Pharmacology & Therapeutics 142 (2014) 99–113

Contents lists available at ScienceDirect

Pharmacology & Therapeutics journal homepage: www.elsevier.com/locate/pharmthera

Associate editor: M. Mouradian

p53 in neurodegenerative diseases and brain cancers Frédéric Checler ⁎, Cristine Alves da Costa ⁎⁎ Institut de Pharmacologie Moléculaire et Cellulaire, UMR7275 CNRS/UNS, “Labex Distalz”, 660 route des Lucioles, 06560, Sophia-Antipolis, Valbonne, France

a r t i c l e

i n f o

a b s t r a c t More than thirty years elapsed since a protein, not yet called p53 at the time, was detected to bind SV40 during viral infection. Thousands of papers later, p53 evolved as the main tumor suppressor involved in growth arrest and apoptosis. A lot has been done but the protein has not yet revealed all its secrets. Particularly important is the observation that in totally distinct pathologies where apoptosis is either exacerbated or impaired, p53 appears to play a central role. This is exemplified for Alzheimer's and Parkinson's diseases that represent the two main causes of age-related neurodegenerative affections, where cell death enhancement appears as one of the main etiological paradigms. Conversely, in cancers, about half of the cases are linked to mutations in p53 leading to the impairment of p53-dependent apoptosis. The involvement of p53 in these pathologies has driven a huge amount of studies aimed at designing chemical tools or biological approaches to rescue p53 defects or over-activity. Here, we describe the data linking p53 to neurodegenerative diseases and brain cancers, and we document the various strategies to interfere with p53 dysfunctions in these disorders. © 2013 Elsevier Inc. All rights reserved.

Keywords: p53 Signaling Alzheimer's disease Parkinson's disease Cerebral cancers Therapeutics

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . 2. p53 and Alzheimer's disease . . . . . . . . . . . . . . 3. p53 and Parkinson's disease . . . . . . . . . . . . . . 4. p53 and prions . . . . . . . . . . . . . . . . . . . . 5. p53 and brain cancers . . . . . . . . . . . . . . . . 6. Pharmacological and gene therapy strategies targeting p53 7. Concluding remarks . . . . . . . . . . . . . . . . . Conflict of interest statement . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction The first report on the presence of a previously unknown protein, the expression of which increased upon SV40 infection, came out over thirty years ago (Lane & Crawford, 1979; Linzer & Levine, 1979). Indeed, a 53 kDa phosphoprotein was consistently detected in a wide series of human tumor cells (Crawford, Pim, Gurney, Goodfellow, & TaylorPapadimitriou, 1981). Interestingly, this protein that was baptized p53 and that is currently recognized as the most famous tumor suppressor ⁎ Corresponding author. Tel.: +33 4 93953460; fax: +33 4 93953408. ⁎⁎ Corresponding author. Tel.: +33 4 93953457; fax: +33 4 93953408. E-mail addresses: [email protected] (F. Checler), [email protected] (C. Alves da Costa). 0163-7258/$ – see front matter © 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.pharmthera.2013.11.009

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

99 100 102 104 105 106 107 107 107 107

was first considered as an oncogene. Thus, the over-expression of p53 cDNA isolated from p53 clonal cells was highly tumorigenic and could transform embryonic cells (Eliyahu, Michalovitz, & Oren, 1985; Eliyahu, Raz, Gruss, Givol, & Oren, 1984). It appeared later that the isolated clones harbored a mutated p53, while wild-type p53 prevented oncogene-induced cell transformation (Finlay, Hinds, & Levine, 1989). These observations indicated that p53 likely behaved as a tumor suppressor in a normal context, and that p53 function could have been impaired by mutations. Obviously, the possibility that such dysfunction could wholly account for, or contribute to, cancer onset and/or progression was rapidly the center of a huge amount of research. Several lines of data supported this view but the likely most convincing evidence came from the demonstration that p53 knockout mice develop tumors very early (Donehower et al., 1992), and that about 50% of human cancers

100

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

were associated with p53 mutations on both alleles of the gene (Olivier, Hollstein, & Hainaut, 2010; Robles & Harris, 2010). Subsequently, p53 was shown to behave as a transcription factor (el-Deiry, Kern, Pietenpol, Kinzler, & Vogelstein, 1992; Funk, Pak, Karas, Wright, & Shay, 1992) able to modulate a series of genes, the list of which is still increasing today. Notably, p53 regulates several genes involved in cell cycle control and apoptotic pathways (Beckerman & Prives, 2010), and its main function is in arresting cell cycle and promoting apoptosis. In this context, it makes sense to envision that a p53 defect could trigger cell transformation and cancer. When it became apparent later that apoptosis exacerbation is a characteristic of several neurodegenerative diseases, p53 came on stage in research concerning several of these neurological disorders. This review will focus particularly on the contribution of p53 in brain cancers and in Alzheimer's, Parkinson's and prion diseases, although p53 is involved in several types of cancers (Lane & Levine, 2010; Olivier et al., 2010; Robles & Harris, 2010) and implicated in other neurodegenerative disorders such as Huntington's disease (Bae et al., 2005; Burton, 2005; Feng et al., 2006).

2. p53 and Alzheimer's disease Alzheimer's disease (AD) is the most frequent age-related neurodegenerative disease (Hebert, Weuve, Scherr, & Evans, 2013; Holtzman, Mandelkow, & Selkoe, 2012; Prestia et al., 2013). Its pathology is characterized by two main histological lesions, the senile plaques and the neurofibrillary tangles (Selkoe, 1991; Suh & Checler, 2002). Tangles (Spillantini & Goedert, 1998) are essentially due to the intracellular accumulation of hyperphosphorylated tau protein (Buee et al., 2010; Mandelkow, 1999), while senile plaques are extracellular deposits mainly due to the aggregation of a set of various hydrophobic peptides semantically gathered under the term of amyloid-β (Αβ) peptides (Glenner & Wong, 1984; Masters et al., 1985). Canonical Aβ consists of 40–42 amino-acid-long sequences derived from a transmembrane precursor βAPP (β-amyloid precursor protein) by sequential proteolytic actions by several proteases named secretases (Checler, 1995). An acidic protease, called β-secretase (identified as BACE1, memapsin 2 or ASP2; Hussain et al., 1999; Lin et al., 2000; Sinha et al., 1999; Vassar et al., 1999; Yan et al., 1999), triggers the first and rate-limiting cleavage liberating C99, the C-terminal fragment of βAPP that harbors the Aβ sequence at its N-terminus. C99 then undergoes an epsilon secretase cleavage (Gu et al., 2001; Passer et al., 2000; Sastre et al., 2001; Weidemann et al., 2002) that releases AICD (APP IntraCellular Domain; Belyaev et al., 2010; Flammang et al., 2012; Goodger et al., 2009) that can act as a transcription factor (Konietzko, 2012; Pardossi-Piquard & Checler, 2012), and APPε. The latter catabolite undergoes an additional C-terminal truncation by both presenilin-dependent (De Strooper et al., 1998) and independent (Armogida et al., 2001) γ-secretases that ultimately liberate “full length” Aβ (flAβ) species (Lefranc-Jullien, Sunyach, & Checler, 2006). The presenilin-dependent γ-secretase is a high molecular weight complex consisting of at least four components, namely presenilin 1 or 2 (Rogaev et al., 1995; Sherrington et al., 1996), nicastrin (Yu et al., 2000), Pen-2 and Aph-1 (Francis et al., 2002). Although the presenilin-independent γ-secretase has yet to be identified, it has been characterized as a serine protease against which a series of isocoumarin blockers have been designed (Petit et al., 2001; Peuchmaur et al., 2013). Once produced, flAβ is bio-transformed by a series of additional modifications by either exopeptidase-mediated N- or C-terminal truncations (Funamoto et al., 2004; Sevalle et al., 2009; Tekirian et al., 1998), glutaminyl cyclization (Hartlage-Rubsamen et al., 2011; Schilling et al., 2008) or internal cleavages (Caillava et al., submitted for publication; Fluhrer et al., 2003; Shi et al., 2003) giving rise to fragments harboring their own aggregation properties, biological functions or toxicities (Wirths et al., 2009). Very little is known about the relative contribution of these various Aβ-related fragments to AD pathology, but several lines of evidence indicate that some of these

fragments or some of the above-described proteins involved in βAPP processing could control p53-dependent cell death. In Alzheimer's disease, neuronal cell death by apoptosis has been a matter of discussion (see http://www.alzforum.org/res/for/journal/detail. asp?liveID=113). The inconsistencies in this field could likely be due to the variability of the biological samples, the post-mortem delays in obtaining anatomical tissues (Anderson, Su, & Cotman, 1996), the wide spreading of the lesions, and the technical limits of sensitivity for in situ detection of apoptotic markers. Nevertheless, several lines of evidence suggest that cell death occurs to a certain extent by apoptosis in AD-affected brains (Kusiak, Izzo, & Zhao, 1996). Thus, in situ examination of the nuclei morphology and DNA fragmentation indicated enhanced apoptotic-like phenotype in the hippocampus and entorhinal cortex of AD-affected patients (Anderson, Stoltzner, Lai, Su, & Nixon, 2000; Su, Anderson, Cummings, & Cotman, 1994). These pro-apoptotic stigmata are in good agreement with a study showing that apoptotic neurons in AD are often associated with enhanced intracellular Aβ42 expression (Chui et al., 2001). Furthermore, a report on the expression of a series of gene products involved in cell death cascades showed an increase of the pro-apoptotic protein Bax-α in AD-affected hippocampus with no alteration of the anti-apoptotic protein Bcl-2 (Sajan et al., 2007). It should be noted, however, that another study reported on a selective increase of Bcl-2 in damaged neurons while it appears decreased in tangles bearing neurons (Su, Satou, Anderson, & Cotman, 1996). Interestingly, not only neuronal cells but also astrocytes could undergo apoptosis in AD-affected cortical areas (Kobayashi et al., 2004). The enhanced cell death possibly occurring in AD is further supported by several observations indicating that βAPP-derived catabolites could indeed trigger apoptosis in cells and animal models. Thus, aggregated Aβ42 correlates with caspase-8 mRNA levels in the temporal region of AD-affected brains (Matsui et al., 2006) and is increased in damaged neurons (Ohyagi et al., 2000). A derivative fragment of Aβ (Aβ25–35), that is not consistently detectable in AD-affected brains but can mimic some of Aβ-related phenotypes (Diaz et al., 2010; Meunier, Leni, & Maurice, 2006), indeed promotes apoptosis in lymphocytes (Velez-Pardo, Ospina, & Jimenez del Rio, 2002) and in PC12 cells (Martin et al., 2001). This Aβ-mediated enhanced neurodegeneration also occurs in vivo in a mouse model of Alzheimer's disease (transgenic mice harboring a presenilin 1 mutation) where Aβ accumulation triggers retinal degeneration (Ning, Cui, To, Ashe, & Matsubara, 2008). Besides Aβ and its derivatives, various C-terminal fragments of βAPP have been shown to induce cell death. Thus, low concentrations of the recombinant C-terminal βAPP fragment CT105 triggers apoptosis in astrocytes (Bach et al., 2001). Higher doses of the same fragment (10 μM) disrupt membrane integrity as assessed by lactate dehydrogenase release from primary cultured neurons (Kim & Suh, 1996), prevents uptake (Kim, Park, & Suh, 1998) and increases intracellular concentrations of calcium (Kim et al., 2000) thereby dramatically perturbing cellular calcium homeostasis. This set of data agrees well with a study demonstrating an association between excitotoxicity, DNA fragmentation and caspase-3 activation (Masliah, Mallory, Alford, Tanaka, & Hansen, 1998) in AD brains and corroborates a study showing that Aβ affects Ca2+ homeostasis and behavioral deficits of transgenic mice via an interplay with ryanodine receptors (Oulès et al., 2012). Finally, we showed that C31, a caspase-3-derived C-terminal fragment of βAPP that accumulates in AD brains (Lu et al., 2000), enhances toxicity in TSM1 neurons although by a caspase-3 independent mechanism (Dumanchin-Njock, Alves da Costa, Mercken, Pradier, & Checler, 2001). βAPP-derived fragments can also modulate p53. For example, Aβ42 directly activates the transactivation of the p53 promoter (Ohyagi et al., 2005). Moreover, oxidative DNA damage triggers Aβ42 translocation into the nucleus along with increased p53 mRNA levels. Finally, of most interest, in transgenic mouse models of AD and in AD brains, neurons that harbor ill-shaped morphology display both Aβ42 and p53 accumulation (Ohyagi et al., 2005). On a more functional stand point, p53 blockade by its transcriptional inhibitor pifithrin-α (Komarov

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

et al., 1999a, 1999b) lowers microglial responsiveness to the toxicity induced by Aβ-related peptides (Davenport, Sevastou, Hooper, & Pocock, 2010). Besides Aβ, the ε-secretase-derived βAPP fragment AICD that can act as a transcription factor (Pardossi-Piquard & Checler, 2012) also controls p53 at a transcriptional level (Alves da Costa, Sunyach et al. 2006; Ozaki et al., 2006). It should be noted that cell cycle re-entry has been postulated to account for exacerbated neuronal apoptosis in various neurodegenerative diseases (Folch et al., 2012) and that soluble Aβ itself could trigger neuronal cell cycle re-entry (Seward et al., 2013). Although the latter phenotype appears to be mediated by tau protein (Seward et al., 2013), the above-described direct link between Aβ and p53, that is clearly involved in cell cycle control (Chandler & Peters, 2013; Nakamura, 2004), could be envisioned as a molecular pathway accounting for, at least in part, the pro-apoptotic lesions observed in AD. The overall above-cited data concur to suggest the possibility of an increased cell death in AD, a process probably linked to the alteration of βAPP processing. They also indicate a functional interplay between the βAPP-related fragments and the tumor suppressor p53, therefore, suggesting a possible direct link between the two cellular partners. Several studies examined more directly the relevance of p53 as a marker/effector of AD pathology. Kitamura and colleagues compared p53-like immunoreactivity in post-mortem AD-affected brains and controls. They reported a consistent increase in p53 expression in the pathological temporal cortex areas, and more precisely at the glial level (Kitamura et al., 1997). This astrocyte-selective p53 label was partly confirmed by a concomitant study showing that indeed, p53 detection by in situ immunolocalization revealed enhanced p53 expression not only in astrocyte and oligodendrocyte populations of frontal and temporal lobes but also in numerous cortical neurons (de la Monte, Sohn, & Wands, 1997). Of interest, this increased p53 expression coincided with enhanced DNA fragmentation and Fas expression (de la Monte et al., 1997), indicating that p53 could contribute to cell death in AD brains. It should be noted that a co-increase in p53 and Fas expressions was also observed in Down syndrome affected brains (Seidl, FangKircher, Bidmon, Cairns, & Lubec, 1999), another pathology displaying AD-like histological alterations (Lott, 1982; Mann, 1988) likely due to the extra copy of the chromosome 21 that bears the βAPP-encoding gene (Holtzman & Epstein, 1992). These studies were confirmed by more recent data showing enhanced levels of p53 in the superior temporal gyrus (Hooper et al., 2007) as well as in the inferior parietal lobule of AD-affected brains (Cenini, Sultana, Memo, & Butterfield, 2008). Interestingly, the latter group correlated the enhanced p53 expression with markers observed in oxidative stress-induced neuronal cell death (Cenini et al., 2008). The parallel increase observed for p53 and cell death in AD-affected brains could be understood with respect to the widely documented function of p53 in cell cycle arrest, DNA repair and apoptosis (Aylon & Oren, 2011; Bourdon, De Laurenzi, Melino, & Lane, 2003; Levine & Oren, 2009; Vogelstein & Kinzler, 1992). Thus, enhanced p53 expression could account for increased cell death observed in AD brains. This apparently clear participation of p53 in AD could likely be discussed when considering structural aspects of p53. Several studies indicated that p53 could be conformationally altered in AD (Buizza et al., 2012). Thus, unfolded p53 was immunologically characterized in fibroblasts of AD-affected patients at early stages of the disease (Uberti, Lanni, Racchi, Govoni, & Memo, 2008; Uberti et al., 2006). Interestingly, these structural alterations of p53 have functional consequences since unfolded p53 species harbor altered DNA binding properties and impairment of its transcriptional activity (Uberti et al., 2006). This agreed well with a previous report indicating that human skin fibroblasts derived from AD-affected patients show drastic alterations in the molecular effectors of p53-dependent cell death pathways. Notably, several of the classic p53-targeted genes were not activated when these fibroblasts were challenged by several pro-apoptotic triggers (Uberti et al., 2002). This could appear at first sight paradoxical when considering enhanced central apoptosis in AD brains. However, it should be noted that such

