Next Article in Journal
Cas3 Protein—A Review of a Multi-Tasking Machine
Next Article in Special Issue
Yeast Genome Maintenance by the Multifunctional PIF1 DNA Helicase Family
Previous Article in Journal
Food and Non-Food-Related Behavior across Settings in Children with Prader–Willi Syndrome
Previous Article in Special Issue
Assembly of a G-Quadruplex Repair Complex by the FANCJ DNA Helicase and the REV1 Polymerase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Maintenance of Yeast Genome Integrity by RecQ Family DNA Helicases

by
Sonia Vidushi Gupta
1 and
Kristina Hildegard Schmidt
1,2,*
1
Department of Cell Biology, Microbiology and Molecular Biology, University of South, Florida, Tampa, FL 33620, USA
2
Cancer Biology and Evolution Program, H. Lee Moffitt Cancer Center and Research, Institute, Tampa, FL 33612, USA
*
Author to whom correspondence should be addressed.
Genes 2020, 11(2), 205; https://doi.org/10.3390/genes11020205
Submission received: 20 January 2020 / Revised: 11 February 2020 / Accepted: 14 February 2020 / Published: 18 February 2020
(This article belongs to the Special Issue DNA Helicases: Mechanisms, Biological Pathways, and Disease Relevance)

Abstract

:
With roles in DNA repair, recombination, replication and transcription, members of the RecQ DNA helicase family maintain genome integrity from bacteria to mammals. Mutations in human RecQ helicases BLM, WRN and RecQL4 cause incurable disorders characterized by genome instability, increased cancer predisposition and premature adult-onset aging. Yeast cells lacking the RecQ helicase Sgs1 share many of the cellular defects of human cells lacking BLM, including hypersensitivity to DNA damaging agents and replication stress, shortened lifespan, genome instability and mitotic hyper-recombination, making them invaluable model systems for elucidating eukaryotic RecQ helicase function. Yeast and human RecQ helicases have common DNA substrates and domain structures and share similar physical interaction partners. Here, we review the major cellular functions of the yeast RecQ helicases Sgs1 of Saccharomyces cerevisiae and Rqh1 of Schizosaccharomyces pombe and provide an outlook on some of the outstanding questions in the field.

1. The RecQ Helicase Family Is Conserved from Bacteria to Humans

To maintain genome integrity, RecQ-like DNA helicases act as key factors in homologous recombination (HR) and recombinational DNA repair and, in some organisms, perform accessory roles in DNA replication and transcription. The RecQ helicase family owes its name to its discovery in Escherichia coli where a new mutation, recQ1, was identified while screening for mutations conferring resistance to thymineless death [1,2]. Thymineless death is the phenomenon of cell death in prokaryotes and eukaryotes that occurs upon depletion of the DNA base thymine. The exact mechanism is still under investigation, but it requires a RecQ-dependent pathway of homologous recombination that also includes RecJ, a RecA-dependent pathway that involves the SOS-response, and possibly a yet unidentified third pathway [3,4,5]. Disruption of RecQ significantly reduces recombination frequency and increases sensitivity to ultraviolet (UV) radiation of cells in which recombination events are prevented from proceeding through RecBCD and SbcBC pathways, identifying RecQ as a member of the alternate RecF recombination pathway [1,6]. In wildtype cells, this RecF pathway appears to mediate genetic exchange at single-strand (ss) DNA gaps and contributes to the repair of collapsed replication forks. Since then, RecQ helicases have been identified in all organisms, from bacteria to plants and humans. Unicellular organisms typically express a single RecQ-like DNA helicase, such as RecQ in E. coli, Sgs1 in S. cerevisiae and Rqh1 in S. pombe, although evidence of additional RecQ helicases in yeasts has recently emerged [7]. In contrast, multiple RecQ-like helicases are present in multicellular eukaryotes, with seven in Arabidopsis thaliana currently holding the top spot [8]. The human genome encodes five RecQ-like DNA helicases (RecQL1 to RecQL5), three of which are associated with genetic disorders. Inactivation of RecQL2 (WRN), RecQL3 (BLM), and RecQL4 leads to Werner’s syndrome, Bloom syndrome and Rothmund–Thomson syndrome, respectively; these disorders are characterized by genome instability, increased cancer risk and, in the case of Werner syndrome, adult-onset premature aging [9,10,11,12]. Additional symptoms of Bloom syndrome include short stature and increased risk for Type-2 diabetes, immunodeficiency, infertility, and sun-sensitivity [13].
Based on mutant phenotypes, protein structure and protein–protein interactions, S. cerevisiae Sgs1 is most closely related to human BLM [9,14]. Loss of Sgs1 results in hypersensitivity to DNA damaging agents such as methylmethanesulfonate (MMS) and hydroxyurea (HU) [15,16], increased gross chromosomal rearrangements (GCRs) [17], reduced replicative lifespan [18], increased rate of mitotic recombination [19] and frequent chromosome missegregation [20], highlighting the importance of Sgs1 for maintaining yeast genome integrity. Similarly, rqh1 mutants of S. pombe are hypersensitive to HU and UV, display chromosome missegregation, elevated recombination and are defective in recovery from S phase arrest [21,22,23]. This review highlights the ever-expanding cellular functions of RecQ helicases in yeast.

2. Domain Structure of the Sgs1 and Rqh1 Helicases

The RecQ helicase family belongs to the superfamily 2 (SF2) helicases. They are motor proteins with 3′-5′ DNA helicase activity and unwind DNA in an ATP-dependent manner, requiring Mg2+ as a cofactor [24]. All RecQ-like DNA helicases possess a helicase domain of nearly 400 amino acids (Figure 1) containing the seven conserved ATPase/helicase motifs (I, Ia, II, III, IV, V, VI) and a characteristic ‘motif 0′ upstream of motif I [21,25,26,27]. Motif 0 is involved in ATP binding in a manner that is similar to that of the Q motif of DEAD-box helicases, which has led to the suggestion of a possible evolutionary connection between the DNA and RNA helicases of the RecQ and DEAD-box families, respectively [28,29]. These conserved motifs form the surface of the catalytic cleft between the two RecA-like lobes where ATP is hydrolyzed in a ssDNA-dependent manner [30,31]. Human BLM possesses a unique proline/lysine-rich loop that protrudes from the surface of the second RecA-like lobe. Although its precise function is unknown, its functional significance is supported by a mutation that partially inactivates BLM, causing increased sister-chromatid exchanges and a slow double-strand break repair phenotype [32,33]. With an allele frequency of 5.3%, this mutation (P868L) is not associated with Bloom syndrome, but may be a candidate for a new cancer risk allele in otherwise healthy individuals [33].
Helicase-and-RNaseD-like-C-terminal (HRDC) and RecQ C-terminal (RQC) domains are present in bacterial RecQ, including RecQ helicases with multiple HRDC domains [34], the yeast RecQ helicases Sgs1 and Rqh1, and in most RecQ helicases of multicellular organisms (Figure 1). The RQC domain consists of a winged-helix (WH) domain and a zinc-binding domain; the latter has been implicated in structural stability of the protein [35,36] whereas the WH domain acts as a DNA binding motif in many proteins [37,38,39,40]. Structural and biochemical analyses for E. coli RecQ and WRN indicate that the WH domain can interact with dsDNA [30,41,42]. The HRDC domain is dispensable for ATPase activity and unwinding of simple double-stranded DNA substrates, but contributes to DNA binding and DNA substrate specificity which, for example, is required for double-Holliday-junction dissolution by BLM [43]. Other structural analyses of BLM and Sgs1 have detected low or no DNA-binding affinity of the HRDC domain, but identified an intramolecular interaction between the BLM HRDC domain and the ATPase domain that may affect helicase activity [31,44]. The HRDC domain in RecQ helicases has not been directly implicated in mediating protein–protein interactions, but it is part of a fragment of human BLM that binds the telomere-associated protein Trf2 [45].
In addition to conserved, structured domains that make up the central helicase core comprised of the ATPase and RQC domains, most eukaryotic RecQ helicases possess long, disordered N-terminal tails. With the exception of WRN, whose N-terminal tail harbors a 3′-5′ exonuclease domain [46,47,48], these disordered tails are devoid of catalytic activity and act as binding sites for the often-large number of interacting proteins (Figure 2). The N-terminus of Sgs1 interacts with at least six proteins, including type I topoisomerase Top3 and Rmi1, forming the STR complex for double Holliday junction dissolution, replication protein A (Rpa) 70, antirecombinase Srs2, global nucleotide excision repair factor Rad16, and the type II topoisomerase Top2. Additional protein interactions for which binding sites on Sgs1 have not yet been narrowed down to specific domains include the DNA double-strand-break (DSB) recognition factor Mre11, the single-strand specific nuclease Dna2, mismatch repair proteins Mlh1/3 and Msh2/6, the SUMO-conjugating enzyme Ubc9 and Smc5 of the Smc5/6 complex, which is a key factor for recombinational DNA repair [49,50,51,52,53,54,55,56]. Although the vast majority of protein interactions occur with the N-terminal tail of Sgs1, the interaction of Sgs1 with the strand exchange protein Rad51 occurs immediately downstream of the RQC domain, with Phe 1192 playing a crucial role in promoting homologous recombination [57].
Sgs1 possesses additional specialized motifs, of which some are conserved in BLM but not in other eukaryotic RecQ helicases. Acidic regions in the disordered N-terminal tail of Sgs1 mediate an interaction with Rpa70 [58]. RPA prevents reannealing of ssDNA and stimulates Sgs1 activity in vitro, and this functional relationship is conserved in human BLM [59,60,61]. It has also been proposed that the acidic regions contribute to the role of Sgs1 in promoting homologous recombination by acting as a DNA mimic that competes with ssDNA for RPA binding, thereby facilitating the initial loading of Sgs1-bound Rad51 onto 3′ ssDNA overhangs of resected DSBs and promoting Rad51 filament initiation and HR [57]. Such a role of the Sgs1/Dna2 resection machinery would be reminiscent of the role of RecBCD in promoting RecA loading in E. coli [62]. Finally, residues 103-322 of Sgs1 display ssDNA binding as well as strand annealing and exchange activities that enable Sgs1 to anneal complementary single DNA strands in vitro [7,63,64]. This activity is conserved in human BLM [65].

3. Post-Translational Modification of Sgs1 and Rqh1

Post-translational modifications regulate interactions with other proteins, cellular localization of proteins, and functional activation [66]. Compared to its human counterparts, Sgs1 and Rqh1 undergo relatively modest post-translational modification. Several [S/T]Q sites in the acidic region between residues 400-600 of Sgs1 are phosphorylated in a Mec1-dependent manner during exponential growth, with T451 phosphorylation being important for Rad53 binding in vitro. Sgs1 was also shown to be phosphorylated at positions S256 and S259 in a Rad53-independent manner under MMS-induced replication stress [67]; however, the functional significance of this event is currently unknown. Several other phosphorylation sites whose effects on Sgs1 function are poorly understood are indicated in Figure 2 [58,68,69,70,71,72]. SUMOylation of lysine 621 is required for the role of Sgs1 in telomere-telomere recombination [70,73,74]. Similarly, SUMOylation of Rqh1 stimulates its activity at telomeres, but also promotes breakage and hyper-recombination of damaged telomeres [75]. Quantitative phosphoproteomics has identified several Rqh1 phosphorylation sites that are currently of unknown functional significance, such as S104, S106, S395, and S1138 [76,77]. Figure 3 presents the domain structure and post-translational modifications of Rqh1.

4. Substrate Preferences of Sgs1 and Rqh1

Sgs1 acts on a wide range of DNA substrates (Figure 4). Sgs1 exhibits affinity for single and double-stranded DNA; however, its affinity for forked structures is markedly higher [73]. Notably, Sgs1 unwinds DNA-RNA duplexes with the same efficiency as DNA-DNA duplexes—an uncommon property for DNA helicases [73]. Sgs1 is able to discriminate between 3′ and 5′ overhangs and unwinds three way junctions and Holliday-junction-like four-way junctions—a key intermediate of homologous recombination [79]. Setting it apart from most RecQ helicases, Sgs1 is able to unwind G-G paired DNA, such as in G-quadruplex (G4) structures, more efficiently than duplex DNA. G4 structures can form in G-rich DNA sequences, which are particularly found in rDNA and telomeric repeats, thereby implicating Sgs1 as a protector of rDNA stability and telomeres [80,81]. The RQC domain of human BLM is a conserved G4-DNA binding domain, but such a determination remains to be made for Sgs1 or Rqh1 [82].

5. Physical Interactions of Sgs1 and Rqh1

Sgs1 interacts with several proteins involved in DNA recombination and the replication stress response (Figure 2). Mutations in SGS1 rescue the severe fitness defect and genomic instability phenotypes of top3 mutants [19]. Similarly, disruption of the rqh1 gene in S. pombe rescues the inviability of top3 mutants [83]. The first evidence of a physical interaction between Sgs1 and Top3 came from a yeast two-hybrid screen, followed by additional studies that mapped the Top3 binding domain to the N-terminal 282 amino acid residues of Sgs1 [19,70,84,85]. Similarly, the Top3-interacting domain in Rqh1 lies within the first 322 residues [78]. Along with Rmi1, Sgs1/Rqh1 and Top3 function as a single heteromeric complex. Rmi1 and Top3 interact with each other and can only bind to Sgs1 as a complex [86,87]. By binding to the topoisomerase gate, Rmi1 is thought to enhance the decatenase activity of Top3 [88,89]. Nuclear magnetic resonance spectroscopy showed that the first 125 N-terminal residues of Sgs1 are intrinsically disordered and contain two transient α-helices comprised of residues 25-38 and residues 88-97 [87]. Disrupting the first α-helix with single proline mutations (e.g., sgs1-F30P) impairs Sgs1 binding to Top3-Rmi1 and results in hypersensitivity to DNA damaging agents and accumulation of gross chromosomal rearrangements [87].
The Rad51 binding domain of Sgs1 was initially mapped to the region C-terminal of the helicase core and one critical region later narrowed down to residues 1187-1207 [57,90]. Mutation of F1192 to asparagine was sufficient to disrupt the Sgs1-Rad51 interaction in vitro [57]. Distinct positive and negative genetic interactions of this sgs1-FD allele with mutations in HR genes, such as mre11Δ, sae2Δ, top3Δ, exo1Δ and srs2Δ, revealed that the Sgs1-Rad51 interaction promotes HR, possibly by facilitating Rad51 filament initiation [57].
An RPA-binding domain has been mapped to amino acid residues 404-560 of Sgs1 just upstream of the helicase domain, overlapping an acidic region and the Rad53-interacting domain between residues 446-456 [58]. Mec1-mediated phosphorylation of [S/T]Q residues in this acidic region promotes Sgs1 binding to Rad53, thereby recruiting the kinase to stalled forks and promoting the replication checkpoint response [58].
Co-immunoprecipitation experiments have also identified a physical interaction of Sgs1 with the Dna2 nuclease, which functions with Sgs1 and the single-strand binding protein RPA in the long-range resection of DSBs [50]. During DSB repair, Sgs1 also interacts with the antirecombinase Srs2 at a region that spans residues 422-722 and with the Mre11 subunit of the DSB recognition complex MRX [91]. All three proteins—Sgs1, Srs2 and Mre11—form a large complex under normal as well as DNA-damaging conditions [49], which likely functions in the recruitment of these HR proteins to DSBs.
The molecular basis and functional significance of physical interactions of Sgs1 with several other proteins is less well-understood. Rad16, a protein involved in lesion recognition during global nucleotide excision repair (NER), physically interacts with the 421–792 amino acid region of Sgs1, possibly implicating Sgs1 in DNA damage recognition [92]. The Rad16-binding region overlaps with the Top2-interacting region, which spans residues 466-746. Top2 is a type-II topoisomerase and a mitotic post-replication decatenase that acts in the same pathway as Sgs1 to prevent chromosome missegregation [20,85,93]. Sgs1 also physically interacts with the mismatch-repair factors Msh2/6 and Mlh1 and this interaction is conserved in human BLM; however, the biological significance of this interaction in yeast and human cells remains to be elucidated [51,56,94,95,96]. That BLM-deficient cells are not defective in DNA mismatch repair and that in late prophase of meiosis I Sgs1 interacts with Mlh3, which forms a heteroduplex with Mlh1 and promotes meiotic crossing-over, may suggest that the interaction of Sgs1/BLM with mismatch repair proteins functions in genetic recombination rather than in DNA replication-coupled mismatch repair [52]. In a yeast two-hybrid assay, Sgs1 residues 420-791 were found to interact with the jumonji domain of Gis1, a transcription factor that has two jumonji subdomains, one of which causes histone demethylation, thereby facilitating transcription [97]. However, whether a link between DNA recombination and histone demethylation exists in yeast still remains to be seen. Using two-hybrid assays, interactions of Sgs1 with Ubc9 and SUMO were also identified, raising the possibility that Sgs1 may also be SUMOylated by Ubc9 [55]. Table 1; Table 2 summarize binding partners of the yeast RecQ helicases Sgs1 and Rqh1.

