• Aucun résultat trouvé

MATERIALS AND METHODS

Dans le document On start and pause of replication fork (Page 127-150)

Budding Yeast Rif1 Controls Genome Integrity by Inhibiting rDNA Replication

MATERIALS AND METHODS

Yeast strains and genetics

Standard genetic methods for budding yeast strains construction and crossing were used (Shyian et al., 2016).

1D and 2D gels and Southern blot

2D gels were performed essentially as in (Shyian et al., 2016) using BglII enzyme for genomic DNA digestion. For 1D gels the first dimension gel was directly transferred to nylon membrane and probed with a radioactively labeled probe specific to Fob1-RFB site (Shyian et al., 2016).

rDNA instability assay (ADE2 marker loss)

ADE2 maker loss assays were performed essentially as in (Shyian et al., 2016).

Serial dilution spot assay

125 Serial dilution spot assays were performed essentially as in (Shyian et al., 2016).

Mcm4 and Cdc45 ChIP

Mcm4-Myc and Cdc45-Myc anti-Myc ChIP assays were performed essentially as in (Mattarocci et al., 2014).

‘Cowcatcher’ screens

Strains containing single copy of ADE2 and URA3 genes inserted into rDNA array were used for mutagenesis with EMS at 50% survival. EMS treated cultures were split in 10 separate tubes, inoculated into SC-ADE-URA and grown overnight (to counter select mutations in ADE2 and URA3). Then, aliquots were inoculated into YPAD and grown overnight (in order to let cell loose makers from rDNA). Dilutions were plated on 5-FOA plates (selection of URA3 loss) and incubated as in “ADE2 loss assay”. After visual inspection, red sectored colonies from 5-FOA plates were manually selected and their white sectors were streaked sequentially 2 times onto SC plates. Of ca. 50’000 colonies from 5-FOA plates, 30 independent reproducibly highly sectoring isolates were chosen. Those were back-crossed, sporulated, dissected and assessed for segregation of the high sectoring phenotype.

Isolates showing 2:2 segregation of sectoring (as Mendelian monoallelic mutations) were subjected to causative mutation identification using pooled linkage analysis, as in (Birkeland et al., 2010; Lang et al., 2015). Briefly, 20 spore colonies with sectoring phenotype were pooled together in one pool (+phenotype) and 20 white spore colonies were pooled in second pool (-phenotype); genomic DNA was isolated with Qiagen genomic tip kit. Total genomic DNA of the two pools was submitted to iGE3 Genomics Platform of University of Geneva for fragmentation, library preparation and whole genome deep sequencing. The FASTQ files with sequencing reads from iGE3 Genomics Platform were mapped to S.

cerevisiae genome with Mapping tool of the ‘HTSstation’ (http://htsstation.epfl.ch/) The SNPs were identified with SNP tool of the same ‘HTSstation’.The SNPs unique/over-represented in plus-phenotype pool compared to minus-phenotype pool were identified in Excel. The mutations in polymerase genes were confirmed by segregation of phenotype in back-crosses with strains harboring C-terminally tagged versions of polymerase genes and by Sanger sequencing of the mutation site.

FACS

FACS assays were performed essentially as in (Mattarocci et al., 2014).

PAGE-WB

126 PAGE-WB assays were performed essentially as in (Shyian et al., 2016).

REFERENCES

Achar, Y.J., and Foiani, M. (2017). Coordinating Replication with Transcription. Adv Exp Med Biol 1042, 455-487.

Akamatsu, Y., and Kobayashi, T. (2015). The Human RNA Polymerase I Transcription Terminator Complex Acts as a Replication Fork Barrier That Coordinates the Progress of Replication with rRNA Transcription Activity. Mol Cell Biol 35, 1871-1881.

Azvolinsky, A., Giresi, P.G., Lieb, J.D., and Zakian, V.A. (2009). Highly transcribed RNA polymerase II genes are impediments to replication fork progression in Saccharomyces cerevisiae. Mol Cell 34, 722-734.

