• No results found

University of Groningen On the molecular biology of telomeres Stinus Ruiz de Gauna, Sonia

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen On the molecular biology of telomeres Stinus Ruiz de Gauna, Sonia"

Copied!
23
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

On the molecular biology of telomeres

Stinus Ruiz de Gauna, Sonia

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Stinus Ruiz de Gauna, S. (2018). On the molecular biology of telomeres: Lessons from budding yeast.

University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Chapter 5

(3)

Chapter 5

Despite the remarkable advances science has made over the last century, many fundamental

questions remain unanswered. In this thesis, the aim is to contribute to a better understanding

of telomere biology. Specifically, we examine how a cell is able to distinguish between a

double-stranded DNA break (DSB) and a telomere, which resemble each other in structure

and share some proteins, but need to be processed in very different ways. The relevance of

the adequate distinction between DSBs and telomeres becomes apparent when looking at

the consequences of a failure in this process. A DSB is very deleterious for yeast cells, to the

extent that one unrepaired DSB is fatal (Resnick and Martin, 1976; Weiffenbach and Haber,

1981). Therefore, tightly regulated repair mechanisms ensure the proper repair of DSBs.

However, if a DSB is improperly recognized as a chromosome end and it is repaired as

such, telomeric sequence will be added at the broken site. Because genes nearby telomeric

sequence are transcriptionally silenced, addition of telomeric sequence at a DSB will lead to

silencing of genetic information. Moreover, two human diseases have been associated with

truncated chromosomes where telomeric sequence addition occurred, mental retardation

and alpha thalassemia (Wilkie et al., 1990; Wong et al., 1997), highlighting the relevance of

fundamental research for human health. The opposite scenario, where a telomere is treated

as a DSB, also has devastating consequences. Telomere shortening is counteracted by the

specialized reverse transcriptase telomerase (Greider and Blackburn, 1985). It is therefore

very important to keep chromosome ends available for telomerase-dependent extension

and, at the same time, to inhibit the repair mechanisms that act at DSBs. Failure to inhibit

DSB repair results in telomere-telomere fusions that can lead to cell death or

breakage-fusion-bridge cycles, which are thought to be important in the formation of some cancers

(Selvarajah et al., 2006).

A ~40 nt length threshold separates telomeres from DSBs

To answer how a cell discerns between a DSB and a telomere, in chapter 2, we have used

two complementary methods. First, we used artificially constructed DNA ends adjacent to

telomeric sequence of increasing length, and monitored whether such ends are repaired

by telomerase-dependent extension. Second, we developed the iSTEX assay, an inducible

method that allows the

in vivo monitoring of telomerase-dependent telomere extension

events at nucleotide resolution and used it to determine what natural ends were sensed

as telomeres and therefore extended by telomerase. Both methods uncovered a

length-dependent threshold of about 40 nt of telomeric sequence that establishes the difference

between telomeres and DSBs. Below the 40 nt threshold, DNA ends contain too little

telomeric sequence and are treated as DSBs and, above it, DNA ends are extended by

telomerase. We have also found what proteins are important for the establishment of this

length-dependent threshold. Pif1 is important to inhibit telomerase at DNA ends below

the DSB-telomere threshold, as evidenced by the increased telomere addition frequency

at TG

18

ends in

pif1-m2 cells. On the other hand, the telomerase recruitment function of

(4)

Discussion and future perspectives

93

Chapter 5

DNA ends above the threshold.

An important question that remains unanswered is what is intrinsically different

between the DNA ends below and above the DSB-telomere threshold. One possibility

is that Cdc13 is recruited less efficiently to DNA ends below the threshold.

In vitro, the

minimum binding site of Cdc13 is 11 nt in length (Hughes et al., 2000), so Cdc13 could

have enough space to bind a TG

18

end; however, the situation in a living cell might be

different. For example, Cdc13 might not be recruited to such short ends. Alternatively, one

Cdc13 molecule might be recruited, but multiple Cdc13 molecules might be required for

telomere extension. Molecular crowding at such a short end could also affect the binding

or function of Cdc13

in vivo. Another possibility is that Cdc13 recruitment at DNA ends

above the threshold overcomes the telomerase inhibition by Pif1. In this case, when Cdc13

is mutated, telomerase recruitment to the TG

34

end might be lost. It is also possible that

Cdc13 recruitment is unaffected at both sides of the threshold, in which case we can

envision at least two options. One is that Pif1 inhibition of telomerase is stronger at DNA

ends below than above the threshold. However, we know that Pif1 inhibits telomerase

both below and above the threshold, indicating that, unless the strength of inhibition is

different below and above the threshold, most likely Cdc13, and not Pif1, is the key player.

Alternatively, it will be important to determine whether the other members of the CST

complex, Stn1 and Ten1, play a role in the establishment of the DSB-telomere threshold.

In short, determining what proteins bind to DNA ends below and above the threshold

would help clarify how telomeres and DSBs are differently recognized and processed.

Development of the inducible STEX assay

Besides defining the length-dependent DSB-telomere threshold, a major contribution

of chapter 2 is the development of the inducible STEX assay (iSTEX). So far, telomere

length has been mostly studied by methods that only provide information about the

average telomere length of a population of cells. However, iSTEX provides information

about telomere extension events on single telomeres at nucleotide resolution. iSTEX

can determine to which telomeres telomerase is recruited and, therefore, whether the

recruitment of telomerase to the shortest telomeres or the recruitment in general is affected.

It also informs about the length of the newly added sequence, and whether the nucleotide

addition processivity or the repeat addition processivity are affected. All these aspects are

informative to understand what is the cause of telomere length change at a molecular level.

Moreover, iSTEX has two improvements with respect to the classic STEX assay

(Teixeira et al., 2004). First, the use of an inducible system largely facilitates the execution

of the experiment because it eliminates the need of the nearly 100% mating efficiency

between the donor (telomerase positive) and the recipient (telomerase negative) strains

required in the original STEX assay (Teixeira et al., 2004). This level of mating efficiency

is rarely observed. Second, the use of the

tlc1-tm template mutant ensures that all telomere

(5)

Chapter 5

classic STEX relies on the imperfect nature of the telomere repeats added by the yeast

telomerase. The telomeric sequences are aligned to a reference sequence. The sequence of

the extended telomeres does not align to the non-extended ones, in an otherwise clonal

population. Those misalignments are called divergent events and are the readout for

telomere extension events. However, telomere divergence events can arise from artefacts

generated during PCR amplification, cloning and sequencing (Claussin and Chang, 2016).

To overcome this issue, iSTEX uses a telomerase template different from the wild type,

ensuring that all telomere extension events are dependent on telomerase and, therefore,

avoiding false positives. Moreover, iSTEX excludes telomeric divergence events arising

from recombination. Deletion of

RAD52 is the common way to exclude recombination

events but, when combined with telomerase-negative cells, results in accelerated replicative

senescence (Le et al., 1999), which can render the assay more challenging. Because

tlc1-tm

telomeric sequence is not present in the cell before the start of the experiment, deletion

of

RAD52 is not required to exclude recombination events. It is formally possible that a

recombination event takes place once the telomeres have been extended by the mutant

telomerase. This is however unlikely, since the experiment allows only one cell cycle, and it

would in any case represent a negligible fraction of the total divergent events.

Cdc13-independent telomerase recruitment to chromosome ends

In the course of the studies described in chapter 2 we found that, while

cdc13-2 mutation

inhibits telomere addition at TG

34

ends, as expected, further mutation of

PIF1 allows

telomerase-dependent extension of these ends. This is striking because the

cdc13-2 mutant

carries a point mutation in the telomerase recruitment domain, which makes it behave

like a telomerase-null mutant (Nugent et al., 1996). In fact, recruitment of the telomerase

subunits Est1 and Est2 to the telomeres has been recently shown to be largely impaired

in

cdc13-2 cells (Chen et al., 2018). This suggests that telomerase can be recruited in a

Cdc13-independent manner to extend the telomeres, probably through the Yku complex,

which can interact with the telomerase RNA subunit (Stellwagen et al., 2003) and has been

proposed to recruit telomerase (Hass and Zappulla, 2015). Alternatively, the interaction

between Cdc13 and Est1 might not be completely lost in

cdc13-2 cells, and mutation of

PIF1 allows just enough telomerase recruitment to prevent senescence.

Preliminary studies to understand how those

cdc13-2 pif1-m2 telomeres are

maintained show that, contrary to

cdc13-2 cells, cdc13-2 pif1-m2 cells do not senesce and,

although shorter than in

pif1-m2 cells, telomeres are stable in length (data not shown).