6. Homologous Recombination-Mediated DNA Double-Strand Break Repair

6.1. Genetic Interactions of SGS1 and RQH1 with HR Genes

DSBs can be repaired by either nonhomologous end joining (NHEJ), which directly ligates the broken ends and is preferred during G1 phase of the cell cycle, or homologous recombination (HR), which utilizes a sister chromatid or a homologous chromosome for restoring genetic information during the S/G2/M phases or meiosis [113]. Studies that first highlighted the roles of Sgs1 and Rqh1 in HR showed that sgs1Δ and rqh1Δ mutants exhibit mitotic hyperrecombination phenotypes [19,23,114]. For DNA-damage sensitivity and growth, homologous recombination genes RAD55, RAD57, RAD54 and RAD51 are epistatic to SGS1 whereas RAD52 is mostly independent of SGS1 [90,115]. However, a genetic background where SGS1 was in a different pathway for the repair of HU-induced DNA damage than RAD51 and RAD54 has also been reported [116]. Accumulation of gross-chromosomal rearrangements, while moderate in the sgs1Δ mutant, are reduced by RAD51 deletion to the rate of the rad51Δ mutant, in support of the function of RAD51 and SGS1 in the same pathway [57].
Deletion of SGS1 suppresses the slow-growth and hyper-recombination phenotypes of the top3Δ mutant, revealing that sgs1Δ is epistatic to top3Δ and suggesting that Top3 resolves a deleterious substrate produced by Sgs1 [19]. Relevant to this model is Pif1, a DNA helicase with opposite polarity of Sgs1 that functions in both nuclear and mitochondrial DNA metabolism [117]. That Pif1 overexpression suppresses top3Δ defects but aggravates sgs1Δ and sgs1Δ top3Δ defects, and that Pif1 localizes to S-phase specific DNA repair foci, suggests that Pif1 may prevent or repair deleterious recombination intermediates that are presumably formed when Sgs1 activity becomes uncoupled from Top3 [118]. Although an SGS1 deletion suppresses the slow growth of the top3Δ mutant, it produces a growth defect in the top1Δ mutant [19,119], which has led to the suggestion that Sgs1-Top3 competes with another pathway involving Top1 and Srs2 for regulating initial recombination events [85]. Mutations in HR repair genes RAD51, RAD55 and RAD57 in S. cerevisiae and rhp51 and rhp57 in S. pombe suppress the synthetic lethality and severe slow growth phenotypes of sgs1Δ srs2Δ and Δrqh1 Δsrs2 mutants, respectively, providing further evidence for the involvement of Sgs1 and Rqh1 in later steps of homologous recombination [120,121].
SGS1 strongly interacts with EXO1. The sgs1Δ exo1Δ mutant exhibits a pronounced fitness defect and accumulates gross-chromosomal rearrangements at one of the highest rates observed to date, exceeding that of the single mutants by 200-fold [122,123,124]. The presence of DSBs in the sgs1Δ exo1Δ mutant that are only minimally trimmed by the MRX-Sae2 complex indicates that the molecular basis for the strongly negative genetic interaction between SGS1 and EXO1 is the disruption of long-range resection of DSBs, which provided the original evidence for redundant roles of Sgs1 and Exo1 in long-range DSB resection [122,125,126]. The dramatic decrease in viability of the sgs1Δ exo1Δ mutant upon deletion of HR genes RAD59 or RAD52, but not RAD51, indicates that a Rad59/Rad52-mediated HR pathway is needed for proper repair of the incompletely processed DSBs in the sgs1Δ exo1Δ mutant [57].
The identification of a role for Sgs1 in homeologous recombination prompted an investigation into genetic interactions of Sgs1 with mismatch repair (MMR) genes, which regulate recombination between nonidentical (homeologous) DNA sequences [17,127]. Indeed, the sgs1Δ msh2Δ double mutant displayed significantly higher rates of gross-chromosomal rearrangements and recombination between homeologous sequences than either single mutant [17,127]. However, while Sgs1 suppressed recombination events between homeologous regions in the divergent genes CAN1, LYP1 and ALP1, the mismatch repair factor Msh6 did not, suggesting that Sgs1 does not suppresses homeologous recombination through a function in the mismatch repair pathway, but plays a MMR-independent role in suppressing translocations between these divergent sequences [128].
Deletion of SGS1 is lethal in combination with numerous mutations (Figure 5), most commonly those that affect recombinational DNA repair, such as SRS2, MUS81/MMS4, SAE2, MRE11/RAD50/XRS2 and RRM3 [49,116,120,129,130,131,132]. The synthetic lethality of the sgs1Δ sae2Δ mutant further underscores the role of Sgs1 in DSB end resection; its suppression by YKU70 deletion suggests that the DSB end-binding protein Ku70 blocks DSB end-resection by the exonuclease Exo1 [133]. Inviability of the sgs1Δ srs2Δ and sgs1Δ mus81Δ mutants and the severe growth defect of the sgs1Δ rrm3Δ mutant are suppressed by deleting RAD51, suggesting that, similar to Srs2, Rrm3 and Mus81/Mms4 prevent the formation of recombination intermediates that require the Sgs1/Top3/Rmi1 complex for dissolution [130,132,134,135]. Similarly, combining the rqh1Δ mutation with mre11Δ or mus81Δ mutations is synthetically lethal in S. pombe [121,136]. Table 3 provides a comparison of the cellular functions of Sgs1 and Rqh1 indicated by the phenotypes and genetic interactions of their deletion mutants.

6.2. Roles of Sgs1 in DSB Repair

Sgs1 participates in both early and late steps of homologous recombination (Figure 6). Following formation of a DSB, phosphorylation of Sae2 by cyclin-dependent kinase Cdk1 stimulates DNA end resection at the 5′ end by the Mre11/Rad50/Xrs2 (MRX) nuclease complex, producing short 3′ ssDNA overhangs [126]. This initial processing is crucial since it directs the DNA break to HR and inhibits NHEJ, but it is not sufficient for the formation of the recombinogenic Rad51 filament for error-free repair. In fact, when resection stops at this step (in the sgs1Δ exo1Δ mutant), repair in the presence of Rad51 is mutagenic [57]. Thus, the trimmed DSB ends produced by MRX/Sae2 subsequently undergo long-range resection either by the 5′-3′ exonuclease Exo1 or the concerted action of the DNA helicase/nuclease Sgs1/Dna2 [125,126,142,143]. Although Sgs1 and Exo1 are equally capable of fully resecting DSB ends, Sgs1 possesses intrinsic nucleosome repositioning activity, implying that the Sgs1-mediated resection machinery would be less prone to encountering nucleosome impediments than Exo1 [144,145]. Notably, in vitro, resection activity of Exo1, but not Sgs1-Dna2, could be stimulated by replacement of H2A with H2A.Z [145]. Together with the finding that H2A.Z is incorporated into chromatin in an Swr1-depedent manner near DSB sites [146,147] and genetic evidence from the sgs1 swr1 mutant, this suggests that the less stable H2A.Z allows efficient chromatin resection by Exo1 specifically at DSBs [145]. Thus, Exo1 nuclease activity appears to be limited to DSBs and to the replication fork where its activity in mismatch repair is not likely to be impeded by nucleosomes. In contrast, Sgs1 appears to resect efficiently through any nucleosomes and may therefore depend on its numerous protein–protein interactions to target its activity carefully throughout the genome.
Sgs1 can be recruited to dsDNA ends either in its inactive ATP-free state or via the Top3/Rmi1 complex, which promotes ATP-dependent recruitment of Sgs1 to DNA ends; however, Sgs1 fails to initiate translocation in both scenarios [144]. Only addition of Dna2 activates long-range translocation activity of Sgs1, both independently as well as in cooperation with Top3-Rmi1. An in vitro study identified Sgs1, Dna2 and RPA as the minimal resection complex where Sgs1, stimulated by RPA, unwinds DNA, producing the substrate for Dna2 to degrade [50]. Another in vitro study showed that RPA, in the presence of ATP, only stimulates short-range helicase activity of Sgs1 [144], suggesting the need for additional factors.
The resected 3′ ssDNA tail is immediately bound by RPA, which subsequently binds Rad52 [151]. Binding of Rad52 to RPA facilitates RPA displacement by Rad51, which assembles as a nucleoprotein filament and facilitates strand invasion into a homologous duplex [151]. The binding site of Rad51 maps to a region immediately downstream of the RQC domain of Sgs1 and the sgs1-F1192D (FD) mutant fails to bind Rad51 [57]. Unlike the SGS1 deletion, the sgs1-FD allele does not affect the viability of the srs2Δ mutant and suppresses its hypersensitivity to HU and MMS, suggesting that sgs1-FD is a hypo-rec allele of SGS1. Additional positive and negative genetic interactions of the sgs1-FD allele that are also distinct from those of the SGS1 deletion suggest that the Sgs1/Rad51 interaction promotes homologous recombination [57]. On the other hand, a recent in vitro study found that similar to Srs2, Sgs1 can dismantle Rad51 filaments, albeit with some mechanistic differences. Sgs1 was able to translocate on RPA-coated ssDNA without dislodging RPA, but stripped off Rad51 [152]. In contrast to Srs2, which strips off RPA and Rad51 by stimulating Rad51 ATP hydrolysis activity, there was no difference in the ability of Sgs1 to dislodge ATPase-deficient or ATPase-proficient Rad51 from ssDNA [152,153]. This Sgs1 activity may be used to prevent the formation of deleterious HR intermediates at stalled replication forks, to disrupt heteroduplexes, or to displace the invading Rad51 filament in D-loops, thereby channeling HR intermediates into the synthesis-dependent strand annealing (SDSA) pathway of DSB repair that produces noncrossover products [152,154,155,156].
If the second end of the processed DSB is also captured to form a double Holliday junction, Sgs1, in a complex with Top3/Rmi1, branch migrates the Holliday junctions to form a hemicatenane structure that is dissolved by the strand passage activity of Top3 [156]. Stimulated by Rmi1, this step produces exclusively non-crossover products [156,157]. In the absence of Sgs1/Top3/Rmi1 complex activity, double Holliday junctions are resolved by endonucleases, which generates both crossovers and noncrossovers.
Recent evidence suggests that the role of Sgs1/Top3/Rmi1 in HR is regulated by the structural maintenance of chromosomes (SMC) complex Smc5/6, which is known to function in DNA repair and replication and whose functions are mediated by the interacting SUMO ligase Mms21 [158]. Sgs1 physically interacts with Smc5/6 and it has been proposed that Smc5/6 binding to DNA damage sites causes its SUMOylation via Mms21, followed by Sgs1 recruitment to Smc5/6 through its SUMO interacting motifs and later Smc5/6-dependent SUMOylation of Sgs1 and Top3 to activate recombination [53,159]. Indeed, disrupting SUMOylation or Sgs1 interaction with Smc5/6 elevates joint molecule accumulation, supporting a role for the Sgs1-Smc5/6 interaction in HR, and reduces Sgs1/Top3/Rmi1 foci formation in the presence of DNA damage [106].
Although mechanistically still unclear, Rqh1 has also been proposed to be involved in early and late steps of HR. Partial suppression of HU and UV sensitivities of rqh1Δ cells by expression of E. coli Holliday junction resolvase RusA implicate Rhq1 in preventing Holliday junction accumulation [160]. However, in vitro studies showing that Rqh1 can process Holliday junctions like Sgs1 and BLM remain to be performed. Loss of rhp55/rhp57 (Rad55/57 homologs) was able to suppress the HU sensitivity of rqh1Δ mutants, which was dependent on rhp51, suggesting a role of Rqh1 downstream of Rhp55/57 [161]. In support of an early role of Rqh1 in HR, it was shown that Rqh1 foci formed earlier than Rhp51 (Rad51 homolog) foci upon exposure to UV [141]. However, S. cerevisiae and S. pombe differ in the contribution of the RecQ helicase to DSB resection. While budding yeast Exo1 and Sgs1 can independently perform sufficient long-range resection, Exo1 is responsible for the majority of extended resection in S. pombe and Rqh1-Dna2 makes only a minor contribution to overall DSB resection [162]. The strong resection defect observed in exo1Δ mutants is partially rescued by loss of Pxd1, which inhibits the nuclease activity of Dna2 [163]. Pxd1 was identified as a regulator of DSB repair by the single-strand annealing (SSA) pathway; it physically interacts not only with Dna2, but also with Rad16, thereby regulating the activity of the two nucleases and improving SSA outcomes [163]. On the other hand, Exo1 and Rqh1-Dna2 play redundant roles in resection of uncapped telomeres arising from deletion of pot1 [164].

6.3. Roles of Sgs1 at the Replication Fork

Although the molecular mechanisms of Sgs1 function at replication forks are still unclear, several studies have linked Sgs1 to the replisome, suggesting roles in sensing DNA damage and overcoming barriers to replication fork progression. The peak of Sgs1 levels in S-phase, colocalization of Sgs1 with the Orc2 subunit of the origin recognition complex, and physical interaction with replisome components, such as Top1, Top2, RPA and DNA polymerase ε support placement of Sgs1 at unperturbed replication forks [19,20,58,165,166]. Chromatin-immunoprecipitation and immunofluorescence microscopy have also placed Sgs1 at sites of de novo DNA synthesis, although it remains to be clarified if Sgs1 directly influences normal DNA replication or if Sgs1 tags along so it is present in the event of fork stalling or collapse [166]. Better established is Sgs1′s importance for the response to replication stress. Sgs1 acts in the intra S-phase checkpoint, which stabilizes the replisome in the event of stalling and halts cell cycle progression until DNA is repaired [16,165,167]. Sgs1 acts upstream of the DNA damage checkpoint kinase Rad53 and colocalizes with it in S-phase specific foci [165]. Sgs1 may also help to prevent DNA breaks at stalled forks by using its ATPase/helicase activity to reverse chicken-foot structures that can form if progression of the replisome is blocked by a DNA lesion on the leading strand [168]. The branch migration activity of Sgs1 could help convert regressed forks into normal fork structures, thereby facilitating replication restart.
sgs1 mutants display faster progression through S-phase, suggesting defective checkpoint signaling, but tend to accumulate in G2/M phase, likely resulting from unrepaired S-phase DNA damage [165,169]. A similar effect is observed in MMS-treated rqh1Δ cells [170] and even though these cells arrest DNA replication in response to HU normally, they exhibit defective chromosome segregation [21]. RusA, an E. coli resolvase, is able to suppress the hypersensitivity of rqh1Δ mutants to fork-stalling agents such as MMS and HU, but not to CPT and γ-rays that cause fork collapse, suggesting different roles for Rqh1 in rescue of stalled and collapsed forks [139,160]. That cells expressing helicase-defective Sgs1 are significantly less sensitive to CPT than sgs1Δ cells suggests that the role of Sgs1 in the repair of collapsed forks does not depend on its helicase activity but may involve recruitment of Top3 or other repair factors. Similarly, the role of Rqh1 in the repair of collapsed forks may involve activation of DNA repair factors, such as Top3, which can modulate DNA topology for recombinational repair [139,171,172].
Holliday junction formation has been demonstrated at stalled forks in checkpoint-deficient cells [173] and there is evidence that Sgs1-Top3 activity is involved in resolving these structures. The X-shaped structures at damaged replication forks in cells lacking Sgs1 or Top3 are also displayed by ubc9 and mms21 mutants [55]). Similar amounts of X-shaped structures in the double mutants indicate that Ubc9/Mms21 cooperates with Sgs1/Top3 to resolve these structures during replication, possibly via Ubc9-mediated SUMOylation of Sgs1 [55,74]. A functional overlap in resolving stalled replication forks has also been observed for Sgs1-Top3 and the structure-specific endonuclease Mus81-Mms4 [131].
In DNA-damage checkpoint activation, Sgs1 works parallel to Rad9, an adaptor for activation of the central effector kinase Rad53 in response to DNA-damage [174]. Genetic studies of intra-S-phase checkpoint activation show that Sgs1 acts in parallel to Rad24, which loads the PCNA-like Mec3/Rad17/Ddc1 DNA-damage sensor onto ssDNA, and is required for full Rad53 activation [58,165]. Further, Sgs1 functionally interacts with Dia2 and Mph1 in the restart of replication forks stalled at MMS-induced DNA lesions [98]. In HR, the Sgs1 and Mph1 helicases act through separate mechanisms in the suppression of crossover products [157,175,176,177,178]. Similarly, during replication, Sgs1/Top3/Rmi1 is thought to dissolve recombination intermediates that arise from sister-chromatid recombination at broken replication forks [49,179]. Sgs1 also acts on damaged replication intermediates to provide substrates for fork rescue mechanisms, such as Mph1- and Mus81/Mms4-mediated pathways [180]. The physical interaction of Sgs1 with Dia2, an F-box protein component of the SCF E3 ubiquitin ligase complex, appears to recruit Sgs1 to stalled forks by mediating degradation of the fork protection complex component Mrc1 and is needed for Rad53 activation in response to replication stress [98]. These insights into roles of Sgs1 at damaged forks, although still limited, are beginning to provide some mechanistic explanations for the impaired recovery of the sgs1Δ mutant from prolonged fork arrest [166].
Synthetic lethality between sgs1Δ and slx4Δ mutations suggests that the Slx1-Slx4 complex functions redundantly with Sgs1/Top3/Rmi1 on stalled forks, cleaving and decatenating them, respectively [181,182]. In addition to its role in dissolving recombination intermediates during replication stress, there is evidence that Sgs1 contributes to the stabilization of polymerases α/primase and ε at stalled forks [58,166]. pol2 mutants are defective in the activation of the intra S-phase checkpoint, and this defect is epistatic with sgs1, suggesting that Sgs1 and DNA pol ε function in the same pathway, in parallel to Rad24/Mec3/Rad17/Ddc1, to activate Rad53 in presence of HU [165]. Deletion of SGS1 in a checkpoint deficient mec1-100 mutant results in complete dissociation of replicative DNA polymerases, impaired recovery from replication fork arrest and a synergistic increase in the GCR rate, providing evidence that Sgs1 and Mec1 independently contribute to genome stability [183]. The suppression by mms2 and ubc13 mutations of the severe growth defects of sgs1Δ mutants lacking the DNA polymerase subunit Pol32 suggests a role for Sgs1 in the error-free Rad6 DNA damage tolerance (DDT) pathway [184,185]. However, requirement of Sgs1 for efficient PCNA monoubiquitination during replication stress also suggests a role for Sgs1 in the error-prone translesion DNA synthesis (TLS) pathway [186]. Similar to sgs1Δ, rqh1 mutations reduce the growth rate of S. pombe lacking DNA polymerase subunits such as those of DNA pol δ and ε [23].

