• Aucun résultat trouvé

Replication stress as a source of telomere recombination during replicative senescence in Saccharomyces cerevisiae

N/A
N/A
Protected

Academic year: 2021

Partager "Replication stress as a source of telomere recombination during replicative senescence in Saccharomyces cerevisiae"

Copied!
23
0
0

Texte intégral

(1)

HAL Id: hal-03085000

https://hal.archives-ouvertes.fr/hal-03085000

Submitted on 21 Dec 2020

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

during replicative senescence in Saccharomyces cerevisiae

Marie-Noëlle Simon, Dmitri Churikov, Vincent Géli

To cite this version:

Marie-Noëlle Simon, Dmitri Churikov, Vincent Géli. Replication stress as a source of telomere recom- bination during replicative senescence in Saccharomyces cerevisiae. FEMS Yeast Research, Oxford University Press (OUP), 2016, 16 (7), pp.fow085. �10.1093/femsyr/fow085�. �hal-03085000�

(2)

1

Replication stress as a source of telomere recombination during replicative senescence in Saccharomyces cerevisiae.

Marie-Noëlle Simon*, Dmitri Churikov and Vincent Géli

Centre de Recherche en Cancérologie de Marseille (CRCM), Inserm, U1068, Marseille, F-13009, France, CNRS, UMR7258, Marseille, F-13009, Institut Paoli-Calmettes, Marseille, F-13009, France, Aix-Marseille

University, UM 105, F-13284, Marseille, France. Equipe labellisée Ligue

*Correspondence to: marie-noelle.simon@inserm.fr

ABSTRACT

Replicative senescence is triggered by short unprotected telomeres that arise in the absence of telomerase. In addition, telomeres are known as difficult regions to replicate due to their repetitive G-rich sequence prone to secondary structures and tightly bound non-histone proteins. Here we review accumulating evidence that telomerase inactivation in yeast immediately unmasks the problems associated with replication stress at telomeres. Early after telomerase inactivation, yeast cells undergo successive rounds of stochastic DNA damages and become dependent on recombination for viability long before the bulk of telomeres are getting critically short. The switch from telomerase to recombination to repair replication stress-induced damage at telomeres creates telomere instability, which may drive further genomic alterations and prepare the ground for telomerase-independent immortalization observed in yeast survivors and in 15% of human cancer.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(3)

INTRODUCTION

In eukaryotic cells the physical ends of linear chromosomes have to be distinguished from accidental DNA double-strand breaks (DSBs). For this purpose, cells have evolved telomeres, the sophisticated structures composed of tandem G-rich DNA repeats bound by a complex of proteins called shelterin that cap chromosome ends. The G-rich strand of the telomeric DNA extends beyond the duplex region making a 3’ end overhang, which serves as substrate for telomerase, a reverse transcriptase that extends the G-rich strand in the 5’ to 3’

direction (Blackburn and Collins 2011). Telomeres prevent DNA damage checkpoint activation (Denchi and de Lange 2007) and undesirable DNA repair activities at the ends of chromosomes. In the budding yeast, Saccharomyces cerevisiae, telomeres are formed by an approximately 300 bp -long array of irregular TG1-3 repeats (McEachern and Blackburn 1994) and a 10-15 bases of 3’ end G-overhang (Wellinger et al. 1993). The capping function of shelterin has been a subject of excellent reviews both in yeast (Dewar and Lydall 2012;

Ribeyre and Shore 2012) and mammals (Giraud-Panis et al. 2010; O'Sullivan and Karlseder 2010). Briefly, in budding yeast, duplex telomeric DNA is bound by Rap1 that serves as a platform to recruit Rif1 and Rif2 (Wotton and Shore 1997). Rap1-Rif1-Rif2 limits the access of exonucleases, binding of replication protein A (RPA) and prevents checkpoint activation (Kupiec, 2014). In addition, Rap1 is essential to block telomere fusions through NHEJ (Pardo and Marcand 2005). The single-stand termini of telomeres are bound by Cdc13, a subunit of the CST (Cdc13/Stn1/Ten1) complex that assumes diverse functions at telomeres including an essential capping function that prevents degradation of the 5’ ends (Churikov et al. 2013;

Greetham et al. 2015). Finally, the yKu70/Ku80 complex contributes to telomere capping by inhibiting resection by Exo1 as it does at DSBs (Bonetti et al. 2010; Vodenicharov et al.

2010). In mammals another level of protection is provided by the t-loop formed by invasion of the double-stranded telomeric region by the G-strand overhang (Griffith et al. 1999;

Doksani et al. 2013). This structure does not appear to occur in wild-type yeast because the limited length of the G-overhangs. Nevertheless, a fold back structure of telomeric heterochromatin has been described in yeast (de Bruin et al. 2000) that reduces the susceptibility to nucleolytic degradation (Poschke et al. 2012).

A unique sequence of telomeric DNA prone to fold into G-quadruplexes (Smith et al.

2011) and stable protein-DNA complexes cause pausing of the replication forks and their slow progression through subtelomeres and telomeres (Ivessa et al. 2002; Makovets et al.

2004). Telomere replication thus requires specific helicases and repair proteins (Ivessa et al.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(4)

2002; Paeschke et al. 2010; Hardy et al. 2014). These proteins are particularly important due to the unidirectional replication of telomeres where no converging forks from a neighboring origin can compensate for defective replication (Gilson and Geli 2007). Nevertheless recombination proteins are not required to maintain telomeres in cells with active telomerase (Shampay and Blackburn 1988). Indeed, intra- or inter-telomere recombination is rare in wild type cell as shown by the low level of recombination foci at telomeres detected in vivo (Khadaroo et al. 2009). This suggests that replication stress-induced damages at telomeres are either limited or compensated by the action of the telomerase in wild type cells. Consistently, it has been shown in S. pombe that stalled replication forks form substrates for telomerase (Dehe et al. 2012).

Evidence is mounting that the first consequence of inactivating telomerase is to unmask the problems that stem from replication stress at telomeres well before the appearance of short uncapped telomeres (Khadaroo et al. 2009; Xie et al. 2015; Xu et al. 2015; Jay et al. 2016). In this review we summarize recent data supporting the idea that in the absence of repair by telomerase, cells become dependent on recombination to repair replication-stress induced damage at telomeres. The switch from telomerase to recombination has a profound impact on telomere (and potentially entire genome) stability, which is manifested by dependency of cell viability on multiple DNA repair pathways. Yeast cells that naturally do not experience the absence of telomerase readily reveal the telomere replication problems, when rendered telomerase-negative, and thus will continue to bring new insight into the mechanisms of telomere replication and recombination at telomeres.

Homologous recombination plays multiple roles at telomeres in the absence of telomerase.

The telomerase enzyme compensates for the loss of terminal DNA which stems from the inability of DNA polymerase to complete the synthesis of the lagging strand at the very end of DNA molecule (Olovnikov 1973) and from 5’ to 3’ exonucleolytic processing of the telomeres generated by the leading strand DNA synthesis (Larrivee et al. 2004; Soudet et al.