Non-essential G-quadruplex-mediated telomere protection

Chapter 4 explores the

in vivo role of yeast telomeric G-quadruplexes. To do so, we

made use of the

tlc1-tm mutant because of the characteristic telomeric sequence that

(6)

Discussion and future perspectives

95

Chapter 5

formation of G-quadruplexes. Our

in vitro and in vivo experiments (circular dichroism and

viability rescue experiments of telomere capping-deficient

cdc13-1 strains upon stabilization

of G-quadruplexes) are in line with the hypothesis that these telomeres are impaired in

G-quadruplex formation. However, these experiments do not rule out the possibility that

the

tlc1-tm phenotype is unrelated to G-quadruplex formation. On the one hand, although

widely used in G-quadruplex studies, circular dichroism is an

in vitro method to measure the

ability of oligonucleotides to fold into G-quadruplex structures, which does not necessarily

represent the

in vivo situation. On the other hand, G-quadruplex stabilization was achieved

by deletion of

PIF1. PIF1 deletion was chosen to stabilize G-quadruplexes because it is the

best G-quadruplex unwinding helicase described so far (Paeschke et al., 2013). However,

the many functions that Pif1 carries out complicate the interpretation of the results, as

discussed in chapter 4. Other methods to stabilize G-quadruplexes, like deletion of the

SGS1 helicase (Sun et al., 1999), overexpression of the G-quadruplex binding protein

Stm1 (Hayashi and Murakami, 2002) or addition of the G-quadruplex stabilizing ligand

PhenDC (Piazza et al., 2010) resulted in inconclusive results (data not shown). The use

of a G-quadruplex specific antibody (Biffi et al., 2013) combined with G-quadruplex

stabilizing ligands and chromatin immunoprecipitation could provide more robust data

about the G-quadruplex forming potential of these telomeres. Yet, our data support the

previously reported rudimentary telomere protection role by G-quadruplexes when Cdc13

is affected (Smith et al., 2011). The fitness of

tlc1-tm cells, indistinguishable from wild-type

cells, indicates that the G-quadruplex-mediated telomere capping function is not essential.

The question arises then, why are telomeric G-quadruplexes conserved? If

organisms of different complexities, from ciliates to humans, have telomeric G-quadruplexes,

it is reasonable to think that they have been evolutionary conserved for a reason. Although

G-quadruplex structures do not seem to play an indispensable role when it comes to

telomere protection, they might influence other telomeric aspects. One could also speculate

that G-quadruplexes were ancient chromosome end protecting structures until telomeres

became specialized, then got obsolete and only worked as a backup telomere capping

mechanism. However, it seems unlikely that such complex structures, as G-quadruplexes

are, were only conserved to work when other telomere protecting structures were lacking.

Characterisation of a Rap1-free telomere

tlc1-tm telomeres show a number of interesting characteristics, the most remarkable probably

being that

tlc1-tm telomeres are not bound by the major telomere binding protein Rap1.

Accordingly, telomere length homeostasis of

tlc1-tm telomeres is altered, leading to long

and heterogeneous sized telomeres. However, although longer than wild-type telomeres,

tlc1-tm telomeres are shorter than those obtained in cells harbouring a RAP1 C-terminal

deletion or

RIF1 RIF2 double deletion (Kyrion et al., 1992; Wotton and Shore, 1997),

which might indicate the presence of residual Rap1 at mutant telomeres. A related question

is how

tlc1-tm telomeres remain stable in length. This might be achieved by Rif1, which can

(7)

Chapter 5

bind the telomeric DNA in a Rap1-independent manner (Mattarocci et al., 2017), or by

the combination of telomerase-dependent extension, facilitated by the absence of Rap1,

and rapid telomere erosion, due to the lack of telomere protection. This is in line with the

strongly increased telomere extension frequency of

tlc1-tm telomeres and the very rapid

senescence and telomere shortening very early after telomerase depletion, that can be

explained by the absence of Rap1 at these telomeres, rendering them unprotected and very

dependent on telomerase-mediated extension. Interestingly, when

tlc1-tm telomeric repeats

are placed internally, followed by wild-type repeats at the distal ends, the bulk telomere

length increases approximately to the same extent as the length of the internally placed

mutant tract. Therefore, the mutant repeats are not sensed by the protein-counting model

that regulates telomere length (Marcand et al., 1997). Moreover, the fact that Rap1-free

tlc1-tm cells are perfectly viable implies that the Rap1 essential function is not its telomeric

function.

Rap1 binding to

tlc1-tm telomeres was determined by chromatin immunoprecipitation

followed by qPCR. However, because

tlc1-tm telomeres are long, the chromatin shearing

process required prior to immunoprecipitation could affect the result. Generation of

shorter

tlc1-tm telomeres by, for example, further mutation of TEL1, could help overcome

this problem. In addition, electrophoretic mobility shift assays (EMSA) with purified Rap1

would further confirm that Rap1 does not bind

tlc1-tm sequence. Mutagenizing the DNA

binding domain of Rap1 to find a Rap1 mutant able to bind

tlc1-tm sequence would be

useful for a complementation assay, which would confirm whether the absence of Rap1

is responsible for the observed phenotype. Determining the protein composition of

tlc1-tm telomeres would be also important. In the absence of Rap1, the recruitlc1-tment of Rif1

and Rif2 (Hardy et al., 1992; Wotton and Shore, 1997) and the components of the Sir

complex (Moretti et al., 1994; Moretti and Shore, 2001) should be affected. We plan to

use biotinylated oligonucleotides with wild-type or mutant sequence incubated with yeast

protein extracts, followed by mass spectrometry to unbiasedly identify what proteins

are bound to mutant telomeres. Assessing telomere fusion occurrence would also be of

interest, since Rap1 protects chromosome ends from NHEJ events (Marcand et al., 2008).

In short, the finding of a Rap1-free telomere opens many questions. Probably the most

important one that needs to be addressed is how telomeres are regulated in the absence of

their major binding protein.

The general aim of the work presented in this thesis was to determine what

minimally constitutes a telomere in

S. cerevisiae. We propose that Cdc13 and a telomeric

sequence of about 40 nt or longer are essential elements of telomeres. G-quadruplexes, on

the other hand, do not seem to play an important role at the chromosome ends. As often

happens, while trying to find an answer, many new questions arise, leaving plenty of room

for further research.

(8)
(9)

Abreu, E., Aritonovska, E., Reichenbach, P., Cristofari, G., Culp, B., Terns, R.M., Lingner, J., and Terns, M.P. (2010). TIN2-tethered TPP1 recruits human telomerase to telomeres in vivo. Mol. Cell. Biol. 30, 2971–2982.

Addinall, S.G., Downey, M., Yu, M., Zubko, M.K., Dewar, J., Leake, A., Hallinan, J., Shaw, O., James, K., Wilkinson, D.J., et al. (2008). A genome-wide suppressor and enhancer analysis of cdc13-1 reveals

varied cellular processes influencing telomere capping

in Saccharomyces cerevisiae. Genetics 180, 2251–2266.

Anbalagan, S., Bonetti, D., Lucchini, G., and Longhese, M.P. (2011). Rif1 supports the function of the CST complex in yeast telomere capping. PLOS Genet. 7, e1002024.

Aparicio, O.M., Billington, B.L., and Gottschling, D.E. (1991). Modifiers of position effect are shared between telomeric and silent mating-type loci in S. cerevisiae. Cell 66, 1279–1287.

Armanios, M., and Blackburn, E.H. (2012). The telomere syndromes. Nat. Rev. Genet. 13, 693–704.

Arnerić, M., and Lingner, J. (2007). Tel1 kinase and subtelomere-bound Tbf1 mediate preferential elongation of short telomeres by telomerase in yeast. EMBO Rep. 8, 1080–1085.

Askree, S.H., Yehuda, T., Smolikov, S., Gurevich, R., Hawk, J., Coker, C., Krauskopf, A., Kupiec, M., and McEachern, M.J. (2004). A genome-wide screen for

Saccharomyces cerevisiae deletion mutants that affect

telomere length. Proc. Natl. Acad. Sci. 101, 8658–

8663.

Azzalin, C.M., Reichenbach, P., Khoriauli, L., Giulotto, E., and Lingner, J. (2007). Telomeric repeat-containing RNA and RNA surveillance factors at

mammalian chromosome ends. Science 318, 798–801.

Balk, B., Maicher, A., Dees, M., Klermund, J., Luke-Glaser, S., Bender, K., and Luke, B. (2013). Telomeric RNA-DNA hybrids affect telomere-length dynamics and senescence. Nat. Struct. Mol. Biol. 20, 1199–1205.

Baur, J.A., Zou, Y., Shay, J.W., and Wright, W.E. (2001). Telomere position effect in human cells. Science 292,

2075–2077.

Berthiau, A.-S., Yankulov, K., Bah, A., Revardel, E., Luciano, P., Wellinger, R.J., Géli, V., and Gilson, E. (2006). Subtelomeric proteins negatively regulate telomere elongation in budding yeast. EMBO J. 25,

846–856.

Biessmann, H., Mason, J.M., Ferry, K., d’Hulst, M., Valgeirsdottir, K., Traverse, K.L., and Pardue, M.-L. (1990). Addition of telomere-associated HeT DNA sequences “heals” broken chromosome ends in

Drosophila. Cell 61, 663–673.

Biffi, G., Tannahill, D., McCafferty, J., and Balasubramanian, S. (2013). Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 5, 182–186.