Barbari, S.R., and Shcherbakova, P.V. (2017). Replicative DNA polymerase defects in human cancers: Consequences, mechanisms, and implications for therapy. DNA Repair (Amst) 56, 16-25.

Bastia, D., and Zaman, S. (2014). Mechanism and physiological significance of programmed replication termination. Semin Cell Dev Biol 30, 165-173.

Bell, S.P., and Labib, K. (2016). Chromosome Duplication in Saccharomyces cerevisiae.

Genetics 203, 1027-1067.

Bentsen, I.B., Nielsen, I., Lisby, M., Nielsen, H.B., Gupta, S.S., Mundbjerg, K., Andersen, A.H., and Bjergbaek, L. (2013). MRX protects fork integrity at protein-DNA barriers, and its absence causes checkpoint activation dependent on chromatin context. Nucleic Acids Res 41, 3173-3189.

Bermejo, R., Doksani, Y., Capra, T., Katou, Y.M., Tanaka, H., Shirahige, K., and Foiani, M.

(2007). Top1- and Top2-mediated topological transitions at replication forks ensure fork progression and stability and prevent DNA damage checkpoint activation. Genes Dev 21, 1921-1936.

Birkeland, S.R., Jin, N., Ozdemir, A.C., Lyons, R.H., Jr., Weisman, L.S., and Wilson, T.E.

(2010). Discovery of mutations in Saccharomyces cerevisiae by pooled linkage analysis and whole-genome sequencing. Genetics 186, 1127-1137.

Brewer, B.J., and Fangman, W.L. (1988). A replication fork barrier at the 3' end of yeast ribosomal RNA genes. Cell 55, 637-643.

Brill, S.J., DiNardo, S., Voelkel-Meiman, K., and Sternglanz, R. (1987). Need for DNA topoisomerase activity as a swivel for DNA replication for transcription of ribosomal RNA. Nature 326, 414-416.

Burkhalter, M.D., and Sogo, J.M. (2004). rDNA enhancer affects replication initiation and mitotic recombination: Fob1 mediates nucleolytic processing independently of replication. Mol Cell 15, 409-421. cluster of Saccharomyces cerevisiae. PLoS Genet 4, e1000105.

Christman, M.F., Dietrich, F.S., and Fink, G.R. (1988). Mitotic recombination in the rDNA of S. cerevisiae is suppressed by the combined action of DNA topoisomerases I and II.

Cell 55, 413-425.

127 Costanzo, M., VanderSluis, B., Koch, E.N., Baryshnikova, A., Pons, C., Tan, G., Wang, W., Usaj, M., Hanchard, J., Lee, S.D., et al. (2016). A global genetic interaction network maps a wiring diagram of cellular function. Science 353.

Dalgaard, J.Z., and Klar, A.J. (2000). swi1 and swi3 perform imprinting, pausing, and termination of DNA replication in S. pombe. Cell 102, 745-751.

Das-Bradoo, S., Nguyen, H.D., Wood, J.L., Ricke, R.M., Haworth, J.C., and Bielinsky, A.K.

(2010). Defects in DNA ligase I trigger PCNA ubiquitylation at Lys 107. Nat Cell Biol 12, 74-79; sup pp 71-20.

Di Felice, F., Cioci, F., and Camilloni, G. (2005). FOB1 affects DNA topoisomerase I in vivo cleavages in the enhancer region of the Saccharomyces cerevisiae ribosomal DNA locus.

Nucleic Acids Res 33, 6327-6337.

Duguet, M. (1997). When helicase and topoisomerase meet! J Cell Sci 110 ( Pt 12), 1345-1350.

Errico, A., Cosentino, C., Rivera, T., Losada, A., Schwob, E., Hunt, T., and Costanzo, V.

(2009). Tipin/Tim1/And1 protein complex promotes Pol alpha chromatin binding and sister chromatid cohesion. EMBO J 28, 3681-3692.

