Oncogene 1999; 18: 66056614.

    103. Ouchi T, Monteiro AN, August A, Aaronson SA,Hanafusa H. BRCA1 regulates p53-dependent geneexpression. Proc Natl Acad Sci USA 1998; 95: 23022306.

    104. MacLachlan TK, Takimoto R, El-Deiry WS. BRCA1directs a selective p53-dependent transcriptionalresponse towards growth arrest and DNA repairtargets. Mol Cell Biol 2002; 22: 42804292.

    105. Wesierska-Gadek J, Schmid G. Poly(ADP-ribose)polymerase-1 regulates the stability of the wild-typep53 protein. Cell Mol Biol Lett 2001; 6: 117140.

    106. Wieler S, Gagne JP, Vaziri H, Poirier GG, BenchimolS. Poly(ADP-ribose) polymerase-1 is a positive reg-ulator of the p53-mediated G1 arrest response fol-lowing ionizing radiation. J Biol Chem 2003; 278:1891418921.

    107. Shimizu H, Burch LR, Smith AJ, et al. The conforma-tionally flexible S9-S10 linker region in the coredomain of p53 contains a novel MDM2 binding sitewhose mutation increases ubiquitination of p53in vivo. J Biol Chem 2002; 277: 2844628458.

    108. Lin J, Chen J, Elenbaas B, Levine AJ. Several hydro-phobic amino acids in the p53 amino-terminaldomain are required for transcriptional activation,binding to mdm-2 and the adenovirus 5 E1B 55-kD protein. Genes Dev 1994; 8: 12351246.

    109. Haupt Y, Maya R, Kazaz A, Oren M. Mdm2 pro-motes the rapid degradation of p53. Nature 1997;387: 296299.

    110. Honda R, Tanaka H, Yasuda H. OncoproteinMDM2 is a ubiquitin ligase E3 for tumor suppressorp53. FEBS Lett 1997; 420: 2527.

    111. Kubbutat MH, Jones SN, Vousden KH. Regulationof p53 stability by Mdm2. Nature 1997; 387: 299303.

    112. Asher G, Lotem J, Kama R, Sachs L, Shaul Y. NQO1stabilizes p53 through a distinct pathway. Proc NatlAcad Sci USA 2002; 99: 30993104.

    316316 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • 113. Asher G, Lotem J, Sachs L, Kahana C, Shaul Y.Mdm-2 and ubiquitin-independent p53 proteaso-mal degradation regulated by NQO1. Proc Natl AcadSci USA 2002; 99: 1312513130.

    114. Freedman DA, Levine AJ. Nuclear export isrequired for degradation of endogenous p53 byMDM2 and human papillomavirus E6. Mol Cell Biol1998; 18: 72887293.

    115. Zhang Y, Xiong Y. A p53 amino-terminal nuclearexport signal inhibited by DNA damage-inducedphosphorylation. Science 2001; 292: 19101915.

    116. Shirangi TR, Zaika A, Moll UM. Nuclear degrada-tion of p53 occurs during down-regulation of thep53 response after DNA damage. FASEB J 2002;16: 420422.

    117. Kobet E, Zeng X, Zhu Y, Keller D, Lu H. MDM2inhibits p300-mediated p53 acetylation and activa-tion by forming a ternary complex with the two pro-teins. Proc Natl Acad Sci USA 2000; 97: 1254712552.

    118. Jin Y, Zeng SX, Dai MS, Yang XJ, Lu H. MDM2 Inhi-bits PCAF (p300/CREB-binding protein-associatedfactor)-mediated p53 acetylation. J Biol Chem 2002;277: 3083830843.

    119. Li M, Luo J, Brooks CL, Gu W. Acetylation of p53inhibits its ubiquitination by Mdm2. J Biol Chem2002; 277: 5060750611.

    120. Ard PG, Chatterjee C, Kunjibettu S, Adside LR,Gralinski LE, McMahon SB. Transcriptional regula-tion of the mdm2 oncogene by p53 requires TRRAPacetyltransferase complexes. Mol Cell Biol 2002; 22:56505661.

    121. Grossman SR, Deato ME, Brignone C, et al. Polyubi-quitination of p53 by a ubiquitin ligase activity ofp300. Science 2003; 300: 342344.

    122. Wu X, Bayle JH, Olson D, Levine AJ. The p53-mdm-2 autoregulatory feedback loop. Genes Dev 1993; 7:11261132.

    123. Chene P. Inhibiting the p53-MDM2 interaction: animportant target for cancer therapy. Nat Rev Cancer2003; 3: 102109.

    124. Shvarts A, Steegenga WT, Riteco N, et al. MDMX: anovel p53-binding protein with some functionalproperties of MDM2. EMBO J 1996; 15: 53495357.