Blasco, M.A. (2005). Telomeres and human disease: ageing, cancer and beyond. Nat. Rev. Genet. 6, 611–

622.

Bochman, M.L., Sabouri, N., and Zakian, V.A. (2010). Unwinding the functions of the Pif1 family helicases. DNA Repair 9, 237–249.

Bochman, M.L., Paeschke, K., and Zakian, V.A. (2012). DNA secondary structures: stability and function of G-quadruplex structures. Nat. Rev. Genet. 13, 770–780.

(10)

99

Bibliography

Bonetti, D., Clerici, M., Anbalagan, S., Martina, M., Lucchini, G., and Longhese, M.P. (2010). Shelterin-like proteins and Yku inhibit nucleolytic processing

of Saccharomyces cerevisiae telomeres. PLoS Genet. 6.

Boulé, J.B., Vega, L.R., and Zakian, V.A. (2005). The yeast Pif1p helicase removes telomerase from telomeric DNA. Nature 438, 57–61.

Boulton, S.J., and Jackson, S.P. (1998). Components of the Ku-dependent non-homologous end-joining pathway are involved in telomeric length maintenance and telomeric silencing. EMBO J. 17, 1819–1828.

Bourns, B.D., Alexander, M.K., Smith, A.M., and Zakian, V.A. (1998). Sir proteins, Rif proteins, and Cdc13p bind Saccharomyces telomeres in vivo. Mol. Cell.

Biol. 18, 5600–5608.

Brigati, C., Kurtz, S., Balderes, D., Vidali, G., and Shore, D. (1993). An essential yeast gene encoding a TTAGGG repeat-binding protein. Mol. Cell. Biol. 13,

1306–1314.

Britt-Compton, B., Capper, R., Rowson, J., and Baird, D. M. (2009). Short telomeres are preferentially elongated by telomerase in human cells. FEBS Lett.

583, 3076–3080.

Bryan, T.M., Englezou, A., Dalla-Pozza, L., Dunham, M.A., and Reddel, R.R. (1997). Evidence for an alternative mechanism for maintaining telomere length in human tumors and tumor-derived cell lines. Nat. Med. 3, 1271–1274.

Budd, M.E., Reis, C.C., Smith, S., Myung, K., and Campbell, J.L. (2006). Evidence suggesting that Pif1 helicase functions in DNA replication with the Dna2 helicase/nuclease and DNA polymerase δ. Mol. Cell. Biol. 26, 2490–2500.

Chandra, A., Hughes, T.R., Nugent, C.I., and Lundblad, V. (2001). Cdc13 both positively and negatively regulates telomere replication. Genes Dev.

15, 404–414.

Chang, M., Arneric, M., and Lingner, J. (2007). Telomerase repeat addition processivity is increased at critically short telomeres in a Tel1-dependent manner

in Saccharomyces cerevisiae. Genes Dev. 21, 2485–2494.

Chang, M., Dittmar, J.C., and Rothstein, R. (2011). Long telomeres are preferentially extended during recombination-mediated telomere maintenance. Nat. Struct. Mol. Biol. 18, 451–456.

Chen, H., Xue, J., Churikov, D., Hass, E.P., Shi, S., Lemon, L.D., Luciano, P., Bertuch, A.A., Zappulla, D.C., Géli, V., et al. (2018). Structural insights into yeast telomerase recruitment to telomeres. Cell 172,

331-343.

Chen, Q., Ijpma, A., and Greider, C.W. (2001). Two survivor pathways that allow growth in the absence of telomerase are generated by distinct telomere recombination events. Mol. Cell. Biol. 21, 1819–1827.

Chu, H.-P., Cifuentes-Rojas, C., Kesner, B., Aeby, E., Lee, H., Wei, C., Oh, H.J., Boukhali, M., Haas, W., and Lee, J.T. (2017). TERRA RNA antagonizes ATRX and protects telomeres. Cell 170, 86-101.e16.

Chung, W.-H., Zhu, Z., Papusha, A., Malkova, A., and Ira, G. (2010). Defective resection at DNA double-strand breaks leads to de novo telomere formation and

enhances gene targeting. PLoS Genet 6, e1000948.

Cimino-Reale, G., Pascale, E., Battiloro, E., Starace, G., Verna, R., and D’Ambrosio, E. (2001). The length of telomeric G-rich strand 3′-overhang measured by oligonucleotide ligation assay. Nucleic Acids Res. 29,

(11)

Claussin, C., and Chang, M. (2016). Multiple Rad52-mediated homology-directed repair mechanisms are required to prevent telomere attrition-induced senescence in Saccharomyces cerevisiae. PLOS Genet 12,

e1006176.

Cooley, C., Davé, A., Garg, M., and Bianchi, A. (2014). Tel1ATM dictates the replication timing of short yeast

telomeres. EMBO Rep. 15, 1093–1101.

Crabbe, L., Verdun, R.E., Haggblom, C.I., and Karlseder, J. (2004). Defective telomere lagging strand synthesis in cells lacking WRN helicase activity. Science 306, 1951–1953.

Daley, J.M., Palmbos, P.L., Wu, D., and Wilson, T.E. (2005). Nonhomologous end joining in yeast. Annu. Rev. Genet. 39, 431–451.

Denchi, E.L., and de Lange, T. (2007). Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 448, 1068–1071.

Dev, H., Chiang, T.-W.W., Lescale, C., Krijger, I. de, Martin, A.G., Pilger, D., Coates, J., Sczaniecka-Clift, M., Wei, W., Ostermaier, M., et al. (2018). Shieldin complex promotes DNA end-joining and counters homologous recombination in BRCA1-null cells. Nat. Cell Biol. 20, 954–965.

Dewar, J.M., and Lydall, D. (2010). Pif1- and Exo1-dependent nucleases coordinate checkpoint activation following telomere uncapping. EMBO J. 29, 4020–

4034.

Dewar, J.M., and Lydall, D. (2012). Similarities and differences between “uncapped” telomeres and DNA double-strand breaks. Chromosoma 121, 117–130.

Diede, S.J., and Gottschling, D.E. (1999). Telomerase-mediated telomere addition in vivo requires DNA

primase and DNA polymerases α and δ. Cell 99,

723–733.

Downey, M., Houlsworth, R., Maringele, L., Rollie, A., Brehme, M., Galicia, S., Guillard, S., Partington, M., Zubko, M.K., Krogan, N.J., et al. (2006). A genome-wide screen identifies the evolutionarily conserved KEOPS complex as a telomere regulator. Cell 124,

1155–1168.

Du, X., Shen, J., Kugan, N., Furth, E.E., Lombard, D.B., Cheung, C., Pak, S., Luo, G., Pignolo, R.J., DePinho, R.A., et al. (2004). Telomere shortening exposes functions for the mouse Werner and Bloom syndrome genes. Mol. Cell. Biol. 24, 8437–8446.

Dunham, M.A., Neumann, A.A., Fasching, C.L., and Reddel, R.R. (2000). Telomere maintenance by recombination in human cells. Nat. Genet. 26, 447–

450.

Dyke, M.W.V., Nelson, L.D., Weilbaecher, R.G., and Mehta, D.V. (2004). Stm1p, a G4 quadruplex and purine motif triplex nucleic acid-binding protein, interacts with ribosomes and subtelomeric Y′ DNA

in Saccharomyces cerevisiae. J. Biol. Chem. 279, 24323–

24333.

Eugster, A., Lanzuolo, C., Bonneton, M., Luciano, P., Pollice, A., Pulitzer, J.F., Stegberg, E., Berthiau, A.-S., Förstemann, K., Corda, Y., et al. (2006). The finger subdomain of yeast telomerase cooperates with Pif1p to limit telomere elongation. Nat. Struct. Mol. Biol.

13, 734–739.

Evans, S.K., and Lundblad, V. (1999). Est1 and Cdc13 as comediators of telomerase access. Science 286,

117–120.

Fang, G., and Cech, T.R. (1993). The β subunit

(12)

101

Bibliography

G-quartet formation by telomeric DNA. Cell 74,

875–885.

Ferreira, M.G., Miller, K.M., and Cooper, J.P. (2004). Indecent exposure: when telomeres become uncapped. Mol. Cell 13, 7–18.

Forstemann, K., and Lingner, J. (2001). Molecular basis for telomere repeat divergence in budding yeast. Mol. Cell. Biol. 21, 7277–7286.

Förstemann, K., Höss, M., and Lingner, J. (2000). Telomerase-dependent repeat divergence at the 3′ ends of yeast telomeres. Nucleic Acids Res. 28, 2690–

2694.

Förstemann, K., Zaug, A.J., Cech, T.R., and Lingner, J. (2003). Yeast telomerase is specialized for C/A-rich RNA templates. Nucleic Acids Res. 31, 1646–1655.

Fouladi, B., Sabatier, L., Miller, D., Pottier, G., and Murnane, J.P. (2000). The relationship between spontaneous telomere loss and chromosome instability in a human tumor cell line. Neoplasia N. Y. N 2, 540–554.