Fachinetti, D., Bermejo, R., Cocito, A., Minardi, S., Katou, Y., Kanoh, Y., Shirahige, K., Azvolinsky, A., Zakian, V.A., and Foiani, M. (2010). Replication termination at eukaryotic chromosomes is mediated by Top2 and occurs at genomic loci containing pausing elements. Mol Cell 39, 595-605.

Feng, W., Di Rienzi, S.C., Raghuraman, M.K., and Brewer, B.J. (2011). Replication stress-induced chromosome breakage is correlated with replication fork progression and is preceded by single-stranded DNA formation. G3 (Bethesda) 1, 327-335.

Fumasoni, M., Zwicky, K., Vanoli, F., Lopes, M., and Branzei, D. (2015). Error-free DNA damage tolerance and sister chromatid proximity during DNA replication rely on the Polalpha/Primase/Ctf4 Complex. Mol Cell 57, 812-823.

Gadaleta, M.C., and Noguchi, E. (2017). Regulation of DNA Replication through Natural Impediments in the Eukaryotic Genome. Genes (Basel) 8.

Gaillard, H., Garcia-Muse, T., and Aguilera, A. (2015). Replication stress and cancer. Nat Rev Cancer 15, 276-289.

Gambus, A., Jones, R.C., Sanchez-Diaz, A., Kanemaki, M., van Deursen, F., Edmondson, R.D., and Labib, K. (2006). GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat Cell Biol 8, 358-366.

Gartenberg, M.R., and Smith, J.S. (2016). The Nuts and Bolts of Transcriptionally Silent Chromatin in Saccharomyces cerevisiae. Genetics 203, 1563-1599.

Haruki, H., Nishikawa, J., and Laemmli, U.K. (2008). The anchor-away technique: rapid, conditional establishment of yeast mutant phenotypes. Mol Cell 31, 925-932.

Hodgson, B., Calzada, A., and Labib, K. (2007). Mrc1 and Tof1 regulate DNA replication forks in different ways during normal S phase. Mol Biol Cell 18, 3894-3902.

Hosono, Y., Abe, T., Higuchi, M., Kajii, K., Sakuraba, S., Tada, S., Enomoto, T., and Seki, M.

(2014). Tipin functions in the protection against topoisomerase I inhibitor. J Biol Chem 289, 11374-11384.

Kaeberlein, M., McVey, M., and Guarente, L. (1999). The SIR2/3/4 complex and SIR2 alone promote longevity in Saccharomyces cerevisiae by two different mechanisms.

Genes Dev 13, 2570-2580.

Kim, R.A., and Wang, J.C. (1989). A subthreshold level of DNA topoisomerases leads to the excision of yeast rDNA as extrachromosomal rings. Cell 57, 975-985.

128 Kobayashi, T., Horiuchi, T., Tongaonkar, P., Vu, L., and Nomura, M. (2004). SIR2 regulates recombination between different rDNA repeats, but not recombination within individual rRNA genes in yeast. Cell 117, 441-453.

Krawczyk, C., Dion, V., Schar, P., and Fritsch, O. (2014). Reversible Top1 cleavage complexes are stabilized strand-specifically at the ribosomal replication fork barrier and contribute to ribosomal DNA stability. Nucleic Acids Res 42, 4985-4995.

Krings, G., and Bastia, D. (2004). swi1- and swi3-dependent and independent replication fork arrest at the ribosomal DNA of Schizosaccharomyces pombe. Proc Natl Acad Sci U S A 101, 14085-14090.

Lang, A.B., John Peter, A.T., Walter, P., and Kornmann, B. (2015). ER-mitochondrial junctions can be bypassed by dominant mutations in the endosomal protein Vps13. J Cell Biol 210, 883-890.

Lemoine, F.J., Degtyareva, N.P., Kokoska, R.J., and Petes, T.D. (2008). Reduced levels of DNA polymerase delta induce chromosome fragile site instability in yeast. Mol Cell Biol 28, 5359-5368.

