Você está na página 1de 13

Journal of Pathology J Pathol 2009; 219: 315 Published online 1 June 2009 in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/path.

2584

Invited Review

Tumour suppression by p53: the importance of apoptosis and cellular senescence


Valentina Zuckerman,1# Kamil Wolyniec,2# Ronit V Sionov,1 Sue Haupt2 and Ygal Haupt1,2 *
1 Lautenberg Centre for General and Tumour Immunology, The Hebrew University Hadassah Medical School, 2 Research Division, The Peter MacCallum Cancer Centre, St. Andrews Place, East Melbourne 3002, Victoria,

Jerusalem 91120, Israel Australia

*Correspondence to: Ygal Haupt, Lautenberg Centre for General and Tumour Immunology, The Hebrew University Hadassah Medical School, Jerusalem 91120, Israel. E-mail: ygal.haupt@petermac.org
# These authors contributed equally to this review.

Abstract
p53 is regarded as a central player in tumour suppression, as it controls programmed cell death (apoptosis) as well as cellular senescence. While apoptosis eliminates cells at high risk for oncogenic transformation, senescence acts as a barrier to tumourigenesis by imposing irreversible cell cycle arrest. p53 can act directly or indirectly at multiple levels of the tumour suppression network by invoking a myriad of mechanisms. p53 induces the extrinsic and intrinsic apoptotic pathways at multiple steps to ensure an efcient death response. This response involves transcriptional activation or repression of target genes, as well as the recently identied microRNAs, and transcription-independent functions. Importantly, p53 loss of function is required for tumour maintenance. Therefore, therapeutic strategies aimed at reactivation of p53 in tumours emerge as a promising approach for the treatment of cancer patients. Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd. Keywords: apoptosis; senescence; tumour suppression; p53; p73; cancer therapy; miRNA

No conicts of interest were declared.


Received: 15 April 2009 Revised: 15 May 2009 Accepted: 19 May 2009

Introduction
Cancer development results from the accumulation of genetic mutations, which lead to uncontrolled and unscheduled proliferation of cells that become immortalized and capable of invading other tissues. The p53 tumour suppressor is the major obstacle in this process. p53 is considered to be a key guardian of the genome. It senses DNA damage and in response induces a transient growth arrest, allowing DNA repair or, in the case of extensive damage, promoting irreversible growth arrest (senescence) or programmed cell death (apoptosis) [1,2]. This tumour-suppressive function of p53 prevents the propagation of abnormal cells at risk of becoming cancer cells. The critical role of p53 in the prevention of cancer development is demonstrated by p53 mutation in approximately 50% of human cancer cases. In the majority of the remaining cases p53 activities are compromised due to the deregulation of downstream or upstream signalling pathways [3]. In addition, LiFraumeni syndrome patients, carrying a mutant p53 allele, develop multiple tumour types at a high rate [4]. The contribution of mutant p53 gain of function to the development of cancer was demonstrated in a knockin mouse model expressing mutant p53 [58]. These mice develop a more metastatic cancer than mice lacking p53 [9]. In this review we will discuss the major mechanism by which p53 exerts its tumour suppression functions.

We will focus mainly on the apoptotic functions of p53 and the contribution of cellular senescence. We will also discuss evidence from recent in vivo studies shedding new light on the temporal restoration of tumour suppression by p53. It is also important to note that p53 has recently been implicated in autophagy and ageing as well as glycolysis [2,10]; however, due to space limitations, these will not be discussed here.

p53 and cellular senescence


Telomere shortening is a universal mechanism that limits the proliferative potential of normal cells (lacking endogenous telomerase) following extensive cell divisions. The process underlying this observation is known as replicative senescence. It is conjectured that telomere erosion beyond a certain limit triggers a DNA damage response and subsequent activation of the ATM/ATRp53 pathway, resulting in growth arrest [11,12]. Cells that fail to senesce and continue to proliferate despite dysfunctional telomeres develop chromosomal aberrations, which can result in malignant transformation [13]. The role of telomere exhaustion in the suppression of tumourigenesis in vivo, by initiating cellular senescence, has recently been demonstrated [1416]. A growing body of evidence suggests that cellular senescence is an important and evolutionarily conserved tumour-suppression mechanism, which

Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd. www.pathsoc.org.uk

V Zuckerman et al

acts as a natural barrier to cell immortalization and transformation [17,18]. The existence of non-telomere-induced senescence was anticipated by the behaviour of primary murine cells in culture. Normal mouse cells, like normal human cells, have a nite replicative potential. However, their replicative lifespan is substantially shorter than that of human cells (1015 population doublings compared to 5070), despite murine broblasts having very long telomeres (60 versus 12 kb) and in some cases expressing telomerase [19]. Thus, it is unlikely that the replicative senescence of primary mouse broblasts is due to telomere shortening. It has been proposed that the senescence phenomenon in primary mouse cells is due to a stress response to culture conditions, which can be overcome by decreasing the oxygen concentration [20]. Senescence mediated by non-telomeric signals is termed premature senescence, accelerated senescence or extrinsic senescence. Another term, stress- or aberrant signalling senescence (STASIS) has been suggested to describe the process of a senescence-like arrest mechanism in response to stress stimuli [21] (Figure 1). Abnormal activation of oncogenes, such as Ras, can promote cellular senescence in mouse and human cells. As long as the programme remains intact, the neoplastic growth process may remain benign for many years, possibly contributing to tumour dormancy [17]. A fundamental role in this fail-safe mechanism is attributed to the p53 and pRb pathways [22]. These pathways are critical for the initiation and maintenance of the senescent phenotype in human and mouse cells. Mutations in p53 or in the pRb pathway, commonly in p16INK4a , is sufcient to prevent cellular senescence, thereby

removing the obstacle from tumour progression [17]. In mouse embryo broblasts (MEFs) disruption of p53 alone is sufcient to prevent senescence. Notably, p53-null MEFs acquire an immortal phenotype and p53-null mice are highly susceptible to spontaneous tumourigenesis [9,23]. While inactivation of pRb on its own is insufcient to overcome senescence, the concomitant inactivation of its family members, p107 and p130, allows MEFs to escape senescence [24,25]. Thus, at least in MEFs, both p53 and the pRb family operate in a linear signalling pathway to induce senescence, whereby stress-activated p53 activates pRb to induce senescence. The p21WAF1 protein, an inhibitor of cyclin E/Cdk2 complexes, which is a direct transcriptional target of p53, links these two pathways. In human cells, however, inactivation of both p53 and pRb is essential to prevent the onset of replicative senescence, whereas disruption of only one of these pathways only delays the onset of senescence [26].

Stimuli and regulation of p53-induced senescence


The signals that induce a DNA damage response, such as sublethal doses of radiation, chemotherapeutic drugs (such as etoposide or cyclophosphamide) or telomere dysfunction, appear to drive senescence primarily via the p53p21 pathway (Figure 1). Disruption of DNA repair genes, such as Brca1 and DNA ligase IV, induces premature ageing in mice and premature senescence of MEFs decient in these genes. Interestingly, many of these senescent phenotypes can be rescued by p53 inactivation, indicating a pivotal role for p53 in DNA damage-induced senescence [27,28]. Moreover, cancer cells that retain

Figure 1. p53-dependent senescence is triggered by a wide spectrum of stimuli. Telomere shortening as well as non-telomeric signals, such as DNA-damaging agents, oncogenic signalling, oxidative stress, -interferon, HDAC inhibitors and depletion of heat shock proteins, induce p53-dependent senescence
J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

intact p53 are much more susceptible to senescence in response to chemotherapy [2931]. The induction of p21 is important for DNA damage-induced senescence, as well as its well-established role in transient growth arrest. The effectors determining the decision between these outcomes remain largely elusive. It has been suggested that efcient DNA repair inhibits p53p21 signalling, allowing cell cycle progression, whereas irreparable DNA lesions sustain the ATM/ATRp53p21 DNA damage response, maintaining the senescent phenotype [32]. A number of oncogenes, such as RAS [33], E2F [34], RUNX1 [35,36] and RUNX1ETO [36], trigger p53-induced senescence. While activation of oncogenes such as RAS or MOS involve a DNA damage response [18], others, such as RUNX1 or RUNX1ETO induce p53-dependent senescence independently of replicative stress and without DNA damage [36]. In addition, hydrogen peroxide-mediated oxidative stress also induces p53-dependent senescence [37]. Intriguingly, chemical inhibition of histone deacetylase, resulting in chromatin decondensation and the formation of euchromatin, was also found to cause p53-mediated senescence in MEFs [38]. Recent ndings demonstrate that specic depletion of chaperones such as members of Hsp70 (heat shock proteins) causes activation of the p53 pathway and subsequently cellular senescence. This may explain why certain cancer cells over-express Hsp70 [39].

between p53 and PML [47]. Further, PRAK (p38regulated/activated protein kinase), a downstream regulator of p38 MAPK, is essential for RAS-induced transcriptional activity of p53. PRAK phosphorylates p53 on Ser37, and p38 phosphorylates p53 on Ser33 and Ser46 during RAS-induced senescence [48].