7. Meiosis

Due to its role in homologous recombination Sgs1 was also expected to be required for successful completion of meiosis. Indeed, deletion of SGS1 reduces meiotic product (tetrad) formation and spore viability; however, unlike in mitotic cells, meiotic crossover formation is not markedly increased [20,187,188,189,190,191], suggesting differences in the roles of Sgs1 in mitotic and meiotic recombination. Notably, the helicase activity was found to be dispensable for Sgs1′s meiotic function, suggesting that Sgs1 serves structural or regulatory functions in interaction with other meiotic factors, such as recruitment of Top3 and Top2 or other factors that interact with the meiosis-essential domain of Sgs1 that spans residues 126–595 [188,189,192]. Sgs1 also colocalizes with Zip3, a component of the synapsis initiation complex, and absence of Sgs1 correlates with an increase in synapsis initiation complexes [187]. Since Zip3 interacts with HR proteins such as Rad51 and Mre11, and its absence delays synaptonemal complex formation [193], Sgs1 may help regulate Zip3 activity. Indeed, longer persistence of fully synapsed chromosomes, suggesting failure to timely exit the meiotic pachytene stage, is thought to be responsible for poor spore formation in the sgs1Δ mutant [187].
Moreover, sgs1Δ mutants exhibit elevated meiotic chromosome missegregation and most of these events are associated with chromosome nondisjunction rather than defects in sister chromatid separation [20]. SGS1/rqh1 deletion is synthetically lethal with mus81Δ and mms4Δ, raising the possibility of overlapping roles of Sgs1 and the structure-specific endonuclease Mus81-Mms4 in meiotic recombination [139,182]. Lethality was shown to result from unresolved joint molecules that formed during meiotic recombination, suggesting that Sgs1/Rqh1 and Mus81/Mms4 collaborate in mediating efficient joint molecule resolution to promote proper segregation of homologs [194,195,196,197].
The ZMM proteins (Zip1/Zip2/Zip3/Zip4, Msh4/Msh5 and Mer3) function in synaptonemal complex assembly and recombination [198]. Synapsis and crossover formation are impaired in zmm mutants, but are restored by deletion of SGS1 [190]. This suggests that by suppressing the anti-crossover activity of Sgs1, ZMM proteins, also known as the synapsis initiation complex (SIC), promote meiotic chromosome synapsis and crossovers [190]. Rqh1, on the other hand, promotes meiotic recombination in S. pombe, which lacks ZMM proteins and crossover interference. Rqh1 may act on fission-yeast-specific meiotic recombination intermediates, such as single Holliday junctions, rather than the double Holliday junctions predominantly found in budding yeast meiosis [197,199,200].

8. Possible Functions in Excision Repair

Nucleotide excision repair (NER) primarily repairs DNA lesions that distort the DNA helix, such as UV-induced pyrimidine dimers or bulky DNA adducts, whereas base excision repair (BER) removes bases damaged primarily by oxidation, alkylation or deamination [201,202,203]. Although human NER and BER proteins interact with RecQ helicases, including BLM, no such evidence exists in yeast [204,205,206,207,208]. sgs1 mutants are only marginally sensitive to UV radiation, but one study suggests that Sgs1 and the NER protein Rad16 may work in a common pathway to repair UV-induced damage [92]. The authors showed that a RAD16 deletion partially rescues the hypersensitivity of the sgs1Δ mutant to MMS and H2O2, suggesting that Rad16 might direct formation of a substrate that requires dissolution by Sgs1.
A contribution of Sgs1 to BER is less likely. Sgs1 functionally interacts with the 5′ flap-endonuclease Rad27, which besides its well-understood function in Okazaki fragment maturation also plays a role in long-patch BER [202,209]. Sgs1 also functionally interacts with the ROS-scavenging enzyme Tsa1, which suppresses genome rearrangements and is required for normal growth in the absence of Sgs1 [210]. Studies are needed to better delineate possible contributions of Sgs1 to excision repair.

9. Telomere Length Maintenance

Several studies suggest a role for RecQ helicases in telomere maintenance, most prominently a role of human WRN whose absence causes telomere defects and rapid, adult-onset aging in Werner syndrome [211,212,213]. Deletion of SGS1 does not lead to telomere length changes [214]. However, its exquisite ability to unwind G-G paired DNA, such as G-quadruplexes, in vitro extends to telomeric sequences and may be important in vivo for ensuring faithful chromosome segregation [80]. Sgs1 also contributes to telomeric end processing, where it might be needed for unwinding, especially when ends form G4 structures, to provide substrates for exonucleases to produce the 3′ G-strand overhangs [214]. Sgs1 has also been proposed to function in telomere lengthening via the alternative lengthening of telomeres (ALT) pathway, which relies on HR and is independent of telomerase [215,216]. Deletion of SGS1 in telomerase mutants est2Δ and tlc1Δ precludes formation of Type II survivors, which amplify TG1-3 -telomeric repeats [217,218,219]. It is thought that Sgs1 performs Holliday junction dissolution or facilitates replication of T-circles that are involved in elongation of telomeres via rolling-circle replication [140,219,220]. Consistently, sgs1Δ tlc1Δ and sgs1Δ est2Δ mutants display rapid telomere shortening compared to the single telomerase mutants, implying that Sgs1 reduces the rate of telomere shortening by homologous recombination [217,218]. Furthermore, sgs1Δ tlc1Δ mutants display reduced recombination frequency and accumulate X-shaped structures at telomeres in a Rad52-dependent manner, indicating a role of Sgs1 in resolving these recombination intermediates and, thus, preventing premature senescence of tlc1Δ mutants [221]. Whereas Sgs1 resolves recombination intermediates and prevents premature senescence of tlc1 mutants, the Mph1 helicase promotes premature senescence in tlc1 mutants by causing telomere uncapping and recombinogenic ssDNA accumulation [222], which may form recombination intermediates that would require Sgs1-Top3-Rmi1 for resolution. Notably, in human cells, a physical interaction between BLM-TOP3A-RMI and FANCM, the human Mph1 homolog, acts to restrict the ALT pathway [223]. It will be interesting to determine if a similar physical interaction also exists between Mph1 and Sgs1, and if it impacts telomere length maintenance in yeast. Finally, SUMOylation of Sgs1 at lysine 621 is required for its role in the survival of telomerase-deficient cells by ALT [70,73,74]. Similarly, S. pombe Rqh1 is SUMOylated and protects telomeres from breaking, loss and entanglement in cells lacking the telomeric maintenance protein Taz1 [75]. The protective role of Rqh1 at telomeres is also supported by the synthetic lethality of the pot1Δ and rqh1Δ mutations [224].

10. Aging and Transcription

The average and maximum life-span of yeast cells lacking Sgs1 helicase activity is reduced by nearly half, suggesting a significant role of Sgs1 in preventing premature cellular aging [18,171]. Increased genomic instability due to recombination defects in the absence of Sgs1 is one of the major causes of reduced lifespan, but there may be additional contributions from other causes [225]. For example, extrachromosomal rDNA circles (ERCs), which were proposed to cause premature aging in yeast, accumulate at an increased rate in sgs1 mutants [226]. However, another study found no significant difference in ERC accumulation between sgs1Δ and wildtype cells, calling into question the association between ERCs and reduced lifespan of sgs1Δ mutants [227]. Increased oxidative stress in sgs1 mutants and mitochondrial dysfunction due to mtDNA mutations could be other potential causes of sgs1Δ mutant’s aging [209,228]. Indeed, human RecQ4 has been shown to localize to mitochondria where it preserves mitochondrial DNA integrity; however, it remains to be determined if Sgs1 contributes to mitochondrial function in yeast [229].
Nucleolar defects arising from the absence of Sgs1 may also contribute to the shortened lifespan of the sgs1Δ mutant. Sir3, which silences transcription at telomeres, rDNA and mating loci, localizes to the nucleolus of sgs1Δ cells after only nine generations [18,230]. Sgs1 localizes to the nucleolus and sgs1Δ cells display nucleolar enlargement and premature nucleolar fragmentation [18]. Nucleolar localization implicates Sgs1 in rRNA synthesis where it may be involved in unwinding of G4-DNA, a common feature of rDNA [80]. Recently, Sgs1 was shown to physically interact with Rio1 to recruit it to rDNA, where Rio1 is involved in rDNA replication [101,102]. Interaction of Sgs1 with the transcription factor Gis1 raises the possibility of additional contributions of Sgs1 to transcription [97].

11. Other RecQ-Like DNA Helicases in Yeast

Both budding and fission yeast were believed to possess a single RecQ helicase, Sgs1 and Rqh1, respectively. However, bioinformatics analysis identified another putative yeast RecQ helicase, Hrq1, based on its similarity to human RecQL4 [231]. In contrast to Sgs1, the 3′-5′ helicase activity of Hrq1 is restricted to unwinding duplex DNA with 3′ overhangs. Hrq1 is also stimulated by a fork structure and possesses DNA strand annealing activity [232]. Deletion of SGS1 in an hrq1Δ mutant has additive effects on the mitotic recombination rate, accumulation of spontaneous mutations and growth, suggesting that Hrq1 functions independently of Sgs1 in DNA recombination and repair [233]. Deletion of fission yeast Hrq1 leads to genome instability and hypersensitivity to DNA damaging agents [234]. Like Sgs1, Hrq1 is important for telomerase-independent telomere maintenance [235]. The epistatic relationship between HRQ1 and RAD4 and the physical interaction between the two proteins in a yeast-two-hybrid assay suggest that Hrq1 may also have an NER-related function [236]. Proteomic screens have identified two ubiquitination (K366 and K872) and one phosphorylation (S17) site in Hrq1, but their functional significance is unknown [72,237]. Future studies aimed at identifying additional genetic interactions and Hrq1-binding partners will help to clarify the roles of the second yeast RecQ helicase family member in DNA metabolism.
Sequence homology searches revealed two additional RecQ-like proteins in fission yeast, Tlh1 and Tlh2, that show significant sequence homology with other known RecQ helicases between residues 1180 to 1820 [238]. They are telomere-linked DNA helicases that contribute to telomere maintenance in the absence of telomerase. Similar to budding yeast Sgs1, deletion of Candida albicans Sgs1 causes hypersensitivity to DNA-damaging agents [239]. In contrast, Candida glabrata sgs1Δ mutants display wildtype growth in the presence of DNA-damage-inducing agents, suggesting that CgSgs1 may not be involved in DSB repair [240].

12. Outlook

Data accumulated over the past 25 years decidedly point to a role for Sgs1 in maintaining genome stability via its roles in DNA recombination and at the replication fork. However, important questions remain to be answered. For example, does the physical interaction of Sgs1 with constitutive components of the replisome [165] imply a function of Sgs1 during unperturbed DNA replication or does Sgs1 simply move along with the replisome to be available if the fork is damaged? How is the interaction of Sgs1 with over a dozen proteins that have functions in several different DNA metabolic pathways regulated? Only a few of these interactions are understood at the molecular level (Top3-Rmi1, RPA, Rad51, Rad53) and a model has been proposed wherein the coordinated interaction of Sgs1 with Rad51 and RPA promotes homologous recombination [57]; however, binding sites for the vast majority of interacting proteins have only been narrowed down to several hundred residues, which has limited our insight into the functional significance of most Sgs1 interactions. Other questions include: How does Sgs1 suppress non-allelic, interchromosomal recombination events [122,241,242]? Does the increased presence of reactive oxygen species in the sgs1Δ mutant [209] point to a role of Sgs1 in maintaining mitochondrial health similar to human RecQL4? What is the helicase-independent role of Sgs1 in meiosis [188,189,192]? Here, yeasts provide unique genetic tools to elucidate fundamental mechanisms of meiotic recombination that may eventually help to better understand the cause of subfertility/infertility in Bloom syndrome. In addition to Sgs1 function, questions regarding the structure of Sgs1 also remain. The N-terminal tail of Sgs1 is one of the longest intrinsically disordered regions in the yeast proteome [87]. Does it simply serve as one of the largest protein binding hubs in yeast or does it have additional biochemical and regulatory functions or engage in intramolecular interactions? Is the ability of Sgs1 to unwind G-quadruplexes based on a similar G-specific pocket that allows E. coli RecQ to unfold G-G paired DNA [243]? Finding answers to these questions and many others that remain about RecQ helicases in yeast will shed light on fundamental mechanisms of eukaryotic DNA metabolism and continue to provide direction to investigations of human RecQ helicases, at least three of which are associated with incurable cancer predisposition syndromes.

