Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Sep 2.
Published in final edited form as: DNA Repair (Amst). 2009 May 27;8(9):974–982. doi: 10.1016/j.dnarep.2009.04.021

Perspectives on the DNA damage and replication checkpoint responses in Saccharomyces cerevisiae

Christopher D Putnam 1, Eric J Jaehnig 1, Richard D Kolodner 1
PMCID: PMC2725198  NIHMSID: NIHMS115434  PMID: 19477695

Abstract

The DNA damage and replication checkpoints are believed to primarily slow the progression of the cell cycle to allow DNA repair to occur. Here we summarize known aspects of the Saccharomyces cerevisiae checkpoints including how these responses are integrated into downstream effects on the cell cycle, chromatin, DNA repair, and cytoplasmic targets. Analysis of the transcriptional response demonstrates that it is far more complex and less relevant to the repair of DNA damage than the bacterial SOS response. We also address more speculative questions regarding potential roles of the checkpoint during the normal S-phase and how current evidence hints at a checkpoint activation mechanism mediated by positive feedback that amplifies initial damage signals above a minimium threshold.

Introduction

The concept of DNA damage checkpoints was first developed through the identification of G2/M arrest after X-ray irradiation in the budding yeast Saccharomyces cerevisiae. This arrest required the RAD9 gene, and the sensitivity of rad9 mutants was reduced by delaying the onset of mitosis after irradiation [1]. This led to the view that RAD9 and similar genes define DNA damage checkpoints, which delay specific cell cycle transitions in response to DNA damage to provide time for DNA repair to occur. In S. cerevisiae, DNA damage checkpoints delay the G1/S transition and block the G2/M transition of the cell cycle [1, 2]. In addition, two types of S-phase checkpoints have been defined: the DNA replication checkpoint, which arrests cell cycle progression and inhibits firing of late replication origins in response to replication stress [3], and the intra-S checkpoint, which slows DNA replication and cell cycle progression in response to DNA damage [4]. The different DNA damage checkpoints share many components and are now known to target many aspects of cellular metabolism besides cell cycle transitions. These checkpoints also likely respond to endogenous sources of DNA damage as well as exogeneous sources, as checkpoint defects result in increased spontaneous genome instability [5].

A “central dogma” for the DNA damage cell cycle checkpoints has been commonly presented:

damage signals -> damage sensors -> signal transducers -> effectors

In this scheme, damage is sensed by sensors, and this information is communicated through signal transducers to effectors that mediate the physiological response of the cell to the damage, including arresting or slowing the cell cycle and activating or repressing other pathways required for the eventual recovery of the cell. It has been convenient to think of checkpoints as unidirectional pathways, but this is an oversimplification. DNA repair proteins, for example, can act as both sensors and effectors. Similarly, checkpoint proteins that are components of replication complexes are both sensors and transducers and might even be effectors. This complexity suggests checkpoint responses likely involve complex regulatory networks that incorporate both feedback loops and threshold responses.

Here, we will summarize the core checkpoint machinery in S. cerevisiae to serve as a framework for examining some key features of checkpoint responses in greater detail. We will consider how checkpoints are activated and what is known about the effectors that are targeted by checkpoint activation. We will also consider the transcriptional response to DNA damage by reviewing which aspects of cell metabolism are transcriptionally regulated and how much of this actually represents a checkpoint response. Finally, we will also address more speculative questions regarding the potential roles of the DNA damage checkpoint during normal regulation and the use of positive feedback and threshold responses by checkpoint functions. We are particularly interested in how these checkpoint responses prevent genome instability. However, our goal is not to provide a comprehensive review of checkpoints but rather to highlight areas that are presently less well understood.

1. The central checkpoint pathway in S. cerevisiae

The key components of the common signaling pathway are several phosphoinositol-3-kinase-related kinase (PIKK) family members. S. cerevisiae has two PIKK proteins, Mec1, the homolog of human ataxia-telangiectasia and Rad3-related (ATR), and Tel1, the homolog of human ataxia telangiectasia mutated (ATM), that function as both damage sensors and signal transducers, but lacks a homolog of the DNA-dependent protein kinase, DNA-PKcs (Table 1). Loading of Mec1 and Tel1 onto damaged DNA does not appear to be mediated by direct recognition of DNA damage but rather by the recognition of complexes recognizing DNA damage and intermediates generated by the activities of DNA repair processes. Mec1 binds to Ddc2, the homolog of the human ATR interacting protein (ATRIP), which recognizes single-stranded DNA (ssDNA) bound by replication protein A (RPA) [6], and Tel1 binds to the DNA end-binding Mre11-Rad50-Xrs2 complex [7]. Genetically MEC1 and TEL1 are partially redundant; the mec1Δ tel1Δ double mutant cannot maintain telomeres via telomerase unlike the single mutants [8, 9], has a synergistically increased sensitivity to DNA damaging agents [10], and has a synergistically increased rate of spontaneous genome rearrangement relative to the respective single mutants [5]. Both kinases preferentially phosphorylate serines and threonines preceding a glutamine residue on numerous target proteins in response to damage. For example, phosphorylation of histone H2A at Ser129 (formation of γ-H2AX) by Mec1 and Tel1 has important roles for the checkpoint response detailed below [11]. The identity of all Mec1 and Tel1 targets and phosphorylation sites is certainly not known; however, improved proteomics approaches have led to considerable progress in identifying these targets. Despite their similarities, Mec1 and Tel1 are not completely redundant. Tel1 is more important for maintaining normal telomere lengths than Mec1 [8, 12]. Tel1 is more important in γ-H2AX formation at sites of DSBs in G1, and Mec1 is more important during S and G2 [13]. These cell cycle-dependent differences may be due to the yeast cyclin-dependent kinase (CDK) Cdk1 (also known as Cdc28) activating resection from DNA breaks via the nuclease Sae2 and thereby generating damage recognized by Mec1-Ddc2 [13, 14].

Table 1. Homologs of the components of the central kinase cascade.

Class S. cerevisiae S. pombe H. sapiens
Sensors MEC1 RAD3 ATR
DDC2 RAD26 ATRIP
RAD24 RAD17 RAD17
DDC1 RAD9 RAD9
MEC3 HUS1 HUS1
RAD17 RAD1 RAD1
DPB11 CUT5/RAD4 TOPBP1
TEL1 TEL1 ATM
MRE11 MRE11 MRE11
RAD50 RAD50 RAD50
XRS2 NBS1 NBS1
- - DNA-PKcs

Adaptors RAD9 CRB2 53BP1
MRC1 MRC1 CLSPN

Effector Kinases CHK1 CHK1 CHK1
RAD53 CDS1 CHK2

Activation of PIKK family members is also influenced by the action of other damage sensors. The PCNA-like Ddc1-Mec3-Rad17 complex (the S. cerevisiae homologs of the Rad9-Hus1-Rad1 or 9-1-1 complex) is loaded onto partial duplex DNA via the Rad24-Rfc2-5 alternative replication factor C (RFC) complex independently of Mec1-Ddc2 [15, 16]. Colocalization of Mec1-Ddc2-RPA and the 9-1-1 complex in the context of partial duplex DNA or to chromosomal arrays of Lac operator sequences in the absence of DNA damage results in Mec1 activation [17, 18], indicating that DNA plays a passive scaffolding role in checkpoint activation. These results make the lack of checkpoint activation from normal telomeres that bind numerous DNA repair and DNA damage checkpoint proteins even more surprising [19]. Activation of Mec1 is also mediated by Dpb11, the S. cerevisiae homolog of TopBP1, and Dpb11 is synergistic with the 9-1-1 complex [20, 21]. The combination of the 9-1-1 complex and Dpb11 in activation of Mec1 is highly conserved, and differs from fission yeast, Xenopus, and human systems only in that the 9-1-1 complex in those organisms cannot activate the Mec1 homolog in the absence of the Dpb11 homolog [22-24].

In addition to phosphorylating effectors of the checkpoint response (described below), the PIKK proteins also activate downstream kinases, including Chk1 and Rad53 (the S. cerevisiae homolog of mammalian Chk2), which presumably diffuse away from the site of their activation. In other organisms, Chk1 homologs are required to inhibit CDKs to prevent cell cycle progression in the presence of DNA damage; however, budding yeast lacks this requirement [25, 26]. Budding yeast Chk1 does, however, have direct roles in suppressing the cell cycle in the context of DNA damage as described below. Phosphorylation of the Rad53 protein kinase by Mec1 and Tel1 leads to its activation and subsequent autophosphorylation; the resulting hyperphosphorylated Rad53 is frequently used as an experimental surrogate for montoring activation of the DNA damage response. At least some inactive, hypophosphorylated Rad53, but not active hyperphosphorylated Rad53, is bound to the chromatin assembly factor Asf1 [27-29]. Rad53 also inhibits Asf1-mediated chromatin deposition in vitro. Despite the the binding of inactive Rad53 by Asf1, deletion of ASF1 does not give rise to a checkpoint response [27], but rather causes defects in checkpoint shut off due to its roles in chromatin assembly and modification [30].