Effectors of p53-mediated growth arrest and senescence


p53 mediates cell growth arrest by inducing the expression of cell cycle regulatory target genes [49]. The p21Waf1 protein inhibits CDK2/Cyclin E activity, while Gadd45 and 1433 inhibit Cdc2 activity (Figure 2). 1433 binds the Cdc2Cyclin B complex and Cdc25C phosphatase, and sequesters these in the cytoplasm [50,51], whereas Gadd45 inhibits Cdc2Cyclin B complex formation [52,53]. p53 is also capable of repressing the expression of cell cycle progression genes, such as CDK4 and Cyclin E2 [54], and the cell cycle phosphatases, Cdc25A and Cdc25C, [55,56] (Figure 2). In addition, a few more p53 target genes have been identied. These include dualspecicity phosphatase 11 (DUSP11 ), response gene to complement 32 (RGC32 ) or protein tyrosine phosphatase (PTPRV ) [5759]. In contrast to the induction of cell cycle arrest, the mechanisms by which p53 induces senescence are partially understood. p53 regulates plasminogen activator inhibitor-1 (PAI-1) expression by stabilizing its mRNA through direct binding to its mRNA 3 -UTR [60]. PAI-1 has been considered a well-established marker of senescent cells [61,62]. Importantly, downregulation of PAI-1 results in escape from replicative senescence in murine and human primary broblasts, through sustained activation of the PI3KAkt survival pathway and nuclear retention of Cyclin D [63]. Another p53-induced gene implicated in senescence is MIC-1, a cytokine of the TGF family. MIC-1 secretion from the cell may propagate a senescence programme by autocrine and paracrine induction [39].

Regulation of p53-induced senescence


Mechanisms responsible for activation of p53 in senescent cells are incompletely understood; however, some molecular details are emerging. One cause of p53 activation appears to be an increase in the expression of ARF (p14ARF in human; p19ARF in mouse), a tumour suppressor encoded by the INK4aARF locus. ARF stimulates p53 activity by sequestrating HDM2 (MDM2 in mice), which is an E3 ubiquitin ligase that targets p53 for proteosome-mediated degradation. Thus, ARF acts to prevent the MDM2-driven negativefeedback regulation of p53 by MDM2 [40,41]. ARF is up-regulated in cultured senescent murine cells as well as during premature senescence induced by HRASV12 in MEFs, and is required for telomeric and non-telomeric senescence in MEFs [42]. In human cells the role of ARF and p53 is more complicated. For example, ARF and p53 are critical regulators of E2F-induced senescence [34], but not RAS-induced senescence in human primary broblasts [43,44]. Another important activator of p53 is the promyelocytic leukaemia (PML) tumour suppressor. PML has been implicated in replicative senescence and in premature senescence in response to oncogenic RAS. PML interacts with CBP/p300 acetyltransferase and stabilizes p53 through acetylation [45,46]. PML has been also recently identied as a direct target of p53, hence revealing a regulatory positive feedback loop

p53 and cell death


Apoptosis is a well-studied and well-understood process that has been considered to play an important role in tumour suppression. Apoptosis is triggered in response to a variety of signals, which can activate the extrinsic and/or intrinsic death pathways; or when cells are deprived of pro-survival signals [1]. p53 acts at multiple levels of the intrinsic and extrinsic pathways [64] through the induction of multiple apoptotic target genes, as well as through transcription-independent mechanisms (Figure 3) [65,66].

Transcription-dependent apoptosis by p53


Stress-activated p53 can trigger apoptosis through the transcriptional activation of pro-apoptotic target

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

V Zuckerman et al

Figure 2. Role of p53 in the regulation of checkpoints in the eukaryotic cell cycle. The cell cycle consists of alternating S phase (DNA synthesis) and M phase (mitosis), separated by two gap phases, G1 and G2 . Cyclin D-dependent kinases (CDKs) accumulate in response to mitogenic signals and initiate the phosphorylation of pRb, a process that is completed by cyclin ECdk2. Once cells enter S phase, cyclin E is degraded and cyclin A enters into complexes with Cdk2. Cyclin BCdc2 mediates G2 M transition. INK4 proteins (eg p16INK4a, p15INK4b, p18INK4c, p19INK4d) oppose the activities of the various cyclin D-dependent kinases, whereas Cip/Kip proteins (eg p21) specically inhibit cyclin E-cdk2. Gadd45 and 1433 , which are transactivated by p53, interfere with cyclin B-Cdc2. Cdc25A and Cdc25C phosphatases, which are repressed by p53, activate cyclin ECDK2 or cyclin Bcdc2, respectively

Figure 3. Multiple mechanisms involved in p53-mediated apoptosis. p53 exerts its pro-apoptotic effect by transcription-dependent and transcription-independent modes of action. The targets of p53 transactivation represent a range of molecules, with various functions, such as the Bcl-2 family (Bax, Bid, Puma, Noxa), the apoptotic effector machinery (Apaf-1, Caspase-8, Caspase-6), cell death receptors (DR5, FAS), cell death ligands (TNFSF10, TNFS6) and other less dened factors (AEN, p53AIP, PERP, PIG3). Well-established transrepressed genes are: Bcl-2, Survivin, ARC and Gelactin-3. The p53 protein can also have direct effects in the mitochondria, as it can facilitate oligomerization of Bax and Bak as well as interact with anti-apoptotic Bcl-2, Bcl-xl and Mcl-1 proteins

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

genes [64]. These target genes belong to the intrinsic and extrinsic apoptotic pathways. Within the intrinsic death pathway, p53 induces the expression of the proapoptotic Bcl-2 family members bax and BH3-only genes bid, puma and noxa, which can function as activators, which directly stimulate Bax/Bak oligomerization, and derepressors, which promote apoptosis through displacement of anti-apoptotic proteins from the BaxBak complex (Figure 3) [67]. Further, p53 represses the apoptosis repressor with caspase recruitment domain (ARC) protein, which counteracts the apoptotic functions of Puma and Bad [68]. p53 can also promote cytochrome c release by inducing the expression of the OKL38 tumour suppressor gene, which localizes to the mitochondria and augments cytochrome c release. Loss of OKL38 correlates with tumourigenesis, and its over-expression induces apoptosis in several carcinoma cell lines [69]. Further, p53 mediates HIPK2 kinase-dependent down-regulation of Galectin-3. Galectin-3 inhibits cytochrome c release from the mitochondria and is over-expressed in a large number of human cancers [70]. Within the extrinsic pathway (Figure 3), p53 induces the expression of the death receptor Fas (CD95) and DR5 (TRAIL receptor 2) [64]. In addition, p53 induces the expression of the TNFSF10 (TRAIL) death ligand itself and the Fas ligand, TNFSF6 (FasL) [71,72]. This induction of apoptotic genes at multiple levels augments the apoptotic signalling. TRAIL protein expression was elevated in adriamycin-treated breast cancer cells and in natural killer cells in vivo after systemic treatment with 5-uorouracil. It has been proposed that TRAIL induction links p53 with the host immune response during cancer therapy [71]. A growing body of evidence implies that p53 is also capable of transactivating several other vital elements of the apoptotic machinery, including apaf-1, caspase 8 and caspase 6 [64], the apoptosis-enhancing nuclease (AEN), which is implicated in DNA fragmentation [73], PIG3, a gene involved in redox metabolism [74], and others [75].

p53 rapidly translocated to the mitochondria prior to its transcriptional function in the nucleus [82]. This was proposed to trigger an early wave of apoptosis, followed by a second wave caused mostly by transcriptionally up-regulated Puma. Mitochondrial p53 was preferentially found in radiosensitive organs and in cultured cells that respond to p53 by undergoing apoptosis [82]. Interestingly, the human p53 Arg72 polymorphic variant, which has a stronger pro-apoptotic capability, localizes better to the mitochondria [83]. Altogether, these studies suggest a role for mitochondrial p53 in the induction of apoptosis [67].

microRNAs and p53-dependent apoptosis and senescence


It was recently shown by several groups that p53 regulates the expression of microRNAs (miRNAs), where a primary role has been attributed to the miR-34 family. Inactivation of miR-34a attenuated p53-mediated apoptosis in cells exposed to genotoxic stress, suggesting a role for this microRNA in regulating p53 responses. Ectopic miR-34 expression affected the transcription of multiple genes and induced apoptosis, cell cycle arrest as well as cellular senescence [84,85]. miR-34 represses genes involved in cell cycle control, such as Cdk4 and Cyclin E2, as well as down-regulating the hepatocyte growth factor receptor, c-Met. In addition, miR-34a down-regulates Notch1 and E2F1/3 transcription factors, which are important for cell cycle progression [84,85]. Notch1 inhibits p53-dependent apoptosis through activating the mTOR-dependent PI3KAkt/PKB survival pathway [86] and by direct interaction with p53, which leads to inhibition of p53 phosphorylation and transcriptional activation [87]. Additional p53-regulated miRs were reported, including miR-15/16, which target the anti-apoptotic Bcl-2 protein, let-7, that downregulates Ras and miR-221, which in turn downregulate the CDK inhibitor p27 [84,85]. It is important to note that miRNAs are implicated in cancer, as their expression is often lost in tumour cells, and forced reduction of global miRNA expression promotes cell transformation [88].