Author Contributions

S.V.G. wrote the manuscript and K.H.S. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Research in the KHS laboratory is supported by National Institutes of Health grant R01GM081425.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nakayama, H.; Nakayama, K.; Nakayama, R.; Irino, N.; Nakayama, Y.; Hanawalt, P.C. Isolation and genetic characterization of a thymineless death-resistant mutant of Escherichia coli K12: Identification of a new mutation (recQ1) that blocks the RecF recombination pathway. Mol. Gen. Genet. MGG 1984, 195, 474–480. [Google Scholar] [CrossRef] [PubMed]
  2. Willis, N.; Rhind, N. Mus81, Rhp51 (Rad51), and Rqh1 form an epistatic pathway required for the S-phase DNA damage checkpoint. Mol. Biol. Cell 2009, 20, 819–833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Nakayama, H.; Nakayama, K.; Nakayama, R.; Nakayama, Y. Recombination-deficient mutations and thymineless death in Escherichia coli K12: Reciprocal effects of recBC and recF and indifference of recA mutations. Can. J. Microbiol. 1982, 28, 425–430. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmad, S.I.; Kirk, S.H.; Eisenstark, A. Thymine metabolism and thymineless death in prokaryotes and eukaryotes. Annu. Rev. Microbiol. 1998, 52, 591–625. [Google Scholar] [CrossRef]
  5. Fonville, N.C.; Bates, D.; Hastings, P.J.; Hanawalt, P.C.; Rosenberg, S.M. Role of RecA and the SOS response in thymineless death in Escherichia coli. PLoS Genet. 2010, 6, e1000865. [Google Scholar] [CrossRef] [Green Version]
  6. Nakayama, K.; Irino, N.; Nakayama, H. The recQ gene of Escherichia coli K12: Molecular cloning and isolation of insertion mutants. Mol. Gen. Genet. MGG 1985, 200, 266–271. [Google Scholar] [CrossRef]
  7. Bernstein, K.A.; Gangloff, S.; Rothstein, R. The RecQ DNA helicases in DNA repair. Annu. Rev. Genet. 2010, 44, 393–417. [Google Scholar] [CrossRef] [Green Version]
  8. Hartung, F.; Puchta, H. The RecQ gene family in plants. J. Plant Physiol. 2006, 163, 287–296. [Google Scholar] [CrossRef]
  9. Ellis, N.A.; Groden, J.; Ye, T.-Z.; Straughen, J.; Lennon, D.J.; Ciocci, S.; Proytcheva, M.; German, J. The Bloom’s syndrome gene product is homologous to RecQ helicases. Cell 1995, 83, 655–666. [Google Scholar] [CrossRef] [Green Version]
  10. Yu, C.-E.; Oshima, J.; Fu, Y.-H.; Wijsman, E.M.; Hisama, F.; Alisch, R.; Matthews, S.; Nakura, J.; Miki, T.; Ouais, S. Positional cloning of the Werner’s syndrome gene. Science 1996, 272, 258–262. [Google Scholar] [CrossRef] [Green Version]
  11. Kitao, S.; Lindor, N.M.; Shiratori, M.; Furuichi, Y.; Shimamoto, A. Rothmund–Thomson syndrome responsible gene, RECQL4: Genomic structure and products. Genomics 1999, 61, 268–276. [Google Scholar] [CrossRef]
  12. Kitao, S.; Shimamoto, A.; Goto, M.; Miller, R.W.; Smithson, W.A.; Lindor, N.M.; Furuichi, Y. Mutations in RECQL4 cause a subset of cases of Rothmund-Thomson syndrome. Nat. Genet. 1999, 22, 82. [Google Scholar] [CrossRef] [PubMed]
  13. Yankiwski, V.; Marciniak, R.A.; Guarente, L.; Neff, N.F. Nuclear structure in normal and Bloom syndrome cells. Proc. Natl. Acad. Sci. USA 2000, 97, 5214–5219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bachrati, C.Z.; Hickson, I.D. RecQ helicases: Suppressors of tumorigenesis and premature aging. Biochem. J. 2003, 374, 577–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kusano, K.; Berres, M.E.; Engels, W.R. Evolution of the RECQ family of helicases: A drosophila homolog, Dmblm, is similar to the human bloom syndrome gene. Genetics 1999, 151, 1027–1039. [Google Scholar]
  16. Yamagata, K.; Kato, J.-I.; Shimamoto, A.; Goto, M.; Furuichi, Y.; Ikeda, H. Bloom’s and Werner’s syndrome genes suppress hyperrecombination in yeast sgs1 mutant: Implication for genomic instability in human diseases. Proc. Natl. Acad. Sci. USA 1998, 95, 8733–8738. [Google Scholar] [CrossRef] [Green Version]
  17. Myung, K.; Datta, A.; Chen, C.; Kolodner, R.D. SGS1, the Saccharomyces cerevisiae homologue of BLM and WRN, suppresses genome instability and homeologous recombination. Nat. Genet. 2001, 27, 113. [Google Scholar] [CrossRef]
  18. Sinclair, D.A.; Mills, K.; Guarente, L. Accelerated aging and nucleolar fragmentation in yeast sgs1 mutants. Science 1997, 277, 1313–1316. [Google Scholar] [CrossRef]
  19. Gangloff, S.; McDonald, J.P.; Bendixen, C.; Arthur, L.; Rothstein, R. The yeast type I topoisomerase Top3 interacts with Sgs1, a DNA helicase homolog: A potential eukaryotic reverse gyrase. Mol. Cell. Biol. 1994, 14, 8391–8398. [Google Scholar] [CrossRef] [Green Version]
  20. Watt, P.M.; Louis, E.J.; Borts, R.H.; Hickson, I.D. Sgs1: A eukaryotic homolog of E. coil RecQ that interacts with topoisomerase II in vivo and is required for faithful chromosome segregation. Cell 1995, 81, 253–260. [Google Scholar] [CrossRef] [Green Version]
  21. Stewart, E.; Chapman, C.R.; Al-Khodairy, F.; Carr, A.M.; Enoch, T. rqh1+, a fission yeast gene related to the Bloom‘s and Werner’s syndrome genes, is required for reversible S phase arrest. EMBO J. 1997, 16, 2682–2692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Davey, S.; Han, C.S.; Ramer, S.A.; Klassen, J.C.; Jacobson, A.; Eisenberger, A.; Hopkins, K.M.; Lieberman, H.B.; Freyer, G.A. Fission yeast rad12+ regulates cell cycle checkpoint control and is homologous to the Bloom’s syndrome disease gene. Mol. Cell. Biol. 1998, 18, 2721–2728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Murray, J.M.; Lindsay, H.D.; Munday, C.A.; Carr, A.M. Role of Schizosaccharomyces pombe RecQ homolog, recombination, and checkpoint genes in UV damage tolerance. Mol. Cell. Biol. 1997, 17, 6868–6875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Larsen, N.B.; Hickson, I.D. RecQ helicases: Conserved guardians of genomic integrity. In DNA Helicases and DNA Motor Proteins; Springer: Berlin/Heidelberg, Germany, 2013; pp. 161–184. [Google Scholar]
  25. Bernstein, D.A.; Keck, J.L. Domain mapping of Escherichia coli RecQ defines the roles of conserved N-and C-terminal regions in the RecQ family. Nucleic Acids Res. 2003, 31, 2778–2785. [Google Scholar] [CrossRef]
  26. Sharma, S.; Doherty, K.M.; Brosh, R.M. Mechanisms of RecQ helicases in pathways of DNA metabolism and maintenance of genomic stability. Biochem. J. 2006, 398, 319–337. [Google Scholar] [CrossRef]
  27. Bansal, R.; Arya, V.; Sethy, R.; Rakesh, R.; Muthuswami, R. RecA-like domain 2 of DNA-dependent ATPase A domain, a SWI2/SNF2 protein, mediates conformational integrity and ATP hydrolysis. Biosci. Rep. 2018, 38, BSR20180568. [Google Scholar] [CrossRef] [Green Version]
  28. Tanner, N.K.; Cordin, O.; Banroques, J.; Doere, M.; Linder, P. The Q motif: A newly identified motif in DEAD box helicases may regulate ATP binding and hydrolysis. Mol. Cell 2003, 11, 127–138. [Google Scholar] [CrossRef]
  29. Killoran, M.P.; Keck, J.L. Sit down, relax and unwind: Structural insights into RecQ helicase mechanisms. Nucleic Acids Res. 2006, 34, 4098–4105. [Google Scholar] [CrossRef]
  30. Bernstein, D.A.; Zittel, M.C.; Keck, J.L. High-resolution structure of the E. coli RecQ helicase catalytic core. EMBO J. 2003, 22, 4910–4921. [Google Scholar] [CrossRef] [Green Version]
  31. Newman, J.A.; Savitsky, P.; Allerston, C.K.; Pike, A.C.W.; Pardon, E.; Steyaert, J.; Arrowsmith, C.H.; Von Delft, F.; Bountra, C.; Edwards, A.; et al. Crystal structure of the Bloom’s syndrome helicase indicates a role for the HRDC domain in conformational changes. Nucleic Acids Res 2013, 10, 5221–5535. [Google Scholar] [CrossRef]
  32. Mirzaei, H.; Schmidt, K.H. Non-Bloom syndrome-associated partial and total loss-of-function variants of BLM helicase. Proc. Natl. Acad. Sci. USA 2012, 109, 19357–19362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Shastri, V.M.; Schmidt, K.H. Cellular defects caused by hypomorphic variants of the Bloom syndrome helicase gene BLM. Mol. Genet. Genom. Med. 2016, 4, 106–119. [Google Scholar] [CrossRef] [PubMed]
  34. Killoran, M.P.; Keck, J.L. Three HRDC domains differentially modulate Deinococcus radiodurans RecQ DNA helicase biochemical activity. J. Biol. Chem. 2006, 281, 12849–12857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Guo, R.-B.; Rigolet, P.; Zargarian, L.; Fermandjian, S.; Xi, X.G. Structural and functional characterizations reveal the importance of a zinc binding domain in Bloom’s syndrome helicase. Nucleic Acids Res. 2005, 33, 3109–3124. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, J.L.; Rigolet, P.; Dou, S.-X.; Wang, P.-Y.; Xi, X.G. The zinc finger motif of Escherichia coli RecQ is implicated in both DNA binding and protein folding. J. Biol. Chem. 2004, 279, 42794–42802. [Google Scholar] [CrossRef] [Green Version]
  37. Gajiwala, K.S.; Burley, S.K. Winged helix proteins. Curr. Opin. Struct. Biol. 2000, 10, 110–116. [Google Scholar] [CrossRef]
  38. Rodríguez, A.C.; Stock, D. Crystal structure of reverse gyrase: Insights into the positive supercoiling of DNA. EMBO J. 2002, 21, 418–426. [Google Scholar] [CrossRef] [Green Version]
  39. Lima, C.D.; Wang, J.C.; Mondragón, A. Three-dimensional structure of the 67K N-terminal fragment of E. coli DNA topoisomerase I. Nature 1994, 367, 138. [Google Scholar] [CrossRef]
  40. Schultz, S.C.; Shields, G.C.; Steitz, T.A. Crystal structure of a CAP-DNA complex: The DNA is bent by 90 degrees. Science 1991, 253, 1001–1007. [Google Scholar] [CrossRef]
  41. Kitano, K.; Kim, S.-Y.; Hakoshima, T. Structural basis for DNA strand separation by the unconventional winged-helix domain of RecQ helicase WRN. Structure 2010, 18, 177–187. [Google Scholar] [CrossRef] [Green Version]
  42. Hoadley, K.A.; Keck, J.L. Werner helicase wings DNA binding. Structure 2010, 18, 149–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Wu, L.; Chan, K.L.; Ralf, C.; Bernstein, D.A.; Garcia, P.L.; Bohr, V.A.; Vindigni, A.; Janscak, P.; Keck, J.L.; Hickson, I.D. The HRDC domain of BLM is required for the dissolution of double Holliday junctions. EMBO J. 2005, 24, 2679–2687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Liu, Z.; Macias, M.J.; Bottomley, M.J.; Stier, G.; Linge, J.; Nilges, M.; Bork, P.; Sattler, M. The three-dimensional structure of the HRDC domain and implications for the Werner and Bloom syndrome proteins. Structure 1999, 7, 1557–1566. [Google Scholar] [CrossRef] [Green Version]
  45. Lillard-Wetherell, K.; Machwe, A.; Langland, G.T.; Combs, K.A.; Behbehani, G.K.; Schonberg, S.A.; German, J.; Turchi, J.J.; Orren, D.K.; Groden, J. Association and regulation of the BLM helicase by the telomere proteins TRF1 and TRF2. Hum. Mol. Genet. 2004, 13, 1919–1932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Huang, S.; Li, B.; Gray, M.D.; Oshima, J.; Mian, I.S.; Campisi, J. The premature ageing syndrome protein, WRN, is a 3’->5’ exonuclease. Nat. Genet. 1998, 20, 114–116. [Google Scholar] [CrossRef]
  47. Shen, J.C.; Gray, M.D.; Oshima, J.; Kamath-Loeb, A.S.; Fry, M.; Loeb, L.A. Werner syndrome protein. I. DNA helicase and DNA exonuclease reside on the same polypeptide. J. Biol. Chem. 1998, 273, 34139–34144. [Google Scholar] [CrossRef] [Green Version]
  48. Choudhary, S.; Sommers, J.A.; Brosh, R.M., Jr. Biochemical and kinetic characterization of the DNA helicase and exonuclease activities of werner syndrome protein. J. Biol. Chem. 2004, 279, 34603–34613. [Google Scholar] [CrossRef] [Green Version]
  49. Liberi, G.; Maffioletti, G.; Lucca, C.; Chiolo, I.; Baryshnikova, A.; Cotta-Ramusino, C.; Lopes, M.; Pellicioli, A.; Haber, J.E.; Foiani, M. Rad51-dependent DNA structures accumulate at damaged replication forks in sgs1 mutants defective in the yeast ortholog of BLM RecQ helicase. Genes Dev. 2005, 19, 339–350. [Google Scholar] [CrossRef] [Green Version]
  50. Cejka, P.; Cannavo, E.; Polaczek, P.; Masuda-Sasa, T.; Pokharel, S.; Campbell, J.L.; Kowalczykowski, S.C. DNA end resection by Dna2–Sgs1–RPA and its stimulation by Top3–Rmi1 and Mre11–Rad50–Xrs2. Nature 2010, 467, 112. [Google Scholar] [CrossRef]
  51. Pedrazzi, G.; Bachrati, C.Z.; Selak, N.; Studer, I.; Petkovic, M.; Hickson, I.D.; Jiricny, J.; Stagljar, I. The Bloom’s syndrome helicase interacts directly with the human DNA mismatch repair protein hMSH6. Biol. Chem. 2003, 384, 1155–1164. [Google Scholar] [CrossRef] [Green Version]
  52. Wang, T.-F.; Kung, W.-M. Supercomplex formation between Mlh1–Mlh3 and Sgs1–Top3 heterocomplexes in meiotic yeast cells. Biochem. Biophys. Res. Commun. 2002, 296, 949–953. [Google Scholar] [CrossRef]
  53. Bermúdez-López, M.; Pociño-Merino, I.; Sánchez, H.; Bueno, A.; Guasch, C.; Almedawar, S.; Bru-Virgili, S.; Garí, E.; Wyman, C.; Reverter, D. ATPase-dependent control of the Mms21 SUMO ligase during DNA repair. PLoS Biol. 2015, 13, e1002089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Bermúdez-López, M.; Villoria, M.T.; Esteras, M.; Jarmuz, A.; Torres-Rosell, J.; Clemente-Blanco, A.; Aragon, L. Sgs1’s roles in DNA end resection, HJ dissolution, and crossover suppression require a two-step SUMO regulation dependent on Smc5/6. Genes Dev. 2016, 30, 1339–1356. [Google Scholar] [CrossRef] [Green Version]
  55. Branzei, D.; Sollier, J.; Liberi, G.; Zhao, X.; Maeda, D.; Seki, M.; Enomoto, T.; Ohta, K.; Foiani, M. Ubc9-and mms21-mediated sumoylation counteracts recombinogenic events at damaged replication forks. Cell 2006, 127, 509–522. [Google Scholar] [CrossRef] [Green Version]
  56. Chakraborty, U.; George, C.M.; Lyndaker, A.M.; Alani, E. A delicate balance between repair and replication factors regulates recombination between divergent DNA sequences in Saccharomyces cerevisiae. Genetics 2016, 202, 525–540. [Google Scholar] [CrossRef] [Green Version]
  57. Campos-Doerfler, L.; Syed, S.; Schmidt, K.H. Sgs1 Binding to Rad51 Stimulates Homology-Directed DNA Repair in Saccharomyces cerevisiae. Genetics 2018, 208, 125–138. [Google Scholar] [CrossRef] [Green Version]
  58. Hegnauer, A.M.; Hustedt, N.; Shimada, K.; Pike, B.L.; Vogel, M.; Amsler, P.; Rubin, S.M.; van Leeuwen, F.; Guenole, A.; van Attikum, H.; et al. An N-terminal acidic region of Sgs1 interacts with Rpa70 and recruits Rad53 kinase to stalled forks. EMBO J. 2012, 31, 3768–3783. [Google Scholar] [CrossRef] [Green Version]
