Open access peer-reviewed chapter

Checkpoint Control of DNA Repair in Yeast

Written By

Michael Fasullo

Submitted: 01 October 2020 Reviewed: 02 March 2021 Published: 29 March 2021

DOI: 10.5772/intechopen.96966

From the Edited Volume

Saccharomyces

Edited by Thalita Peixoto Basso and Luiz Carlos Basso

Chapter metrics overview

449 Chapter Downloads

View Full Metrics

Abstract

Budding yeast has been a model organism for understanding how DNA damage is repaired and how cells minimize genetic instability caused by arresting or delaying the cell cycle at well-defined checkpoints. However, many DNA damage insults are tolerated by mechanisms that can both be error-prone and error-free. The mechanisms that tolerate DNA damage and promote cell division are less well-understood. This review summarizes current information known about the checkpoint response to agents that elicit both the G2/M checkpoint and the intra-S phase checkpoint and how cells adapt to unrepaired DNA damage. Tolerance to particular bulky DNA adducts and radiomimetic agents are discussed, as well as possible mechanisms that may control phosphatases that deactivate phosphorylated proteins.

Keywords

  • DNA damage tolerance
  • checkpoint
  • budding yeast
  • phosphatase

1. Introduction

DNA repair involves the recognition and excision of DNA damage followed by template-directed DNA synthesis using an undamaged strand (for reviews see [1]). Major repair mechanisms include base excision repair (BER, [2]), nucleotide excision repair (NER, [3]), and double-strand break (DSB) repair [4]; In budding yeast, homologous recombination (HR) is the preferred pathway for repair of DSBs. When the DNA replication apparatus bypasses DNA lesions on either the leading or lagging strand, single-strand gaps are created, and the resulting repair of gapped sister chromatids is referred to as postreplication repair (PRR); postreplication repair pathway involves DNA synthesis by low fidelity polymerases or template switch mechanisms. These studies have determined the identity of multiple components of these DNA repair pathways and demonstrated a remarkable conservation between both “simple” eukaryotes, such as budding yeast, and higher eukaryotes, including humans. Yeast studies, therefore, have a direct impact on understanding the molecular basis of inheritable DNA repair deficiencies in humans, many of which are associated with cancer (for review, see [5]).

One unifying theme of DNA repair is the redundancy of DNA repair mechanisms for specific DNA lesions. In budding yeast, the pathway choice may depend on the context of the DNA lesion, the stage of the cell cycle, and the ploidy of the strain. For example, DSBs can be repaired by non-homologous end joining (NHEJ), single-strand annealing (SSA), and HR [4]). HR is enhanced [6, 7] while NHEJ is suppressed in MATa/MATα diploid strains [8] where the HR repair can use an undamaged homolog as a repair template. Secondly, specific DNA base lesions, such as abasic sites can be repaired by both bases excision repair (BER) or nucleotide excision repair (NER, [2, 9, 10]). Thirdly, postreplication repair (PPR) pathways can be both error-prone and error-free [11]. While some DNA lesions are repaired by redundant mechanisms, others, such as inter-strand DNA cross-links that impede DNA polymerases, require components of multiple pathways, including NER, postreplication repair, and HR [12].

Cells adapt to unrepaired DNA lesions and rely on DNA damage tolerance mechanisms to maintain viability. For example, exposure to 150 J/m2 UV, generates 3 x 104 cyclobutane pyrimidine dimers (CPD) per yeast cell, or approximately one UV-induced dimer per 400 bp of yeast DNA [13]. The pyrimidine-pyrimidone (6–4) photoproduct is also abundant but eightfold less present after exposure to UVB (280–320) [14]. The efficiency of CPD repair depends on the surrounding chromatin and whether the damaged strand is transcribed (for review, [15]). Since not all UV-induced damage is repaired within the period of a single cell cycle, cell viability depends on DNA damage tolerance and adaptation.

The purpose of this review is to summarize mechanisms by which checkpoint activation and DNA damage tolerance confer resistance for particular DNA lesions, and to summarize more recent data concerning complex carcinogen-associated lesions. The importance of this topic is underscored by observations that tolerance of DNA damage may reduce the efficacy of chemotherapeutic drugs, such as cisplatin, while increasing genetic instability. We present studies that suggest that DNA damage tolerance can be influenced by multiple factors, including the nutritional status of the cell and signaling from both the Target of Rapamycin (TOR) and the protein kinase A (PKA) pathways.

The orchestration of DNA repair pathways is especially critical when the replication fork progression is blocked by bulky DNA adducts [16]. Stalled replication forks can generate DNA secondary structures that trigger genome instability. While particular mechanisms to bypass such adducts, such as template switching and translesion DNA synthesis, may be error free, there is a risk that toxic recombination intermediates can either impede DNA replication progression or lead to replication fork collapse [17]. Replication-associated DNA breaks, in turn, may re-initiate replication on an undamaged chromosome or chromatid, referred to as break-induced replication (BIR, [18]). Chromosome breaks, if unrepaired, can be aberrantly rejoined forming dicentric chromosomes, leading to further breakage and instability, often referred to as the breakage-fusion-bridge cycle, a phenomena suggested to account for gross genome rearrangements in yeast [19] and in cancer cells [20, 21].

To suppress genetic instability and facilitate DNA repair, cell cycle checkpoints trigger arrest at defined stages in the cell cycle to ensure that DNA damage is repaired before the damage is replicated or inherited in the next cell cycle (for review, see [22]). These checkpoints are referred to as the G1-S checkpoint, the intra S checkpoint, and the G2/M checkpoint. In brief, phosphoinositide three-kinase-related kinase (PI3K)-like kinases, referred to as apical or sensor kinases, initiate signaling after recruitment to DNA damage or stalled replication forks, checkpoint kinases, referred to as effector or downstream kinases, then amplify and transmit the checkpoint signal, and effectors that catalyze covalent protein modifications [23], resulting in activation or degradation of cellular targets. Besides ensuring that the cell cycle is delayed so that adequate time is available for DNA repair [24], effectors also modify DNA repair proteins [25, 26]), upregulate the synthesis of deoxynucleotides (dNTPs, [27]), regulate transport of tRNA from the nucleus to and from the cytoplasm, trigger autophagy [28], regulate histone levels [29], and cross-talk with other stress-induced pathways to ensure survival [30]. The totality of the response is generally referred to as the DNA damage response (DDR) [22]. Checkpoint recovery occurs once the DNA damage is repaired or the replication block has been circumvented [31]; however, adaptation to persistent DNA lesions, such as DSBs, can also occur. In either case, the phosphorylated and activated checkpoint proteins are either dephosphorylated or degraded and subsequently rendered inactive. Depending on the DNA damage and time period of exposure, peak activation occurs within 2–4 hours after acute DNA damage exposure [32], with simultaneous upregulation of dNTP levels and DNA damage-inducible genes. Repression of late origins of replication, inhibition of replication, and prevention of anaphase can last for additional hours [33], while adaptation can be observed after 12 hrs. The presence of DNA repair foci, such as Rad51, can thus last hours after the initiation of the DNA damage insult. Thus, aspects of the DNA damage response can persist hours after the initial genomic insult and after DNA repair is completed.

Advertisement

2. Checkpoint activation initiated by DSBs

Checkpoint activation triggered by DSBs, and subsequent recovery or adaptation has been extensively studied in strains containing either uncapped telomeres or chromosomal DSBs that cannot be repaired by HR. An unrepaired DSB occurs when HO endonuclease cleaves the recognition sequence at the MAT locus but silent mating type locus has been deleted [34]. Uncapped telomeres occur when either the Cdc13-Stn1-Ten1 (CST) complex or the Ku complex, composed of yKu70 and yKu80, is defective. At restrictive (elevated) temperatures in either cdc13 or yku70 mutants, extensive tracts of single-stranded DNA complex are generated [35]. While two DNA ends are revealed by a single unrepaired DSB at the MAT locus, at the non-permissive temperature in cdc13 mutants single-stranded DNA is revealed at the telomeres of sixteen chromosomes, thus amplifying the DNA damage signal.

A single DSB occurring in G1 does not trigger cell cycle arrest at the G1/S or intra S phase checkpoint [36], but instead the cell cycle progresses through S phase and into G2 phase, where cells arrest. Repair of DSBs can occur at any time in the cell cycle by NHEJ; however, in budding yeast, NHEJ is favorable when the single-strand overhangs are short [37]. However, DSBs, will trigger a partial DNA damage response in G1 cells [38], and recombination proteins, such as Rad51 and Rad54, are still induced [39] and Rad55 is phosphorylated [38] .

The orchestration of checkpoint signaling has been well described in current reviews [40] and is briefly summarized (Figure 1). Mre11/Rad50/Xrs2 (MRX) and Tel1 (ataxia telangiectasia mutated (ATM) ortholog) bind to the ends of the DSB, which facilitates the juxtaposition of the ends of the breaks. NHEJ requires yKu70 and yKu80. However, if NHEJ is not successful, cyclin dependent kinase I (Cdk1 or Cdc28), which has high activity in G2, phosphorylates Sae2 and the 5′ to 3′ exonuclease Dna2 [41]. Sae2 phosphorylation activates the Mre11 endonuclease activity that ejects yKu70 from the ends of the DSB. Together with Sgs1/Dna2 and Exo1, the ends are further degraded in a 5′ to 3′ direction. NHEJ and resection require chromatin remodeling factors, including the Ino80 complex [42, 43], Rsc complex [44], and Fun30 [45, 46]. Resection is generally slow and proceeds at 1–2 nucleotide per minute [40]. Resection of the ends reveals single stranded DNA (ssDNA), which is then coated by single strand binding protein (RPA), which serves as a general sensor for DNA damage. The RPA-coated ssDNA is a binding site for Ddc2-Mec1 (ataxia telangiectasia mutated and rad3-related (ATR ortholog). Rad24/Rfc facilitates the binding of trimeric Rad17/Mec3/Ddc1 (9–1-1) protein which recognizes the junction between the single stranded DNA and the double-stranded DNA [47]. Thus, ssDNA serves as a general signal for checkpoint signaling [48].

Figure 1.

A pathway for checkpoint pathway commences with a DNA damage signal that triggers the PIKK kinase, Mec1. Downstream checkpoint kinases are activated, as facilitated by the adaptors Rad9 and Mrc1. Kinase substrates are identified for Rad53 and Dun1, but both Mec1 and Rad53 phosphorylate multiple substrates that are not shown.

Mec1, a sensor or apical serine/threonine kinase phosphorylates downstream kinases, DNA repair proteins, and histones, preferably at SQ/TQ sites [49]. Both Mec1 and Tel1 phosphorylate histone γ-H2A for ~50 kb on either side of the DSB, which serves to recruit other checkpoint protein, such as the adaptor, Rad9 (53BP1 ortholog). Mec1 regulates checkpoint signaling by autophosphorylation on the S1964 residue [50] and phosphorylation of Ddc2, which destabilizes unbound Ddc2 and limits the amount of bound Ddc2-Mec1. Mec1 also phosphorylates Exo1 [51], which limits the amount of single-stranded DNA that could serve as a signal for checkpoint activation. Thus, Mec1’s activity serves to not only activate downstream kinases but also dampen the checkpoint response.