Transcription-independent regulation of apoptosis by p53


A number of early studies suggested that under certain conditions p53 can promote apoptosis without the transcriptional activation of target genes [76,77]. The current notion is that p53 can have direct apoptogenic effects at the mitochondria (Figure 3) [67]. Mitochondrially targeted wild-type p53 fusion proteins that bypass the nucleus were sufcient to trigger effective apoptosis in p53-decient cells [78,79] and to exert tumour-suppressive activities in vivo on p53-null or mutant backgrounds [80,81]. p53 directly activates the pro-apoptotic function of Bax, Bak and VDAC by inducing their oligomerizations. In addition, p53 has been shown to associate with anti-apoptotic Bcl-2, Bcl-xl and Mcl-1, leading to the release of pro-apoptotic BH3-only proteins [67]. In mice subjected to DNA damage treatment, a subpopulation of

Decision between apoptosis and growth arrest or senescence


The determinants of cell fate upon the activation of p53 are only patially understood. Notably, it seems that certain oncogenes, such as Myc, preferentially induce ARFp53-dependent apoptosis, whereas Ras is a prototypical mediator of senescence in primary broblasts [62,89]. Also, it appears that lymphocytes are intrinsically predisposed to apoptosis, whereas broblasts and epithelial cell undergo senescence. The crucial challenge is to understand what the molecular mechanisms are that determine whether cells undergo senescence

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

V Zuckerman et al

or apoptosis, and how this knowledge could be utilized to modulate p53 responses in cancer cells. The stress-induced cellular response depends on the nature, intensity (the threshold level of stress signal) and duration of the stress signal, as well as the cell type [90]. As p53 is an integral and central part of a network of proteins responding to genotoxic stimuli, the decision between growth arrest, senescence or apoptosis is believed to be determined by the appropriate qualitative status of p53, which we term p53 conformation, localization, activity and stability status (CLASS). As suggested, the p53 CLASS can be modied by certain post-translational modications (PTMs) or by direct proteinprotein interactions [9092]. Signicantly, there are multitudes of other gene-specic regulatory components within the p53 network that are independent of and do not affect p53 status, but can shift the balance towards specic cellular outcome in response to stress [90]. p53 undergoes multiple modications in the N- and C-terminal regions [92] that mostly contribute to generic activation of p53 CLASS. This p53 activation is achieved by PTMs through a variety of mechanisms; eg by inhibiting p53 interaction with its negative regulators Mdm2 and Mdmx (Mdm4), by recruiting essential co-activators, by affecting its localization or by promoting subsequent p53 conformational changes. The dissociation from Mdm2, for example, leads to p53 stabilization and an increase in its nuclear and mitochondrial concentration. As p53 has different afnities for the promoters of distinct p53-target genes, its nuclear concentration per se can modulate the transcriptional programme [90]. Thus, the level of generic activation of p53 may be decisive for the cellular response: low levels of p53 may favour growth arrest, whereas higher levels would trigger apoptosis. Alternatively, some of p53 PTMs were proposed to enhance p53 transcriptional activity towards selective target promoters. For example, one factor that may inuence the decision to preferentially undergo apoptosis after severe DNA damage is the acetylation of p53 at Lys120, within its DNA-binding domain mediated by the MYST histone acetyltransferases (HAT) Tip60 and hMOF [93,94]. Prevention of this acetylation by mutation selectively reduced p53-induction of the pro-apoptotic bax and puma genes, without affecting p21 or mdm2. The stress-induced phosphorylation of p53 at Ser46 by either p38, DYRK2 or HIPK2 kinases also selectively enhanced direct p53 apoptotic promoter activity [9598]. One possible mechanism for this enhancement is that phosphorylation of Ser46 enables prolyl isomerase Pin1-mediated conformational change and dissociation of p53 from the inhibitor of apoptosis, iASPP, thereby promoting cell death [99]. A recent study suggested that modications of K320 and K373 modulate p53 N-terminus phosphorylations and affect the repertoire of genes induced by p53 [100]. Acetylation of K320 by PCAF causes hypophosphorylation of N-terminal residues and allows activation of p21, whereas acetylation of K373 enhances N-terminal

phosphorylations and the induction of pro-apoptotic genes. Intriguingly, mono-ubiquitylation of K320 by the zinc-nger protein E4F1 enhances the specicity of p53 towards the induction of cell cycle arrestpromoting genes [101]. A recent study using chromatin immunoprecipitation (ChIP) on Chip revealed that the binding of p53 to its target promotes is irrespective of the cellular outcome [102], suggesting that the selection for promoter specicity may occur, if at all, at the transcriptional level rather than promoter recognition. Gaining insight into the critical determinants that inuence the cellular outcome to p53 activation is particularly important for effective therapeutics by favouring apoptosis over growth arrest.

Role of p53 in tumour maintenance


The important question related to p53-based cancer therapies and tumour suppression is whether p53 loss is essential for the maintenance of established tumours. This issue has been recently addressed by several groups, using elegant mouse models to control the temporal activation of p53. Martins et al [103] demonstrated that restoring p53 function in p53-decient Emyc-driven lymphomas triggered a rapid and extensive cell death [103]. Seven days of p53 restoration conferred a 50% increase in mean survival. These mice developed secondary lymphomas by escaping from p53-mediated tumour suppression, which involved either deletion of the p53ER (oestrogen receptor fusion) allele or loss of p19ARF. Importantly, a p53 response could still be activated in E-myc secondary lymphomas with inactivated p19ARF. The observation that p53 restoration in lymphomas causes cancer regression was supported by Dickins et al [104], who used the RNAi strategy to reset active p53 in tumours. Interestingly, restoration of p53 in two types of solid tumour, radiation-induced sarcomas and hepatocellular carcinoma, led to cellular senescence and tumour clearance [105,106]. In the liver model the senescent cells were effectively eliminated by an innate immune-mediated mechanism, stimulated by pro-inammatory cytokines, which may explain the shrinkage of tumours without apoptosis being involved. Intriguingly, despite the restoration of p53 in normal and neoplastic tissues, only in the latter was p53 activated [105]. These experiments support the notion that p53 activating signals exist and persist in tumour cells, making the strategy of p53 restoration very attractive for cancer therapy. It is important to note that in these experimental models p53 was restored in the context of a p53-null background. p53 is infrequently absent from human cancers but instead is directly mutated, or its function is compromised indirectly. In both cases, the activity of restored p53 is anticipated to be at least partially compromised. Another important question to address was to determine the relative importance of DNA damage and oncogenic stress in p53-mediated

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

tumour suppression. Using switchable p53 ER knockin mice, Christophorou et al [107] showed that restoration of p53 immediately prior to whole-body irradiation resulted in an expected massive apoptosis response in all radiosensitive tissues. However, it did not protect the mice from radiation-induced lymphomagenesis. Thus, an immediate apoptotic response did not eliminate existing tumour cells. By contrast, delayed restoration of p53 after acute DNA damage responses resulted in suppression of tumourigenesis and increased survival, despite the absence of an immediate apoptotic response to irradiation. These results imply that signals activating p53 persist in lymphoma cells long after the DNA damage signals have ceased. One potential signal that may be generated as a consequence of the original genotoxic stress is oncogenic activation, capable of stimulating p19ARF induction [62,89]. Critically, delayed p53 restoration did not protect mice from radiation-induced lymphomas when ARF was absent [107]. This is consistent with the essential role of ARF in tumour suppression by p53 [108]. This does not exclude the contribution of other p53-activating signals in human tumours [109]. Overall, these important studies demonstrated that inactivation of the p53 tumour suppressor pathway is essential for tumour maintenance and that restoration of p53 can promote tumour regression. The particular tumour suppression response of p53 is tumour type- and context-dependent.

p53 and cancer therapy


Instilling malignancies with p53 that is competent for tumour suppression continues to be a primary ambition of extensive research initiatives (refer to earlier reviews for the broader historical context [110,111]). The potential for such approaches will depend upon the nature of the endogenous p53 gene. Furthermore, it is now clear that p73, the p53 family member which functions as a critical determinate of chemosensitivity, is also frequently activated by drugs designed to activate p53. This observation has led to the design of p73 activators.