Frank, C.J., Hyde, M., and Greider, C.W. (2006). Regulation of telomere elongation by the cyclin-dependent kinase CDK1. Mol. Cell 24, 423–432.

Frantz, J.D., and Gilbert, W. (1995). A yeast gene product, G4p2, with a specific affinity for quadruplex nucleic acids. J. Biol. Chem. 270, 9413–9419.

Fukunaga, K., Hirano, Y., and Sugimoto, K. (2012). Subbinding protein Tbf1 and telomere-binding protein Rap1 collaborate to inhibit localization of the Mre11 complex to DNA ends in budding yeast. Mol. Biol. Cell 23, 347–359.

Gao, H., Toro, T.B., Paschini, M.,

Braunstein-Ballew, B., Cervantes, R.B., and Lundblad, V. (2010). Telomerase recruitment in Saccharomyces cerevisiae is

not dependent on Tel1-mediated phosphorylation of Cdc13. Genetics 186, 1147–1159.

Garvik, B., Carson, M., and Hartwell, L. (1995). Single-stranded DNA arising at telomeres in cdc13

mutants may constitute a specific signal for the RAD9

checkpoint. Mol. Cell. Biol. 15, 6128–6138.

Gatbonton, T., Imbesi, M., Nelson, M., Akey, J.M., Ruderfer, D.M., Kruglyak, L., Simon, J.A., and Bedalov, A. (2006). Telomere length as a quantitative trait: genome-wide survey and genetic mapping of telomere length-control genes in yeast. PLoS Genet

2, e35.

Gellert, M., Lipsett, M.N., and Davies, D.R. (1962). Helix formation by guanylic acid. Proc. Natl. Acad. Sci. U. S. A. 48, 2013–2018.

Geronimo, C.L., and Zakian, V.A. (2016). Getting it done at the ends: Pif1 family DNA helicases and telomeres. DNA Repair.

Ghezraoui, H., Oliveira, C., Becker, J.R., Bilham, K., Moralli, D., Anzilotti, C., Fischer, R., Deobagkar-Lele, M., Sanchiz-Calvo, M., Fueyo-Marcos, E., et al. (2018). 53BP1 cooperation with the REV7–shieldin complex underpins DNA structure-specific NHEJ. Nature 560, 122–127.

Gilson, E., Roberge, M., Giraldo, R., Rhodes, D., and Gasser, S.M. (1993). Distortion of the DNA double helix by RAP1 at silencers and multiple telomeric binding sites. J. Mol. Biol. 231, 293–310.

Giraldo, R., and Rhodes, D. (1994). The yeast telomere-binding protein RAP1 binds to and promotes the formation of DNA quadruplexes in telomeric DNA. EMBO J. 13, 2411–2420.

(13)

Giraldo, R., Suzuki, M., Chapman, L., and Rhodes, D. (1994). Promotion of parallel DNA quadruplexes by a yeast telomere binding protein: a circular dichroism study. Proc. Natl. Acad. Sci. 91, 7658–7662.

Gomez, D., O’Donohue, M.-F., Wenner, T., Douarre, C., Macadré, J., Koebel, P., Giraud-Panis, M.-J., Kaplan, H., Kolkes, A., Shin-ya, K., et al. (2006a). The G-quadruplex ligand telomestatin inhibits POT1 binding to telomeric sequences in vitro and induces

GFP-POT1 dissociation from telomeres in human cells. Cancer Res. 66, 6908–6912.

Gomez, D., Wenner, T., Brassart, B., Douarre, C., O’Donohue, M.-F., Khoury, V.E., Shin-ya, K., Morjani, H., Trentesaux, C., and Riou, J.-F. (2006b). Telomestatin-induced telomere uncapping is modulated by POT1 through G-overhang extension in HT1080 human tumor cells. J. Biol. Chem. 281,

38721–38729.

Gonzalo, S., Jaco, I., Fraga, M.F., Chen, T., Li, E., Esteller, M., and Blasco, M.A. (2006). DNA methyltransferases control telomere length and telomere recombination in mammalian cells. Nat. Cell Biol. 8, 416–424.

Gottschling, D.E., Aparicio, O.M., Billington, B.L., and Zakian, V.A. (1990). Position effect at S. cerevisiae telomeres: reversible repression of Pol II

transcription. Cell 63, 751–762.

Graf, M., Bonetti, D., Lockhart, A., Serhal, K., Kellner, V., Maicher, A., Jolivet, P., Teixeira, M.T., and Luke, B. (2017). Telomere length determines TERRA and R-Loop regulation through the cell cycle. Cell

170, 72-85. e14.

Graham, I.R., and Chambers, A. (1994). Use of a selection technique to identify the diversity of binding sites for the yeast RAP1 transcription factor. Nucleic

Acids Res. 22, 124–130.

Grandin, N., Reed, S.I., and Charbonneau, M. (1997). Stn1, a new Saccharomyces cerevisiae protein, is implicated

in telomere size regulation in association with Cdc13. Genes Dev. 11, 512–527.

Grandin, N., Damon, C., and Charbonneau, M. (2000). Cdc13 cooperates with the yeast Ku proteins and Stn1 to regulate telomerase recruitment. Mol. Cell. Biol. 20, 8397–8408.

Grandin, N., Damon, C., and Charbonneau, M. (2001). Ten1 functions in telomere end protection and length regulation in association with Stn1 and Cdc13. EMBO J. 20, 1173–1183.

Greenwell, P.W., Kronmal, S.L., Porter, S.E., Gassenhuber, J., Obermaier, B., and Petes, T.D. (1995). TEL1, a gene involved in controlling telomere

length in S. cerevisiae, is homologous to the human

ataxia telangiectasia gene. Cell 82, 823–829.

Greider, C.W. (2016). Regulating telomere length from the inside out: the replication fork model. Genes Dev. 30, 1483–1491.

Greider, C.W., and Blackburn, E.H. (1985). Identification of a specific telomere terminal transferase activity in Tetrahymena extracts. Cell 43,

405–413.

Griffith, J.D., Comeau, L., Rosenfield, S., Stansel, R.M., Bianchi, A., Moss, H., and Lange, T. de (1999). Mammalian telomeres end in a large duplex loop. Cell

97, 503–514.

Grossi, S., Bianchi, A., Damay, P., and Shore, D. (2001). Telomere formation by Rap1p binding site arrays reveals end-specific length regulation requirements and active telomeric recombination. Mol. Cell. Biol.

(14)

103

Bibliography

21, 8117–8128.

Gupta, R., Somyajit, K., Narita, T., Maskey, E., Stanlie, A., Kremer, M., Typas, D., Lammers, M., Mailand, N., Nussenzweig, A., et al. (2018). DNA repair network analysis reveals shieldin as a key regulator of NHEJ and PARP inhibitor sensitivity. Cell 173, 972-988.e23.

Hanahan, D., and Weinberg, R.A. (2011). Hallmarks of cancer: the next generation. Cell 144, 646–674.

Hardy, C.F., Sussel, L., and Shore, D. (1992). A RAP1-interacting protein involved in transcriptional silencing and telomere length regulation. Genes Dev.

6, 801–814.

Hass, E.P., and Zappulla, D.C. (2015). The Ku subunit of telomerase binds Sir4 to recruit telomerase to lengthen telomeres in S. cerevisiae. ELife 4.

Hayashi, N., and Murakami, S. (2002). STM1, a gene

which encodes a guanine quadruplex binding protein, interacts with CDC13 in Saccharomyces cerevisiae. Mol.

Genet. Genomics 267, 806–813.

Hecht, A., Laroche, T., Strahl-Bolsinger, S., Gasser, S.M., and Grunstein, M. (1995). Histone H3 and H4 N-termini interact with SIR3 and SIR4 proteins: a molecular model for the formation of heterochromatin in yeast. Cell 80, 583–592.

Hector, R.E., Shtofman, R.L., Ray, A., Chen, B.-R., Nyun, T., Berkner, K.L., and Runge, K.W. (2007). Tel1p preferentially associates with short telomeres to stimulate their elongation. Mol. Cell 27, 851–858.

Hemann, M.T., Strong, M.A., Hao, L.-Y., and Greider, C.W. (2001). The shortest telomere, not average telomere length, is critical for cell viability and chromosome stability. Cell 107, 67–77.

Henderson, E., Hardin, C.C., Walk, S.K., Tinoco, I., and Blackburn, E.H. (1987). Telomeric DNA oligonucleotides form novel intramolecular structures containing guanine·guanine base pairs. Cell 51, 899–

908.

Hirano, Y., and Sugimoto, K. (2007). Cdc13 telomere capping decreases Mec1 association but does not affect Tel1 association with DNA ends. Mol. Biol. Cell 18, 2026–2036.

Hirano, Y., Fukunaga, K., and Sugimoto, K. (2009). Rif1 and Rif2 inhibit localization of Tel1 to DNA ends. Mol. Cell 33, 312–322.

Hoppe, G.J., Tanny, J.C., Rudner, A.D., Gerber, S.A., Danaie, S., Gygi, S.P., and Moazed, D. (2002). Steps in assembly of silent chromatin in yeast: Sir3-independent binding of a Sir2/Sir4 complex to silencers and role for Sir2-dependent deacetylation. Mol. Cell. Biol. 22, 4167–4180.

Horowitz, H., Thorburn, P., and Haber, J.E. (1984). Rearrangements of highly polymorphic regions near telomeres of Saccharomyces cerevisiae. Mol. Cell. Biol. 4,

2509–2517.

Huang, P.-H., Pryde, F.E., Lester, D., Maddison, R.L., Borts, R.H., Hickson, I.D., and Louis, E.J. (2001).

SGS1 is required for telomere elongation in the

absence of telomerase. Curr. Biol. 11, 125–129.

Huber, M.D., Lee, D.C., and Maizels, N. (2002). G4 DNA unwinding by BLM and Sgs1p: substrate specificity and substrate-specific inhibition. Nucleic Acids Res. 30, 3954–3961.

Hug, N., and Lingner, J. (2006). Telomere length homeostasis. Chromosoma 115, 413–425.

(15)

M., and Lundblad, V. (2000). Identification of the single-strand telomeric DNA binding domain of

the Saccharomyces cerevisiae Cdc13 protein. Proc. Natl.

Acad. Sci. U. S. A. 97, 6457–6462.

Iglesias, N., Redon, S., Pfeiffer, V., Dees, M., Lingner, J., and Luke, B. (2011). Subtelomeric repetitive elements determine TERRA regulation by Rap1/Rif and Rap1/Sir complexes in yeast. EMBO Rep. 12,

587–593.

Janke, C., Magiera, M.M., Rathfelder, N., Taxis, C., Reber, S., Maekawa, H., Moreno-Borchart, A., Doenges, G., Schwob, E., Schiebel, E., et al. (2004). A versatile toolbox for PCR-based tagging of yeast genes: new fluorescent proteins, more markers and promoter substitution cassettes. Yeast 21, 947–962.

Ji, H., Adkins, C.J., Cartwright, B.R., and Friedman, K.L. (2008). Yeast Est2p affects telomere length by influencing association of Rap1p with telomeric chromatin. Mol. Cell. Biol. 28, 2380–2390.

Johnson, F.B., Marciniak, R.A., McVey, M., Stewart, S.A., Hahn, W.C., and Guarente, L. (2001). The

Saccharomyces cerevisiae WRN homolog Sgs1p

participates in telomere maintenance in cells lacking telomerase. EMBO J. 20, 905–913.

Karlseder, J., Broccoli, D., Dai, Y., Hardy, S., and Lange, T. de (1999). p53- and ATM-dependent apoptosis induced by telomeres lacking TRF2. Science 283, 1321–1325.

Kaul, Z., Cesare, A.J., Huschtscha, L.I., Neumann, A.A., and Reddel, R.R. (2012). Five dysfunctional telomeres predict onset of senescence in human cells. EMBO Rep. 13, 52–59.

Kim, N.W., Piatyszek, M.A., Prowse, K.R., Harley, C.B., West, M.D., Ho, P.L., Coviello, G.M., Wright,

W.E., Weinrich, S.L., and Shay, J.W. (1994). Specific association of human telomerase activity with immortal cells and cancer. Science 266, 2011–2015.

Koering, C.E., Pollice, A., Zibella, M.P., Bauwens, S., Puisieux, A., Brunori, M., Brun, C., Martins, L., Sabatier, L., Pulitzer, J.F., et al. (2002). Human telomeric position effect is determined by chromosomal context and telomeric chromatin integrity. EMBO Rep. 3, 1055–1061.

König, P., and Rhodes, D. (1997). Recognition of telomeric DNA. Trends Biochem. Sci. 22, 43–47.

Kramer, K.M., and Haber, J.E. (1993). New telomeres in yeast are initiated with a highly selected subset of TG1-3 repeats. Genes Dev. 7, 2345–2356.

Kyrion, G., Boakye, K.A., and Lustig, A.J. (1992). C-terminal truncation of RAP1 results in the deregulation of telomere size, stability, and function

in Saccharomyces cerevisiae. Mol. Cell. Biol. 12, 5159–

5173.

Lachner, M., O’Carroll, D., Rea, S., Mechtler, K., and Jenuwein, T. (2001). Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins. Nature 410,

116–120.

de Lange, T. (2005). Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110.

de Lange, T. (2009). How Telomeres Solve the End-Protection Problem. Science 326, 948.

Larcher, M.V., Pasquier, E., MacDonald, R.S., and Wellinger, R.J. (2016). Ku binding on telomeres occurs at sites distal from the physical chromosome ends. PLOS Genet. 12, e1006479.

(16)

105

Bibliography

Larrivée, M., LeBel, C., and Wellinger, R.J. (2004). The generation of proper constitutive G-tails on yeast telomeres is dependent on the MRX complex. Genes Dev. 18, 1391–1396.

Le, S., Moore, J.K., Haber, J.E., and Greider, C.W. (1999). RAD50 and RAD51 define two pathways that

collaborate to maintain telomeres in the absence of telomerase. Genetics 152, 143–152.

Lee, J.S., Mandell, E.K., Tucey, T.M., Morris, D.K., and Victoria, L. (2008). The Est3 protein associates with yeast telomerase through an OB-fold domain. Nat. Struct. Mol. Biol. 15, 990–997.

Lendvay, T.S., Morris, D.K., Sah, J., Balasubramanian, B., and Lundblad, V. (1996). Senescence mutants

of Saccharomyces cerevisiae with a defect in telomere

replication identify three additional EST genes.

Genetics 144, 1399–1412.

Levy, D.L., and Blackburn, E.H. (2004). Counting of Rif1p and Rif2p on Saccharomyces cerevisiae telomeres

regulates telomere length. Mol. Cell. Biol. 24, 10857–

10867.

Li, J.R., Yu, T.Y., Chien, I.C., Lu, C.Y., Lin, J.J., and Li, H.W. (2014). Pif1 regulates telomere length by preferentially removing telomerase from long telomere ends. Nucleic Acids Res. 42, 8527–8536.

Li, J.S.Z., Fuste, J.M., Simavorian, T., Bartocci, C., Tsai, J., Karlseder, J., and Denchi, E.L. (2017). TZAP: a telomere-associated protein involved in telomere length control. Science aah6752.

Li, Q.-J., Tong, X.-J., Duan, Y.-M., and Zhou, J.-Q. (2013). Characterization of the intramolecular G-quadruplex promoting activity of Est1. FEBS Lett.

587, 659–665.

Lin, J.-J., and Zakian, V.A. (1996). The Saccharomyces

CDC13 protein is a single-strand TG1–3 telomeric

DNA-binding protein in vitro that affects telomere

behavior in vivo. Proc. Natl. Acad. Sci. U. S. A. 93,

13760–13765.

Lingner, J., Hughes, T.R., Shevchenko, A., Mann, M., Lundblad, V., and Cech, T.R. (1997). Reverse transcriptase motifs in the catalytic subunit of telomerase. Science 276, 561–567.

Liu, C., Mao, X., and Lustig, A.J. (1994). Mutational analysis defines a C-terminal tail domain of Rap1 essential for telomeric silencing in Saccharomyces cerevisiae. Genetics 138, 1025–1040.

Lue, N.F., Yu, E.Y., and Lei, M. (2013). A popular engagement at the ends. Nat. Struct. Mol. Biol. 20,

10–12.

Luke, B., Panza, A., Redon, S., Iglesias, N., Li, Z., and Lingner, J. (2008). The Rat1p 5′ to 3′ exonuclease degrades telomeric repeat-containing RNA and promotes telomere elongation in Saccharomyces cerevisiae. Mol. Cell 32, 465–477.

Lundblad, V., and Blackburn, E.H. (1993). An alternative pathway for yeast telomere maintenance rescues est1- senescence. Cell 73, 347–360.

Lundblad, V., and Szostak, J.W. (1989). A mutant with a defect in telomere elongation leads to senescence in yeast. Cell 57, 633–643.

Lustig, A.J. (1992). Hoogsteen G-G base pairing is dispensable for telomere healing in yeast. Nucleic Acids Res. 20, 3021–3028.

Lustig, A.J., Kurtz, S., and Shore, D. (1990). Involvement of the silencer and UAS binding protein RAP1 in regulation of telomere length. Science 250,

(17)

549–553.

Lydeard, J.R., Jain, S., Yamaguchi, M., and Haber, J.E. (2007). Break-induced replication and telomerase-independent telomere maintenance require Pol32. Nature 448, 820–823.

Makovets, S., and Blackburn, E.H. (2009). DNA damage signalling prevents deleterious telomere addition at DNA breaks. Nat. Cell Biol. 11, 1383–

1386.

Mangahas, J.L., Alexander, M.K., Sandell, L.L., and Zakian, V.A. (2001). Repair of chromosome ends after telomere loss in Saccharomyces. Mol. Biol. Cell 12,

4078–4089.