Activation of Rad53 also depends upon Mec1 or Tel1 phosphorylating a scaffolding protein, Rad9 or Mrc1, which then binds the forkhead-associated (FHA) domains of Rad53; simultaneous deletion of both RAD9 and MRC1 extensively phenocopies a deletion of RAD53 [31], and mutagenesis of Mec1 and Tel1 consensus phosphorylation sites in Rad9 and Mrc1 prevent Rad53 activation [32, 33]. A Ddc2-Rad53 fusion construct alleviates many of the defects of a rad9Δ mrc1Δ double mutation [34], and a similar construct suppresses mrc1Δ defects in fission yeast [35]. Mrc1 is a component of the replication fork, has no identifiable protein domains or motifs, and seems to specifically signal replication stress [31, 33, 36]. Rad9, on the other hand, is more important for other types of DNA damage, although it is required for Rad53 hyperphosphorylation late in S-phase in mrc1Δ strains [33], which could indicate that DNA repair processes convert replication stress into damage recognized by Rad9. Rad9 contains tandem BRCT domains that bind phosphorylated residues, mediate dimerization after DNA damage-induced phosphorylation [37], and direct binding to phosphorylated histone H2A [38]. Rad9 also contains tandem Tudor domains, the latter of which recognize methylated histones [39]. In the G1/S and intra-S checkpoints, phosphorylation of Rad9 is dependent upon methylation of histone H3 at Lys79 by Dot1 or Lys4 by Set1, which in turn is dependent upon ubiquitination of histone H2B at Lys123 by Rad6 and the Bre1/Lge1 complex [40, 41]. Strains with deletion of RAD6, BRE1, both DOT1 and SET1, or strains encoding the histone H2B Lys123Arg variant have defects in the G1/S and the intra-S checkpoint after UV treatment and fail to hyperphosphorylate Rad53 [40, 41]. Surprisingly, deletion of BRE1 or DOT1 causes only minor defects in the G2/M checkpoint in response to a DSB, despite the role of BRE1 in G1/S and the intra-S checkpoints [40], and its role in conjunction with DPB11 in M phase [42]. Intriguingly, the mitotic roles for a phosphorylated adaptor protein appears to be conserved for the activation of the meiotic Rad53 homolog, Mek1, by phosphorylation of Hop1 by Mec1/Tel1 [43].

Rad53 recruits Dun1 through binding by the Dun1 FHA domain [44], and Rad53 activates the Dun1 protein kinase through phosphorylation of the activation loop [45]. Activation by Rad53 is the principle way in which Dun1 is activated; however, the phenotypes of mutations affecting the FHA domain of Dun1 have been used to argue for a some Rad53-independent roles of Dun1 [44]. The best characterized Dun1 substrate is the ribonucleotide reductase (RNR) inhibitor Sml1, which upon phosphorylation is targeted for degradation; consequently Sml1 is a key experimental target for following DNA damage signaling [46]. Deletion of RAD53 or MEC1 is lethal without co-deletion of SML1 or overexpression of RNR [47]; however, deletion of DUN1 does not require co-deletion of SML1 for viability, suggesting either that deficiency in Mec1 or Rad53 but not Dun1 causes defects requiring increased RNR function or that other pathways redundant with Dun1 inactivate Sml1, albeit without mediating Sml1 degradation. The net effect of activating the central kinase cascade in the DNA damage response is to amplify an initial chromatin-associated signal via the action of chromatin-bound and diffusible protein kinases that affect other processes to facilitate repair of the damage.

2. Targets of the DNA damage and replication checkpoint pathways

Identifying targets of checkpoint pathways is complicated by the fact that protein phosphorylation is a ubiquitous modification in the cell. In addition to genetic screens and candidate gene studies, a number of mass spectrometry experiments have identified phosphorylation sites dependent upon checkpoint activation [48, 49]. The overlap between these global screens and traditional studies is imperfect, possibly due to the complexity of the peptide mixtures being analyzed by mass spectrometry; however, these studies have identified new candidate targets for checkpoint pathways. Checkpoint-dependent phosphorylation sites are not necessarily functionally significant, however, as some proteins may be adventitious targets. Verification of the roles of individual targets is complicated by several factors, including multiple redundant phosphorylation sites, cell-cycle phase dependence, potential dependence on the types of DNA lesions, and the need to analyze both non-phosphorylatable and phosphomimicking amino acid substitutions at many different positions within an individual protein. The identification and validation of checkpoint targets remains an important area of study.

2.1 Cell-cycle targets of the checkpoint response

The arrest of the cell cycle is believed to be a primary role of checkpoint responses, and the checkpoint protein kinases play multiple roles in suppressing progression of the cell cycle through mitosis. Phosphorylation of Pds1, an anaphase suppressing protein, by Chk1 blocks its APC-dependent degradation [50]. Pds1 is further stabilized through Rad53-dependent inhibition of the Chk1-Cdc20 interaction, which recruits the APC ubiquitin ligase [51]. In addition, both Rad53 and Chk1 also appear to suppress the later stage of mitotic exit by preventing the release of the Cdc14 phosphatase from a complex with Net1 in the nucleolus thus inhibiting the mitotic exit network (MEN) and Cdc-fourteen early anaphase release (FEAR) pathways, respectively [52]. An additional role for Rad53 in suppressing MEN through regulation of the Cdc5 POLO-like kinase has been proposed [50]; however, this is controversial [52]. Release of Cdc14 marks the end of mitosis through dephosphorylation of Cdk1 targets, leading to the degradation of mitotic cyclins, induction of transcription of the SIC1 gene encoding a Cdk1 inhibitor, and stabilization of the expressed Sic1 protein [53]. The DNA damage checkpoint also appears to interact with the spindle assembly checkpoint, as Rad9 and Rad53 are phosphorylated after nocodazole arrest, which both activates the spindle assembly checkpoint and prevents the degradation of Pds1. This phosphorylation is independent of Mec1 and Tel1, but is abolished by deletion of the spindle assembly checkpoint genes MAD2 or BUB1 [54]. These types of checkpoint crosstalk may provide a mechanism for generating a more robust cell cycle arrest phenotype.

Checkpoint activation also impinges upon the Dbf4-dependent kinase (DDK), which activates origins of replication [55]. In budding yeast, Dbf4 is phosphorylated in a MEC1-dependent fashion in strains with persistant DNA damage [56]. Similarly, Dbf4 phosphorylation after HU treatment is RAD53-dependent [57]. The in vivo role of these modifications has not been established, but it is attractive to speculate that this modification might suppress late origin firing in the replication checkpoint response. Remarkably in both budding and fission yeasts, it has been suggested that complete activation of Rad53 requires DDK [58, 59], consistent with a notion that Rad53 activation is sensitive to the efficiency of origin firing [60]. This result, however, is controversial and several studies have shown efficient cell cycle arrest in the DDK deletion background [57, 61]. Remarkably, bypassing the lethality of DDK deletions via mcm5bob1 requires the kinase activity of Rad53, but not Mec1 or Tel1, which has been used to argue that Rad53 has a DNA damage-independent role in cellular replication [60]; however, it is not clear if this activity is important in wild-type cells.

2.2 DNA repair targets of checkpoint activation

The arrest of the cell cycle is likely required to allow DNA repair to occur. Numerous proteins directly involved in this repair have been identified as targets of the checkpoint kinases (Table 2). In some cases, such as Rad55, Rtt107, and Nej1, the phosphorylation appears to play important functional roles [62-65]. In others, such as Rfa1, Rfa2, and Xrs2, the role of the phosphorylation during mitotic DNA repair is less clear [66-69]. Taken as a whole, the data accumulated to date (Table 2) are insufficient to define the specific pathways targeted by checkpoint activation. Similarly, the data are not sufficient to discern potential differences in the downstream responses to specific types of DNA damage. The data do, however, indicate that the activities of at least some DNA repair proteins in S. cerevisiae are modified after DNA damage, suggesting a post-translational regulatory mechanism for DNA repair proteins rather distinct from the classic transcriptional SOS response in bacteria driven by DNA damage [70].

Table 2. Known DNA repair substrates of checkpoint activation.

Substrate Kinase Effect of mutagenizing sites References
Rfa1 Mec1, Tel1 Meiotic crossover effects (122SQ->DQ); no effect (122SQ->AQ) [66, 67]
Rfa2 Mec1, Tel1 No effect (SQ/TQ->AQ)* [67, 68]
Xrs2 Tel1 No effect (SQ/TQ->AQ) [68, 69]
Mre11 Tel1 Untested [68, 131]
Rad55 Rad53; possibly Mec1, Tel1 Slow S-phase traversal during damage (S->A) [62]
Slx4 Mec1, Tel1 Defect in single-strand annealing, but no defect with sgs1Δ or sensitivity to DNA damaging agents (SQ->AQ) [132]
Rtt107 (Esc4) Mec1; requires Slx4 Sensitivity to DNA damaging agents (S->A) [63, 64]
Nej1 Dun1 Decrease in NHEJ (S->A), No effect (S->E) [65]
Exo1 Kinase pathway; possibly Rad53 Growth defects with cdc13-1 (S->A); No effect with cdc13-1 (S->E) [133]
Srs2 Kinase pathway + Cdk1 Untested [134]
Sae2 Mec1, Tel1 MMS hypersensitivity, synthetic lethality rad27Δ, decreased mitotic recombination (SQ/TQ->AQ) [135]
Cdc13 Mec1, Tel1 Telomere shortening, senescence (SQ/TQ->AQ); Stable short telomeres (SQ/TQ->DQ/EQ) [136]
Mdc1 (Scc1) Chk1 Defect in forming DSB-induced cohesion (83S->A); Cohesion formed in absence of DSB (83S->D) [85]
*

In mammalian cells, phosphomimicking mutations have impaired ability to associate with replication centers; checkpoint phosphorylation proposed to contribute to inhibition of DNA replication [137].

2.3 Chromatin targets of the damage checkpoint

One of the more important direct substrates for the activated Mec1 and Tel1 is histone H2A Ser129 [11], which is equivalent to phosphorylation of the alternative metazoan H2A subunit H2AX on Ser 139, called γ-H2AX, and has been extensively used as an experimental surrogate for checkpoint activation. In response to an induced DSB, Mec1 and Tel1 generate a unique γ-H2AX-containing chromatin domain that can extend at least 10 kb from the lesion [71]. In metazoans, this modification suppresses the formation of chromosomal rearrangements due to DSBs [72] but as yet there is little evidence for this in budding yeast [73].