101

unfolded p53 species were not detected in the brain and that they were observed very early in the blood of MCI-affected patients (Lanni et al., 2010). Therefore, one can consider that unfolded p53 could be a peripheral and early signature of AD, independent of the cell death observed in the central nervous system. Alternatively, one can also envision a dual involvement of p53, the structural modifications of which could be seen as a protective mechanism at early stages of the disease (Lanni et al., 2010), while the increased levels of residual wild-type biologically active p53 would overcome this protective phenotype and contribute to the neuropathological process later. p53 can also control the expression of proteins involved in AD pathology and indirectly influence the course of the disease without directly affecting proteins considered as its classical targets. Thus, the four proteins involved in the build-up of a biologically active high molecular weight γ-secretase complex (Edbauer, Kaether, Steiner, & Haass, 2004; Kimberly et al., 2003; Takasugi et al., 2003) are regulated or regulate p53 (Checler, Dunys, Pardossi-Piquard, & Alves da Costa, 2010). AD-linked pathogenic mutations in presenilin 1 (Guo et al., 1999) or presenilin-1 depletion (Roperch et al., 1998) activate caspases and thereby, enhance cell vulnerability to apoptotic stimuli by yet poorly understood mechanisms. Conversely, presenilin-2-associated proapoptotic phenotype has been consistently linked to p53. Thus, presenilin 2 and its presenilinase-derived C-terminal counterpart increase p53 expression (Alves da Costa, 2005; da Costa, Masliah, & Checler, 2003; Alves da Costa, Mattson, Ancolio, & Checler, 2003; Alves Da Costa, Paitel, Vincent, & Checler, 2002; Alves da Costa et al., 2002) likely via the release of AICD that acts as a transcriptional activator of p53 promoter transactivation (da Costa, Dunys, et al., 2006; Alves da Costa, Sunyach, et al., 2006). This presenilin-driven modulation of p53 accounts for the functional interplay observed between presenilins and another member of the γ-secretase complex, namely Pen-2 (Dunys et al., 2009). Conversely, we recently documented a direct p53-dependent control of mRNA and protein levels of both presenilins 1 and 2 (Duplan, Sevalle, et al., 2013) in agreement with a previous study showing that p53-dependent cell death was accompanied by a modulation of presenilin-1 expression (Roperch et al., 1998). Finally, the other members of the presenilin-dependent γ-secretase complex, namely nicastrin (Pardossi-Piquard et al., 2009), Aph-1 and Pen-2 (Campbell et al., 2006; Dunys et al., 2007) are able to repress p53 in various cell and animal models. Overall, this complicated scheme (Fig. 1) that links p53 to all the members of the γ-secretase complex and thereby to the production of various βAPP-catabolites such as Aβ and AICD that can both act as p53 modulators (da Costa, Dunys, et al., 2006; Alves da Costa, Sunyach, et al., 2006; Ohyagi et al., 2005) likely accounts, at least in part, to the observation of concomitant enhancements of p53 and cell death in AD. One cannot preclude the possibility that tau protein that accumulates in the intracellular compartments of dying neurons in AD (Lee & Trojanowski, 1999; Spillantini & Goedert, 1998; Yankner, 1996) could contribute to AD-associated apoptosis. Thus, tau hyperphosphorylation is associated with, and even precedes, apoptosis in neuroblastoma cells (Mookherjee & Johnson, 2001). It is noteworthy that p53 was shown to increase tau phosphorylation in human cells (Hooper et al., 2007). Interestingly, the peptidyl prolyl cis/trans isomerase Pin1 could be seen as a potential regulator of p53-dependent phosphorylation of tau as it was shown to increase AP-1-mediated stabilization of p53 and to influence the ability of tau to bind microtubules (Takahashi, Uchida, Shin, Shimazaki, & Uchida, 2008). Furthermore, Pin1 was also reported to modulate the phosphorylation state of βAPP (Takahashi et al., 2008) at threonine 668 (Ma, Pastorino, Zhou, & Lu, 2012), which is recognized as a key residue for βAPPassociated control of synaptic plasticity, memory process and signaling properties (Lombino, Biundo, Tamayev, Arancio, & D'Adamio, 2013). Nevertheless, it remains to be established definitively whether Pin1 expression is modulated at early or late stages of AD pathology since previous studies led to apparent conflicting data probably due to differences in post-mortem samples. Thus, MCI-affected brains appear

102

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

Fig. 1. Functional interplay between p53 and members of the γ-secretase complex. Cross-talk and feed-back between p53 and the four members of the γ-secretase complex (PS1 or PS2, Aph-1, nicastrin (Nct) and Pen-2). Arrows indicate modulations observed following over-expression of the proteins indicated at the starting points of arrows (red arrows, inhibition (−); blue arrows, activation ( )). PS2NNPS1 means that PS2 is dominant over PS1 for p53-dependent proapoptotic phenotype (see Alves da Costa, Paitel, Mattson, et al., 2002).

to display higher levels of Pin1 than controls (Pastorino et al., 2012), while Pin1 expression apparently decreases as long as AD progresses (Wang et al., 2007), indicating that Pin1-linked modulation of p53, if it has a genuine role in AD pathology, should be considered as an early event in the course of the disease. 3. p53 and Parkinson's disease Parkinson's disease (PD) comes in second rank of neurodegenerative diseases in terms of number of affected patients worldwide. The histopathological hallmarks were described about a century ago and it is now admitted that the main brain locus degenerating in the disease is a small area in the ventral midbrain, the pars compacta of the substantia nigra that is mainly composed of dopaminergic neurons (Barzilai & Melamed, 2003). PD-affected substantia nigra harbors intracellular inclusions named Lewy bodies (Goedert, Spillantini, Del Tredici, & Braak, 2013) filled with a protein named α-synuclein (Gwinn-Hardy & Singleton, 2002; Trojanowski & Lee, 1998). Unlike AD, the occurrence of cell death in PD is a matter of consensus (Hartmann & Hirsch, 2001; Lev, Melamed, & Offen, 2003; Maruyama & Naoi, 2002; Nagatsu, 2002; Schapira & Jenner, 2011). Histological and molecular biologic approaches consistently revealed alterations in in situ DNA fragmentation and the number of Tunel-positive cells (Mochizuki, Goto, Mori, & Mizuno, 1996; Tatton & Kish, 1997) as well as in the levels and activities of various cellular effectors participating in cell death control such as caspase-3, caspase-9, cytochrome C and p38 MAPK (Blandini et al., 2006; Junn & Mouradian, 2001). This was confirmed at the level of a single neuronal cell by microarray gene profiling (Simunovic et al., 2009). The main conclusion of these studies was corroborated by a vast number of works performed on toxin-based cellular and animal models that exhibit exacerbated cell death (Gandhi & Wood, 2005). Thus, chronic infusion of both rotenone and MPTP toxins

in murine species mimics the histological features observed in PD (Betarbet et al., 2000; Fornai et al., 2005). It should be noted that protein aggregates often occur in neurodegenerative diseases (Checler, Alves da Costa, Ancolio, Lopez-Perez, & Marambaud, 2000; Huang & Figueiredo-Pereira, 2010) which lead to the impairment of the proteasomal machinery (Bence, Sampat, & Kopito, 2001). Proteasomal dysfunction is often accompanied by cell death as has been evidenced by the pro-apoptotic phenotype triggered by the proteasome inhibitor lactacystin (Fenteany & Schreiber, 1998; Fenteany et al., 1995; Zhou et al., 2010). Interestingly, several studies indicate that indeed, proteasome-mediated degrading process is impaired in PD. Thus, Blandini and colleagues reported on the reduction of the proteasome activity in PD and unraveled an inverse relationship between PD duration and severity with 20S proteasome activity (Blandini et al., 2006). p53 is among the various substrates that could accumulate in response to proteasome defect (Vogelstein, Lane, & Levine, 2000), as it undergoes degradation by Mdm2 in a proteasomedependent manner (Haupt, Maya, Kazaz, & Oren, 1997; Kubbutat, Jones, & Vousden, 1997). In good agreement with the above-cited observations, p53 is affected in PD experimental models and in PD-affected brains. Mogi and Colleagues showed that p53-like immunoreactivity is higher in the caudate nucleus but not in the substantia nigra of PD-affected brains (Mogi, Kondo, Mizuno, & Nagatsu, 2007). Lee et al. showed that a metabolically stable nitrated form of p53 is detected in dopaminergic cells dying in response to nitric oxide stress (Lee, Kim, Choi, Ha, & Kim, 2006) and the dopaminergic toxin 6-hydroxydopamine (Nair, 2006). Again, proteasomal inhibition in dopaminergic cells triggers p53-dependent cell death (Nair, McNaught, Gonzalez-Maeso, Sealfon, & Olanow, 2006). The involvement of p53 in 6-hydroxydopamine-linked dopaminergic cell death was confirmed by genetic inactivation of p53 expression by siRNA (Biswas, Ryu, Park, Malagelada, & Greene, 2005) and by

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

a pharmacological approach aimed at blocking p53 transcriptional activity by pifithrin-α (Komarov et al., 1999a, 1999b). The latter inhibitor proved useful in prolonging the survival of dopaminergic neural grafts in 6-hydroxydopamine-lesioned rat model of PD (Chou, Greig, Reiner, Hoffer, & Wang, 2011). Other toxin-based models of PD corroborated these data. The treatment of PC12 cells with MPP+, the toxic metabolite of MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine), triggers an activation of p53 transcriptional activity (Xu, Cawthon, McCastlain, Duhart, et al., 2005; Xu, Cawthon, McCastlain, Slikker, et al., 2005), and p53 knockout neurons show resistance to MPTP treatment (Trimmer, Smith, Jung, & Bennett, 1996). Recently, p53 was shown to be involved in growth arrest triggered by another dopaminergic neuronal toxin, rotenone, (Wu et al., 2013) and to contribute to rotenoneinduced degeneration in rat neurons and PC12 cells (Goncalves et al., 2011). The postulate of a contribution of p53 to PD pathology was significantly reinforced by data derived from genetic aspects of the disease (Corti, Lesage, & Brice, 2011). PD could be of sporadic or familial origin (Gasser, 1998; Gosal, Ross, & Toft, 2006; Houlden & Singleton, 2012; Morris, 2005; Rosner, Giladi, & Orr-Urtreger, 2008). Familial PD could follow either autosomal dominant or autosomal recessive inheritance (Houlden & Singleton, 2012; Kumar, Lohmann, & Klein, 2012; Trinh & Farrer, 2013). Autosomal recessive cases of PD are essentially due to mutations in three genes encoding parkin (Kitada et al., 1998), PINK-1 (for PTEN-induced kinase-1, Valente et al., 2004) or DJ-1 (Bonifati et al., 2003) while autosomal dominant inherited cases are linked to mutations principally located on genes encoding α-synuclein (ChartierHarlin et al., 2004; Polymeropoulos et al., 1997; Singleton et al., 2003) and leucine-rich repeat kinase 2 (LRRK2) (Paisan-Ruiz et al., 2004; Zimprich et al., 2004). The main functions of the above-cited proteins and the alterations associated with pathogenic mutations have been extensively studied and reviewed. Parkin has been described as an E3-ubiquitin ligase (Mizuno, Hattori, Mori, Suzuki, & Tanaka, 2001; Shimura et al., 2000) but could also act as a transcription factor (da Costa et al., 2009; Duplan, Sevalle, et al., 2013; Sunico et al., in press). PINK-1 is a serine/ threonine protein kinase (Klein & Schlossmacher, 2007), while DJ-1 is mainly described as an oxidative stress sensor and chaperone protein (da Costa, 2007; Olzmann et al., 2003). Concerning α-synuclein, several studies consistently showed that this protein regulates cell death (Albani et al., 2004; da Costa, Ancolio, & Checler, 2000; Burke, 1999; Chandra, Gallardo, Fernàndez-Chàcon, Schlüter, & Südhof, 2005; Choi et al., 2006; Suh & Checler, 2002) and many additional functions (Alves da Costa, 2003; Surguchov, 2008), while LRKK2 is a kinase that controls the mitogen-activated protein kinase-related pathway (Gloeckner, Schumacher, Boldt, & Ueffing, 2009). Overall, the functions ascribed to the proteins involved in familial PD are closely linked to the control of protein degradation and cell death. In addition, there are several experimental clues linking these proteins to p53 (da Costa & Checler, 2011). α-Synuclein protects neuronal cells challenged with pro-apoptotic effectors (da Costa et al., 2000) by lowering p53 expression and transcriptional activity in cells (Alves Da Costa, Paitel, Vincent, et al., 2002), a phenotype that is abrogated by the A53T pathogenic mutation (da Costa et al., 2000). These observations are supported by the demonstration that p53 accumulates in mitochondria in A53T-expressing transgenic mice (Martin et al., 2006). There exists an apparent paradox between the fact that α-synuclein appears to be toxic rather than protective at high concentrations in dopaminergic cell lines and in tissues. We showed that 6-hydroxydopamine, (6OHDA) which is a natural pro-oxidant and toxic dopamine derivative, triggers both α-synuclein aggregation and proteasomal inhibition and blocks α-synucleinmediated protection against staurosporine-induced caspase-3 activation (da Costa, Dunys, et al., 2006; Alves Da Costa, Paitel, Vincent, et al., 2002; Alves da Costa, Sunyach, et al., 2006). The 6OHDAassociated blockade of α-synuclein function could be prevented by

103

anti-oxidant (da Costa, Dunys, et al., 2006) and reversed by overexpression of the α-synuclein congener, β-synuclein (da Costa, Masliah, et al., 2003; Alves da Costa, Mattson, et al., 2003). Interestingly, the latter protein is protective, lowers staurosporine-induced p53 activation and restores α-synuclein's ability to repress p53 likely by modulating Mdm2mediated p53 degradation (da Costa, Masliah, et al., 2003; Alves da Costa, Mattson, et al., 2003). This agrees well with a study showing that β-synuclein could control Akt that acts as an upstream effector of the Mdm2-linked cascade ultimately leading to p53 phosphorylation and nuclear degradation (Hashimoto et al., 2004). Finally, synphilin-1, which is a component of Lewy bodies (Wakabayashi et al., 2000) and interacts with α-synuclein physically and functionally (Engelender et al., 1999; Smith et al., 2010), undergoes a caspase-3 mediated cleavage giving rise to a C-terminal product able to lower p53 at the level of the promoter, mRNA and protein expression (Giaime et al., 2006). Several lines of evidence indicate that proteins involved in recessive cases of PD also interact functionally with p53. The most convincing evidence came from our recent report on a novel function of parkin as a transcription factor able to repress p53 (da Costa & Checler, 2010). This phenotype was independent of parkin-linked ubiquitin ligase activity and was abrogated by pathogenic mutations (da Costa et al., 2009). Indeed, parkin physically interacts via its RING1 domain with the p53 promoter region (da Costa et al., 2009). Interestingly, parkinassociated control of p53 could trigger downstream events, and notably, could control DJ-1 by a molecular cascade involving first p53 followed by another transcription factor involved in cellular response to endoplasmic reticular stress, XBP-1S (Duplan, Giaime, et al., 2013). Obviously, the direct or indirect participation of the ubiquitin-ligase-dependent function of parkin in the control of the degradation process of p53 could not be totally precluded. In this context, it is interesting to note that several parkin substrates are indeed upstream regulators of p53-dependent pathways. Also interesting was the demonstration that a cytosolic protein displaying high homology with parkin C-terminus (thus referred to as PARC, parkin-like ubiquitin ligase) controls the subcellular localization of p53 and p53-mediated apoptosis even in unstressed cells (Nikolaev, Li, Puskas, Qin, & Gu, 2003). Moreover, p53 trafficking could be proteolytically regulated by calpain-mediated PARC degradation (Woo, Xue, Liu, McBride, & Tsang, 2012). A functional crosstalk between p53 and parkin was also recently documented. Accordingly, p53 directly regulates the transcription of parkin (Viotti et al., in press; Zhang et al., 2011) and could account for the interplay taking place in brain cancers (see below). The functional interaction of DJ-1 with p53 and notably, its ability to repress p53, has been reported by two groups (da Costa, 2007; Ottolini, Cali, Negro, & Brini, 2013) although one study showed that DJ-1 could positively regulate p53 transcriptional activity (Shinbo, Taira, Niki, Iguchi-Ariga, & Ariga, 2005). More generally, DJ-1 was found to lower p53 expression and activity in both mammalian cells (Fan, Ren, Fei, et al., 2008; Fan, Ren, Jia, et al., 2008; Giaime et al., 2010) and zebrafish (Bretaud, Allen, Ingham, & Bandmann, 2007). Thus, mutations that render DJ-1 resistant to caspase-6-mediated proteolysis abolish the production of a DJ-1 C-terminal fragment that is able to repress p53 at both transcriptional and post-transcriptional levels (Giaime et al., 2010). This was corroborated by knockout (Bretaud et al., 2007) and over-expression approaches (Fan, Ren, Fei, et al., 2008; Fan, Ren, Jia, et al., 2008) showing that DJ-1 could also control the levels of the p53 transcriptional target, Bax. Again, a feed-back mechanism by which p53 apparently controls DJ-1 was demonstrated by a proteomic study designed to identify new p53 phosphorylated targets in HCT116 cancer cells (Rahman-Roblick et al., 2008). Also interesting was the observation that caspase-6 could yield DJ-1 N- and C-terminal fragments with distinct functions as it is the case for cellular prion protein PrPc (GuillotSestier & Checler, 2012a) since the caspase-6-derived N-terminal product triggers apoptosis by increasing reactive oxygen species (Robert et al., 2012).

104

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

Few reports link LRRK2 and p53. It should be noted, however, that LRRK2 phosphorylates various kinases of the mitogen-activated protein kinase cascades (Gloeckner et al., 2009). Since some of these kinases are upstream effectors of p38 MAP Kinase (p38 MAPK) that phosphorylates p53 (Perfettini et al., 2005), one could speculate about a LRRK2associated and p38 MAPK-linked modulation of the p53 cellular pathways, although this requires final demonstration. Supporting this hypothesis, it was recently demonstrated that LRRK2-deficiency apparently triggers enhanced expression of phosphorylated p53 in human neural progenitor cells, indicating that LRRK2 could indeed exert a p53-dependent protective role in dopaminergic neurons (Milosevic et al., 2009). Finally, there is no clear evidence for the control of p53 by PINK-1. However, the simple assertion that PINK-1 induces PTEN, which undergoes p53-mediated transactivation (Stambolic et al., 2001), makes plausible a direct or indirect functional link between p53 and PINK-1. Overall, both histological and biological evidence concur to propose p53 as a main contributor of exacerbated cell death taking place in PD (Fig. 2).

4. p53 and prions Prion diseases have been at the origin of a revolutionary concept stating that pathogenic proteins could be, per se, considered as infectious triggers that can accumulate, spread and propagate without contribution of any nucleic acid (Prusiner, 1998). In animals and humans, prion-related pathologies mainly correspond to either scrapie, bovine spongiform encephalopathies, Creutzfeldt–Jakob and Gerstmann– Sträussler–Scheinker diseases (Aguzzi & Polymenidou, 2004; Collinge, 2005; Weissmann, Enari, Klöhn, Rossi, & Flechsig, 2002). Pathogenic prion (PrPsc) is a proteolytically resistant protein that results from the biophysical conversion of a physiological and ubiquitous protein called cellular prion (PrPc). The latter congener is essential for the infectious process since several studies indicate that cells or animals devoid of PrPc fully resist infection by PrPsc-enriched preparations (Brandner et al., 1996; Büeler et al., 1993; Sailer, Bueler, Fischer, Aguzzi, & Weissmann, 1994).

Fig. 2. Functional interplay between p53 and various proteins responsible for familial Parkinson's disease. Cross-talk and feed-back between p53 and proteins involved in autosomal dominant (α-synuclein and LRRK2) and autosomal recessive (parkin and DJ-1) forms of Parkinson's disease. The interplay between p53 and proteins indirectly linked to PD (synphilin-1 and HtrA2/Omi) as well as the influence of several toxins linked to PD-like models (chemicals) are also indicated. Arrows indicate modulations observed following over-expression of the proteins indicated at the starting points of arrows or after toxin treatments (red arrows, inhibition (−); blue arrows, activation ( )).