Marcand, S., Gilson, E., and Shore, D. (1997). A protein-counting mechanism for telomere length regulation in yeast. Science 275, 986–990.

Marcand, S., Brevet, V., and Gilson, E. (1999). Progressive cis-inhibition of telomerase upon

telomere elongation. EMBO J. 18, 3509–3519.

Marcand, S., Pardo, B., Gratias, A., Cahun, S., and Callebaut, I. (2008). Multiple pathways inhibit NHEJ at telomeres. Genes Dev. 22, 1153–1158.

Maringele, L., and Lydall, D. (2002). EXO1-dependent

single-stranded DNA at telomeres activates subsets of DNA damage and spindle checkpoint pathways in budding yeast yku70Δ mutants. Genes Dev. 16,

1919–1933.

Mason, M., Wanat, J.J., Harper, S., Schultz, D.C., Speicher, D.W., Johnson, F.B., and Skordalakes, E. (2013). Cdc13 OB2 dimerization required for productive Stn1 binding and efficient telomere maintenance. Struct. Lond. Engl. 1993 21, 109–120.

Mateyak, M.K., and Zakian, V.A. (2006). Human PIF helicase is cell cycle regulated and associates with telomerase. Cell Cycle 5, 2796–2804.

Mattarocci, S., Reinert, J.K., Bunker, R.D., Fontana, G.A., Shi, T., Klein, D., Cavadini, S., Faty, M., Shyian, M., Hafner, L., et al. (2017a). Rif1 maintains telomeres and mediates DNA repair by encasing DNA ends. Nat. Struct. Mol. Biol. 24, 588–595.

Mattarocci, S., Reinert, J.K., Bunker, R.D., Fontana, G.A., Shi, T., Klein, D., Cavadini, S., Faty, M., Shyian, M., Hafner, L., et al. (2017b). Rif1 maintains telomeres and mediates DNA repair by encasing DNA ends. Nat. Struct. Mol. Biol. advance online publication.

McClintock, B. (1941). The stability of broken ends of chromosomes in Zea mays. Genetics 26, 234–282.

McGee, J.S., Phillips, J.A., Chan, A., Sabourin, M., Paeschke, K., and Zakian, V.A. (2010). Reduced Rif2 and lack of Mec1 target short telomeres for elongation rather than double-strand break repair. Nat. Struct. Mol. Biol. 17, 1438–1445.

Mersaoui, S.Y., Bonnell, E., and Wellinger, R.J. (2018). Nuclear import of Cdc13 limits chromosomal capping. Nucleic Acids Res. 46, 2975–2989.

Mirman, Z., Lottersberger, F., Takai, H., Kibe, T., Gong, Y., Takai, K., Bianchi, A., Zimmermann, M., Durocher, D., and Lange, T. de (2018). 53BP1–RIF1– shieldin counteracts DSB resection through CST- and Polα-dependent fill-in. Nature 560, 112–116.

Mitchell, M.T., Smith, J.S., Mason, M., Harper, S., Speicher, D.W., Johnson, F.B., and Skordalakes, E. (2010). Cdc13 N-terminal dimerization, DNA binding, and telomere length regulation. Mol. Cell. Biol. 30, 5325–5334.

(18)

107

Bibliography

Miyake, Y., Nakamura, M., Nabetani, A., Shimamura, S., Tamura, M., Yonehara, S., Saito, M., and Ishikawa, F. (2009). RPA-like mammalian Ctc1-Stn1-Ten1 complex binds to single-stranded DNA and protects telomeres independently of the Pot1 pathway. Mol. Cell 36, 193–206.

Moretti, P., and Shore, D. (2001). Multiple interactions in Sir protein recruitment by Rap1p at silencers and telomeres in yeast. Mol. Cell. Biol. 21, 8082–8094.

Moretti, P., Freeman, K., Coodly, L., and Shore, D. (1994). Evidence that a complex of SIR proteins interacts with the silencer and telomere-binding protein RAP1. Genes Dev. 8, 2257–2269.

Moye, A.L., Porter, K.C., Cohen, S.B., Phan, T., Zyner, K.G., Sasaki, N., Lovrecz, G.O., Beck, J.L., and Bryan, T.M. (2015). Telomeric G-quadruplexes are a substrate and site of localization for human telomerase. Nat. Commun. 6, 7643.

Muller, H. J. (1927). Artificial transmutation of the gene. Science 66, 84–87.

Muller, H. J. (1938). The remaking of chromosomes. Collect. Net 13, 181–195.

Myung, K., Chen, C., and Kolodner, R.D. (2001). Multiple pathways cooperate in the suppression of genome instability in Saccharomyces cerevisiae. Nature 411, 1073–1076.

Nandakumar, J., Bell, C.F., Weidenfeld, I., Zaug, A.J., Leinwand, L.A., and Cech, T.R. (2012). The TEL patch of telomere protein TPP1 mediates telomerase recruitment and processivity. Nature 492, 285–289.

Negrini, S., Ribaud, V., Bianchi, A., and Shore, D. (2007). DNA breaks are masked by multiple Rap1 binding in yeast: implications for telomere capping

and telomerase regulation. Genes Dev. 21, 292–302.

Nikolova, E.N., Kim, E., Wise, A.A., O’Brien, P.J., Andricioaei, I., and Al-Hashimi, H.M. (2011). Transient Hoogsteen base pairs in canonical duplex DNA. Nature 470, 498–502.

Noordermeer, S.M., Adam, S., Setiaputra, D., Barazas, M., Pettitt, S.J., Ling, A.K., Olivieri, M., Álvarez-Quilón, A., Moatti, N., Zimmermann, M., et al. (2018). The shieldin complex mediates 53BP1-dependent DNA repair. Nature 560, 117–121.

Nugent, C.I., Hughes, T.R., Lue, N.F., and Lundblad, V. (1996). Cdc13p: a single-strand telomeric DNA-binding protein with a dual role in yeast telomere maintenance. Science 274, 249–252.

Obodo, U.C., Epum, E.A., Platts, M.H., Seloff, J., Dahlson, N.A., Velkovsky, S.M., Paul, S.R., and Friedman, K.L. (2016). Endogenous hot spots of de novo telomere addition in the yeast genome contain

proximal enhancers that bind Cdc13. Mol. Cell. Biol.

36, 1750–1763.

Oganesian, L., Moon, I.K., Bryan, T.M., and Jarstfer, M.B. (2006). Extension of G-quadruplex DNA by ciliate telomerase. EMBO J. 25, 1148–1159.

Oganesian, L., Graham, M.E., Robinson, P.J., and Bryan, T.M. (2007). Telomerase recognizes G-quadruplex and linear DNA as distinct substrates. Biochemistry 46, 11279–11290.

Olovnikov, A.M. (1973). A theory of marginotomy. J. Theor. Biol. 41, 181–190.

Paeschke, K., Simonsson, T., Postberg, J., Rhodes, D., and Lipps, H.J. (2005). Telomere end-binding proteins control the formation of G-quadruplex DNA structures in vivo. Nat. Struct. Mol. Biol. 12,

(19)

847–854.

Paeschke, K., Juranek, S., Simonsson, T., Hempel, A., Rhodes, D., and Lipps, H.J. (2008). Telomerase recruitment by the telomere end binding protein-β facilitates G-quadruplex DNA unfolding in ciliates. Nat. Struct. Mol. Biol. 15, 598–604.

Paeschke, K., McDonald, K.R., and Zakian, V.A. (2010). Telomeres: structures in need of unwinding. FEBS Lett. 584, 3760–3772.

Paeschke, K., Capra, J.A., and Zakian, V.A. (2011). DNA replication through G-quadruplex motifs is promoted by the Saccharomyces cerevisiae Pif1 DNA

helicase. Cell 145, 678–691.

Paeschke, K., Bochman, M.L., Garcia, P.D., Cejka, P., Friedman, K.L., Kowalczykowski, S.C., and Zakian, V.A. (2013). Pif1 family helicases suppress genome instability at G-quadruplex motifs. Nature 497, 458–

462.

Palm, W., and Lange, T. de (2008). How shelterin protects mammalian telomeres. Annu. Rev. Genet.

42, 301–334.

Parenteau, J., and Wellinger, R.J. (1999). Accumulation of single-stranded DNA and destabilization of telomeric repeats in yeast mutant strains carrying a deletion of RAD27. Mol. Cell. Biol. 19, 4143–4152.

Park, P.U., Defossez, P.-A., and Guarente, L. (1999). Effects of Mutations in DNA repair genes on formation of ribosomal DNA circles and life span in

Saccharomyces cerevisiae. Mol. Cell. Biol. 19, 3848–3856.

Parkinson, G.N., Lee, M.P.H., and Neidle, S. (2002). Crystal structure of parallel quadruplexes from human telomeric DNA. Nature 417, 876–880.

Paschini, M., Toro, T.B., Lubin, J.W., Braunstein-Ballew, B., Morris, D.K., and Lundblad, V. (2012). A naturally thermolabile activity compromises genetic analysis of telomere function in Saccharomyces cerevisiae.