In yeast, γ-H2AX domains recruit the Ino80, Swr1, and NuA4 chromatin modifying complexes via the Nhp10 and/or Arp4 subunits [71, 74, 75]. It has been suggested that Ino80 and the Swr1 chromatin remodeling complexes act antagonistically, with Swr1 replacing γ-H2AX with the alternative H2A subunit Htz1, although accumulation of Htz1 near the break only occurs in an ino80Δ strain [76]. In contrast, other experiments have suggested that Ino80 facilitates ssDNA formation by Mre11-Rad50-Xrs2 at DSBs to promote HR and that Swr1 complex recruits Ku80 to promote NHEJ [77]. Disruption of the NuA4, Swr1, and Ino80 complexes had no effect on cell cycle arrest at G1/S, although a NuA4 mutation caused a persistent checkpoint response [78]. The Ino80 complex subunit Ies4 is also a Mec1 and Tel1 substrate [79]; however, the role of this phosphorylation remains unclear. Overall, the exact roles of these chromatin remodeling complexes as part of the damage response is also currently unclear; mutations affecting these complexes have had little or no effect on assayed DNA repair reactions.

Post-replicative assembly of cohesin domains holding together sister chromatids occurs after treatment with ionizing radiation or induction of a DSB and is also dependent upon γ-H2AX domain formation and is important for efficient homologous recombination [80, 81]. Cohesin recruitment at the break site is also dependent upon the cohesin loading gene Scc2, Mre11 and partially on Rad53 [81]. Cohesin recruitment induced by breaks is propagated throughout the entire genome in a MRE11-, MEC1-dependent and a RAD52- and RAD9-independent fashion and also requires the acetyltransferase function of Eco1/Ctf7 [82, 83], which is dispensable for normal cohesin loading during S-phase [84]. Inhibiting genome-wide cohesion between sister chromatids causes unbroken chromosomes to be lost at three-fold higher rates [83]. Phosphorylation of Mdc1 (Scc1) by Chk1 also appears to be required for DSB-induced sister chromatid cohesion [85] and directly links the checkpoint kinases to this process.

2.4 Cytoplasmic targets of the checkpoint response

The known checkpoint targets described above are nuclear, and the introduction of a single unrepairable DSB within one nucleus of a budding yeast heterokaryon has established that the resulting G2/M arrest does not influence DNA replication in the other nucleus in the cell [86]. Despite this, the effects of the checkpoint activation are integrated with a cytoplasmic response; rad53Δ strains have significant morphological defects [87], and Rad53 phosphorylates a number of cytoplasmic targets including septins [48]. Deletion of RAD53 or CHK1 also cause inappropriate movement of the nucleus into the bud neck in a dynein-dependent fashion prior to anaphase [88]. Thus it is clear that Rad53 is a component that helps coordinate the checkpoint response with the larger cell cycle response of the entire cell.

2.5 Transcriptional targets of the damage checkpoint

The transcriptional response of a number of individual S. cerevisiae genes relevant to repair of DNA damage has been well characterized as part of the checkpoint response. The RNR1, RNR2, RNR3, and RNR4 genes, which encode subunits of ribonucleotide reductase, are normally repressed by Crt1, and this repression is alleviated by hyperphosphorylation by Dun1 [89, 90]. Similarly, repression of PHR1, encoding photolyase, is alleviated during DNA damage by phosphorylation of Rph1 in a manner dependent on RAD17, MEC1, RAD9, and RAD53 and independent of CHK1 [91]. MAG1, encoding 3-methyladenine DNA glycosylase, and DDI1 are expressed from a single divergent promoter, and MAG1 expression has some dependence on the central checkpoint kinases and Rad6-Rad18-dependent ubiquitination of Rad17 [92-94], although transcription factors mediating this response have not been identified. Additionally, expression of CLN1 and CLN2, encoding cyclins, and transition through start are inhibited by methylmethanesulfonate (MMS) treatment in a mechanism partially dependent upon RAD53 and putative Rad53 phosphorylation sites on the Swi6 transcription factor [95, 96].

The relevance of the individually analyzed transcriptional targets to the DNA repair response has led to a number of unfortunate parallels with the highly specific and coordinated bacterial SOS response to DNA damage [70], which uses a single mechanism to activate transcription of a number of critical operons encoding DNA repair factors. An early observation arguing against yeast employing a SOS-like response was a lack of concordance between genes upregulated in response to DNA damaging agents and the genes that conferred resistance to those agents [97]. Numerous large-scale transcriptional profiles of S. cerevisiae under conditions of DNA damaging agents have now been undertaken (for example [89, 98, 99]), and analysis of one study [99] that identified 293 up-regulated genes (Table 3) and 220 down-regulated genes (Table 4) in the cellular response to MMS exposure is illustrative.

Table 3. Genes upregulated under MMS treatment.

Category Total Cell cycle* ESR** TF***
Yes No Yes No Yes No
Energy metabolism 37 6 31 25 12 22 15
Fatty acid and lipid metabolism 5 0 5 2 3 1 4
Amino acid biosynthesis and metabolism 11 3 8 1 10 9 2
Nucleotide biosynthesis and metabolism 6 3 3 1 5 6 0
DNA replication/repair and chromosome structure 10 7 3 1 9 4 4
Transcription, RNA metabolism, and modification 9 1 8 5 4 8 1
Protein translation and ribosome biogenesis 0 0 0 0 0 0 0
Protein folding and trafficking 14 0 14 8 6 6 8
Proteolysis 32 2 30 6 26 9 23
Signaling 10 1 9 5 5 9 1
Stress responses (oxidative, osmotic, drug, general) 39 7 32 22 17 26 13
Organelle and vesicle transport 12 0 12 3 9 4 8
Cell structure 7 2 5 1 6 4 3
Other 18 5 13 3 15 8 10
Unknown 83 10 73 42 41 56 27
All 293 47 246 125 168 174 119
*

Genes shown to have cell-cycle regulated transcription [100].

**

The ∼900 genes associated with the Environmental Stress Response [101].

***

Transcriptional response affected by deletion of one of several genes encoding transcription factors that contribute to MMS resistance [99].

Table 4. Genes downregulated under MMS treatment.

Category Total 1 Cell cycle* ESR** TF***
Yes No Yes No Yes No
Energy metabolism 17 4 13 1 16 8 9
Fatty acid and lipid metabolism 12 5 7 2 10 5 7
Amino acid biosynthesis and metabolism 9 4 5 2 7 5 4
Nucleotide biosynthesis and metabolism 12 1 11 6 6 7 5
DNA replication/repair and chromosome structure 4 4 0 0 4 1 3
Transcription, metabolism, and modification 18 6 12 7 11 7 11
Protein translation and ribosome biogenesis 65 0 65 53 12 32 33
Protein folding and trafficking 2 0 2 0 2 1 1
Proteolysis 5 2 3 1 4 3 2
Signaling 11 8 3 1 10 4 7
Stress responses (oxidative, osmotic, drug, general) 3 1 2 0 3 2 1
Organelle and vesicle transport 3 2 1 1 2 2 1
Cell structure 16 13 3 3 13 4 12
Other 8 4 4 2 6 11 5
Unknown 35 10 25 12 23 15 20
All 220 64 156 91 129 106 114
*

Genes shown to have cell-cycle regulated transcription [100].

**

The ∼900 genes associated with the Environmental Stress Response [101].

***

Transcriptional response affected by deletion of one of several genes encoding transcription factors that contribute to MMS resistance [99].

In the 513 genes with altered transcription, only 10 up-regulated genes are involved in DNA replication and repair processes (CDC13, DIN7, MAG1, MDC1, MSH2, POL30, RAD51, RAD54, RFA1, and RFA2; Table 3). The vast majority of genes appear to result from multiple independent responses. Around a fourth of the genes are cell-cycle regulated [100] and therefore might be affected indirectly by DNA damage, including 6 of the 10 DNA replication and repair genes (MSH2, POL30, RAD51, RAD54, RFA1, and RFA2). The MMS response also partially overlaps with the ∼900 genes in the “environmental stress response” (ESR) that display differential expression in response to a variety of external stresses [101]. Rougly half of the genes affected by MMS are independent of transcription factors encoded by genes that when deleted cause MMS sensitivity [99]. In fact, a major aspect of the MMS response is halting protein synthesis and activating proteolysis, consistent with results from other studies involving multiple DNA damaging agents [98, 102].

Some, but not all, of the above transcriptional responses are likely due the influence of the checkpoint response. Mutations of MEC1 and DUN1 affected the transcriptional response of roughly half of the ∼2000 genes altered by ionizing radiation and MMS treatment in another study [89]. Of the ∼1000 checkpoint-dependent genes, only 9 (DIN7, DUN1, PCM2, RAD51, RAD54, RNR2, RNR4, YBR070C, and YER004W) could not be attributed to a general stress response or to indirect regulation through loss of checkpoint function. Importantly, induction of the ESR genes was MEC1-dependent, suggesting that the DNA damage and replication checkpoints are responsible for more than simply halting the cell cycle, whereas other responses such as induction of protein degradation genes were unaffected by deletion of MEC1 or DUN1.

Taken together, these data provide evidence that distinguish the eukaryotic transcriptional response to DNA damage from the bacterial SOS response, despite claims to the contrary [94]. Overall, the induced genes have very little to do with DNA repair per se, which suggests that the checkpoint control of DNA repair is primarily mediated by post-translational modifications of pre-existing proteins (described above). Additionally, these transcriptional responses lack a single master regulator of the response similar to LexA in the bacterial SOS response. Neither MEC1 [89] nor RAD6 [94] are completely responsible for the diverse range of observed transcriptional effects, and there is evidence for multiple downstream transcription factors [89]. Finally, the transcriptional response appears coordinated with more general stress responses and may reflect the tendancy for DNA damage to coincide with other stresses in environments outside of the laboratory, similar to the fact that the resistance of the bacterium Deinococcus radiodurans to ionizing radiation is likely incidental to its ability to survive prolonged desiccation [103].