Rad9, as an adaptor protein and 53BP1 ortholog, is required to bring the effector (transducer) kinases in contact with Mec1. Rad9 binding to chromatin is mediated by its BRCT and tudor domains that interact with phosphorylated and trimethylated histone H3, respectively [52]. While histone phosphorylation is induced by DNA damage, Dot1-mediated histone H3 methylation is constitutive [53]. Localization to damaged DNA is facilitated by binding to Rtt107/Dbp11. Both Mec1 and Cdk1 phosphorylate Rad9 on separate domains [40]. In turn, oligomers of phosphorylated Rad9 bind to Rad53 and facilitate Mec1-mediated phosphorylation. A Rad53 phosphorylated heterodimer then autophosphorylates; the hyper-phosphorylated Rad53 can, in turn, rapidly diffuse throughout the nucleus and phosphorylate multiple substrates, including Dun1 and Asf1. Similarly, Rad9 facilitates Mec1-mediated Chk1 phosphorylation; the activated Chk1 phosphorylates Pds1, which prevents its degradation by the anaphase promoting complex (APC). In turn, sister chromatid cohesion is maintained and anaphase is prevented [54]. Activated Rad53 also inhibits the APC from degrading securin [54]. A Rad53-mediated pathway inhibits Cdc5, a polo-like kinase that functions in mitotic exit through the regulation of spindle pole body separation [30]. While Pds1 phosphorylation can also be triggered by the Mad2-mediated spindle checkpoints, Rad53 phosphorylation is only triggered by DNA damage or stalled replication forks. While the single mutants are partially defective in DNA damage-induced G2 arrest, rad53 pds1 and rad53 dun1 double mutants are fully deficient [55].

Partial to full checkpoint activation will also occur when DNA damage processing is rendered less efficient. For example, mating-type switching in a rad1 mutant, defective in removal of 3′ non-homologous ends, will trigger a checkpoint-dependent cell cycle delay [56]. Interestingly, Rad53 phosphorylation was not abundant in the rad1 mutant; however, the cell cycle delay was not observed in the rad9 strain, and was shortened in the mad3 mutants, defective in spindle checkpoint. The authors speculated that checkpoint activation occurred when H2A phosphorylation extended through centromeric chromatin, triggering a spindle pole checkpoint response [56]. These studies indicate that spindle pole checkpoints also participate in the DNA damage response, depending on the context of the DSB.

While the G2 checkpoint is critical for HR repair of DSBs, Mec1 and Mec1-signaling pathway also phosphorylate additional DNA repair functions that facilitate DSB repair and damage incurred by radiomimetic agents [40]. For example, Mec1 phosphorylates Rad51 [57] and Rad55 [58]. Phosphorylation of Rad51 enhances its activity and is required for resistance to recombinagens, such as methyl methane sulfonate (MMS) [57]. In addition, Mec1 phosphorylates Slx4, which binds to Rad1/Rad10 and facilitates single-strand annealing by cleaving non-homologous tails [59]. The studies indicate that the checkpoint pathway directly phosphorylates repair proteins to enhance their function. While there are many proteins that are phosphorylated in response to DNA damage [60, 61], the functional significance of the phosphorylation of many of these proteins has yet to be determined (Table 1).

Protein PhosphorylatedKinaseEffectReference
Checkpoint Signaling
Rad9Mec1 /Tel1Rad53 docking and Rad9 multimerization[40, 60]
Rad53Mec1Activation of Rad53 autophosphorylation[40, 60]
Rad53Tel1Activation of Rad53[40, 60]
Chk1Mec1Phosphorylation of Pds1[30, 54]
Pds1Chk1APC-associated degradation of Pds1 is inhibited[54, 60]
Mec1-Ddc2Mec1Attenuation of Mec1 kinase activity[50]
Dun1Rad53Activation of Dun1 kinase activity[40]
Nucleases
Sae2Cdk1Cell cycle regulation limiting resection to G2/M[41]
Dna2Cdk1Cell cycle regulation limiting resection to G2/M[62]
Exo1Mec1Inhibition of Exo1 5′-3′ exonuclease activity[51]
Transcription inhibitors
Crt1Dun1Crt1 phosphorylation leads to degradation, and subsequent Rnr transcriptional activation[60, 63]
Protein Inhibitors
Sml1Dun1Sml1 phosphorylation leads to degradation and release from Rnr1 subunit and subsequent increase in dNTPs[27]
Dif1Dun1Allows for transport of RNR into the cytoplasm[64]
Recombination Proteins
Rad55Rad53Enhances recombination in rad5 mutants[58]
Rad51Mec1/Rad53Enhances activity[57]
Rev1Mec1Facilitates binding to ssDNA[26]
Resolvases
Yen1Cdk1Inhibits function in S phase by transportation to the cytoplasm[62]
Mus81/Mms4Cdk1, Dbf4 Cdc7Regulation of cleavage of Holliday and branch junctions[62, 65]
Helicases
Pif1Rad53Inhibits fork unwinding, promotes DNA damage tolerance by HR[66]
Rrm3Rad53Inhibits fork unwinding[66]
Srs2Cdk1Promotes adaptation by removal of Rad51 filaments[62, 67]

Table 1.

Proteins phosphorylated by DDR.

Advertisement

3. Checkpoint recovery and adaptation from double-strand break

Once cells have repaired the DSB, recovery involves reversal of protein modifications and chromatin restoration. While the DNA damage may no longer be present, protein modifications are still present that signal checkpoint activation. To inactivate the G2/M checkpoint and resume division, Rad53 must be dephosphorylated. Two phosphatases involved in the inactivation of Rad53 include phosphorylated versions of the type 2C protein phosphatases (PP2C), Ptc2 and Ptc3 [68, 69, 70]; these phosphatases are also involved in inactivating other stress induced pathways, such as the Hog1-mediated osmotic stress induced pathway [71], while Ptc2 dephosphorylates Cdk1. Casein kinase II (Ck2) phosphorylates Ptc2, which specifically binds to the Rad53 FHA1 domains [72]. Interestingly, CK2 mutants are more defective in adaptation than ptc2 mutants, suggesting that CK2 may control additional genes involved in adaptation [68].

Pph3, a member of the PP4 family, is important in maintaining full recovery; the triple mutant (ptc2, ptc3, pph3) is severely defective in DSB repair when the repair pathway is slow [70]. This may be partially explained by observations that Pph3 functions to dephosphorylate γ-H2A, which serves as a signal for activation of checkpoint proteins, cohesins, and chromatin remodelers [73]. However, the mechanism by with chromatin associated gamma γ-H2A is fully dephosphorylated is still being explored.

Chromatin restoration requires Asf1 and Caf1 which reassemble chromatin on DNA (Kim and Haber [74]). Asf1 binds histone H3 triggering acetylation by the histone acetyltransferase, Rtt109, and further ubiquitylation by Rtt101 [75]. This, in turn promotes the binding of the histone H3 and H4 heterodimer by Caf1. Interestingly, Asf1 also functions to bind Rad53, thus serving a role to sequester dephosphorylated Rad53. Thus Asf1 functions both in reassembling chromatin and stabilizing dephosphorylated Rad53 [75].

If a DSB is not repaired, cells will either resume the cell cycle or die. The resumption of the cell cycle is referred to as adaptation. Similar to recovery, adaptation involves both chromatin remodeling and phosphatases that deactivate the Rad53 kinase and Cdk1 kinase. This adaptation is blocked in yku80 mutants [69], deficient in NHEJ, and cdc5-ad, which is defective in mitotic exit. Yku80 mutants exhibit twice the rate of resection of the DSBs, resulting in more single-stranded DNA and thus the potential for more Rad53 checkpoint signaling. This interpretation is supported by observations that overexpression of Ptc2 is sufficient to suppress the adaptation defect of both yku70 and yku80 [69]. However, the role of resection in checkpoint adaptation is complicated by the identification of chromatin remodelers, such as Fun30 [76, 77, 78], which are required for adaptation but enhance resection. One possibility is that Fun30-associated resection in γ -H2A-modified chromatin antagonizes the checkpoint protein Rad9 from binding and signaling downstream checkpoint effectors [78].

Additional genes function to remove recombination proteins from chromatin. Removal of Rad51 filaments is facilitated by the chromatin modifier Tid1 (Rdh54) [79] and the Srs2 helicase, the former is phosphorylated by Mec1 and the latter is phosphorylated by Cdk1 [67]. Both rdh54 and srs2 mutants are defective in adaptation (reference [79, 80]). These studies present additional evidence MEC1 functions both in the triggering of checkpoint arrest as well as recovery from checkpoint arrest.

Advertisement

4. Uncapped telomeres, checkpoint activation, and adaptation

While single-stranded DNA is present at telomeres, it is normally “capped” by a RPA-like structure, referred to as Cdc13-Stn1-Ten1 (CST), and by Ku (yKu70/yKu80) complex [35]. During replication, Cdc13 is phosphorylated by Cdk1 and recruits telomerase [81]. Telomere ends are susceptible to nucleases in yeast mutants defective in proteins that bind to chromosome ends, such as yku70, and that are defective in recruiting telomerase, such as the cdc13–1 mutant at the restrictive temperature. Pif1 helicase inhibits telomerase and leads to slow resection at the telomere end [82, 83]. Resection is also slowed by binding of Rif1 and Rap1, which bind specifically to single stranded telomere sequences and inhibit the binding of the checkpoint activators, RPA and Rad24 [84, 85]. In the cdc13–1 mutant, resection is extensive and largely performed by Exo1, leading to ssDNA bound to RPA, the 9–1-1 complex, and Rad9. Similarly, in yku70 mutants, ssDNA is generated, but it takes several generations for ssDNA to accumulate [35]. The 9–1-1 complex is apparently not involved in eliciting a checkpoint response but Chk1 activation is required for Exo1-mediated resection [86]. In yku70 mutants, resected telomeres elicit both a spindle and DNA damage checkpoint activation. However, unlike HO-induced DSBs, Mec1 binding does not lead to rapid resection but rather an inhibition of resection through subsequent binding of Rad9 and Rad53 [87]). Resection of the telomere, in turn, may facilitate recombination or break-induced replication (BIR, [88]) using an undamaged chromosomal end as a template for replication to the end of the chromosome. BIR is facilitated by activated Pif1 [89]. Thus, checkpoint activation at uncapped telomeres enables alternative mechanisms of telomere lengthening.

Adaptation to shortened telomeres was first noted by Sandell and Zakian [90] and require CKII and Cdc5 [91]. CKII directly phosphorylates Ptc2, which is required for tolerating shortened telomeres [92]. In addition, phosphorylated Cdc13 can be dephosphorylated by Pph3/Psy3, resulting in the segregation of uncapped chromosomal ends [35]. Over-expression of Cdc5 also decreases Rad53 phosphorylation [93]. Thus, as in HO-induced DSBs, there are multiple phosphatases and kinases that modulate adaptation.