2014). Upon inactivation of telomerase in yeast, telomeres shorten on average 3-5 bp per population doubling (PD) (Marcand et al. 1999). Proliferation rate is not immediately affected, but it begins to decline after about 20-30 PDs until it reaches a minimum that is referred to as crisis driven by telomere erosion (Figure 1). When telomeres reach a critical length, cells

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(5)

enter an irreversible G2/M arrest due to activation of Mec1, an ortholog of the ataxia telangiectasia- and Rad3-related (ATR) phosphatidylinositol 3 kinases (Enomoto et al. 2002;

IPjma and Greider 2003). It is thought that very short telomeres lose their capping function and are detected as and processed largely as DSBs. Following permanent activation of the checkpoint, few cells can recover growth capacity by regenerating functional telomeres through recombination-based mechanisms. Two types of survivors are described in yeast based on genetic requirements and telomere organization (Le et al. 1999; Chen et al. 2001) (Figure 1). The formation of both types appears to involve break-induced replication (BIR) since both depend on Rad52 and Pol32 (Lydeard et al. 2007). Type I survivors rely on the strand invasion activity of Rad51 and show an amplification of the Y’ subtelomeric elements together with the interstitial telomeric sequences (ITS). Type II survivors carry long heterogeneous in length telomere repeats and depend on Rad59 (Chen et al. 2001) as well as checkpoint activation (Grandin and Charbonneau 2007a) and proteins required for resection of the 5’ end of DSBs (Costelloe et al. 2012; Hardy et al. 2014).

The idea of replicative senescence being triggered by telomere attrition that results exclusively from the “end replication problem” was challenged by unprecedented heterogeneity of the population of senescing cells even when they all descended from a single telomerase minus cell (and thus with the same heritage of telomere length). Recent technical advances in the use of microfluidic devises made it possible to directly observe the dynamics of cell divisions in individual cell lineages upon telomerase loss (Xie et al. 2015; Xu et al.

2015). These studies revealed that heterogeneity of the cell cycle duration is already evident within the first 10 divisions without telomerase (Xie et al. 2015). Stochastically some lineages encounter transient cell cycle arrest followed by resumption of divisions, a pattern totally distinct from the short-telomere driven permanent G2/M arrest observed later during terminal senescence (Xu et al. 2015). Unfortunately, microfluidic devises do not yet allow relating transient arrests to specific telomere defects. Nevertheless, much can be learned from analyses of the mutants. For example, deletion of RAD51 substantially decreases the frequency of early lags in cell cycle progression (Xu et al. 2015) suggesting that Rad51-dependent recombination is responsible for them. This observation is in agreement with relocalization of the essential HR factor Rad52 into telomere foci as early as 20 PDs after inactivation of telomerase (Khadaroo et al. 2009). Premature senescence of the rad52 and rad51 cells in the absence of telomerase (Chen et al. 2001; Lundblad and Blackburn 1993) suggests that these telomere recombination events are essential to sustain proliferation early after telomerase inactivation.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(6)

Telomere repair activities that precede survivor formation rely at least in part on non- reciprocal translocation of terminal telomeric repeats (Figure 2A). Sequencing of telomere DNA revealed unequal inter- and intra-telomeric exchanges that occur before the onset of survivor formation (Teixeira et al. 2004; Lee et al. 2007; Abdallah et al. 2009; Kozak et al.

2010; Chang et al. 2011). In yeast, individual telomeres differ in their precise sequence because telomerase RNA template can be aligned to the G-strand overhang in multiple registers (Forstemann et al. 2000). As a consequence, TG1-3 tracts are well suited for homeologous or microhomology-based recombination. Consistently, inactivation of mismatch repair, which suppresses recombination between sequences that contain mismatches (Spell and Jinks-Robertson 2003), delays senescence (Rizki and Lundblad 2001).

Recently, we analyzed in details another mode of telomere repair that operates in pre- senescent cells (Churikov et al. 2014). We showed that in the absence of telomerase, short terminal TG1-3 repeats engage in pairing with ITSs between subtelomeric X and Y’ elements and initiate Pol32-dependent BIR, which results in non-reciprocal translocation of the entire Y’ element and terminal TG1-3 tract from the donor chromosome (Figure 2B). This repair event is associated with a transient arrest of cell cycle progression but the checkpoint function of Mec1 is not essential for efficient translocation. Y’ acquisition has genetic requirements that are distinct from those described for Y’ amplification in type I post senescence survivors.

First, Y’ translocation is mainly promoted by Rad52 associated with Rad59, while Rad51 has only a modest effect in this pathway (Churikov et al. 2014). Second, the Bloom RecQ helicase homolog Sgs1 favors Y’ translocation while it has an inhibitory role on recombination between Y’ elements in wild type cells (Watt et al. 1996). The dependency on Rad59 and positive effect of Sgs1 are in keeping with recombination between homeologous sequences, but how the G-rich overhang anneals to internal TG1-3 tracts that are mainly double stranded remains to be characterized. Possibly, partial opening of the double helix during replication, especially when the fork pauses at subtelomeric region (Makovets et al. 2004), creates permissive condition for the initiation of recombination.

The involvement of Rad51 in recombination between TG1-3 tracts is uncertain. Yeast Rad51 has a preference for GT-rich sequences (Tracy et al. 1997), but at telomeres, single- stranded TG1-3 repeats are bound by Cdc13 (Mitton-Fry et al. 2002). It remains unclear whether Rad51, even assisted by Rad52 and other mediators, would efficiently strip Cdc13 from the single-stranded TG1-3 repeats. One possibility is that Rad51 may support telomere replication rather than recombination in pre-senescing cells. Indeed, Rad51 is now recognized as one of the factors that escort replication fork (Carr and Lambert 2013), where it protects

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(7)

the nascent DNA at stalled replication forks from Mre11-dependent degradation and facilitates replication fork restart (Hashimoto et al. 2010; Gonzalez-Prieto et al. 2013).

Telomerase inactivation aggravates replication stress at the telomeres.

The experiments conducted in the laboratory of Teresa Teixeira showed that the onset of senescence is similar in two populations of cells descended from one mitosis after inactivation of telomerase but the pattern of transient arrests differs between the two sister populations (Xu et al. 2013; Xu et al. 2015). As the starting populations shared the same set of telomeres, these observations suggest that the terminal arrest is determined by initial telomere length while the transient arrests are not (Xu et al. 2015). Therefore, transient arrests might occur in response to stochastic damage that is independent of gradual telomere shortening. In fact, the frequency of transient cell cycle arrest is increased in a strain with elongated as compared to wild-type telomeres (Xu et al., 2015), indicating that the probability of stochastic damage may positively correlate with telomere length. Alterations of the dynamics of DNA replication have emerged as a major source of genome instability (Magdalou et al. 2014). This raises the possibility that loss of telomerase activity further aggravates replication stress that is already present within telomeric regions in wild-type cells (Geronimo and Zakian 2016). In this speculative model, telomerase might repair damage resulting from replication stress at telomeres either by elongating accidentally broken telomere (Figure 3A) or the G-rich single strand exposed at the regressed replication fork (Figure 3B). In the absence of telomerase, either accidental breakage or enzymatic processing of the stalled fork would generate short recombinogenic telomeres (Figure 3).