Lemoine, F.J., Degtyareva, N.P., Lobachev, K., and Petes, T.D. (2005). Chromosomal translocations in yeast induced by low levels of DNA polymerase a model for chromosome fragile sites. Cell 120, 587-598.

Macheret, M., and Halazonetis, T.D. (2015). DNA replication stress as a hallmark of cancer. Annu Rev Pathol 10, 425-448.

Mattarocci, S., Shyian, M., Lemmens, L., Damay, P., Altintas, D.M., Shi, T., Bartholomew, C.R., Thoma, N.H., Hardy, C.F., and Shore, D. (2014). Rif1 controls DNA replication timing in yeast through the PP1 phosphatase Glc7. Cell Rep 7, 62-69.

Mekhail, K., Seebacher, J., Gygi, S.P., and Moazed, D. (2008). Role for perinuclear chromosome tethering in maintenance of genome stability. Nature 456, 667-670.

Mohanty, B.K., Bairwa, N.K., and Bastia, D. (2006). The Tof1p-Csm3p protein complex counteracts the Rrm3p helicase to control replication termination of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 103, 897-902.

Morawska, M., and Ulrich, H.D. (2013). An expanded tool kit for the auxin-inducible degron system in budding yeast. Yeast 30, 341-351.

Mundbjerg, K., Jorgensen, S.W., Fredsoe, J., Nielsen, I., Pedersen, J.M., Bentsen, I.B., Lisby, M., Bjergbaek, L., and Andersen, A.H. (2015). Top2 and Sgs1-Top3 Act Redundantly to Ensure rDNA Replication Termination. PLoS Genet 11, e1005697.

Nedelcheva, M.N., Roguev, A., Dolapchiev, L.B., Shevchenko, A., Taskov, H.B., Shevchenko, A., Stewart, A.F., and Stoynov, S.S. (2005). Uncoupling of unwinding from DNA synthesis implies regulation of MCM helicase by Tof1/Mrc1/Csm3 checkpoint complex. J Mol Biol 347, 509-521.

Park, H., and Sternglanz, R. (1999). Identification and characterization of the genes for two topoisomerase I-interacting proteins from Saccharomyces cerevisiae. Yeast 15, 35-41.

Pasero, P., Bensimon, A., and Schwob, E. (2002). Single-molecule analysis reveals clustering and epigenetic regulation of replication origins at the yeast rDNA locus. Genes Dev 16, 2479-2484.

Perera, R.L., Torella, R., Klinge, S., Kilkenny, M.L., Maman, J.D., and Pellegrini, L. (2013).

Mechanism for priming DNA synthesis by yeast DNA polymerase alpha. Elife 2, e00482.

Pommier, Y., Sun, Y., Huang, S.N., and Nitiss, J.L. (2016). Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat Rev Mol Cell Biol 17, 703-721.

129 Redon, C., Pilch, D.R., and Bonner, W.M. (2006). Genetic analysis of Saccharomyces cerevisiae H2A serine 129 mutant suggests a functional relationship between H2A and the sister-chromatid cohesion partners Csm3-Tof1 for the repair of topoisomerase I-induced DNA damage. Genetics 172, 67-76.

Reid, R.J., Gonzalez-Barrera, S., Sunjevaric, I., Alvaro, D., Ciccone, S., Wagner, M., and Rothstein, R. (2011). Selective ploidy ablation, a high-throughput plasmid transfer protocol, identifies new genes affecting topoisomerase I-induced DNA damage. Genome Res 21, 477-486.

Salim, D., Bradford, W.D., Freeland, A., Cady, G., Wang, J., Pruitt, S.C., and Gerton, J.L.

(2017). DNA replication stress restricts ribosomal DNA copy number. PLoS Genet 13, e1007006.

Schalbetter, S.A., Mansoubi, S., Chambers, A.L., Downs, J.A., and Baxter, J. (2015). Fork rotation and DNA precatenation are restricted during DNA replication to prevent chromosomal instability. Proc Natl Acad Sci U S A 112, E4565-4570.