Advertisement

5. Intra-S phase checkpoint and stabilization of the replication fork

The purpose of the intra-S phase checkpoint is to maintain replication fork integrity so that replication can be completed; collapsed replication forks are a major source of genetic instability [94]. Replication forks stall because of limiting amounts of dNTPs or when DNA damage, resulting from a bulky adduct or cross-links, block progression of a high fidelity polymerase (for review, see [95]). The precise number of stalled forks to trigger the intra S-phase checkpoint is unknown [96]. Uncoupling of the helicase and DNA polymerase activity generates single strand gaps on both the leading and lagging strands. Checkpoint responses serve to maintain the stability of the replication fork in part by blocking the formation of toxic secondary DNA structures and replication fork reversal, degrading excessive histones, and inhibiting the firing of late replication forks so that replication can be resumed if stalled replication forks become permanently arrested [97]. In S phase checkpoint mutants, such as rad53, regressed replication forks, referred to as “chicken feet” structures, can be visualized [98]. Severe deficiencies can lead to mitotic catastrophe and subsequent lethality.

The extensive tracts of single-stranded DNA generated at stalled forks signal a checkpoint response. Similar, to checkpoint signaling at DSBs, the 9–1-1 complex is loaded and facilitates binding of Ddc2-Mec1. Rad18, which monoubiquitinates trimeric PCNA at K164 position at stalled replication forks, also monoubiquitinates the 9–1-1 complex leading to enhanced recruitment of Ddc2-Mec1 [99]. Checkpoint activation at stalled forks can also be facilitated by Elg1, which removes PCNA from stalled forks [100]. The Mrc1 (claspin) functions as an adaptor, analogous to Rad9, in the phosphorylation of Rad53; however, unlike Rad9, Mrc1 is associated with the replication forks [101]. Full checkpoint activation requires BLM homolog Sgs1 [102]. Rad53 phosphorylation in turn serves to promote histone degradation, inhibit late origin firing, and increase the levels of dNTPs. The inhibition of late origin firing maintains RPA and allows replication restart from other replication origins.

Deoxynucleotide levels (dNTPs) increase 7–8 fold after DNA damage by upregulating the activity of ribonucleotide reductase (Rnr) activity [103]. Upregulation of Rnr activity is achieved at the transcriptional, translational, and the posttranslational levels. At the transcriptional level, phosphorylated Rad53 activates Dun1 kinase, which deactivates Crt1 transcriptional repressor [63]. At the translational level, TRM9, which functions to methylate the uridine wobble base of tRNA-Arg (UCU) and tRNA-Glu (UUC), facilitates the translation of the Rnr1 transcript [104]; however, how checkpoint signaling enhances TRM9 function is unclear. At the post-translational level, Dun1 inactivates the Sml1 protein inhibitor by phosphorylation [27]. Besides increasing the transcription of Rnr subunits, the Rnr inhibitor Sml1 and its paralog Dif1 are degraded, subsequently the Rnr subunits are shuttled to the cytoplasm where they form an active enzyme complex [64]. In addition to increasing the overall level of dNTPs, the Rnr3 large subunit forms an alternative ribonucleotide reductase complex that has relaxed dATP negative feedback regulation [105]. This ensures that adequate levels of dNTPs are available during times of unscheduled DNA synthesis.

While high levels of dNTPs facilitate replication fork progression [106]; abnormally high or low levels of dNTPs can promote genetic instability. High levels of dNTPs reduce the fidelity of polymerase epsilon [107]. Low levels of dNTPs correlate with hyper-recombination, as has been observed in dun1 null mutants and in mec1 hypomorphs; these phenotypes can be suppressed by higher basal levels of dNTPs conferred by a SML1 deletion [108]. These studies suggest that there is a range of dNTP concentrations that correlate with replication fork stability; however, the mechanisms by which higher dNTP levels decrease replication fork collapse are unclear.

In contrast to DSBs, where HR processes are facilitated, there are redundant mechanisms to prevent recombination at stalled forks; these mechanisms include disassembly of Rad51 filaments, helicases that abort recombination intermediates, and nucleases that degrade aberrant structures. PCNA is SUMOylated (SUMO-PCNA), facilitating the binding of the helicase Srs2, which suppresses recombination by disassembling Rad51 filaments (for review, see [109]. Recombination intermediates are aborted by the BLM ortholog Sgs1 helicase; indeed, the sgs1 srs2 double mutants is not viable but viability is rescued when HR is defective [110]. Additional helicases, including Pif1 and Rrm3 can unwind and reverse “chicken feet” structures and reversed forks [66]. Nucleases, such as Exo1, function to degrade reversed forks, although excessive Exo1 activity can lead to replication fork collapse [62]. Finally, enzymes which cleave aberrant secondary structures, such as Yen1 and Mus4/Mus81, are inhibited or rendered less active by phosphorylation and Sumoylation (SUMO) [65]). Thus, there are multiple mechanisms that prevent aberrant structures from accumulating at replication blocks (Figure 2).

Figure 2.

The pathways of tolerating DNA damage at a stalled fork are shown by the arrows. The 5′ to 3′ polarity of the DNA is designated by an arrow. The Rrm3 and Pif1 helicases inhibit replication fork reversal, while the Pif1 helicase promotes template switching. Phosphorylated Rev1 binds to single-stranded DNA and promotes replication bypass.

Generally, replication blocks that impede DNA polymerases can be bypassed by two mechanisms: 1) lesion bypass inserts a base opposite the replication block using error-prone or error-free translesion polymerases, and 2) template switch mechanisms [111] utilize recombination so that DNA polymerase bypasses DNA lesions on an undamaged template. Factors recruited to stalled forks would initially suggest that checkpoint signaling might favor lesion bypass by translesion polymerases. For example, Rad5 binding to stalled forks facilitates the recruitment of Rev1, even in the absence of DNA damage [112]. In addition, Rev1 is also phosphorylated by Mec1, which increases its affinity to ssDNA [26]. In vitro replication experiments have demonstrated that other error-prone polymerases can “jump start” replication, even without association of the replicative helicase [113]. However, there is no evidence that high levels of dNTPs would facilitate translesion synthesis mediated by error-prone polymerases, such as polζ [114], and the mechanism for jump start is unclear.

Advertisement

6. Bypass of single-strand gaps and replication blocks by template switch mechanisms

Template switch mechanisms also allow polymerases to bypass replication forks and resume DNA synthesis; these mechanisms are generally thought to occur on both leading and lagging strands. Template switching is orchestrated by proteins that modify the DNA polymerase processivity factor, PCNA. When the high fidelity polymerase stalls at the replication block, Rad18/Rad6 monoubiquitinates PCNA at the K164 position; monoubiquitinated PCNA can facilitate polymerase switching from to a translesion polymerase of lower fidelity and processivity. PCNA may further become polyubiquitinated at position K164 by combined action of Ubc13/Mms2/Rad5 (for review, see [115]). Rad5 also contains a helicase function that catalyzes replication fork reversal and is required for template switch mechanisms on the lagging strand [116]. While Rad5 does not require DNA damage at the stalled replication fork for recruitment [112], Rad5 over-expression can trigger genome instability [117]. The checkpoint signaling cascade, mediated by the Dun1 kinase, regulates Rad5 at the post-transcriptional level by destabilizing Rad5 mRNA [118]. These studies indicate that RAD5 function is regulated.

However, checkpoint signaling may also facilitate template switch mechanisms. Rad53 is required for DNA damage-associated unequal SCE after exposure to MMS [119] and Rad53-mediated Rad55 phosphorylation confers enhanced MMS resistance when RAD5 is also defective [25]. The Rad9 checkpoint protein binds to persistent single strand gaps on the lagging strand, inhibiting the RecQ-like Sgs1 anti-recombination function. In addition Pif1, which is phosphorylated by Rad53, functions in template switching [120]. With longer term checkpoint-mediated G2 arrest, however, Rev1 protein levels accumulate [121, 122], suggesting that error-prone polymerases may serve as the ultimate backup in postreplication repair after error-free mechanisms have failed.

Advertisement

7. Choice of DNA damage tolerance pathway is influenced by the DNA lesion

Multiple tolerance pathways can confer resistance to particular types of DNA damage and the pathway preference depends on the DNA damaging agent. For example, MMS exposure generates by 7Me-Guanine and 3Me-Adenine lesions; while the 7Me-Guanine is mutagenic, the 3Me-Adenine blocks replication [123]. Replication bypass can occur by error-prone or error-free polymerases, or by template switching. While all three pathways are involved in bypass of 3Me-Adenine lesions [124], template switch mechanisms are preferred [125]. Checkpoint signaling facilitates template switch mechanisms after exposure to MMS [125, 126]. These studies suggest that template switch mechanisms may be the preferred pathway for bypassing particular lesions that block DNA replication.

The preference of template switch mechanisms or translesion pathways may depend on the efficiency of bypass and repair for large bulky adduct or cross-links. Particular UV-associated DNA cross-links are efficiently bypassed using either polε [127] or a two-step mechanism involving polε and polζ [128]. However, error-free bypass of 4–6 pyrmidine-pyrimidone lesions, present on a plasmid, occurs by template switch mechanisms after their introduction in a NER deficient yeast strain [129]. Likewise, 4-NQO induces bulky damage and stimulates template switch mechanisms [126]. These studies indicate that template switch mechanisms are likely used in error-free postreplication repair pathways [130].

Advertisement

8. Attenuation of the S phase checkpoint activation

In order for the cell cycle to resume and chromatids to separate the checkpoint activation needs to be downregulated and joint molecules need to be resolved. Once replication is completed, Mrc1 functions as an adaptor for Mec1-mediated checkpoint signaling is diminished since there are no more replication forks [33]. Resumption of the cell cycle is accomplished by dephosphorylating Rad53 [131]. However, single-strand gaps on sister chromatids can still function to trigger Rad9-mediated checkpoint signaling. To dampen Rad9’s adaptor function in mediating Mec1 catalyzed Rad53 phosphorylation, competitive scaffolds compete with Rad9 binding to chromatin [132]. For example, the Mec1-mediated phosphorylation of Slx4 enables an association with Rtt107/Dpb11, which provides a competitive scaffold for the interaction of Rad9 with Dpb11 [133]. These mechanisms thus prevent Rad53 hyperphosphorylation.

Cleavage of joint DNA strands, or Holliday structures, is timed just before anaphase so that cleavage does not occur during S phase. Both kinases and phosphatases fine tune the timing of joint molecule cleavage. Cdk phosphorylates structure-specific nucleases Slx1/Slx4 and Mus81/Mus4 in late G2 and M phases respectively [134]. Whereas Mec1 phosphorylates and subsequently inactivates Yen1, Cdc14 dephosphorylates the inactivated form in mitosis, ensuring that joint molecules do not hinder sister chromatid division [134].