    125. Sabbatini P, McCormick F. MDMX inhibits thep300/CBP-mediated acetylation of p53. DNA CellBiol 2002; 21: 519525.

    126. Yamaguchi A, Tamatani M, Matsuzaki H, et al. Aktactivation protects hippocampal neurons fromapoptosis by inhibiting transcriptional activity ofp53. J Biol Chem 2001; 276: 52565264.

    127. Fortunato EA, Spector DH. Viral induction of site-specific chromosome damage. Rev Med Virol 2003;13: 2137.

    128. Mackey JK, Rigden PM, Green M. Do highly onco-genic group A human adenoviruses cause human

    cancer? Analysis of human tumors for adenovirus12 transforming DNA sequences. Proc Natl AcadSci USA 1976; 73: 46574661.

    129. Subramanian T, Malstrom SE, Chinnadurai G.Requirement of the C-terminal region of adenovirusE1a for cell transformation in cooperation with E1b.Oncogene 1991; 6: 11711173.

    130. Sarnow P, Ho YS, Williams J, Levine AJ. Adeno-virus E1b-58kd tumor antigen and SV40 largetumor antigen are physically associated with thesame 54 kd cellular protein in transformed cells. Cell1982; 28: 387394.

    131. Kao CC, Yew PR, Berk AJ. Domains required for invitro association between the cellular p53 and theadenovirus 2 E1B 55K proteins. Virology 1990; 179:806814.

    132. Yew PR, Berk AJ. Inhibition of p53 transactivationrequired for transformation by adenovirus early1B protein. Nature 1992; 357: 8285.

    133. Sabbatini P, Chiou SK, Rao L, White E. Modulationof p53-mediated transcriptional repression andapoptosis by the adenovirus E1B 19K protein. MolCell Biol 1995; 15: 10601070.

    134. Liu Y, Colosimo AL, Yang XJ, Liao D. AdenovirusE1B 55-kilodalton oncoprotein inhibits p53 acetyla-tion by PCAF. Mol Cell Biol 2000; 20: 55405553.

    135. Dobner T, Horikoshi N, Rubenwolf S, Shenk T.Blockage by adenovirus E4orf6 of transcriptionalactivation by the p53 tumor suppressor. Science1996; 272: 14701473.

    136. Steegenga WT, Riteco N, Jochemsen AG, Fallaux FJ,Bos JL. The large E1B protein together with theE4orf6 protein target p53 for active degradationin adenovirus infected cells. Oncogene 1998; 16:349357.

    137. Querido E, Blanchette P, Yan Q, et al. Degradation ofp53 by adenovirus E4orf6 and E1B55K proteinsoccurs via a novel mechanism involving a Cullin-containing complex. Genes Dev 2001; 15: 31043117.

    138. Lowe SW, Ruley HE. Stabilization of the p53 tumorsuppressor is induced by adenovirus 5 E1A andaccompanies apoptosis. Genes Dev 1993; 7: 535545.

    139. Steegenga WT, van Laar T, Riteco N, et al. Adeno-virus E1A proteins inhibit activation of transcrip-tion by p53. Mol Cell Biol 1996; 16: 21012109.

    140. Sang N, Avantaggiati ML, Giordano A. Roles ofp300, pocket proteins, and hTBP in E1A-mediatedtranscriptional regulation and inhibition of p53transactivation activity. J Cell Biochem 1997; 66:277285.

    141. Shah K, Nathanson N. Human exposure to SV40:review and comment. Am J Epidemiol 1976; 103:112.

    142. Zhu J, Rice PW, Gorsch L, Abate M, Cole CN. Trans-formation of a continuous rat embryo fibroblast cell

    Human tumor suppressor p53Human tumor suppressor p53 317317

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • line requires three separate domains of simian virus40 large T antigen. J Virol 1992; 66: 27802791.

    143. Pipas JM, Peden KW, Nathans D. Mutational analy-sis of simian virus 40 T antigen: isolation and char-acterization of mutants with deletions in theT-antigen gene. Mol Cell Biol 1983; 3: 203213.

    144. Chen J, Tobin GJ, Pipas JM, Van Dyke T. T-antigenmutant activities in vivo: roles of p53 and pRB bind-ing in tumorigenesis of the choroid plexus.Oncogene 1992; 7: 11671175.

    145. Bargonetti J, Reynisdottir I, Friedman PN, Prives C.Site-specific binding of wild-type p53 to cellularDNA is inhibited by SV40 T antigen and mutantp53. Genes Dev 1992; 6: 18861898.

    146. Chen JY, Funk WD, Wright WE, Shay JW, Minna JD.Heterogeneity of transcriptional activity of mutantp53 proteins and p53 DNA target sequences. Onco-gene 1993; 8: 21592166.