Several studies indicated that scrapie-mediated infection of cells, animals and humans triggers cell death markers enhancement. Thus, various regions of Creutzfeldt–Jakob-affected brains show increased expression of caspases-3 and -9, Bax and the Poly-ADP-ribose polymerase PARP (Kovacs & Budka, 2010). Furthermore, immunohistochemical analysis of natural scrapie-infected sheep indicates that Bax co-distributes with PrPsc deposition (Lyahyai et al., 2007). This was corroborated with the demonstration that active caspase-3 expression was increased in mice inoculated with infectious scrapie-enriched preparations (Drew et al., 2011). Finally, a scrapie-infected neuronal cell line shows morphological and biochemical alterations reminiscent of apoptotic hallmarks (Schatzl et al., 1997). The above-described set of observations did not delineate the relative contribution of PrPc vs PrPsc to the apoptotic process. Several studies based on the modulation of cellular PrPc levels or on the use of recombinant PrPc support the view that this protein could indeed harbor a proapoptotic potential. Thus over-expression of PrP23–230 reduces viability and increases apoptosis of HeLa cells (Li et al., 2011). This agreed well with the demonstration that the reduction of endogenous PrPc by RNA silencing indeed lowers T lymphoblastoid CEM cell responsiveness to Fas-induced apoptosis (Mattei et al., 2011). This effect appears to be linked to a PrPc redistribution at the ER-mitochondrial contacts that ultimately lead to an alteration of mitochondrial membrane potentials and to the release of cytochrome c (Sorice et al., 2012). Interestingly, an unbiased search for PrP-related toxic species led to the identification of an alpha helical-enriched form of PrPc that triggers apoptosis and autophagy (Zhou, Ottenberg, Sferrazza, & Lasmezas, 2012). It should also be noted that several studies aimed at delineating PrPc-related toxicity used a 21-amino-acid long peptide referred to as PrPc106–126 that mimics some of the PrPsc-associated phenotypes (Brown, 2000; Singh et al., 2002) and that was shown to be neurotoxic both in vitro and in vivo. PrPc106–126 increases DNA fragmentation, the number of Tunel-positive cells, p38 MAPK and active form of caspase-3 (Jeong et al., 2011). Consistent with the idea that PrPc could undergo proteolysis by the proteasome (Amici et al., 2010; Cecarini et al., 2010; Wang, Wang, Sy, & Ma, 2005), it was demonstrated that proteasomal inhibition potentiated PrPc-associated caspase-3 activation (Paitel, Alves da Costa, Vilette, Grassi, & Checler, 2002). It should be noted that a study indicated that PrPc could lower apoptotic response in a bacterial model of infection (Ding et al., 2013). In a more general cellular context, PrPc could behave as a pro-survival effector instead of triggering cell death (Linden et al., 2008). Particularly interesting is a series of studies showing that PrPc physically interacts with the STI1 (Lopes et al., 2005; Zanata et al., 2002), thereby initiating a signaling cascade (Chiarini et al., 2002) ultimately leading to PKA and ERK1 activation (Caetano et al., 2008). These studies, which could appear at first sight contradictory with the PrPc-related proapoptotic hypothesis, could likely be explained by pro-apoptotic challenges, cell specificity and proteolytic events. Thus, ER-stressinduced caspase activation and apoptosis could be enhanced by PrPc while conversely, PrP c protects cells against H 2 O2 treatment (Anantharam et al., 2008). PrPc-associated resistance to apoptosis could be directly linked to its glycosylation state (Yap & Say, 2011) and cell type (Yap & Say, 2012). More recently, several studies indicated that proteolytic events could participate in the control of PrPc function (Guillot-Sestier & Checler, 2012b). Thus, PrPc is cleaved within the 106–126 toxic sequence by ADAM proteases belonging to the class of disintegrins (Vincent et al., 2000; Vincent et al., 2001). One of them also contributes to the shedding of PrPc, thereby releasing the extra domain within the extracellular space (Altmeppen et al., 2011; Taylor et al., 2009). It is of interest to note that the same set of secretases participates in the α-cleavage of both βAPP and PrPc, and that the kinase PDK-1 contributes to a functional cross-talk between βAPP and PrPc via the phosphorylation of ADAM17 and its depletion from the membrane (Checler, 2013; Pietri et al., 2013).

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

Two distinct lines of evidence indicate that the N-terminal moiety of PrPc is important for its function/toxicity. First, the over-expression of PrPc deleted in its N-terminal region of 80 to 100 amino-acid long segments in PrPc knockout mice leads to a severe ataxia and neuronal loss (Shmerling et al., 1998). Accordingly, neuronal treatment with recombinant human PrPc-90–231 triggers cytosolic accumulation of aggregated fragment that is accompanied by neuronal apoptosis (Thellung et al., 2011). Second, The N1 fragment liberated by ADAM proteases is neuroprotective in vitro and in vivo (Guillot-Sestier, Sunyach, Druon, Scarzello, & Checler, 2009). Therefore, one can envision that α-secretase-mediated cleavage of PrPc is a central mechanism conditioning the overall phenotype associated with intact PrPc levels, and that the cell specific enzymatic machinery targeting PrPc could directly drive the levels of N1/intact PrPc and thereby, could explain the apparently discrepant pro-apoptotic/pro-survival phenotypes observed in the above-described studies. Several lines of evidence indicate that PrPc and its fragments could modulate p53 expression and activity (Guillot-Sestier & Checler, 2012a). First, it was shown that PrPc over-expression sensitized neurons to pro-apoptotic triggers (Paitel, Fahraeus, & Checler, 2003) and increased p53 expression at a post-transcriptional level via the modulation of Mdm2 (Paitel et al., 2003). In agreement, primary cultured neurons derived from PrPc-depleted mice display reduced responsiveness to staurosporine-induced p53 activation (Paitel et al., 2004). This phenotype required PrPc endocytosis as was demonstrated by various mutational and cell biological approaches (Sunyach & Checler, 2005). We showed that the C-terminal counterpart of N1 named C1 yielded by α-secretase cleavage potentiated p53 activation, while another fragment (C2) recovered in PrPsc-affected brains (Chen et al., 1995) and resulting from a downstream cleavage by a yet unknown protease remained biologically inert (Sunyach, Cisse, da Costa, Vincent, & Checler, 2007). Interestingly, N1 protects cells from C1- and PrPcinduced activation of p53 (Guillot-Sestier et al., 2009). This indicates first that ADAM-mediated proteolysis of PrPc could produce fragments with distinct functions (Guillot-Sestier & Checler, 2012a) and second, that the proteolytic events yielding PrPc fragments and regulating the PrPc levels could indeed directly modulate the function of PrPc and its catabolites. Additionally, the 106–126 PrPc fragment increases p53 and lowers the expression of the anti-apoptotic protein Bcl-2 (Jeong et al., 2011). Furthermore, DNA microarray analyses indicated that PrPc triggers significant up-regulation of p53 in PrP-mediated myopathy (Liang et al., 2009). As it has been described for Alzheimer's and Parkinson's diseases, there exists a feed-back control by which p53 controls PrPc levels. Thus, p53 could directly control PrPc promoter transactivation and a functional p53 responsive element has been identified in the promoter region of PrPc (Vincent, Sunyach, Orzechowski, St George-Hyslop, & Checler, 2009). 5. p53 and brain cancers Cancer is a huge public health problem, the prevalence of which is also associated with aging. In contrast to neurodegenerative diseases, the tumorigenesis process is associated with increased proliferation concomitant with drastic reduction of apoptotic responses. Of most interest, several epidemiological studies have found an inverse correlation between the risk of developing several neurodegenerative disorders and cancer (Becker, Brobert, Johansson, Jick, & Meier, 2010; Musicco et al., 2013; Roe, Behrens, Xiong, Miller, & Morris, 2005). This negative relationship raises the question of whether despite being phenotypically divergent, these multifactorial chronic pathologies could share common protein effectors, the function of which would be either altered or exacerbated. Given the major role of p53 in neurodegenerative diseases and brain cancer, one could easily envision p53 as a serious molecular candidate. Brain tumors of glial origin are among the top ten most deadly cancers. Gliomas represent ~80% of all malignant brain tumors (Ohgaki &

105

Kleihues, 2007; Wrensch, Minn, Chew, Bondy, & Berger, 2002) and can be classified according to cytological type and grade (Furnari et al., 2007). The most up-to-date classification of gliomas recognized by the World Health Organization (WHO) combines cytological origin and malignancy grade criteria (Louis et al., 2007). Accordingly, gliomas originating from astrocytes, oligodendrocytes or a mix of different glial cell types are respectively named astrocytomas, oligodendrogliomas and mixed gliomas. This cytology-based glioma classification is then refined according to malignancy severity in either low-grade (WHO grade II) or high grade infiltrating gliomas (WHO grades III and IV) (Lu et al., 2000). The low-grade astrocytomas (grade II) are characterized by a faint increase in cellular density in comparison with white substance, the presence of cytological atypias, usually lack of mitotic activity and by the absence of vascular proliferation and necrosis. The anaplastic astrocytoma of grade III is characterized by a more pronounced augmentation of cell density, number of nuclear atypias and mitosis. The astrocytomas of grade IV, better known as glioblastoma multiforme (GBM), present all the characteristics of grade III astrocytomas in addition to a vascular proliferation and necrosis. GBM may be further classified as primary, which develops rapidly de novo, or secondary, which develops through progression from a low-grade tumor. GBM is the most aggressive malignant tumors and is associated with 50% of all gliomas (Ohgaki & Kleihues, 2005a, 2007). Grade I gliomas are mainly represented by pilocytic astrocytomas and represent 50% of gliomas affecting preferentially young children (Villarejo, de Diego, & de la Riva, 2008). They are considered benign and in contrast to infiltrating gliomas, they have a good prognosis. The oligodendrogliomas may be of grade II or III and, even if these two grades present the same cytological characteristics, the WHO classification highlights that an increased mitotic activity and a possible presence of necrosis indicate a progression to a grade III. Moreover, this type of glioma may be distinguished by the strong labeling of the oligo2 biomarker (Colin et al., 2007; Lu et al., 2001; Maries, Dass, Collier, Kordower, & Steece-Collier, 2003). The mixed gliomas, given the difficulty to accurately estimate their composition in astrocytes versus oligodendrocytes, are much more difficult to diagnose. Nevertheless, the WHO recommends the classification of mixed gliomas in grades II to III depending on the malignancy criteria previously described (cellular density, nuclear atypias, mitotic index and presence of vascularization/necrosis). Tumors of distinct cytological type respond differently to the available clinical treatments. Thus, oligodendrogliomas are much more sensitive to chemotherapy than astrocytomas (Behin, Hoang-Xuan, Carpentier, & Delattre, 2003; Cairncross & Macdonald, 1988). The limits of the WHO gliomas classification, based on subjective morphological characteristics, are outlined by the rather large discordance rate (20– 60%) for the diagnosis made by neuropathologists (Coons, Johnson, Scheithauer, Yates, & Pearl, 1997; Mittler, Walters, & Stopa, 1996). This observation drove a huge research effort aimed at delineating key glioma molecular determinants to better understand the causes of glioma genesis and to identify reliable glioma biomarkers to be used in clinics. The search for molecular effectors of gliomas rapidly converged to the demonstration of a key role for p53 in their etiology. p53 is the most frequently mutated gene product in human cancers and especially in nervous system-associated cancers as indicated by the recently released IARC p53 database (Petitjean et al., 2007). According to the R16 IARC TP53 database (R16, November, 2012), the p53 mutation mostly represented in brain tumors is C:G → A:T at CpG sites (39%), located mainly in exons associated with its DNA binding properties (e5–e8), and preferentially affect codons 175, 248 and 273. Moreover, 90% of these mutations are missense mutations that inactivate p53 transcriptional activity and are very deleterious. The prevalence of p53 mutations varies according to the cytological type of the tumor. Thus, this prevalence is high in secondary GBM (70%), moderate in mixed gliomas (40%) and low-grade astrocytomas (50%), and low in pure oligodendrogliomas (Hagel et al., 1996; Kim et al.,

106

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

2010; Louis et al., 2007; Maintz et al., 1997; Mendrysa, Ghassemifar, & Malek, 2011; Ohgaki & Kleihues, 2007, 2011; Watanabe et al., 1997). Moreover, both frequencies of p53 mutations and transcriptional activity are correlated with tumor grade in astrocytomas. Interestingly, p53 mutations are already present in low-grade astrocytomas suggesting that this molecular alteration is an early event in secondary glioblastoma genesis (Mendrysa et al., 2011; Stander, Peraud, Leroch, & Kreth, 2004). This observation also suggests that the evaluation of p53 mutation status may have a prognostic value. Accordingly, even if it is still a matter of debate, the presence of p53 mutations seems to be associated with a poor prognostic in low grade astrocytomas (Kim et al., 2010; Stander et al., 2004) and is irrelevant in GBM (Bleeker, Molenaar, & Leenstra, 2012). Interestingly, several studies have shown a link between p53 mutations and p53 protein accumulation suggesting that an accumulation of p53 would be predictive of p53 mutations (Nozaki et al., 1999; Watanabe et al., 1997). Although the predictive value of p53 mutations due to its protein accumulation is still questioned (Levidou et al., 2010; Stander et al., 2004), we have recently shown that p53 protein levels correlate with tumor grade and p53 mutation frequency in oligodendrogliomas and mixed gliomas (Viotti et al., in press), in agreement with several works based on immunohistochemical approaches (Fulci, Ishii, & Van Meir, 1998; Watanabe et al., 1997; Weller et al., 2009). It is worth noting that, even if the prevalence of p53 mutations is predominantly associated with secondary (65%) compared to primary (28%) GBM (Ohgaki & Kleihues, 2005b; Ohgaki et al., 2004), more recent studies, that included additional sequencing of TP53, revealed that mutations are prevalent in primary GBMs as well (Parsons et al., 2008; Zheng et al., 2008). Not only the frequency but also the distribution of p53 mutations in primary versus secondary GBM is distinct. In secondary GBM, most of p53 mutations (57%) will occur in the hot-spot codons 248 and 273, whereas in primary GBM they appear equally distributed (Ohgaki et al., 2004). As described in previous sections, p53 is a transcription factor that responds to several tumor-related stress conditions by halting the cell cycle in the G1 phase in order to allow DNA repair or increase cell death response (Vousden & Lu, 2002), thereby counteracting putative tumor development. It was not surprising, therefore, to note a tight correlation between cell cycle deregulation and inactivation of p53 by mutations in sporadic glioma genesis. It has been shown that the expression of p21, a p53 transcriptional target and key piece of p53associated cell cycle regulation, is reduced in high-grade astrocytomas (Kirla, Haapasalo, Kalimo, & Salminen, 2003). Moreover, in addition to mutation-linked p53 inactivation, other mechanisms may contribute to loss of function of p53 in gliomas. For example, MDM-2, an E3ligase implicated in p53 targeting to proteasomal degradation (Fakharzadeh, Trusko, & George, 1991; Momand, Zambetti, Olson, George, & Levine, 1992; Oliner et al., 1993), p14ARF, an MDM-2 blocker (Kamijo et al., 1998; Pomerantz et al., 1998) and MDM-4, a negative regulator of p53 transcriptional activity, were shown to be genetically altered in gliomas. MDM2 (14%) and MDM4 (7%) genes are amplified whereas p14ARF is homozygously deleted or mutated in 49% of the cases (“Comprehensive genomic characterization defines human glioblastoma genes and core pathways”, Cancer Genome Atlas Research Network, 2008; Toledo & Wahl, 2007). Interestingly, even if the mutation of p53 is a rare event in pure oligodendrogliomas, the overall contribution of p53 mutations, p14ARFdeletions and MDM2/MDM4 amplifications correlates to tumor grade in this cytological glioma type (Watanabe et al., 2001). Finally, several carcinomas may be associated with inactivation of wild-type p53 due to its sequestration into the cytoplasm (Moll, LaQuaglia, Benard, & Riou, 1995; Moll, Riou, & Levine, 1992). This possibility is specially highlighted by the fact that despite a low p53 mutation rate, primary GBM (Watanabe et al., 1997) manifests a huge accumulation of p53 (Von Eckardstein et al., 1997). Interestingly, it has been shown that p53 mutations are associated with nuclear p53 expression, while wild-type p53 is associated with cytoplasmic expression (Sembritzki, Hagel, Lamszus, Deppert, & Bohn, 2002).