3. Do DNA damage and replication checkpoints have non-damage related roles during DNA replication?

To date, most studies on induction of checkpoint responses have relied upon a number of methods for artificially perturbing cells, such as induction of DSBs via endonucleases, induction of DNA damage, or induction of replication stress. These methods have been very useful for identifying key genes and mechanistic features of checkpoint responses; however, the artificial nature of these treatments may be obscuring the normal role of checkpoint signaling in a typical cell cycle.

Intermediates formed during the normal replication process, such as ssDNA, are substrates known to activate the DNA damage checkpoints. This might be a coincidence due to the structure of DNA itself; however, Mec1 and Rad53 appear to play roles affecting DNA replication in the absence of exogenous damage. These kinases are antagonistic to origin firing: late firing replication origins fire earlier in strains with rad53 and mec1 mutations [3], the mec1-21 mutation suppresses the origin firing defects of an orc2-1 mutation [104], and strains with deletions of MEC1 or RAD53 complete S-phase more quickly than wild-type cells during MMS treatment only due to the firing of late and cryptic origins [105]. These kinases also have roles in replication fork progression: mec1 and rad53 mutations have slower fork progression, elongation defects, and difficulty in completely duplicating DNA [105, 106], and MEC1 prevents fork stalling and chromosome fragmentation at “replication slow zones”, which may be due to dNTP depletion at the end of S-phase as these defects are suppressed by SML1 deletion [107]. In addition to Mec1 and Rad53, Dun1 increases dNTP pools by degrading Sml1 and by promoting the transcription of RNR subunits [46, 89, 90]. Moreover, numerous proteins implicated as damage sensors or signal transducers are components of the replication fork or are involved in polymerase loading, including Mrc1, Tof1, Csm3, Pol2, Pri1, Rfc2, Rfc5, Dpb11, and Sld2 [108-110]. Components of the DNA replication checkpoint may be part of a normal mechanism that allows active replication forks to regulate the replication of adjacent DNA, but likely do not signal the completion of DNA replication. Strains with smc6 mutations initiate anaphase prior to completion of the replication of the entire genome, despite the presence of a functional DNA damage checkpoint response [111]. One intriguing possibility is that the difference between “replication control” and “damage signaling” may be the persistence of the signal at any site along the DNA, providing an elegant mechanism for recognizing inappropriately halted replication forks. Consistent with this, checkpoint defects result in spontaneous genome rearrangements in the absence of exogenously induced DNA damage [5].

Insights into the role of checkpoint signaling during normal replication may be provided through studies that stall replication forks via nucleotide depletion due to HU treatment. Mutants lacking RAD53 and MEC1 are highly sensitive to transient exposure to HU [112], and treatment of rad53 mutants with HU is associated with the formation of “regressed forks” observed by electron microscopy [113]; however, these structures are formed in cells that never recover from HU and therefore do not necessarily represent repair intermediates. In the presence of functional checkpoints, HU treatment does not induce increased negative supercoiling during plasmid replication, indicating that the Mcm2-7 replicative helicase is associated with the rest of the fork [114], and genetic results are consistent with a role of MRC1 in arresting Mcm2-7 and Cdc45 in response to HU treatment in fission yeast [115]. Remarkably, deletion of MEC1 or RAD53 does not affect pausing of replisomes stalled at protein barriers or recovery from this pausing [116], which contrasts greatly from the failure of mec1 and rad53 mutants to recover after transient HU exposure. Thus, the ultimate target for replisome stabilization by checkpoints may be maintenance of the association of the Mcm2-7 and/or Cdc45 with the rest of the replication fork: the helicase is present at stably stalled forks [116]; these complexes are the major regulatory target for preventing re-replication of the genome [117]; and temperature shifting the degron-tagged version of the Mcm2-7 complex in S-phase, but not G1-phase is lethal, indicating that loss of the complex during DNA replication is irreversible [118]. That S. cerevisiae appears to lack a reloading mechanism for the repliactive helicase reminiscent of the multiple PriABC pathways in bacteria [119] may not be so surprising given that bacteria have only a single bidirectional origin whereas S. cerevisiae has replication origins roughly every 40 kbp [120], making avoidance of re-replication, and not fork reloading, the primary concern during DNA replication in S. cerevisiae.

4. Do DNA damage and replication checkpoints function by positive feedback?

The checkpoint signal leading to cell cycle arrest must either be on or off. Mechanistically, both electronic and biological circuits mediate these responses through positive feedback [121], which feeds the downstream signal back into upstream amplification components. Positive feedback has recently been established in S. cerevisiae in the control of the G1/S transition in coordinating the expression of S-phase specific genes [122] and the initiation of anaphase in coordinating the degradation of Pds1, leading to sister chromatid separation [123]. Although this has not been directly tested, there is some evidence that would be consistent with a positive feedback mechanism in DNA damage and replication checkpoint responses.

The first feature of positive feedback is that the signal needs to be amplified by upstream components. Maintenance of a MEC1- and RAD53-dependent G2/M arrest due to an induced double-strand break is dependent upon resection of the break in a Cdk1 and Mre11-Rad50-Xrs2 dependent fashion, which presumably continuously generates more RPA-ssDNA filaments [13, 14]. Elimination of Cdk1 activity abolishes the checkpoint and continued resection from the DSB; however, it remains unclear whether newly liberated ssDNA and additional Mec1-Ddc2-RPA binding or some aspect of continued resection is the amplified signal. Similarly in metazoans, ATM (the Tel1-homolog) generates phosphorylated γ-H2AX, and ATM is also recruited to DNA marked by phosphorylated γ-H2AX via the adaptor protein MDC1, which provides a mechanism for amplifying the ATM signal [124]. Moreover, signal amplification would be anticipated to combine multiple input signals. In an orc2-1 background, which has 30% of the normal replication forks, activation of the intra-S checkpoint shows a dosage dependence for the number of stalled forks and can be combined with bleomycin-induced damage to enhance the overall signaling [60].

The second feature of biological positive feedback is a mechanism to squelch the signal until it crosses an activating threshold. An argument can be made for this aspect of the mechanism based on the fact that the DNA damage driving the signal appears to function only as a passive scaffold to mediate colocalization of DNA damage sensors [18, 125]. Thus, transient colocalization of damage sensors in cells could lead to inappropriate activation of the checkpoint if a silencing mechanism did not exist to dampen the positive feedback response until the initiating signal reached an activating threshold. A number of candidate proteins for suppressing the DNA damage signal exist, including the PP2A protein phosphatase complex Psy2-Psy4-Pph3, which dephosphorylates Rad53 and γ-H2AX when γ-H2AX is displaced from chromatin [126, 127], and the PPC2-like protein phosphatases Ptc2 and Ptc3, which are required for “adaptation” to a constitutively activated DNA damage checkpoint and may also dephosphorylate Rad53 [128]. Remarkably, overexpression of Rad53 suppresses the ability of cells to adapt [129]; however, it remains to be seen if the genetic requirements for adaptation match those for suppression of a DNA damage positive feedback loop.

The third and final feature required for positive feedback during checkpoint activation is distinct from the positive feedback loops identified in the G1/S transition and the initiation of anaphase: the DNA damage signal must be reversible. Thus, the “sensors” for the DNA damage must also be an integral part of the feedback amplification so that the signal can be silenced when repair is complete. Intriguingly, shifting a temperature-sensitive degron-Mec1 fusion to the non-permissive temperature decreases Rad53 activity even in the presence of persistent DNA damage [129], and inhibition of oligomerization of Rad9 via interactions of Rad9 BRCT domains with Rad9 molecules phosphorylated at Mec1/Tel1 consensus sites is required for the maintenance but not initial activation of a DNA damage checkpoint reponse [130].

The presence of a positive feedback loop in the DNA damage checkpoint changes the nature of the experiments that need to be pursued. First, the proteins that control the normally suppressive environment need to be closely investigated. Although we know some details of the phenotypes of some of the candidate genes that could be involved in suppressing the positive feedback signal, until their roles are clearly established, a diagnostic phenotype caused by loss of function of the genes mediating basal checkpoint suppression has not been identified; it is currently unclear if these genes should be essential, show synthetic lethality with adaptation mutations, show basal activation of the DNA damage response, and/or show hypersensitivity to DNA damaging agents due to an inability to inactivate a checkpoint after damage is repaired. Second, the biological implications of modification of “upstream” sensors in checkpoint maintenance, such as RPA and replication fork subunits, need to be explored. Finally, in both studies that identified a positive feedback response in the G1/S transition and anaphase initiation, the heterogeneity in population studies averaged out the evidence for positive feedback, suggesting that future studies may also require monitoring individual cells [122, 123].

Acknowledgments

The authors thank Jorrit Ensernik, Huilin Zhou, and Wolf-Dietrich Heyer for their generous comments. Supported by NIH grants GM26017 and ES014811.