Similar to adaptions to DSBs, phosphatases deactivate Rad53 (Figure 3). These phosphatases include Pph3/Psy2 complex and Ptc1, 2. Interestingly, Pph3 directly interacts with Mec1/Ddc2 [135] at the replication fork, although the interaction does not rely on DNA damage [135]. Besides Rad53, other Mec1 substrates are likely dephosphorylated by Pph3, including phosphorylated Mec1. Thus Pph3 could potentially upregulate Mec1. However, the full range of Pph3 substrates is unknown [135].

Figure 3.

Double-strand breaks, collapsed replication forks, and replication stress lead to checkpoint activation. Activated Rad53 is dephosphorylated by a series of phosphatases, depending on the signal induced by the DNA damaging agent, shown below the DNA damage. The 5′ to 3′ polarity of the DNA is designated by an arrow. The substrates of the phosphatases Pph3/Psy2 and Ptc2/Ptc3 include Mec1, Rad53, and Exo1.

Mutations in different phosphatases may confer sensitivities to different DNA damaging agents (Table 2). For example, pph3 and psy2 mutants are hypersensitive to phelomycin but not 4NQO, while ptc2, ptc3 and ptc2 ptc3 double mutants are not phleomycin sensitive and are not required for recovery from MMS-associated checkpoint delay [136, 142]. On the other hand ptc2 ptc3 double mutants are hypersensitive to 4-NQO while pph3 and psy2 mutants are not sensitive. However, for particular agents, such as cisplatin, the triple pph3 ptc2 ptc3 mutant, is synergistically more sensitive [140]. One idea is that phosphorylation of Rad53 is differentially patterned by particular DNA damaging agents, and that the phosphatases, Ptc2/Ptc3 and Pph3/Psy2 recognize different patterns [142]. This notion is supported by the identification of different MMS and 4-NQO associated Rad53 phosphorylation sites. The connection between Ptc2 and checkpoint activation is further strengthened by observations that over-expression of Ptc2 suppresses the lethality in a Rad53 dominant lethality mutant [69].

DNA damaging agent or environmental conditionDNA damagePhosphatase
Required for Resistance or Adaptation
Phenotype of
Δ ptc2 ptc3 pph3
Reference
Restrictive temperature for cdc13 mutantLong tracts of ssDNAPtc2Unknown[84]
HO EndonucleaseDouble-strand breakPtc2/Ptc3Completely deficient in adaptation[70]
4-Nitroquinoline oxideBulky adduct and oxidative damagePtc2/Ptc3, Sit4Unknown[136, 137]
PhleomycinSingle-strand and double-strand breaksPph3,
Ptc2/Ptc3
Unknown[136]
Methyl methanesulfonateMajor and minor groove alkylationsPph3/Psy2Synergistically sensitive[70]
Aflatoxin B1AFB1-N7-Gua, and AFB1
formamidopyrimidine
Pph3/Psy2Unknown[138]
Ultraviolet radiationPyrimidine dimers and Pyrimidine-pyrimidone cross linksNot requiredModerate sensitivity[70]
HydroxyureaStalled DNA replication forks, double-strand breaksPph3/Psy2, Pph2Synergistically sensitive[70, 139]
CamptothecinTopo1 cross-link with DNAPtc2, Ptc3Synergistically sensitive[70]
CisplatinDNA cross linksPph3/Psy2Synergistically sensitive[140]
BleomycinSingle and double-strand break sPph3, Rts1 (regulator of Cdc55)Unknown[73, 141]

Table 2.

Phosphatases that function in checkpoint adaptation to specific DNA damaging agents.

Tolerance to MMS-induced DNA damage includes reactivation of stalled replication forks, which depends on the level of Rad53 phosphorylation [143]. Pph3/Psy2 phosphatase is the principle phosphatase that deactivates Rad53. In the absence Pph3/Psy2 replication restart can occur; however late origins are used to complete DNA replication. Interestingly, downregulation of Rad53 phosphorylation by a HA-Rad53 or a dot1 deletion confers higher levels of MMS resistance, although at the sake of more Rev1 foci and mutagenesis [144]. These studies would suggest that MMS-induced checkpoint activation is a double-edged sword; limiting MMS-induced mutation may come at the cost of toxic recombination intermediates.

While tolerance to MMS-induced DNA damage relies on dampening the checkpoint response, UV resistance heavily relies on checkpoint activation, as illustrated by observations that the rad14 mec1 double mutant, defective in both NER and checkpoint signaling, is synergistically more UV sensitive [145]. In yeast, UV triggers the G1-S checkpoint when NER is functional, but unrepaired UV lesions trigger checkpoint responses in S and G2 cells [146]. Interestingly, chronic exposure to low dose UV does not elicit cell cycle arrest at the G2 checkpoint, suggesting that DNA replication machinery is not significantly impeded during chronic exposure [147].

P450-acitvated carcinogens may also elicit a strong DNA damage inducible effect. For example, aflatoxin B1 (AFB1), induces strong Rad53 activation in budding yeast, which generally occurs within two hours of exposure and then is gradually attenuated [148]). AFB1 exposure also upregulates the expression of DNA repair genes, including Rad51, Csm2, and Rad16 [149, 150]. Interestingly, AFB1 exposure elicits an S phase delay coinciding with the appearance of Rad51 foci [148]. This is consistent with AFB1 being a strong recombinagen but weak mutagen in yeast [151]. Interestingly, checkpoint signaling is required for stimulation of both AFB1-associated unequal sister chromatid recombination and mutation [152]. By profiling the yeast genome for AFB1 resistance using next generation sequencing, St. John et al. [138] identified both HR genes, including Rad54, Rad55, and Csm2, and those encoding error-prone polymerases. Similar to alkylated induced damage, the Csm2(Shu) complex favors an error-free template switch mechanism [153]; thus, csm2 mutants are deficient in sister chromatid recombination but exhibit higher frequencies of AFB1-associated mutations.

Genes the confer AFB1 resistance included PSY3, CKB1 and CKB2, which function in DNA damage tolerance [138]. While the genes encoding the CKII substrates, Ptc2 and Ptc3, did not appear in the screen, the identification of both CKII and Pph3 suggest that tolerance to AFB1-associated DNA damage requires both phosphorylation and dephosphorylation of multiple proteins. The identity of these proteins may further elucidate how AFB1-associated DNA damage is tolerated.

Additional phosphatases that function in DNA damage tolerance include PP2A and PP2A-like phosphatases. These phosphatases are composed of catalytic subunits, such as Pph21 and Pph22, scaffolding subunits, and regulatory subunits, such as Cdc55 and Rts1. While a direct interaction with phosphorylated Rad53 has not been demonstrated, the PP2A phosphatase suppresses the checkpoint response after HU exposure [139]. While the identity of all of the PP2A substrates is unknown, PP2A is involved in both cytokinesis and mitosis [154]. Particular regulatory subunits are required for tolerance to different DNA damaging agents. For example, Rts1 is required for DNA damage tolerance after rad51 cells are exposed to bleomycin [141] and Sit4, a PP2A-like phosphatase, is required for tolerance to 4NQO [137].

Advertisement

9. Nutrient sensing and the regulation of adaptation and the checkpoint response

One unifying theme in DNA damage tolerance to multiple types of DNA lesions is that nutrient sensing plays an important role in promoting downregulation of the checkpoint response. Deregulation of IRA1 and IRA2, which control glucose-growth signaling, prevent adaptation to uncapped telomeres in cdc13 strains [155]. Inhibition of TOR1 by rapamycin prevents adaptation and aneuploidy in rad52 diploid strains exposed to DNA damaging agents [156].

Nutrient sensing is also important in controlling the checkpoint response through type 2A protein phosphatases. In the presence of plentiful carbon and nitrogen, target of rapamycin (TORC1) activates Mec1-signaling pathway by inhibiting PP2A and PP2A-like phosphatases. PP2A activators include ceramide and S-adenosyl methionine (SAM) [139]. The effect of this signaling on the PP2C and PP4 phosphatases is unclear. Nonetheless, these studies illustrate that the DNA damage response requires an active growth signaling response [139]. Recent data also suggests that TORC1 inhibition results in lower levels of checkpoint proteins [157]. Thus, it may appear that TORC1 may be required for both checkpoint activation and for adaptation.

Advertisement

10. Concluding remarks

Adaptation to DNA damage is critical for cell survival. The simple, straightforward model is that DNA damage activates checkpoint signaling kinases and that phosphatases, which are constitutively expressed, serve to dephosphorylate and deactivate phosphorylated proteins. Once the DNA damage is repaired, the checkpoint signaling ceases and activated proteins are dephosphorylated. However, yeast can adapt to DNA damage caused by diverse damaging agents and individual phosphatases are controlled by different kinases. In addition, cells exposed to different DNA damaging agents exhibit distinct Rad53 phosphorylation patterns and rely more on particular phosphatases for checkpoint adaptation. The checkpoint pathway also autoregulates itself and dampens its signaling in coordination with Cdk1. Finally, adaptation to particular DNA damage requires TORC1 function, which senses nutrient abundance. Thus, simple models are likely complicated by the complexity of the checkpoint responses elicited by distinct DNA damaging agents.

11. Future directions and implications

Understanding DNA damage tolerance and repair will have a significant impact on elucidating the mechanisms by which DNA adducts cause mutations and genome instability. While DNA damage tolerance has been well-studied for particular types of DNA damage, such as that caused by UV and MMS, the mechanisms for DNA damage tolerance of more complex lesions is still in its infancy. The importance of studying DNA damage tolerance mechanisms for complex agents is underscored by the importance of cross-linking agents, such as cisplatin, in cancer chemotherapy. In addition, understanding how DNA damage is tolerated may have important impacts in assessing the efficacy of antifungal agents. Elucidating DNA damage tolerance mechanisms will also be important in understanding how mutations and genetic instabilities arise when cells are exposed to low doses of the DNA damaging agent. These studies should elucidate mechanisms on how cellular aging, ploidy, and cell type may affect DNA damage tolerance pathways.