The checkpoint response to eroded telomeres share many common partners with the response to DSBs by inducing a cascade of events that leads to the activation of Rad53, Mec1, Mec3, Chk1 and Dun1 that are all required for the terminal cell cycle arrest in G2/M (Enomoto et al. 2002; IPjma and Greider 2003; Grandin and Charbonneau 2007a).

Nevertheless the phosphorylation of Rad53 upon inactivation of telomerase depends on both Rad9 and Mrc1 (Grandin et al. 2005), two adaptor proteins implicated respectively in the response to DSB and replication stress (Vialard et al. 1998; Alcasabas et al. 2001; Osborn and Elledge 2003). Furthermore, inactivation of telomerase immediately creates dependence of the viability on the DNA damage adaptors (Jay et al. 2016). Importantly, this dependence can be alleviated by elevation of dNTPs pools (Jay et al. 2016) that facilitates replication

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(8)

indicating that telomerase is indeed essential for resolution of the replication stress at telomeres. Interestingly, MRC1, in contrast to other DDR genes, is not required for the cell cycle arrest elicited by eroded telomeres and instead, its deletion accelerates senescence (Grandin and Charbonneau 2007b). Moderate replication stress linked to a reduction of fork speed lead to recruitment of mediators that are crucial for fork protection but do not block S- phase progression and mitotic onset (Koundrioukoff et al. 2013). It is thus possible that cells are under pressure of moderate replicative stress that requires Mrc1 and recombination to be settled in the absence of telomerase. In line with this possibility, the checkpoint deficient mrc1AQ allele or deletion of the ATM-homolog TEL1, but not RAD9, exacerbates the precocious heterogeneity of cell cycle duration during pre-senescence (Xie et al. 2015).

The importance of counter-acting replication stress to delay the onset of senescence is further supported by the phenotype of various mutants affecting the repair of replication- generated damages. In budding yeast, fork restart depends on the Smc5/6 and Rtt101-Mms1- Mms22 (Luke et al. 2006; Duro et al. 2008; Irmisch et al. 2009). Notably Smc5/6 is required to slow senescence in telomerase defective cells (Chavez et al. 2010; Noel and Wellinger 2011). smc5-6 mutant cells accumulate recombination intermediates at telomeres in the absence of telomerase suggesting commitment of template switch mechanism to bypass a stall-induced lesion (Chavez et al. 2010). Similarly, MMS1 is required to maintain viability before the onset of senescence (Abdallah et al. 2009; Noel and Wellinger 2011). RAD5 deletion as well accelerates senescence (Fallet et al. 2014) suggesting that the error-free DNA damage tolerance (DDT) and template switch are engaged to prevent loss of viability due to unresolved replication stress at telomeres. Involvement of template switch in bypassing replication stress during senescence is consistent with the observations described above including the accumulation of X-shaped structures at telomeres during senescence and the dissolution of these sister chromatid junctions by Sgs1 (Lee et al. 2007; Branzei et al. 2008;

Karras and Jentsch 2010). It is also consistent with dependence of pre-senescing cells on DNA polymerase Pol (Lydeard et al. 2010; Fallet et al. 2014).

Alternatively, stalled forks might also be stabilized by Rad5-dependent regression (see Figure 3B) although it remains unclear whether regression of the replication forks occurs in checkpoint-proficient S. cerevisiae (Sogo et al. 2002).

Although the pathways responsible for resolution of the replication stress are likely to be shared between the telomeres and other regions of the genome, unique properties of the telomere and subtelomere sequences could impact on both the response to replication stress and the outputs of the repair. For example, it remains unclear to which extent RPA and CST

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(9)

would intersect or compete to bind internal and terminal TG1-3 sequences when single stranded (Gao et al. 2007; Lewis et al., 2013). Considering the central role of RPA not only in checkpoint activation but also in the choice of the repair pathway (Marechal and Zou 2015), very little is known whether CST shares or interferes with some of these functions. Another difference that is likely to impact on the resolution of replication stress at telomeres is their terminal position. One could imagine that uncoupling of leading and lagging strand synthesis could generate a long terminal G-tail (Audry et al., 2015) that would favor invasion and copy of either sister chromatid or other telomere in a process involving either Rad51- or Rad59- dependent BIR. This possibility is supported by the synthetic growth defect between rad52 and rad5 deletions in the absence of telomerase (Fallet et al. 2014).

Factors influencing telomere replication in the absence of telomerase

As mentioned above, telomeres and subtelomeres are intrinsic obstacles for replication-forks in both yeast (Ivessa and Zakian 2002; Makovets et al. 2004) and mammals (Sfeir et al. 2009).

The precise hitch to telomere replication is not fully defined. In S. cerevisiae, the telomeric repeats are assembled into a compact nucleoprotein structure maintained by the Rap1 as well as Sir and Rif proteins. Subtelomeric nucleosomes at X and Y’ elements also differ from nucleosomes in most of the others regions of the genome by the binding of Sir proteins that impose transcriptional repression and hetero-chromatization (Kueng et al. 2013). Tight binding of Rap1 to telomere repeats likely creates a barrier to fork progression (Makovets et al. 2004). In addition, due to the uni-directionality of telomere replication, the G-rich strand is replicated by the lagging strand machinery ie by the DNA polymerase  that tends to stall on telomeric G-rich repeats (Lormand et al. 2013; Audry et al. 2015). Concomitantly, chromatin structure might be another actor of either the replication stress or the consecutive recombination at telomeres. Senescence is accompanied by an increase in subtelomeric H4K16 acetylation and inactivation of the H4K16 acetylase Sas2 delays senescence and stimulates HR-mediated telomere elongation (Kozak et al. 2010).

It is possible that modification of the telomeric structure as telomeres shorten in the absence of telomerase further aggravates the replication stress at telomeres. It has been shown that the S. pombe Taz1 protein, which binding is expected to be decreased at short telomeres, facilitates telomere replication (Dehe et al., 2012). Another factor with a potential to impair the passage of the replication fork through subtelomere and telomere sequences is the

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(10)

interference between the replication fork and the presence of TERRA transcripts. TERRA are long non-coding heterogeneous in length RNA transcribed in the 5’-3’ direction from the subtelomeric region toward chromosome ends in yeast as well as in mammals (Maicher et al.

2014) and which transcription is increased at short telomeres in both S. cerevisiae and S.

pombe (Iglesias et al. 2011; Pfeiffer and Lingner 2012; Cusanelli et al. 2013; Moravec et al., 2016). The level of TERRA transcripts is very low in wild type cells. TERRA transcription is repressed by the Sir complex and Rif1/2 at X-only telomeres but mainly by Rap1-associated Rif1/2 at telomeres containing Y’ elements (Iglesias et al. 2011). The level of TERRA is further controlled by the 5’-3’ RNA exonuclease Rat1 (Luke et al. 2008) and RNase H suggesting that TERRA forms RNA-DNA hybrids with the C-rich telomere strand (Iglesias et al. 2011). Increased production of TERRA upon telomere shortening might be the consequence of a loss of Rif repression associated with the decreased number of telomere repeats. Accumulation of TERRA RNA/DNA hybrids correlates with an increased frequency of telomere recombination in the absence of telomerase (Balk et al. 2013; Yu et al. 2014).