  59. Bachrati, C.Z.; Hickson, I.D. RecQ helicases: Guardian angels of the DNA replication fork. Chromosoma 2008, 117, 219–233. [Google Scholar] [CrossRef]
  60. Machwe, A.; Lozada, E.M.; Xiao, L.; Orren, D.K. Competition between the DNA unwinding and strand pairing activities of the Werner and Bloom syndrome proteins. BMC Mol. Biol. 2006, 7, 1. [Google Scholar] [CrossRef] [Green Version]
  61. Niu, H.; Chung, W.H.; Zhu, Z.; Kwon, Y.; Zhao, W.; Chi, P.; Prakash, R.; Seong, C.; Liu, D.; Lu, L.; et al. Mechanism of the ATP-dependent DNA end-resection machinery from Saccharomyces cerevisiae. Nature 2010, 467, 108–111. [Google Scholar] [CrossRef] [Green Version]
  62. Anderson, D.G.; Kowalczykowski, S.C. The translocating RecBCD enzyme stimulates recombination by directing RecA protein onto ssDNA in a chi-regulated manner. Cell 1997, 90, 77–86. [Google Scholar] [CrossRef] [Green Version]
  63. Chen, C.F.; Brill, S.J. An essential DNA strand-exchange activity is conserved in the divergent N-termini of BLM orthologs. EMBO J. 2010, 29, 1713–1725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Machwe, A.; Xiao, L.; Groden, J.; Matson, S.W.; Orren, D.K. RecQ family members combine strand pairing and unwinding activities to catalyze strand exchange. J. Biol. Chem. 2005, 280, 23397–23407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cheok, C.F.; Wu, L.; Garcia, P.L.; Janscak, P.; Hickson, I.D. The Bloom’s syndrome helicase promotes the annealing of complementary single-stranded DNA. Nucleic Acids Res 2005, 33, 3932–3941. [Google Scholar] [CrossRef]
  66. Beltrao, P.; Albanèse, V.; Kenner, L.R.; Swaney, D.L.; Burlingame, A.; Villén, J.; Lim, W.A.; Fraser, J.S.; Frydman, J.; Krogan, N.J. Systematic functional prioritization of protein posttranslational modifications. Cell 2012, 150, 413–425. [Google Scholar] [CrossRef] [Green Version]
  67. Swaney, D.L.; Beltrao, P.; Starita, L.; Guo, A.; Rush, J.; Fields, S.; Krogan, N.J.; Villén, J. Global analysis of phosphorylation and ubiquitylation cross-talk in protein degradation. Nat. Methods 2013, 10, 676. [Google Scholar] [CrossRef]
  68. Holt, L.J.; Tuch, B.B.; Villén, J.; Johnson, A.D.; Gygi, S.P.; Morgan, D.O. Global analysis of Cdk1 substrate phosphorylation sites provides insights into evolution. Science 2009, 325, 1682–1686. [Google Scholar] [CrossRef] [Green Version]
  69. Fricke, W.M.; Kaliraman, V.; Brill, S.J. Mapping the DNA topoisomerase III binding domain of the Sgs1 DNA helicase. J. Biol. Chem. 2001, 276, 8848–8855. [Google Scholar] [CrossRef] [Green Version]
  70. Bodenmiller, B.; Campbell, D.; Gerrits, B.; Lam, H.; Jovanovic, M.; Picotti, P.; Schlapbach, R.; Aebersold, R. PhosphoPep-a database of protein phosphorylation sites in model organisms. Nat. Biotechnol. 2008, 26, 1339–1340. [Google Scholar] [CrossRef] [Green Version]
  71. Bodenmiller, B.; Wanka, S.; Kraft, C.; Urban, J.; Campbell, D.; Pedrioli, P.G.; Gerrits, B.; Picotti, P.; Lam, H.; Vitek, O.; et al. Phosphoproteomic analysis reveals interconnected system-wide responses to perturbations of kinases and phosphatases in yeast. Sci. Signal. 2010, 3, rs4. [Google Scholar] [CrossRef] [Green Version]
  72. Zhou, C.; Elia, A.E.; Naylor, M.L.; Dephoure, N.; Ballif, B.A.; Goel, G.; Xu, Q.; Ng, A.; Chou, D.M.; Xavier, R.J. Profiling DNA damage-induced phosphorylation in budding yeast reveals diverse signaling networks. Proc. Natl. Acad. Sci. USA 2016, 113, E3667–E3675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Bennett, R.J.; Sharp, J.A.; Wang, J.C. Purification and characterization of the Sgs1 DNA helicase activity of Saccharomyces cerevisiae. J. Biol. Chem. 1998, 273, 9644–9650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Lu, C.-Y.; Tsai, C.-H.; Brill, S.J.; Teng, S.-C. Sumoylation of the BLM ortholog, Sgs1, promotes telomere–telomere recombination in budding yeast. Nucleic Acids Res. 2009, 38, 488–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Rog, O.; Miller, K.M.; Ferreira, M.G.; Cooper, J.P. Sumoylation of RecQ helicase controls the fate of dysfunctional telomeres. Mol. Cell 2009, 33, 559–569. [Google Scholar] [CrossRef]
  76. Swaffer, M.P.; Jones, A.W.; Flynn, H.R.; Snijders, A.P.; Nurse, P. Quantitative phosphoproteomics reveals the signaling dynamics of Cell-Cycle kinases in the fission Yeast Schizosaccharomyces pombe. Cell Rep. 2018, 24, 503–514. [Google Scholar] [CrossRef] [Green Version]
  77. Kettenbach, A.N.; Deng, L.; Wu, Y.; Baldissard, S.; Adamo, M.E.; Gerber, S.A.; Moseley, J.B. Quantitative phosphoproteomics reveals pathways for coordination of cell growth and division by the conserved fission yeast kinase pom1. Mol. Cell. Proteom. 2015, 14, 1275–1287. [Google Scholar] [CrossRef] [Green Version]
  78. Ahmad, F.; Stewart, E. The N-terminal region of the Schizosaccharomyces pombe RecQ helicase, Rqh1p, physically interacts with Topoisomerase III and is required for Rqh1p function. Mol. Genet. Genom. 2005, 273, 102–114. [Google Scholar] [CrossRef]
  79. Bennett, R.J.; Keck, J.L.; Wang, J.C. Binding specificity determines polarity of DNA unwinding by the Sgs1 protein of S. cerevisiae. J. Mol. Biol. 1999, 289, 235–248. [Google Scholar] [CrossRef]
  80. Sun, H.; Bennett, R.J.; Maizels, N. The Saccharomyces cerevisiae Sgs1 helicase efficiently unwinds GG paired DNAs. Nucleic Acids Res. 1999, 27, 1978–1984. [Google Scholar] [CrossRef]
  81. Sen, D.; Gilbert, W. A sodium-potassium switch in the formation of four-stranded G4-DNA. Nature 1990, 344, 410. [Google Scholar] [CrossRef]
  82. Huber, M.D.; Duquette, M.L.; Shiels, J.C.; Maizels, N. A conserved G4 DNA binding domain in RecQ family helicases. J. Mol. Biol. 2006, 358, 1071–1080. [Google Scholar] [CrossRef] [PubMed]
  83. Goodwin, A.; Wang, S.-W.; Toda, T.; Norbury, C.; Hickson, I.D. Topoisomerase III is essential for accurate nuclear division in Schizosaccharomyces pombe. Nucleic Acids Res. 1999, 27, 4050–4058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Bennett, R.J.; Noirot-Gros, M.-F.; Wang, J.C. Interaction between yeast sgs1 helicase and DNA topoisomerase III. J. Biol. Chem. 2000, 275, 26898–26905. [Google Scholar] [CrossRef] [PubMed]
  85. Dunø, M.; Thomsen, B.; Westergaard, O.; Krejci, L.; Bendixen, C. Genetic analysis of the Saccharomyces cerevisiae Sgs1 helicase defines an essential function for the Sgs1-Top3 complex in the absence of SRS2 or TOP1. Mol. Gen. Genet. MGG 2000, 264, 89–97. [Google Scholar] [CrossRef]
  86. Mullen, J.R.; Nallaseth, F.S.; Lan, Y.Q.; Slagle, C.E.; Brill, S.J. Yeast Rmi1/Nce4 controls genome stability as a subunit of the Sgs1-Top3 complex. Mol. Cell. Biol. 2005, 25, 4476–4487. [Google Scholar] [CrossRef] [Green Version]
  87. Kennedy, J.A.; Daughdrill, G.W.; Schmidt, K.H. A transient α-helical molecular recognition element in the disordered N-terminus of the Sgs1 helicase is critical for chromosome stability and binding of Top3/Rmi1. Nucleic Acids Res. 2013, 41, 10215–10227. [Google Scholar] [CrossRef]
  88. Wang, F.; Yang, Y.; Singh, T.R.; Busygina, V.; Guo, R.; Wan, K.; Wang, W.; Sung, P.; Meetei, A.R.; Lei, M. Crystal structures of RMI1 and RMI2, two OB-fold regulatory subunits of the BLM complex. Structure 2010, 18, 1159–1170. [Google Scholar] [CrossRef] [Green Version]
  89. Bocquet, N.; Bizard, A.H.; Abdulrahman, W.; Larsen, N.B.; Faty, M.; Cavadini, S.; Bunker, R.D.; Kowalczykowski, S.C.; Cejka, P.; Hickson, I.D.; et al. Structural and mechanistic insight into Holliday-junction dissolution by topoisomerase IIIalpha and RMI1. Nat. Struct. Mol. Biol. 2014, 21, 261–268. [Google Scholar] [CrossRef] [Green Version]
  90. Wu, L.; Davies, S.L.; Levitt, N.C.; Hickson, I.D. Potential role for the BLM helicase in recombinational repair via a conserved interaction with RAD51. J. Biol. Chem. 2001, 276, 19375–19381. [Google Scholar] [CrossRef] [Green Version]
  91. Chiolo, I.; Carotenuto, W.; Maffioletti, G.; Petrini, J.H.; 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, 5738–5751. [Google Scholar] [CrossRef] [Green Version]
  92. Saffi, J.; Feldmann, H.; Winnacker, E.-L.; Henriques, J.A. Interaction of the yeast Pso5/Rad16 and Sgs1 proteins: Influences on DNA repair and aging. Mutat. Res. DNA Repair 2001, 486, 195–206. [Google Scholar] [CrossRef]
  93. Mullen, J.R.; Kaliraman, V.; Brill, S.J. Bipartite structure of the SGS1 DNA helicase in Saccharomyces cerevisiae. Genetics 2000, 154, 1101–1114. [Google Scholar] [PubMed]
  94. Pedrazzi, G.; Perrera, C.; Blaser, H.; Kuster, P.; Marra, G.; Davies, S.L.; Ryu, G.-H.; Freire, R.; Hickson, I.D.; Jiricny, J. Direct association of Bloom’s syndrome gene product with the human mismatch repair protein MLH1. Nucleic Acids Res. 2001, 29, 4378–4386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Langland, G.; Kordich, J.; Creaney, J.; Goss, K.H.; Lillard-Wetherell, K.; Bebenek, K.; Kunkel, T.A.; Groden, J. The Bloom’s syndrome protein (BLM) interacts with MLH1 but is not required for DNA mismatch repair. J. Biol. Chem. 2001, 276, 30031–30035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Sugawara, N.; Goldfarb, T.; Studamire, B.; Alani, E.; Haber, J.E. Heteroduplex rejection during single-strand annealing requires Sgs1 helicase and mismatch repair proteins Msh2 and Msh6 but not Pms1. Proc. Natl. Acad. Sci. USA 2004, 101, 9315–9320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Tronnersjö, S.; Hanefalk, C.; Balciunas, D.; Hu, G.-Z.; Nordberg, N.; Murén, E.; Ronne, H. The jmjN and jmjC domains of the yeast zinc finger protein Gis1 interact with 19 proteins involved in transcription, sumoylation and DNA repair. Mol. Genet. Genom. 2007, 277, 57–70. [Google Scholar] [CrossRef]
  98. Chaudhury, I.; Koepp, D.M. Degradation of Mrc1 promotes recombination-mediated restart of stalled replication forks. Nucleic Acids Res. 2016, 45, 2558–2570. [Google Scholar] [CrossRef] [Green Version]
  99. Albers, M.; Diment, A.; Muraru, M.; Russell, C.S.; Beggs, J.D. Identification and characterization of Prp45p and Prp46p, essential pre-mRNA splicing factors. RNA 2003, 9, 138–150. [Google Scholar] [CrossRef] [Green Version]
  100. Piya, G.; Mueller, E.N.; Haas, H.K.; Ghospurkar, P.L.; Wilson, T.M.; Jensen, J.L.; Colbert, C.L.; Haring, S.J. Characterization of the Interaction between Rfa1 and Rad24 in Saccharomyces cerevisiae. PLoS ONE 2015, 10, e0116512. [Google Scholar] [CrossRef] [Green Version]
  101. Iacovella, M.G.; Golfieri, C.; Massari, L.F.; Busnelli, S.; Pagliuca, C.; Dal Maschio, M.; Infantino, V.; Visintin, R.; Mechtler, K.; Ferreira-Cerca, S. Rio1 promotes rDNA stability and downregulates RNA polymerase I to ensure rDNA segregation. Nat. Commun. 2015, 6, 6643. [Google Scholar] [CrossRef] [Green Version]
  102. Iacovella, M.G.; Bremang, M.; Basha, O.; Giacò, L.; Carotenuto, W.; Golfieri, C.; Szakal, B.; Dal Maschio, M.; Infantino, V.; Beznoussenko, G.V. Integrating Rio1 activities discloses its nutrient-activated network in Saccharomyces cerevisiae. Nucleic Acids Res. 2018, 46, 7586–7611. [Google Scholar] [CrossRef]
  103. Chin, J.K.; Bashkirov, V.I.; Heyer, W.-D.; Romesberg, F.E. Esc4/Rtt107 and the control of recombination during replication. DNA Repair 2006, 5, 618–628. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Sollier, J.; Driscoll, R.; Castellucci, F.; Foiani, M.; Jackson, S.P.; Branzei, D. The Saccharomyces cerevisiae Esc2 and Smc5-6 proteins promote sister chromatid junction-mediated intra-S repair. Mol. Biol. Cell 2009, 20, 1671–1682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Böhm, S.; Mihalevic, M.J.; Casal, M.A.; Bernstein, K.A. Disruption of SUMO-targeted ubiquitin ligases Slx5–Slx8/RNF4 alters RecQ-like helicase Sgs1/BLM localization in yeast and human cells. DNA Repair 2015, 26, 1–14. [Google Scholar] [CrossRef] [Green Version]
  106. Bonner, J.N.; Choi, K.; Xue, X.; Torres, N.P.; Szakal, B.; Wei, L.; Wan, B.; Arter, M.; Matos, J.; Sung, P. Smc5/6 mediated sumoylation of the Sgs1-Top3-Rmi1 complex promotes removal of recombination intermediates. Cell Rep. 2016, 16, 368–378. [Google Scholar] [CrossRef] [Green Version]
  107. Wong, J.; Nakajima, Y.; Westermann, S.; Shang, C.; Kang, J.-S.; Goodner, C.; Houshmand, P.; Fields, S.; Chan, C.S.; Drubin, D. A protein interaction map of the mitotic spindle. Mol. Biol. Cell 2007, 18, 3800–3809. [Google Scholar] [CrossRef] [Green Version]
  108. Pietrobon, V.; Freon, K.; Hardy, J.; Costes, A.; Iraqui, I.; Ochsenbein, F.; Lambert, S.A. The chromatin assembly factor 1 promotes Rad51-dependent template switches at replication forks by counteracting D-loop disassembly by the RecQ-type helicase Rqh1. PLoS Biol. 2014, 12, e1001968. [Google Scholar] [CrossRef] [Green Version]
  109. Kibe, T.; Ono, Y.; Sato, K.; Ueno, M. Fission yeast Taz1 and RPA are synergistically required to prevent rapid telomere loss. Mol. Biol. Cell 2007, 18, 2378–2387. [Google Scholar] [CrossRef] [Green Version]
  110. Vo, T.V.; Das, J.; Meyer, M.J.; Cordero, N.A.; Akturk, N.; Wei, X.; Fair, B.J.; Degatano, A.G.; Fragoza, R.; Liu, L.G. A proteome-wide fission yeast interactome reveals network evolution principles from yeasts to human. Cell 2016, 164, 310–323. [Google Scholar] [CrossRef] [Green Version]
  111. Watts, F.; Skilton, A.; Ho, J.-Y.; Boyd, L.K.; Trickey, M.; Gardner, L.; Ogi, F.-X.; Outwin, E. The role of Schizosaccharomyces pombe SUMO ligases in genome stability. Biochem. Soc. Trans. 2007. [Google Scholar] [CrossRef] [Green Version]
  112. McDonald, K.R.; Guise, A.J.; Pourbozorgi-Langroudi, P.; Cristea, I.M.; Zakian, V.A.; Capra, J.A.; Sabouri, N. Pfh1 is an accessory replicative helicase that interacts with the replisome to facilitate fork progression and preserve genome integrity. PLoS Genet. 2016, 12, e1006238. [Google Scholar] [CrossRef] [PubMed]
  113. Barlow, J.H.; Rothstein, R. Timing is everything: Cell cycle control of Rad52. Cell Div. 2010, 5, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Watt, P.M.; Hickson, I.D.; Borts, R.H.; Louis, E.J. SGS1, a homologue of the Bloom’s and Werner’s syndrome genes, is required for maintenance of genome stability in Saccharomyces cerevisiae. Genetics 1996, 144, 935–945. [Google Scholar] [PubMed]