Footnotes

5. Conflict of interest statement. The authors declare that there are no conflicts of interest.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Weinert TA, Hartwell LH. The RAD9 gene controls the cell cycle response to DNA damage in Saccharomyces cerevisiae. Science. 1988;241:317–322. doi: 10.1126/science.3291120. [DOI] [PubMed] [Google Scholar]
  • 2.Siede W, Friedberg AS, Friedberg EC. RAD9-dependent G1 arrest defines a second checkpoint for damaged DNA in the cell cycle of Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 1993;90:7985–7989. doi: 10.1073/pnas.90.17.7985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Santocanale C, Diffley JF. A Mec1- and Rad53-dependent checkpoint controls late-firing origins of DNA replication. Nature. 1998;395:615–618. doi: 10.1038/27001. [DOI] [PubMed] [Google Scholar]
  • 4.Paulovich AG, Margulies RU, Garvik BM, Hartwell LH. RAD9, RAD17, and RAD24 are required for S phase regulation in Saccharomyces cerevisiae in response to DNA damage. Genetics. 1997;145:45–62. doi: 10.1093/genetics/145.1.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Myung K, Kolodner RD. Suppression of genome instability by redundant S-phase checkpoint pathways in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2002;99:4500–4507. doi: 10.1073/pnas.062702199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Zou L, Elledge SJ. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science. 2003;300:1542–1548. doi: 10.1126/science.1083430. [DOI] [PubMed] [Google Scholar]
  • 7.Nakada D, Matsumoto K, Sugimoto K. ATM-related Tel1 associates with double-strand breaks through an Xrs2-dependent mechanism. Genes Dev. 2003;17:1957–1962. doi: 10.1101/gad.1099003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Ritchie KB, Mallory JC, Petes TD. 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. 1999;19:6065–6075. doi: 10.1128/mcb.19.9.6065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Chan SW, Chang J, Prescott J, Blackburn EH. Altering telomere structure allows telomerase to act in yeast lacking ATM kinases. Curr Biol. 2001;11:1240–1250. doi: 10.1016/s0960-9822(01)00391-8. [DOI] [PubMed] [Google Scholar]
  • 10.Morrow DM, Tagle DA, Shiloh Y, Collins FS, Hieter P. TEL1, an S. cerevisiae homolog of the human gene mutated in ataxia telangiectasia, is functionally related to the yeast checkpoint gene MEC1. Cell. 1995;82:831–840. doi: 10.1016/0092-8674(95)90480-8. [DOI] [PubMed] [Google Scholar]
  • 11.Downs JA, Lowndes NF, Jackson SP. A role for Saccharomyces cerevisiae histone H2A in DNA repair. Nature. 2000;408:1001–1004. doi: 10.1038/35050000. [DOI] [PubMed] [Google Scholar]
  • 12.Lustig AJ, Petes TD. Identification of yeast mutants with altered telomere structure, Proc. Natl Acad Sci U S A. 1986;83:1398–1402. doi: 10.1073/pnas.83.5.1398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ira G, Pellicioli A, Balijja A, Wang X, Fiorani S, Carotenuto W, Liberi G, Bressan D, Wan L, Hollingsworth NM, Haber JE, Foiani M. DNA end resection, homologous recombination and DNA damage checkpoint activation require CDK1. Nature. 2004;431:1011–1017. doi: 10.1038/nature02964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Huertas P, Cortes-Ledesma F, Sartori AA, Aguilera A, Jackson SP. CDK targets Sae2 to control DNA-end resection and homologous recombination. Nature. 2008;455:689–692. doi: 10.1038/nature07215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kondo T, Wakayama T, Naiki T, Matsumoto K, Sugimoto K. Recruitment of Mec1 and Ddc1 checkpoint proteins to double-strand breaks through distinct mechanisms. Science. 2001;294:867–870. doi: 10.1126/science.1063827. [DOI] [PubMed] [Google Scholar]
  • 16.Melo JA, Cohen J, Toczyski DP. Two checkpoint complexes are independently recruited to sites of DNA damage in vivo. Genes Dev. 2001;15:2809–2821. doi: 10.1101/gad.903501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Majka J, Niedziela-Majka A, Burgers PM. The checkpoint clamp activates Mec1 kinase during initiation of the DNA damage checkpoint. Mol Cell. 2006;24:891–901. doi: 10.1016/j.molcel.2006.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Bonilla CY, Melo JA, Toczyski DP. Colocalization of sensors is sufficient to activate the DNA damage checkpoint in the absence of damage. Mol Cell. 2008;30:267–276. doi: 10.1016/j.molcel.2008.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Lydall D. Hiding at the ends of yeast chromosomes: telomeres, nucleases and checkpoint pathways. J Cell Sci. 2003;116:4057–4065. doi: 10.1242/jcs.00765. [DOI] [PubMed] [Google Scholar]
  • 20.Mordes DA, Nam EA, Cortez D. Dpb11 activates the Mec1-Ddc2 complex, Proc. Natl Acad Sci U S A. 2008;105:18730–18734. doi: 10.1073/pnas.0806621105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Navadgi-Patil VM, Burgers PM. Yeast DNA replication protein Dpb11 activates the Mec1/ATR checkpoint kinase. J Biol Chem. 2008;283:35853–35859. doi: 10.1074/jbc.M807435200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Delacroix S, Wagner JM, Kobayashi M, Yamamoto K, Karnitz LM. The Rad9-Hus1-Rad1 (9-1-1) clamp activates checkpoint signaling via TopBP1. Genes Dev. 2007;21:1472–1477. doi: 10.1101/gad.1547007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Furuya K, Poitelea M, Guo L, Caspari T, Carr AM. Chk1 activation requires Rad9 S/TQ-site phosphorylation to promote association with C-terminal BRCT domains of Rad4TOPBP1. Genes Dev. 2004;18:1154–1164. doi: 10.1101/gad.291104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Lee J, Kumagai A, Dunphy WG. The Rad9-Hus1-Rad1 checkpoint clamp regulates interaction of TopBP1 with ATR. J Biol Chem. 2007;282:28036–28044. doi: 10.1074/jbc.M704635200. [DOI] [PubMed] [Google Scholar]
  • 25.Amon A, Surana U, Muroff I, Nasmyth K. Regulation of p34CDC28 tyrosine phosphorylation is not required for entry into mitosis in S. cerevisiae. Nature. 1992;355:368–371. doi: 10.1038/355368a0. [DOI] [PubMed] [Google Scholar]
  • 26.Sorger PK, Murray AW. S-phase feedback control in budding yeast independent of tyrosine phosphorylation of p34cdc28. Nature. 1992;355:365–368. doi: 10.1038/355365a0. [DOI] [PubMed] [Google Scholar]
  • 27.Emili A, Schieltz DM, Yates JR, 3rd, Hartwell LH. Dynamic interaction of DNA damage checkpoint protein Rad53 with chromatin assembly factor Asf1. Mol Cell. 2001;7:13–20. doi: 10.1016/s1097-2765(01)00150-2. [DOI] [PubMed] [Google Scholar]
  • 28.Hu F, Alcasabas AA, Elledge SJ. Asf1 links Rad53 to control of chromatin assembly. Genes Dev. 2001;15:1061–1066. doi: 10.1101/gad.873201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Sharp JA, Rizki G, Kaufman PD. Regulation of histone deposition proteins Asf1/Hir1 by multiple DNA damage checkpoint kinases in Saccharomyces cerevisiae. Genetics. 2005;171:885–899. doi: 10.1534/genetics.105.044719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Chen CC, Carson JJ, Feser J, Tamburini B, Zabaronick S, Linger J, Tyler JK. Acetylated lysine 56 on histone H3 drives chromatin assembly after repair and signals for the completion of repair. Cell. 2008;134:231–243. doi: 10.1016/j.cell.2008.06.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Alcasabas AA, Osborn AJ, Bachant J, Hu F, Werler PJ, Bousset K, Furuya K, Diffley JF, Carr AM, Elledge SJ. Mrc1 transduces signals of DNA replication stress to activate Rad53. Nat Cell Biol. 2001;3:958–965. doi: 10.1038/ncb1101-958. [DOI] [PubMed] [Google Scholar]
  • 32.Schwartz MF, Duong JK, Sun Z, Morrow JS, Pradhan D, Stern DF. Rad9 phosphorylation sites couple Rad53 to the Saccharomyces cerevisiae DNA damage checkpoint. Mol Cell. 2002;9:1055–1065. doi: 10.1016/s1097-2765(02)00532-4. [DOI] [PubMed] [Google Scholar]
  • 33.Osborn AJ, Elledge SJ. Mrc1 is a replication fork component whose phosphorylation in response to DNA replication stress activates Rad53. Genes Dev. 2003;17:1755–1767. doi: 10.1101/gad.1098303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Lee SJ, Duong JK, Stern DF. A Ddc2-Rad53 fusion protein can bypass the requirements for RAD9 and MRC1 in Rad53 activation. Mol Biol Cell. 2004;15:5443–5455. doi: 10.1091/mbc.E04-07-0608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Tanaka K, Russell P. Cds1 phosphorylation by Rad3-Rad26 kinase is mediated by forkhead-associated domain interaction with Mrc1. J Biol Chem. 2004;279:32079–32086. doi: 10.1074/jbc.M404834200. [DOI] [PubMed] [Google Scholar]
  • 36.Katou Y, Kanoh Y, Bando M, Noguchi H, Tanaka H, Ashikari T, Sugimoto K, Shirahige K. S-phase checkpoint proteins Tof1 and Mrc1 form a stable replication-pausing complex. Nature. 2003;424:1078–1083. doi: 10.1038/nature01900. [DOI] [PubMed] [Google Scholar]
  • 37.Soulier J, Lowndes NF. The BRCT domain of the S. cerevisiae checkpoint protein Rad9 mediates a Rad9-Rad9 interaction after DNA damage. Curr Biol. 1999;9:551–554. doi: 10.1016/s0960-9822(99)80242-5. [DOI] [PubMed] [Google Scholar]