Genetics 191, 79–93.

Pennock, E., Buckley, K., and Lundblad, V. (2001). Cdc13 delivers separate complexes to the telomere for end protection and replication. Cell 104, 387–396.

Pfeiffer, V., and Lingner, J. (2012). TERRA promotes telomere shortening through Exonuclease 1– mediated resection of chromosome ends. PLOS Genet. 8, e1002747.

Phillips, J.A., Chan, A., Paeschke, K., and Zakian, V.A. (2015). The Pif1 helicase, a negative regulator of telomerase, acts preferentially at long telomeres. PLOS Genet. 11, e1005186.

Piazza, A., Boulé, J.-B., Lopes, J., Mingo, K., Largy, E., Teulade-Fichou, M.-P., and Nicolas, A. (2010). Genetic instability triggered by G-quadruplex interacting Phen-DC compounds in Saccharomyces cerevisiae. Nucleic Acids Res. 38, 4337–4348.

Piazza, A., Serero, A., Boulé, J.-B., Legoix-Né, P., Lopes, J., and Nicolas, A. (2012). Stimulation of gross chromosomal rearrangements by the human CEB1 and CEB25 minisatellites in Saccharomyces cerevisiae

depends on G-quadruplexes or Cdc13. PLOS Genet.

8, e1003033.

Poon, B.P.K., and Mekhail, K. (2012). Effects of perinuclear chromosome tethers in the telomeric

URA3/5FOA system reflect changes to gene silencing

and not nucleotide metabolism. Genet. Aging 3, 144.

Porro, A., Feuerhahn, S., Delafontaine, J., Riethman, H., Rougemont, J., and Lingner, J. (2014). Functional characterization of the TERRA transcriptome at

(20)

109

Bibliography

damaged telomeres. Nat. Commun. 5, 5379.

Postberg, J., Tsytlonok, M., Sparvoli, D., Rhodes, D., and Lipps, H.J. (2012). A telomerase-associated RecQ protein-like helicase resolves telomeric G-quadruplex structures during replication. Gene 497, 147–154.

Pryde, F.E., and Louis, E.J. (1999). Limitations of silencing at native yeast telomeres. EMBO J. 18,

2538–2550.

Puglisi, A., Bianchi, A., Lemmens, L., Damay, P., and Shore, D. (2008). Distinct roles for yeast Stn1 in telomere capping and telomerase inhibition. EMBO

J. 27, 2328–2339.

Qi, H., and Zakian, V.A. (2000). The Saccharomyces

telomere-binding protein Cdc13p interacts with both the catalytic subunit of DNA polymerase α and the telomerase-associated Est1 protein. Genes Dev. 14,

1777–1788.

Ray, A., and Runge, K.W. (1998). The C terminus of the major yeast telomere binding protein Rap1p enhances telomere formation. Mol. Cell. Biol. 18,

1284–1295.

Ray, A., and Runge, K.W. (1999). The yeast telomere length counting machinery is sensitive to sequences at the telomere-nontelomere junction. Mol. Cell. Biol.

19, 31–45.

Resnick, M.A., and Martin, P. (1976). The repair of double-strand breaks in the nuclear DNA of

Saccharomyces cerevisiae and its genetic control. Mol.

Gen. Genet. MGG 143, 119–129.

Rhodes, D., and Lipps, H.J. (2015). G-quadruplexes and their regulatory roles in biology. Nucleic Acids Res. 43, 8627–8637.

Ritchie, K.B., Mallory, J.C., and Petes, T.D. (1999). Interactions of TLC1 (which encodes the RNA

subunit of telomerase), TEL1, and MEC1 in

regulating telomere length in the yeast Saccharomyces cerevisiae. Mol. Cell. Biol. 19, 6065–6075.

Rizzo, A., Salvati, E., Porru, M., D’Angelo, C., Stevens, M.F., D’Incalci, M., Leonetti, C., Gilson, E., Zupi, G., and Biroccio, A. (2009). Stabilization of quadruplex DNA perturbs telomere replication leading to the activation of an ATR-dependent ATM signaling pathway. Nucleic Acids Res. 37, 5353–5364.

Rodriguez, R., Miller, K.M., Forment, J.V., Bradshaw, C.R., Nikan, M., Britton, S., Oelschlaegel, T., Xhemalce, B., Balasubramanian, S., and Jackson, S.P. (2012). Small-molecule–induced DNA damage identifies alternative DNA structures in human genes. Nat. Chem. Biol. 8, 301–310.

Rossmann, M.P., Luo, W., Tsaponina, O., Chabes, A., and Stillman, B. (2011). A common telomeric gene silencing assay is affected by nucleotide metabolism. Mol. Cell 42, 127–136.

Roy, R., Meier, B., McAinsh, A.D., Feldmann, H.M., and Jackson, S.P. (2004). Separation-of-function mutants of yeast Ku80 reveal a Yku80p-Sir4p interaction involved in telomeric silencing. J. Biol. Chem. 279, 86–94.

Sabourin, M., Tuzon, C.T., and Zakian, V.A. (2007). Telomerase and Tel1p preferentially associate with short telomeres in S. cerevisiae. Mol. Cell 27, 550–561.

Schaffitzel, C., Berger, I., Postberg, J., Hanes, J., Lipps, H.J., and Plückthun, A. (2001). In vitro generated

antibodies specific for telomeric guanine-quadruplex DNA react with Stylonychia lemnae macronuclei. Proc.

(21)

Schulz, V.P., and Zakian, V.A. (1994). The Saccharomyces PIF1 DNA helicase inhibits telomere elongation and de novo telomere formation. Cell 76, 145–155.

Selvarajah, S., Yoshimoto, M., Park, P.C., Maire, G., Paderova, J., Bayani, J., Lim, G., Al-Romaih, K., Squire, J.A., and Zielenska, M. (2006). The breakage–fusion– bridge (BFB) cycle as a mechanism for generating genetic heterogeneity in osteosarcoma. Chromosoma

115, 459–467.

Sfeir, A., Kosiyatrakul, S.T., Hockemeyer, D., MacRae, S.L., Karlseder, J., Schildkraut, C.L., and Lange, T. de (2009). Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138,

90–103.

Sfeir, A., Kabir, S., Overbeek, M. van, Celli, G.B., and Lange, T. de (2010). Loss of Rap1 induces telomere recombination in the absence of NHEJ or a DNA damage signal. Science 327, 1657–1661.

Shampay, J., Szostak, J.W., and Blackburn, E.H. (1984). DNA sequences of telomeres maintained in yeast. Nature 310, 154–157.

Sherman, F. (2002). Getting started with yeast. Methods Enzymol. 194, 3–21.

Shi, T., Bunker, R.D., Mattarocci, S., Ribeyre, C., Faty, M., Gut, H., Scrima, A., Rass, U., Rubin, S.M., Shore, D., et al. (2013). Rif1 and Rif2 shape telomere function and architecture through multivalent Rap1 interactions. Cell 153, 1340–1353.

Singer, M., and Gottschling, D. (1994). TLC1:

template RNA component of Saccharomyces cerevisiae

telomerase. Science 266, 404–409.

Smith, J.S., Chen, Q., Yatsunyk, L.A., Nicoludis, J.M., Garcia, M.S., Kranaster, R., Balasubramanian, S.,

Monchaud, D., Teulade-Fichou, M.-P., Abramowitz, L., et al. (2011). Rudimentary G-quadruplex–based telomere capping in Saccharomyces cerevisiae. Nat. Struct.

Mol. Biol. 18, 478–485.

Smogorzewska, A., Steensel, B. van, Bianchi, A., Oelmann, S., Schaefer, M.R., Schnapp, G., and Lange, T. de (2000). Control of human telomere length by TRF1 and TRF2. Mol. Cell. Biol. 20, 1659–1668.

van Steensel, B., and de Lange, T. (1997). Control of telomere length by the human telomeric protein TRF1. Nature 385, 740–743.

van Steensel, B., Smogorzewska, A., and de Lange, T. (1998). TRF2 protects human telomeres from end-to-end fusions. Cell 92, 401–413.

Stellwagen, A.E., Haimberger, Z.W., Veatch, J.R., and Gottschling, D.E. (2003). Ku interacts with telomerase RNA to promote telomere addition at native and broken chromosome ends. Genes Dev. 17,

2384–2395.

Stinus, S., Paeschke, K., and Chang, M. (2017). Telomerase regulation by the Pif1 helicase: a length-dependent effect? Curr. Genet. 64, 509–513.

Strecker, J., Stinus, S., Caballero, M.P., Szilard, R.K., Chang, M., and Durocher, D. (2017). A sharp Pif1-dependent threshold separates DNA double-strand breaks from critically short telomeres. ELife 6,

e23783.

Sun, H., Bennett, R.J., and Maizels, N. (1999). The

Saccharomyces cerevisiae Sgs1 helicase efficiently unwinds

G-G paired DNAs. Nucleic Acids Res. 27, 1978–1984.

Sun, J., Yang, Y., Wan, K., Mao, N., Yu, T.-Y., Lin, Y.-C., DeZwaan, D.C., Freeman, B.C., Lin, J.-J., Lue, N.F., et al. (2011). Structural bases of dimerization of

(22)

111

Bibliography

yeast telomere protein Cdc13 and its interaction with the catalytic subunit of DNA polymerase α. Cell Res.

21, 258–274.

Sussel, L., and Shore, D. (1991). Separation of transcriptional activation and silencing functions of

the RAP1-encoded repressor/activator protein 1:

isolation of viable mutants affecting both silencing and telomere length. Proc. Natl. Acad. Sci. 88, 7749–

7753.

Takai, K.K., Kibe, T., Donigian, J.R., Frescas, D., and de Lange, T. (2011). Telomere protection by TPP1/ POT1 requires tethering to TIN2. Mol. Cell 44, 647–

659.

Teixeira, M.T., Arneric, M., Sperisen, P., and Lingner, J. (2004). Telomere length homeostasis is achieved via a switch between telomerase- extendible and -nonextendible states. Cell 117, 323–335.

Teng, S.-C., and Zakian, V.A. (1999). Telomere-telomere recombination is an efficient bypass pathway for telomere maintenance in Saccharomyces cerevisiae.

Mol. Cell. Biol. 19, 8083–8093.

Treco, D.A., and Lundblad, V. (1993). Preparation of yeast media. In Current Protocols in Molecular Biology, (John Wiley & Sons, Inc.), pp. 13.1.1-13.1.7. Tsai, Y.-C., Qi, H., and Liu, L.F. (2007). Protection of DNA ends by telomeric 3′ G-tail sequences. J. Biol. Chem. 282, 18786–18792.

Tsai, Y.-L., Tseng, S.-F., Chang, S.-H., Lin, C.-C., and Teng, S.-C. (2002). Involvement of replicative polymerases, Tel1p, Mec1p, Cdc13p, and the Ku complex in telomere-telomere recombination. Mol. Cell. Biol. 22, 5679–5687.

Tseng, S.-F., Lin, J.-J., and Teng, S.-C. (2006). The

telomerase-recruitment domain of the telomere binding protein Cdc13 is regulated by Mec1p/Tel1p-dependent phosphorylation. Nucleic Acids Res. 34,

6327–6336.

Ungar, L., Yosef, N., Sela, Y., Sharan, R., Ruppin, E., and Kupiec, M. (2009). A genome-wide screen for essential yeast genes that affect telomere length maintenance. Nucleic Acids Res. 37, 3840–3849.

Vannier, J.-B., Pavicic-Kaltenbrunner, V., Petalcorin, M.I.R., Ding, H., and Boulton, S.J. (2012). RTEL1 dismantles T loops and counteracts telomeric G4-DNA to maintain telomere integrity. Cell 149, 795–

806.

Vodenicharov, M.D., and Wellinger, R.J. (2006). DNA degradation at unprotected telomeres in yeast is regulated by the CDK1 (Cdc28/Clb) cell-cycle kinase. Mol. Cell 24, 127–137.

Wahlin, J., and Cohn, M. (2000). Saccharomyces cerevisiae

RAP1 binds to telomeric sequences with spatial flexibility. Nucleic Acids Res. 28, 2292–2301.

Watson, J.D. (1972). Origin of concatemeric T7 DNA. Nature. New Biol. 239, 197–201.

Weiffenbach, B., and Haber, J.E. (1981). Homothallic mating type switching generates lethal chromosome breaks in rad52 strains of Saccharomyces cerevisiae. Mol.

Cell. Biol. 1, 522–534.

Wellinger, R.J., and Zakian, V.A. (2012). Everything you ever wanted to know about Saccharomyces cerevisiae

telomeres: beginning to end. Genetics 191, 1073–

1105.

Wellinger, R.J., Wolf, A.J., and Zakian, V.A. (1993).

Saccharomyces telomeres acquire single-strand TG1–3

(23)

Wilkie, A.O.M., Lamb, J., Harris, P.C., Finney, R.D., and Higgs, D.R. (1990). A truncated human chromosome 16 associated with α thalassaemia is stabilized by addition of telomeric repeat (TTAGGG)

n. Nature 346, 868–871.

Williamson, J.R., Raghuraman, M.K., and Cech, T.R. (1989). Monovalent cation-induced structure of telomeric DNA: The G-quartet model. Cell 59,

871–880.

Wong, A.C., Ning, Y., Flint, J., Clark, K., Dumanski, J.P., Ledbetter, D.H., and McDermid, H.E. (1997). Molecular characterization of a 130-kb terminal microdeletion at 22q in a child with mild mental retardation. Am. J. Hum. Genet. 60, 113–120.

Wotton, D., and Shore, D. (1997). A novel Rap1p-interacting factor, Rif2p, cooperates with Rif1p to regulate telomere length in Saccharomyces cerevisiae.

Genes Dev. 11, 748–760.

Xu, F., Zhang, Q., Zhang, K., Xie, W., and Grunstein, M. (2007). Sir2 deacetylates histone H3 lysine 56 to regulate telomeric heterochromatin structure in yeast. Mol. Cell 27, 890–900.

Xu, Z., Duc, K.D., Holcman, D., and Teixeira, M.T. (2013). The length of the shortest telomere as the major determinant of the onset of replicative senescence. Genetics 194, 847–857.

Yu, E.Y., Wang, F., Lei, M., and Lue, N.F. (2008). A proposed OB-fold with a protein-interaction surface

in Candida albicans telomerase protein Est3. Nat.

Struct. Mol. Biol. 15, 985–989.

Zahler, A.M., Williamson, J.R., Cech, T.R., and Prescott, D.M. (1991). Inhibition of telomerase by G-quartet DNA structures. Nature 350, 718–720.

Zakian, V.A. (1996). Structure, function, and replication of Saccharomyces cerevisiae telomeres. Annu.

Rev. Genet. 30, 141–172.

Zhang, W., and Durocher, D. (2010). De novo telomere

formation is suppressed by the Mec1-dependent inhibition of Cdc13 accumulation at DNA breaks. Genes Dev. 24, 502–515.

Zhang, D.-H., Zhou, B., Huang, Y., Xu, L.-X., and Zhou, J.-Q. (2006). The human Pif1 helicase, a potential Escherichia coli RecD homologue, inhibits

telomerase activity. Nucleic Acids Res. 34, 1393–1404.

Zhang, M.-L., Tong, X.-J., Fu, X.-H., Zhou, B.O., Wang, J., Liao, X.-H., Li, Q.-J., Shen, N., Ding, J., and Zhou, J.-Q. (2010). Yeast telomerase subunit Est1p has guanine quadruplex-promoting activity that is required for telomere elongation. Nat. Struct. Mol. Biol. 17, 202–209.

Zhong, F.L., Batista, L.F.Z., Freund, A., Pech, M.F., Venteicher, A.S., and Artandi, S.E. (2012). TPP1 OB-fold domain controls telomere maintenance by recruiting telomerase to chromosome ends. Cell 150,

481–494.

Zhou, J.Q., Monson, E.K., Teng, S.C., Schulz, V.P., and Zakian, V.A. (2000). Pif1p helicase, a catalytic inhibitor of telomerase in yeast. Science 289, 771–

774.

Zimmer, J., Tacconi, E.M.C., Folio, C., Badie, S., Porru, M., Klare, K., Tumiati, M., Markkanen, E., Halder, S., Ryan, A., et al. (2016). Targeting BRCA1 and BRCA2 deficiencies with G-quadruplex-interacting compounds. Mol. Cell 61, 1–12.

Zou, L., and Elledge, S.J. (2003). Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 300, 1542–1548.

Referenties

GERELATEERDE DOCUMENTEN

Cloete I, BSc, BSc RD(SA), Dietitian, False Bay Hospital Daniels L, BSc, MPH RD(SA), RNT(SA), Lecturer Jordaan J, BSc RD(SA), Dietitian, Gordonia Hospital Derbyshire C, BSc

Here, a simplified version of the Gielis equation was shown to be an excellent model for describing the foliage leaf blades of bamboo with lanceolate characteristics and

On the molecular biology of telomeres Stinus Ruiz de Gauna, Sonia.. IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite

cerevisiae telomerase Est1 subunit and the major telomeric binding protein Rap1 have been reported to promote the formation of parallel G-quadruplexes (Giraldo and Rhodes,

At both the V-R and VI-R telomeres, the percentage of elongated telomeres below the DSB-telomere transition length determined in PIF1 cells increased in the pif1-m2

We defined this tract length as the DSB-telomere threshold, below which Pif1 actively supresses telomere addition (Strecker et al., 2017). This result suggests that Pif1

Given that the main features of telomeres, namely the general structure and function, are conserved among eukaryotes, and that the sequence composition required for G-quadruplex

Moreover, we found that Pif1 inhibits telomere addition at DNA ends that have less than ~40 base pairs of telomeric sequence, but not above this threshold, suggesting that,