  115. Ii, M.; Ii, T.; Mironova, L.I.; Brill, S.J. Epistasis analysis between homologous recombination genes in Saccharomyces cerevisiae identifies multiple repair pathways for Sgs1, Mus81-Mms4 and RNase H2. Mutat. Res. 2011, 714, 33–43. [Google Scholar] [CrossRef] [Green Version]
  116. Shor, E.; Gangloff, S.; Wagner, M.; Weinstein, J.; Price, G.; Rothstein, R. Mutations in homologous recombination genes rescue top3 slow growth in Saccharomyces cerevisiae. Genetics 2002, 162, 647–662. [Google Scholar]
  117. Bessler, J.B.; Torre, J.Z.; Zakian, V.A. The Pif1p subfamily of helicases: Region-specific DNA helicases? Trends Cell Biol. 2001, 11, 60–65. [Google Scholar] [CrossRef]
  118. Wagner, M.; Price, G.; Rothstein, R. The absence of Top3 reveals an interaction between the Sgs1 and Pif1 DNA helicases in Saccharomyces cerevisiae. Genetics 2006, 174, 555–573. [Google Scholar] [CrossRef] [Green Version]
  119. Lu, J.; Mullen, J.R.; Brill, S.J.; Kleff, S.; Romeo, A.M.; Sternglanz, R. Human homologues of yeast helicase. Nature 1996, 383, 678. [Google Scholar] [CrossRef]
  120. 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, 192. [Google Scholar] [CrossRef]
  121. Maftahi, M.; Hope, J.C.; Delgado-Cruzata, L.; Han, C.S.; Freyer, G.A. The severe slow growth of Δ srs2 Δ rqh1 in Schizosaccharomyces pombe is suppressed by loss of recombination and checkpoint genes. Nucleic Acids Res. 2002, 30, 4781–4792. [Google Scholar] [CrossRef] [Green Version]
  122. Doerfler, L.; Harris, L.; Viebranz, E.; Schmidt, K.H. Differential genetic interactions between Sgs1, DNA-damage checkpoint components and DNA repair factors in the maintenance of chromosome stability. Genome Integr. 2011, 2, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Doerfler, L.; Schmidt, K.H. Exo1 phosphorylation status controls the hydroxyurea sensitivity of cells lacking the Pol32 subunit of DNA polymerases delta and zeta. DNA Repair 2014, 24, 26–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Gravel, S.; Chapman, J.R.; Magill, C.; Jackson, S.P. DNA helicases Sgs1 and BLM promote DNA double-strand break resection. Genes Dev. 2008, 22, 2767–2772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Zhu, Z.; Chung, W.-H.; Shim, E.Y.; Lee, S.E.; Ira, G. Sgs1 helicase and two nucleases Dna2 and Exo1 resect DNA double-strand break ends. Cell 2008, 134, 981–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Mimitou, E.P.; Symington, L.S. Sae2, Exo1 and Sgs1 collaborate in DNA double-strand break processing. Nature 2008, 455, 770. [Google Scholar] [CrossRef] [Green Version]
  127. Spell, R.M.; Jinks-Robertson, S. Examination of the Roles of Sgs1 and Srs2 Helicases in the Enforcement of Recombination Fidelity in Saccharomyces cerevisiae. Genetics 2004, 168, 1855–1865. [Google Scholar] [CrossRef] [Green Version]
  128. Schmidt, K.H.; Kolodner, R.D. Suppression of spontaneous genome rearrangements in yeast DNA helicase mutants. Proc. Natl. Acad. Sci. USA 2006, 103, 18196–18201. [Google Scholar] [CrossRef] [Green Version]
  129. Lee, S.-K.; Johnson, R.E.; Yu, S.-L.; Prakash, L.; Prakash, S. Requirement of yeast SGS1 and SRS2 genes for replication and transcription. Science 1999, 286, 2339–2342. [Google Scholar] [CrossRef]
  130. Schmidt, K.H.; Kolodner, R.D. Requirement of Rrm3 helicase for repair of spontaneous DNA lesions in cells lacking Srs2 or Sgs1 helicase. Mol. Cell. Biol. 2004, 24, 3213–3226. [Google Scholar] [CrossRef] [Green Version]
  131. Kaliraman, V.; Mullen, J.R.; Fricke, W.M.; Bastin-Shanower, S.A.; Brill, S.J. Functional overlap between Sgs1–Top3 and the Mms4–Mus81 endonuclease. Genes Dev. 2001, 15, 2730–2740. [Google Scholar] [CrossRef] [Green Version]
  132. Ooi, S.L.; Shoemaker, D.D.; Boeke, J.D. DNA helicase gene interaction network defined using synthetic lethality analyzed by microarray. Nat. Genet. 2003, 35, 277. [Google Scholar] [CrossRef]
  133. Mimitou, E.P.; Symington, L.S. Ku prevents Exo1 and Sgs1-dependent resection of DNA ends in the absence of a functional MRX complex or Sae2. EMBO J. 2010, 29, 3358–3369. [Google Scholar] [CrossRef] [Green Version]
  134. Torres, J.Z.; Schnakenberg, S.L.; Zakian, V.A. Saccharomyces cerevisiae Rrm3p DNA helicase promotes genome integrity by preventing replication fork stalling: Viability of rrm3 cells requires the intra-S-phase checkpoint and fork restart activities. Mol. Cell. Biol. 2004, 24, 3198–3212. [Google Scholar] [CrossRef] [Green Version]
  135. Fabre, F.; Chan, A.; Heyer, W.-D.; Gangloff, S. Alternate pathways involving Sgs1/Top3, Mus81/Mms4, and Srs2 prevent formation of toxic recombination intermediates from single-stranded gaps created by DNA replication. Proc. Natl. Acad. Sci. USA 2002, 99, 16887–16892. [Google Scholar] [CrossRef] [Green Version]
  136. Boddy, M.N.; Lopez-Girona, A.; Shanahan, P.; Interthal, H.; Heyer, W.-D.; Russell, P. Damage tolerance protein Mus81 associates with the FHA1 domain of checkpoint kinase Cds1. Mol. Cell. Biol. 2000, 20, 8758–8766. [Google Scholar] [CrossRef] [Green Version]
  137. Anderson, R.M.; Sinclair, D.A. Yeast RecQ Helicases: Clues to DNA Repair, Genome Stability and Aging. Mol. Mech. Werner’s Syndr. 2004, 78. [Google Scholar]
  138. Caspari, T.; Murray, J.M.; Carr, A.M. Cdc2—Cyclin B kinase activity links Crb2 and Rqh1—Topoisomerase III. Genes Dev. 2002, 16, 1195–1208. [Google Scholar] [CrossRef] [Green Version]
  139. Doe, C.L.; Ahn, J.S.; Dixon, J.; Whitby, M.C. Mus81-Eme1 and Rqh1 involvement in processing stalled and collapsed replication forks. J. Biol. Chem. 2002, 277, 32753–32759. [Google Scholar] [CrossRef] [Green Version]
  140. Ashton, T.M.; Hickson, I.D. Yeast as a model system to study RecQ helicase function. DNA Repair 2010, 9, 303–314. [Google Scholar] [CrossRef]
  141. Laursen, L.V.; Ampatzidou, E.; Andersen, A.H.; Murray, J.M. Role for the fission yeast RecQ helicase in DNA repair in G2. Mol. Cell. Biol. 2003, 23, 3692–3705. [Google Scholar] [CrossRef] [Green Version]
  142. Budd, M.E.; Campbell, J.L. A yeast gene required for DNA replication encodes a protein with homology to DNA helicases. Proc. Natl. Acad. Sci. USA 1995, 92, 7642–7646. [Google Scholar] [CrossRef] [Green Version]
  143. Gnügge, R.; Symington, L.S. Keeping it real: MRX—Sae2 clipping of natural substrates. Genes Dev. 2017, 31, 2311–2312. [Google Scholar] [CrossRef] [Green Version]
  144. Xue, C.; Wang, W.; Crickard, J.B.; Moevus, C.J.; Kwon, Y.; Sung, P.; Greene, E.C. Regulatory control of Sgs1 and Dna2 during eukaryotic DNA end resection. Proc. Natl. Acad. Sci. USA 2019, 116, 6091–6100. [Google Scholar] [CrossRef] [Green Version]
  145. Adkins, N.L.; Niu, H.; Sung, P.; Peterson, C.L. Nucleosome dynamics regulates DNA processing. Nat. Struct. Mol. Biol. 2013, 20, 836. [Google Scholar] [CrossRef] [Green Version]
  146. Kalocsay, M.; Hiller, N.J.; Jentsch, S. Chromosome-wide Rad51 spreading and SUMO-H2A.Z-dependent chromosome fixation in response to a persistent DNA double-strand break. Mol. Cell 2009, 33, 335–343. [Google Scholar] [CrossRef]
  147. Papamichos-Chronakis, M.; Krebs, J.E.; Peterson, C.L. Interplay between Ino80 and Swr1 chromatin remodeling enzymes regulates cell cycle checkpoint adaptation in response to DNA damage. Genes Dev. 2006, 20, 2437–2449. [Google Scholar] [CrossRef] [Green Version]
  148. Bernstein, K.A.; Reid, R.J.; Sunjevaric, I.; Demuth, K.; Burgess, R.C.; Rothstein, R. The Shu complex, which contains Rad51 paralogues, promotes DNA repair through inhibition of the Srs2 anti-recombinase. Mol. Biol. Cell 2011, 22, 1599–1607. [Google Scholar] [CrossRef]
  149. Godin, S.; Wier, A.; Kabbinavar, F.; Bratton-Palmer, D.S.; Ghodke, H.; Van Houten, B.; VanDemark, A.P.; Bernstein, K.A. The Shu complex interacts with Rad51 through the Rad51 paralogues Rad55-Rad57 to mediate error-free recombination. Nucleic Acids Res. 2013, 41, 4525–4534. [Google Scholar] [CrossRef] [Green Version]
  150. She, Z.; Gao, Z.Q.; Liu, Y.; Wang, W.J.; Liu, G.F.; Shtykova, E.V.; Xu, J.H.; Dong, Y.H. Structural and SAXS analysis of the budding yeast SHU-complex proteins. FEBS Lett. 2012, 586, 2306–2312. [Google Scholar] [CrossRef] [Green Version]
  151. Sugiyama, T.; Kowalczykowski, S.C. Rad52 protein associates with replication protein A (RPA)-single-stranded DNA to accelerate Rad51-mediated displacement of RPA and presynaptic complex formation. J. Biol. Chem. 2002, 277, 31663–31672. [Google Scholar] [CrossRef] [Green Version]
  152. Crickard, J.B.; Xue, C.; Wang, W.; Kwon, Y.; Sung, P.; Greene, E.C. The RecQ helicase Sgs1 drives ATP-dependent disruption of Rad51 filaments. Nucleic Acids Res. 2019, 47, 4694–4706. [Google Scholar] [CrossRef] [PubMed]
  153. Antony, E.; Tomko, E.J.; Xiao, Q.; Krejci, L.; Lohman, T.M.; Ellenberger, T. Srs2 disassembles Rad51 filaments by a protein-protein interaction triggering ATP turnover and dissociation of Rad51 from DNA. Mol. Cell 2009, 35, 105–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Kodadek, T. Inhibition of protein-mediated homologous pairing by a DNA helicase. J. Biol. Chem. 1991, 266, 9712–9718. [Google Scholar] [PubMed]
  155. Morel, P.; Hejna, J.A.; Ehrlich, S.D.; Cassuto, E. Antipairing and strand transferase activities of E. coli helicase II (UvrD). Nucleic Acids Res. 1993, 21, 3205–3209. [Google Scholar] [CrossRef] [Green Version]
  156. Fasching, C.L.; Cejka, P.; Kowalczykowski, S.C.; Heyer, W.-D. Top3-Rmi1 dissolve Rad51-mediated D loops by a topoisomerase-based mechanism. Mol. Cell 2015, 57, 595–606. [Google Scholar] [CrossRef] [Green Version]
  157. Cejka, P.; Plank, J.L.; Bachrati, C.Z.; Hickson, I.D.; Kowalczykowski, S.C. Rmi1 stimulates decatenation of double Holliday junctions during dissolution by Sgs1–Top3. Nat. Struct. Mol. Biol. 2010, 17, 1377–1382. [Google Scholar] [CrossRef] [Green Version]
  158. Menolfi, D.; Delamarre, A.; Lengronne, A.; Pasero, P.; Branzei, D. Essential roles of the Smc5/6 complex in replication through natural pausing sites and endogenous DNA damage tolerance. Mol. Cell 2015, 60, 835–846. [Google Scholar] [CrossRef] [Green Version]
  159. Bermúdez-López, M.; Villoria, M.T.; Esteras, M.; Jarmuz, A.; Torres-Rosell, J.; Clemente-Blanco, A.; Aragon, L. Sgs1’s Roles in DNA End Resection, Hj Dissolution, and Crossover Suppression Require a Two-Step Sumo Regulation Dependent on Smc5/6. Genes Dev. 2016, 11, 1339–1356. [Google Scholar]
  160. Doe, C.L.; Dixon, J.; Osman, F.; Whitby, M.C. Partial suppression of the fission yeast rqh1− phenotype by expression of a bacterial Holliday junction resolvase. EMBO J. 2000, 19, 2751–2762. [Google Scholar] [CrossRef] [Green Version]
  161. Hope, J.C.; Maftahi, M.; Freyer, G.A. A postsynaptic role for Rhp55/57 that is responsible for cell death in Δrqh1 mutants following replication arrest in Schizosaccharomyces pombe. Genetics 2005, 170, 519–531. [Google Scholar] [CrossRef] [Green Version]
  162. Langerak, P.; Mejia-Ramirez, E.; Limbo, O.; Russell, P. Release of Ku and MRN from DNA ends by Mre11 nuclease activity and Ctp1 is required for homologous recombination repair of double-strand breaks. PLoS Genet. 2011, 7, e1002271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Zhang, J.-M.; Liu, X.-M.; Ding, Y.-H.; Xiong, L.-Y.; Ren, J.-Y.; Zhou, Z.-X.; Wang, H.-T.; Zhang, M.-J.; Yu, Y.; Dong, M.-Q. Fission yeast Pxd1 promotes proper DNA repair by activating Rad16XPF and inhibiting Dna2. PLoS Biol. 2014, 12, e1001946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Nanbu, T.; Nguyễn, L.C.; Habib, A.G.; Hirata, N.; Ukimori, S.; Tanaka, D.; Masuda, K.; Takahashi, K.; Yukawa, M.; Tsuchiya, E. Fission yeast Exo1 and Rqh1-Dna2 redundantly contribute to resection of uncapped telomeres. PLoS ONE 2015, 10, e0140456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Frei, C.; Gasser, S.M. The yeast Sgs1p helicase acts upstream of Rad53p in the DNA replication checkpoint and colocalizes with Rad53p in S-phase-specific foci. Genes Dev. 2000, 14, 81–96. [Google Scholar] [PubMed]
  166. Cobb, J.A.; Bjergbaek, L.; Shimada, K.; Frei, C.; Gasser, S.M. DNA polymerase stabilization at stalled replication forks requires Mec1 and the RecQ helicase Sgs1. EMBO J. 2003, 22, 4325–4336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Saffi, J.; Pereira, V.R.; Henriques, J.A.P. Importance of the Sgs1 helicase activity in DNA repair of Saccharomyces cerevisiae. Curr. Genet. 2000, 37, 75–78. [Google Scholar] [CrossRef] [PubMed]
  168. Ui, A.; Seki, M.; Ogiwara, H.; Lai, M.S.; Yamamoto, K.; Tada, S.; Enomoto, T. Activation of a novel pathway involving Mms1 and Rad59 in sgs1 cells. Biochem. Biophys. Res. Commun. 2007, 356, 1031–1037. [Google Scholar] [CrossRef]
  169. Versini, G.; Comet, I.; Wu, M.; Hoopes, L.; Schwob, E.; Pasero, P. The yeast Sgs1 helicase is differentially required for genomic and ribosomal DNA replication. EMBO J. 2003, 22, 1939–1949. [Google Scholar] [CrossRef] [Green Version]
  170. Marchetti, M.A.; Kumar, S.; Hartsuiker, E.; Maftahi, M.; Carr, A.M.; Freyer, G.A.; Burhans, W.C.; Huberman, J.A. A single unbranched S-phase DNA damage and replication fork blockage checkpoint pathway. Proc. Natl. Acad. Sci. USA 2002, 99, 7472–7477. [Google Scholar] [CrossRef] [Green Version]
  171. Mankouri, H.W.; Morgan, A. The DNA helicase activity of yeast Sgs1p is essential for normal lifespan but not for resistance to topoisomerase inhibitors. Mech. Ageing Dev. 2001, 122, 1107–1120. [Google Scholar] [CrossRef]
  172. Wang, J.C. DNA topoisomerases. Annu. Rev. Biochem. 1996, 65, 635–692. [Google Scholar] [CrossRef] [PubMed]
  173. Sogo, J.M.; Lopes, M.; Foiani, M. Fork reversal and ssDNA accumulation at stalled replication forks owing to checkpoint defects. Science 2002, 297, 599–602. [Google Scholar] [CrossRef] [PubMed]
  174. Nielsen, I.; Bentsen, I.B.; Andersen, A.H.; Gasser, S.M.; Bjergbaek, L. A Rad53 independent function of Rad9 becomes crucial for genome maintenance in the absence of the Recq helicase Sgs1. PLoS ONE 2013, 8, e81015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Ira, G.; Malkova, A.; Liberi, G.; Foiani, M.; Haber, J.E. Srs2 and Sgs1-Top3 suppress crossovers during double-strand break repair in yeast. Cell 2003, 115, 401–411. [Google Scholar] [CrossRef] [Green Version]