  • 38.Hammet A, Magill C, Heierhorst J, Jackson SP. Rad9 BRCT domain interaction with phosphorylated H2AX regulates the G1 checkpoint in budding yeast. EMBO Rep. 2007;8:851–857. doi: 10.1038/sj.embor.7401036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Grenon M, Costelloe T, Jimeno S, O'Shaughnessy A, Fitzgerald J, Zgheib O, Degerth L, Lowndes NF. Docking onto chromatin via the Saccharomyces cerevisiae Rad9 Tudor domain. Yeast. 2007;24:105–119. doi: 10.1002/yea.1441. [DOI] [PubMed] [Google Scholar]
  • 40.Giannattasio M, Lazzaro F, Plevani P, Muzi-Falconi M. The DNA damage checkpoint response requires histone H2B ubiquitination by Rad6-Bre1 and H3 methylation by Dot1. J Biol Chem. 2005;280:9879–9886. doi: 10.1074/jbc.M414453200. [DOI] [PubMed] [Google Scholar]
  • 41.Wysocki R, Javaheri A, Allard S, Sha F, Cote J, Kron SJ. Role of Dot1-dependent histone H3 methylation in G1 and S phase DNA damage checkpoint functions of Rad9. Mol Cell Biol. 2005;25:8430–8443. doi: 10.1128/MCB.25.19.8430-8443.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Puddu F, Granata M, Nola LDi, Balestrini A, Piergiovanni G, Lazzaro F, Giannattasio M, Plevani P, Muzi-Falconi M. Phosphorylation of the budding yeast 9-1-1 complex is required for Dpb11 function in the full activation of the UV-induced DNA damage checkpoint. Mol Cell Biol. 2008;28:4782–4793. doi: 10.1128/MCB.00330-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Carballo JA, Johnson AL, Sedgwick SG, Cha RS. Phosphorylation of the axial element protein Hop1 by Mec1/Tel1 ensures meiotic interhomolog recombination. Cell. 2008;132:758–770. doi: 10.1016/j.cell.2008.01.035. [DOI] [PubMed] [Google Scholar]
  • 44.Bashkirov VI, Bashkirova EV, Haghnazari E, Heyer WD. Direct kinase-to-kinase signaling mediated by the FHA phosphoprotein recognition domain of the Dun1 DNA damage checkpoint kinase. Mol Cell Biol. 2003;23:1441–1452. doi: 10.1128/MCB.23.4.1441-1452.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Chen SH, Smolka MB, Zhou H. Mechanism of Dun1 activation by Rad53 phosphorylation in Saccharomyces cerevisiae. J Biol Chem. 2007;282:986–995. doi: 10.1074/jbc.M609322200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zhao X, Rothstein R. The Dun1 checkpoint kinase phosphorylates and regulates the ribonucleotide reductase inhibitor Sml1. Proc Natl Acad Sci U S A. 2002;99:3746–3751. doi: 10.1073/pnas.062502299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Zhao X, Muller EG, Rothstein R. A suppressor of two essential checkpoint genes identifies a novel protein that negatively affects dNTP pools. Mol Cell. 1998;2:329–340. doi: 10.1016/s1097-2765(00)80277-4. [DOI] [PubMed] [Google Scholar]
  • 48.Smolka MB, Chen SH, Maddox PS, Enserink JM, Albuquerque CP, Wei XX, Desai A, Kolodner RD, Zhou H. An FHA domain-mediated protein interaction network of Rad53 reveals its role in polarized cell growth. J Cell Biol. 2006;175:743–753. doi: 10.1083/jcb.200605081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Smolka MB, Albuquerque CP, Chen SH, Zhou H. Proteome-wide identification of in vivo targets of DNA damage checkpoint kinases. Proc Natl Acad Sci U S A. 2007;104:10364–10369. doi: 10.1073/pnas.0701622104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Sanchez Y, Bachant J, Wang H, Hu F, Liu D, Tetzlaff M, Elledge SJ. Control of the DNA damage checkpoint by chk1 and rad53 protein kinases through distinct mechanisms. Science. 1999;286:1166–1171. doi: 10.1126/science.286.5442.1166. [DOI] [PubMed] [Google Scholar]
  • 51.Agarwal R, Tang Z, Yu H, Cohen-Fix O. Two distinct pathways for inhibiting pds1 ubiquitination in response to DNA damage. J Biol Chem. 2003;278:45027–45033. doi: 10.1074/jbc.M306783200. [DOI] [PubMed] [Google Scholar]
  • 52.Liang F, Wang Y. DNA damage checkpoints inhibit mitotic exit by two different mechanisms. Mol Cell Biol. 2007;27:5067–5078. doi: 10.1128/MCB.00095-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Visintin R, Craig K, Hwang ES, Prinz S, Tyers M, Amon A. The phosphatase Cdc14 triggers mitotic exit by reversal of Cdk-dependent phosphorylation. Mol Cell. 1998;2:709–718. doi: 10.1016/s1097-2765(00)80286-5. [DOI] [PubMed] [Google Scholar]
  • 54.Clemenson C, Marsolier-Kergoat MC. The spindle assembly checkpoint regulates the phosphorylation state of a subset of DNA checkpoint proteins in Saccharomyces cerevisiae. Mol Cell Biol. 2006;26:9149–9161. doi: 10.1128/MCB.00310-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Duncker BP, Brown GW. Cdc7 kinases (DDKs) and checkpoint responses: lessons from two yeasts. Mutat Res. 2003;532:21–27. doi: 10.1016/j.mrfmmm.2003.08.007. [DOI] [PubMed] [Google Scholar]
  • 56.Cheng L, Collyer T, Hardy CF. Cell cycle regulation of DNA replication initiator factor Dbf4p. Mol Cell Biol. 1999;19:4270–4278. doi: 10.1128/mcb.19.6.4270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Weinreich M, Stillman B. Cdc7p-Dbf4p kinase binds to chromatin during S phase and is regulated by both the APC and the RAD53 checkpoint pathway. Embo J. 1999;18:5334–5346. doi: 10.1093/emboj/18.19.5334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Ogi H, Wang CZ, Nakai W, Kawasaki Y, Masumoto H. The role of the Saccharomyces cerevisiae Cdc7-Dbf4 complex in the replication checkpoint. Gene. 2008;414:32–40. doi: 10.1016/j.gene.2008.02.010. [DOI] [PubMed] [Google Scholar]
  • 59.Takeda T, Ogino K, Tatebayashi K, Ikeda H, Arai K, Masai H. Regulation of initiation of S phase, replication checkpoint signaling, and maintenance of mitotic chromosome structures during S phase by Hsk1 kinase in the fission yeast. Mol Biol Cell. 2001;12:1257–1274. doi: 10.1091/mbc.12.5.1257. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Shimada K, Pasero P, Gasser SM. ORC and the intra-S-phase checkpoint: a threshold regulates Rad53p activation in S phase. Genes Dev. 2002;16:3236–3252. doi: 10.1101/gad.239802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Pessoa-Brandao L, Sclafani RA. CDC7/DBF4 functions in the translesion synthesis branch of the RAD6 epistasis group in Saccharomyces cerevisiae. Genetics. 2004;167:1597–1610. doi: 10.1534/genetics.103.021675. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Herzberg K, Bashkirov VI, Rolfsmeier M, Haghnazari E, McDonald WH, Anderson S, Bashkirova EV, Yates JR, 3rd, Heyer WD. Phosphorylation of Rad55 on serines 2, 8, and 14 is required for efficient homologous recombination in the recovery of stalled replication forks. Mol Cell Biol. 2006;26:8396–8409. doi: 10.1128/MCB.01317-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Rouse J. Esc4p, a new target of Mec1p (ATR), promotes resumption of DNA synthesis after DNA damage. Embo J. 2004;23:1188–1197. doi: 10.1038/sj.emboj.7600129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Roberts TM, Kobor MS, Bastin-Shanower SA, Ii M, Horte SA, Gin JW, Emili A, Rine J, Brill SJ, Brown GW. Slx4 regulates DNA damage checkpoint-dependent phosphorylation of the BRCT domain protein Rtt107/Esc4. Mol Biol Cell. 2006;17:539–548. doi: 10.1091/mbc.E05-08-0785. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Ahnesorg P, Jackson SP. The non-homologous end-joining protein Nej1p is a target of the DNA damage checkpoint. DNA Repair (Amst) 2007;6:190–201. doi: 10.1016/j.dnarep.2006.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Bartrand AJ, Iyasu D, Marinco SM, Brush GS. Evidence of meiotic crossover control in Saccharomyces cerevisiae through Mec1-mediated phosphorylation of replication protein A. Genetics. 2006;172:27–39. doi: 10.1534/genetics.105.047845. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Brush GS, Kelly TJ. Phosphorylation of the replication protein A large subunit in the Saccharomyces cerevisiae checkpoint response. Nucleic Acids Res. 2000;28:3725–3732. doi: 10.1093/nar/28.19.3725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Mallory JC, Bashkirov VI, Trujillo KM, Solinger JA, Dominska M, Sung P, Heyer WD, Petes TD. Amino acid changes in Xrs2p, Dun1p, and Rfa2p that remove the preferred targets of the ATM family of protein kinases do not affect DNA repair or telomere length in Saccharomyces cerevisiae. DNA Repair (Amst) 2003;2:1041–1064. doi: 10.1016/s1568-7864(03)00115-0. [DOI] [PubMed] [Google Scholar]
  • 69.D'Amours D, Jackson SP. The yeast Xrs2 complex functions in S phase checkpoint regulation. Genes Dev. 2001;15:2238–2249. doi: 10.1101/gad.208701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Sutton MD, Smith BT, Godoy VG, Walker GC. The SOS response: recent insights into umuDC-dependent mutagenesis and DNA damage tolerance. Annu Rev Genet. 2000;34:479–497. doi: 10.1146/annurev.genet.34.1.479. [DOI] [PubMed] [Google Scholar]