References

  1. 1. Chatterjee N, Walker GC. Mechanisms of DNA damage, repair, and mutagenesis. Environ Mol Mutagen. 2017; 58(5):235-263. DOI: 10.1002/em.22087
  2. 2. Bauer NC, Corbett AH, Doetsch PW. The current state of eukaryotic DNA base damage and repair. Nucleic Acids Res. 2015; 43(21):10083-10101. DOI: 10.1093/nar/gkv1136
  3. 3. Schärer OD. Nucleotide excision repair in eukaryotes. Cold Spring Harb Perspect Biol. 2013; 5(10):a012609. DOI: 10.1101/cshperspect.a012609
  4. 4. Wright WD, Shah SS, Heyer WD. Homologous recombination and the repair of DNA double-strand breaks. J Biol Chem. 2018; 293(27):10524-10535. DOI: 10.1074/jbc.TM118.000372
  5. 5. Hoeijmakers JH. Genome maintenance mechanisms for preventing cancer. Nature. 2001;411(6835):366-374. DOI: 10.1038/35077232
  6. 6. Heude M, Fabre F. a/alpha-control of DNA repair in the yeast Saccharomyces cerevisiae: genetic and physiological aspects. Genetics. 1993; 133(3):489-498
  7. 7. Fasullo M, Dave P. Mating type regulates the radiation-associated stimulation of reciprocal translocation events in Saccharomyces cerevisiae. Mol Gen Genet. 1994; 243(1):63-70. DOI: 10.1007/BF00283877
  8. 8. Kegel A, Sjöstrand JO, Aström SU. Nej1p, a cell type-specific regulator of nonhomologous end joining in yeast. Curr Biol. 2001; 11(20):1611-1617. DOI: 10.1016/s0960-9822(01)00488-2
  9. 9. Torres-Ramos CA, Johnson RE, Prakash L, Prakash S. Evidence for the involvement of nucleotide excision repair in the removal of abasic sites in yeast. Mol Cell Biol. 2000;20(10):3522-3528. DOI: 10.1128/mcb.20.10.3522-3528.2000
  10. 10. Kim N, Jinks-Robertson S. Abasic sites in the transcribed strand of yeast DNA are removed by transcription-coupled nucleotide excision repair. Mol Cell Biol. 2010; 30(13):3206-3215. DOI: 10.1128/MCB.00308-10
  11. 11. Gangavarapu V, Santa Maria SR, Prakash S, Prakash L. Requirement of replication checkpoint protein kinases Mec1/Rad53 for postreplication repair in yeast. mBio. 2011; 2(3):e00079–e00011. DOI: 10.1128/mBio.00079-11
  12. 12. Lehoczký P, McHugh PJ, Chovanec M. DNA interstrand cross-link repair in Saccharomyces cerevisiae. FEMS Microbiol Rev. 2007 31(2):109-133. DOI: 10.1111/j.1574-6976.2006.00046.x
  13. 13. Resnick MA, Setlow JK. Repair of pyrimidine dimer damage induced in yeast by ultraviolet light. J Bacteriol. 1972; 109(3):979-986. DOI: 10.1128/JB.109.3.979-986.1972
  14. 14. Hung KF, Sidorova JM, Nghiem P, Kawasumi M. The 6-4 photoproduct is the trigger of UV-induced replication blockage and ATR activation. Proc Natl Acad Sci U S A. 2020; 117(23):12806-12816. DOI: 10.1073/pnas.1917196117
  15. 15. Boiteux S, Jinks-Robertson S. DNA repair mechanisms and the bypass of DNA damage in Saccharomyces cerevisiae. Genetics. 2013;193(4):1025-1064. DOI: 10.1534/genetics.112.145219
  16. 16. Lehmann CP, Jiménez-Martín A, Branzei D, Tercero JA. Prevention of unwanted recombination at damaged replication forks. Curr Genet. 2020; 66(6):1045-1051. DOI: 10.1007/s00294-020-01095-7
  17. 17. Xu X, Lin A, Zhou C, Blackwell SR, Zhang Y, Wang Z, Feng Q, Guan R, Hanna MD, Chen Z, Xiao W. Involvement of budding yeast Rad5 in translesion DNA synthesis through physical interaction with Rev1. Nucleic Acids Res. 2016 Jun 20; 44(11):5231-5245. DOI: 10.1093/nar/gkw183
  18. 18. Kramara J, Osia B, Malkova A. Break-Induced Replication: The Where, The Why, and The How. Trends Genet. 2018; 34(7):518-531. DOI: 10.1016/j.tig.2018.04.002
  19. 19. Paek AL, Kaochar S, Jones H, Elezaby A, Shanks L, Weinert T. Fusion of nearby inverted repeats by a replication-based mechanism leads to formation of dicentric and acentric chromosomes that cause genome instability in budding yeast. Genes Dev. 2009; 23(24):2861-2875. DOI: 10.1101/gad.1862709
  20. 20. Kitada K, Yamasaki T. The complicated copy number alterations in chromosome 7 of a lung cancer cell line is explained by a model based on repeated breakage-fusion-bridge cycles. Cancer Genet Cytogenet. 2008; 185(1):11-19. DOI: 10.1016/j.cancergencyto.2008.04.005
  21. 21. Plug-DeMaggio AW, Sundsvold T, Wurscher MA, Koop JI, Klingelhutz AJ, McDougall JK. Telomere erosion and chromosomal instability in cells expressing the HPV oncogene 16E6. Oncogene. 2004; 23(20):3561-3571. DOI: 10.1038/sj.onc.1207388
  22. 22. Lanz MC, Dibitetto D, Smolka MB. DNA damage kinase signaling: checkpoint and repair at 30 years. EMBO J. 2019; 38(18):e101801. DOI: 10.15252/embj.2019101801
  23. 23. Zhou C, Elia AE, Naylor ML, Dephoure N, Ballif BA, Goel G, Xu Q, Ng A, Chou DM, Xavier RJ, Gygi SP, Elledge SJ. Profiling DNA damage-induced phosphorylation in budding yeast reveals diverse signaling networks. Proc Natl Acad Sci U S A. 2016; 113(26):E3667–E3675. DOI: 10.1073/pnas.1602827113
  24. 24. Weinert TA, Hartwell LH. The RAD9 gene controls the cell cycle response to DNA damage in Saccharomyces cerevisiae. Science. 1988; 241(4863):317-322. DOI: 10.1126/science.3291120
  25. 25. 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(22):8396-8409. DOI: 10.1128/MCB.01317-06
  26. 26. Sabbioneda S, Bortolomai I, Giannattasio M, Plevani P, Muzi-Falconi M. Yeast Rev1 is cell cycle regulated, phosphorylated in response to DNA damage and its binding to chromosomes is dependent upon MEC1. DNA Repair (Amst). 2007;6(1):121-127. DOI: 10.1016/j.dnarep.2006.09.002
  27. 27. Zhao X, Chabes A, Domkin V, Thelander L, Rothstein R. The ribonucleotide reductase inhibitor Sml1 is a new target of the Mec1/Rad53 kinase cascade during growth and in response to DNA damage. EMBO J. 2001;20(13):3544-3553. DOI: 10.1093/emboj/20.13.3544
  28. 28. Dotiwala F, Eapen VV, Harrison JC, Arbel-Eden A, Ranade V, Yoshida S, Haber JE. DNA damage checkpoint triggers autophagy to regulate the initiation of anaphase. Proc Natl Acad Sci U S A. 2013;110(1):E41–E49. DOI: 10.1073/pnas.1218065109
  29. 29. Singh RK, Kabbaj MH, Paik J, Gunjan A. Histone levels are regulated by phosphorylation and ubiquitylation-dependent proteolysis. Nat Cell Biol. 2009;11(8):925-933. DOI: 10.1038/ncb1903
  30. 30. 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(5442):1166-1171. DOI: 10.1126/science.286.5442.1166
  31. 31. Clémenson C, Marsolier-Kergoat MC. DNA damage checkpoint inactivation: adaptation and recovery. DNA Repair (Amst). 2009 Sep 2;8(9):1101-1109. DOI: 10.1016/j.dnarep.2009.04.008
  32. 32. Yeung M, Durocher D. Srs2 enables checkpoint recovery by promoting disassembly of DNA damage foci from chromatin. DNA Repair (Amst). 2011;10(12):1213-1222. DOI: 10.1016/j.dnarep.2011.09.005
  33. 33. Moriel-Carretero M, Pasero P, Pardo B. DDR Inc., one business, two associates. Curr Genet. 2019; 65(2):445-451. DOI: 10.1007/s00294-018-0908-7
  34. 34. Lee SE, Pellicioli A, Malkova A, Foiani M, Haber JE. The Saccharomyces recombination protein Tid1p is required for adaptation from G2/M arrest induced by a double-strand break. Curr Biol. 2001;11(13):1053-1057. DOI: 10.1016/s0960-9822(01)00296-2
  35. 35. Dewar JM, Lydall D. Similarities and differences between "uncapped" telomeres and DNA double-strand breaks. Chromosoma. 2012;121(2):117-130. DOI: 10.1007/s00412-011-0357-2
  36. 36. Doksani Y, Bermejo R, Fiorani S, Haber JE, Foiani M. Replicon dynamics, dormant origin firing, and terminal fork integrity after double-strand break formation. Cell. 2009; 137(2):247-258. DOI: 10.1016/j.cell.2009.02.016
  37. 37. Daley JM, Wilson TE. Rejoining of DNA double-strand breaks as a function of overhang length. Mol Cell Biol. 2005 Feb;25(3):896-906. doi: 10.1128/MCB.25.3.896-906.2005. PMID: 15657419; PMCID: PMC544009
  38. 38. Janke R, Herzberg K, Rolfsmeier M, Mar J, Bashkirov VI, Haghnazari E, Cantin G, Yates JR 3rd, Heyer WD. A truncated DNA-damage-signaling response is activated after DSB formation in the G1 phase of Saccharomyces cerevisiae. Nucleic Acids Res. 2010;38(7):2302-2313. DOI: 10.1093/nar/gkp1222
  39. 39. Basile G, Aker M, Mortimer RK. Nucleotide sequence and transcriptional regulation of the yeast recombinational repair gene RAD51. Mol Cell Biol. 1992; 12(7):3235-3246. DOI: 10.1128/mcb.12.7.3235
  40. 40. Waterman DP, Haber JE, Smolka MB. Checkpoint Responses to DNA Double-Strand Breaks. Annu Rev Biochem. 2020; 89:103-133. DOI: 10.1146/annurev-biochem-011520-104722
  41. 41. 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(7011):1011-1017. DOI: 10.1038/nature02964
  42. 42. 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(6):777-788. DOI: 10.1016/j.cell.2004.11.033
  43. 43. 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(18):4113-4125. DOI: 10.1038/sj.emboj.7601835