Conversely, overexpression of RNH1 decreases telomeric recombination (Balk et al. 2013), an effect in line with the ability of Rnh1 to decrease Transcription-Associated recombination (TAR) (Huertas and Aguilera 2003; El Hage et al. 2010). Further work will be needed to understand the mechanistic consequences of the presence of R-loops at telomeres. It would be for example crucial to determine whether the presence of R-loops enhances the replication fork pause at telomeres and subtelomeres and whether it is causatively linked to an accumulation of G-rich ssDNA.

Spatial regulation of telomere recombination during survivor formation

The fact that type I and II survivors appear at an estimated frequency of 10-4 or less (Lundblad and Blackburn 1993) suggests that they result from rare recombination events rather than dedicated pathways. Type I survivor formation has specific genetic requirements.

Rad59-dependent translocation of Y’ elements onto X-only telomeres constitutes a first step of their production (Churikov et al. 2014), while spreading and amplification of Y’ elements at all telomeres next requires Rad52, Rad51, Rad54 and Rad55 and Pol32 (Le et al. 1999;

Lydeard et al. 2007). This suggests that type I survivors are formed by successive rounds of Rad51-catalyzed strand invasion followed by migration of the D-loop and DNA synthesis up to the end of the chromosome, essentially by multiple rounds of Rad51-dependent BIR. In keeping with BIR mechanism, type I survivor formation also requires Pif1 (Hu et al. 2013),

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(11)

which is required for long-range synthesis via bubble migration (Saini et al. 2013; Wilson et al. 2013).

Recombination dependent telomere elongation (RTE) observed in type II survivors has drawn more attention because it shares similarity with the ALT (Alternative Lengthening of Telomere) phenotype that accounts for about 15% of human cancers (Pickett and Reddel 2015). RTE required the strand annealing function of Rad52 reinforced by Rad59 as well as components of the BIR pathway (McEachern and Haber 2006; Lydeard et al. 2010). In contrast to Y’ amplification, which is initiated early after telomerase inactivation, the sudden amplification of TG1-3 repeats takes place only after senescence (Chang et al. 2011; Teng et al.

2000). Most probably, RTE occurs at eroded telomeres that are recognized and processed as DSBs, an assumption in accord with type II recombination dependence on activation of the DNA damage checkpoint (Grandin and Charbonneau 2007a). Homologous recombination at DSBs is initiated by resection of the DNA 5’-ends initiated by MRX (Mre11-Rad50-Xrs2) and Sae2 and extended by the exonuclease Exo1 and/or by the combined action of Dna2 and Sgs1 (Symington, 2014). Inactivation of any of these proteins strongly affects formation of type II survivors (Huang et al. 2001; Bertuch and Lundblad 2004; Hardy et al. 2014; Johnson et al. 2001; Maringele and Lydall 2002).

Another level of regulation of telomere recombination is related to the spatial control of DNA damage repair. In wild type cells, telomeres cluster in 6 to 8 foci at the nuclear periphery in the recombination prohibitive zone enriched in Sir proteins during most of the cell cycle (Gotta et al. 1996). Perinuclear telomere localization depends on two redundant pathways involving Sir4 and the yKu70/80 complex, which interact with the nuclear membrane anchored proteins Esc1 and Mps3, respectively (Hediger et al. 2002; Taddei et al.

2004; Taddei et al. 2010). Although the telomeres remain by and large clustered during senescence, a subset of eroded recombinogenic telomeres revealed by foci of colocalization of Cdc13 and Rad52, relocate from their membrane anchor to the Nuclear Pore Complex (NPC) (Khadaroo et al. 2009). Specialization of the nuclear compartments for specific repair and/or protection has emerged as a major actor of HR regulation (Geli and Lisby 2015). It has long been though that NPCs play a role in tethering hard-to-repair DNA damage, such as DSBs without a donor template for HR, collapsed replication forks (Nagai et al. 2008), and eroded telomeres (Khadaroo et al. 2009). This has been recently extended to tandem CAG repeats prone to secondary structures, which localize to NPCs during their replication in repeat array size-dependent manner (Su et al. 2015). Similarly to other types of damages (Nagai et al.

2008; Su et al. 2015; Horigome et al. 2016), localization of telomeres to NPC, as well as

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(12)

formation of type II survivors, depends on the small ubiquitin-like modifier (SUMO)-targeted ubiquitin ligase (STUbL) Slx5-Slx8 (Churikov et al. 2016). Accordingly, SUMOylation of proteins bound to telomeres increases as telomeres shorten in the absence of telomerase and correlates with the recruitment of STUbL to telomeres (Churikov et al. 2016). Since Slx5- Slx8 interacts with the pore protein Nup84 (Nagai et al. 2008), it became apparent that it may tether the telomeres to NPC. The current model for spatial regulation of type II recombination is depicted in the figure 4. A challenging but important task for the future will be to identify the proteins, which SUMOylation regulates NPC localization and the ones that are deSUMOylated or degraded at the pores. Although multiple proteins are likely involved, one prominent candidate is RPA, which (1) binds resected telomeres, (2) is SUMOylated in response to DNA damage and telomere shortening (Cremona et al. 2012; Psakhye and Jentsch 2012; Churikov et al. 2016), and (3) constitutively interacts with Slx5 (Churikov et al. 2016), therefore capable to stabilize transient interactions of the STUbL with polySUMOylated substrates.

Perhaps most enigmatic aspect of type II recombination is the nature of the template for telomere elongation that occurs at a time where the reserve of telomeric sequence is exhausted.

The finding that telomere localization to NPC promotes type II recombination implies that such template might be enriched at the NPCs. The presence of extrachromosomal telomeric circles (t-circles) in type II survivors (Larrivee and Wellinger 2006), K. lactis mutants (Basenko et al. 2009) and ALT cancer cells (Henson et al. 2009) led to a model of roll-and- spread mechanism for fast amplification of telomeric repeats (Natarajan and McEachern, 2002; McEachern and Haber 2006). It is thus tempting to speculate that the t-circles are trapped to the NPC as it has been shown for other circular byproducts of recombination (Denoth-Lippuner et al. 2014).

Concluding remarks.

Unlike most human somatic cells, budding yeast constitutively express telomerase, and therefore, did not evolve specific adaptations to sustain telomere maintenance in the absence of telomerase. As a result, inactivation of telomerase via genetic manipulations in budding yeast not only recapitulates the process of replicative senescence observed in human somatic cells, but might also reveal those problems with telomere maintenance that are either counteracted by specific adaptive mechanisms in human cells or are masked by the action of telomerase in yeast. One such problem appears to be related to replication stress at telomeres,

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(13)

which is revealed immediately after inactivation of telomerase and concomitantly with the dependence of yeast growth on checkpoint mediators, recombination factors and post- replication repair. Therefore, telomerase itself is essential for repair of damage resulting from replication stress at telomeres, whereas multiple repair mechanisms are likely required to compensate for the absence of telomerase. In fact, unleashed recombination observed at the telomeres in yeast growing in the absence of telomerase appears to stem from the unresolved replication stress. Similar relationship between replication stress and recombination at telomeres likely exist in human cancers that bypassed senescence barrier and activated recombination-based telomere maintenance mechanism known as ALT and possibly in some genetic diseases linked to accelerated aging. Thus, telomerase-negative yeast serve as a useful model for deciphering the causes and consequences of replication stress at telomeres and its bearing on telomere maintenance by recombination.

FUNDING

VG is supported by the ‘‘Ligue Nationale Contre le Cancer’’ and by ‘‘L’Institut National Du Cancer, « TELOCHROM and PLBIO14-012 ». Dmitri Churikov is supported by the SIRIC (INCa-DGOS-Inserm 6038 grant).

References

Abdallah P, Luciano P, Runge KW et al. A two-step model for senescence triggered by a single critically short telomere. Nat Cell Biol 2009;11:988-93.

Alcasabas AA, Osborn AJ, Bachant J et al. Mrc1 transduces signals of DNA replication stress to activate Rad53. Nat Cell Biol 2001;3:958-65.

Audry J, Maestroni L, Delagoutte E et al. RPA prevents G-rich structure formation at lagging-strand telomeres to allow maintenance of chromosome ends. EMBO J 2015;34:1942- 58.

Balk B, Maicher A, Dees M et al. Telomeric RNA-DNA hybrids affect telomere-length dynamics and senescence. Nat Struct Mol Biol 2013;20:1199-1205.

Basenko EY, Cesare AJ, Iyer S et al. Telomeric circles are abundant in the stn1-M1 mutant that maintains its telomeres through recombination. Nucleic Acids Res 2009;38:182-89.

Bertuch AA, Lundblad V. EXO1 contributes to telomere maintenance in both telomerase- proficient and telomerase-deficient Saccharomyces cerevisiae, Genetics 2004;166:1651-1659.

Blackburn EH, Collins K. Telomerase: an RNP enzyme synthesizes DNA. Cold Spring Harb Perspect Biol 2011;3. doi: 10.1101/cshperspect.a003558.

Bonetti D, Clerici M, Anbalagan S et al. Shelterin-like proteins and Yku inhibit nucleolytic processing of Saccharomyces cerevisiae telomeres. PLoS Genet 2010;6: doi:

10.1371/journal.pgen.1000966.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(14)

Branzei D, Vanoli F, Foiani M. SUMOylation regulates Rad18-mediated template switch.

Nature 2008;456:915-20.

Carr AM, Lambert S. Relication stress-induced genome instability: the dark side of replication maintenance by homologous recombination. J Mol Biol 2013;425:4733-44.

Chang M, Dittmar JC, Rothstein R. Long telomeres are preferentially extended during recombination-mediated telomere maintenance. Nat Struct Mol Biol 2011;18:451-56.

Chavez A, George V, Agrawal V et al. Sumoylation and the structural maintenance of chromosomes (Smc) 5/6 complex slow senescence through recombination intermediate resolution. J Biol Chem 2010;285:11922-30.

Chen Q, Ijpma A, Greider CW. Two survivor pathways that allow growth in the absence of telomerase are generated by distinct telomere recombination events. Mol Cell Biol 2001;

21:1819-27.

Churikov D, Charifi F, Eckert-Boulet N et al. SUMO-Dependent Relocalization of Eroded Telomeres to Nuclear Pore Complexes Controls Telomere Recombination. Cell Rep.

2016;15:1242-53.

Churikov D, Charifi F, Simon MN et al. Rad59-facilitated acquisition of Y' elements by short telomeres delays the onset of senescence. PLoS Genet 2014;10:doi:

10.1371/journal.pgen.1004736.

Churikov D, Corda Y, Luciano P et al. Cdc13 at a crossroads of telomerase action. Front Oncol 2013;3:39. doi: 10.3389/fonc.2013.00039.

Costelloe T, Louge R, Tomimatsu N et al. The yeast Fun30 and human SMARCAD1 chromatin remodellers promote DNA end resection. Nature 2012;489:581-84.

Cremona CA, Sarangi P, Yang Y et al. Extensive DNA damage-induced sumoylation contributes to replication and repair and acts in addition to the mec1 checkpoint. Mol Cell 2012;45:422-32.

Cusanelli E, Romero CA, Chartrand P. Telomeric noncoding RNA TERRA is induced by telomere shortening to nucleate telomerase molecules at short telomeres. Mol Cell 2013;51:

780-91.

de Bruin D, Kantrow SM, Liberatore RA et al. Telomere folding is required for the stable maintenance of telomere position effects in yeast. Mol Cell Biol 2000;20:7991-8000.

Dehe PM, Rog O, Ferreira MG et al. Taz1 enforces cell-cycle regulation of telomere synthesis. Mol Cell 2012;46:797-808.

Denchi EL, de Lange T. Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 2007;448:1068-1071.

Denoth-Lippuner A, Krzyzanowski MK, Stober C et al. Role of SAGA in the asymmetric segregation of DNA circles during yeast ageing. Elife 2014;3:300-25.

Dewar JM, Lydall D. Similarities and differences between "uncapped" telomeres and DNA double-strand breaks. Chromosoma 2012;121:117-30.

Doksani Y, Wu JY, de Lange T et al. Super-resolution fluorescence imaging of telomeres reveals TRF2-dependent T-loop formation. Cell 2013;155:345-56.

Duro E, Vaisica JA, Brown GW et al. Budding yeast Mms22 and Mms1 regulate homologous recombination induced by replisome blockage. DNA Repair (Amst) 2008;7:811- 18.

El Hage A, French SL, Beyer AL et al. Loss of Topoisomerase I leads to R-loop-mediated transcriptional blocks during ribosomal RNA synthesis. Genes Dev 2010;24:1546-58.

Enomoto S, Glowczewski L, Berman J. MEC3, MEC1, and DDC2 are essential components of a telomere checkpoint pathway required for cell cycle arrest during senescence in Saccharomyces cerevisiae. Mol Biol Cell 2002;13:2626-38.

Fallet E, Jolivet P, Soudet J et al. Length-dependent processing of telomeres in the absence of telomerase. Nucleic Acids Res. 2014;42:3648-65.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(15)

Forstemann K, Hoss M, Lingner J. Telomerase-dependent repeat divergence at the 3' ends of yeast telomeres. Nucleic Acids Res 2000;28:2690-94.

Gao H, Cervantes RB, Mandell EK et al. RPA-like proteins mediate yeast telomere function. Nat Struct Mol Biol. 2007;14:208-214.

Geli V, Lisby M. Recombinational DNA repair is regulated by compartmentalization of DNA lesions at the nuclear pore complex. Bioessays 2015;37:1287-92.