When p53 is wild-type


The most extensively studied small molecule therapy applied to cancer cells bearing wt p53 is nutlin-3 (used in the following to also refer to its active enantiomer, Nutlin-3a) (Figure 4). Nutlin-3 binds to the p53-binding pocket of Mdm2, inhibits p53 attachment and consequently promotes p53 accumulation [112]. Nutlin-3-mediated elevation of p53 levels, activation of transcriptional targets and induction of apoptosis occurs independently of p53 N-terminal phosphorylation associated with DNA damage-induced p53 stabilization [113]. Efcacy of nutlin-3 has been demonstrated in a xenograft mouse model [114]. A nongenotoxic therapy that is able to selectively target cancer cells without detriment to the healthy cells of the body is the ambition of Nutlin therapy.

Figure 4. The tumour-suppressive functions of p53 can be imposed on cancer cells that have lost these capabilities, using gene therapy and pharmacological intervention. Nutlin-3 and RITA promote wild-type p53 activation by inhibiting interaction with its negative regulator, Mdm2. Mutant p53 can be converted to perform wild-type functions in the presence of CP-31398, PRIMA. p73 chemosensitivity can be restored by exposure to RETRA or 37AA, if its activities are inhibited through sequestration by mutant p53 or iASPP, respectively
J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

10

V Zuckerman et al

Importantly, Nutlin-3 was also found to inhibit the Mdmxp53 interaction (despite its lower binding afnity compared to that for Mdm2), due to conservation of the p53-binding domain between the two molecules. Retinoblastomas (RBs), initiated by the inactivation of the retinoblastoma protein, commonly advance due to the selective inactivation of the p53 pathway, mediated through the up-regulation of Mdmx. Local application of Nutlin-3 and the p53 activator topotecan were found to synergize to induce a dramatic (82-fold) reduction in RB tumour burden in a mouse model, in the absence of systemic or ocular side-effects (compared with the maximum ve-fold reduction achieved through systemic application, with signicant detrimental side-effects [115]). The advantage of local application of Nutlin-3 to target sites, as compared with systemic delivery, is further suggested by the ability of Nutlin-3 to induce cellular senescence in mouse [116] and human primary broblasts [117]. An important identication of Nutlin-3 enhancement has been demonstrated using subgenotoxic levels of compounds known to induce DNA damage at higher concentrations. The cyclin-dependent kinase (CDK) inhibitors roscovitine and DRB (which inhibits CDK9 and hence RNA polymerase II-dependent transcription) synergize with Nutlin-3 to induce apoptosis and contribute to the reduced cell viability. Importantly, the combined effect of these inhibitors is greater than either alone and, while active in promoting p53 transcription, does not stimulate a DNA damage response involving Ser15 phosphorylation [118]. Signicantly, as Nutlin-3 targets Mdm2 and Mdmx in a region that is not exclusive to p53 association, it was suspected that it may affect p53-independent functions, related to other partners that associate in the same binding domain. Additional partners that have been identied to bind Mdm2 in this region include p73, p63, E2F-1, numb and the transcription factors TFIIE [119] and HIF1a [120]. Nutlin-3 was also demonstrated to inhibit the binding between Mdm2 and E2F-1 and subsequently induce the transcriptional activation of E2F-1 in the context of DNA damage, and requires wt p53 context. Thus, the actions of Nutlins are still being delineated, with recent additional studies suggesting that at least in some leukaemic and colonic lines, Nutlin can trigger rapid apoptosis onset through the induction of p53 activity in the mitochondria, without the requirement for transcriptional target activation [121]. Another small molecular activator of wt p53 is referred to as RITA, which is the acronym for reactivation of p53 and induction of tumour-cell apoptosis (Figure 4). In contrast to Nutlin-3, RITA was characterized for its ability to bind to the Mdm2binding pocket of p53 and consequently promote the accumulation of Mdm2-free p53 molecules and induce apoptosis [122]. Subsequent studies have identied that RITA can also interact with the p53-binding domain of Mdm2, although at lower afnity, thus

suggesting that RITA has at least two target molecules [123].

When p53 is mutant


Restoring wt p53 conformation and consequently DNA binding capacity to mutant p53 through chemical correction is the ambition of the p53 reactivation approach. Intriguingly, while p53 mutation alone is insufcient for inducing cancer onset, converting a mutant structure to a wild-type form can be growth-constrictive, presumably because of the abnormally high levels and the activation signals permeating cancer cells [105]. A small molecule selected for restoring a wt p53 transcriptional transactivation function to mutant p53 is referred to as PRIMA1 (Figure 4), the acronym for p53 reactivation and initiation of massive apoptosis [124]. PRIMA has been demonstrated to induce p53-dependent growth arrest and apoptosis (although this capacity appears to be mutation- and cellular context-dependent [125]. A structural derivative PRIMA-1 (MET) reduced tumour growth in a mouse xenograft model [126]. CP-31 398 is a small molecule, styrylquinazoline, selected for its capacity to restore a wt p53 DNAbinding epitope to mutant p53 (Figure 4). CP-31 398 has been demonstrated to not only reactivate the wt p53 functions of cell cycle arrest and apoptosis in a mutant p53 cellular context [127], but also to induce elevated wt p53 levels. Intriguingly, this effect appears to be Mdm2-independent and p53Mdm2 binding does not appear to be disrupted during CP-31 398 activation of p53. A novel mode of action has been proposed for CP-31 398, in which it protects p53 from ubiquitination and thus high levels of transcriptionally active p53 accumulate [128]. The in vivo efcacy of CP-31 398 has been demonstrated in both an athymic nude mouse model [129] and in a UVB-induced skin cancer mouse model [130].

p73 activation
The ability of mutant p53 to sequester its family member p73 has been identied to enhance chemoresistance [131]. Specic activation of p73 has recently gained attention as a potential tumour therapy. Reactivation of transcriptional reporter activity (RETRA; Figure 4) is a synthetic small molecule that releases p73 from inactivating mutant p53 sequestration. A p73-dependent up-regulation of p21 and PUMA transcription was induced in response. The efcacy has been demonstrated in vivo in a xenograft mouse model [132]. A p53-derived peptide, 37AA, binds iASPP, an inhibitor of p53 family members p53, p63 and p73. In the absence of p53, 37AA binding was noted to disrupt iASPP interaction with p73 37AA, activate p73 transcriptional activity and induce cell death. Further, in vivo administration, involving transgene expression

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

11

of this peptide, led to a p73-dependent tumour repression. In tumours where iASPP inactivates p53, p63 or p73, a potential therapeutic value of iASPP inhibition has been predicted [133]. In the absence of p53, it has been demonstrated that Nutlin-3 promotes p73Mdm2 dissociation, stabilizes p73 and enhances p73 transcriptional activity, resulting in the up-regulation of the p73 target genes noxa, puma and p21 (common targets of p53) and enhances apoptosis [119]. In both a mutant p53 and a p53-null context, apoptosis was provoked through the synergistic action of Nutlin-3 and DNA damage (druginduced). Thus, a potential application for Nutlin-3 to promote chemosensitivity has been suggested also in a mutant p53 context [134], apparently by harnessing p73. Another avenue to be explored is the potential disruption of the interaction between p73 and Mdmx. Additional screening for small molecular activators of p53 transcription have also identied compounds that activate p53 transcriptional targets even in the absence of either wt or mutant p53 and this is also believed to be mediated through p73 activation [129].

Gene therapy for wt and mutant p53-bearing tumours


Adenoviral-mediated p53 gene therapy (Ad-p53), in which replication-incompetent recombinant adenovirus carries normal p53 directly into tumours, has been clinically trialled in a conned study in America (Advexin) [135] and more extensively in China (Gendicine) [136]. Advexin is currently in Phase III trials [137] after demonstrating some effect in limited earlier studies. Gendicine (injectable), launched for the treatment of squamous cell carcinoma of the head and neck, was more effective when combined with chemotherapy or radiotherapy and its therapeutic potential is currently being tested for other types of cancer, such as ovarian, non-small cell lung and many other solid tumours [136]. The endogenous p53 status, whether wt or mutant, also appears to be critically decisive to the response of Ad-p53 therapy, where endogenous wt p53 is associated with reversible cell growth arrest and mutant p53 with apoptosis. Thus, ironically, a poorer response and greater resistance to Ad-p53 therapy is observed in the wt p53 context, due to DNA repair and recovery from cell cycle arrest. Further, in cells undergoing apoptosis, phosphorylation of p53 at Thr18 and Ser20 was identied, but not in those arresting. Strikingly, enhanced levels of apoptosis were identied following the introduction of Thr18/Ser20 p53 phosphorylation mimic into wt p53 cells, and this was associated with enhanced expression of apoptosis-related genes. While the risk to normal, healthy cells harbouring wt p53 is obvious, the restricted distribution of the virus would be anticipated to act as a natural barrier to broader body damage. In fact, in brain tumour therapy, poor tumour penetration appears at

least partially responsible for the reported limited efcacy of this approach. However, in the case of gliomas, in which 70% bear wt p53, this innovative approach does appear to offer therapeutic promise [138]. Interestingly, a complementary study identied that acetylation of p53, which also promotes p53 transcription and apoptosis, induced through histone deacetylase inhibition (FK228), in conjunction with Ad-p53, enhanced the therapeutic efcacy in human cancer (expressing wt p53) xenograft mouse model [139]. Another interesting possibility for Ad-p53 therapy was demonstrated by the ability of retrovirally transferred, mitochondrially-targeted p53 (which is excluded from the nucleus and thus unable to act as a transcription factor), to induce tumour cell death in a mouse xenograft model [140]. An additional approach to adenoviral gene therapy is the introduction of AD-p73 into the HPV-wtp53 context, where p53 levels are low through the combined effect of viral E6 proteins and host E6-associated protein. As p73 is not targeted by E6, it is stable and able to induce apoptosis. Selectivity for cancer cells over normal was proposed through the application of a tumour-specic promoter (ESM6), suggested as a therapy for life-threatening uterine cervical cancers [141].

The signicance and biology of p53 isoforms


Bourdon and colleagues have identied nine novel transcripts of p53: TAp53, TAp53, TAp53 , Np53 ( Np53), Np53, Np53 , 133p53 ( 133p53), 133p53 and 133p53 , which arise from internal promoter and alternative splicing [142]. An additional transcript, p53, which lacks the Cterminal 65 amino acids, has been also identied [143]. Much of the information to date on the function of these isoforms is based on their ectopic expression in culture. For instance, Np53 inhibits p53 transcriptional activity [142] and suppression of colony growth. The Np53 transgenic mice expressed higher levels of p53 and presented decreased body mass and premature ageing associated with slower proliferation and enhanced senescence [144]. On the other hand, the TAp53 isoform elevates p53-mediated bax induction, but not p21, and increases the induction of apoptosis [142]. By contrast, p53 elevates p21 levels and induces an intra-S-phase checkpoint arrest through a p53-independent transcriptional activity [143]. Thus, it appears that different isoforms of p53 exert different effects on p53 activities and signalling. Importantly, aberrant expression of some p53 isoforms was observed in certain types of cancer. For example, elevated expression of 133p53 and reduced TAp53 expression was observed in breast cancer patients [142]; however, a different expression pattern was observed in head and neck cancer patients [145]. While it appears that p53 isoforms can affect tumour suppression by p53, additional studies are required to

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

12

V Zuckerman et al

determine their precise role and involvement in human cancer.

Conclusion
It is well established that p53 is a key tumour suppressor, integrating multiple stress conditions into appropriate cellular responses. The molecular basis for how p53 induces a transient growth arrest is well understood. The explanation for the induction of apoptosis has been extensively studied and a wealth of information on the induction of apoptosis by p53 is available. The relative contribution of each downstream effector is still to be explored. Likewise, the relative contribution of the transcriptional-independent roles of p53 requires further study. A major challenge that has been tackled by many laboratories concerns the critical determinants that inuence the particular cellular response to p53 activation. Recent studies have shed some light on post-translational modications and factors that can inuence this decision; however, additional studies are required to further explore this important issue. The recent identications of microRNAs as vital effectors of p53 growth-inhibitory functions are likely to shed light on these issues. While much of this review has focused on the role of p53 in tumour suppression in a cell autonomous manner, it is important to note that initial studies suggest that p53 may act also in a non-cell autonomous manner. Future studies will explore the contribution of p53 to tumour suppression by acting at the microinvironment in healthy cells. While the answers to many questions still need to be completed, it is now apparent that tumours are addicted to a loss of p53 function, providing a rationale for therapeutic reactivation of p53 in cancer patients.

Acknowledgements
Due to space limitations, many original important studies have not been cited directly but rather through recent reviews. Work in the authors laboratory was supported by the Israel Science Foundation (Grant No. 1341/05), by NHMRC project grants (Grant Nos 509196 and 509197), by the VESKI award, and by the EC FP6 funding of the European Commission (Contract No. 503576). This publication reects only the authors views. The European Commission is not liable for any use that may be made of the information herein. We thank Gerard Tarulli for kindly helping with formatting the gures.

References
1. Lowe SW, Cepero E, Evan G. Intrinsic tumour suppression. Nature 2004;432:307315. 2. Vousden KH, Prives C. Blinded by the light: the growing complexity of p53. Cell 2009;137:413431. 3. Vousden KH, Prives C. p53 and prognosis: new insights and further complexity. Cell 2005;120:710. 4. Royds JA, Iacopetta B. p53 and disease: when the guardian angel fails. Cell Death Differ 2006;13:10171026.

5. Blandino G, Levine AJ, Oren M. Mutant p53 gain of function: differential effects of different p53 mutants on resistance of cultured cells to chemotherapy. Oncogene 1999;18:477485. 6. Vikhanskaya F, Lee MK, Mazzoletti M, Broggini M, Sabapathy K. Cancer-derived p53 mutants suppress p53 target gene expression potential mechanism for gain of function of mutant p53. Nucleic Acids Res 2007;35:20932104. 7. Lang GA, Iwakuma T, Suh YA, Liu G, Rao VA, Parant JM, et al. Gain of function of a p53 hot spot mutation in a mouse model of Li-Fraumeni syndrome. Cell 2004;119:861872. 8. Olive KP, Tuveson DA, Ruhe ZC, Yin B, Willis NA, Bronson RT, et al. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 2004;119:847860. 9. Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA Jr, Butel JS, et al. Mice decient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 1992;356:215221. 10. Vousden KH, Lane DP. p53 in health and disease. Nat Rev Mol Cell Biol 2007;8:275283. 11. dAdda di Fagagna F, Reaper PM, Clay-Farrace L, Fiegler H, Carr P, Von Zglinicki T, et al. A DNA damage checkpoint response in telomere-initiated senescence. Nature 2003; 426:194198. 12. Herbig U, Jobling WA, Chen BP, Chen DJ, Sedivy JM. Telomere shortening triggers senescence of human cells through a pathway involving ATM, p53, and p21(CIP1), but not p16(INK4a). Mol Cell 2004;14:501513. 13. Artandi SE, DePinho RA. A critical role for telomeres in suppressing and facilitating carcinogenesis. Curr Opin Genet Dev 2000;10:3946. 14. Cosme-Blanco W, Shen MF, Lazar AJ, Pathak S, Lozano G, Multani AS, et al. Telomere dysfunction suppresses spontaneous tumorigenesis in vivo by initiating p53-dependent cellular senescence. EMBO Rep 2007;8:497503. 15. Feldser DM, Greider CW. Short telomeres limit tumor progression in vivo by inducing senescence. Cancer Cell 2007;11:461469. 16. Guo X, Deng Y, Lin Y, Cosme-Blanco W, Chan S, He H, et al. Dysfunctional telomeres activate an ATMATR-dependent DNA damage response to suppress tumorigenesis. EMBO J 2007;26:47094719. 17. Campisi J. Suppressing cancer: the importance of being senescent. Science 2005;309:886887. 18. Di Micco R, Fumagalli M, dAdda di Fagagna F. Breaking news: high-speed race ends in arrest how oncogenes induce senescence. Trends Cell Biol 2007;17:529536. 19. Sherr CJ, DePinho RA. Cellular senescence: mitotic clock or culture shock? Cell 2000;102:407410. 20. Parrinello S, Samper E, Krtolica A, Goldstein J, Melov S, Campisi J. Oxygen sensitivity severely limits the replicative lifespan of murine broblasts. Nat Cell Biol 2003;5:741747. 21. Drayton S, Peters G. Immortalisation and transformation revisited. Curr Opin Genet Dev 2002;12:98104. 22. Dimri GP. What has senescence got to do with cancer? Cancer Cell 2005;7:505512. 23. Harvey M, Sands AT, Weiss RS, Hegi ME, Wiseman RW, Pantazis P, et al. In vitro growth characteristics of embryo broblasts isolated from p53-decient mice. Oncogene 1993;8:24572467. 24. Dannenberg JH, van Rossum A, Schuijff L, te Riele H. Ablation of the retinoblastoma gene family deregulates G1 control causing immortalization and increased cell turnover under growthrestricting conditions. Genes Dev 2000;14:30513064. 25. Sage J, Miller AL, Perez-Mancera PA, Wysocki JM, Jacks T. Acute mutation of retinoblastoma gene function is sufcient for cell cycle re-entry. Nature 2003;424:223228. 26. Smogorzewska A, de Lange T. Different telomere damage signaling pathways in human and mouse cells. EMBO J 2002;21:43384348. 27. Frank KM, Sharpless NE, Gao Y, Sekiguchi JM, Ferguson DO, Zhu C, et al. DNA ligase IV deciency in mice leads to defective neurogenesis and embryonic lethality via the p53 pathway. Mol Cell 2000;5:9931002.

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

13

28. Ongusaha PP, Ouchi T, Kim KT, Nytko E, Kwak JC, Duda RB, et al. BRCA1 shifts p53-mediated cellular outcomes towards irreversible growth arrest. Oncogene 2003;22:37493758. 29. Roberson RS, Kussick SJ, Vallieres E, Chen SY, Wu DY. Escape from therapy-induced accelerated cellular senescence in p53-null lung cancer cells and in human lung cancers. Cancer Res 2005;65:27952803. 30. Roninson IB. Tumor cell senescence in cancer treatment. Cancer Res 2003;63:27052715. 31. Shay JW, Roninson IB. Hallmarks of senescence in carcinogenesis and cancer therapy. Oncogene 2004;23:29192933. 32. Campisi J, dAdda di Fagagna F. Cellular senescence: when bad things happen to good cells. Nat Rev Mol Cell Biol 2007;8:729740. 33. Ferbeyre G, de Stanchina E, Lin AW, Querido E, McCurrach ME, Hannon GJ, et al. Oncogenic ras and p53 cooperate to induce cellular senescence. Mol Cell Biol 2002;22:34973508. 34. Dimri GP, Itahana K, Acosta M, Campisi J. Regulation of a senescence checkpoint response by the E2F1 transcription factor and p14(ARF) tumor suppressor. Mol Cell Biol 2000;20:273285. 35. Wotton SF, Blyth K, Kilbey A, Jenkins A, Terry A, BernardinFried F, et al. RUNX1 transformation of primary embryonic broblasts is revealed in the absence of p53. Oncogene 2004;23:54765486. 36. Wolyniec K, Wotton S, Kilbey A, Terry A, Jenkins A, Peters G, et al. RUNX1 and its fusion oncoprotein derivative, RUNX1 ETO, induce senescence-like growth arrest independently of replicative stress. Oncogene 2009;DOI: 10.1038/onc.2009.101. 37. Chen QM, Bartholomew JC, Campisi J, Acosta M, Reagan JD, Ames BN. Molecular analysis of H2O2-induced senescent-like growth arrest in normal human broblasts: p53 and Rb control G1 arrest but not cell replication. Biochem J 1998;332(1):4350. 38. Munro J, Barr NI, Ireland H, Morrison V, Parkinson EK. Histone deacetylase inhibitors induce a senescence-like state in human cells by a p16-dependent mechanism that is independent of a mitotic clock. Exp Cell Res 2004;295:525538. 39. Sherman MY, Gabai V, OCallaghan C, Yaglom J. Molecular chaperones regulate p53 and suppress senescence programs. FEBS Lett 2007;581:37113715. 40. Gil J, Peters G. Regulation of the INK4bARFINK4a tumour suppressor locus: all for one or one for all. Nat Rev Mol Cell Biol 2006;7:667677. 41. Haupt Y. Certainly no ARFterthought: oncogenic cooperation in ARF induction a key step in tumor suppression. Cell Cycle 2003;2:113115. 42. Sharpless NE, Ramsey MR, Balasubramanian P, Castrillon DH, DePinho RA. The differential impact of p16(INK4a) or p19(ARF) deciency on cell growth and tumorigenesis. Oncogene 2004;23:379385. 43. Brookes S, Rowe J, Ruas M, Llanos S, Clark PA, Lomax M, et al. INK4a-decient human diploid broblasts are resistant to RAS-induced senescence. EMBO J 2002;21:29362945. 44. Wei W, Hemmer RM, Sedivy JM. Role of p14(ARF) in replicative and induced senescence of human broblasts. Mol Cell Biol 2001;21:67486757. 45. Ferbeyre G, de Stanchina E, Querido E, Baptiste N, Prives C, Lowe SW. PML is induced by oncogenic ras and promotes premature senescence. Genes Dev 2000;14:20152027. 46. Pearson M, Carbone R, Sebastiani C, Cioce M, Fagioli M, Saito S, et al. PML regulates p53 acetylation and premature senescence induced by oncogenic Ras. Nature 2000; 406:207210. 47. de Stanchina E, Querido E, Narita M, Davuluri RV, Pandol PP, Ferbeyre G, et al. PML is a direct p53 target that modulates p53 effector functions. Mol Cell 2004;13:523535. 48. Sun P, Yoshizuka N, New L, Moser BA, Li Y, Liao R, et al. PRAK is essential for ras-induced senescence and tumor suppression. Cell 2007;128:295308. 49. Levine AJ, Hu W, Feng Z. The p53 pathway: what questions remain to be explored? Cell Death Differ 2006;13:10271036.

50. Chan TA, Hermeking H, Lengauer C, Kinzler KW, Vogelstein B. 1433 is required to prevent mitotic catastrophe after DNA damage. Nature 1999;401:616620. 51. Peng CY, Graves PR, Thoma RS, Wu Z, Shaw AS, PiwnicaWorms H. Mitotic and G2 checkpoint control: regulation of 1433 protein binding by phosphorylation of Cdc25C on serine216. Science 1997;277:15011505. 52. Jin S, Antinore MJ, Lung FD, Dong X, Zhao H, Fan F, et al. The GADD45 inhibition of Cdc2 kinase correlates with GADD45-mediated growth suppression. J Biol Chem 2000;275:1660216608. 53. Zhan Q, Antinore MJ, Wang XW, Carrier F, Smith ML, Harris CC, et al. Association with Cdc2 and inhibition of Cdc2/Cyclin B1 kinase activity by the p53-regulated protein Gadd45. Oncogene 1999;18:28922900. 54. Spurgers KB, Gold DL, Coombes KR, Bohnenstiehl NL, Mullins B, Meyn RE, et al. Identication of cell cycle regulatory genes as principal targets of p53-mediated transcriptional repression. J Biol Chem 2006;281:2513425142. 55. Rother K, Kirschner R, Sanger K, Bohlig L, Mossner J, Engeland K. p53 downregulates expression of the G1 /S cell cycle phosphatase Cdc25A. Oncogene 2007;26:19491953. 56. St Clair S, Giono L, Varmeh-Ziaie S, Resnick-Silverman L, Liu WJ, Padi A, et al. DNA damage-induced downregulation of Cdc25C is mediated by p53 via two independent mechanisms: one involves direct binding to the cdc25C promoter. Mol Cell 2004;16:725736. 57. Caprara G, Zamponi R, Melixetian M, Helin K. Isolation and characterization of DUSP11, a novel p53 target gene. J Cell Mol Med 2008;Dec 16 [E-pub ahead of print]. 58. Doumont G, Martoriati A, Beekman C, Bogaerts S, Mee PJ, Bureau F, et al. G1 checkpoint failure and increased tumor susceptibility in mice lacking the novel p53 target Ptprv. EMBO J 2005;24:30933103. 59. Saigusa K, Imoto I, Tanikawa C, Aoyagi M, Ohno K, Nakamura Y, et al. RGC32, a novel p53-inducible gene, is located on centrosomes during mitosis and results in G2 /M arrest. Oncogene 2007;26:11101121. 60. Shetty S, Shetty P, Idell S, Velusamy T, Bhandary YP, Shetty RS. Regulation of plasminogen activator inhibitor-1 expression by tumor suppressor protein p53. J Biol Chem 2008;283:1957019580. 61. Mu XC, Higgins PJ. Differential growth state-dependent regulation of plasminogen activator inhibitor type-1 expression in senescent IMR-90 human diploid broblasts. J Cell Physiol 1995;165:647657. 62. Serrano M, Lin AW, McCurrach ME, Beach D, Lowe SW. Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell 1997;88:593602. 63. Kortlever RM, Higgins PJ, Bernards R. Plasminogen activator inhibitor-1 is a critical downstream target of p53 in the induction of replicative senescence. Nat Cell Biol 2006;8:877884. 64. Haupt S, Berger M, Goldberg Z, Haupt Y. Apoptosis the p53 network. J Cell Sci 2003;116:40774085. 65. Chipuk JE, Green DR. Dissecting p53-dependent apoptosis. Cell Death Differ 2006;13:9941002. 66. Meulmeester E, Jochemsen AG. p53: a guide to apoptosis. Curr Cancer Drug Targets 2008;8:8797. 67. Vaseva AV, Moll UM. The mitochondrial p53 pathway. Biochim Biophys Acta 2009;1787:414420. 68. Li YZ, Lu DY, Tan WQ, Wang JX, Li PF. p53 initiates apoptosis by transcriptionally targeting the antiapoptotic protein ARC. Mol Cell Biol 2008;28:564574. 69. Yao H, Li P, Venters BJ, Zheng S, Thompson PR, Pugh BF, et al. Histone Arg modications and p53 regulate the expression of OKL38, a mediator of apoptosis. J Biol Chem 2008;283:2006020068. 70. Cecchinelli B, Lavra L, Rinaldo C, Iacovelli S, Gurtner A, Gasbarri A, et al. Repression of the antiapoptotic molecule galectin-3 by homeodomain-interacting protein kinase 2-activated p53 is required for p53-induced apoptosis. Mol Cell Biol 2006;26:47464757.

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

14

V Zuckerman et al

71. Kuribayashi K, Krigsfeld G, Wang W, Xu J, Mayes PA, Dicker DT, et al. TNFSF10 (TRAIL), a p53 target gene that mediates p53-dependent cell death. Cancer Biol Ther 2008;7:20342038. 72. Maecker HL, Koumenis C, Giaccia AJ. p53 promotes selection for Fas-mediated apoptotic resistance. Cancer Res 2000;60:46384644. 73. Kawase T, Ichikawa H, Ohta T, Nozaki N, Tashiro F, Ohki R, et al. p53 target gene AEN is a nuclear exonuclease required for p53-dependent apoptosis. Oncogene 2008;27:37973810. 74. Polyak K, Xia Y, Zweier JL, Kinzler KW, Vogelstein B. A model for p53-induced apoptosis. Nature 1997;389:300305. 75. Riley T, Sontag E, Chen P, Levine A. Transcriptional control of human p53-regulated genes. Nat Rev Mol Cell Biol 2008;9:402412. 76. Moll UM, Wolff S, Speidel D, Deppert W. Transcriptionindependent pro-apoptotic functions of p53. Curr Opin Cell Biol 2005;17:631636. 77. Green DR, Kroemer G. Cytoplasmic functions of the tumour suppressor p53. Nature 2009;458:11271130. 78. Marchenko ND, Zaika A, Moll UM. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling. J Biol Chem 2000;275:1620216212. 79. Mihara M, Erster S, Zaika A, Petrenko O, Chittenden T, Pancoska P, et al. p53 has a direct apoptogenic role at the mitochondria. Mol Cell 2003;11:577590. 80. Palacios G, Moll UM. Mitochondrially targeted wild-type p53 suppresses growth of mutant p53 lymphomas in vivo. Oncogene 2006;25:61336139. 81. Talos F, Petrenko O, Mena P, Moll UM. Mitochondrially targeted p53 has tumor suppressor activities in vivo. Cancer Res 2005;65:99719981. 82. Erster S, Mihara M, Kim RH, Petrenko O, Moll UM. In vivo mitochondrial p53 translocation triggers a rapid rst wave of cell death in response to DNA damage that can precede p53 target gene activation. Mol Cell Biol 2004;24:67286741. 83. Dumont P, Leu JI, Della Pietra AC, 3rd, George DL, Murphy M. The codon 72 polymorphic variants of p53 have markedly different apoptotic potential. Nat Genet 2003;33:357365. 84. He L, He X, Lowe SW, Hannon GJ. microRNAs join the p53 network another piece in the tumour-suppression puzzle. Nat Rev Cancer 2007;7:819822. 85. Hermeking H. p53 enters the microRNA world. Cancer Cell 2007;12:414418. 86. Mungamuri SK, Yang X, Thor AD, Somasundaram K. Survival signaling by Notch1: mammalian target of rapamycin (mTOR)dependent inhibition of p53. Cancer Res 2006;66:47154724. 87. Kim SB, Chae GW, Lee J, Park J, Tak H, Chung JH, et al. Activated Notch1 interacts with p53 to inhibit its phosphorylation and transactivation. Cell Death Differ 2007;14:982991. 88. Kumar MS, Lu J, Mercer KL, Golub TR, Jacks T. Impaired microRNA processing enhances cellular transformation and tumorigenesis. Nat Genet 2007;39:673677. 89. Zindy F, Eischen CM, Randle DH, Kamijo T, Cleveland JL, Sherr CJ, et al. Myc signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev 1998;12:24242433. 90. Espinosa JM. Mechanisms of regulatory diversity within the p53 transcriptional network. Oncogene 2008;27:40134023. 91. Das S, Boswell SA, Aaronson SA, Lee SW. P53 promoter selection: choosing between life and death. Cell Cycle 2008;7:154157. 92. Lavin MF, Gueven N. The complexity of p53 stabilization and activation. Cell Death Differ 2006;13:941950. 93. Sykes SM, Mellert HS, Holbert MA, Li K, Marmorstein R, Lane WS, et al. Acetylation of the p53 DNA-binding domain regulates apoptosis induction. Mol Cell 2006;24:841851. 94. Tang Y, Luo J, Zhang W, Gu W. Tip60-dependent acetylation of p53 modulates the decision between cell-cycle arrest and apoptosis. Mol Cell 2006;24:827839. 95. Bulavin DV, Saito S, Hollander MC, Sakaguchi K, Anderson CW, Appella E, et al. Phosphorylation of human p53 by p38

96.

97.

98.

99.

100.

101.

102.

103. 104.

105.

106.

107.

108.

109. 110. 111. 112.

113.

114.

115.

116.

kinase coordinates N-terminal phosphorylation and apoptosis in response to UV radiation. EMBO J 1999;18:68456854. DOrazi G, Cecchinelli B, Bruno T, Manni I, Higashimoto Y, Saito S, et al. Homeodomain-interacting protein kinase-2 phosphorylates p53 at Ser 46 and mediates apoptosis. Nat Cell Biol 2002;4:1119. Feng L, Hollstein M, Xu Y. Ser46 phosphorylation regulates p53-dependent apoptosis and replicative senescence. Cell Cycle 2006;5:28122819. Taira N, Nihira K, Yamaguchi T, Miki Y, Yoshida K. DYRK2 is targeted to the nucleus and controls p53 via Ser46 phosphorylation in the apoptotic response to DNA damage. Mol Cell 2007;25:725738. Mantovani F, Tocco F, Girardini J, Smith P, Gasco M, Lu X, et al. The prolyl isomerase Pin1 orchestrates p53 acetylation and dissociation from the apoptosis inhibitor iASPP. Nat Struct Mol Biol 2007;14:912920. Knights CD, Catania J, Di Giovanni S, Muratoglu S, Perez R, Swartzbeck A, et al. Distinct p53 acetylation cassettes differentially inuence gene-expression patterns and cell fate. J Cell Biol 2006;173:533544. Le Cam L, Linares LK, Paul C, Julien E, Lacroix M, Hatchi E, et al. E4F1 is an atypical ubiquitin ligase that modulates p53 effector functions independently of degradation. Cell 2006;127:775788. Shaked H, Shiff I, Kott-Gutkowski M, Siegfried Z, Haupt Y, Simon I. Chromatin immunoprecipitation-on-chip reveals stressdependent p53 occupancy in primary normal cells but not in established cell lines. Cancer Res 2008;68:96719677. Martins CP, Brown-Swigart L, Evan GI. Modeling the therapeutic efcacy of p53 restoration in tumors. Cell 2006;127:13231334. Dickins RA, McJunkin K, Hernando E, Premsrirut PK, Krizhanovsky V, Burgess DJ, et al. Tissue-specic and reversible RNA interference in transgenic mice. Nat Genet 2007;39:914921. Ventura A, Kirsch DG, McLaughlin ME, Tuveson DA, Grimm J, Lintault L, et al. Restoration of p53 function leads to tumour regression in vivo. Nature 2007;445:661665. Xue W, Zender L, Miething C, Dickins RA, Hernando E, Krizhanovsky V, et al. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature 2007;445:656660. Christophorou MA, Ringshausen I, Finch AJ, Swigart LB, Evan GI. The pathological response to DNA damage does not contribute to p53-mediated tumour suppression. Nature 2006;443:214217. Efeyan A, Garcia-Cao I, Herranz D, Velasco-Miguel S, Serrano M. Tumour biology: policing of oncogene activity by p53. Nature 2006;443:159. Kastan MB. Wild-type p53: tumors cant stand it. Cell 2007;128:837840. Dey A, Verma CS, Lane DP. Updates on p53: modulation of p53 degradation as a therapeutic approach. Br J Cancer 2008;98:48. Haupt S, Haupt Y. Importance of p53 for cancer onset and therapy. Anticancer Drugs 2006;17:725732. Fry DC, Graves B, Vassilev LT. Development of E3substrate (MDM2p53)-binding inhibitors: structural aspects. Methods Enzymol 2005;399:622633. Thompson T, Tovar C, Yang H, Carvajal D, Vu BT, Xu Q, et al. Phosphorylation of p53 on key serines is dispensable for transcriptional activation and apoptosis. J Biol Chem 2004;279:5301553022. Vassilev LT, Vu BT, Graves B, Carvajal D, Podlaski F, Filipovic Z, et al. In vivo activation of the p53 pathway by smallmolecule antagonists of MDM2. Science 2004;303:844848. Laurie NA, Donovan SL, Shih CS, Zhang J, Mills N, Fuller C, et al. Inactivation of the p53 pathway in retinoblastoma. Nature 2006;444:6166. Efeyan A, Ortega-Molina A, Velasco-Miguel S, Herranz D, Vassilev LT, Serrano M. Induction of p53-dependent senescence by the MDM2 antagonist nutlin-3a in mouse cells of broblast origin. Cancer Res 2007;67:73507357.

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Tumour suppression by p53: importance of apoptosis and cellular senescence

15

117. Kumamoto K, Spillare EA, Fujita K, Horikawa I, Yamashita T, Appella E, et al. Nutlin-3a activates p53 to both down-regulate inhibitor of growth 2 and up-regulate mir-34a, mir-34b, and mir-34c expression, and induce senescence. Cancer Res 2008;68:31933203. 118. Cheok CF, Dey A, Lane DP. Cyclin-dependent kinase inhibitors sensitize tumor cells to nutlin-induced apoptosis: a potent drug combination. Mol Cancer Res 2007;5:11331145. 119. Lau LM, Nugent JK, Zhao X, Irwin MS. HDM2 antagonist Nutlin-3 disrupts p73-HDM2 binding and enhances p73 function. Oncogene 2008;27:9971003. 120. LaRusch GA, Jackson MW, Dunbar JD, Warren RS, Donner DB, Mayo LD. Nutlin3 blocks vascular endothelial growth factor induction by preventing the interaction between hypoxia inducible factor 1 and Hdm2. Cancer Res 2007;67:450454. 121. Vaseva AV, Marchenko ND, Moll UM. The transcriptionindependent mitochondrial p53 program is a major contributor to nutlin-induced apoptosis in tumor cells. Cell Cycle 2009;8:(in press). 122. Issaeva N, Bozko P, Enge M, Protopopova M, Verhoef LG, Masucci M, et al. Small molecule RITA binds to p53, blocks p53-HDM-2 interaction and activates p53 function in tumors. Nat Med 2004;10:13211328. 123. Espinoza-Fonseca LM. Targeting MDM2 by the small molecule RITA: towards the development of new multi-target drugs against cancer. Theor Biol Med Model 2005;2:38. 124. Bykov VJ, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E, Chumakov P, et al. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat Med 2002;8:282288. 125. Weinmann L, Wischhusen J, Demma MJ, Naumann U, Roth P, Dasmahapatra B, et al. A novel p53 rescue compound induces p53-dependent growth arrest and sensitises glioma cells to Apo2L/TRAIL-induced apoptosis. Cell Death Differ 2008;15:718729. 126. Bykov VJ, Zache N, Stridh H, Westman J, Bergman J, Selivanova G, et al. PRIMA-1(MET) synergizes with cisplatin to induce tumor cell apoptosis. Oncogene 2005;24:34843491. 127. Foster BA, Coffey HA, Morin MJ, Rastinejad F. Pharmacological rescue of mutant p53 conformation and function. Science 1999;286:25072510. 128. Wang W, Takimoto R, Rastinejad F, El-Deiry WS. Stabilization of p53 by CP-31398 inhibits ubiquitination without altering phosphorylation at serine 15 or 20 or MDM2 binding. Mol Cell Biol 2003;23:21712181. 129. Wang W, Kim SH, El-Deiry WS. Small-molecule modulators of p53 family signaling and antitumor effects in p53-decient human colon tumor xenografts. Proc Natl Acad Sci USA 2006;103:1100311008. 130. Tang X, Zhu Y, Han L, Kim AL, Kopelovich L, Bickers DR, et al. CP-31398 restores mutant p53 tumor suppressor function

131.

132.

133.

134.

135. 136.

137.

138.

139.

140.

141.

142.

143.

144.

145.

and inhibits UVB-induced skin carcinogenesis in mice. J Clin Invest 2007;117:37533764. Irwin MS, Kondo K, Marin MC, Cheng LS, Hahn WC, Kaelin WG Jr. Chemosensitivity linked to p73 function. Cancer Cell 2003;3:403410. Kravchenko JE, Ilyinskaya GV, Komarov PG, Agapova LS, Kochetkov DV, Strom E, et al. Small-molecule RETRA suppresses mutant p53-bearing cancer cells through a p73-dependent salvage pathway. Proc Natl Acad Sci USA 2008;105:63026307. Bell HS, Dufes C, OPrey J, Crighton D, Bergamaschi D, Lu X, et al. A p53-derived apoptotic peptide derepresses p73 to cause tumor regression in vivo. J Clin Invest 2007;117:10081018. Ambrosini G, Sambol EB, Carvajal D, Vassilev LT, Singer S, Schwartz GK. Mouse double minute antagonist Nutlin-3a enhances chemotherapy-induced apoptosis in cancer cells with mutant p53 by activating E2F1. Oncogene 2007;26:34733481. Lo HW, Day CP, Hung MC. Cancer-specic gene therapy. Adv Genet 2005;54:235255. Peng Z. Current status of gendicine in China: recombinant human Ad-p53 agent for treatment of cancers. Hum Gene Ther 2005;16:10161027. Olivier M, Petitjean A, Marcel V, Petre A, Mounawar M, Plymoth A, et al. Recent advances in p53 research: an interdisciplinary perspective. Cancer Gene Ther 2009;16:112. Nakamizo A, Amano T, Zhang W, Zhang XQ, Ramdas L, Liu TJ, et al. Phosphorylation of Thr18 and Ser20 of p53 in Adp53-induced apoptosis. Neuro Oncol 2008;10:275291. Sasaki Y, Negishi H, Idogawa M, Suzuki H, Mita H, Toyota M, et al. Histone deacetylase inhibitor FK228 enhances adenovirusmediated p53 family gene therapy in cancer models. Mol Cancer Ther 2008;7:779787. Palacios G, Crawford HC, Vaseva A, Moll UM. Mitochondrially targeted wild-type p53 induces apoptosis in a solid human tumor xenograft model. Cell Cycle 2008;7:25842590. Lee JJ, Kim S, Yeom YI, Heo DS. Enhanced specicity of the p53 family proteins-based adenoviral gene therapy in uterine cervical cancer cells with E2F1-responsive promoters. Cancer Biol Ther 2006;5:15021510. Bourdon JC, Fernandes K, Murray-Zmijewski F, Liu G, Diot A, Xirodimas DP, et al. p53 isoforms can regulate p53 transcriptional activity. Genes Dev 2005;19:21222137. Rohaly G, Chemnitz J, Dehde S, Nunez AM, Heukeshoven J, Deppert W, et al. A novel human p53 isoform is an essential element of the ATR-intra-S phase checkpoint. Cell 2005;122:2132. Maier B, Gluba W, Bernier B, Turner T, Mohammad K, Guise T, et al. Modulation of mammalian life span by the short isoform of p53. Genes Dev 2004;18:306319. Boldrup L, Bourdon JC, Coates PJ, Sjostrom B, Nylander K. Expression of p53 isoforms in squamous cell carcinoma of the head and neck. Eur J Cancer 2007;43:617623.

Teaching Materials
Power Point slides of the gures from this Review may be found in the supporting information.

J Pathol 2009; 219: 315 DOI: 10.1002/path Copyright 2009 Pathological Society of Great Britain and Ireland. Published by John Wiley & Sons, Ltd.

Você também pode gostar