  44. 44. Chambers AL, Downs JA. The RSC and INO80 chromatin-remodeling complexes in DNA double-strand break repair. Prog Mol Biol Transl Sci. 2012; 110:229-261. DOI: 10.1016/B978-0-12-387665-2.00009-2
  45. 45. Costelloe T, Louge R, Tomimatsu N, Mukherjee B, Martini E, Khadaroo B, Dubois K, Wiegant WW, Thierry A, Burma S, van Attikum H, Llorente B. The yeast Fun30 and human SMARCAD1 chromatin remodellers promote DNA end resection. 2012; Nature 489:581-584. DOI: org/10.1038/nature11353.
  46. 46. Eapen VV, Sugawara N, Tsabar M, Wu WH, Haber JE. The Saccharomyces cerevisiae chromatin remodeler Fun30 regulates DNA end resection and checkpoint deactivation. Mol Cell Biol. 2012; 32(22):4727-4740. DOI: 10.1128/MCB.00566-12
  47. 47. Melo JA, Cohen J, Toczyski DP. Two checkpoint complexes are independently recruited to sites of DNA damage in vivo. Genes Dev. 2001; 15(21):2809-2821. DOI: 10.1101/gad.903501
  48. 48. Garvik B, Carson M, Hartwell L. Single-stranded DNA arising at telomeres in cdc13 mutants may constitute a specific signal for the RAD9 checkpoint. Mol Cell Biol. 1995; 15(11):6128-6138. DOI: 10.1128/mcb.15.11.6128
  49. 49. Traven A, Heierhorst J. SQ/TQ cluster domains: concentrated ATM/ATR kinase phosphorylation site regions in DNA-damage-response proteins. Bioessays 2005; 27:397-407. DOI://dx.doi.org/10.1002/bies.20204
  50. 50. Memisoglu G, Lanz MC, Eapen VV, Jordan JM, Lee K, Smolka MB, Haber JE. Mec1ATR Autophosphorylation and Ddc2ATRIP Phosphorylation Regulates DNA Damage Checkpoint Signaling. Cell Rep. 2019; 28(4):1090-1102.e3. DOI: 10.1016/j.celrep.2019.06.068
  51. 51. 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(18):2400-2410. DOI: 10.1038/emboj.2008.171
  52. 52. Blankley RT, Lydall D. A domain of Rad9 specifically required for activation of Chk1 in budding yeast. J Cell Sci. 2004; 117(4):601-608. DOI: 10.1242/jcs.00907
  53. 53. Conde F, Ontoso D, Acosta I, Gallego-Sánchez A, Bueno A, San-Segundo PA. Regulation of tolerance to DNA alkylating damage by Dot1 and Rad53 in Saccharomyces cerevisiae. DNA Repair (Amst). 2010; 9(10):1038-1049. DOI: 10.1016/j.dnarep.2010.07.003
  54. 54. 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(45):45027-45033. DOI: 10.1074/jbc.M306783200
  55. 55. Gardner R, Putnam CW, Weinert T. RAD53, DUN1 and PDS1 define two parallel G2/M checkpoint pathways in budding yeast. EMBO J. 1999;18(11):3173-3185. DOI: 10.1093/emboj/18.11.3173
  56. 56. Lyndaker AM, Goldfarb T, Alani E. Mutants defective in Rad1-Rad10-Slx4 exhibit a unique pattern of viability during mating-type switching in Saccharomyces cerevisiae. Genetics. 2008; 179(4):1807-1821. DOI: 10.1534/genetics.108.090654
  57. 57. Flott S, Kwon Y, Pigli YZ, Rice PA, Sung P, Jackson SP. Regulation of Rad51 function by phosphorylation. EMBO Rep. 2011; 12(8):833-839. DOI: 10.1038/embor.2011.127
  58. 58. Bashkirov VI, King JS, Bashkirova EV, Schmuckli-Maurer J, Heyer WD. DNA repair protein Rad55 is a terminal substrate of the DNA damage checkpoints. Mol Cell Biol. 2000; 20(12):4393-4404. DOI: 10.1128/mcb.20.12.4393-4404.2000
  59. 59. Toh GW, Sugawara N, Dong J, Toth R, Lee SE, Haber JE, Rouse J. Mec1/Tel1-dependent phosphorylation of Slx4 stimulates Rad1-Rad10-dependent cleavage of non-homologous DNA tails. DNA Repair (Amst). 2010; 9(6):718-726. DOI: 10.1016/j.dnarep.2010.02.013
  60. 60. Zhou C, Elia AE, Naylor ML, Dephoure N, Ballif BA, Goel G, Xu Q, Ng A, Chou DM, Xavier RJ, Gygi SP, Elledge SJ. Profiling DNA damage-induced phosphorylation in budding yeast reveals diverse signaling networks. Proc Natl Acad Sci U S A. 2016 Jun 28;113(26):E3667-75. doi: 10.1073/pnas.1602827113. Epub 2016 Jun 13. PMID: 27298372; PMCID: PMC4932963
  61. 61. Breitkreutz A, Choi H, Sharom JR, Boucher L, Neduva V, Larsen B, Lin ZY, Breitkreutz BJ, Stark C, Liu G, Ahn J, Dewar-Darch D, Reguly T, Tang X, Almeida R, Qin ZS, Pawson T, Gingras AC, Nesvizhskii AI, Tyers M. A global protein kinase and phosphatase interaction network in yeast. Science. 2010 May 21;328(5981):1043-1046. DOI: 10.1126/science.1176495
  62. 62. Ubersax JA, Woodbury EL, Quang PN, Paraz M, Blethrow JD, Shah K, Shokat KM, Morgan DO. Targets of the cyclin-dependent kinase Cdk1. Nature 2003; 425:859-864
  63. 63. Huang M, Zhou Z, Elledge SJ. The DNA replication and damage checkpoint pathways induce transcription by inhibition of the Crt1 repressor. Cell. 1998; 94(5):595-605. DOI: 10.1016/s0092-8674(00)81601-3
  64. 64. Lee YD, Wang J, Stubbe J, Elledge SJ. Dif1 is a DNA-damage-regulated facilitator of nuclear import for ribonucleotide reductase. Mol Cell. 2008; 32(1):70-80. DOI: 10.1016/j.molcel.2008.08.018
  65. 65. Ramos F, Villoria MT, Alonso-Rodríguez E, Clemente-Blanco A. Role of protein phosphatases PP1, PP2A, PP4 and Cdc14 in the DNA damage response. Cell Stress. 2019; 3(3):70-85. DOI:10.15698/cst2019.03.178
  66. 66. Rossi SE, Ajazi A, Carotenuto W, Foiani M, Giannattasio M. Rad53-Mediated Regulation of Rrm3 and Pif1 DNA Helicases Contributes to Prevention of Aberrant Fork Transitions under Replication Stress. Cell Rep. 2015; 13(1):80-92. DOI: 10.1016/j.celrep.2015.08.073
  67. 67. Chiolo I, Carotenuto W, Maffioletti G, Petrini JH, Foiani M, Liberi G. Srs2 and Sgs1 DNA helicases associate with Mre11 in different subcomplexes following checkpoint activation and CDK1-mediated Srs2 phosphorylation. Mol Cell Biol. 2005;25(13):5738-5751. DOI:
  68. 68. Marsolier MC, Roussel P, Leroy C, Mann C. Involvement of the PP2C-like phosphatase Ptc2p in the DNA checkpoint pathways of Saccharomyces cerevisiae. Genetics. 2000;154(4):1523-1532
  69. 69. Leroy C, Lee SE, Vaze MB, Ochsenbein 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(3):827-835. doi: 10.1016/s1097-2765(03)00058-3
  70. 70. Kim JA, Hicks WM, Li J, Tay SY, Haber JE. Protein phosphatases pph3, ptc2, and ptc3 play redundant roles in DNA double-strand break repair by homologous recombination. Mol Cell Biol. 2011; 31(3):507-516. DOI: 10.1128/MCB.01168-10
  71. 71. Johnson AJ, Veljanoski F, O’Doherty PJ, Zaman MS, Petersingham G, Bailey TD, Münch G, Kersaitis C, Wu MJ. Molecular insight into arsenic toxicity via the genome-wide deletion mutant screening of Saccharomyces cerevisiae. Metallomics. 2016; 8(2):228-235. DOI: 10.1039/c5mt00261c
  72. 72. Pike BL, Tenis N, Heierhorst J. Rad53 kinase activation-independent replication checkpoint function of the N-terminal forkhead-associated (FHA1) domain. J Biol Chem. 2004;279(38):39636-33944. DOI: 10.1074/jbc.M405080200
  73. 73. 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(7075):497-501. DOI: 10.1038/nature04384
  74. 74. Kim JA, Haber JE. Chromatin assembly factors Asf1 and CAF-1 have overlapping roles in deactivating the DNA damage checkpoint when DNA repair is complete. Proc Natl Acad Sci U S A. 2009; 106(4):1151-1161. DI: 10.1073/pnas.0812578106
  75. 75. Tsabar M, Waterman DP, Aguilar F, Katsnelson L, Eapen VV, Memisoglu G, Haber JE. Asf1 facilitates dephosphorylation of Rad53 after DNA double-strand break repair. Genes Dev. 2016;30(10):1211-1224. DOI: 10.1101/gad.280685.116
  76. 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(17):2437-2449. DOI: 10.1101/gad.1440206
  77. 77. Tong ZB, Ai HS, Li JB. The Mechanism of Chromatin Remodeler SMARCAD1/Fun30 in Response to DNA Damage. Front Cell Dev Biol. 2020 Sep 25;8:560098. DOI: 10.3389/fcell.2020.560098
  78. 78. Costelloe T, Louge R, Tomimatsu N, Mukherjee B, Martini E, Khadaroo B, Dubois K, Wiegant WW, Thierry A, Burma S, van Attikum H, Llorente B. The yeast Fun30 and human SMARCAD1 chromatin remodellers promote DNA end resection. Nature. 2012; 489(7417):581-584. DOI: 10.1038/nature11353
  79. 79. Ferrari M, Nachimuthu BT, Donnianni RA, Klein H, Pellicioli A. Tid1/Rdh54 translocase is phosphorylated through a Mec1- and Rad53-dependent manner in the presence of DSB lesions in budding yeast. DNA Repair (Amst). 2013; 12(5):347-355. DOI: 10.1016/j.dnarep.2013.02.004
  80. 80. Vaze MB, Pellicioli A, Lee SE, Ira G, Liberi G, Arbel-Eden A, Foiani M, Haber JE. Recovery from checkpoint-mediated arrest after repair of a double-strand break requires Srs2 helicase. Mol Cell. 2002; 10(2):373-385. DOI: 10.1016/s1097-2765(02)00593-2
  81. 81. Li S, Makovets S, Matsuguchi T, Blethrow JD, Shokat KM, Blackburn EH. Cdk1-dependent phosphorylation of Cdc13 coordinates telomere elongation during cell-cycle progression. Cell. 2009 Jan 9;136(1):50-61. doi: 10.1016/j.cell.2008.11.027. PMID: 19135888; PMCID: PMC2642970
  82. 82. Phillips JA, Chan A, Paeschke K, Zakian VA. The pif1 helicase, a negative regulator of telomerase, acts preferentially at long telomeres. PLoS Genet. 2015;11(4):e1005186. DOI: 10.1371/journal.pgen.1005186