6. Pharmacological and gene therapy strategies targeting p53 As indicated above, p53 plays a key role in both neurodegenerative diseases and brain cancer and remains a premier target of strategies aimed at interfering with either exacerbated cell death or defective apoptosis. Theoretically, in brain tumors where wild-type and mutated p53 are inactivated by either accumulation/aggregation or mutation, respectively, the main approach actually envisioned consists of preventing p53 degradation or restoring its transcriptional activity by means of chaperone molecules allowing the recovery of a structurally and biologically active p53. Conversely, in neurodegenerative diseases where one must counteract enhanced p53-dependent cell death, the main strategy aims at developing pharmacological blockers of p53 transcriptional activity or genetic depletion of p53 by siRNA approaches. In the case of both neurodegenerative diseases and cancer, additional therapeutic strategies could consist of interacting with upstream effectors of the p53-dependent cellular pathways or interfering with downstream effectors of p53 such as microRNAs. p53 can undergo ubiquitination (Lee & Gu, 2010) and, accordingly, its levels are tightly regulated by proteasomal degradation (Scheffner, 1998). Overall, targeting the ubiquitin-dependent proteasomal system has been envisioned as a means to activate the p53 pathway in cancer therapy (Allende-Vega & Saville, 2010; Devine & Dai, 2013). More precisely, it was shown about 20 years ago that p53 physically and functionally interacts with several E3-ligases of the Mdm-X (murine double minute gene) family. Thus, MDM-2 was identified as an oncogenic factor able to inhibit p53 transcriptional activity (Momand et al., 1992) by tightly interacting with its acidic activation domain (Oliner et al., 1993). Thus, Mdm-2 blocks p53-linked cell growth arrest and pro-apoptotic phenotype (Kubbutat et al., 1997). Haupt and colleague indeed demonstrated that MDM-2-mediated inhibition of p53 function resulted from the enhancement of its proteasomal inactivation (Haupt et al., 1997). An additional member of the MDM-X family of ligases, MDM-4, also contributes to the stabilization of p53 (Toledo & Wahl, 2007). Prior to proteasomal degradation, Mdm-X could also behave as a p53 molecular chaperone inducing structural changes and inactivation (Sasaki, Nie, & Maki, 2007; Wawrzynow, Zylicz, Wallace, Hupp, & Zylicz, 2007). Thus, via multi-step mechanisms, Mdm-2 and Mdm-4 act as p53 antagonists and, therefore, have been chemically targeted in order to restore p53-mediated loss of function and enhance cell death in cancer (Maslon & Hupp, 2010); Brown, Cheok, Verma, & Lane, 2011). Chaperoning can translate into various phenotypes. As mentioned above, it can lead to functional inactivation by inducing conformational changes in proteins. Conversely, chaperone molecules can be designed to fix a deleterious conformation and restore native structure and function of denatured proteins. This strategy has been envisioned for p53 since various cancer-related mutations can disrupt its conformation and activity. Natural proteins and chemically designed compounds have been developed with the aim of restoring conformationally native and bioactive p53 species (Bossi & Sacchi, 2007; Brown et al., 2011; Bullock & Fersht, 2001; Friedler, Veprintesev, Hansson, & Fersht, 2003). For example, heat shock protein 90 (HSP90) stabilizes mutant p53 (Blagosklonny, Toretsky, Bohen, & Neckers, 1996; Muller, Ceskova, & Vojtesek, 2005), and its natural inhibitor geldanamycin abolishes HSP90 activity and thereby triggers mutant p53 destabilization and proteasomal degradation (Blagosklonny et al., 1996). Interestingly, geldanamycin apparently affects mutant but not wild-type p53 (Dasgupta & Momand, 1997). Numerous synthetic compounds have been shown to potently restore the defective function of mutant p53 (Brown et al., 2011; Maslon & Hupp, 2010). Examples include a nonapeptide CDB3, which has been shown to help maintain the native conformation of newly synthesized mutant p53 and can restore the DNA-binding ability of mutant p53 (Friedler et al., 2002). More recently, a small molecule physically interacting with the DNA binding domain of p53 was shown to restore the normal function of various p53 mutants and, interestingly also to prevent the ubiquitination and proteasomal

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

degradation of wild-type p53 by HDM-2 (the human counterpart of Mdm-2) (Demma et al., 2010). In neurodegenerative diseases, putative neuroprotective therapeutic approaches include consideration of the hypothesis that p53 plays a key role in the observed exacerbated cell death. Accordingly, the main strategy consists of the design and study of inhibitors of p53 transcriptional activity, in agreement with the fact that this function controls the levels of its pro-apoptotic transcriptional targets (Bourdon et al., 2003; Brown, Lain, Verma, Fersht, & Lane, 2009; Levine & Oren, 2009). Komarov and colleagues described the first potent inhibitor of p53 transcriptional activity and apoptosis named pifithrin-α (Komarov et al., 1999a, b). Its very first utilization was not linked to neurodegenerative disease treatment but to the cancer field. In mice, pifithrin-α prevents the side effects of deleterious apoptosis associated with chemotherapy and radiotherapy in healthy cells surrounding the treated tumor (Gudkov & Komarova, 2005; Komarov et al., 1999a, 1999b). Other inhibitors of p53 have been described as well (Zhu et al., 2002). Most of these p53 inactivators have been validated in AD and PD models. MPTP administration mimics PD-related histological hallmarks and increases DNA fragmentation and p53/Bax signaling (Eberhardt & Schulz, 2003). Intraperitoneal injection of pifithrin-α in MPTP-treated mice reduced nigrostriatal dopaminergic cell damage and rescued motor impairment (Duan et al., 2002). Interestingly, daily administration of pifithrin-α for five days also favors the survival of grafted dopaminergic cells in the striatum of the rats 6-OHDA-induced lesion model of PD. In this animal model, pifithrin-α reinforced the graft-associated improvement of cognitive deficit, reduced the number of Tunel-positive cells and increased tyrosine hydroxylase expression in dopaminergic grafts. Interestingly, this protective phenotype was not observed after grafting of cortical tissue transplants (Chou et al., 2011). In experimental models of AD, pifithrin-α has also proved effective in blocking various Aβ-related toxic phenotypes. This inhibitor prevents Aβ-induced cell death, stabilizes mitochondrial function and reduces caspase activation in hippocampal cell cultures (Culmsee & Landshamer, 2006). Furthermore, in cortical neurons, Aβ increases p53 and its target Bax and perturbs lysosomal function. All these effects are rescued by pifithrin-α and p53 siRNA approach (Fogarty et al., 2010). Finally, it has been reported that Aβ could increase apoptosis in microglia. In AD afflicted brains, the number of apoptotic microglial cells increases (Lassmann et al., 1995; Yang et al., 1998) and this is accompanied by enhanced expression of p53 (Davenport et al., 2010; Kitamura et al., 1997). Both Aβ-related phenotypes are reduced by pifithrin-α (Davenport et al., 2010). Other p53-directed therapeutic strategies applicable to both neurodegenerative diseases and cancer could consist of increasing/repressing its expression by genetic approaches or by interfering with either upstream effectors of the p53 pathway or downstream targets of this tumor suppressor. Targeting the p53 gene was envisioned about twenty years ago (Fulci et al., 1998). Adenoviral expression of p53 revealed the effectiveness of such approach in cells (Fujiwara et al., 1994) without apparent drastic side effects even in normal cells (Bossi et al., 2004; Scardigli et al., 1997). Although viral expression of p53 led to contrasting results in human trials (Moon, Oh, & Roth, 2003; Zeimet and Marth, 2003), the clinical interest in such an approach remains, particularly when combined with conventional chemo- or radio-therapies (Bossi et al., 2004). Conversely, siRNA-linked silencing of mutant p53 reduces cell proliferation and therefore could appear as a promising strategy for both bladder (Zhu et al., 2013) and breast cancer therapy (Braicu, Pileczki, Irimie, & Berindan-Neagoe, 2013). Numerous and interconnected molecular pathways control p53 cellular homeostasis, subcellular localization and transcriptional/ post-transcriptional events that drive several fundamental p53dependent cellular functions via the modulation of various p53 transcriptional targets. These effectors and targets often interplay and contribute to positive or negative regulatory loops (Harris & Levine, 2005) building a very complex cellular network. It is,

107

therefore, difficult to envision a selective pharmacological modulation of a given target, particularly because compensatory mechanisms could occur through molecular intermediates that could contribute to additional p53-independent phenotypes. Another opportunity for a more selective approach came from the recent interest in microRNA research (Chen & Rajewsky, 2007; Kloosterman & Plasterk, 2006; Valencia-Sanchez, Liu, Hannon, & Parker, 2006) and the demonstration that p53 indeed can control microRNA levels (Hermeking, 2007). Several reports concomitantly and consistently documented microRNA of the miR-34 family as p53 targets (Bommer et al., 2007; Chang et al., 2007; Corney, Flesken-Nikitin, Godwin, Wang, & Nikitin, 2007; He et al., 2007; Raver-Shapira et al., 2007). Interestingly, miR-34 appears inactivated in cancer cells (Hermeking, 2010). Therefore, restoring miR-34 expression could prove useful to circumscribe chemotherapy resistance in cancer cells. Whether inactivating miR-34 could allow reduction of exacerbated p53-dependent cell death in neurodegenerative disease remains to be examined. 7. Concluding remarks Since the initial clues for the presence of a new protein modulated by SV40 cell infection, more than 68,000 papers have been published on p53, and still 3500 in 2013. This clearly underlines the huge interest in this protein now considered as a main molecular factor defective in carcinogenesis. The key role of p53 now spreads over many other fundamental processes, and the control of its homeostasis by transcriptional and post-transcriptional regulatory mechanisms represents central means to maintain normal cellular physiology. More recently, it appeared that p53 dysfunction could contribute to exacerbated cell death observed in various neurodegenerative diseases. Thus, either increasing/restoring normal p53 function in cancers or reducing p53dependent cell death in neurodegenerative diseases could be seen as tempting therapeutic strategies, and multiple approaches described above could be envisioned. The real challenge remains to modulate p53 function in pathological contexts without triggering serious putative deleterious effects due to the impairment of physiological functions of this pleiotropic protein. Progresses in the design of more selective, potent and bio-available compounds or improvements in alternative genetic approaches ultimately leading to the modulation of p53 function can be reasonably but optimistically envisioned in a rather close future with the aim to prevent, arrest or at least slow down the onset and/ or progression of tumorigenesis and neurodegenerative processes. Conflict of interest statement The authors declare that there are no conflicts of interest. Acknowledgments This work has been developed and supported through the LABEX “Excellence Laboratory Program Investment for the Future” Development of Innovative Strategies for Transdisciplinary Approach to Alzheimer's disease. We wish to thank the Conseil Général des Alpes Maritimes for its constant support. CAC is a recipient of a Hospital Contract for translational research (CHRT) between INSERM and the Hospices civils de Lyon. References Aguzzi, A., & Polymenidou, M. (2004). Mammalian prion biology: one century of evolving concepts. Cell 116, 1–20. Albani, D., Peverelli, E., Rametta, R., Batelli, S., Veschini, L., Negro, A., et al. (2004). Protective effect of TAT-delivered a-synuclein: relevance of the C-terminal domain and involvement of HSP70. FASEB J 18, 1713–1715. Allende-Vega, N., & Saville, M. K. (2010). Targeting the ubiquitin–proteasome system to activate wild-type p53 for cancer therapy. Semin Cancer Biol 20, 29–39.

108

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

Altmeppen, H. C., Prox, J., Puig, B., Kluth, M.A., Bernreuther, C., Thurm, D., et al. (2011). Lack of a-disintegrin-and-metalloproteinase ADAM10 leads to intracellular accumulation and loss of shedding of the cellular prion protein in vivo. Mol Neurodegener 6, 36. Alves da Costa, C. (2003). Recent advances on a-synuclein cell biology: functions and dysfunctions. Curr Mol Med 3, 17–24. Alves da Costa, C. (2005). Recent insights on the pro-apoptotic phenotype elicited by presenilin 2 and its caspase and presenilinase-derived fragments. Curr Alzheimer Res 2, 507–514. Alves da Costa, C., Mattson, M. P., Ancolio, K., & Checler, F. (2003). The C-terminal fragment of presenilin 2 triggers p53-mediated staurosporine-induced apoptosis, a function independent of the presenilinase-derived N-terminal counterpart. J Biol Chem 278, 12064–12069. Alves da Costa, C., Paitel, E., Mattson, M. P., Amson, R., Telerman, A., Ancolio, K., et al. (2002). Wild-type and mutated presenilins 2 trigger p53-dependent apoptosis and down-regulate presenilin 1 expression in HEK293 human cells and in murine neurons. Proc Natl Acad Sci U S A 99, 4043–4048. Alves Da Costa, C., Paitel, E., Vincent, B., & Checler, F. (2002). Alpha-synuclein lowers p53-dependent apoptotic response of neuronal cells. Abolishment by 6-hydroxydopamine and implication for Parkinson's disease. J Biol Chem 277, 50980–50984. Alves da Costa, C., Sunyach, C., Pardossi-Piquard, R., Sevalle, J., Vincent, B., Boyer, N., et al. (2006). Presenilin-dependent gamma-secretase-mediated control of p53-associated cell death in Alzheimer's disease. J Neurosci 26, 6377–6385. Amici, M., Cecarini, V., Cuccioloni, M., Angeletti, M., Barocci, S., Rossi, G., et al. (2010). Interplay between 20S proteasomes and prion proteins in scrapie disease. J Neurosci Res 88, 191–201. Anantharam, V., Kanthasamy, A., Choi, C. J., Martin, D. P., Latchoumycandane, C., Richt, J. A., et al. (2008). Opposing roles of prion protein in oxidative stress- and ER stress-induced apoptotic signaling. Free Radic Biol Med 45, 1530–1541. Anderson, A. J., Stoltzner, S., Lai, F., Su, J., & Nixon, R. A. (2000). Morphological and biochemical assessment of DNA damage and apoptosis in Down syndrome and Alzheimer disease, and effect of postmortem tissue archival on TUNEL. Neurobiol Aging 21, 511–524. Anderson, A. J., Su, J. H., & Cotman, C. W. (1996). DNA damage and apoptosis in Alzheimer's disease: colocalization with c-Jun immunoreactivity, relationship to brain area, and effect of postmortem delay. J Neurosci 16, 1710–1719. Armogida, M., Petit, A., Vincent, B., Scarzello, S., da Costa, C. A., & Checler, F. (2001). Endogenous beta-amyloid production in presenilin-deficient embryonic mouse fibroblasts. Nat Cell Biol 3, 1030–1033. Aylon, Y., & Oren, M. (2011). p53: guardian of ploidy. Mol Oncol 5, 315–323. Bach, J. H., Chae, H. S., Rah, J. C., Lee, M. W., Park, C. H., Choi, S. H., et al. (2001). C-terminal fragment of amyloid precursor protein induces astrocytosis. J Neurochem 78, 109–120. Bae, B. I., Xu, H., Igarashi, S., Fujimuro, M., Agrawal, N., Taya, Y., et al. (2005). p53 mediates cellular dysfunction and behavioral abnormalities in Huntington's disease. Neuron 47, 29–41. Barzilai, A., & Melamed, E. (2003). Molecular mechanisms of selective dopaminergic neuronal death in Parkinson's disease. Trends Mol Med 9, 126–132. Becker, C., Brobert, G. P., Johansson, S., Jick, S. S., & Meier, C. R. (2010). Cancer risk in association with Parkinson disease: a population-based study. Parkinsonism Relat Disord 16, 186–190. Beckerman, R., & Prives, C. (2010). Transcriptional regulation by p53. Cold Spring Harb Perspect Biol 2, a000935. Behin, A., Hoang-Xuan, K., Carpentier, A. F., & Delattre, J. Y. (2003). Primary brain tumours in adults. Lancet 361, 323–331. Belyaev, N. D., Kellett, K. A., Beckett, C., Makova, N. Z., Revett, T. J., Nalivaeva, N. N., et al. (2010). The transcriptionally active amyloid precursor protein (APP) intracellular domain is preferentially produced from the 695 isoform of APP in a {beta}-secretasedependent pathway. J Biol Chem 285, 41443–41454. Bence, N. F., Sampat, R. M., & Kopito, R. R. (2001). Impairment of the ubiquitin–proteasome system by protein aggregation. Science 292, 1552–1555. Betarbet, R., Sherer, T. B., MacKenzie, G., Garcia-Osuna, M., Panov, A. V., & Greenamyre, J. T. (2000). Chronic systemic pesticide exposure reproduces features of Parkinson's disease. Nat Neurosci 3, 1301–1306. Biswas, S.C., Ryu, E., Park, C., Malagelada, C., & Greene, L. A. (2005). Puma and p53 play required roles in death evoked in a cellular model of Parkinson disease. Neurochem Res 30, 839–845. Blagosklonny, M. V., Toretsky, J., Bohen, S., & Neckers, L. (1996). Mutant conformation of p53 translated in vitro or in vivo requires functional HSP90. Proc Natl Acad Sci U S A 93, 8379–8383. Blandini, F., Sinforiani, E., Pacchetti, C., Samuele, A., Bazzini, E., Zangaglia, R., et al. (2006). Peripheral proteasome and caspase activity in Parkinson disease and Alzheimer disease. Neurology 66, 529–534. Bleeker, F. E., Molenaar, R. J., & Leenstra, S. (2012). Recent advances in the molecular understanding of glioblastoma. J Neurooncol 108, 11–27. Bommer, G. T., Gerin, I., Feng, Y., Kaczorowski, A. J., Kuick, R., Love, R. E., et al. (2007). p53-mediated activation of miRNA34 candidate tumor-suppressor genes. Curr Biol 17, 1298–1307. Bonifati, V., Rizzu, P., van Baren, M. J., Schaap, O., Breedveld, G. J., Krieger, E., et al. (2003). Mutations in the DJ-1 gene associated with autosomal recessive early-onset parkinsonism. Science 299, 256–259. Bossi, G., Mazzaro, G., Porrello, A., Crescenzi, M., Soddu, S., & Sacchi, A. (2004). Wild-type p53 gene transfer is not detrimental to normal cells in vivo: implications for tumor gene therapy. Oncogene 23, 418–425. Bossi, G., & Sacchi, A. (2007). Restoration of wild-type p53 function in human cancer: relevance for tumor therapy. Head Neck 29, 272–284.