Segurado, M., and Diffley, J.F. (2008). Separate roles for the DNA damage checkpoint protein kinases in stabilizing DNA replication forks. Genes Dev 22, 1816-1827.

Sekedat, M.D., Fenyo, D., Rogers, R.S., Tackett, A.J., Aitchison, J.D., and Chait, B.T. (2010).

GINS motion reveals replication fork progression is remarkably uniform throughout the yeast genome. Mol Syst Biol 6, 353.

Shah, K.A., Shishkin, A.A., Voineagu, I., Pavlov, Y.I., Shcherbakova, P.V., and Mirkin, S.M.

(2012). Role of DNA polymerases in repeat-mediated genome instability. Cell Rep 2, 1088-1095.

Shyian, M., Mattarocci, S., Albert, B., Hafner, L., Lezaja, A., Costanzo, M., Boone, C., and Shore, D. (2016). Budding Yeast Rif1 Controls Genome Integrity by Inhibiting rDNA Replication. PLoS Genet 12, e1006414.

Smith, D.J., and Whitehouse, I. (2012). Intrinsic coupling of lagging-strand synthesis to chromatin assembly. Nature 483, 434-438.

Song, W., Dominska, M., Greenwell, P.W., and Petes, T.D. (2014). Genome-wide high-resolution mapping of chromosome fragile sites in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A 111, E2210-2218.

Strumberg, D., Pilon, A.A., Smith, M., Hickey, R., Malkas, L., and Pommier, Y. (2000).

Conversion of topoisomerase I cleavage complexes on the leading strand of ribosomal DNA into 5'-phosphorylated DNA double-strand breaks by replication runoff. Mol Cell Biol 20, 3977-3987.

Swan, M.K., Johnson, R.E., Prakash, L., Prakash, S., and Aggarwal, A.K. (2009). Structural basis of high-fidelity DNA synthesis by yeast DNA polymerase delta. Nat Struct Mol Biol 16, 979-986.

Tourriere, H., Versini, G., Cordon-Preciado, V., Alabert, C., and Pasero, P. (2005). Mrc1 and Tof1 promote replication fork progression and recovery independently of Rad53.

Mol Cell 19, 699-706.

Valjavec-Gratian, M., Henderson, T.A., and Hill, T.M. (2005). Tus-mediated arrest of DNA replication in Escherichia coli is modulated by DNA supercoiling. Mol Microbiol 58, 758-773.

Voineagu, I., Narayanan, V., Lobachev, K.S., and Mirkin, S.M. (2008). Replication stalling at unstable inverted repeats: interplay between DNA hairpins and fork stabilizing proteins. Proc Natl Acad Sci U S A 105, 9936-9941.

Voineagu, I., Surka, C.F., Shishkin, A.A., Krasilnikova, M.M., and Mirkin, S.M. (2009).

Replisome stalling and stabilization at CGG repeats, which are responsible for chromosomal fragility. Nat Struct Mol Biol 16, 226-228.

130 Willer, M., Rainey, M., Pullen, T., and Stirling, C.J. (1999). The yeast CDC9 gene encodes both a nuclear and a mitochondrial form of DNA ligase I. Curr Biol 9, 1085-1094.

Xu, H., Boone, C., and Brown, G.W. (2007). Genetic dissection of parallel sister-chromatid cohesion pathways. Genetics 176, 1417-1429.

Zheng, D.Q., Zhang, K., Wu, X.C., Mieczkowski, P.A., and Petes, T.D. (2016). Global analysis of genomic instability caused by DNA replication stress in Saccharomyces cerevisiae.

Proc Natl Acad Sci U S A 113, E8114-E8121.

Zou, H., and Rothstein, R. (1997). Holliday junctions accumulate in replication mutants via a RecA homolog-independent mechanism. Cell 90, 87-96.