  • 71.Downs JA, Allard S, Jobin-Robitaille O, Javaheri A, Auger A, Bouchard N, Kron SJ, Jackson SP, Cote J. Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites. Mol Cell. 2004;16:979–990. doi: 10.1016/j.molcel.2004.12.003. [DOI] [PubMed] [Google Scholar]
  • 72.Franco S, Gostissa M, Zha S, Lombard DB, Murphy MM, Zarrin AA, Yan C, Tepsuporn S, Morales JC, Adams MM, Lou Z, Bassing CH, Manis JP, Chen J, Carpenter PB, Alt FW. H2AX prevents DNA breaks from progressing to chromosome breaks and translocations. Mol Cell. 2006;21:201–214. doi: 10.1016/j.molcel.2006.01.005. [DOI] [PubMed] [Google Scholar]
  • 73.Myung K, Pennaneach V, Kats ES, Kolodner RD. Saccharomyces cerevisiae chromatin-assembly factors that act during DNA replication function in the maintenance of genome stability. Proc Natl Acad Sci U S A. 2003;100:6640–6645. doi: 10.1073/pnas.1232239100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Morrison AJ, Highland J, Krogan NJ, Arbel-Eden A, Greenblatt JF, Haber JE, Shen X. INO80 and gamma-H2AX interaction links ATP-dependent chromatin remodeling to DNA damage repair. Cell. 2004;119:767–775. doi: 10.1016/j.cell.2004.11.037. [DOI] [PubMed] [Google Scholar]
  • 75.van Attikum H, Fritsch O, Hohn B, Gasser SM. Recruitment of the INO80 complex by H2A phosphorylation links ATP-dependent chromatin remodeling with DNA double-strand break repair. Cell. 2004;119:777–788. doi: 10.1016/j.cell.2004.11.033. [DOI] [PubMed] [Google Scholar]
  • 76.Papamichos-Chronakis M, Krebs JE, Peterson CL. Interplay between Ino80 and Swr1 chromatin remodeling enzymes regulates cell cycle checkpoint adaptation in response to DNA damage. Genes Dev. 2006;20:2437–2449. doi: 10.1101/gad.1440206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.van Attikum H, Fritsch O, Gasser SM. Distinct roles for SWR1 and INO80 chromatin remodeling complexes at chromosomal double-strand breaks. Embo J. 2007;26:4113–4125. doi: 10.1038/sj.emboj.7601835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Javaheri A, Wysocki R, Jobin-Robitaille O, Altaf M, Cote J, Kron SJ. Yeast G1 DNA damage checkpoint regulation by H2A phosphorylation is independent of chromatin remodeling. Proc Natl Acad Sci U S A. 2006;103:13771–13776. doi: 10.1073/pnas.0511192103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Morrison AJ, Kim JA, Person MD, Highland J, Xiao J, Wehr TS, Hensley S, Bao Y, Shen J, Collins SR, Weissman JS, Delrow J, Krogan NJ, Haber JE, Shen X. Mec1/Tel1 phosphorylation of the INO80 chromatin remodeling complex influences DNA damage checkpoint responses. Cell. 2007;130:499–511. doi: 10.1016/j.cell.2007.06.010. [DOI] [PubMed] [Google Scholar]
  • 80.Strom L, Lindroos HB, Shirahige K, Sjogren C. Postreplicative recruitment of cohesin to double-strand breaks is required for DNA repair. Mol Cell. 2004;16:1003–1015. doi: 10.1016/j.molcel.2004.11.026. [DOI] [PubMed] [Google Scholar]
  • 81.Unal E, Arbel-Eden A, Sattler U, Shroff R, Lichten M, Haber JE, Koshland D. DNA damage response pathway uses histone modification to assemble a double-strand break-specific cohesin domain. Mol Cell. 2004;16:991–1002. doi: 10.1016/j.molcel.2004.11.027. [DOI] [PubMed] [Google Scholar]
  • 82.Strom L, Karlsson C, Lindroos HB, Wedahl S, Katou Y, Shirahige K, Sjogren C. Postreplicative formation of cohesion is required for repair and induced by a single DNA break. Science. 2007;317:242–245. doi: 10.1126/science.1140649. [DOI] [PubMed] [Google Scholar]
  • 83.Unal E, Heidinger-Pauli JM, Koshland D. DNA double-strand breaks trigger genome-wide sister-chromatid cohesion through Eco1 (Ctf7) Science. 2007;317:245–248. doi: 10.1126/science.1140637. [DOI] [PubMed] [Google Scholar]
  • 84.Brands A, Skibbens RV. Ctf7p/Eco1p exhibits acetyltransferase activity--but does it matter? Curr Biol. 2005;15:R50–51. doi: 10.1016/j.cub.2004.12.052. [DOI] [PubMed] [Google Scholar]
  • 85.Heidinger-Pauli JM, Unal E, Guacci V, Koshland D. The kleisin subunit of cohesin dictates damage-induced cohesion. Mol Cell. 2008;31:47–56. doi: 10.1016/j.molcel.2008.06.005. [DOI] [PubMed] [Google Scholar]
  • 86.Demeter J, Lee SE, Haber JE, Stearns T. The DNA damage checkpoint signal in budding yeast is nuclear limited. Mol Cell. 2000;6:487–492. doi: 10.1016/s1097-2765(00)00047-2. [DOI] [PubMed] [Google Scholar]
  • 87.Enserink JM, Smolka MB, Zhou H, Kolodner RD. Checkpoint proteins control morphogenetic events during DNA replication stress in Saccharomyces cerevisiae. J Cell Biol. 2006;175:729–741. doi: 10.1083/jcb.200605080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Dotiwala F, Haase J, Arbel-Eden A, Bloom K, Haber JE. The yeast DNA damage checkpoint proteins control a cytoplasmic response to DNA damage. Proc Natl Acad Sci U S A. 2007;104:11358–11363. doi: 10.1073/pnas.0609636104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Gasch AP, Huang M, Metzner S, Botstein D, Elledge SJ, Brown PO. Genomic expression responses to DNA-damaging agents and the regulatory role of the yeast ATR homolog Mec1p. Mol Biol Cell. 2001;12:2987–3003. doi: 10.1091/mbc.12.10.2987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Jelinsky SA, Samson LD. Global response of Saccharomyces cerevisiae to an alkylating agent. Proc Natl Acad Sci U S A. 1999;96:1486–1491. doi: 10.1073/pnas.96.4.1486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Kim EM, Jang YK, Park SD. Phosphorylation of Rph1, a damage-responsive repressor of PHR1 in Saccharomyces cerevisiae, is dependent upon Rad53 kinase. Nucleic Acids Res. 2002;30:643–648. doi: 10.1093/nar/30.3.643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Zhu Y, Xiao W. Two alternative cell cycle checkpoint pathways differentially control DNA damage-dependent induction of MAG1 and DDI1 expression in yeast. Mol Genet Genomics. 2001;266:436–444. doi: 10.1007/s004380100538. [DOI] [PubMed] [Google Scholar]
  • 93.Zhu Y, Xiao W. Differential regulation of two closely clustered yeast genes, MAG1 and DDI1, by cell-cycle checkpoints. Nucleic Acids Res. 1998;26:5402–5408. doi: 10.1093/nar/26.23.5402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Fu Y, Zhu Y, Zhang K, Yeung M, Durocher D, Xiao W. Rad6-Rad18 mediates a eukaryotic SOS response by ubiquitinating the 9-1-1 checkpoint clamp. Cell. 2008;133:601–611. doi: 10.1016/j.cell.2008.02.050. [DOI] [PubMed] [Google Scholar]
  • 95.Sidorova JM, Breeden LL. Rad53-dependent phosphorylation of Swi6 and down-regulation of CLN1 and CLN2 transcription occur in response to DNA damage in Saccharomyces cerevisiae. Genes Dev. 1997;11:3032–3045. doi: 10.1101/gad.11.22.3032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Sidorova JM, Breeden LL. Rad53 checkpoint kinase phosphorylation site preference identified in the Swi6 protein of Saccharomyces cerevisiae. Mol Cell Biol. 2003;23:3405–3416. doi: 10.1128/MCB.23.10.3405-3416.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Birrell GW, Brown JA, Wu HI, Giaever G, Chu AM, Davis RW, Brown JM. Transcriptional response of Saccharomyces cerevisiae to DNA-damaging agents does not identify the genes that protect against these agents. Proc Natl Acad Sci U S A. 2002;99:8778–8783. doi: 10.1073/pnas.132275199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Jelinsky SA, Estep P, Church GM, Samson LD. Regulatory networks revealed by transcriptional profiling of damaged Saccharomyces cerevisiae cells: Rpn4 links base excision repair with proteasomes. Mol Cell Biol. 2000;20:8157–8167. doi: 10.1128/mcb.20.21.8157-8167.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Workman T, Mak HC, McCuine S, Tagne JB, Agarwal M, Ozier O, Begley TJ, Samson LD, Ideker T. A systems approach to mapping DNA damage response pathways. Science. 2006;312:1054–1059. doi: 10.1126/science.1122088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Spellman PT, Sherlock G, Zhang MQ, Iyer VR, Anders K, Eisen MB, Brown PO, Botstein D, Futcher B. Comprehensive identification of cell cycle-regulated genes of the yeast Saccharomyces cerevisiae by microarray hybridization. Mol Biol Cell. 1998;9:3273–3297. doi: 10.1091/mbc.9.12.3273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Gasch AP, Spellman PT, Kao CM, Carmel-Harel O, Eisen MB, Storz G, Botstein D, Brown PO. Genomic expression programs in the response of yeast cells to environmental changes. Mol Biol Cell. 2000;11:4241–4257. doi: 10.1091/mbc.11.12.4241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Burgis NE, Samson LD. The protein degradation response of Saccharomyces cerevisiae to classical DNA-damaging agents. Chem Res Toxicol. 2007;20:1843–1853. doi: 10.1021/tx700126e. [DOI] [PubMed] [Google Scholar]
  • 103.Mattimore V, Battista JR. Radioresistance of Deinococcus radiodurans: functions necessary to survive ionizing radiation are also necessary to survive prolonged desiccation. J Bacteriol. 1996;178:633–637. doi: 10.1128/jb.178.3.633-637.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Desany BA, Alcasabas AA, Bachant JB, Elledge SJ. Recovery from DNA replicational stress is the essential function of the S-phase checkpoint pathway. Genes Dev. 1998;12:2956–2970. doi: 10.1101/gad.12.18.2956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Tercero JA, Diffley JF. Regulation of DNA replication fork progression through damaged DNA by the Mec1/Rad53 checkpoint. Nature. 2001;412:553–557. doi: 10.1038/35087607. [DOI] [PubMed] [Google Scholar]
  • 106.Raveendranathan M, Chattopadhyay S, Bolon YT, Haworth J, Clarke DJ, Bielinsky AK. Genome-wide replication profiles of S-phase checkpoint mutants reveal fragile sites in yeast. Embo J. 2006;25:3627–3639. doi: 10.1038/sj.emboj.7601251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Cha RS, Kleckner N. ATR homolog Mec1 promotes fork progression, thus averting breaks in replication slow zones. Science. 2002;297:602–606. doi: 10.1126/science.1071398. [DOI] [PubMed] [Google Scholar]
  • 108.Araki H, Leem SH, Phongdara A, Sugino A. Dpb11, which interacts with DNA polymerase II(epsilon) in Saccharomyces cerevisiae, has a dual role in S-phase progression and at a cell cycle checkpoint. Proc Natl Acad Sci U S A. 1995;92:11791–11795. doi: 10.1073/pnas.92.25.11791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Marini F, Pellicioli A, Paciotti V, Lucchini G, Plevani P, Stern DF, Foiani M. A role for DNA primase in coupling DNA replication to DNA damage response. Embo J. 1997;16:639–650. doi: 10.1093/emboj/16.3.639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Sugimoto K, Ando S, Shimomura T, Matsumoto K. Rfc5, a replication factor C component, is required for regulation of Rad53 protein kinase in the yeast checkpoint pathway. Mol Cell Biol. 1997;17:5905–5914. doi: 10.1128/mcb.17.10.5905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Torres-Rosell J, Piccoli GDe, Cordon-Preciado V, Farmer S, Jarmuz A, Machin F, Pasero P, Lisby M, Haber JE, Aragon L. Anaphase onset before complete DNA replication with intact checkpoint responses. Science. 2007;315:1411–1415. doi: 10.1126/science.1134025. [DOI] [PubMed] [Google Scholar]
  • 112.Allen JB, Zhou Z, Siede W, Friedberg EC, Elledge SJ. The SAD1/RAD53 protein kinase controls multiple checkpoints and DNA damage-induced transcription in yeast. Genes Dev. 1994;8:2401–2415. doi: 10.1101/gad.8.20.2401. [DOI] [PubMed] [Google Scholar]
  • 113.Sogo JM, Lopes M, Foiani M. Fork reversal and ssDNA accumulation at stalled replication forks owing to checkpoint defects. Science. 2002;297:599–602. doi: 10.1126/science.1074023. [DOI] [PubMed] [Google Scholar]
  • 114.Nedelcheva MN, Roguev A, Dolapchiev LB, Shevchenko A, Taskov HB, Shevchenko A, Stewart AF, Stoynov SS. Uncoupling of unwinding from DNA synthesis implies regulation of MCM helicase by Tof1/Mrc1/Csm3 checkpoint complex. J Mol Biol. 2005;347:509–521. doi: 10.1016/j.jmb.2005.01.041. [DOI] [PubMed] [Google Scholar]
  • 115.Nitani N, Nakamura K, Nakagawa C, Masukata H, Nakagawa T. Regulation of DNA replication machinery by Mrc1 in fission yeast. Genetics. 2006;174:155–165. doi: 10.1534/genetics.106.060053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Calzada A, Hodgson B, Kanemaki M, Bueno A, Labib K. Molecular anatomy and regulation of a stable replisome at a paused eukaryotic DNA replication fork. Genes Dev. 2005;19:1905–1919. doi: 10.1101/gad.337205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Arias EE, Walter JC. Strength in numbers: preventing rereplication via multiple mechanisms in eukaryotic cells. Genes Dev. 2007;21:497–518. doi: 10.1101/gad.1508907. [DOI] [PubMed] [Google Scholar]
  • 118.Labib K, Tercero JA, Diffley JF. Uninterrupted MCM2-7 function required for DNA replication fork progression. Science. 2000;288:1643–1647. doi: 10.1126/science.288.5471.1643. [DOI] [PubMed] [Google Scholar]
  • 119.Sandler SJ, Marians KJ. Role of PriA in replication fork reactivation in Escherichia coli. J Bacteriol. 2000;182:9–13. doi: 10.1128/jb.182.1.9-13.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Raghuraman MK, Winzeler EA, Collingwood D, Hunt S, Wodicka L, Conway A, Lockhart DJ, Davis RW, Brewer BJ, Fangman WL. Replication dynamics of the yeast genome. Science. 2001;294:115–121. doi: 10.1126/science.294.5540.115. [DOI] [PubMed] [Google Scholar]
  • 121.Mitrophanov AY, Groisman EA. Positive feedback in cellular control systems. Bioessays. 2008;30:542–555. doi: 10.1002/bies.20769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Skotheim JM, Talia SDi, Siggia ED, Cross FR. Positive feedback of G1 cyclins ensures coherent cell cycle entry. Nature. 2008;454:291–296. doi: 10.1038/nature07118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Holt LJ, Krutchinsky AN, Morgan DO. Positive feedback sharpens the anaphase switch. Nature. 2008;454:353–357. doi: 10.1038/nature07050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA, Celeste A, Manis JP, van Deursen J, Nussenzweig A, Paull TT, Alt FW, Chen J. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol Cell. 2006;21:187–200. doi: 10.1016/j.molcel.2005.11.025. [DOI] [PubMed] [Google Scholar]
  • 125.Soutoglou E, Misteli T. Activation of the cellular DNA damage response in the absence of DNA lesions. Science. 2008;320:1507–1510. doi: 10.1126/science.1159051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.O'Neill BM, Szyjka SJ, Lis ET, Bailey AO, Yates JR, 3rd, Aparicio OM, Romesberg FE. Pph3-Psy2 is a phosphatase complex required for Rad53 dephosphorylation and replication fork restart during recovery from DNA damage. Proc Natl Acad Sci U S A. 2007;104:9290–9295. doi: 10.1073/pnas.0703252104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Keogh MC, Kim JA, Downey M, Fillingham J, Chowdhury D, Harrison JC, Onishi M, Datta N, Galicia S, Emili A, Lieberman J, Shen X, Buratowski S, Haber JE, Durocher D, Greenblatt JF, Krogan NJ. A phosphatase complex that dephosphorylates gammaH2AX regulates DNA damage checkpoint recovery. Nature. 2006;439:497–501. doi: 10.1038/nature04384. [DOI] [PubMed] [Google Scholar]
  • 128.Leroy C, Lee SE, Vaze MB, Ochsenbien F, Guerois R, Haber JE, Marsolier-Kergoat MC. PP2C phosphatases Ptc2 and Ptc3 are required for DNA checkpoint inactivation after a double-strand break. Mol Cell. 2003;11:827–835. doi: 10.1016/s1097-2765(03)00058-3. [DOI] [PubMed] [Google Scholar]
  • 129.Pellicioli A, Lee SE, Lucca C, Foiani M, Haber JE. Regulation of Saccharomyces Rad53 checkpoint kinase during adaptation from DNA damage-induced G2/M arrest. Mol Cell. 2001;7:293–300. doi: 10.1016/s1097-2765(01)00177-0. [DOI] [PubMed] [Google Scholar]
  • 130.Usui T, Foster SS, Petrini JH. Maintenance of the DNA-damage checkpoint requires DNA-damage-induced mediator protein oligomerization. Mol Cell. 2009;33:147–159. doi: 10.1016/j.molcel.2008.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Usui T, Ogawa H, Petrini JH. A DNA damage response pathway controlled by Tel1 and the Mre11 complex. Mol Cell. 2001;7:1255–1266. doi: 10.1016/s1097-2765(01)00270-2. [DOI] [PubMed] [Google Scholar]
  • 132.Flott S, Alabert C, Toh GW, Toth R, Sugawara N, Campbell DG, Haber JE, Pasero P, Rouse J. Phosphorylation of Slx4 by Mec1 and Tel1 regulates the single-strand annealing mode of DNA repair in budding yeast. Mol Cell Biol. 2007;27:6433–6445. doi: 10.1128/MCB.00135-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Morin I, Ngo HP, Greenall A, Zubko MK, Morrice N, Lydall D. Checkpoint-dependent phosphorylation of Exo1 modulates the DNA damage response. Embo J. 2008;27:2400–2410. doi: 10.1038/emboj.2008.171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Liberi G, Chiolo I, Pellicioli A, Lopes M, Plevani P, Muzi-Falconi M, Foiani M. Srs2 DNA helicase is involved in checkpoint response and its regulation requires a functional Mec1-dependent pathway and Cdk1 activity. Embo J. 2000;19:5027–5038. doi: 10.1093/emboj/19.18.5027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Baroni E, Viscardi V, Cartagena-Lirola H, Lucchini G, Longhese MP. The functions of budding yeast Sae2 in the DNA damage response require Mec1- and Tel1-dependent phosphorylation. Mol Cell Biol. 2004;24:4151–4165. doi: 10.1128/MCB.24.10.4151-4165.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Tseng SF, Lin JJ, Teng SC. The telomerase-recruitment domain of the telomere binding protein Cdc13 is regulated by Mec1p/Tel1p-dependent phosphorylation. Nucleic Acids Res. 2006;34:6327–6336. doi: 10.1093/nar/gkl786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Olson E, Nievera CJ, Klimovich V, Fanning E, Wu X. RPA2 is a direct downstream target for ATR to regulate the S-phase checkpoint. J Biol Chem. 2006;281:39517–39533. doi: 10.1074/jbc.M605121200. [DOI] [PubMed] [Google Scholar]

RESOURCES