Bourdon, J. C., De Laurenzi, V., Melino, G., & Lane, D. P. (2003). p53: 25 years of research and more questions to answer. Cell Death Diff 10, 397–399. Braicu, C., Pileczki, V., Irimie, A., & Berindan-Neagoe, I. (2013). p53siRNA therapy reduces cell proliferation, migration and induces apoptosis in triple negative breast cancer cells. Mol Cell Biochem 381, 61–68. Brandner, S., Isenmann, S., Raeber, A., Fischer, M., Sailer, A., Kobayashi, Y., et al. (1996). Normal host prion protein necessary for scrapie-induced neurotoxicity. Nature 379, 339–343. Bretaud, S., Allen, C., Ingham, P. W., & Bandmann, O. (2007). p53-dependent neuronal cell death in a DJ-1-deficient zebrafish model of Parkinson's disease. J Neurochem 100, 1626–1635. Brown, D. R. (2000). PrPSc-like prion protein peptide inhibits the function of cellular prion protein. Biochem J 352(Pt 2), 511–518. Brown, C. J., Cheok, C. F., Verma, C. S., & Lane, D. P. (2011). Reactivation of p53: from peptides to small molecules. Trends Pharmacol Sci 32, 53–62. Brown, C. J., Lain, S., Verma, C. S., Fersht, A.R., & Lane, D. P. (2009). Awakening guardian angels: drugging the p53 pathway. Nat Rev Cancer 9, 862–873. Buee, L., Troquier, L., Burnouf, S., Belarbi, K., Van der Jeugd, A., Ahmed, T., et al. (2010). From tau phosphorylation to tau aggregation: what about neuronal death? Biochem Soc Trans 38, 967–972. Büeler, H., Aguzzi, A., Sailer, A., Greiner, R., Autenried, P., Aguet, M., et al. (1993). Mice devoid of PrP are resistant to scrapie. Cell 73, 1339–1347. Buizza, L., Cenini, G., Lanni, C., Ferrari-Toninelli, G., Prandelli, C., Govoni, S., et al. (2012). Conformational altered p53 as an early marker of oxidative stress in Alzheimer's disease. PLoS One 7, e29789. Bullock, A. N., & Fersht, A.R. (2001). Rescuing the function of mutant p53. Nat Rev Cancer 1, 68–76. Burke, R. E. (1999). alpha-Synuclein and Parkinson's disease. Brain Res Bull 50, 465–466. Burton, A. (2005). Possible role for p53 in Huntington's disease. Lancet Neurol 4, 528–529. Caetano, F. A., Lopes, M. H., Hajj, G. N., Machado, C. F., Pinto Arantes, C., Magalhaes, A.C., et al. (2008). Endocytosis of prion protein is required for ERK1/2 signaling induced by stress-inducible protein 1. J Neurosci 28, 6691–6702. Caillava, C., Ranaldi, S., Lauritzen, I., Bauer, C., Fareh, J., Abraham, J.D., et al. (submitted for publication). Study on Ab34 Biology and Detection in Transgenic Mice Brain and Human CSF. (Submitted for publication). Cairncross, J. G., & Macdonald, D. R. (1988). Successful chemotherapy for recurrent malignant oligodendroglioma. Ann Neurol 23, 360–364. Campbell, W. A., Yang, H., Zetterberg, H., Baulac, S., Sears, J. A., Liu, T., et al. (2006). Zebrafish lacking Alzheimer presenilin enhancer 2 (Pen-2) demonstrate excessive p53-dependent apoptosis and neuronal loss. J Neurochem 96, 1423–1440. Cancer Genome Atlas Research Network (2008). Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061–1068. Cecarini, V., Bonfili, L., Cuccioloni, M., Mozzicafreddo, M., Angeletti, M., & Eleuteri, A.M. (2010). The relationship between the 20S proteasomes and prion-mediated neurodegenerations: potential therapeutic opportunities. Apoptosis 15, 1322–1335. Cenini, G., Sultana, R., Memo, M., & Butterfield, D. A. (2008). Elevated levels of pro-apoptotic p53 and its oxidative modification by the lipid peroxidation product, HNE, in brain from subjects with amnestic mild cognitive impairment and Alzheimer's disease. J Cell Mol Med 12, 987–994. Chandler, H., & Peters, G. (2013). Stressing the cell cycle in senescence and aging. Curr Opin Cell Biol 25, 765–771. Chandra, S., Gallardo, G., Fernàndez-Chàcon, R., Schlüter, O. M., & Südhof, T. C. (2005). a-synuclein cooperates with CSPa in preventing neurodegeneration. Cell 123, 383–396. Chang, T. C., Wentzel, E. A., Kent, O. A., Ramachandran, K., Mullendore, M., Lee, K., et al. (2007). Transactivation of miR-34a by p53 broadly influences gene expression and promotes apoptosis. Mol Cell 26, 745–752. Chartier-Harlin, M., Kachergus, J., Roumier, C., Mouroux, V., Douay, X., Lincoln, S., et al. (2004). Alpha-synuclein locus duplication as a cause of familial Parkinson's disease. Lancet 364, 1167–1169. Checler, F. (1995). Processing of the b-amyloid precursor protein and its regulation in Alzheimer's disease. J Neurochem 65, 1431–1444. Checler, F. (2013). Alzheimer's and prion diseases: PDK1 at the crossroads. Nat Med 19, 1088–1090. Checler, F., Alves da Costa, C., Ancolio, K., Lopez-Perez, E., & Marambaud, P. (2000). Role of the proteasome in Alzheimer's disease. Biochim Biophys Acta 1502, 133–138. Checler, F., Dunys, J., Pardossi-Piquard, R., & Alves da Costa, C. (2010). p53 is regulated by and regulates members of the gamma-secretase complex. Neurodegener Dis 7, 50–55. Chen, K., & Rajewsky, N. (2007). The evolution of gene regulation by transcription factors and microRNAs. Nat Rev Genet 8, 93–103. Chen, S. G., Teplow, D. B., Parchi, P., Teller, J. K., Gambetti, P., & Autilio-Gambetti, L. (1995). Truncated forms of the human prion protein in normal brain and in prion disease. J Biol Chem 270, 19173–19180. Chiarini, L. B., Freitas, A.R., Zanata, S. M., Brentani, R. R., Martins, V. R., & Linden, R. (2002). Cellular prion protein transduces neuroprotective signals. EMBO J 21, 3317–3326. Choi, H. S., Lee, S. H., Kim, S. Y., An, J. J., Hwang, S. -I., Kim, D. W., et al. (2006). Transduced Tat-a-synuclein protects against oxidative stress in vitro and in vivo. Biochem Mol Biol 39, 253–262. Chou, J., Greig, N. H., Reiner, D., Hoffer, B. J., & Wang, Y. (2011). Enhanced survival of dopaminergic neuronal transplants in hemiparkinsonian rats by the p53 inactivator PFT-alpha. Cell Transplant 20, 1351–1359. Chui, D. H., Dobo, E., Makifuchi, T., Akiyama, H., Kawakatsu, S., Petit, A., et al. (2001). Apoptotic neurons in Alzheimer's disease frequently show intracellular Abeta42 labeling. J Alzheimers Dis 3, 231–239.

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113 Colin, C., Virard, I., Baeza, N., Tchoghandjian, A., Fernandez, C., Bouvier, C., et al. (2007). Relevance of combinatorial profiles of intermediate filaments and transcription factors for glioma histogenesis. Neuropathol Appl Neurobiol 33, 431–439. Collinge, J. (2005). Molecular neurology of prion disease. J Neurol Neurosurg Psychiatry 76, 906–919. Coons, S. W., Johnson, P. C., Scheithauer, B. W., Yates, A. J., & Pearl, D. K. (1997). Improving diagnostic accuracy and interobserver concordance in the classification and grading of primary gliomas. Cancer 79, 1381–1393. Corney, D. C., Flesken-Nikitin, A., Godwin, A. K., Wang, W., & Nikitin, A. Y. (2007). MicroRNA-34b and MicroRNA-34c are targets of p53 and cooperate in control of cell proliferation and adhesion-independent growth. Cancer Res 67, 8433–8438. Corti, O., Lesage, S., & Brice, A. (2011). What genetics tells us about the causes and mechanisms of Parkinson's disease. Physiol Rev 91, 1161–1218. Crawford, L. V., Pim, D. C., Gurney, E. G., Goodfellow, P., & Taylor-Papadimitriou, J. (1981). Detection of a common feature in several human tumor cell lines—a 53,000-dalton protein. Proc Natl Acad Sci U S A 78, 41–45. Culmsee, C., & Landshamer, S. (2006). Molecular insights into mechanisms of the cell death program: role in the progression of neurodegenerative disorders. Curr Alzheimer Res 3, 269–283. da Costa, C. A. (2007). DJ-1: a newcomer in Parkinson's disease pathology. Curr Mol Med 7, 650–657. da Costa, C. A., Ancolio, K., & Checler, F. (2000). Wild-type but not Parkinson's disease-related Ala53Thr-a-synuclein protect neuronal cells from apoptotic stimuli. J Biol Chem 275, 24065–24069. da Costa, C. A., & Checler, F. (2010). A novel parkin-mediated transcriptional function links p53 to familial Parkinson's disease. Cell Cycle 9, 16–17. da Costa, C. A., & Checler, F. (2011). Apoptosis in Parkinson's disease: is p53 the missing link between genetic and sporadic Parkinsonism? Cell Signal 23, 963–968. da Costa, C. A., Dunys, J., Brau, F., Wilk, S., Cappai, R., & Checler, F. (2006). 6-Hydroxydopamine but not 1-methyl-4-phenylpyridinium abolishes alpha-synuclein anti-apoptotic phenotype by inhibiting its proteasomal degradation and by promoting its aggregation. J Biol Chem 281, 9824–9831. da Costa, C. A., Masliah, E., & Checler, F. (2003). β-Synuclein displays an antiapoptotic p53-dependent phenotype and protects neurons from 6-hydroxydopamine-induced caspase-3 activation. J Biol Chem 278, 37330–37335. da Costa, C. A., Sunyach, C., Giaime, E., West, A., Corti, O., Brice, A., et al. (2009). Transcriptional repression of p53 by parkin and impairment by mutations associated with autosomal recessive juvenile Parkinson's disease. Nat Cell Biol 11, 1370–1375. Dasgupta, G., & Momand, J. (1997). Geldanamycin prevents nuclear translocation of mutant p53. Exp Cell Res 237, 29–37. Davenport, C. M., Sevastou, I. G., Hooper, C., & Pocock, J. M. (2010). Inhibiting p53 pathways in microglia attenuates microglial-evoked neurotoxicity following exposure to Alzheimer peptides. J Neurochem 112, 552–563. de la Monte, S. M., Sohn, Y. K., & Wands, J. R. (1997). Correlates of p53- and Fas (CD95)-mediated apoptosis in Alzheimer's disease. J Neurol Sci 152, 73–83. De Strooper, B., Saftig, P., Craessaerts, K., Vanderstichele, H., Guhde, G., Annaert, W., et al. (1998). Deficiency of presenilin-1 inhibits the normal cleavage of amyloid precursor protein. Nature 391, 387–390. Demma, M., Maxwell, E., Ramos, R., Liang, L., Li, C., Hesk, D., et al. (2010). SCH529074, a small molecule activator of mutant p53, which binds p53 DNA binding domain (DBD), restores growth-suppressive function to mutant p53 and interrupts HDM2-mediated ubiquitination of wild type p53. J Biol Chem 285, 10198–10212. Devine, T., & Dai, M. S. (2013). Targeting the ubiquitin-mediated proteasome degradation of p53 for cancer therapy. Curr Pharm Des 19, 3248–3262. Diaz, A., De Jesus, L., Mendieta, L., Calvillo, M., Espinosa, B., Zenteno, E., et al. (2010). The amyloid-beta25-35 injection into the CA1 region of the neonatal rat hippocampus impairs the long-term memory because of an increase of nitric oxide. Neurosci Lett 468, 151–155. Ding, T., Zhou, X., Kouadir, M., Shi, F., Yang, Y., Liu, J., et al. (2013). Cellular prion protein participates in the regulation of inflammatory response and apoptosis in BV2 microglia during infection with Mycobacterium bovis. J Mol Neurosci 51, 118–126. Donehower, L. A., Harvey, M., Slagle, B.L., McArthur, M. J., Montgomery, C. A., Jr., Butel, J. S., et al. (1992). Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 356, 215–221. Drew, S.C., Haigh, C. L., Klemm, H. M., Masters, C. L., Collins, S. J., Barnham, K. J., et al. (2011). Optical imaging detects apoptosis in the brain and peripheral organs of prion-infected mice. J Neuropathol Exp Neurol 70, 143–150. Duan, W., Zhu, X., Ladenheim, B., Yu, Q. S., Guo, Z., Oyler, J., et al. (2002). p53 inhibitors preserve dopamine neurons and motor function in experimental parkinsonism. Ann Neurol 52, 597–606. Dumanchin-Njock, C., Alves da Costa, C. A., Mercken, L., Pradier, L., & Checler, F. (2001). The caspase-derived C-terminal fragment of betaAPP induces caspase-independent toxicity and triggers selective increase of Abeta42 in mammalian cells. J Neurochem 78, 1153–1161. Dunys, J., Kawarai, T., Sevalle, J., Dolcini, V., George-Hyslop, P.S., da Costa, C. A., et al. (2007). p53-dependent Aph-1 and Pen-2 anti-apoptotic phenotype requires the integrity of the {gamma}-secretase complex but is independent of its activity. J Biol Chem 282, 10516–10525. Dunys, J., Sevalle, J., Giaime, E., Pardossi-Piquard, R., Vitek, M. P., Renbaum, P., et al. (2009). p53-dependent control of transactivation of the Pen2 promoter by presenilins. J Cell Sci 122, 4003–4008. Duplan, E., Giaime, E., Viotti, J., Sevalle, J., Corti, O., Brice, A., et al. (2013). ER-stress-associated functional link between Parkin and DJ-1 via a transcriptional cascade involving the tumor suppressor p53 and the spliced X-box binding protein XBP-1. J Cell Sci 126, 2124–2133.

109

Duplan, E., Sevalle, J., Viotti, J., Goiran, T., Druon, C., Renbaum, P., et al. (2013). Parkin differently regulates presenilin-1 and presenilin-2 function by direct control of their promoter transcription. J Mol Cell Biol 5, 132–142. Eberhardt, O., & Schulz, J. B. (2003). Apoptotic mechanisms and antiapoptotic therapy in the MPTP model of Parkinson's disease. Toxicol Lett 139, 135–151. Edbauer, D., Kaether, C., Steiner, H., & Haass, C. (2004). Co-expression of nicastrin and presenilin rescues a loss of function mutant of APH-1. J Biol Chem 279, 37311–37315. el-Deiry, W. S., Kern, S. E., Pietenpol, J. A., Kinzler, K. W., & Vogelstein, B. (1992). Definition of a consensus binding site for p53. Nat Genet 1, 45–49. Eliyahu, D., Michalovitz, D., & Oren, M. (1985). Overproduction of p53 antigen makes established cells highly tumorigenic. Nature 316, 158–160. Eliyahu, D., Raz, A., Gruss, P., Givol, D., & Oren, M. (1984). Participation of p53 cellular tumour antigen in transformation of normal embryonic cells. Nature 312, 646–649. Engelender, S., Kaminsky, Z., Guo, X., Sharp, A. H., Amaravi, R. K., Kleiderlein, J. J., et al. (1999). Synphilin-1 associates with a-synuclein and promotes the formation of cytosolic inclusions. Nat Genet 22, 110–114. Fakharzadeh, S. S., Trusko, S. P., & George, D. L. (1991). Tumorigenic potential associated with enhanced expression of a gene that is amplified in a mouse tumor cell line. EMBO J 10, 1565–1569. Fan, J., Ren, H., Fei, E., Jia, N., Ying, Z., Jiang, P., et al. (2008). Sumoylation is critical for DJ-1 to repress p53 transcriptional activity. FEBS Lett 582, 1151–1156. Fan, J., Ren, H., Jia, N., Fei, E., Zhou, T., Jiang, P., et al. (2008). DJ-1 decreases Bax expression through repressing p53 transcriptional activity. J Biol Chem 283, 4022–4030. Feng, Z., Jin, S., Zupnick, A., Hoh, J., de Stanchina, E., Lowe, S., et al. (2006). p53 tumor suppressor protein regulates the levels of huntingtin gene expression. Oncogene 25, 1–7. Fenteany, G., & Schreiber, S. L. (1998). Lactacystin, proteasome function, and cell fate. J Biol Chem 273, 8545–8548. Fenteany, G., Standaert, R., Lane, W. S., Choi, S., Corey, E. J., & Schreiber, S. L. (1995). Inhibition of proteasome activities and subunit-specific amino-terminal threonine modification by lactacystin. Science 268, 726–731. Finlay, C. A., Hinds, P. W., & Levine, A. J. (1989). The p53 proto-oncogene can act as a suppressor of transformation. Cell 57, 1083–1093. Flammang, B., Pardossi-Piquard, R., Sevalle, J., Debayle, D., Dabert-Gay, A. S., Thevenet, A., et al. (2012). Evidence that the amyloid-beta protein precursor intracellular domain, AICD, derives from beta-secretase-generated C-terminal fragment. J Alzheimers Dis 30, 145–153. Fluhrer, R., Multhaup, G., Schlicksupp, A., Okochi, M., Takeda, M., Lammich, S., et al. (2003). Identification of a b-secretase activity, which truncates amyloid b-peptides after its presenilin-dependent generation. J Biol Chem 278, 5531–5538. Fogarty, M. P., McCormack, R. M., Noonan, J., Murphy, D., Gowran, A., & Campbell, V. A. (2010). A role for p53 in the beta-amyloid-mediated regulation of the lysosomal system. Neurobiol Aging 31, 1774–1786. Folch, J., Junyent, F., Verdaguer, E., Auladell, C., Pizarro, J. G., Beas-Zarate, C., et al. (2012). Role of cell cycle re-entry in neurons: a common apoptotic mechanism of neuronal cell death. Neurotox Res 22, 195–207. Fornai, F., Schluter, O. M., Lenzi, P., Gesi, M., Ruffoli, R., Ferrucci, M., et al. (2005). Parkinson-like syndrome induced by continuous MPTP infusion: convergent roles of the ubiquitin-proteasome system and alpha-synuclein. Proc Natl Acad Sci U S A 102, 3413–3418. Francis, R., McGrath, G., Zhang, J., Ruddy, D. A., Sym, M., Apfeld, J., et al. (2002). aph-1 and pen2 are required for notch pathway signaling, g-secretase cleavage of bAPP and presenilin protein accumulation. Dev Cell 3, 85–97. Friedler, A., Hansson, L. O., Veprintsev, D. B., Freund, S. M., Rippin, T. M., Nikolova, et al. (2002). A peptide that binds and stabilizes p53 core domain: chaperone strategy for rescue of oncogenic mutants. Proc Natl Acad Sci U S A 99, 937–942. Friedler, A., Veprintesev, D. B., Hansson, L. O., & Fersht, A.R. (2003). Kinetic instability of p53 core domain mutants. Implications for rescue by small molecules. J Biol Chem 278, 24108–24112. Fujiwara, T., Cai, D. W., Georges, R. N., Mukhopadhyay, T., Grimm, E. A., & Roth, J. A. (1994). Therapeutic effect of a retroviral wild-type p53 expression vector in an orthotopic lung cancer model. J Natl Cancer Inst 86, 1458–1462. Fulci, G., Ishii, N., & Van Meir, E. G. (1998). p53 and brain tumors: from gene mutations to gene therapy. Brain Pathol 8, 599–613. Funamoto, S., Morishima-Kawashima, M., Tanimura, Y., Hirotani, N., Saido, T. C., & Ihara, M. (2004). Truncated carboxyl-terminal fragments of b-amyloid precursor protein are processed to amyloid b-proteins 40 and 42. Biochemistry 43, 13532–13540. Funk, W. D., Pak, D. T., Karas, R. H., Wright, W. E., & Shay, J. W. (1992). A transcriptionally active DNA-binding site for human p53 protein complexes. Mol Cell Biol 12, 2866–2871. Furnari, F. B., Fenton, T., Bachoo, R. M., Mukasa, A., Stommel, J. M., Stegh, A., et al. (2007). Malignant astrocytic glioma: genetics, biology, and paths to treatment. Genes Dev 21, 2683–2710. Gandhi, S., & Wood, N. W. (2005). Molecular pathogenesis of Parkinson's disease. Hum Mol Genet 14(Spec No. 2), 2749–2755. Gasser, T. (1998). Genetics of Parkinson's disease. Clin Genet 54, 259–265. Giaime, E., Sunyach, C., Druon, C., Scarzello, S., Robert, G., Grosso, S., et al. (2010). Loss of function of DJ-1 triggered by Parkinson's disease-associated mutation is due to proteolytic resistance to caspase-6. Cell Death Differ 17, 158–169. Giaime, E., Sunyach, C., Herrant, M., Grosso, S., Auberger, P., McLean, P. J., et al. (2006). Caspase-3-derived C-terminal product of synphilin-1 displays antiapoptotic function via modulation of the p53-dependent cell death pathway. J Biol Chem 281, 11515–11522. Glenner, G. G., & Wong, C. W. (1984). Alzheimer's disease and Down's syndrome: sharing of a unique cerebrovascular amyloid fibril protein. Biochem Biophys Res Commun 120, 885–890.

110

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

Gloeckner, C. J., Schumacher, A., Boldt, K., & Ueffing, M. (2009). The Parkinson disease-associated protein kinase LRRK2 exhibits MAPKKK activity and phosphorylates MKK3/6 and MKK4/7, in vitro. J Neurochem 109, 959–968. Goedert, M., Spillantini, M. G., Del Tredici, K., & Braak, H. (2013). 100 years of Lewy pathology. Nat Rev Neurol 9, 13–24. Goncalves, A. P., Maximo, V., Lima, J., Singh, K. K., Soares, P., & Videira, A. (2011). Involvement of p53 in cell death following cell cycle arrest and mitotic catastrophe induced by rotenone. Biochim Biophys Acta 1813, 492–499. Goodger, Z. V., Rajendran, L., Trutzel, A., Kohli, B.M., Nitsch, R. M., & Konietzko, U. (2009). Nuclear signaling by the APP intracellular domain occurs predominantly through the amyloidogenic processing pathway. J Cell Sci 122, 3703–3714. Gosal, D., Ross, O. A., & Toft, M. (2006). Parkinson's disease: the genetics of a heterogeneous disorder. Eur J Neurol 13, 616–627. Gu, Y., Misonou, H., Sato, T., Dohmae, N., Takio, K., & Ihara, Y. (2001). Distinct intramembrane cleavages of the b-amyloid precursor protein family resembling g-secretase-like cleavage of Notch. J Biol Chem 276, 35235–35238. Gudkov, A. V., & Komarova, E. A. (2005). Prospective therapeutic applications of p53 inhibitors. Biochem Biophys Res Commun 331, 726–736. Guillot-Sestier, M. V., & Checler, F. (2012a). alpha-Secretase-derived cleavage of cellular prion yields biologically active catabolites with distinct functions. Neurodegener Dis 10, 294–297. Guillot-Sestier, M. V., & Checler, F. (2012b). Cellular prion and its catabolites in the brain: production and function. Curr Mol Med 12, 304–315. Guillot-Sestier, M. V., Sunyach, C., Druon, C., Scarzello, S., & Checler, F. (2009). The alpha-secretase-derived N-terminal product of cellular prion, N1, displays neuroprotective function in vitro and in vivo. J Biol Chem 284, 35973–35986. Guo, Q., Sebastian, L., Sopher, B.L., Miller, M. W., Ware, C. B., Martin, G. M., et al. (1999). Increased vulnerability of hippocampal neurons from presenilin-1 mutant knock-in mice to amyloid b-peptide toxicity: central roles of superoxide production and caspase activation. J Neurochem 72, 1019–1029. Gwinn-Hardy, K., & Singleton, A. A. (2002). Familial Lewy body diseases. J Geriatr Psychiatry Neurol 15, 217–223. Hagel, C., Laking, G., Laas, R., Scheil, S., Jung, R., Milde-Langosch, K., et al. (1996). Demonstration of p53 protein and TP53 gene mutations in oligodendrogliomas. Eur J Cancer 32A, 2242–2248. Harris, S. L., & Levine, A. J. (2005). The p53 pathway: positive and negative feedback loops. Oncogene 24, 2899–2908. Hartlage-Rubsamen, M., Morawski, M., Waniek, A., Jager, C., Zeitschel, U., Koch, B., et al. (2011). Glutaminyl cyclase contributes to the formation of focal and diffuse pyroglutamate (pGlu)-Abeta deposits in hippocampus via distinct cellular mechanisms. Acta Neuropathol 121, 705–719. Hartmann, A., & Hirsch, E. C. (2001). Parkinson's disease. The apoptosis hypothesis revisited. Adv Neurol 86, 143–153. Hashimoto, M., Bar-on, P., Ho, G., Takenouchi, T., Rockenstein, E., Crews, L., et al. (2004). b-synuclein regulates Akt activity in neuronal cells: a possible mechanism for neuroprotection in Parkinson's disease. J Biol Chem 279, 23622–23629. Haupt, Y., Maya, R., Kazaz, A., & Oren, M. (1997). Mdm2 promotes the rapid degradation of p53. Nature 387, 296–299. He, L., He, X., Lim, L. P., de Stanchina, E., Xuan, Z., Liang, Y., et al. (2007). A microRNA component of the p53 tumour suppressor network. Nature 447, 1130–1134. Hebert, L. E., Weuve, J., Scherr, P. A., & Evans, D. A. (2013). Alzheimer disease in the United States (2010–2050) estimated using the 2010 census. Neurology 80, 1778–1783. Hermeking, H. (2007). p53 enters the microRNA world. Cancer Cell 12, 414–418. Hermeking, H. (2010). The miR-34 family in cancer and apoptosis. Cell Death Differ 17, 193–199. Holtzman, D.M., & Epstein, C. J. (1992). The molecular genetics of Down syndrome. Mol Genet Med 2, 105–120. Holtzman, D.M., Mandelkow, E., & Selkoe, D. J. (2012). Alzheimer disease in 2020. Cold Spring Harb Perspect Med 2. Hooper, C., Meimaridou, E., Tavassoli, M., Melino, G., Lovestone, S., & Killick, R. (2007). p53 is upregulated in Alzheimer's disease and induces tau phosphorylation in HEK293a cells. Neurosci Lett 418, 34–37. Houlden, H., & Singleton, A.B. (2012). The genetics and neuropathology of Parkinson's disease. Acta Neuropathol 124, 325–338. Huang, Q., & Figueiredo-Pereira, M. E. (2010). Ubiquitin/proteasome pathway impairment in neurodegeneration: therapeutic implications. Apoptosis 15, 1292–1311. Hussain, I., Powell, D., Howlett, D. R., Tew, D.G., Meek, T. D., Chapman, C., et al. (1999). Identification of a novel aspartic protease (Asp2) as b-secretase. Mol Cell Neurosci 14, 419–427. Jeong, J. K., Moon, M. H., Seol, J. W., Seo, J. S., Lee, Y. J., & Park, S. Y. (2011). Prion peptide-mediated cellular prion protein overexpression and neuronal cell death can be blocked by aspirin treatment. Int J Mol Med 27, 689–693. Junn, E., & Mouradian, M. M. (2001). Apoptotic signaling in dopamine-induced cell death: the role of oxidative stress, p38 mitogen-activated protein kinase, cytochrome c and caspases. J Neurochem 78, 374–383. Kamijo, T., Weber, J.D., Zambetti, G., Zindy, F., Roussel, M. F., & Sherr, C. J. (1998). Functional and physical interactions of the ARF tumor suppressor with p53 and Mdm2. Proc Natl Acad Sci U S A 95, 8292–8297. Kim, Y. H., Nobusawa, S., Mittelbronn, M., Paulus, W., Brokinkel, B., Keyvani, K., et al. (2010). Molecular classification of low-grade diffuse gliomas. Am J Pathol 177, 2708–2714. Kim, H. S., Park, C. H., Cha, S. H., Lee, J. H., Lee, S., Kim, Y., et al. (2000). Carboxyl-terminal fragment of Alzheimer's APP destabilizes calcium homeostasis and renders neuronal cells vulnerable to excitotoxicity. FASEB J 14, 1508–1517. Kim, H. S., Park, C. H., & Suh, Y. H. (1998). C-terminal fragment of amyloid precursor protein inhibits calcium uptake into rat brain microsomes by Mg2+-Ca2+ ATPase. Neuroreport 9, 3875–3879.

Kim, S. H., & Suh, Y. H. (1996). Neurotoxicity of a carboxyl-terminal fragment of the Alzheimer's amyloid precursor protein. J Neurochem 67, 1172–1182. Kimberly, W. T., LaVoie, M. J., Ostaszewski, B.L., Ye, W., Wolfe, M. S., & Selkoe, D. J. (2003). g-secretase is a membrane protein complex comprised of presenilin, nicastrin, aph-1 and pen-2. Proc Natl Acad Sci U S A 100, 6382–6387. Kirla, R. M., Haapasalo, H. K., Kalimo, H., & Salminen, E. K. (2003). Low expression of p27 indicates a poor prognosis in patients with high-grade astrocytomas. Cancer 97, 644–648. Kitada, T., Asakawa, S., Hattori, N., Matsumine, H., Yamamura, Y., Minoshima, S., et al. (1998). Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 392, 605–608. Kitamura, Y., Shimohama, S., Kamoshima, W., Matsuoka, Y., Nomura, Y., & Taniguchi, T. (1997). Changes of p53 in the brains of patients with Alzheimer's disease. Biochem Biophys Res Commun 232, 418–421. Klein, C., & Schlossmacher, M. G. (2007). Parkinson disease, 10 years after its genetic revolution: multiple clues to a complex disorder. Neurology 69, 2093–2104. Kloosterman, W. P., & Plasterk, R. H. (2006). The diverse functions of microRNAs in animal development and disease. Dev Cell 11, 441–450. Kobayashi, K., Hayashi, M., Nakano, H., Shimazaki, M., Sugimori, K., & Koshino, Y. (2004). Correlation between astrocyte apoptosis and Alzheimer changes in gray matter lesions in Alzheimer's disease. J Alzheimers Dis 6, 623–632 (discussion 673–681). Komarov, P. G., Komarova, E. A., Kondratov, R. V., Christov-Tselkov, K., Coon, J. S., Chernov, M. V., et al. (1999a). A chemical inhibitor of p53 that protects mice from the side effect of cancer therapy. Science 285, 1733–1737. Komarov, P. G., Komarova, E. A., Kondratov, R. V., Christov-Tselkov, K., Coon, J. S., Chernov, M. V., et al. (1999b). A chemical inhibitor of p53 that protects mice from the side effects of cancer therapy. Science 285, 1733–1737. Konietzko, U. (2012). AICD nuclear signaling and its possible contribution to Alzheimer's disease. Curr Alzheimer Res 9, 200–216. Kovacs, G. G., & Budka, H. (2010). Distribution of apoptosis-related proteins in sporadic Creutzfeldt–Jakob disease. Brain Res 1323, 192–199. Kubbutat, M. H., Jones, S. N., & Vousden, K. H. (1997). Regulation of p53 stability by Mdm2. Nature 387, 299–303. Kumar, K. R., Lohmann, K., & Klein, C. (2012). Genetics of Parkinson disease and other movement disorders. Curr Opin Neurol 25, 466–474. Kusiak, J. W., Izzo, J. A., & Zhao, B. (1996). Neurodegeneration in Alzheimer disease. Is apoptosis involved? Mol Chem Neuropathol 28, 153–162. Lane, D. P., & Crawford, L. V. (1979). T antigen is bound to a host protein in SV40-transformed cells. Nature 278, 261–263. Lane, D., & Levine, A. (2010). p53 research: the past thirty years and the next thirty years. Cold Spring Harb Perspect Biol 2, a000893. Lanni, C., Racchi, M., Stanga, S., Mazzini, G., Ranzenigo, A., Polotti, R., et al. (2010). Unfolded p53 in blood as a predictive signature signature of the transition from mild cognitive impairment to Alzheimer's disease. J Alzheimers Dis 20, 97–104. Lassmann, H., Bancher, C., Breitschopf, H., Wegiel, J., Bobinski, M., Jellinger, K., et al. (1995). Cell death in Alzheimer's disease evaluated by DNA fragmentation in situ. Acta Neuropathol 89, 35–41. Lee, J. T., & Gu, W. (2010). The multiple levels of regulation by p53 ubiquitination. Cell Death Differ 17, 86–92. Lee, S. J., Kim, D. C., Choi, B. H., Ha, H., & Kim, K. T. (2006). Regulation of p53 by activated protein kinase C-delta during nitric oxide-induced dopaminergic cell death. J Biol Chem 281, 2215–2224. Lee, V. M. -Y., & Trojanowski, J. Q. (1999). Neurodegenerative tauopathies: human disease and transgenic mouse models. Neuron 24, 507–510. Lefranc-Jullien, S., Sunyach, C., & Checler, F. (2006). The ε-secretase-derived N-terminal product of the β-amyloid precursor protein behaves as a membrane-bound protein and undergoes α-, β-, and γ-secretases cleavages. J Neurochem 97, 807–817. Lev, N., Melamed, E., & Offen, D. (2003). Apoptosis and Parkinson's disease. Prog Neuropsychopharmacol Biol Psychiatry 27, 245–250. Levidou, G., El-Habr, E., Saetta, A. A., Bamias, C., Katsouyanni, K., Patsouris, E., et al. (2010). P53 immunoexpression as a prognostic marker for human astrocytomas: a meta-analysis and review of the literature. J Neurooncol 100, 363–371. Levine, A. J., & Oren, M. (2009). The first 30 years of p53: growing ever more complex. Nat Rev Cancer 9, 749–758. Li, X. L., Wang, G. R., Jing, Y. Y., Pan, M. M., Dong, C. F., Zhou, R. M., et al. (2011). Cytosolic PrP induces apoptosis of cell by disrupting microtubule assembly. J Mol Neurosci 43, 316–325. Liang, J., Parchaliuk, D., Medina, S., Sorensen, G., Landry, L., Huang, S., et al. (2009). Activation of p53-regulated pro-apoptotic signaling pathways in PrP-mediated myopathy. BMC Genomics 10, 201. Lin, X., Koelsch, G., Wu, S., Downs, D., Dashti, A., & Tang, J. (2000). Human aspartyl protease memapsin 2 cleaves the beta-secretase site of beta-amyloid precursor protein. Proc Natl Acad Sci U S A 97, 1456–1460. Linden, R., Martins, V. R., Prado, M.A., Cammarota, M., Izquierdo, I., & Brentani, R. R. (2008). Physiology of the prion protein. Physiol Rev 88, 673–728. Linzer, D. I., & Levine, A. J. (1979). Characterization of a 54 K dalton cellular SV40 tumor antigen present in SV40-transformed cells and uninfected embryonal carcinoma cells. Cell 17, 43–52. Lombino, F., Biundo, F., Tamayev, R., Arancio, O., & D'Adamio, L. (2013). An intracellular threonine of amyloid-beta precursor protein mediates synaptic plasticity deficits and memory loss. PLoS One 8, e57120. Lopes, M. H., Hajj, G. N., Muras, A. G., Mancini, G. L., Castro, R. M., Ribeiro, K. C., et al. (2005). Interaction of cellular prion and stress-inducible protein 1 promotes neuritogenesis and neuroprotection by distinct signaling pathways. J Neurosci 25, 11330–11339. Lott, I. T. (1982). Down's syndrome, aging, and Alzheimer's disease: a clinical review. Ann N Y Acad Sci 396, 15–27.