131 DISCUSSION

DNA replication initiation control by conserved protein Rif1

Origin firing starts from DDK-dependent phosphorylation of inactive Mcm2-7 double hexamer and recruitment of origin firing factors Sld3/7 and Cdc45 (Bell and Labib, 2016). Rif1 appears to act exactly at this DDK-controlled step of DNA replication initiation. Genetically, loss of RIF1 compensates for the deficiencies of DDK components (ts alleles of cdc7 and dbf4) and origin firing factors ‘reading’ Mcm2-7 phosphorylation (sld3, cdc45, dpb11).

Biochemically, absence of Rif1 leads to increase in Mcm4 and Sld3 phosphorylation.

Mechanistically, Rif1 recruits PP1 phosphatase, which in turn wipes off the phosphate residues from firing factors Mcm4 and Sld3. Several important questions to this model of Rif1 action remain. First, what is the complete set of firing factors regulated by DDK and Rif1-Glc7? Second, is Rif1 acting in cis at regulated origins by recruiting PP1 there or in trans (e.g.

by influencing limiting firing factors distribution due to competition of early, late and rARS origins)?

Rif1-PP1 targets relevant for DNA replication

Which proteins harbor the phosphosites relevant for Rif1-PP1 dependent DNA replication initiation regulation?

By candidate approach, we and others (Dave et al., 2014; Hiraga et al., 2014;

Mattarocci et al., 2014) have discovered that out of the pre-replicative complex specifically the Mcm4 subunit of the Mcm2-7 complex is regulated through Rif-PP1 dependent dephosphorylation. A recent phosphoproteomic analysis of chromatin fraction in HEK293 cells treated with siRNA targeting RIF1 also found that Mcm4 is the most hyperphosphorylated in this condition (Hiraga et al., 2017). Additionally, this study found that some residues in Mcm2 and Mcm6 are also more phosphorylated upon depletion of RIF1, albeit to lower extent.

Moreover, while searching for Rif1-PP1 targets, we observed that Sld3 protein in cells lacking RIF1 exhibit a slower-migrating set of bands (Mattarocci et al., 2014). These bands appeared in G1 arrested cells in DDK-dependent manner, persisted in early S phase and

132 disappeared in later S phase and we presumed these Sld3 protein species to be phosphorylated forms. These observations are consistent with recent report describing a specific interaction between DDK regulatory subunit Dbf4 and Sld3 (Fang et al., 2017).

Interestingly, the mcm4-NΔ allele of Mcm4 lacking the replication initiation inhibitory domain (Sheu and Stillman, 2010) has additive effect with rif1Δ mutation on the temperature sensitive growth of cdc7-4 cells (Stefano Mattarocci, personal communication), highlighting a possibility of additional relevant DDK-dependent sites in Mcm4, involvement of Sld3 phosphosites, other targets of DDK or altogether DDK-independent additional effect of RIF1 deletion in this case. Further deep phosphoproteomic investigation of cells lacking Rif1 in various physiological conditions should help to clarify this question.

Molecular mechanism of selectivity in Rif1 action at origins

Chromatin immunoprecipitation methods in budding yeast detect Rif1 only on telomeres and silent mating type loci (HML and HMR), where it is recruited through Rif1 C-terminus via interaction with Rap1 (Smith et al., 2003). Curiously, in mouse embryonic stem cells Rif1 could be detected at megabase-sized domains throughout chromosomes (Foti et al., 2016). Moreover, these Rif1 associated sites in mouse overlap with late-replicating domains, suggesting that Rif1 could indeed act in cis at the affected origins. Recently, our laboratory used a different method, chromosome endogenous cleavage (ChEC-seq), to detect regions of Rif1 association with chromosomes in budding yeast (Hafner et al., 2018).

This method allowed detecting Rif1 association at the late origins regulated by Rif1, including rARS at the rDNA locus.