  176. Wu, L.; Hickson, I.D. The Bloom’s syndrome helicase suppresses crossing over during homologous recombination. Nature 2003, 426, 870–874. [Google Scholar] [CrossRef]
  177. Krejci, L.; Van Komen, S.; Li, Y.; Villemain, J.; Reddy, M.S.; Klein, H.; Ellenberger, T.; Sung, P. DNA helicase Srs2 disrupts the Rad51 presynaptic filament. Nature 2003, 423, 305–309. [Google Scholar] [CrossRef]
  178. Prakash, R.; Satory, D.; Dray, E.; Papusha, A.; Scheller, J.; Kramer, W.; Krejci, L.; Klein, H.; Haber, J.E.; Sung, P.; et al. Yeast Mph1 helicase dissociates Rad51-made D-loops: Implications for crossover control in mitotic recombination. Genes Dev. 2009, 23, 67–79. [Google Scholar] [CrossRef] [Green Version]
  179. Mankouri, H.W.; Ngo, H.P.; Hickson, I.D. Shu proteins promote the formation of homologous recombination intermediates that are processed by Sgs1-Rmi1-Top3. Mol. Biol. Cell 2007, 18, 4062–4073. [Google Scholar] [CrossRef] [Green Version]
  180. Panico, E.R.; Ede, C.; Schildmann, M.; Schurer, K.A.; Kramer, W. Genetic evidence for a role of Saccharomyces cerevisiae Mph1 in recombinational DNA repair under replicative stress. Yeast 2010, 27, 11–27. [Google Scholar] [CrossRef]
  181. Fricke, W.M.; Brill, S.J. Slx1—Slx4 is a second structure-specific endonuclease functionally redundant with Sgs1—Top3. Genes Dev. 2003, 17, 1768–1778. [Google Scholar] [CrossRef] [Green Version]
  182. Mullen, J.R.; Kaliraman, V.; Ibrahim, S.S.; Brill, S.J. Requirement for three novel protein complexes in the absence of the Sgs1 DNA helicase in Saccharomyces cerevisiae. Genetics 2001, 157, 103–118. [Google Scholar] [PubMed]
  183. Cobb, J.A.; Schleker, T.; Rojas, V.; Bjergbaek, L.; Tercero, J.A.; Gasser, S.M. Replisome instability, fork collapse, and gross chromosomal rearrangements arise synergistically from Mec1 kinase and RecQ helicase mutations. Genes Dev. 2005, 19, 3055–3069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Branzei, D.; Vanoli, F.; Foiani, M. SUMOylation regulates Rad18-mediated template switch. Nature 2008, 456, 915. [Google Scholar] [CrossRef]
  185. Karras, G.I.; Jentsch, S. The RAD6 DNA damage tolerance pathway operates uncoupled from the replication fork and is functional beyond S phase. Cell 2010, 141, 255–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Li, F.; Ball, L.G.; Fan, L.; Hanna, M.; Xiao, W. Sgs1 helicase is required for efficient PCNA monoubiquitination and translesion DNA synthesis in Saccharomyces cerevisiae. Curr. Genet. 2018, 64, 459–468. [Google Scholar] [CrossRef]
  187. Rockmill, B.; Fung, J.C.; Branda, S.S.; Roeder, G.S. The Sgs1 helicase regulates chromosome synapsis and meiotic crossing over. Curr. Biol. 2003, 13, 1954–1962. [Google Scholar] [CrossRef] [Green Version]
  188. Miyajima, A.; Seki, M.; Onoda, F.; Ui, A.; Satoh, Y.; Ohno, Y.; Enomoto, T. Different domains of Sgs1 are required for mitotic and meiotic functions. Genes Genet. Syst. 2000, 75, 319–326. [Google Scholar] [CrossRef]
  189. Gangloff, S.; de Massy, B.; Arthur, L.; Rothstein, R.; Fabre, F. The essential role of yeast topoisomerase III in meiosis depends on recombination. EMBO J. 1999, 18, 1701–1711. [Google Scholar] [CrossRef] [Green Version]
  190. Jessop, L.; Rockmill, B.; Roeder, G.S.; Lichten, M. Meiotic chromosome synapsis-promoting proteins antagonize the anti-crossover activity of sgs1. PLoS Genet. 2006, 2, e155. [Google Scholar] [CrossRef]
  191. Oh, S.D.; Lao, J.P.; Hwang, P.Y.; Taylor, A.F.; Smith, G.R.; Hunter, N. BLM ortholog, Sgs1, prevents aberrant crossing-over by suppressing formation of multichromatid joint molecules. Cell 2007, 130, 259–272. [Google Scholar] [CrossRef] [Green Version]
  192. Miyajima, A.; Seki, M.; Onoda, F.; Shiratori, M.; Odagiri, N.; Ohta, K.; Kikuchi, Y.; Ohno, Y.; Enomoto, T. Sgs1 helicase activity is required for mitotic but apparently not for meiotic functions. Mol. Cell. Biol. 2000, 20, 6399–6409. [Google Scholar] [CrossRef] [PubMed]
  193. Agarwal, S.; Roeder, G.S. Zip3 provides a link between recombination enzymes and synaptonemal complex proteins. Cell 2000, 102, 245–255. [Google Scholar] [CrossRef] [Green Version]
  194. Oh, S.D.; Lao, J.P.; Taylor, A.F.; Smith, G.R.; Hunter, N. RecQ helicase, Sgs1, and XPF family endonuclease, Mus81-Mms4, resolve aberrant joint molecules during meiotic recombination. Mol. Cell 2008, 31, 324–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Jessop, L.; Lichten, M. Mus81/Mms4 endonuclease and Sgs1 helicase collaborate to ensure proper recombination intermediate metabolism during meiosis. Mol. Cell 2008, 31, 313–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Boddy, M.N.; Gaillard, P.-H.L.; McDonald, W.H.; Shanahan, P.; Yates, J.R., III; Russell, P. Mus81-Eme1 are essential components of a Holliday junction resolvase. Cell 2001, 107, 537–548. [Google Scholar] [CrossRef] [Green Version]
  197. Cromie, G.A.; Hyppa, R.W.; Taylor, A.F.; Zakharyevich, K.; Hunter, N.; Smith, G.R. Single Holliday junctions are intermediates of meiotic recombination. Cell 2006, 127, 1167–1178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Lynn, A.; Soucek, R.; Börner, G.V. ZMM proteins during meiosis: Crossover artists at work. Chromosome Res. 2007, 15, 591–605. [Google Scholar] [CrossRef] [Green Version]
  199. Cromie, G.A.; Hyppa, R.W.; Smith, G.R. The fission yeast BLM homolog Rqh1 promotes meiotic recombination. Genetics 2008, 179, 1157–1167. [Google Scholar] [CrossRef] [Green Version]
  200. Cromie, G.A.; Smith, G.R. Branching out: Meiotic recombination and its regulation. Trends Cell Biol. 2007, 17, 448–455. [Google Scholar] [CrossRef]
  201. De Boer, J.; Hoeijmakers, J.H. Nucleotide excision repair and human syndromes. Carcinogenesis 2000, 21, 453–460. [Google Scholar] [CrossRef] [Green Version]
  202. Chalissery, J.; Jalal, D.; Al-Natour, Z.; Hassan, A.H. Repair of oxidative DNA damage in Saccharomyces cerevisiae. DNA Repair 2017, 51, 2–13. [Google Scholar] [CrossRef] [PubMed]
  203. Krokan, H.E.; Bjoras, M. Base excision repair. Cold Spring Harb. Perspect. Biol. 2013, 5, a012583. [Google Scholar] [CrossRef] [PubMed]
  204. Brosh, R.M.; von Kobbe, C.; Sommers, J.A.; Karmakar, P.; Opresko, P.L.; Piotrowski, J.; Dianova, I.; Dianov, G.L.; Bohr, V.A. Werner syndrome protein interacts with human flap endonuclease 1 and stimulates its cleavage activity. EMBO J. 2001, 20, 5791–5801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Sharma, S.; Sommers, J.A.; Wu, L.; Bohr, V.A.; Hickson, I.D.; Brosh, R.M. Stimulation of flap endonuclease-1 by the Bloom’s syndrome protein. J. Biol. Chem. 2004, 279, 9847–9856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Sharma, S.; Sommers, J.A.; Gary, R.K.; Friedrich-Heineken, E.; Hübscher, U.; Brosh, R.M., Jr. The interaction site of Flap Endonuclease-1 with WRN helicase suggests a coordination of WRN and PCNA. Nucleic Acids Res. 2005, 33, 6769–6781. [Google Scholar] [CrossRef] [Green Version]
  207. Trego, K.S.; Chernikova, S.B.; Davalos, A.R.; Perry, J.J.P.; Finger, L.D.; Ng, C.; Tsai, M.-S.; Yannone, S.M.; Tainer, J.A.; Campisi, J. The DNA repair endonuclease XPG interacts directly and functionally with the WRN helicase defective in Werner syndrome. Cell Cycle 2011, 10, 1998–2007. [Google Scholar] [CrossRef] [Green Version]
  208. Fan, W.; Luo, J. RecQ4 facilitates UV light-induced DNA damage repair through interaction with nucleotide excision repair factor xeroderma pigmentosum group A (XPA). J. Biol. Chem. 2008, 283, 29037–29044. [Google Scholar] [CrossRef] [Green Version]
  209. Ringvoll, J.; Uldal, L.; Roed, M.A.; Reite, K.; Baynton, K.; Klungland, A.; Eide, L. Mutations in the RAD27 and SGS1 genes differentially affect the chronological and replicative lifespan of yeast cells growing on glucose and glycerol. FEMS Yeast Res. 2007, 7, 848–859. [Google Scholar] [CrossRef]
  210. Huang, M.-E.; Kolodner, R.D. A biological network in Saccharomyces cerevisiae prevents the deleterious effects of endogenous oxidative DNA damage. Mol. Cell 2005, 17, 709–720. [Google Scholar] [CrossRef]
  211. Schulz, V.P.; Zakian, V.A.; Ogburn, C.E.; McKay, J.; Jarzebowicz, A.A.; Martin, G.; Edland, S. Accelerated loss of telomeric repeats may not explain accelerated replicative decline of Werner syndrome cells. Hum. Genet. 1996, 97, 750–754. [Google Scholar] [CrossRef]
  212. Tahara, H.; Tokutake, Y.; Maeda, S.; Kataoka, H.; Watanabe, T.; Satoh, M.; Matsumoto, T.; Sugawara, M.; Ide, T.; Goto, M. Abnormal telomere dynamics of B-lymphoblastoid cell strains from Werner’s syndrome patients transformed by Epstein–Barr virus. Oncogene 1997, 15, 1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Wyllie, F.S.; Jones, C.J.; Skinner, J.W.; Haughton, M.F.; Wallis, C.; Wynford-Thomas, D.; Faragher, R.G.; Kipling, D. Telomerase prevents the accelerated cell ageing of Werner syndrome fibroblasts. Nat. Genet. 2000, 24, 16. [Google Scholar] [CrossRef] [PubMed]
  214. Bonetti, D.; Martina, M.; Clerici, M.; Lucchini, G.; Longhese, M.P. Multiple pathways regulate 3’ overhang generation at S. cerevisiae telomeres. Mol. Cell 2009, 35, 70–81. [Google Scholar] [CrossRef] [PubMed]
  215. Bryan, T.M.; Englezou, A.; Dalla-Pozza, L.; Dunham, M.A.; Reddel, R.R. Evidence for an alternative mechanism for maintaining telomere length in human tumors and tumor-derived cell lines. Nat. Med. 1997, 3, 1271. [Google Scholar] [CrossRef] [PubMed]
  216. Cesare, A.J.; Reddel, R.R. Telomere uncapping and alternative lengthening of telomeres. Mech. Ageing Dev. 2008, 129, 99–108. [Google Scholar] [CrossRef] [PubMed]
  217. Johnson, F.B.; Marciniak, R.A.; McVey, M.; Stewart, S.A.; Hahn, W.C.; Guarente, L. The Saccharomyces cerevisiae WRN homolog Sgs1p participates in telomere maintenance in cells lacking telomerase. EMBO J. 2001, 20, 905–913. [Google Scholar] [CrossRef] [Green Version]
  218. Cohen, H.; Sinclair, D.A. Recombination-mediated lengthening of terminal telomeric repeats requires the Sgs1 DNA helicase. Proc. Natl. Acad. Sci. USA 2001, 98, 3174–3179. [Google Scholar] [CrossRef] [Green Version]
  219. Huang, P.-H.; Pryde, F.E.; Lester, D.; Maddison, R.L.; Borts, R.H.; Hickson, I.D.; Louis, E.J. SGS1 is required for telomere elongation in the absence of telomerase. Curr. Biol. 2001, 11, 125–129. [Google Scholar] [CrossRef] [Green Version]
  220. Hardy, J.; Churikov, D.; Géli, V.; Simon, M.-N. Sgs1 and Sae2 promote telomere replication by limiting accumulation of ssDNA. Nat. Commun. 2014, 5, 5004. [Google Scholar] [CrossRef] [Green Version]
  221. Lee, J.Y.; Kozak, M.; Martin, J.D.; Pennock, E.; Johnson, F.B. Evidence that a RecQ helicase slows senescence by resolving recombining telomeres. PLoS Biol. 2007, 5, e160. [Google Scholar] [CrossRef]
  222. Luke-Glaser, S.; Luke, B. The Mph1 helicase can promote telomere uncapping and premature senescence in budding yeast. PLoS ONE 2012, 7, e42028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Lu, R.; O’Rourke, J.J.; Sobinoff, A.P.; Allen, J.A.; Nelson, C.B.; Tomlinson, C.G.; Lee, M.; Reddel, R.R.; Deans, A.J.; Pickett, H.A. The FANCM-BLM-TOP3A-RMI complex suppresses alternative lengthening of telomeres (ALT). Nat. Commun. 2019, 10, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Nanbu, T.; Takahashi, K.; Murray, J.M.; Hirata, N.; Ukimori, S.; Kanke, M.; Masukata, H.; Yukawa, M.; Tsuchiya, E.; Ueno, M. Fission yeast RecQ helicase Rqh1 is required for the maintenance of circular chromosomes. Mol. Cell. Biol. 2013, 33, 1175–1187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. McVey, M.; Kaeberlein, M.; Tissenbaum, H.A.; Guarente, L. The short life span of Saccharomyces cerevisiae sgs1 and srs2 mutants is a composite of normal aging processes and mitotic arrest due to defective recombination. Genetics 2001, 157, 1531–1542. [Google Scholar]
  226. Sinclair, D.A.; Guarente, L. Extrachromosomal rDNA circles—A cause of aging in yeast. Cell 1997, 91, 1033–1042. [Google Scholar] [CrossRef] [Green Version]
  227. Heo, S.J.; Tatebayashi, K.; Ohsugi, I.; Shimamoto, A.; Furuichi, Y.; Ikeda, H. Bloom’s syndrome gene suppresses premature ageing caused by Sgs1 deficiency in yeast. Genes Cells 1999, 4, 619–625. [Google Scholar] [CrossRef]
  228. Lee, H.-C.; Wei, Y.-H. Mitochondria and aging. In Advances in Mitochondrial Medicine; Springer: Berlin/Heidelberg, Germany, 2012; pp. 311–327. [Google Scholar]
  229. Croteau, D.L.; Rossi, M.L.; Canugovi, C.; Tian, J.; Sykora, P.; Ramamoorthy, M.; Wang, Z.; Singh, D.K.; Akbari, M.; Kasiviswanathan, R. RECQL4 localizes to mitochondria and preserves mitochondrial DNA integrity. Aging Cell 2012, 11, 456–466. [Google Scholar] [CrossRef] [Green Version]
  230. Smeal, T.; Claus, J.; Kennedy, B.; Cole, F.; Guarente, L. Loss of transcriptional silencing causes sterility in old mother cells of S. cerevisiae. Cell 1996, 84, 633–642. [Google Scholar] [CrossRef] [Green Version]
  231. Barea, F.; Tessaro, S.; Bonatto, D. In silico analyses of a new group of fungal and plant RecQ4-homologous proteins. Comput. Biol. Chem. 2008, 32, 349–358. [Google Scholar] [CrossRef]
  232. Kwon, S.-H.; Choi, D.-H.; Lee, R.; Bae, S.-H. Saccharomyces cerevisiae Hrq1 requires a long 3′-tailed DNA substrate for helicase activity. Biochem. Biophys. Res. Commun. 2012, 427, 623–628. [Google Scholar] [CrossRef]
  233. Choi, D.-H.; Lee, R.; Kwon, S.-H.; Bae, S.-H. Hrq1 functions independently of Sgs1 to preserve genome integrity in Saccharomyces cerevisiae. J. Microbiol. 2013, 51, 105–112. [Google Scholar] [CrossRef] [PubMed]
  234. Groocock, L.M.; Prudden, J.; Perry, J.J.P.; Boddy, M.N. The RecQ4 orthologue Hrq1 is critical for DNA interstrand cross-link repair and genome stability in fission yeast. Mol. Cell. Biol. 2012, 32, 276–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Bochman, M.L.; Paeschke, K.; Chan, A.; Zakian, V.A. Hrq1, a homolog of the human RecQ4 helicase, acts catalytically and structurally to promote genome integrity. Cell Rep. 2014, 6, 346–356. [Google Scholar] [CrossRef] [Green Version]
  236. Choi, D.-H.; Min, M.-H.; Kim, M.-J.; Lee, R.; Kwon, S.-H.; Bae, S.-H. Hrq1 facilitates nucleotide excision repair of DNA damage induced by 4-nitroquinoline-1-oxide and cisplatin in Saccharomyces cerevisiae. J. Microbiol. 2014, 52, 292–298. [Google Scholar] [CrossRef] [PubMed]