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113 Louis, D. N., Ohgaki, H., Wiestler, O. D., Cavenee, W. K., Burger, P. C., Jouvet, A., et al. (2007). The 2007 WHO classification of tumours of the central nervous system. Acta Neuropathol 114, 97–109. Lu, Q. R., Park, J. K., Noll, E., Chan, J. A., Alberta, J., Yuk, D., et al. (2001). Oligodendrocyte lineage genes (OLIG) as molecular markers for human glial brain tumors. Proc Natl Acad Sci U S A 98, 10851–10856. Lu, D. C., Rabizadeh, S., Chandra, S., Shayya, R. F., Ellerby, L. M., Ye, X., et al. (2000). A second cytotoxic proteolytic peptide derived from amyloid beta-protein precursor. Nat Med 6, 397–404. Lyahyai, J., Bolea, R., Serrano, C., Vidal, E., Pumarola, M., Badiola, J. J., et al. (2007). Differential expression and protein distribution of Bax in natural scrapie. Brain Res 1180, 111–120. Ma, S. L., Pastorino, L., Zhou, X. Z., & Lu, K. P. (2012). Prolyl isomerase Pin1 promotes amyloid precursor protein (APP) turnover by inhibiting glycogen synthase kinase-3beta (GSK3beta) activity: novel mechanism for Pin1 to protect against Alzheimer disease. J Biol Chem 287, 6969–6973. Maintz, D., Fiedler, K., Koopmann, J., Rollbrocker, B., Nechev, S., Lenartz, D., et al. (1997). Molecular genetic evidence for subtypes of oligoastrocytomas. J Neuropathol Exp Neurol 56, 1098–1104. Mandelkow, E. (1999). The tangled tale of tau. Nature 402, 588–589. Mann, D.M. (1988). Alzheimer's disease and Down's syndrome. Histopathology 13, 125–137. Maries, E., Dass, B., Collier, T. J., Kordower, J. H., & Steece-Collier, K. (2003). The role of alpha-synuclein in Parkinson's disease: insights from animal models. Nat Rev Neurosci 4, 727–738. Martin, L. J., Pan, Y., Price, A.C., Sterling, W., Copeland, N. G., Jenkins, N. A., et al. (2006). Parkinson's disease alpha-synuclein transgenic mice develop neuronal mitochondrial degeneration and cell death. J Neurosci 26, 41–50. Martin, D., Salinas, M., Lopez-Valdaliso, R., Serrano, E., Recuero, M., & Cuadrado, A. (2001). Effect of the Alzheimer amyloid fragment Abeta(25–35) on Akt/PKB kinase and survival of PC12 cells. J Neurochem 78, 1000–1008. Maruyama, W., & Naoi, M. (2002). Cell death in Parkinson's disease. J Neurol 249(Suppl. 2), II6–II10. Masliah, E., Mallory, M., Alford, M., Tanaka, S., & Hansen, L. A. (1998). Caspase dependent DNA fragmentation might be associated with excitotoxicity in Alzheimer disease. J Neuropathol Exp Neurol 57, 1041–1052. Maslon, M. M., & Hupp, T. R. (2010). Drug discovery and mutant p53. Trends Cell Biol 20, 542–555. Masters, C. L., Simons, G., Weinman, N. A., Multhaup, G., Mc Donald, B.L., & Beyreuther, K. (1985). Amyloid plaque core protein in Alzheimer's Disease and Down Syndrome. Proc Natl Acad Sci U S A 82, 4245–4249. Matsui, T., Ramasamy, K., Ingelsson, M., Fukumoto, H., Conrad, C., Frosch, M. P., et al. (2006). Coordinated expression of caspase 8, 3 and 7 mRNA in temporal cortex of Alzheimer disease: relationship to formic acid extractable abeta42 levels. J Neuropathol Exp Neurol 65, 508–515. Mattei, V., Matarrese, P., Garofalo, T., Tinari, A., Gambardella, L., Ciarlo, L., et al. (2011). Recruitment of cellular prion protein to mitochondrial raft-like microdomains contributes to apoptosis execution. Mol Biol Cell 22, 4842–4853. Mendrysa, S. M., Ghassemifar, S., & Malek, R. (2011). p53 in the CNS: perspectives on development, stem cells, and cancer. Genes Cancer 2, 431–442. Meunier, J., Ieni, J., & Maurice, T. (2006). The anti-amnesic and neuroprotective effects of donepezil against amyloid beta25-35 peptide-induced toxicity in mice involve an interaction with the sigma1 receptor. Br J Pharmacol 149, 998–1012. Milosevic, J., Schwarz, S.C., Ogunlade, V., Meyer, A. K., Storch, A., & Schwarz, J. (2009). Emerging role of LRRK2 in human neural progenitor cell cycle progression, survival and differentiation. Mol Neurodegener 4, 25. Mittler, M.A., Walters, B. C., & Stopa, E. G. (1996). Observer reliability in histological grading of astrocytoma stereotactic biopsies. J Neurosurg 85, 1091–1094. Mizuno, Y., Hattori, N., Mori, H., Suzuki, T., & Tanaka, K. (2001). Parkin and Parkinson's disease. Curr Opin Neurol 14, 477–482. Mochizuki, H., Goto, K., Mori, H., & Mizuno, Y. (1996). Histochemical detection of apoptosis in Parkinson's disease. J Neurol Sci 137, 120–123. Mogi, M., Kondo, T., Mizuno, Y., & Nagatsu, T. (2007). p53 protein, interferon-gamma, and NF-kappaB levels are elevated in the parkinsonian brain. Neurosci Lett 414, 94–97. Moll, U. M., LaQuaglia, M., Benard, J., & Riou, G. (1995). Wild-type p53 protein undergoes cytoplasmic sequestration in undifferentiated neuroblastomas but not in differentiated tumors. Proc Natl Acad Sci U S A 92, 4407–4411. Moll, U. M., Riou, G., & Levine, A. J. (1992). Two distinct mechanisms alter p53 in breast cancer: mutation and nuclear exclusion. Proc Natl Acad Sci U S A 89, 7262–7266. Momand, J., Zambetti, G. P., Olson, D. C., George, D., & Levine, A. J. (1992). The mdm-2 oncogene product forms a complex with the p53 protein and inhibits p53-mediated transactivation. Cell 69, 1237–1245. Mookherjee, P., & Johnson, G. V. (2001). Tau phosphorylation during apoptosis of human SH-SY5Y neuroblastoma cells. Brain Res 921, 31–43. Moon, C., Oh, Y., & Roth, J. A. (2003). Current status of gene therapy for lung cancer and head and neck cancer. Clin Cancer Res 9, 5055–5067. Morris, H. R. (2005). Genetics of Parkinson's disease. Ann Med 37, 86–96. Muller, P., Ceskova, P., & Vojtesek, B. (2005). Hsp90 is essential for restoring cellular functions of temperature-sensitive p53 mutant protein but not for stabilization and activation of wild-type p53: implications for cancer therapy. J Biol Chem 280, 6682–6691. Musicco, M., Adorni, F., Di Santo, S., Prinelli, F., Pettenati, C., Caltagirone, C., et al. (2013). Inverse occurrence of cancer and Alzheimer disease: a population-based incidence study. Neurology 81, 322–328. Nagatsu, T. (2002). Parkinson's disease: changes in apoptosis-related factors suggesting possible gene therapy. J Neural Transm 109, 731–745.

111

Nair, V. D. (2006). Activation of p53 signaling initiates apoptotic death in a cellular model of Parkinson's disease. Apoptosis 11, 955–966. Nair, V. D., McNaught, K. S., Gonzalez-Maeso, J., Sealfon, S.C., & Olanow, C. W. (2006). p53 mediates nontranscriptional cell death in dopaminergic cells in response to proteasome inhibition. J Biol Chem 281, 39550–39560. Nakamura, Y. (2004). Isolation of p53-target genes and their functional analysis. Cancer Sci 95, 7–11. Nikolaev, A. Y., Li, M., Puskas, N., Qin, J., & Gu, W. (2003). Parc: a cytoplasmic anchor for p53. Cell 112, 29–40. Ning, A., Cui, J., To, E., Ashe, K. H., & Matsubara, J. (2008). Amyloid-beta deposits lead to retinal degeneration in a mouse model of Alzheimer disease. Invest Ophthalmol Vis Sci 49, 5136–5143. Nozaki, M., Tada, M., Kobayashi, H., Zhang, C. L., Sawamura, Y., Abe, H., et al. (1999). Roles of the functional loss of p53 and other genes in astrocytoma tumorigenesis and progression. Neuro Oncol 1, 124–137. Ohgaki, H., Dessen, P., Jourde, B., Horstmann, S., Nishikawa, T., Di Patre, P. L., et al. (2004). Genetic pathways to glioblastoma: a population-based study. Cancer Res 64, 6892–6899. Ohgaki, H., & Kleihues, P. (2005a). Epidemiology and etiology of gliomas. Acta Neuropathol 109, 93–108. Ohgaki, H., & Kleihues, P. (2005b). Population-based studies on incidence, survival rates, and genetic alterations in astrocytic and oligodendroglial gliomas. J Neuropathol Exp Neurol 64, 479–489. Ohgaki, H., & Kleihues, P. (2007). Genetic pathways to primary and secondary glioblastoma. Am J Pathol 170, 1445–1453. Ohgaki, H., & Kleihues, P. (2011). Genetic profile of astrocytic and oligodendroglial gliomas. Brain Tumor Pathol 28, 177–183. Ohyagi, Y., Asahara, H., Chui, D. H., Tsuruta, Y., Sakae, N., Miyoshi, K., et al. (2005). Intracellular Abeta42 activates p53 promoter: a pathway to neurodegeneration in Alzheimer's disease. FASEB J 19, 255–257. Ohyagi, Y., Yamada, T., Nishioka, K., Clarke, N. J., Tomlinson, A. J., Naylor, S., et al. (2000). Selective increase in cellular A beta 42 is related to apoptosis but not necrosis. Neuroreport 11, 167–171. Oliner, J.D., Pietenpol, J. A., Thiagalingam, S., Gyuris, J., Kinzler, K. W., & Vogelstein, B. (1993). Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature 362, 857–860. Olivier, M., Hollstein, M., & Hainaut, P. (2010). TP53 mutations in human cancers: origins, consequences, and clinical use. Cold Spring Harb Perspect Biol 2, a001008. Olzmann, J. A., Brown, K., Wilkinson, K. D., Rees, H. D., Huai, Q., Ke, H., et al. (2003). Familial Parkinson's disease-associated L166P mutation disrupts DJ-1 protein folding and function. J Biol Chem 279, 8506–8515. Ottolini, D., Cali, T., Negro, A., & Brini, M. (2013). The Parkinson disease-related protein DJ-1 counteracts mitochondrial impairment induced by the tumour suppressor protein p53 by enhancing endoplasmic reticulum-mitochondria tethering. Hum Mol Genet 22, 2152–2168. Oulès, B., Del Prete, D., Greco, B., Zhang, X., Lauritzen, I., Sevalle, J., et al. (2012). Ryanodine receptor blockade reduces amyloid-beta load and memory impairments in Tg2576 mouse model of Alzheimer disease. J Neurosci 32, 11820–11834. Ozaki, T., Li, Y., Kikuchi, H., Tomita, T., Iwatsubo, T., & Nakagawara, A. (2006). The intracellular domain of the amyloid precursor protein (AICD) enhances the p53-mediated apoptosis. Biochem Biophys Res Commun 351, 57–63. Paisan-Ruiz, C., Jain, S., Evans, E. W., Gilks, W. P., Simon, J., van den Brug, M., et al. (2004). Cloning of the gene containing mutations that cause PARK8-linked Parkinson's disease. Neuron 44, 595–600. Paitel, E., Alves da Costa, C., Vilette, D., Grassi, J., & Checler, F. (2002). Overexpression of PrPc triggers caspase 3 activation: potentiation by proteasome inhibitors and blockade by anti-PrP antibodies. J Neurochem 83, 1208–1214. Paitel, E., Fahraeus, R., & Checler, F. (2003). Cellular prion protein sensitizes neurons to apoptotic stimuli through Mdm2-regulated and p53-dependent caspase 3-like activation. J Biol Chem 278, 10061–10066. Paitel, E., Sunyach, C., Alves da Costa, C., Bourdon, J. C., Vincent, B., & Checler, F. (2004). Primary cultured neurons devoid of cellular prion display lower responsiveness to staurosporine through the control of p53 at both transcriptional and posttranscriptional levels. J Biol Chem 279, 612–618. Pardossi-Piquard, R., & Checler, F. (2012). The physiology of the beta-amyloid precursor protein intracellular domain AICD. J Neurochem 120(Suppl. 1), 109–124. Pardossi-Piquard, R., Dunys, J., Giaime, E., Guillot-Sestier, M. V., St George-Hyslop, P., Checler, F., et al. (2009). p53-dependent control of cell death by nicastrin: lack of requirement for presenilin-dependent gamma-secretase complex. J Neurochem 109, 225–237. Parsons, D. W., Jones, S., Zhang, X., Lin, J. C., Leary, R. J., Angenendt, P., et al. (2008). An integrated genomic analysis of human glioblastoma multiforme. Science 321, 1807–1812. Passer, B., Pellegrini, L., Russo, C., Siegel, R. M., Lenardio, M. J., Schettini, G., et al. (2000). Generation of an apoptotic intracellular peptide by g-secretase cleavage of Alzheimer's amyloid b protein precursor. J Alzheimers Dis 2, 289–301. Pastorino, L., Ma, S. L., Balastik, M., Huang, P., Pandya, D., Nicholson, L., et al. (2012). Alzheimer's disease-related loss of Pin1 function influences the intracellular localization and the processing of AbetaPP. J Alzheimers Dis 30, 277–297. Perfettini, J. L., Castedo, M., Nardacci, R., Ciccosanti, F., Boya, P., Roumier, T., et al. (2005). Essential role of p53 phosphorylation by p38 MAPK in apoptosis induction by the HIV-1 envelope. J Exp Med 201, 279–289. Petit, A., Bihel, F., Alves da Costa, C., Pourquie, O., Checler, F., & Kraus, J. L. (2001). New protease inhibitors prevent gamma-secretase-mediated production of Abeta40/42 without affecting Notch cleavage. Nat Cell Biol 3, 507–511. Petitjean, A., Mathe, E., Kato, S., Ishioka, C., Tavtigian, S. V., Hainaut, P., et al. (2007). Impact of mutant p53 functional properties on TP53 mutation patterns and tumor

112

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113

phenotype: lessons from recent developments in the IARC TP53 database. Hum Mutat 28, 622–629. Peuchmaur, M., Lacour, M.A., Sevalle, J., Lisowski, V., Touati-Jallabe, Y., Rodier, F., et al. (2013). Further characterization of a putative serine protease contributing to the gamma-secretase cleavage of beta-amyloid precursor protein. Bioorg Med Chem 21, 1018–1029. Pietri, M., Dakowski, C., Hannoui, S., Alleaume-Butaux, A., Hernandez-Rapp, J., Ragagnin, A., et al. (2013). PDK1 decreases TACE-mediated alpha-secretase activity and promotes disease progression in prion and Alzheimer's diseases. Nat Med 19, 1124–1131. Polymeropoulos, M. H., Lavedan, C., Leroy, E., Ide, S. E., Dehejia, A., Dutra, A., et al. (1997). Mutation in the a-synuclein gene identified in families with Parkinson Disease. Science 276, 2045–2047. Pomerantz, J., Schreiber-Agus, N., Liegeois, N. J., Silverman, A., Alland, L., Chin, L., et al. (1998). The Ink4a tumor suppressor gene product, p19Arf, interacts with MDM2 and neutralizes MDM2's inhibition of p53. Cell 92, 713–723. Prestia, A., Caroli, A., van der Flier, W. M., Ossenkoppele, R., Van Berckel, B., Barkhof, F., et al. (2013). Prediction of dementia in MCI patients based on core diagnostic markers for Alzheimer disease. Neurology 80, 1048–1056. Prusiner, S. B. (1998). Prions. Proc Natl Acad Sci U S A 95, 13363–13383. Rahman-Roblick, R., Hellman, U., Becker, S., Bader, F. G., Auer, G., Wiman, K. G., et al. (2008). Proteomic identification of p53-dependent protein phosphorylation. Oncogene 27, 4854–4859. Raver-Shapira, N., Marciano, E., Meiri, E., Spector, Y., Rosenfeld, N., Moskovits, N., et al. (2007). Transcriptional activation of miR-34a contributes to p53-mediated apoptosis. Mol Cell 26, 731–743. Robert, G., Puissant, A., Dufies, M., Marchetti, S., Jacquel, A., Cluzeau, T., et al. (2012). The caspase 6 derived N-terminal fragment of DJ-1 promotes apoptosis via increased ROS production. Cell Death Differ 19, 1769–1778. Robles, A. I., & Harris, C. C. (2010). Clinical outcomes and correlates of TP53 mutations and cancer. Cold Spring Harb Perspect Biol 2, a001016. Roe, C. M., Behrens, M. I., Xiong, C., Miller, J. P., & Morris, J. C. (2005). Alzheimer disease and cancer. Neurology 64, 895–898. Rogaev, E. I., Sherrington, R., Rogaeva, E. A., Levesque, G., Ikeda, M., Liang, Y., et al. (1995). Familial Alzheimer's disease in kindreds with missense mutations in a gene on chromosome 1 related to the Alzheimer's disease type 3 gene. Nature 376, 775–778. Roperch, J. P., Alvaro, V., Prieur, S., Tuynder, M., Nemani, M., Lethrosne, F., et al. (1998). Inhibition of presenilin 1 expression is promoted by p53 and p21WAF-1 and results in apoptosis and tumor suppression. Nat Med 4, 835–838. Rosner, S., Giladi, N., & Orr-Urtreger, A. (2008). Advances in the genetics of Parkinson's disease. Acta Pharmacol Sin 29, 21–34. Sailer, A., Bueler, H., Fischer, M., Aguzzi, A., & Weissmann, C. (1994). No propagation of prions in mice devoid of PrP. Cell 77, 967–968. Sajan, F. D., Martiniuk, F., Marcus, D. L., Frey, W. H., 2nd, Hite, R., Bordayo, E. Z., et al. (2007). Apoptotic gene expression in Alzheimer's disease hippocampal tissue. Am J Alzheimers Dis Other Demen 22, 319–328. Sasaki, M., Nie, L., & Maki, C. G. (2007). MDM2 binding induces a conformational change in p53 that is opposed by heat-shock protein 90 and precedes p53 proteasomal degradation. J Biol Chem 282, 14626–14634. Sastre, M., Steiner, H., Fuchs, K., Capell, A., Multhaup, G., Condron, M. M., et al. (2001). Presenilin dependent g-secretase processing of b-amyloid precursor protein at a site corresponding to the S3 cleavage of Notch. EMBO Rep 2, 835–841. Scardigli, R., Bossi, G., Blandino, G., Crescenzi, M., Soddu, S., & Sacchi, A. (1997). Expression of exogenous wt-p53 does not affect normal hematopoiesis: implications for bone marrow purging. Gene Ther 4, 1371–1378. Schapira, A. H., & Jenner, P. (2011). Etiology and pathogenesis of Parkinson's disease. Mov Disord 26, 1049–1055. Schatzl, H. M., Laszlo, L., Holtzman, D.M., Tatzelt, J., DeArmond, S. J., Weiner, R. I., et al. (1997). A hypothalamic neuronal cell line persistently infected with scrapie prions exhibits apoptosis. J Virol 71, 8821–8831. Scheffner, M. (1998). Ubiquitin, E6-AP, and their role in p53 inactivation. Pharmacol Ther 78, 129–139. Schilling, S., Zeitschel, U., Hoffmann, T., Heiser, U., Francke, M., Kehlen, A., et al. (2008). Glutaminyl cyclase inhibition attenuates pyroglutamate Abeta and Alzheimer's disease-like pathology. Nat Med 14, 1106–1111. Seidl, R., Fang-Kircher, S., Bidmon, B., Cairns, N., & Lubec, G. (1999). Apoptosis-associated proteins p53 and APO-1/Fas (CD95) in brains of adult patients with Down syndrome. Neurosci Lett 260, 9–12. Selkoe, D. J. (1991). The molecular pathology of Alzheimer's disease. Neuron 6, 487–498. Sembritzki, O., Hagel, C., Lamszus, K., Deppert, W., & Bohn, W. (2002). Cytoplasmic localization of wild-type p53 in glioblastomas correlates with expression of vimentin and glial fibrillary acidic protein. Neuro Oncol 4, 171–178. Sevalle, J., Amoyel, A., Robert, P., Fournie-Zaluski, M. C., Roques, B., & Checler, F. (2009). Aminopeptidase A contributes to the N-terminal truncation of amyloid beta-peptide. J Neurochem 109, 248–256. Seward, M. E., Sanson, E., Norambuena, A., Reimann, A., Cochran, J. N., Li, R., et al. (2013). Amyloid-beta signas through tau to drive ectopic neuronal cell cycle re-entry in Alzheimer's disease. J Cell Sci 126, 1278–1286. Sherrington, R., Froelich, S., Sorbi, S., Campion, D., Chi, H., Rogaeva, E. A., et al. (1996). Alzheimer's disease associated with mutations in presenilin 2 is rare and variably penetrant. Hum Mol Genet 5, 985–988. Shi, X. P., Tugusheva, K., Bruce, J. E., Lucka, A., Wu, G. X., Chen-Dodson, E., et al. (2003). Beta-secretase cleavage at amino acid residue 34 in the amyloid beta peptide is dependent upon gamma-secretase activity. J Biol Chem 278, 21286–21294.