How does Rif1 recognize the origins then? The N-terminal HEAT repeats of Rif1 is an avid DNA-binding module (Mattarocci et al., 2017). However, budding yeast Rif1 doesn’t appear to show sequence-specific binding to DNA (Mattarocci et al., 2017), although the ARS sequence was not assessed in this study. Masai laboratory presented several reports for preferential binding by fission yeast Rif1 to G-quadruplex (G4) forming DNA sequences (Kanoh et al., 2015) both in vitro and in vivo. These sequences were adjacent to Rif1-regulated dormant or late origins in vivo and mutation of some of them led to enhanced origin firing in a big domain in the vicinity of the cite (Kanoh et al., 2015). A more recent report confirmed that murine Rif1 is also capable of G4 sequence binding and bridging (Moriyama et al., 2018). It is conceivable therefore that Rif1 might either access origins by

133 binding to G4 sequences in their vicinity in cis or by affecting the genome architecture and thereby origin clustering and accessibility to firing factors in trans.

Alternatively, some structural or topological transitions at origin DNA during its activation could recruit Rif1. Indeed, Rif1 has preference in binding toward 3’ ssDNA junctions with dsDNA (Mattarocci et al., 2017) and ssDNA or ssDNA-dsDNA junctions could be formed during initial origin activation. Going further in this line of thoughts, we could imagine that initial origin melting at the earliest steps of origin firing could create a favorable substrate for Rif1-Glc7 recruitment and reversal of the activation. However, once the DDK-CDK activities overcome some threshold, their activation of origins would win over the inhibitory reversing activity of Rif1-Glc7. Still alternatively, protein factors binding to the origins could be responsible for Rif1 recruitment. For instance, as Rif1 interacts with Dbf4 component of DDK (Mattarocci et al., 2014), it could be expected to follow DDK at the sites of its localization at chromatin. Further studies will undoubtedly help to resolve which (if any) of these models is/are correct.

It is worth noting that one should not necessarily expect a direct interaction of Rif1 with all the origins it regulates. Indeed, the replication initiation limiting factors model predicts that a change in firing of a set of origins could lead to concomitant change of an opposite sign in firing of another set. This was proven to be the case with Sir2 and Rpd3 effects on origin firing throughout the genome that are mostly secondary to the primary effect of these two factors on the rDNA origins (rARS) (Yoshida et al., 2014). Indeed, deletion or knockdown of RIF1 leads not only to advancement of late origins, but also to a delay in firing of some of the earlies origins. For instance, centromere-proximal origins in budding yeast are delayed in activation in rif1Δ (Peace et al., 2014). It would be of interest then to see how many of the Rif1-dependent changes in replication timing are direct effects due to Rif1 binding and PP1 recruitment and how many are secondary downstream effects due to firing factors re-distribution. We imagine that rDNA locus may play a significant role in this interplay, as it contains a lion’s share of all potential origins in yeast genome and would be expected therefore to contribute to the competition. Specifically, we imagine that one of the biological roles of Rif1 in DNA replication and genome integrity could be to spread the firing of rDNA origins over a broad time period in the middle S phase. Yeast cell starts to duplicate genome from early origins (especially centromeric, where DDK level is high in G1, which is necessary to maintain genome integrity (Natsume et al., 2013) followed by rARS and then

134 late origins (and even dormant in case of replication delay). Longer rDNA replication could therefore be partially responsible for the separation in time of early and late origins. A cell lacking Rif1 would have higher DDK-dependent pre-RC priming at both rARS and late/dormant origins. This, in turn, would lead to advancement of rARS and late origins firing time and overall more stochastic firing of all origins (i.e. loss of spatiotemporal control).

Rif1 is a phosphoprotein regulated by CDK-DDK, Mec1-Rad53 and Glc7

Rif1 phosphoregulation by CDK-DDK, Mec1-Rad53 and Glc7 could ensure an ON-OFF switch mechanism for Rif1-Glc7 interaction. In this model, Glc7 association with Rif1 is

Rif1 phosphoregulation by CDK-DDK, Mec1-Rad53 and Glc7 could ensure an ON-OFF switch mechanism for Rif1-Glc7 interaction. In this model, Glc7 association with Rif1 is

Dans le document On start and pause of replication fork (Page 127-150)

Documents relatifs