  83. 83. Dewar JM, Lydall D. Pif1- and Exo1-dependent nucleases coordinate checkpoint activation following telomere uncapping. EMBO J. 2010 Dec 1;29(23):4020-34. doi: 10.1038/emboj.2010.267. Epub 2010 Nov 2. PMID: 21045806; PMCID: PMC3020640
  84. 84. Fontana GA, Reinert JK, Thomä NH, Rass U. Shepherding DNA ends: Rif1 protects telomeres and chromosome breaks. Microb Cell. 2018; 5(7):327-343. DOI: 10.15698/mic2018.07.639.37
  85. 85. Mattarocci S, Reinert JK, Bunker RD, Fontana GA, Shi T, Klein D, Cavadini S, Faty M, Shyian M, Hafner L, Shore D, Thomä NH, Rass U. Rif1 maintains telomeres and mediates DNA repair by encasing DNA ends. Nat Struct Mol Biol. 2017; 24(7):588-595. DOI: 10.1038/nsmb.3420
  86. 86. Maringele L, Lydall D. EXO1-dependent single-stranded DNA at telomeres activates subsets of DNA damage and spindle checkpoint pathways in budding yeast yku70Delta mutants. Genes Dev. 2002 Aug 1;16(15):1919-1933
  87. 87. Jia X, Weinert T, Lydall D. Mec1 and Rad53 inhibit formation of single-stranded DNA at telomeres of Saccharomyces cerevisiae cdc13-1 mutants. Genetics. 2004; 166(2):753-764. DOI: 10.1534/genetics.166.2.753
  88. 88. Holstein EM, Ngo G, Lawless C, Banks P, Greetham M, Wilkinson D, Lydall D. Systematic Analysis of the DNA Damage Response Network in Telomere Defective Budding Yeast. G3 (Bethesda). 2017; 7(7):2375-2389. DOI: 10.1534/g3.117.042283
  89. 89. Vasianovich Y, Harrington LA, Makovets S. Break-induced replication requires DNA damage-induced phosphorylation of Pif1 and leads to telomere lengthening. PLoS Genet. 2014; 10(10):e1004679. DOI: 10.1371/journal.pgen.1004679
  90. 90. Sandell LL, Zakian VA. Loss of a yeast telomere: arrest, recovery, and chromosome loss. Cell. 1993; 75(4):729-739. DOI: 10.1016/0092-8674(93)90493-a
  91. 91. Toczyski DP, Galgoczy DJ, Hartwell LH. CDC5 and CKII control adaptation to the yeast DNA damage checkpoint. Cell. 1997; 90(6):1097-1106. DOI: 10.1016/s0092-8674(00)80375-x
  92. 92. Mersaoui SY, Gravel S, Karpov V, Wellinger RJ. DNA damage checkpoint adaptation genes are required for division of cells harbouring eroded telomeres. Microb Cell. 2015; 2(10):394-405. DOI: 10.15698/mic2015.10.229
  93. 93. Vidanes GM, Sweeney FD, Galicia S, Cheung S, Doyle JP, Durocher D, Toczyski DP. CDC5 inhibits the hyperphosphorylation of the checkpoint kinase Rad53, leading to checkpoint adaptation. PLoS Biol. 2010 Jan 26;8(1):e1000286. DOI: 10.1371/journal.pbio.1000286
  94. 94. Myung K, Chen C, Kolodner RD. Multiple pathways cooperate in the suppression of genome instability in Saccharomyces cerevisiae. Nature. 2001; 411(6841):1073-1076. DOI 10.1038/35082608. PMID: 11429610
  95. 95. Marians KJ. Lesion Bypass and the Reactivation of Stalled Replication Forks. Annu Rev Biochem. 2018; 87:217-238. DOI: 10.1146/annurev-biochem-062917-011921
  96. 96. 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(24):3236-3252. DOI: 10.1101/gad.239802
  97. 97. Tercero JA, Diffley JF. Regulation of DNA replication fork progression through damaged DNA by the Mec1/Rad53 checkpoint. Nature. 2001 Aug 2;412(6846):553-557. DOI: 10.1038/35087607
  98. 98. Lopes M, Foiani M, Sogo JM. Multiple mechanisms control chromosome integrity after replication fork uncoupling and restart at irreparable UV lesions. Mol Cell. 2006; 21(1):15-27. DOI: 10.1016/j.molcel.2005.11.015.0
  99. 99. 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(4):601-611. DOI: 10.1016/j.cell.2008.02.050
  100. 100. Sau S, Liefshitz B, Kupiec M. The Yeast PCNA Unloader Elg1 RFC-Like Complex Plays a Role in Eliciting the DNA Damage Checkpoint. mBio. 2019; 10(3):e01159–e01119. DOI: 10.1128/mBio.01159-19
  101. 101. Osborn AJ, Elledge SJ. Mrc1 is a replication fork component whose phosphorylation in response to DNA replication stress activates Rad53. Genes Dev. 2003; 17(14):1755-1767. DOI: 10.1101/gad.1098303
  102. 102. Bjergbaek L, Cobb JA, Tsai-Pflugfelder M, Gasser SM. Mechanistically distinct roles for Sgs1p in checkpoint activation and replication fork maintenance. EMBO J. 2005; 24(2):405-417. DOI: 10.1038/sj.emboj.7600511
  103. 103. Chabes A, Georgieva B, Domkin V, Zhao X, Rothstein R, Thelander L. Survival of DNA damage in yeast directly depends on increased dNTP levels allowed by relaxed feedback inhibition of ribonucleotide reductase. Cell. 2003; 112(3):391-401. DOI: 10.1016/s0092-8674(03)00075-8
  104. 104. Begley U, Dyavaiah M, Patil A, Rooney JP, DiRenzo D, Young CM, Conklin DS, Zitomer RS, Begley TJ. Trm9-catalyzed tRNA modifications link translation to the DNA damage response. Mol Cell. 2007; 28(5):860-870. DOI: 10.1016/j.molcel.2007.09.021
  105. 105. Domkin V, Thelander L, Chabes A. Yeast DNA damage-inducible Rnr3 has a very low catalytic activity strongly stimulated after the formation of a cross-talking Rnr1/Rnr3 complex. J Biol Chem. 2002; 277(21):18574-18578. DOI: 10.1074/jbc.M201553200
  106. 106. Poli J, Tsaponina O, Crabbé L, Keszthelyi A, Pantesco V, Chabes A, Lengronne A, Pasero P. dNTP pools determine fork progression and origin usage under replication stress. EMBO J. 2012;31(4):883-894. doi: 10.1038/emboj.2011.470
  107. 107. Williams LN, Marjavaara L, Knowels GM, Schultz EM, Fox EJ, Chabes A, Herr AJ. dNTP pool levels modulate mutator phenotypes of error-prone DNA polymerase ε variants. Proc Natl Acad Sci U S A. 2015; 112(19):E2457–E2466. DOI: 10.1073/pnas.1422948112
  108. 108. Fasullo M, Tsaponina O, Sun M, Chabes A. Elevated dNTP levels suppress hyper-recombination in Saccharomyces cerevisiae S-phase checkpoint mutants. Nucleic Acids Res. 2010; 38(4):1195-1203. DOI: 10.1093/nar/gkp1064
  109. 109. Arbel M, Liefshitz B, Kupiec M. How yeast cells deal with stalled replication forks. Curr Genet. 2020; 66(5):911-915. DOI: 10.1007/s00294-020-01082-y
  110. 110. Gangloff S, Soustelle C, Fabre F. Homologous recombination is responsible for cell death in the absence of the Sgs1 and Srs2 helicases. Nat Genet. 2000; 25(2):192-194. DOI: 10.1038/76055
  111. 111. Rupp WD, Howard-Flanders P. Discontinuities in the DNA synthesized in an excision-defective strain of Escherichia coli following ultraviolet irradiation. J Mol Biol. 1968 Jan 28;31(2):291-304. doi: 10.1016/0022-2836(68)90445-2
  112. 112. Gallo D, Kim T, Szakal B, Saayman X, Narula A, Park Y, Branzei D, Zhang Z, Brown GW. Rad5 Recruits Error-Prone DNA Polymerases for Mutagenic Repair of ssDNA Gaps on Undamaged Templates. Mol Cell. 2019; 73(5):900-914.e9. DOI: 10.1016/j.molcel.2019.01.001
  113. 113. Guilliam TA, Yeeles JTP. Reconstitution of translesion synthesis reveals a mechanism of eukaryotic DNA replication restart. Nat Struct Mol Biol. 2020; 27(5):450-460. DOI: 10.1038/s41594-020-0418-4
  114. 114. Kochenova OV, Bezalel-Buch R, Tran P, Makarova AV, Chabes A, Burgers PM, Shcherbakova PV. Yeast DNA polymerase ζ maintains consistent activity and mutagenicity across a wide range of physiological dNTP concentrations. Nucleic Acids Res. 2017; 45(3):1200-1218. DOI: 10.1093/nar/gkw1149
  115. 115. Andersen PL, Xu F, Xiao W. Eukaryotic DNA damage tolerance and translesion synthesis through covalent modifications of PCNA. Cell Res. 2008; 18(1):162-173. DOI: 10.1038/cr.2007.114
  116. 116. Blastyák A, Pintér L, Unk I, Prakash L, Prakash S, Haracska L. Yeast Rad5 protein required for postreplication repair has a DNA helicase activity specific for replication fork regression. Mol Cell. 2007; 28(1):167-175. DOI: 10.1016/j.molcel.2007.07.030
  117. 117. Bryant EE, Šunjevarić I, Berchowitz L, Rothstein R, Reid RJD. Rad5 dysregulation drives hyperactive recombination at replication forks resulting in cisplatin sensitivity and genome instability. Nucleic Acids Res. 2019; 47(17):9144-9159. DOI: 10.1093/nar/gkz631
  118. 118. Hammet A, Pike BL, Heierhorst J. Posttranscriptional regulation of the RAD5 DNA repair gene by the Dun1 kinase and the Pan2-Pan3 poly(A)-nuclease complex contributes to survival of replication blocks. J Biol Chem. 2002; 277(25):22469-22474. DOI: 10.1074/jbc.M202473200
  119. 119. Fasullo M, Dong Z, Sun M, Zeng L. Saccharomyces cerevisiae RAD53 (CHK2) but not CHK1 is required for double-strand break-initiated SCE and DNA damage-associated SCE after exposure to X rays and chemical agents. DNA Repair (Amst). 2005; 4(11):1240-1251. DOI: 10.1016/j.dnarep.2005.06.006
  120. 120. Waters LS, Walker GC. The critical mutagenic translesion DNA polymerase Rev1 is highly expressed during G(2)/M phase rather than S phase. Proc Natl Acad Sci U S A. 2006; 103(24):8971-8976. DOI: 10.1073/pnas.0510167103
  121. 121. García-Rodríguez N, Wong RP, Ulrich HD. The helicase Pif1 functions in the template switching pathway of DNA damage bypass. Nucleic Acids Res. 2018; 46(16):8347-8356. DOI: 10.1093/nar/gky648