    147. Sheppard HM, Corneillie SI, Espiritu C, Gatti A, LiuX. New insights into the mechanism of inhibition ofp53 by simian virus 40 large T antigen. Mol Cell Biol1999; 19: 27462753.

    148. Herzig M, Novatchkova M, Christofori G. An unex-pected role for p53 in augmenting SV40 large T anti-gen-mediated tumorigenesis. Biol Chem 1999; 380:203211.

    149. Eckner R, Ludlow JW, Lill NL, et al. Association ofp300 and CBP with simian virus 40 large T antigen.Mol Cell Biol 1996; 16: 34543464.

    150. Lill NL, Tevethia MJ, Eckner R, Livingston DM,Modjtahedi N. p300 family members associate withthe carboxyl terminus of simian virus 40 largetumor antigen. J Virol 1997; 71: 129137.

    151. Damania B, Alwine JC. TAF-like function of SV40large T antigen. Genes Dev 1996; 10: 13691381.

    152. Zhen HN, Zhang X, Bu XY, et al. Expression of thesimian virus 40 large tumor antigen (Tag) and for-mation of Tag-p53 and Tag-pRb complexes inhuman brain tumors. Cancer 1999; 86: 21242132.

    153. Carbone M, Rizzo P, Grimley PM, et al. Simianvirus-40 large-T antigen binds p53 in humanmesotheliomas. Nat Med 1997; 3: 908912.

    154. Pipas JM, Levine AJ. Role of T antigen interactionswith p53 in tumorigenesis. Semin Cancer Biol 2001;11: 2330.

    155. zur Hausen H. Papillomavirus and p53. Nature1998; 393: 217.

    156. Watanabe S, Kanda T, Yoshiike K. Human papillo-mavirus type 16 transformation of primary humanembryonic fibroblasts requires expression of openreading frames E6 and E7. J Virol 1989; 63: 965969.

    157. Thomas M, Massimi P, Jenkins J, Banks L. HPV-18E6 mediated inhibition of p53 DNA binding activityis independent of E6 induced degradation.Oncogene 1995; 10: 261268.

    158. Werness BA, Levine AJ, Howley PM. Association ofhuman papillomavirus types 16 and 18 E6 proteinswith p53. Science 1990; 248: 7679.

    159. Oren M, Maltzman W, Levine AJ. Post-translationalregulation of the 54K cellular tumor antigen in nor-mal and transformed cells. Mol Cell Biol 1981; 1:101110.

    160. Reich NC, Oren M, Levine AJ. Two distinct mechan-isms regulate the levels of a cellular tumor antigen,p53. Mol Cell Biol 1983; 3: 21432150.

    161. Huibregtse JM, Scheffner M, Howley PM. A cellularprotein mediates association of p53 with the E6oncoprotein of human papillomavirus types 16 or18. EMBO J 1991; 10: 41294135.

    162. Huibregtse JM, Scheffner M, Howley PM. Localiza-tion of the E6-AP regions that direct human papillo-mavirus E6 binding, association with p53, andubiquitination of associated proteins. Mol Cell Biol1993; 13: 49184927.

    163. Scheffner M, Huibregtse JM, Vierstra RD, HowleyPM. The HPV-16 E6 and E6-AP complex functionsas a ubiquitin-protein ligase in the ubiquitinationof p53. Cell 1993; 75: 495505.

    164. Munger K, Howley PM. Human papillomavirusimmortalization and transformation functions.Virus Res 2002; 89: 213228.

    165. Camus S, Higgins M, Lane DP, Lain S. Differencesin the ubiquitination of p53 by Mdm2 and the HPVprotein E6. FEBS Lett 2003; 536: 220224.

    166. Patel D, Huang SM, Baglia LA, McCance DJ. The E6protein of human papillomavirus type 16 binds toand inhibits co-activation by CBP and p300.EMBO J 1999; 18: 50615072.

    167. Zimmermann H, Degenkolbe R, Bernard HU,OConnor MJ. The human papillomavirus type 16E6 oncoprotein can down-regulate p53 activity bytargeting the transcriptional coactivator CBP/p300. J Virol 1999; 73: 62096219.

    168. Shirakata Y, Kawada M, Fujiki Y, et al. The X gene ofhepatitis B virus induced growth stimulation andtumorigenic transformation of mouse NIH3T3 cells.Jpn J Cancer Res 1989; 80: 617621.

    169. Kim CM, Koike K, Saito I, Miyamura T, Jay G. HBxgene of hepatitis B virus induces liver cancer intransgenic mice. Nature 1991; 351: 317320.

    170. Takada S, Tsuchida N, Kobayashi M, Koike K. Dis-ruption of the function of tumor-suppressor genep53 by the hepatitis B virus X protein and hepatocar-cinogenesis. J Cancer Res Clin Oncol 1995; 121: 593601.

    171. Elmore LW, Hancock AR, Chang SF, et al. HepatitisB virus X protein and p53 tumor suppressor interac-tions in the modulation of apoptosis. Proc Natl AcadSci USA 1997; 94: 1470714712.

    172. Prost S, Ford JM, Taylor C, Doig J, Harrison DJ.Hepatitis B x protein inhibits p53-dependent DNA

    318318 S. Collot-TeixeiraS. Collot-Teixeira et al.et al.

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.

  • repair in primary mouse hepatocytes. J Biol Chem1998; 273: 3332733332.

    173. Kwon JA, Rho HM. Transcriptional repression ofthe human p53 gene by hepatitis B viral core protein(HBc) in human liver cells. Biol Chem 2003; 384:203212.

    174. Singh M, Kumar V. Transgenic mouse models ofhepatitis B virus-associated hepatocellular carci-noma. Rev Med Virol 2003; 13: 243253.

    175. Shen Y, Zhu H, Shenk T. Human cytomagalovirusIE1 and IE2 proteins are mutagenic and mediatehit-and-run oncogenic transformation in coopera-tion with the adenovirus E1A proteins. Proc NatlAcad Sci USA 1997; 94: 33413345.

    176. Castillo JP, Kowalik TF. Human cytomegalovirusimmediate early proteins and cell growth control.Gene 2002; 290: 1934.

    177. Speir E, Huang ES, Modali R, et al. Interaction ofhuman cytomegalovirus with p53: possible role incoronary restenosis. Scand J Infect Dis Suppl 1995;99: 7881.

    178. Bonin LR, McDougall JK. Human cytomegalovirusIE2 86-kilodalton protein binds p53 but does notabrogate G1 checkpoint function. J Virol 1997; 71:58615870.

    179. Albrecht T, Rapp F. Malignant transformation ofhamster embryo fibroblasts following exposure toultraviolet-irradiated human cytomegalovirus.Virology 1973; 55: 5361.

    180. Harkins L, Volk AL, Samanta M, et al. Specific loca-lisation of human cytomegalovirus nucleic acidsand proteins in human colorectal cancer. Lancet2002; 16: 15571563.

    181. Maeda A, Sata T, Enzan H, et al. The evidence ofhuman herpesvirus 6 infection in the lymph nodesof Hodgkins disease. Virchows Arch A Pathol AnatHistopathol 1993; 423: 7175.

    182. Kashanchi F, Araujo J, Doniger J, et al. Human her-pesvirus 6 (HHV-6) ORF-1 transactivating gene

    exhibits malignant transforming activity and itsprotein binds to p53. Oncogene 1997; 14: 359367.

    183. Zhang Q, Gutsch D, Kenney S. Functional and phy-sical interaction between p53 and BZLF1: implica-tions for Epstein-Barr virus latency. Mol Cell Biol1994; 14: 19291938.

    184. Mauser A, Saito S, Appella E, Anderson CW,Seaman WT, Kenney S. The Epstein-Barr virusimmediate-early protein BZLF1 regulates p53 func-tion through multiple mechanisms. J Virol 2002; 76:1250312512.

    185. Liebowitz D. Epstein-Barr virus and a cellular sig-naling pathway in lymphomas from immunosup-pressed patients. N Engl J Med 1998; 338: 14131421.

    186. Chang Y, Cesarman E, Pessin MS, et al. Identifica-tion of herpesvirus-like DNA sequences in AIDS-associated Kaposis sarcoma. Science 1994; 266:18651869.

    187. Burysek L, Yeow WS, Lubyova B, et al. Functionalanalysis of human herpesvirus 8-encoded viralinterferon regulatory factor 1 and its associationwith cellular interferon regulatory factors andp300. J Virol 1999; 73: 73347342.

    188. Katano H, Ogawa-Goto K, Hasegawa H, Kurata T,Sata T. Human-herpesvirus-8-encoded K8 proteincolocalizes with the promyelocytic leukemia pro-tein (PML) bodies and recruits p53 to the PMLbodies. Virology 2001; 286: 446455.

    189. Gwack Y, Hwang S, Byun H, et al. Kaposis sar-coma-associated herpesvirus open reading frame50 represses p53-induced transcriptional activityand apoptosis. J Virol 2001; 75: 62456248.

    190. White E. Regulation of p53-dependent apoptosis byE1A and E1B. Curr Top Microbiol Immunol 1995;199(3): 3458.

    191. Cassetti MC, McElhiney SP, Shahabi V, et al. Antitu-mor efficacy of Venezuelan equine encephalitisvirus replicon particles encoding mutated HPV16E6 and E7 genes. Vaccine 2004; 22: 520527.

    Human tumor suppressor p53Human tumor suppressor p53 319319

    Copyright # 2004 John Wiley & Sons, Ltd. Rev. Med. Virol. 2004; 14: 301319.