Geronimo CL, Zakian VA. Getting it done at the ends: Pif1 family DNA helicases and telomeres. DNA Repair (Amst). 2016; DOI 10.1080/15476286.2016.1191725

Gilson E, Geli V. How telomeres are replicated. Nat Rev Mol Cell Biol 2007;8:825-38.

Giraud-Panis MJ, Teixeira MT, Geli V et al. CST meets shelterin to keep telomeres in check. Mol Cell 2010;39:665-76.

Gonzalez-Prieto R, Munoz-Cabello AM, Cabello-Lobato MJ et al. Rad51 replication fork recruitment is required for DNA damage tolerance. EMBO J 2013;32:1307-21.

Gotta M, Laroche T, Formenton A et al. The clustering of telomeres and colocalization with Rap1, Sir3, and Sir4 proteins in wild-type Saccharomyces cerevisiae. J Cell Biol 1996;

134:1349-63.

Grandin N, Bailly A, Charbonneau M. Activation of Mrc1, a mediator of the replication checkpoint, by telomere erosion. Biol Cell 2005;97:799-814.

Grandin N, Charbonneau M. Control of the yeast telomeric senescence survival pathways of recombination by the Mec1 and Mec3 DNA damage sensors and RPA. Nucleic Acids Res 2007a;35:822-38.

Grandin N, Charbonneau M. Mrc1, a non-essential DNA replication protein, is required for telomere end protection following loss of capping by Cdc13, Yku or telomerase. Mol Genet Genomics 2007b;277:685-99.

Greetham M, Skordalakes E, Lydall D et al. The Telomere Binding Protein Cdc13 and the Single-Stranded DNA Binding Protein RPA Protect Telomeric DNA from Resection by Exonucleases. J Mol Biol 2015;427:3023-30.

Griffith JD, Comeau L, Rosenfield S et al. Mammalian telomeres end in a large duplex loop. Cell 1999;97:503-14.

Hardy J, Churikov D, Geli V et al. Sgs1 and Sae2 promote telomere replication by limiting accumulation of ssDNA. Nat Commun 2014;5:5004.

Hashimoto Y, Ray Chaudhuri A, Lopes M et al. Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis. Nat Struct Mol Biol 2010 17:1305-11.

Hediger F, Neumann FR, Van Houwe G et al. Live imaging of telomeres: yKu and Sir proteins define redundant telomere-anchoring pathways in yeast. Curr Biol 2002;12:2076-89.

Henson JD, Cao Y, Huschtscha LI et al. DNA C-circles are specific and quantifiable markers of alternative-lengthening-of-telomeres activity. Nat Biotechnol 2009;27:1181-85.

Horigome C, Bustard DE, Marcomini I et al. PolySUMOylation by Siz2 and Mms21 triggers relocation of DNA breaks to nuclear pores through the Slx5/Slx8 STUbL. Genes Dev 2016;30:931-45.

Hu Y, Tang HB, Liu NN et al. Telomerase-null survivor screening identifies novel telomere recombination regulators. PLoS Genet 2013;9: doi: 10.1371/journal.pgen.1003208.

Huang P, Pryde FE, Lester D et al. SGS1 is required for telomere elongation in the absence of telomerase.2001;11:125-29.

Huertas P, Aguilera A. Cotranscriptionally formed DNA:RNA hybrids mediate transcription elongation impairment and transcription-associated recombination. Mol Cell 2003;12:711-721.

Iglesias N, Redon S, Pfeiffer V et al. Subtelomeric repetitive elements determine TERRA regulation by Rap1/Rif and Rap1/Sir complexes in yeast. EMBO Rep 2011;12:587-93.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(16)

IPjma AS, Greider CW. Short telomeres induce a DNA damage response in Saccharomyces cerevisiae. Mol Biol Cell 2003;14:987-1001.

Irmisch A, Ampatzidou E, Mizuno K et al. Smc5/6 maintains stalled replication forks in a recombination-competent conformation. EMBO J 2009;28:144-155.

Ivessa AS, Zakian VA. To fire or not to fire: origin activation in Saccharomyces cerevisiae ribosomal DNA. Genes Dev 2002;16:2459-64.

Ivessa AS, Zhou JQ, Schulz VP et al. Saccharomyces Rrm3p, a 5' to 3' DNA helicase that promotes replication fork progression through telomeric and subtelomeric DNA. Genes Dev 2002;16:1383-96.

Jay KA, Smith DL, Blackburn E. Early Loss of Telomerase Action in Yeast Creates a Dependence on the DNA Damage Response Adaptor Proteins. Mol Cell Biol. 2016;36:1908- 19.

Johnson FB, Marciniak RA, McVey M et al. The Saccharomyces cerevisiae WRN homolog Sgs1p participates in telomere maintenance in cells lacking telomerase. EMBO J.

2001;20:905-913.

Karras GI, Jentsch S. The RAD6 DNA damage tolerance pathway operates uncoupled from the replication fork and is functional beyond S phase. Cell 2010;141:255-67.

Khadaroo B, Teixeira MT, Luciano P et al. The DNA damage response at eroded telomeres and tethering to the nuclear pore complex. Nat Cell Biol 2009;11:980-87.

Koundrioukoff S, Carignon S, Techer H et al. Stepwise activation of the ATR signaling pathway upon increasing replication stress impacts fragile site integrity. PLoS Genet. 2013;9:

doi: 10.1371/journal.pgen.1003643.

Kozak ML, Chavez A, Dang W et al. Inactivation of the Sas2 histone acetyltransferase delays senescence driven by telomere dysfunction. EMBO J 2010;29:158-70.

Kueng S, Oppikofer M, Gasser SM. SIR proteins and the assembly of silent chromatin in budding yeast. Annu Rev Genet 2013;47:275-306.

Kupiec M. Biology of telomeres: lessons from budding yeast. FEMS Microbiol Rev 2014;

38:144-171.

Larrivee M, LeBel C, Wellinger RJ. The generation of proper constitutive G-tails on yeast telomeres is dependent on the MRX complex. Genes Dev 2004;18:1391-96.

Larrivee M, Wellinger RJ. Telomerase- and capping-independent yeast survivors with alternate telomere states. Nat Cell Biol 2006;8:741-47.

Le S, Moore JK, Haber JE et al. RAD50 and RAD51 define two pathways that collaborate to maintain telomeres in the absence of telomerase. Genetics 1999;152:143-52.

Lee JY, Kozak M, Martin JD et al. Evidence that a RecQ helicase slows senescence by resolving recombining telomeres. PLoS Biol 2007;5: e160.

Lewis KA, Pfaff DA, Earley JN, Altschuler SE, Wuttke DS. The tenacious recognition of yeast telomere sequence by Cdc13 is fully exerted by a single OB-fold domain. Nucleic Acids Res 2013;42:475-84.

Lormand JD, Buncher N, Murphy CT et al. DNA polymerase delta stalls on telomeric lagging strand templates independently from G-quadruplex formation. Nucleic Acids Res 2013;41:10323-33.

Louis EJ, Haber JE. The structure and evolution of subtelomeric Y' repeats in Saccharomyces cerevisiae. Genetics 1992;131:559-74.

Luke B, Versini G, Jaquenoud M et al. The cullin Rtt101p promotes replication fork progression through damaged DNA and natural pause sites. Curr Biol 2006;16:786-92.

Luke B, Panza A, Redon S et al. The Rat1p 5' to 3' exonuclease degrades telomeric repeat- containing RNA and promotes telomere elongation in Saccharomyces cerevisiae. Mol Cell 2008;32:465-77.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(17)

Lundblad V, Blackburn EH. An alternative pathway for yeast telomere maintenance rescues est1- senescence. Cell 1993;73:347-60.

Lydeard JR, Lipkin-Moore Z, Sheu YJ et al. Break-induced replication requires all essential DNA replication factors except those specific for pre-RC assembly. Genes Dev 2010;24:1133-44.

Magdalou I, Lopez BS, Pasero P et al. The causes of replication stress and their consequences on genome stability and cell fate. Semin Cell Dev Biol 2014;30:154-64.

Maicher A, Lockhart A, Luke B. Breaking new ground: digging into TERRA function.

Biochim Biophys Acta 2014;1839:387-94.

Makovets S, Herskowitz I, Blackburn EH. Anatomy and dynamics of DNA replication fork movement in yeast telomeric regions. Mol Cell Biol 2004;24:4019-31.

Marcand S, Brevet V, Gilson E. Progressive cis-inhibition of telomerase upon telomere elongation. EMBO J. 199;18:3509-19.

Marechal A, Zou L. RPA-coated single-stranded DNA as a platform for post-translational modifications in the DNA damage response. Cell Res. 2015;25:9-23.

Maringele L, Lydall D. EXO1-dependent single-stranded DNA at telomeres activates subsets of DNA damage and spindle checkpoint pathways in budding yeast yku70Delta mutants. Genes Dev. 2002;16:1919-1933.

McEachern MJ, Blackburn EH. A conserved sequence motif within the exceptionally diverse telomeric sequences of budding yeasts. Proc Natl Acad Sci U S A 1994;91:3453-57.

McEachern MJ, Haber JE. Break-induced replication and recombinational telomere elongation in yeast. Annu Rev Biochem 2006;75:111-35.

Mitton-Fry RM, Anderson EM, Hughes TR et al. Conserved structure for single-stranded telomeric DNA recognition. Science 2002;296:145-47.

Moravec M, Wischnewski H, Bah A et al. TERRA promotes telomerase-medited telomere elongation in Schizosaccharomyces pombe. EMBO rep. 2016;17:999-1012.

Mullen JR, Brill SJ. Activation of the Slx5-Slx8 ubiquitin ligase by poly-small ubiquitin- like modifier conjugates. J Biol Chem 2008;283:19912-21.

Nagai S, Dubrana K, Tsai-Pflugfelder M et al. Functional targeting of DNA damage to a nuclear pore-associated SUMO-dependent ubiquitin ligase. Science 2008;322:597-602.

Natarajan S, McEachern MJ. Recombinational Telomere Elongation Promoted by DNA Circles. Mol Cell Biol 2002;22:4512-21.

Niepel M, Molloy KR, Williams R et al. The nuclear basket proteins Mlp1p and Mlp2p are part of a dynamic interactome including Esc1p and the proteasome. Mol Biol Cell 2013;24:3920-38.

Noel JF, Wellinger RJ. The Smc5/6 complex and the difficulties cutting the ties of twin sisters. Aging (Albany NY) 2011;3:186-88.

O'Sullivan RJ, Karlseder J. Telomeres: protecting chromosomes against genome instability. Nat Rev Mol Cell Biol 2010;11:171-81.

Olovnikov AM. A theory of marginotomy. The incomplete copying of template margin in enzymic synthesis of polynucleotides and biological significance of the phenomenon. J Theor Biol 1973;41:181-90.

Osborn AJ, Elledge SJ. Mrc1 is a replication fork component whose phosphorylation in response to DNA replication stress activates Rad53. Genes Dev 2003;17: 1755-67.

Paeschke K, McDonald KR, Zakian VA. Telomeres: structures in need of unwinding.

FEBS Lett 2010;584:3760-72.

Pardo B, Marcand S. Rap1 prevents telomere fusions by nonhomologous end joining.

EMBO J 2005;24:3117-27.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(18)

Pfeiffer V, Lingner J. TERRA promotes telomere shortening through exonuclease 1- mediated resection of chromosome ends. PLoS Genet 2012;8:doi:

10.1371/journal.pgen.1002747.

Pickett HA, Reddel RR. Molecular mechanisms of activity and derepression of alternative lengthening of telomeres. Nat Struct Mol Biol 2015;22:875-80.

Poschke H, Dees M, Chang M et al. Rif2 promotes a telomere fold-back structure through Rpd3L recruitment in budding yeast. PLoS Genet 2012;8: doi: 10.1371/journal.pgen.1002960.

Psakhye I, Jentsch S. Protein group modification and synergy in the SUMO pathway as exemplified in DNA repair. Cell 2012;151:807-20.

Ribeyre C, Shore D. Anticheckpoint pathways at telomeres in yeast. Nat Struct Mol Biol 2012;19:307-13.

Rizki A, Lundblad V. Defects in mismatch repair promote telomerase-independent proliferation. Nature 2001;411:713-16.

Saini N, Ramakrishnan S, Elango R et al. Migrating bubble during break-induced replication drives conservative DNA synthesis. Nature 2013;502:389-92.

Sarangi P, Zhao X. SUMO-mediated regulation of DNA damage repair and responses.

Trends Biochem Sci 2015;40:233-42.

Sfeir A, Kosiyatrakul ST, Hockemeyer D et al. Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 2009;138:90-103.

Shampay J, Blackburn EH. Generation of telomere-length heterogeneity in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 1988;85:534-38.

Smith JS, Chen Q, Yatsunyk LA et al. Rudimentary G-quadruplex-based telomere capping in Saccharomyces cerevisiae. Nat Struct Mol Biol 2011;18:478-85.

Sogo JM, Lopez M, Foiani M. Fork reversal and ssDNA accumulation at stalled replication forks owing to checkpoint defects. Science 2002;297:599-602.

Soudet J, Jolivet P, Teixeira MT. Elucidation of the DNA end-replication problem in Saccharomyces cerevisiae. Mol Cell 2014;53:954-64.

Spell RM, Jinks-Robertson S. Role of mismatch repair in the fidelity of RAD51- and RAD59-dependent recombination in Saccharomyces cerevisiae. Genetics 2003;165:1733-44.

Su XA, Dion V, Gasser SM et al. Regulation of recombination at yeast nuclear pores controls repair and triplet repeat stability. Genes Dev 2015;29:1006-17.

Symington L. End resection at double-strand breaks : mechanism and regulation. Cold Spring Harb Perspect Biol. 2014;6: doi: 10.1101/cshperspect.a016436.

Taddei A, Hediger F, Neumann FR et al. Separation of silencing from perinuclear anchoring functions in yeast Ku80, Sir4 and Esc1 proteins. EMBO J 2004;23:1301-12.

Taddei A, Schober H, Gasser SM. The budding yeast nucleus. Cold Spring Harb Perspect Biol 2010;2: doi: 10.1101/cshperspect.a000612.

Teixeira MT, Arneric M, Sperisen P et al. Telomere length homeostasis is achieved via a switch between telomerase- extendible and -nonextendible states. Cell 2004;117:323-35.

Teng SC, Chang J, McCowan B et al. Telomerase-independent lengthening of yeast telomeres occurs by an abrupt Rad50p-dependent, Rif-inhibited recombinational process. Mol Cell 2000;6:947-952.

Tracy RB, Baumohl JK, Kowalczykowski SC. The preference for GT-rich DNA by the yeast Rad51 protein defines a set of universal pairing sequences. Genes Dev 1997;11:3423- 31.

Vialard JE, Gilbert CS, Green CM et al. The budding yeast Rad9 checkpoint protein is subjected to Mec1/Tel1-dependent hyperphosphorylation and interacts with Rad53 after DNA damage. EMBO J 1998;17:5679-88.

Vodenicharov MD, Laterreur N, Wellinger RJ. Telomere capping in non-dividing yeast cells requires Yku and Rap1. EMBO J 2010;29:3007-19.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(19)

Watt PM, Hickson ID, Borts RH et al. SGS1, a homologue of the Bloom's and Werner's syndrome genes, is required for maintenance of genome stability in Saccharomyces cerevisiae. Genetics 1996;144:935-45.

Wellinger RJ, Wolf AJ, Zakian VA. Saccharomyces telomeres acquire single-strand TG1-3 tails late in S phase. Cell 1993;72:51-60.

Wilson MA, Kwon Y, Xu Y et al. Pif1 helicase and Poldelta promote recombination- coupled DNA synthesis via bubble migration. Nature 2013;502:393-96.

Wotton D, Shore D. A novel Rap1p-interacting factor, Rif2p, cooperates with Rif1p to regulate telomere length in Saccharomyces cerevisiae. Genes Dev 1997;11:748-60.

Xie Z, Jay KA, Smith DL et al. Early telomerase inactivation accelerates aging independently of telomere length. Cell 2015;160:928-39.

Xu Z, Duc KD, Holcman D et al. The length of the shortest telomere as the major determinant of the onset of replicative senescence. Genetics 2013;194:847-57.

Xu Z, Fallet E, Paoletti C et al. Two routes to senescence revealed by real-time analysis of telomerase-negative single lineages. Nat Commun 2015;6:7680.

Yu TY, Kao YW, Lin JJ. Telomeric transcripts stimulate telomere recombination to suppress senescence in cells lacking telomerase. Proc Natl Acad Sci U S A 2014;111:3377- 3382.

Zhao X, Wu CY, Blobel G. Mlp-dependent anchorage and stabilization of a desumoylating enzyme is required to prevent clonal lethality. J Cell Biol 2004;167:605-11.

Legend of the figures:

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(20)

Figure 1. Replicative senescence in Saccharomyces cerevisiae. (A) In yeast, subtelomeric regions contain repeated sequences called X and Y’ elements. The core X element is present at all chromosome ends with limited sequence and size variations and is proximal to either terminal TG1-3 repeats or Y’ elements that are present in two third of the telomeres as 1 to 4 in tandem repeats (Louis and Haber 1992). Upon inactivation of telomerase, the telomere sequences shorten on average 3-5bp per population doubling (PD). Depending on the telomere length in the initial cell, telomeres reach a critical length (about 70 to 100bp) in about 50 to 60 PDs. Uncapped telomeres activate the DNA damage checkpoint leading to a permanent arrest in G2/M. G2/M arrested cells, often referred to as “monsters”, are characterized by a dumbbell morphology and the localization of the undivided nucleus at the bud neck. Among the population of G2/M arrested cells, few cells succeed in reconstituting functional telomeres by homologous recombination, recover growth capacity and give rise to post-senescence survivors. (B) A schematic curve of the growth capacity of the cells in liquid culture upon inactivation of telomerase is shown. The growth rate of the population starts to decline about 20 PDs after telomerase inactivation until it reaches a minimum after 50-60 PDs where the presence of “monsters” is at its maximum in the cell population (senescence).

Videomicroscopy of telomerase-minus cells shows an enrichment of cells undergoing transient cell cycle arrest during the initial period after telomerase inactivation (Xu et al.

2015). This period, defined here as pre-senescence, lasts until the cell population accumulates the highest proportion of monsters due to permanent G2/M arrest and is immediately followed by the appearance of survivors. C) Telomere organization in the survivors. Type I survivors are characterized by an amplification of Y’ elements separated by ITS and a very short array of terminal TG1-3 repeats. Type II survivors carry long (up to 10 kb) and heterogeneous extensions of telomere repeats.

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

(21)

Figure 2: Models of the short telomere repair in the absence of telomerase. Short telomeres generated as a result of either breakage or enzymatic cleavage of the replication forks can be repaired by BIR after completion of the replication. Depending on the location of the donor (terminal or interstitial TG1-3 repeats), BIR at the telomeres may result either in the elongation of the terminal TG1-3 tract (A) or in the non-reciprocal translocation of the Y’

element along with the terminal TG1-3 repeats on the recipient short telomere (B). Since heteroduplexes between TG1-3 tracts are expected to contain multiple mismatches, they are likely to be targeted by mismatch repair and prone to unwinding by Sgs1. Therefore multiple rounds of alignment, synthesis and dissociation could be expected during BIR at telomeres.

The involvement of Rad51 in recombination between short TG1-3 repeats remains uncertain (see text for details).

by guest on September 30, 2016http://femsyr.oxfordjournals.org/Downloaded from

Références

Documents relatifs

In the absence of rad51, only Type II survivors form and if MUS81 was required for Type II growth, the combined est2Δ mus81Δ rad51Δ triple mutant would be lethal as cells would

evaluated joining single rupture forces obtained using several AFM probes over different ex- perimental sessions. Since the total number of molecules N elastically compressed by the

Al- though Digital Mammography (DM) remains the reference imaging modal- ity, Ultra-Sound (US) imaging has proven to be a successful adjunct image modality for breast cancer

fraction-line proportion matching task, b) the line proportion comparison task, c) the symbolic fraction magnitude 357. comparison task, and d) the scrambled letters

In order to study replication initiation in plasmids, the two sequences ORI3018 and ORI1068 were cloned in a SalI-linearized pINA732 vector (a centromere- containing plasmid; see

Conversely, AKR1C2 overexpression increased the frequency of PSNE, therefore validating the PSNE signature as a signature of the process of neoplastic emergence by senescence evasion

- Group P2 and P6 (23 strains; 7 from sourdough) correspond to Mixed origin clade which also include beer strains, clinical strains, water, fruits, tree leaves and natural