  122. 122. Beranek DT. Distribution of methyl and ethyl adducts following alkylation with monofunctional alkylating agents. Mutat Res. 1990; 231(1):11-30. DOI: 10.1016/0027-5107(90)90173-2
  123. 123. Plosky BS, Frank EG, Berry DA, Vennall GP, McDonald JP, Woodgate R. Eukaryotic Y-family polymerases bypass a 3-methyl-2′-deoxyadenosine analog in vitro and methyl methanesulfonate-induced DNA damage in vivo. Nucleic Acids Res. 2008 Apr;36(7):2152-2162. doi: 10.1093/nar/gkn058
  124. 124. Johnson RE, Yu SL, Prakash S, Prakash L. A role for yeast and human translesion synthesis DNA polymerases in promoting replication through 3-methyl adenine. Mol Cell Biol. 2007; 27(20):7198-7205. DOI: 10.1128/MCB.01079-07
  125. 125. Vanoli F, Fumasoni M, Szakal B, Maloisel L, Branzei D. Replication and recombination factors contributing to recombination-dependent bypass of DNA lesions by template switch. PLoS Genet. 2010; 6(11):e1001205. DOI: 10.1371/journal.pgen.1001205
  126. 126. Minca EC, Kowalski D. Multiple Rad5 activities mediate sister chromatid recombination to bypass DNA damage at stalled replication forks. Mol Cell. 2010; 38(5):649-661. DOI: 10.1016/j.molcel.2010.03.020
  127. 127. Yu SL, Johnson RE, Prakash S, Prakash L. Requirement of DNA polymerase eta for error-free bypass of UV-induced CC and TC photoproducts. Mol Cell Biol. 2001; 21(1):185-1888. DOI: 10.1128/MCB.21.1.185-188.2001
  128. 128. Johnson RE, Haracska L, Prakash S, Prakash L. Role of DNA polymerase eta in the bypass of a (6-4) TT photoproduct. Mol Cell Biol. 2001; 21(10):3558-3563. DOI: 10.1128/MCB.21.10.3558-3563.2001
  129. 129. Zhang H, Lawrence CW. The error-free component of the RAD6/RAD18 DNA damage tolerance pathway of budding yeast employs sister-strand recombination. Proc Natl Acad Sci U S A. 2005 Nov 1;102(44):15954-15959. DOI: 10.1073/pnas.0504586102
  130. 130. Xu X, Blackwell S, Lin A, Li F, Qin Z, Xiao W. Error-free DNA-damage tolerance in Saccharomyces cerevisiae. Mutat Res Rev Mutat Res. 2015; 764:43-50. DOI: 10.1016/j.mrrev.2015.02.001
  131. 131. Jablonowski CM, Cussiol JR, Oberly S, Yimit A, Balint A, Kim T, Zhang Z, Brown GW, Smolka MB. Termination of Replication Stress Signaling via Concerted Action of the Slx4 Scaffold and the PP4 Phosphatase. Genetics. 2015; 201(3):937-949. DOI: 10.1534/genetics.115.181479
  132. 132. Dibitetto D, Ferrari M, Rawal CC, Balint A, Kim T, Zhang Z, Smolka MB, Brown GW, Marini F, Pellicioli A. Slx4 and Rtt107 control checkpoint signalling and DNA resection at double-strand breaks. Nucleic Acids Res. 2016; 44(2):669-682. DOI: 10.1093/nar/gkv1080
  133. 133. Ohouo PY, Bastos de Oliveira FM, Liu Y, Ma CJ, Smolka MB. DNA-repair scaffolds dampen checkpoint signalling by counteracting the adaptor Rad9. Nature. 2013;493(7430):120-124. DOI: 10.1038/nature11658
  134. 134. Campos A, Clemente-Blanco A. Cell Cycle and DNA Repair Regulation in the Damage Response: Protein Phosphatases Take Over the Reins. Int J Mol Sci. 2020; 21(2):446. DOI: 10.3390/ijms21020446
  135. 135. Hustedt N, Seeber A, Sack R, Tsai-Pflugfelder M, Bhullar B, Vlaming H, van Leeuwen F, Guénolé A, van Attikum H, Srivas R, Ideker T, Shimada K, Gasser SM. Yeast PP4 interacts with ATR homolog Ddc2-Mec1 and regulates checkpoint signaling. Mol Cell. 2015; 57(2):273-289. DOI: 10.1016/j.molcel.2014.11.016
  136. 136. Heideker J, Lis ET, Romesberg FE. Phosphatases, DNA damage checkpoints and checkpoint deactivation. Cell Cycle. 2007; 6(24):3058-3064. DOI: 10.4161/cc.6.24.5100
  137. 137. Douville J, David J, Fortier PK, Ramotar D. The yeast phosphotyrosyl phosphatase activator protein, yPtpa1/Rrd1, interacts with Sit4 phosphatase to mediate resistance to 4-nitroquinoline-1-oxide and UVA. Curr Genet. 2004 Aug;46(2):72-81. doi: 10.1007/s00294-004-0513
  138. 138. St John N, Freedland J, Baldino H, Doyle F, Cera C, Begley T, Fasullo M. Genome Profiling for Aflatoxin B1 Resistance in Saccharomyces cerevisiae Reveals a Role for the CSM2/SHU Complex in Tolerance of Aflatoxin B1-Associated DNA Damage. G3 (Bethesda). 2020; 10(11):3929-3947. DOI: 10.1534/g3.120.401723
  139. 139. Ferrari E, Bruhn C, Peretti M, Cassani C, Carotenuto WV, Elgendy M, Shubassi G, Lucca C, Bermejo R, Varasi M, Minucci S, Longhese MP, Foiani M. PP2A Controls Genome Integrity by Integrating Nutrient-Sensing and Metabolic Pathways with the DNA Damage Response. Mol Cell. 2017; 67(2):266-281.e4. DOI: 10.1016/j.molcel.2017.05.027
  140. 140. Jain D, Patel N, Shelton M, Basu A, Roque R, Siede W. Enhancement of cisplatin sensitivity by NSC109268 in budding yeast and human cancer cells is associated with inhibition of S-phase progression. Cancer Chemother Pharmacol. 2010; 66(5):945-952. DOI: 10.1007/s00280-010-1246-1248
  141. 141. Aouida M, Eshrif A, Ramotar D. Yeast Lacking the PP2A Phosphatase Regulatory Subunit Rts1 Sensitizes rad51 Mutants to Specific DNA Damaging Agents. Front Genet. 2019; 10:1117. DOI: 10.3389/fgene.2019.01117
  142. 142. Travesa A, Duch A, Quintana DG. Distinct phosphatases mediate the deactivation of the DNA damage checkpoint kinase Rad53. J Biol Chem. 2008; 283(25):17123-17130. DOI: 10.1074/jbc.M801402200
  143. 143. 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(22):9290-9295. DOI: 10.1073/pnas.0703252104
  144. 144. Conde F, Ontoso D, Acosta I, Gallego-Sánchez A, Bueno A, San-Segundo PA. Regulation of tolerance to DNA alkylating damage by Dot1 and Rad53 in Saccharomyces cerevisiae. DNA Repair (Amst). 2010; 9(10):1038-1049. DOI: 10.1016/j.dnarep.2010.07.003
  145. 145. Gangavarapu V, Santa Maria SR, Prakash S, Prakash L. Requirement of replication checkpoint protein kinases Mec1/Rad53 for postreplication repair in yeast. mBio. 2011; 2(3):e00079–e00011. DOI: 10.1128/mBio.00079-11
  146. 146. Neecke H, Lucchini G, Longhese MP. Cell cycle progression in the presence of irreparable DNA damage is controlled by a Mec1- and Rad53-dependent checkpoint in budding yeast. EMBO J. 1999; 18(16):4485-4497. DOI: 10.1093/emboj/18.16.4485
  147. 147. Hishida T, Kubota Y, Carr AM, Iwasaki H. RAD6-RAD18-RAD5-pathway-dependent tolerance to chronic low-dose ultraviolet light. Nature. 2009; 457(7229):612-615. DOI: 10.1038/nature07580
  148. 148. Fasullo M, Chen Y, Bortcosh W, Sun M, Egner PA. Aflatoxin B(1)-Associated DNA Adducts Stall S Phase and Stimulate Rad51 foci in Saccharomyces cerevisiae. J Nucleic Acids. 2010 Dec 2;2010:456487. DOI: 10.4061/2010/456487
  149. 149. Guo Y, Breeden LL, Fan W, Zhao LP, Eaton DL, Zarbl H. Analysis of cellular responses to aflatoxin B(1) in yeast expressing human cytochrome P450 1A2 using cDNA microarrays. Mutat Res. 2006; 593(1-2):121-142. DOI: 10.1016/j.mrfmmm.2005.07.001
  150. 150. Keller-Seitz MU, Certa U, Sengstag C, Würgler FE, Sun M, Fasullo M. Transcriptional response of yeast to aflatoxin B1: recombinational repair involving RAD51 and RAD1. Mol Biol Cell. 2004; 15(9):4321-4336. DOI: 10.1091/mbc.e04-05-0375
  151. 151. Sengstag C, Weibel B, Fasullo M. Genotoxicity of aflatoxin B1: evidence for a recombination-mediated mechanism in Saccharomyces cerevisiae. Cancer Res. 1996; 56:5457-5465
  152. 152. Fasullo M, Sun M, Egner P. Stimulation of sister chromatid exchanges and mutation by aflatoxin B1-DNA adducts in Saccharomyces cerevisiae requires MEC1 (ATR), RAD53, and DUN1. Mol Carcinog. 2008 Aug;47(8):608-15. DOI: 10.1002/mc.20417
  153. 153. Godin S, Wier A, Kabbinavar F, Bratton-Palmer DS, Ghodke H, Van Houten B, VanDemark AP, Bernstein KA. The Shu complex interacts with Rad51 through the Rad51 paralogues Rad55-Rad57 to mediate error-free recombination. Nucleic Acids Res. 2013; 41(8):4525-4534. DOI: 10.1093/nar/gkt138
  154. 154. Moyano-Rodriguez Y, Queralt E. PP2A Functions during Mitosis and Cytokinesis in Yeasts. Int J Mol Sci. 2019 Dec 30;21(1):264. DOI: 10.3390/ijms21010264
  155. 155. Wood MD, Sanchez Y. Deregulated Ras signaling compromises DNA damage checkpoint recovery in S. cerevisiae. Cell Cycle. 2010; 9(16):3353-3363. DOI: 10.4161/cc.9.16.12713
  156. 156. Vydzhak O, Bender K, Klermund J, Busch A, Reimann S, Luke B. Checkpoint adaptation in recombination-deficient cells drives aneuploidy and resistance to genotoxic agents. DNA Repair (Amst). 2020; 95:102939. DOI: 10.1016/j.dnarep.2020.102939
  157. 157. Miyamoto I, Ozaki R, Yamaguchi K, Yamamoto K, Kaneko A, Ushimaru T. TORC1 regulates the DNA damage checkpoint via checkpoint protein levels. Biochem Biophys Res Commun. 2019; 510(4):629-635. DOI: 10.1016/j.bbrc.2019.02.010

Written By

Michael Fasullo

Submitted: 01 October 2020 Reviewed: 02 March 2021 Published: 29 March 2021