Shimura, H., Hattori, N., Kubo, S., Mizuno, Y., Asakawa, S., Minoshima, S., et al. (2000). Familial Parkinson disease gene product, parkin, is a ubiquitin-protein ligase. Nat Genet 25, 302–305. Shinbo, Y., Taira, T., Niki, T., Iguchi-Ariga, S. M., & Ariga, H. (2005). DJ-1 restores p53 transcription activity inhibited by Topors/p53BP3. Int J Oncol 26, 641–648. Shmerling, D., Hegyi, I., Fischer, M., Blattler, T., Brandner, S., Gotz, J., et al. (1998). Expression of amino-terminally truncated PrP in the mouse leading to ataxia and specific cerebellar lesions. Cell 93, 203–214. Simunovic, F., Yi, M., Wang, Y., Macey, L., Brown, L. T., Krichevsky, A.M., et al. (2009). Gene expression profiling of substantia nigra dopamine neurons: further insights into Parkinson's disease pathology. Brain 132, 1795–1809. Singh, N., Gu, Y., Bose, S., Kalepu, S., Mishra, R. S., & Verghese, S. (2002). Prion peptide 106–126 as a model for prion replication and neurotoxicity. Front Biosci 7, a60–a71. Singleton, A.B., Farrer, M., Johnson, J., Singleton, A., Hague, S., Kachergus, J., et al. (2003). Alpha-synuclein locus triplication causes Parkinson's disease. Science 302. Sinha, S., Anderson, J. P., Barbour, R., Basi, G. S., Caccavello, R., Davis, D., et al. (1999). Purification and cloning of amyloid precursor protein b-secretase from human brain. Nature 402, 537–540. Smith, W. W., Liu, Z., Liang, Y., Masuda, N., Swing, D. A., Jenkins, N. A., et al. (2010). Synphilin-1 attenuates neuronal degeneration in the A53T alpha-synuclein transgenic mouse model. Hum Mol Genet 19, 2087–2098. Sorice, M., Mattei, V., Tasciotti, V., Manganelli, V., Garofalo, T., & Misasi, R. (2012). Trafficking of PrPc to mitochondrial raft-like microdomains during cell apoptosis. Prion 6, 354–358. Spillantini, M. G., & Goedert, M. (1998). Tau protein pathology in neurodegenerative diseases. TIN 21, 428–433. Stambolic, V., MacPherson, D., Sas, D., Lin, Y., Snow, B., Jang, Y., et al. (2001). Regulation of PTEN transcription by p53. Mol Cell 8, 317–325. Stander, M., Peraud, A., Leroch, B., & Kreth, F. W. (2004). Prognostic impact of TP53 mutation status for adult patients with supratentorial World Health Organization Grade II astrocytoma or oligoastrocytoma: a long-term analysis. Cancer 101, 1028–1035. Su, J. H., Anderson, A. J., Cummings, B. J., & Cotman, C. W. (1994). Immunohistochemical evidence for apoptosis in Alzheimer's disease. Neuroreport 5, 2529–2533. Su, J. H., Satou, T., Anderson, A. J., & Cotman, C. W. (1996). Up-regulation of Bcl-2 is associated with neuronal DNA damage in Alzheimer's disease. Neuroreport 7, 437–440. Suh, Y. H., & Checler, F. (2002). Amyloid precursor protein, presenilins, and alpha-synuclein: molecular pathogenesis and pharmacological applications in Alzheimer's disease. Pharmacol Rev 54, 469–525. Sunico, C., Nakamura, T., Rockenstein, E., Mante, M., Adame, A., Chan, S., et al. (2013). S-nitrosylaton of parkin as a novel regulator of p53-mediated neuronal cell death in sporadic Parkinson's disease. Mol Neurodegener 8. http://dx.doi.org/10. 1186/1750-1326-8-29 (in press). Sunyach, C., & Checler, F. (2005). Combined pharmacological, mutational and cell biology approaches indicate that p53-dependent caspase 3 activation triggered by cellular prion is dependent on its endocytosis. J Neurochem 92, 1399–1407. Sunyach, C., Cisse, M.A., da Costa, C. A., Vincent, B., & Checler, F. (2007). The C-terminal products of cellular prion protein processing, C1 and C2, exert distinct influence on p53-dependent staurosporine-induced caspase-3 activation. J Biol Chem 282, 1956–1963. Surguchov, A. (2008). Molecular and cellular biology of synucleins. Int Rev Cell Mol Biol 270, 225–317. Takahashi, K., Uchida, C., Shin, R. W., Shimazaki, K., & Uchida, T. (2008). Prolyl isomerase, Pin1: new findings of post-translational modifications and physiological substrates in cancer, asthma and Alzheimer's disease. Cell Mol Life Sci 65, 359–375. Takasugi, N., Tomita, T., Hayashi, I., Tsuruoka, M., Niimura, M., Takahashi, Y., et al. (2003). The role of presenilin cofactors in the g-secretase complex. Nature 422, 438–441. Tatton, N., & Kish, S. (1997). In situ detection of apoptotic nuclei in the substantia nigra compacta of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated mice using terminal deoxynucleotidyl transferase labelling and acridine orange staining. Neuroscience 77, 1037–1048. Taylor, D. R., Parkin, E. T., Cocklin, S. L., Ault, J. R., Ashcroft, A. E., Turner, A. J., et al. (2009). Role of ADAMs in the ectodomain shedding and conformational conversion of the prion protein. J Biol Chem 284, 22590–22600. Tekirian, T. L., Saido, T. C., Markesbery, W. R., Russell, M. J., Wekstein, D. R., Patel, E., et al. (1998). N-terminal heterogeneity of parenchymal and cerebrovascular Abeta deposit. J Neuropathol Exp Neurol 57, 76–94. Thellung, S., Corsaro, A., Villa, V., Simi, A., Vella, S., Pagano, A., et al. (2011). Human PrP90-231-induced cell death is associated with intracellular accumulation of insoluble and protease-resistant macroaggregates and lysosomal dysfunction. Cell Death Dis 2, e138. Toledo, F., & Wahl, G. M. (2007). MDM2 and MDM4: p53 regulators as targets in anticancer therapy. Int J Biochem Cell Biol 39, 1476–1482. Trimmer, P. A., Smith, T. S., Jung, A.B., & Bennett, J. P. J. (1996). Dopaminergic neurons from transgenic mice with a knockout of the p53 gene resist MPTP neurotoxicity. Neurodegeneration 5, 233–239. Trinh, J., & Farrer, M. (2013). Advances in the genetics of Parkinson disease. Nat Rev Neurol 9, 445–454. Trojanowski, J. Q., & Lee, V. M. (1998). Aggregation of neurofilament and alpha-synuclein proteins in Lewy bodies: implications for the pathogenesis of Parkinson disease and Lewy body dementia. Arch Neurol 55, 151–152. Uberti, D., Lanni, C., Carsana, T., Francisconi, S., Missale, C., Racchi, M., et al. (2006). Identification of a mutant-like conformation of p53 in fibroblasts from sporadic Alzheimer's disease patients. Neurobiol Aging 27, 1193–1201. Uberti, D., Lanni, C., Racchi, M., Govoni, S., & Memo, M. (2008). Conformationally altered p53: a putative peripheral marker for Alzheimer's disease. Neurodegener Dis 5, 209–211.

F. Checler, C. Alves da Costa / Pharmacology & Therapeutics 142 (2014) 99–113 Uberti, D., Carsana, T., Bernardi, E., Rodella, L., Grigolato, P., Lanni, C., et al. (2002). Selective impairment of p53-mediated cell death in fibroblasts from sporadic Alzheimer's disease patients. J Cell Sci 115, 3131–3138. Valencia-Sanchez, M.A., Liu, J., Hannon, G. J., & Parker, R. (2006). Control of translation and mRNA degradation by miRNAs and siRNAs. Genes Dev 20, 515–524. Valente, E. M., Abou-Sleiman, P.M., Caputo, V., Muqit, M. M., Harvey, K., Gispert, S., et al. (2004). Hereditary early-onset Parkinson's disease caused by mutations in PINK1. Science 304, 1158–1160. Vassar, R., Bennett, B.D., Babu-Khan, S., Khan, S., Mendiaz, E. A., Denis, P., et al. (1999). β-Secretase cleavage of Alzheimer's amyloid precursor protein by the transmembrane aspartic protease BACE. Science 286, 735–741. Velez-Pardo, C., Ospina, G. G., & Jimenez del Rio, M. (2002). Abeta[25–35] peptide and iron promote apoptosis in lymphocytes by an oxidative stress mechanism: involvement of H2O2, caspase-3, NF-kappaB, p53 and c-Jun. Neurotoxicology 23, 351–365. Villarejo, F., de Diego, J. M., & de la Riva, A. G. (2008). Prognosis of cerebellar astrocytomas in children. Childs Nerv Syst 24, 203–210. Vincent, B., Paitel, E., Frobert, Y., Lehmann, S., Grassi, J., & Checler, F. (2000). Phorbol ester-regulated cleavage of normal prion protein in HEK293 human cells and murine neurons. J Biol Chem 275, 35612–35616. Vincent, B., Paitel, E., Saftig, P., Frobert, Y., Hartmann, D., De Strooper, B., et al. (2001). The disintegrins ADAM10 and TACE contribute to the constitutive and phorbol ester-regulated normal cleavage of the cellular prion protein. J Biol Chem 276, 37743–37746. Vincent, B., Sunyach, C., Orzechowski, H. D., St George-Hyslop, P., & Checler, F. (2009). p53-Dependent transcriptional control of cellular prion by presenilins. J Neurosci 29, 6752–6760. Viotti, J., Duplan, E., Caillava, C., Condat, J., Goiran, T., Giordano, C., et al. (2013). Glioma Tumor Grades Correlates With Parkin Depletion in Mutant p53-linked Tumors and Results From Loss of Function of p53 Transcriptional Activity. http://dx.doi.org/10. 1038/onc.2013.124 (in press). Vogelstein, B., & Kinzler, K. W. (1992). p53 function and dysfunction. Cell 70, 523–526. Vogelstein, B., Lane, D., & Levine, A. J. (2000). Surfing the p53 network. Nature 408, 307–310. Von Eckardstein, K., Gries, H., Bolik, E., Cervos-Navarro, J., Tschairkin, I. N., & Patt, S. (1997). p53 mutation and protein alteration in 50 gliomas. Retrospective study by DNA-sequencing techniques and immunohistochemistry. Histol Histopathol 12(3), 611–616. Vousden, K. H., & Lu, X. (2002). Live or let die: the cell's response to p53. Nat Rev Cancer 2, 594–604. Wakabayashi, K., Engelender, S., Yoshimoto, M., Tsuji, S., Ross, C. A., & Takahashi, H. (2000). Synphilin-1 is present in Lewy bodies in Parkinson's disease. Ann Neurol 47, 521–523. Wang, S., Simon, B. P., Bennett, D. A., Schneider, J. A., Malter, J. S., & Wang, D. S. (2007). The significance of Pin1 in the development of Alzheimer's disease. J Alzheimers Dis 11, 13–23. Wang, X., Wang, F., Sy, M. S., & Ma, J. (2005). Calpain and other cytosolic proteases can contribute to the degradation of retro-translocated prion protein in the cytosol. J Biol Chem 280, 317–325. Watanabe, K., Sato, K., Biernat, W., Tachibana, O., von Ammon, K., Ogata, N., et al. (1997). Incidence and timing of p53 mutations during astrocytoma progression in patients with multiple biopsies. Clin Cancer Res 3, 523–530. Watanabe, T., Yokoo, H., Yokoo, M., Yonekawa, Y., Kleihues, P., & Ohgaki, H. (2001). Concurrent inactivation of RB1 and TP53 pathways in anaplastic oligodendrogliomas. J Neuropathol Exp Neurol 60, 1181–1189. Wawrzynow, B., Zylicz, A., Wallace, M., Hupp, T., & Zylicz, M. (2007). MDM2 chaperones the p53 tumor suppressor. J Biol Chem 282, 32603–32612. Weidemann, A., Eggert, S., Reinhard, F. B.M., Vogel, M., Paliga, K., Baier, G., et al. (2002). A novel e-cleavage within the transmembrane domain of the Alzheimer amyloid precursor protein demonstrates homology with the Notch processing. Biochemistry 41, 2825–2835. Weissmann, C., Enari, M., Klöhn, P. -C., Rossi, D., & Flechsig, E. (2002). Molecular biology of prions. Acta Neurobiol Exp 62, 153–166. Weller, M., Felsberg, J., Hartmann, C., Berger, H., Steinbach, J. P., Schramm, J., et al. (2009). Molecular predictors of progression-free and overall survival in patients with newly

113

diagnosed glioblastoma: a prospective translational study of the German Glioma Network. J Clin Oncol 27, 5743–5750. Wirths, O., Breyhan, H., Cynis, H., Schilling, S., Demuth, H. U., & Bayer, T. A. (2009). Intraneuronal pyroglutamate-Abeta 3-42 triggers neurodegeneration and lethal neurological deficits in a transgenic mouse model. Acta Neuropathol 118, 487–496. Woo, M. G., Xue, K., Liu, J., McBride, H., & Tsang, B. K. (2012). Calpain-mediated processing of p53-associated parkin-like cytoplasmic protein (PARC) affects chemosensitivity of human ovarian cancer cells by promoting p53 subcellular trafficking. J Biol Chem 287, 3963–3975. Wrensch, M., Minn, Y., Chew, T., Bondy, M., & Berger, M. S. (2002). Epidemiology of primary brain tumors: current concepts and review of the literature. Neuro Oncol 4, 278–299. Wu, F., Wang, Z., Gu, J. H., Ge, J. B., Liang, Z. Q., & Qin, Z. H. (2013). p38(MAPK)/p53-Mediated Bax induction contributes to neurons degeneration in rotenone-induced cellular and rat models of Parkinson's disease. Neurochem Int 63, 133–140. Xu, Z., Cawthon, D., McCastlain, K. A., Duhart, H. M., Newport, G. D., Fang, H., et al. (2005). Selective alterations of transcription factors in MPP+-induced neurotoxicity in PC12 cells. Neurotoxicology 26, 729–737. Xu, Z., Cawthon, D., McCastlain, K. A., Slikker, W., Jr., & Ali, S. F. (2005). Selective alterations of gene expression in mice induced by MPTP. Synapse 55, 45–51. Yan, R., Bienkowski, M. J., Shuck, M. E., Miao, H., Tory, M. C., Pauley, A.M., et al. (1999). Membrane-anchored aspartyl protease with Alzheimer's disease b-secretase activity. Nature 402, 533–537. Yang, F., Sun, X., Beech, W., Teter, B., Wu, S., Sigel, J., et al. (1998). Antibody to caspase-cleaved actin detects apoptosis in differentiated neuroblastoma and plaque-associated neurons and microglia in Alzheimer's disease. Am J Pathol 152, 379–389. Yankner, B.A. (1996). New clues to Alzheimer's disease: unraveling the roles of amyloid and tau. Nat Med 2, 850–852. Yap, Y. H., & Say, Y. H. (2011). Resistance against apoptosis by the cellular prion protein is dependent on its glycosylation status in oral HSC-2 and colon LS 174T cancer cells. Cancer Lett 306, 111–119. Yap, Y. H., & Say, Y. H. (2012). Resistance against tumour necrosis factor alpha apoptosis by the cellular prion protein is cell-specific for oral, colon and kidney cancer cell lines. Cell Biol Int 36, 273–277. Yu, G., Nishimura, M., Arawaka, S., Levitan, D., Zhang, L., Tandon, A., et al. (2000). Nicastrin modulates presenilin-mediated notch/glp1 signal transduction and bAPP processing. Nature 407, 48–54. Zanata, S. M., Lopes, M. H., Mercadante, A. F., Hajj, G. N., Chiarini, L. B., Nomizo, R., et al. (2002). Stress-inducible protein 1 is a cell surface ligand for cellular prion that triggers neuroprotection. EMBO J 21, 3307–3316. Zeimet, A. G., & Marth, C. (2003). Why did p53 gene therapy fail in ovarian cancer? Lancet Oncol 4, 415–422. Zhang, C., Lin, M., Wu, R., Wang, X., Yang, B., Levine, A. J., et al. (2011). Parkin, a p53 target gene, mediates the role of p53 in glucose metabolism and the Warburg effect. Proc Natl Acad Sci U S A 108, 16259–16264. Zheng, H., Ying, H., Yan, H., Kimmelman, A.C., Hiller, D. J., Chen, A. J., et al. (2008). p53 and Pten control neural and glioma stem/progenitor cell renewal and differentiation. Nature 455, 1129–1133. Zhou, M., Ottenberg, G., Sferrazza, G. F., & Lasmezas, C. I. (2012). Highly neurotoxic monomeric alpha-helical prion protein. Proc Natl Acad Sci U S A 109, 3113–3118. Zhou, H. Y., Tan, Y. Y., Wang, Z. Q., Wang, G., Lu, G. Q., & Chen, S. D. (2010). Proteasome inhibitor lactacystin induces cholinergic degeneration. Can J Neurol Sci 37, 229–234. Zhu, H. B., Yang, K., Xie, Y. Q., Lin, Y. W., Mao, Q. Q., & Xie, L. P. (2013). Silencing of mutant p53 by siRNA induces cell cycle arrest and apoptosis in human bladder cancer cells. World J Surg Oncol 11, 22. Zhu, X., Yu, Q. S., Cutler, R. G., Culmsee, C. W., Holloway, H. W., Lahiri, D. K., et al. (2002). Novel p53 inactivators with neuroprotective action: syntheses and pharmacological evaluation of 2-imino-2,3,4,5,6,7-hexahydrobenzothiazole and 2-imino-2,3,4,5,6,7hexahydrobenzoxazole derivatives. J Med Chem 45, 5090–5097. Zimprich, A., Biskup, S., Leitner, P., Lichtner, P., Farrer, M., Lincoln, S., et al. (2004). Mutations in LRRK2 cause autosomal-dominant Parkinsonism with pleomorphic pathology. Neuron 44, 601–607.

p53 in neurodegenerative diseases and brain cancers.

More than thirty years elapsed since a protein, not yet called p53 at the time, was detected to bind SV40 during viral infection. Thousands of papers ...
689KB Sizes 0 Downloads 0 Views