  237. Mayor, T.; Graumann, J.; Bryan, J.; MacCoss, M.J.; Deshaies, R.J. Quantitative profiling of ubiquitylated proteins reveals proteasome substrates and the substrate repertoire influenced by the Rpn10 receptor pathway. Mol. Cell. Proteom. 2007, 6, 1885–1895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Mandell, J.G.; Goodrich, K.J.; Bähler, J.; Cech, T.R. Expression of a RecQ helicase homolog affects progression through crisis in fission yeast lacking telomerase. J. Biol. Chem. 2005, 280, 5249–5257. [Google Scholar] [CrossRef] [Green Version]
  239. Legrand, M.; Chan, C.L.; Jauert, P.A.; Kirkpatrick, D.T. The contribution of the S-phase checkpoint genes MEC1 and SGS1 to genome stability maintenance in Candida albicans. Fungal Genet. Biol. 2011, 48, 823–830. [Google Scholar] [CrossRef] [Green Version]
  240. Rai, M.N.; Balusu, S.; Gorityala, N.; Dandu, L.; Kaur, R. Functional genomic analysis of Candida glabrata-macrophage interaction: Role of chromatin remodeling in virulence. PLoS Pathog. 2012, 8, e1002863. [Google Scholar] [CrossRef] [Green Version]
  241. Schmidt, K.H.; Viebranz, E.; Doerfler, L.; Lester, C.; Rubenstein, A. Formation of complex and unstable chromosomal translocations in yeast. PLoS ONE 2010, 5, e12007. [Google Scholar] [CrossRef] [Green Version]
  242. Schmidt, K.H.; Wu, J.; Kolodner, R.D. Control of Translocations between Highly Diverged Genes by Sgs1, the Saccharomyces cerevisiae Homolog of the Bloom’s Syndrome Protein. Mol. Cell. Biol. 2006, 26, 5406–5420. [Google Scholar] [CrossRef] [Green Version]
  243. Voter, A.F.; Qiu, Y.; Tippana, R.; Myong, S.; Keck, J.L. A guanine-flipping and sequestration mechanism for G-quadruplex unwinding by RecQ helicases. Nat. Commun. 2018, 9, 4201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Conserved domain structure of RecQ helicases from major model systems. RecQ helicases are conserved from bacteria to mammals. Proteins are aligned by their conserved helicase domains. The respective organism is shown on the left and the protein length in amino acids is indicated on the right. Human RECQ5α, RECQ5β and RECQ5γ are isoforms resulting from alternative splicing of the RECQ5 gene.
Figure 1. Conserved domain structure of RecQ helicases from major model systems. RecQ helicases are conserved from bacteria to mammals. Proteins are aligned by their conserved helicase domains. The respective organism is shown on the left and the protein length in amino acids is indicated on the right. Human RECQ5α, RECQ5β and RECQ5γ are isoforms resulting from alternative splicing of the RECQ5 gene.
Genes 11 00205 g001
Figure 2. Structure of Sgs1 and its binding partners. Sgs1 is composed of the structurally ordered ATPase/helicase core and HRDC domains in the C-terminal half of the protein and a 645-residue long intrinsically disordered N-terminal tail. This domain structure is conserved in human BLM and several other RecQ family helicases. In addition to SUMOylation at K621, SUMOylation at K175 and K831 has also been reported [54]. Pink boxes indicate phosphorylation sites. The helicase core contains the ATPase domain, consisting of two RecA-like lobes, 1A and 2A, which harbor the eight RecQ conserved helicase motifs indicated in black (motifs 0 to VI), and the RQC-domain, consisting of the zinc-binding and winged-helix subdomains. For interacting proteins, the binding region in Sgs1 is shown in brackets after the protein name, followed by the assay of detection. Interacting proteins for which binding sites on Sgs1 have not been narrowed down are listed in the gray box. SE, strand-exchange domain; SIMs, SUMO-interacting motifs; RQC, RecQ-C-terminal domain; HRDC, helicase-and-RNaseD-like-C-terminal domain; Y2H, yeast-two-hybrid assay; co-IP, co-immunoprecipitation; ITC, isothermal titration calorimetry.
Figure 2. Structure of Sgs1 and its binding partners. Sgs1 is composed of the structurally ordered ATPase/helicase core and HRDC domains in the C-terminal half of the protein and a 645-residue long intrinsically disordered N-terminal tail. This domain structure is conserved in human BLM and several other RecQ family helicases. In addition to SUMOylation at K621, SUMOylation at K175 and K831 has also been reported [54]. Pink boxes indicate phosphorylation sites. The helicase core contains the ATPase domain, consisting of two RecA-like lobes, 1A and 2A, which harbor the eight RecQ conserved helicase motifs indicated in black (motifs 0 to VI), and the RQC-domain, consisting of the zinc-binding and winged-helix subdomains. For interacting proteins, the binding region in Sgs1 is shown in brackets after the protein name, followed by the assay of detection. Interacting proteins for which binding sites on Sgs1 have not been narrowed down are listed in the gray box. SE, strand-exchange domain; SIMs, SUMO-interacting motifs; RQC, RecQ-C-terminal domain; HRDC, helicase-and-RNaseD-like-C-terminal domain; Y2H, yeast-two-hybrid assay; co-IP, co-immunoprecipitation; ITC, isothermal titration calorimetry.
Genes 11 00205 g002
Figure 3. Structure of Rqh1. Rqh1 shares the ATPase/helicase, RQC and HRDC domains with Sgs1 and other RecQ helicases [78]. Red boxes indicate predicted SUMOylation sites with residues 724-727 (PKKD) being the predominant site. Pink boxes represent phosphorylation sites [76,77].
Figure 3. Structure of Rqh1. Rqh1 shares the ATPase/helicase, RQC and HRDC domains with Sgs1 and other RecQ helicases [78]. Red boxes indicate predicted SUMOylation sites with residues 724-727 (PKKD) being the predominant site. Pink boxes represent phosphorylation sites [76,77].
Genes 11 00205 g003
Figure 4. DNA substrates that Sgs1 binds and unwinds. Sgs1 preferentially unwinds Holliday junctions in vitro and G4 quadruplex DNA with more efficiency than duplex DNA. It can displace the D-loop formed during strand invasion and also unwind forked and ssDNA overhang structures.
Figure 4. DNA substrates that Sgs1 binds and unwinds. Sgs1 preferentially unwinds Holliday junctions in vitro and G4 quadruplex DNA with more efficiency than duplex DNA. It can displace the D-loop formed during strand invasion and also unwind forked and ssDNA overhang structures.
Genes 11 00205 g004
Figure 5. Genetic interactions of SGS1. Negative genetic interactions of the SGS1 deletion mutation sgs1Δ are shown in red. Mutants that are rescued by sgs1Δ are shown in green, and mutations that rescue the sgs1Δ mutant are shown in blue. Orange, red, green and gray circles group genes involved in DNA repair, DNA damage and replication stress checkpoints, replication, and ‘other’ functions, respectively. Genetic interactions are primarily based on growth rates, hypersensitivity to DNA-damaging agents, DNA resection rates, mutation frequencies and checkpoint activation.
Figure 5. Genetic interactions of SGS1. Negative genetic interactions of the SGS1 deletion mutation sgs1Δ are shown in red. Mutants that are rescued by sgs1Δ are shown in green, and mutations that rescue the sgs1Δ mutant are shown in blue. Orange, red, green and gray circles group genes involved in DNA repair, DNA damage and replication stress checkpoints, replication, and ‘other’ functions, respectively. Genetic interactions are primarily based on growth rates, hypersensitivity to DNA-damaging agents, DNA resection rates, mutation frequencies and checkpoint activation.
Genes 11 00205 g005
Figure 6. Early and late roles for Sgs1 in DSB repair by homologous recombination. In addition to long-range resection and double-Holliday-junction dissolution, Sgs1-Top3-Rmi1 has also been implicated in D-loop reversal. Genetic evidence suggests that the Sgs1-Rad51 interaction promotes homologous recombination; a model has been proposed wherein the acidic region of Sgs1 acts as a DNA mimic that can compete with ssDNA for RPA binding, thereby facilitating initial loading of Rad51 [57]. Shu complex proteins promote Rad51 filament formation by antagonizing the antirecombinase Srs2 and by interacting with Rad51-Rad55-Rad57 [148,149,150].
Figure 6. Early and late roles for Sgs1 in DSB repair by homologous recombination. In addition to long-range resection and double-Holliday-junction dissolution, Sgs1-Top3-Rmi1 has also been implicated in D-loop reversal. Genetic evidence suggests that the Sgs1-Rad51 interaction promotes homologous recombination; a model has been proposed wherein the acidic region of Sgs1 acts as a DNA mimic that can compete with ssDNA for RPA binding, thereby facilitating initial loading of Rad51 [57]. Shu complex proteins promote Rad51 filament formation by antagonizing the antirecombinase Srs2 and by interacting with Rad51-Rad55-Rad57 [148,149,150].
Genes 11 00205 g006
Table 1. Physical interactions of Sgs1.
Table 1. Physical interactions of Sgs1.
Binding PartnerName DescriptionAssayProtein FunctionSgs1-Interacting DomainReferences
Bud27Bud site selectionTwo-hybridTOR-dependent gene expression421–791[97]
Dia2 aDigs Into AgarAffinity capture-westernComponent of SCF E3 ubiquitin ligase complexFull length[98]
Dna2DNA synthesis defectiveCoIPDNA-dependent ATPase, helicase & nucleaseFull length[51]
Gis1Glg1-2 SuppressorY2HHistone demethylase & transcription factor420–791[98]
Mlh1MutL HomologY2HDNA mismatch repair784–1447[95]
Mlh3MutL HomologCoIPDNA mismatch repairFull length[53]
Mre11Meiotic REcombinationCoIP, Y2HDSBR; nuclease subunit of MRX Full length[50]
Prp45Pre-mRNA processingTwo-hybridPre-mRNA splicing503–739[99]
Rad16RADiation sensitiveY2HNucleotide excision repair421–792[93]
Rad51RADiation sensitiveY2H, pull-downDSBR; strand exchange protein1187–1318[58,91]
Rad53RADiation sensitiveITCDNA damage response kinase446–456[60]
Rpa70/Rfa1Replication Factor AY2H, affinity capture-western, reconstituted complexSubunit of heterotrimeric RPA (ssDNA binding protein)421–792[58,100]
Rio1RIght Open reading frameY2H, CoIPSerine kinase involved in cell cycle regulation and rDNA integrityFull length[101,102]
Rmi1RecQ Mediated genome InstabilityCoIPDSBR; subunit of Sgs1-Top3-Rmi1 complex1–100[87]
Rtt107/Esc4Regulator of Ty1 TranspositionY2HDNA repair during S phase; recruits Smc5/6 to DSBsFull length[103]
Smt3Suppressor of Mif TwoY2HUbiquitin like proteinFull length[104,105,106]
Srs2Suppressor of Rad Six DNA helicase; antirecombinase422–722[92]
Stu2Suppressor of TUbulinY2HMicrotubule associated proteinFull length[107]
Top2TOPoisomeraseY2HRelaxes both positively & negatively supercoiled DNA466–746[20,86,94]
Top3TOPoisomeraseY2HRelaxes negatively supercoiled DNA, subunit of Sgs1-Top3-Rmi1 complex1–282[19,71,85,86]
Ubc9UBiquitin-ConjugatingY2HSUMO-conjugating enzymeFull length[56]
a interaction observed during HU/MMS exposure.
Table 2. Physical interactions of Rqh1.
Table 2. Physical interactions of Rqh1.
Binding PartnerName DescriptionAssayProtein DescriptionRqh1-Interacting DomainReference
CAF-1 a (subunit Pcf1)Chromatin assembly factor CoIPHistone chaperone promotes chromatin assembly during DNA repair and replicationFull length[108]
Top3Topoisomerase IIICoIPRelaxes negatively supercoiled DNA1–322[78]
RPA (Rad11)Replication protein ACoIPBinds to ssDNA Full length[109]
Cdc23MCM associated protein Mcm10Two-hybridEfficient phosphorylation of MCM complex and pre-RC activationFull length[110]
Nse2 bNon-SMC element SUMO ligase Biochemical activityComponent of Smc5-6 required for DNA damage responseFull length[111]
Cbh1CENP-B homologTwo-hybridPromotes Swi6 association with centromere causing increased silencingFull length[110]
Rcl1rRNA processing proteinTwo-hybridNuclease for 18S rRNA productionFull length[110]
Rmi1RecQ mediated genome instability proteinTwo-hybridHolliday junction dissolutionFull length[110]
Usp104U1 snRNP-associated proteinTwo-hybridSplicing factorFull length[110]
Pfh1PiF1 Helicase homologAffinity capture-MS5′-3′ DNA helicase promotes fork progressionFull length[112]
Spo7Sporulation proteinTwo-hybridmeiotic spindle pole body componentFull length[110]
Atg11Autophagy associated proteinTwo-hybridScaffold protein in mitophagyFull length[110]
Pli1 bSUMO E3 ligaseBiochemical activityMajor SUMO ligase, role in telomere maintenanceFull length[111]
a protects recombination intermediates from disassembly by Rqh1 b can both SUMOylate Rqh1.
Table 3. Comparison of budding yeast Sgs1 and fission yeast Rqh1 properties.
Table 3. Comparison of budding yeast Sgs1 and fission yeast Rqh1 properties.
Property aSgs1Rqh1
ATPase/Helicase activityYesYes
PTMs bPhosphorylation, SUMOylationPhosphorylation, SUMOylation
Nucleolar localizationYesYes
Localization in nuclear fociYesYes c(NLS:1294YSRKRKYSTS1303)
Homologous Recombination YesYes
Suppression of homeologous recombinationYesn.d. d
Suppression of GCRs eYesYes (only between gene duplications)
Maintenance of normal lifespanYesYes
Suppression of meiotic defectsYesYes
MeiosisYesYes
Meiotic recombinationYesNo
Activation of intra-S phase checkpoint fYesYes
Stabilization of replicative polymerases at stalled forksYesn.d.
Telomere maintenance gYesYes
rDNA maintenanceYesYes
Resistance to DNA damaging agentsMMS, HU, CPT h, IR, H2O2, Cisplatin, MMC iMMS, HU, CPT, MMC, IR, UV
SubstratesssDNA, dsDNA, four-way junctions, forked and G4 DNAdsDNA
a Functions and properties of Sgs1 and Rqh1 are from references [2,120,137,138,139,140,141] b post-translational modifications c predicted using cNLS Mapper d n.d., not determined e GCR, gross chromosomal rearrangement f observed in presence of DNA damage g in the absence of telomerase activity h CPT, camptothecin, topoisomerase I inhibitor i MMC, mitomycin C, DNA inter-strand crosslinking agent.

Share and Cite

MDPI and ACS Style

Gupta, S.V.; Schmidt, K.H. Maintenance of Yeast Genome Integrity by RecQ Family DNA Helicases. Genes 2020, 11, 205. https://doi.org/10.3390/genes11020205

AMA Style

Gupta SV, Schmidt KH. Maintenance of Yeast Genome Integrity by RecQ Family DNA Helicases. Genes. 2020; 11(2):205. https://doi.org/10.3390/genes11020205

Chicago/Turabian Style

Gupta, Sonia Vidushi, and Kristina Hildegard Schmidt. 2020. "Maintenance of Yeast Genome Integrity by RecQ Family DNA Helicases" Genes 11, no. 2: 205. https://doi.org/10.3390/genes11020205

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop