Next Article in Journal
MCAM/MUC18/CD146 as a Multifaceted Warning Marker of Melanoma Progression in Liquid Biopsy
Next Article in Special Issue
Focal Point of Fanconi Anemia Signaling
Previous Article in Journal
A Systematic Approach to the Development of Cilostazol Nanosuspension by Liquid Antisolvent Precipitation (LASP) and Its Combination with Ultrasound
Previous Article in Special Issue
Mutagenic Consequences of Sublethal Cell Death Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

DNA Repair in Haploid Context

by
Loïs Mourrain
and
Guylain Boissonneault
*
Department of Biochemistry and Functional Genomics, Université de Sherbrooke, Sherbrooke, QC J1K 2R1, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(22), 12418; https://doi.org/10.3390/ijms222212418
Submission received: 1 October 2021 / Revised: 8 November 2021 / Accepted: 14 November 2021 / Published: 17 November 2021
(This article belongs to the Special Issue DNA Damage Response, an Update)

Abstract

:
DNA repair is a well-covered topic as alteration of genetic integrity underlies many pathological conditions and important transgenerational consequences. Surprisingly, the ploidy status is rarely considered although the presence of homologous chromosomes dramatically impacts the repair capacities of cells. This is especially important for the haploid gametes as they must transfer genetic information to the offspring. An understanding of the different mechanisms monitoring genetic integrity in this context is, therefore, essential as differences in repair pathways exist that differentiate the gamete’s role in transgenerational inheritance. Hence, the oocyte must have the most reliable repair capacity while sperm, produced in large numbers and from many differentiation steps, are expected to carry de novo variations. This review describes the main DNA repair pathways with a special emphasis on ploidy. Differences between Saccharomyces cerevisiae and Schizosaccharomyces pombe are especially useful to this aim as they can maintain a diploid and haploid life cycle respectively.

1. Introduction

Maintenance of genome integrity is a permanent cell challenge as both intra- and extracellular conditions can lead to chemical alterations of nucleotides or their sequence. Proper repair mechanisms have evolved so as to maintain a balance between maintenance of cellular function and adaptative processes improving fitness. For obvious reasons, germline cells must be especially proficient at this task as diversity must be transmitted while maintaining the gametes’ integrity through the many differentiation steps. In addition, anisogamy, whereby the two sexes produced highly different and specialized haploid gametes, enhances diversity. Indeed, while diploidy allows gene versions compensation and error-free repair, haploidy is expected to create more variation as there is no templated repair for homologous recombination. Ploidy is therefore a key element to consider for genome plasticity or stability. The following review discusses the DNA repair and genome maintenance processes linked to the ploidy state by comparing two unicellular eukaryotes models, yeasts Saccharomyces cerevisiae (S. cerevisiae) and Schizosaccharomyces pombe (S. pombe), while considering the case of the mammalian gametes.

2. General Consideration Regarding Ploidy States

Since the turn of the century, knowledge regarding the advantages of different ploidy states has been increasing thanks to the study of lower eukaryotes. Thus, fission and budding yeasts have proved to be valuable models because they, respectively, represent the impact of haploid and diploid genomes on evolution and genetic integrity. The ploidy choice for lifespan appears to be strongly related to the evolution and maintenance of genetic integrity [1,2,3,4]. It is well known that diploidy confers greater fidelity to replication of both genomic and mitochondrial DNA, in particular, to attenuate single-nucleotide mutations. However, large chromosomal rearrangements or deletions occur more frequently in the diploid budding yeast impacting cellular fitness. Although there is twice the amount of genetic material in diploid cells, therefore statistically twice as many possible mutations [5,6], observations show that there are only 1.4 times more mutations in diploid cells than in haploid cells [7]. Such a bias is perhaps linked to the faster replication in the haploid state allowing a much shorter repair time. Errors mainly involve AT or CG transversions in late replicating genes [8]. In addition to this mutation bias, acquisition of a recessive mutation that would clearly alter the cell survival or mutation load is reduced by diploidy because two such similar recessive mutations must happen over both alleles to alter the phenotype. However, such recessive mutations will be eliminated in the population, a process coined “heterozygote advantage”.
On the other hand, this would prevent diploid cells from positively adapting to a changing environment [9]. Diploid cells can nevertheless accumulate heterozygous mutations, creating a pool of sequence variants that can be useful in reaction to a changing environment, whereas haploid cells either die or survive. Larger chromosome rearrangements in the diploid cells should be explained by the homologous recombination between chromosomes, which represents the main difference from haploid cells. However, the deletion of RDH54, a protein necessary for recombination between homologous chromosomes, does not alter the mutation rate of diploid cells [7]. This observation may be reconciled when considering the fission yeast S. Pombe, which maintains a haploid life cycle. S. pombe seems to be protected against the instability of the haploid life cycle by spending most of its time in the G2 phase. Indeed, in the G2 phase, sister chromatids are present allowing error-free repair of Double Strand Break (DSB), Homologous Recombination (HR). Although the alternative to HR is the error-prone Non-Homologous End-Joining (NHEJ). S. pombe promotes HR through stalling at the G2 phase. Deletion of NHEJ repair protein (PKu70Sp or Lig4Sp) displays low sensitivity to DSB-inducing genotoxins, indicating that even outside the G2 phases, S. pombe may not rely on NHEJ for damage repair. However, these observations are in contrast to the higher mutation rate observed in haploid cells compared to diploids. For example, DNA microsatellites are 100 times more stable in diploid cells [10]. Another example is the higher rate of mutation found in mitochondrial DNA of haploid cells [7].
One potential explanation is the higher number of mitochondria in diploid cells while mitochondria often mate, mixing alleles versions, rapidly reducing heteroplasmia. In addition, polyploid cells evolve faster than haploids, in relation to chromosomal rearrangement and aberrant condition of polyploid cells [11]. It, therefore, appears that the difference in the amount of DNA between the ploidy states only partially explains the difference in the mutation rate. Although haploidy allows a faster evolution, this condition is protected by some elements preventing excessive genome instability. Similarly, anisogamy in mammals results in very numerous haploids spermatozoa versus a single diploid ovum before fertilization. As stated above, genetic stability of the ovum is provided by diploidy and a complete array of repair systems, while the haploid character of sperm and the condensed chromatin may provide the proper context for DNA repair errors at the origin of the well-known male bias in the transmission of de novo mutations. In that sense, the ovum stands as the guardian of heredity while spermatozoa are providers of diversity.

3. DNA Repair Systems Comparison between S. cerevisiae and Sc. pombe

3.1. Nucleotide Excision Repair

3.1.1. General Mechanism

Nucleotide excision repair (NER) resolves various substrates resulting in structural alterations of DNA helix (Figure 1) [12]. NER can be divided into two sub-pathways: transcription-coupled NER (TC-NER) and global-genome NER (GG-NER). Although GG-NER uses self-assessment of DNA damage, TC-NER uses hindrance of the RNA polymerase transit to detect bulky adducts (Figure 2). This transcription-coupled mechanism ensures the elimination of mutations during transcription by the so-called transcriptional mutagenesis. Whereas TC-NER is dedicated to transcribed strands of active genes, GG-NER detects DNA helix disturbing damages in both transcribed and non-transcribed strands [13].
TC-NER is triggered by RNA polymerase blockage. RNA polymerases are exceptionally processive, and cover between 50 up to 90% of the genome depending on the cell’s transcriptional state, providing a reliable scanning process for damage. In addition to bulky adducts, pol II arrest can result from a wide variety of potential hindrances (Figure 1) including nicks, gaps, abasic sites, collisions with the replication complex, stable DNA-RNA dimers, or non-B DNA [14]. CSBHs (Cockayne Syndrome B protein) is the first NER factor to be recruited. When the pol II stops, the CSBHs-pol II interaction becomes stronger and CSBHs wraps DNA around itself hence decreasing pol II-DNA interaction [15]. Then CSBHs recruits CSAHs (Cockayne Syndrome A protein) [16] at the lesion site allowing the docking of UVSSA-USP7Hs. The latter is a ubiquitin ligase complex that activates CSAHs, in contrast to CSBHs which is ubiquitinated and targeted for degradation [17]. Following repair, USP7Hs removes ubiquitination of all NER components so they can be recycled. The second step in TC-NER is done by the transcription factor TFIIHHs, which binds DNA downstream of pol IIHs and provides an “assessment” of the lesion [18].
GG-NER targets larger lesions, typically more than one nucleotide in size, as this dramatically changes the DNA helix thermodynamics [19]. XPC-Rad23bHs (XPCHs complex, Xeroderma Pigmentum C protein) is the first complex to bind the DNA lesion. The XPCHs complex may detect DNA lesions as they create a single strand local transition [20]. Other than the scanning mechanism, alternative recognition mechanisms by XPCHs do exist [21,22,23,24]. XPCHs SUMOylation, and re-ubiquitinoylation, are then required for GG-NER progression [25,26]. Once activated, XPCHs complex recruits TFIIHHs. TFIIHHs can help XPCHs to discriminate between lesion types since XPCHs is easily misled by a base mismatch [27].
TC-NER and GG-NER repair pathways converge once TFIIHHs is recruited. TFIIHHs core subunits XPBHs and XPDHs create a “repair bubble” of 20-30 nucleotides surrounding the lesion [28]. A second proofreading process involves XPAHs. XPAHs normally stimulates the helicase activity of TFIIHHs but can decrease this activity in the presence of a lesion. XPAHs recruits RECCI-XPFHs [29] a nuclease that will cleave 5′ of the lesion [30]. From its interaction with the TFIIHHs complex, XPGHs endonuclease can then cleaves 3′ of the lesion [31,32] and replaces XPCHs. XPGHs plays a structural role in complex with TFIIHHs but acquires endonuclease activity once ERCC1-XPFHs is recruited by XPAHs 5′ to the lesion. ERCC1-XPFHs is the first to nick DNA followed by XPGHs [33]. After removal of the damaged nucleotide, DNA polymerases can fill the gap, while ligase seals the nick.

3.1.2. NER in Yeast

As for mammals, NER in yeasts is divided into two sub-pathways (Figure 2). However, Sc. pombe NER is less documented as redundancy with UVDER (UV Damage Excision Repair, see below) leads to phenotypic confusion. In S. cerevisiae, TC-NER is triggered after RNA polymerase stops, activating either Rad26Sc (CSBHs homolog) or Rpd9Sc, an RNA pol IISc subunit-specific of budding yeast [34]. Rad26Sc deletion decreases mRNA synthesis, in agreement with its transcription promoting activity over direct damage repair (8-oxoG, 3-MeA, abasic site) [35]. This may, however, lead to “transcriptional mutagenesis”. If the lesion creates a major hindrance, Rad26Sc then initiates chromatin remodelling and the recruitment of NER partners. An alternative TC-NER mechanism involves RNA pol II subunit Rpb9Sc, which can also recruit the NER machinery. In Sc. pombe, Rhp26Sp (CSBHs homolog) deletion displays an important increase in UV sensitivity [36], demonstrating the prominent role of the TC-NER pathway in UV resistance. In S. cerevisiae, GG-NER starts, with the Rad23-Rad33Sc (Rad23bHs homologs) complex that can transit over DNA until a UV lesion is encountered. Rad4Sc (XPCHs homolog) is then recruited to initiate NER [37,38]. While TFIIHSc is already positioned at the lesion in TC-NER, a control step is needed in GG-NER. This control step uses RPASc (Replication protein A), which binds intact DNA, and Rad14Sc (XPAHs homolog) which binds damaged DNA [39]. Without such a control step, NER will abort before the final dual incision [40]. GG-NER in Sc. pombe uses two XPCHs homologs: Rhp41Sp and Rhp42Sp [41,42]. Whereas Rhp41Sp has a major role in Cyclo-Pyrimidine Dimers (CPDs) removal in both transcribed and non-transcribed genes, Rhp42Sp is only active over non-transcribed genes [41,42,43]. Following damage detection by either GG-NER or TC-NER, Rad2Sc/Rad13Sp (XPGHs homologs) then cleaves 3′ of the lesion, whereas Rad1Sc-Rad10Sc/Rhp16Sp-Swi10Sp (ERCC1-XPFHs homologs) cleave in 5′, respectively in S. cerevisiae and in Sc. pombe [44,45,46].

3.2. Base Excision Repair

3.2.1. General Mechanism

Base excision repair (BER) is the DNA damage repair pathway in charge of resolving chromatin lesions not associated with structural alterations of the DNA helix [47]. BER mainly recognizes oxidized base (8-oxoG), deaminated base (uracil), and alkylated base (Figure 1). BER also takes charge of the abasic sites. Apuric or apyrimidic sites (AP) are indeed the most frequent lesions encountered in a stationary state [48,49]. BER pathway uses glycosylases which target abnormal bases and cleave β glycosidic bonds between riboses and bases. Glycosylases are generally evolutionary conserved, from bacteria to higher eukaryotes [50]. In mitochondria, BER also exists and uses different isoforms, resulting in the most robust DNA repair in mitochondria [51,52].
Glycosylases are generally small proteins harbouring a unique catalytic domain with high substrate specificity. Although most of these enzymes have unique glycosylase activity (monofunctional), some also have β lyase, or β/Δ lyase activities able to cleave in 3′ or 5′P, or both (bi- trifunctional) [51,52]. Glycosylases first bind the minor groove, inducing a kink in DNA, then flip the faulty base out of the helix and separate the base from the ribose [53]. Irrespective of their host organism, glycosylases are fast to catalyse the glycosidic bond cleavage, but then disengage from DNA with much slower kinetics [54,55]. This dual kinetic mode may therefore hide AP sites, which can lead to apoptosis when reaching a given threshold. If the glycosylase is monofunctional, AP site-specific endonuclease will assist in cleaving 3′ of the deoxyribose, and DNA pol will remove the 5′ deoxyribose phosphate (dRP) during DNA neosynthesis [51,56,57]. On the other hand, if the glycosylase also possesses a lyase activity, this will result in DNA nicking 5′ to the damaged deoxyribose. The 3′ dRP is then processed by AP endonuclease, leaving a functional 3′OH, the initiation site for DNA neosynthesis, and a 5′-diphosphate (5′-PP) thereby facilitating DNA pol binding (Figure 2) [58]. BER comprises two sub pathways: short and long patch repair. Either short or long patch repair pathways are used by the cell depending on the lesion size or the glycosylase involved. One single nucleotide is excised in short patch repair whereas long patch BER involves two or more nucleotides [59]. The main BER pathway is usually the short patch as it does not require the replication machinery typical of long patch repair. The short patch process uses glycosylase, the AP endonuclease, DNA polymerase, and DNA ligase. Long patch repair is mainly used during replication and uses a glycosylase and the AP endonuclease, but usually relies on DNA polymerase, PCNA (Proliferating Cell Nuclear Antigen), flap endonuclease from the replication machinery, as well as DNA ligase [51]. In a long patch repair pathway, the basic catalytic pocket of DNA polymerase can contain some ssDNA while the repair synthesis bypasses the lesion site creating a 5′ overhang tail with the old DNA. The Flap endonuclease (mammalian FEN1Hs, S. cerevisiae Rad27Sc, Sc. pombe Rad2Sp), which possesses a 5′flap endonuclease activity and 5′ to 3′ exonuclease activity, will then remove this 5′ tail [60].

3.2.2. BER in Yeast

Six DNA glycosylases have been characterized in S. cerevisiae: monofunctionals, Ung1Sc and Mag1Sc, Mag2Sc; and bifunctionals, Ngt1Sc, Ngt2Sc, and Ogg1Sc. In Sc. Pombe, monofunctionals, Ung1Sp, Thp1Sp, Mag1Sp, Mag2Sp, and Myh1Sp, and the bifunctional, Nth1Sp have been characterized (Table 1). Ung1Sc, Ung1Sp, and Thp1Sp are Uracil DNA Glycosylases (UDG) recognising uracil resulting from cytosine deamination [61]. Thp1Sp also recognizes T:G or U:G mismatches, but is also known to process 5-fluorouracil, 3,N4-ethanocytosine and 5-hydroxyuracil [62]. In addition, Thp1Sp recognizes deaminated purines (xanthine and oxanine from guanine, hypoxanthine from adenine) [63]. S. cerevisiae’s Mag1Sc and Mag2Sc are expressed after exposure to alkylating agents [64]. Mag1Sc is also able to cleave normal bases, usually guanine [65]. In Sc. Pombe, Mag1Sp and Mag2Sp share 44,8% identity. Although Mag1Sp initiates the alkylation repair, Mag2Sp may also plays a role in the process [66]. Unlike S. cerevisiae Mag1Sc, deletion of Mag1Sp only slightly increases alkylating agents’ sensitivity [67]. However, Rad13Sp, Rad16Sp or Rhp51Sp deletion mutants are hypersensitive to alkylation, suggesting an important contribution of NER or HR to this type of damage in Sc. pombe [67,68]. Although, mutants of both Sc. pombe BER endonucleases, Apn2Sp or Nth1Sp, are hypersensitive to alkylation, confirming BER in alkylation repair [69,70,71]. Surprisingly, Mag1Sp deletion leads to a 3-fold increase in intrachromosomic recombinations [68], suggesting that HR is somehow involved in alkylation repair in Sc. pombe. In S. cerevisiae Ntg1Sc is a bifunctional glycosylase located in the mitochondrion whereas Ntg2Sc is nucleolar. However, only Ntg1Sc is upregulated during oxidative stress [72]. These enzymes can process a wide variety of the oxidized pyrimidine and purine lesions [73]. However, Ntg1Sc is the sole enzyme able to resolve 8-oxoG whereas Ogg1Sc processes only Fapy-G and 7,8 dihydro-8-oxoG [74]. Surprisingly, an Ogg1Sc homolog is not found in Sc. Pombe [75]. Instead, Nth1Sp is in charge of resolving 8-oxoG in a mismatch situation. Nth1Sp seems to process a wide variety of substrates including 8-hydroxycytosine, thimidineglycol, and 8-hydroxyuracil [76].
Due to its bifunctional activity, Nth1Sp acts upstream of the Apn2Sp AP endonuclease, as it leaves a 3′ protected end, which is removed by the phosphoesterase activity of Apn2Sp [69]. In fission yeast, Nth1Sp also resolves AP sites resulting from the activity of both Mag1Sp and Mag2Sp, therefore playing a central role in BER [71]. However, in contrast to mammals NTHL1Hs, or S. cerevisiae Ntg1Sc, Nth1Sp is seemingly absent from mitochondrion. Finally, Myh1Sp, is a glycosylase specific to Sc. pombe and in charge of mismatches found at adenines (A:G, A:8-oxoG, 2-aminopurine:G, 2-aminopurine:A) [77]. Myh1Sp may also process G:8-oxoG, preventing C:G to G:C transversion [78]. A prominent role for Myh1Sp is binding to PCNASp and the PCNA-like heterotrimer Rad9Sp/Rad10Sp/Hus1Sp [79,80], which are involved in DNA lesion checkpoint. After base eviction, the AP site is processed by the two AP endonucleases found in yeasts. Apn1Sc and Apn2Sc are the two AP endonucleases found in S. cerevisiae although Apn1Sc is responsible for >95% of the total in vivo activity [81]. Apn1Sc cleaves 5′ to the lesion and possess a 3′ phospho-esterase activity able to resolve 3′dRP, 3′ phospho-glycolate (3′PGA) or 3′P 3′ dirty ends. Both S. cerevisiae AP endonucleases are 3′ tyrosyl DNA phospho-esterase with a weak 3′ exonuclease activity so they can resolve Topo1Sc-DNA adducts [82]. In Sc. pombe, Apn2Sp is the major AP endonuclease. Unlike Apn2Sp, Apn1Sp does not need specific subdomains to process PCNASp or Top3Sp catalytic hindrance, Apn1Sp may represent a proper backup to Apn2Sp at 3′ blocked end, but to a limited extent as Apn1Sp is present in both nucleus and cytoplasm, unlike Apn2Sp [83]. Uve1Sp, an UVDER enzyme, may also be involved in the processing of AP sites [84], especially as it is present in both the nucleus and mitochondrion [85].

3.3. Mismatch Repair

3.3.1. General Mechanism

DNA polymerases can induce from 10−5 to ~10−4 mismatch per base given their own exonucleolytic proof-reading activity. The mismatch repair (MMR) system increases repair fidelity by about ~100–1000 times [86]. The MMR pathway is in charge of mismatches but also resolves intra-strand loops. The MMR pathway was first described in Escherichia coli (E. coli) which represented a key model to decipher the sequential events of MMR: MutS first detects the mismatch, MutL confirms the helix deformation induced by the mismatch and MutH endonuclease cleaves the faulty base. DNA resection then occurs, followed by DNA polymerase gap filing and nick sealing by DNA ligase (Figure 2). While eukaryotic MutS and MutL exist as heterodimers [87,88], the MMR-specific MutH endonuclease is seemingly absent in eukaryotes and some bacteria. Lesion recognition by MMR is achieved by two heterodimeric homologs: MutSα and MutSβ. Whereas MutSα is in charge of small mismatches and helix extrusions of about 1–4 nt in size, MutSβ exclusively recognizes larger extrusions of 2–10 nt [89]. Loss of MutSβ can easily be compensated for by MutSα. Lesion recognition by MutSα is indeed prominent in the MMR pathway [90]. Once the MutS complex becomes bound to DNA, it recruits MutLα or MutLβ. During recombination, MMR proteins act to repair mismatch in DNA heteroduplexes, remove 3′ non-complementary tails produced during resection, and prevent recombination between divergent sequences. MutLγ and MutLβ have been shown to play a role in meiotic HR [91]. MutLγ is mainly used during meiotic HR, where MutSγ activates it. In the mismatch repair, the endonucleolytic cleavage is done over the weakest strand, admitting the one containing a mismatch, in both 3′ and 5′ flanking the lesion. The mechanism directing strand selection is still unknown but may involve PCNA, which is still bound to the neosynthetised strand or may result from hypomethylation of the new strand [92]. Although a eukaryotic MutH has yet to be identified, the human PMS2Hs, part of MutLαHs, have been shown to possess a latent endonuclease activity. Removing its endonuclease motif abolishes the mitotic function of MutLαHs [93]. After incision, a fragment of DNA is excised by Exo1Hs. Resection by Exo1Hs is greatly enhanced by association with MutSαHs and ATP [94], as it removes only a few nucleotides when acting independently. After resection, DNA polymerase fills in the gap while DNA ligase ligates the 3′ end to complete the process. An alternative pathway to MMR involved DNA polymerase, whereby the faulty strand is displaced during DNA synthesis [95].

3.3.2. MMR in Yeast

In S. cerevisiae, while MutSαSc targets mismatches, MutSβSc is mainly in charge of long insertion-deletion loops (IDLs), although redundancy exists between both as MutSβSc may substitute for defective MutSαSc but to a limited extent [96]. MutSαSc initiates repair of all mismatches except C:C [97]. IDLs are repaired more efficiently than mismatches as MutS functional redundancy facilitates the repair of small IDLs [98,99]. Sc. pombe possesses four MutS homologs (Msh1Sp, Msh2Sp, Swi4Sp, Msh6Sp) [100]. MutSαSp (Msh2Sp-Msh6Sp) is sufficient to support Sc. pombe MMR, regardless of the lesion type. However, Swi4Sp, likely acting in complex with Msh2Sp, is involved in mating-type switching and some recombination processes, as demonstrated for MutSβSc in S. cerevisiae [101,102]. Nevertheless, although Sc. pombe MutSβSp is solely implicated in recombination, the MMR function, typical of the S. cerevisiae MutSβSc, has apparently been lost. The MutSγSc complex in S. cerevisiae is only expressed during meiosis and is linked to recombination [103,104]. In yeast, MutLγSc is the main Holliday Junctions resolvase [105], symmetrically nicking Holiday Junctions. MutL homologs in budding yeast includes MutLαSc (Mlh1Sc-Pms1Sc), MutLβSc (Mlh1Sc-Mlh2Sc) and MutLγSc (Mlh1Sc-Mlh3Sc) [106]. While MutLαSc interacts with MutSαSc and MutSβSc to process most of MMR, MutLγSc only interacts with MutSγSc to resolve meiotic crossing-over. Sc. pombe contains only one MutL homolog (Mlh1Sp-Pms1Sp) [100]. Thus, MutSαSp and MutLαSp should be sufficient for a functional MMR machinery in fission yeast, as it seems to be the case in Drosophila and Caenorhabditis. In S. cerevisiae, Exo1Sc mutants were found to be slightly mutagenic, owing to its “error-free damage bypass” function, as described by Tan et al. [107]. Considering the weak impact of Exo1Sc deletion, one may surmise that other nucleases may be involved such as Rad27Sc or from the proofreading activity of DNA polymerases Δ or εSc [107,108]. In contrast, Exo1Sp deletion in Sc. pombe generates a strong mutator phenotype, probably as a result of its role in GT repeat maintenance in an MMR-independent pathway [109].
A short patch MMR pathway is described in fission yeast and involves the NER machinery [110]. Indeed, Rhp14Sp, Swi10Sp, Rad13Sp, Rad15Sp, and Rad16Sp are important NER effectors components in this short patch repair pathway [111,112,113]. The NER machinery may therefore play a broader role and supports MMR pathways in specific cases. The first clear evidence is the NER requirement for C:C mismatches, which cannot be repaired by MMR [111,113]. Hence, in the absence of MMR, C:C mismatches are quickly repaired, so NER seemingly processes such damage with high efficiency. Interestingly, the NER Rhp14Sp binds to various mismatches with similar affinity to C:C mismatches [112]. Defective NER has also been shown to induce a slight increase in reversion of GT repeats, suggesting a role for NER in Sc. pombe microsatellites maintenance [112]. Short patch MMR exists in S. cerevisiae but is seemingly independent of NER [114]. In S. cerevisiae, NER mutants show an increase in reversion rate but are mainly associated with flanking nucleotides substitutions, suggesting a more complex mechanism, likely originating from a bypass by the error-prone DNA pol ζ [115]. In vitro evidence indicated that NER may also play a role in mismatch repair in other species including drosophila [116,117], E. coli, and humans [118].

3.4. Double Strand Breaks (DSBs) Repair Pathways Selection

3.4.1. DSB Regulation in Mammals

In humans, about 1 double-strand break (DSB) arises in 108 base pairs, so around 10–30 DSBs are generated per cell, per cell cycle, resulting from endogenous or exogenous factors [119,120]. The two main DSBs repair pathways are Homologous Recombination (HR) and Non-Homologous End-Joining (NHEJ). These pathways are quite contrasted in terms of quality and timing of repair during the cell cycle. HR relies on sister chromatids to repair DNA, providing an almost error-free mechanism. HR is therefore dependent on the cell cycle, being most active during S-G2 phases. In contrast, NHEJ is an error-prone system, inducing sequence variations, and is mainly active during G1-G2 phases, although its components are found during the whole cell cycle [121,122]. Whereas HR needs hours to complete, NHEJ can repair DSBs within half an hour given the high processivity of its enzymes [123]. Upon inactivation of NHEJ (mutation of Ku, Lig4, or XRCC4 proteins), HR becomes the prominent DSBs repair pathway [124]. If sequence homologies however exist at DSBs ends, other repair pathways may be used such as Single Strand Annealing (SSA), Micro-homology End-Joining (MMEJ), and an Alternative NHEJ (Alt-NHEJ) pathway. SSA requires at least a repetition of two nucleotides to be present at both ends of the DSB. However, the two-nucleotides homology can also be used by MMEJ or alt-NHEJ. MMEJ is frequently used even though HR proteins are present or in cases of defective HR [125]. A choice must be made between NHEJ, HR, SSA, MMEJ, and alt-NEHJ (Figure 2). Initially, the DSB may undergo a direct repair by NHEJ, or resection may occur. If resection is chosen, then a competition takes place between HR, SSA, MMEJ, and alt-NHEJ depending on resection size [126,127,128]. Resection is inhibited either by the NHEJ KuHs binding at DNA ends or by both the p53 binding protein 1 (53BP1Hs) or the Replication timing regulatory Factor 1 (RIF1Hs), which are known to be resection antagonists [129]. Many such down regulators of resection lead to the predominance of NHEJ. If a resection still occurs, a competition between other repair systems begins. The shift between HR and SSA, MMEJ, and Alt-NHEJ is dependent on the cell cycle, and particularly the cyclin-dependent kinases (CDKHs). Indeed, during S or G2 phases, CDKHs can phosphorylate CtIPHs (C-terminal-binding protein-interacting-protein) thereby facilitating its interaction with BRCA1Hs to promote HR [130]. In the case of resection, it is done before the S phase and DSB repair will be achieved by SSA or MMEJ [131]. Although most of the MMEJ promoting factors also promote SSA repair [132], SSA will be preferred for smaller resection sizes. Alternative NHEJ also needs a micro-homology of 2–20 nt [133].

3.4.2. DSB Regulation in Yeast

Yeast regulation of DSBs repair is relatively similar to that of mammals. However, in this case, the presence of sister chromatid is not the main factor in choosing between HR and NHEJ. Indeed, HR is regulated through the cell cycle by post-translational modifications, degradation, and re-localization of its effectors. In S. cerevisiae, some HR genes are even regulated by the presence of DNA damage itself [134,135]. In contrast, YKu70Sc is mainly transcribed in early G1 and almost undetectable in the S phase. In post-replicative cells, HR proteins are downregulated by Clb-CDK-inhibitorsSc or Sic1Sc overexpression [136]. A few minutes after DSB formation, YKuSc heterodimers and MRX complexes become recruited independently [137]. MRX complexes are essential to maintain both DNA ends physically closed and protected [138], whereas YKuSc inhibits DNA end resection [139]. As in mammals, HR-NHEJ balance is determined by resection initiation, with YKuSc competing with MRX complex and Sae2Sc endonuclease (CtIPHs homolog) [140]. Moreover, MRX is ambivalent regarding YKuSc. It seems to stabilise YKuSc on DSBs, but also destabilising it [141]. In addition, YKuSc dissociation from unrepaired DSBs is dependent on the formation of the MRX complex, suggesting that the latter can complete the NHEJ repair process [137]. However, YKuSc is able to prevent extensive resections in the absence of either the MRX complex or Sae2Sc, while the MRX complex is dispensable to initiate HR in the absence of YKuSc since Exo1Sc will carry out resection [139]. In the particular case of the fission yeast, despite being a haploid organism, it spends most of its time in the G2 phase, where sister chromatids are present allowing preferential selection of HR. Thus, in contrast to mammals, HR is the major DSBs repair pathway in yeast. Unlike NHEJ mutants (PKu70Sp or lig4Sp mutants), HR mutants were shown to be much more sensitive to genotoxins [142]. However, in G1 synchronised cells, NHEJ is 7 to 10 times more important than in asynchronous cells, and PKu70Sp mutants become far more sensitive to ionising radiations [143]. This demonstrates the primary role of NHEJ in G1 phasis when sister chromatids are absent although HR remains the predominant DSBs repair pathway in yeast. However, loss of PKu70Sp increases chromosomal rearrangements from DSBs without changing the gene conversion ratio, suggesting competition between HR and NHEJ despite the lack of PKu70Sp [144]. In contrast to S. cerevisiae, MRXSp mutation (deletion of Rad32Sp or Rad50Sp) does not alter NHEJ activity [145]. Moreover, Rad32-Rad50Sp double mutants are unable to proceed beyond the S-phase DNA damage checkpoint [146] showing a clear prevalence of HR. In S. cerevisiae, some proteins become expressed in order to promote HR. The rad55-Rad57Sc complex is also found to promote HR in S. cerevisiae [147]. In S. pombe, two Rad52Sc homologs are known: Rad22Sp and Rti1Sp. Rad22Sp is required for mating-type switching and DNA repair [148] and shares high sequence homology with Rad52Hs [143]. By itself, Rad22Sp is able to bind DSB. Rti1Sp mutant increases sensitivity to ionising radiations as much as in Rti1/Rad22Sp [149,150] double mutants. Thus, although NHEJ, SSA, MMEJ, and Alt-NHEJ do exist in yeast, HR remains the major repair mechanism regardless of the ploidy state. Details of DSB repair processes are provided in the following sections.

3.5. Homologous Recombination

3.5.1. General Mechanism

The first step in HR is the formation of a single strand 3′ tail by resection from the original DSB (Figure 3). This initial step is mediated by the Mre11Hs-Rad50Hs-NBS1Hs (MRNHs) complex and CtIPHs in humans. The 3′ tail resulting from large 5′->3′ DNA resection is performed by Exo1Hs [151]. The main characteristic of HR is the single strand invasion over the sister chromatid. Rad51Hs is the RecA bacterial orthologue in eukaryotes with a DNA-dependent ATPase activity able to form a nucleoprotein-DNA filament [152]. HR proteins form supersized structures containing hundreds to thousands of proteins. Rad51Hs may bind B-DNA formed by trinucleotide repeats in a single strand 3′ tail. These triplets could then form Watson-Crick interactions with complementary dsDNA present in sister chromatids. After ATP hydrolysis, Rad51Hs releases a stabilized triplex DNA. By competing with Rad51Hs, RPAHs (Replication Protein A) binds to single-stranded DNA with high affinity providing protection against nucleases [153]. Strand invasion can lead to crossover resolution and or to non-crossover resolution (patch) sequence exchange at Holiday junctions [154].

3.5.2. HR in Yeast

As in mammals, HR is initiated by resection, which is a two steps process. Initially, a 50–200 nt resection is performed by the MRXSc (Mre11Sc/Rad50Sc/Xrs1Sc) complex and Sae2Sc in S. cerevisiae, and MRNSp (Mre11/Rad50/Nbs1Sp) complex and Ctp1Sp in Sc. pombe, which are respectively MRNHs and CtIPHs homologs [155,156]. Then, resection extends to kilobases through the action of Exo1Sc or Sgs1Sc-Top3Sc-Rmi1Sc (STRSc) complex with Dna2Sc endonuclease in S. cerevisiae [157]. Exo1Sp is likely the HR exonuclease in Sc. pombe, but Exo1Sp deletion does not increase sensitivity to ionizing radiation, suggesting functional redundancy. Exo1Sp-Rad50Sp double mutant displays enhanced sensitivity than individual mutant, showing distinct activities [156]. After resection, single-strand tails are coated by RPA homologs (RPASc in S. cerevisiae and Rad11Sp in Sc. pombe) in order to protect them and prevent untimely Rad51Hs homologs (Rad51Sc in S. cerevisiae and Rhp51Sp in Sc. pombe) binding [158]. In Sc. pombe, deletion of the Rad11Sp gene is lethal, and the temperature-sensitive mutant (Rad11-404) shows high sensitivity to genotoxins [159]. In S. cerevisiae, Rad52Sc displaces some RPASc molecules to allow Rad51Sc binding to ssDNA and to initiate the nucleoprotein filament formation [160]. Rad52Sc can be assisted by Rad51Sc paralogues, including Rad55Sc-Rad57Sc, which has been shown to stabilize Rad51Sc filaments [161]. The ssDNA invasion by Rhp51Sp nucleofilaments, in Sc. pombe, results from a competition between Rhp51Sp and Rad11Sp for ssDNA binding. Deletion of the Rhp51Sp gene leads to much higher sensitivity to ionizing radiations [162]. Rph51Sp binds to ssDNA to form the nucleoprotein filaments in an ATP-dependent manner [163]. As in mammals, the Rad51Sc/Rhp51Sp filament promotes strand invasion to find a DNA duplex with sufficient homology to form a synapse. Although very complex, the process can take from 20 to 60 min to complete [164,165]. In Sc. pombe, HR is involved in processing stalled replication forks [166]. Not surprisingly, mutations of HR proteins lead to alteration of DNA replication. Moreover, Segurado et al. [167] have shown that DNA replication is associated with physical interactions and annealing between sister chromatids. Such structures are not found in Rad22Sp (Rad52Hs homolog), or Rhp51Sp mutants, clearly linking HR to replication. Finally, Rhp51Sp and Rad22Sp foci induced by fork stalling, result in chromosomal rearrangements from an increase in recombination activity [168].

3.6. Non-Homologous End-Joining

3.6.1. General Mechanism

In mammals, Non-Homologous End-Joining (NHEJ) is a DSBs repair pathway that allows direct ligation of DNA ends in an error-prone manner (Figure 3). NHEJ proceeds in three tightly regulated steps. First Ku70/80Hs heterodimers bind to free DNA ends and act as protein hubs [169]. DNA-PkcHs (DNA dependant protein kinase catalytic subunit) is recruited given its high affinity for the Ku-DNA complex [170]. DNA-PkcHs undergo an auto-phosphorylation that recruits and activates other repair partners [171]. The second step consists of DNA end processing by the ArtemisHs endonuclease [172]. A third step involves XRCC4Hs and XLFHs (XRCC4-Like factor or CernunnosHs), which interact together to form a small filament [173]. This allows interaction between the two DNA ends and the action of DNA ligase IVHs. These canonical NHEJ proteins are sufficient to repair some DSBs, but a number of DSBs require more proteins including nucleases, DNA polymerases, and some functional homologs of XRCC4Hs and XLFHs [174].

3.6.2. NHEJ in Yeast

NHEJ Ku70Hs, Ku80Hs, and lig IVHs homologs are present in both S. cerevisiae (Yku70/80Sc, Dnl4Sc) and Sc. pombe (Pku70/80Sp, Lig4Sp), but DNA-PKcsHs and ArtemisHs homologs proteins have yet to be identified. Whereas Sc. pombe only possesses an XRCC4Hs homolog, Xlf1Sp, S. cerevisiae possesses both XRCC4Hs (Nej1Sc) and XLFHs (Lif1Sc) homologs [175,176,177,178,179,180]. In yeasts, MRX is also related to NHEJ, as recently shown in Sc. pombe, where it seems to play a similar role as ArtemisHs and processes dirty ends before ligation. This function is enhanced by Sae2Sc/Ctp1Sp (CtIPHs homologs) [178,181,182].

3.7. Single-Strand Annealing

3.7.1. General Mechanism

Single Stranded Annealing (SSA) arises when a DSB is formed between two homologous regions. Once the repair is completed, the sequence between repeats is lost. Repeats must be separated by at least 400 bp and up to 15 kbp [183]. The first step in SSA depends on CtIPHs and Mre11Hs resection (Figure 3). This resection is around 100 nt in length and could be targeted by MMEJ [184]. Once generated, a 3′ single-strand tail is covered by RPAHs and Rad51Hs [185]. RPAHs is progressively displaced to allow Rad51Hs filaments formation. The scanning for strand homology is provided by the activity of Rad52Hs. After annealing, the two 3′ tails emerging from the hybridized region are targeted by ERCC1Hs-XPFHs for cleavage with the help of Rad52Hs [186]. Gaps between newly hybridized and the original dsDNA are then filled by a combination of DNA polymerase and ligase [133].

3.7.2. SSA in Yeast

S. cerevisiae and Sc. pombe SSA involves a similar sequence of events: DNA resection, annealing, 3′ overhang tail removal by Rad1Sc-Rad10Sc/Rad16Sp-Swi10Sp flap endonucleases (Figure 3). Sc. pombe SSA is functionally linked to Rad16Sp and Swi10Sp [187], as well as Rad22Sp but not to Rhp51Sp, suggesting a specific SSA pathway in this model. However, in Sc. pombe, mutation of Rad16Sp-Swi10Sp results in a very small increase in sensitivity to ionizing radiations [188,189], suggesting that SSA is not a major repair pathway in this experimental model.

3.8. Yeast Specific DNA Repair Mechanisms

3.8.1. UV Damage Excision Repair

UV-induced damages can also be resolved by an alternative pathway discovered in yeast Sc. pombe [190]. Since NER genes deletion, mutants of either Rad16-Swi10Sp, Rad13Sp, Rad15Sp, or Rad16Sp show unaltered sensitivity to UV, an alternative system has been proposed and termed “UV Damage Excision Repair” (UVDER) that relies on the UV Damage Endonuclease, UVDE. UvdeSp enzyme can perform 5′ incision just next to CPD, 6-4 photoproducts, or to Dewar isomers [191]. In contrast to photolyases or glycosylases, UvdeSp recognises both CPD and 6–4 photoproducts and may process other substrates, as it is compatible with a nuclease activity reported over apurinic/apyrimidinic sites. However, the deletion of UVDE shows a weak increase in sensitivity to chemical genotoxins [192,193]. In Sc. pombe, the Rad2Sp protein, a homolog of mammalian FEN1Hs or S. cerevisiae Rad27Sc, could be involved in a related NER alternative pathway. Indeed, Rad2Sp mutants are far more sensitive to UV than UvdeSp mutants whereas Uvde-Rad2Sp double mutants are less sensitive than single Rad2Sp mutants [194]. This is compatible with a mechanism whereby Rad2Sp could help to process UvdeSp nicked DNA, and where the lack of Rad2Sp leads to a potentially lethal accumulation of UvdeSp-lesions [195]. Whatever the activation mechanism, UVDER-induced lesions can be repaired by HR through the activity of both Rhp51Sp and Rad18Sp [196]. Moreover, it has been shown that Exo1Sp endonuclease may also process UvdeSp nicked intermediates, as Exo1Sp and Exo1-UvdeSp mutants have similar phenotypes [197].

3.8.2. Photolyases

In contrast to mammals and fission yeast, budding yeast possesses two unique DNA repair enzymes: Phr1Sc and Mgt1Sc. Phr1Sc is a DNA photolyase, or a so-called CPD-specific photolyase [198,199]. It is produced in response to UV-C, or to exposure to alkylating agents [200]. Phr1Sc is a monomeric protein with two prosthetic residues and uses near-UV photon (300–500 nm) to split off the cyclobutane ring of CPD, repairing DNA [201]. The second unique enzyme, Mgt1Sc, is a methyltransferase, targeting O6-MeG and O4-MeT [202]. O6-MeG can mismatch with thymine whereas O4-MeT mismatches with a guanidine, leading to G:C->A:T and to A:T->G:C transversions. In a suicide reaction, Mgt1Sc transfers the faulty methyl to its cysteine, leading to its inactivation [202,203]. Despite its repair abilities, Mgt1Sc has a higher affinity for O6-MeG than O4-MeT, resulting in a bias in repair kinetics.

3.9. Alternative DNA Repair Approaches in Yeast

Despite their evolutionary distance, NER repair systems are relatively similar between the two yeast models. However, in Sc. pombe, the presence of UVDER conceals certain phenotypes of NER proteins mutants, suggesting a strong functional redundancy. However, this is nuanced in the presence of UV, because of S. cerevisiae Phr1Sc photolyase which processes NER substrates and somewhat hides NER deficient phenotypes. The TC-NER sub-pathway is also favoured in S. cerevisiae by the Rpb9Sc subunit of RNA polymerase, allowing rapid identification of RNA pol II arrests leading to redundancy in TC-NER initiation. On the other hand, Sc. pombe has two homologs of XPCSp (Rhp41Sp and Rhp42Sp). Although their targets are slightly different, this redundancy seems to indicate the greater importance of GG-NER.
The case of the BER system is more difficult to compare. Although the general process is similar, specificity in glycosylases creates subtle differences in processes. Thus, even if Sc. pombe has two UDG, against one in S. cerevisiae, it does not have Ogg1Sc. Oddly enough, Ogg1Hs is in charge of the most common damage, 8-oxoG, in mammals, but this does not appear to be the case in budding yeast. This useful function is performed by Ntg1Sc/Ntg2Sc and by Nth1Sp, respectively in S. cerevisiae and Sc. pombe. The redundant NtgSc homologs in budding yeast are required because of their different subcellular localization and not to improve repair fidelity. Bifunctional glycosylases seem to be more efficient for cells as it reduces the number of steps in the BER process. Thus, S. cerevisiae has 3 bifunctional enzymes, compared to only one in fission yeast. However, Nth1Sp, the bifunctional enzyme in Sc. pombe, is a key component of BER, apparently acting upstream of AP endonucleases. Regarding alkylation, although Sc. pombe has two glycosylases dedicated to this task, Mag1Sc from S. cerevisiae seems much more efficient. Perhaps this is due to the prominent role of NER and HR in the repair of alkylation in fission yeast. However, in addition to having a more efficient enzyme, S. cerevisiae has a photolyase, Mgt1Sc, specific to this damage, and provides a more efficient BER activity in this model. Sc. pombe has one more enzyme, Myh1Sp, which is in charge of resolving mismatches in front of an adenine. This may be related to a simpler MMR mechanism in this model. Few differences seem to emerge regarding AP endonucleases, although abasic sites may require processing from activation of Uve1Sp, in the Sc. pombe UVDER system to complete BER more efficiently.
As outlined previously, the MMR system in fission yeast seems much simpler considering that MutSβSp has lost its function in MMR and that MutLαSp is the only homolog of MutL. This observation was also made in other species, such as Drosophila and Caenorhabditis, suggesting a lesser role for MMR in maintaining genomic integrity. On the other hand, Sc. pombe can rely on alternative pathways, for example using NER machinery. However, this is compensated for by the existence of short patches repair in S. cerevisiae, although it does not involve the machinery of NER.
The two yeasts respond to UV damage in different ways. Response to UV in S. cerevisiae is elegant because it uses the energy of the toxic UV, to repair the damage with Phr1Sc photolyase. This however is too specific to CPDs, as UV can induce other damage such as 6–4 photoproducts and Dewar isomers. Sc. pombe adapted by using the UVDER system which can recognize all these different substrates. UVDER is so efficient that it compensates for a defective NER and can support BER in the elimination of abasic sites.
The HR system is predominant in both yeasts and appears to work in similar ways. However, the STRSc complex confers a greater resection activity to S. cerevisiae. The filament also appears to be better stabilized by Rad51Sc homologs. In Sc. pombe, stalls of the replication fork are repaired by HR, so it ensures proper replication in this model.
The NHEJ systems are seemingly different between these two models. In S. cerevisiae, the MRXSc complex seems very important to NHEJ. Since MRXSc is known to direct resection, its role in NHEJ is rather unexpected. In S. cerevisiae, MRXSc is used mainly to process dirty DNA ends, as ArtemisHs does in mammals. This suggests that resection is important in S. cerevisiae for repairing DSBs regardless of the repair mechanism. Recently, MRNSp has been shown to influence the NHEJ system in Sc. pombe, albeit with limited action, as shown by mutants of Rad50Sp, which do not change NHEJ repair capacity. It is likely that the lesser dependence on MRNSp in Sc. pombe protects against extensive resection which, in a haploid cell, could be dangerous.
SSA systems of the two yeast models appear to also be different but have been studied to a much lesser extent. In fact, in Sc. pombe SSA was shown not to rely on Rhp51Sp, the core protein of HR. However, since SSA does not depend on the presence of sister chromatids, its regulation is somehow opposite to the activities of HR proteins. An HR-independent pathway, like SSA, should therefore be more valuable in the haploid fission yeast. Unfortunately, as the activation of SSA is related to the presence of MRNSp, but this protein is mostly expressed in the presence of sister chromatid, the moment where SSA competes with HR, SSA is relegated to a secondary system in fission yeast.

4. Repair in Human Germ Cells

4.1. Gametes

Most vertebrates are anisogametic whereby gametes from the opposite sex are specialized and differ in morphology and number. They are therefore expected to respond differently to DNA damage inducers. This is especially the case for mammalian gametes where spermatozoa are ex-soma cells produced in large numbers whereas the oocyte is a resident cell. One additional major difference is the much greater level of DNA compaction provided by protamines in mature sperm following the eviction of most histones whereas oocytes maintain histone-based chromatin.
In addition to the greater number of cell divisions required to produce mature sperm, the peculiar steps of the sperm differentiation program may be more vulnerable to both exogenous and endogenous DNA damage inducers. For instance, endogenous DNA breaks may arise from the change in chromatin structure in post-meiotic spermatids coincident with a decreased DNA repair activity of these cells, whereas oocytes possess a far greater DNA repair capacity and harbour much fewer DNA breaks.

4.1.1. Spermatozoa

Male gametogenesis is a complex differentiation program producing mature spermatozoa from spermatogonia. During meiosis, HR takes place during the late phase of prophase I. At the beginning of prophase I, programmed DSBs allow parental alleles to recombine [204]. After meiosis, spermiogenesis undergoes a Golgi phase whereby spermatozoa head and axoneme are formed. Then chromatin is remodelled with Transition Proteins (TP1 and TP2) followed by protamination where histones are replaced by protamines (PRM1 and PRM2) to achieve the highest level of DNA compaction known to the eukaryotic world.
In spermatozoa, DNA damages, strand breakage, or base alteration may result from (1) Faulty compaction/protamination of DNA [205]; (2) Abortive apoptosis, or “anastasis”, in late spermiogenesis [206]; (3) Oxidative stress [207]; (4) Persistence of enzymatic DNA breaks induce during chromatin remodelling. The contribution of each to sperm DNA damages is still unknown but it becomes clear that an elevated number of DNA breaks is associated with male infertility [208] or with impaired development of the early embryo. Chromatin remodelling in spermatids leads to the replacement of 90–95% of histones by protamines [209]. During the eviction of histones, it is assumed that free DNA supercoils are formed that can be eliminated by the action of topoisomerases [210]. Potential hindrance of topoisomerases catalytic cycle in this context may lead to SSBs in the case of topoisomerase I, or DSBs in the case of a topoisomerase II activity. A possibility also exists that mechanical breaks resulting from the major chromatin remodelling can also be created. Protamination itself can induce damage [211,212] as the PRM1/PRM2 ratio can impact proper DNA packaging [209,211,213]. We, however, generated evidence that DSBs were indeed generated during this process in the whole cell population and harbour a 3′ OH so is consistent with endonuclease digestion. In this haploid context, only direct end ligation repair mechanisms such as NHEJ are expected [214]. Although apoptosis was shown to maintain the germ cell vs Sertoli cell ratio [215,216], canonical apoptosis may not be responsible for the transient DNA breaks shown in spermatids as these are observed in 100% of spermatids. This could represent an “apoptosis-like” reversible mechanism that is reminiscent of “anastasis”, or the recovery from apoptosis described recently [217]. Various degrees of recovery, or repairs, would then modulate the persistence of DNA breaks in the mature sperm. At other spermatogenesis steps, DSBs could be linked to the altered balance of anti and pro-apoptotic factors through different steps of spermatogenesis [218] and to the compartmentalisation of the mitochondria which prevents nucleus-organelles exchanges. Reactive oxygen species (ROS) are a common source of DNA damage. ROS create abasic sites, base modifications, inter-strand crosslinking, and both single and double-strand breaks. However, ROS are essential for some sperm functions including capacitation or acrosome formation and are produced in the sperm mitochondria [219,220]. Sperm and white blood cells are the main sources of ROS in semen. Sperm cells are very sensitive to ROS because of their high content in unsaturated fatty acids and their weak DNA repair activity [207,219]. Antioxidants appear to balance a high level of ROS [221], for this purpose, seminal fluid possesses catalases, superoxide dismutases, and glutathione peroxidases and is rich in vitamins C, E, A, lactoferrin, and Q10. Ultimately, prostasomes can decrease the release of ROS from leukocytes [222]. The relevance of ROS in male infertility is emphasized as 20 to 88% of sub-fertile men show an elevated level of ROS [223].
So far, evidence of NER, BER, SSA, MMR, HR, and NHEJ DNA repair processes has been reported during spermatogenesis.
Important variations in NER activity were reported throughout spermatogenesis. For instance, spermatogonia were shown to be more sensitive to UV than meiotic or post-meiotic cells [224]. 6–4 photoproducts repair activity is also reduced significantly in post-meiotic round spermatids and during aging [224]. This may be linked to a decrease in NER protein expression [225]. Various levels of GG- or TC-NER also exist as 16.8% of CPD are removed from transcribed Scp1 gene within 16 h in rat spermatocytes [226], while 50% of CPD are removed from Dhfr and Dazl transcribed genes in spermatogonia [224]. However, spermatogonia can repair both active and inactive genes on both strands, in contrast to meiotic cells where transcribed genes are repaired. TC-NER is lower in round spermatids than in spermatogonia. These differences may reflect the greater mitotic activity of spermatogonia, and the fact that they are lying outside of the blood-testis barrier and so are potentially more exposed to genotoxins. Chromatin compaction during spermiogenesis may also adversely impact NER [227] as chromatin access is known to hinder NER. NER is therefore present at all stages of the male germ cells differentiation albeit with various efficiency.
OGG1Hs glycosylase is the only BER factor found in spermatozoa and recognizes 8-oxoG [219]. However, as none of the AP endonucleases exist in spermatozoa, AP sites will likely be processed during the first zygotic mitosis [219].
HR is known for its primary role during meiosis but is also important for DSBs and inter-strand lesions repair in spermatogonia. Outside of meiosis, HR acts in the S and G2 phases of the cell cycle, when sister chromatids are present. HR is obviously absent in haploid spermatids but resumes in the diploid zygote.
In male germ cells, some canonical components of NHEJ are missing resulting in an alternative NHEJ pathway [228]. For instance, Ku70Hs and 53BP1Hs are not expressed therefore preventing the DNA-PKcHs pathway. PARP-1Hs and XRCC1Hs are however expressed in elongating spermatids supporting alt-NHEJ pathway as described by Ahmed et al. [228]. PARP-1Hs may bind to DSB as a catalytic homodimer bringing together both DNA ends. Then, PARP-1Hs may recruit the XRCC1Hsc/DNA ligase IIIHs complex at the DSB site to complete the end-joining.

4.1.2. Oocyte

Since a limited number of oocytes can be retrieved, the DNA damage response in this gamete is studied to a lesser extent. DNA damages arise mainly during the active life of oocytes in prenatal stages, before prophase I arrest, and during recovery from pre-ovulatory meiosis [229].
From the primary oocyte stage to meiosis II and ovulation, oocytes can actively repair DNA [230]. During oogenesis, repair genes are overexpressed, and their mRNA are accumulated to be used in the zygote [231]. NER, BER, MMR, HR or NHEJ repair pathways have been described in oocytes from humans to mice [231,232,233,234]. Proper DNA repair machinery must operate as oocytes undergo chromatin remodelling during fertilisation and early steps of a zygote. Oocytes are diploid cells, that possess homologous chromosomes until meiosis II that arises after fertilisation. Thus, the oocyte is virtually never haploid. The oocyte becomes transiently haploid between meiosis II end and the fusion of parental genomes (syngamy). Thus, following meiosis II, only NHEJ should operate as shown in mice [235], although variations exist between species as rodent oocytes have better DNA repair abilities than primate ones [232]. The repair capacity of the oocyte and mRNA levels for repair proteins also decreases with female age [236]. After fertilisation, syngamy is a critical step since (1) Chromatids from each parent exist in different cellular compartments; (2) Chromatin undergoes major remodelling associated with demethylation; (3) Male derived chromatids are repaired by the oocyte DNA repair machinery [237].
As for spermatogenesis, ROS are produced during oogenesis. ROS production is necessary for folliculogenesis, oocyte maturation, ovulation, and luteal function [238]. Maturation and ovulation are linked to inflammatory processes, characterized by high ROS production [239]. Unlike sperm, oocytes are more resistant to ROS due to an environment rich in antioxidants in the follicular fluid [240]. However, ROS imbalance in the female tract decreases fertility and oocyte quality [241]. In addition to its impact on genetic integrity, oxidative stress is known to alter microtubule function and spindle morphology leading to aneuploidy [242].
In contrast to spermatozoa, OGG1Hs glycosylase is only weakly expressed in oocytes, as 8-oxoG damages are not present in spermatozoa after fertilization [243]. In many aspects, the early zygote may still be considered as an oocyte since only mRNAs for the oocyte are used up until the 4-cell stage [244]. This “zygotic oocyte” is in charge of repairing DNA damages in spermatozoa but may be overloaded if DNA damages occur in more than 8% of the nucleotide genome amount [245,246].
Both the number and the nature of DNA breaks are of concern for fertility. Whereas SSBs are easily repaired as in somatic cells, DSBs can overload the DNA repair capacity of the oocyte. After fertilisation, the persistence of DSBs in spermatozoa will delay DNA replication or can even be lethal for the zygote [247]. Finally, DSBs are mainly repaired by NHEJ during zygote [214], as for somatic cells.

5. Conclusions

Ploidy is a key element to consider in the regulation of DNA repair systems. In view of the aforementioned details about the two yeast models, it seems that species can however adapt their uses of DNA repair systems to overcome the weaknesses of either haploidy or diploidy. Thus, many redundancies exist within and between DNA repair systems that are linked together into a complex and integrated overarching regulatory mechanism that ensures the maintenance of genomic integrity regardless of the ploidy state. This may of course represent a functional bias for a single-cell organism as they are solely responsible to generate genetically intact progeny. In fact, major differences in repair efficiency between diploid and haploid cells are mainly observed among mammalian metazoan gametes. Hence, unlike oocytes, mature sperm clearly lacks proper DNA repair activity. However, the many replication/differentiation steps, change in DNA topology and highly condensed and specialized chromatin structure observed during spermatogenesis offers an appropriate context to generate adaptative variations.

Funding

This work was funded by The Natural Sciences and Engineering Research Council of Canada grant #06089.

Conflicts of Interest

Authors declare no conflict of interest.

References

  1. Gerstein, A.C.; Otto, S.P. Ploidy and the causes of genomic evolution. J. Hered. 2009, 100, 571–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Gresham, D.; Desai, M.M.; Tucker, C.M.; Jenq, H.T.; Pai, D.A.; Ward, A.; DeSevo, C.G.; Botstein, D.; Dunham, M.J. The repertoire and dynamics of evolutionary adaptations to controlled nutrient-limited environments in yeast. PLoS Genet. 2008, 4, e1000303. [Google Scholar] [CrossRef] [Green Version]
  3. Zeyl, C.; Vanderford, T.; Carter, M. An evolutionary advantage of haploidy in large yeast populations. Science 2003, 299, 555–558. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, H.; Zeidler, A.F.B.; Song, W.; Puccia, C.M.; Malc, E.; Greenwell, P.W.; Mieczkowski, P.A.; Petes, T.D.; Argueso, J.L. Gene copy-number variation in haploid and diploid strains of the yeast saccharomyces cerevisiae. Genetics 2013, 193, 785–801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Crow, J.F.; Kimura, M. Evolution in sexual and asexual populations. Am. Nat. 1965, 99, 439–450. [Google Scholar] [CrossRef]
  6. Kondrashov, A.S.; Crow, J.F. Haploidy or Diploidy: Which Is Better? Nature 1991, 351, 314–315. [Google Scholar] [CrossRef]
  7. Sharp, N.P.; Sandell, L.; James, C.G.; Otto, S.P. The Genome-Wide Rate and Spectrum of Spontaneous Mutations Differ between Haploid and Diploid Yeast. Proc. Natl. Acad. Sci. USA 2018, 115, E5046–E5055. [Google Scholar] [CrossRef] [Green Version]
  8. Lang, G.I.; Murray, A.W. Mutation rates across budding yeast chromosome VI are correlated with replication timing. Genome Biol. Evol. 2011, 3, 799–811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Orr, H.A.; Otto, S.P. Does Diploidy Increase the Rate of Adaptation? Genetics 1994, 136, 1475–1480. [Google Scholar] [CrossRef]
  10. Sia, E.A.; Butler, C.A.; Dominska, M.; Greenwell, P.; Fox, T.D.; Petes, T.D. Analysis of Microsatellite Mutations in the Mitochondrial DNA of Saccharomyces Cerevisiae. Proc. Natl. Acad. Sci. USA 2000, 97, 250–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Selmecki, A.M. Polyploidy Can Drive Rapid Adaptation in Yeast. Nature 2015, 519, 349–352. [Google Scholar] [CrossRef]
  12. Spivak, G. Nucleotide Excision Repair in Humans. DNA Repair 2015, 36, 13–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Hanawalt, P.C.; Spivak, G. Transcription-Coupled DNA Repair: Two Decades of Progress and Surprises. Nat. Rev. Mol. Cell Biol. 2008, 9, 958–970. [Google Scholar] [CrossRef] [PubMed]
  14. Shin, J.H.; Xu, L.; Wang, D. Mechanism of Transcription-Coupled DNA Modification Recognition. Cell Biosci. 2017, 7, 9. [Google Scholar] [CrossRef] [Green Version]
  15. Dawlaty, M.M.; Malureanu, L.; Jeganathan, K.B. Resolution of Sister Centromeres Requires RanBP2-Mediated SUMOylation of Topoisomerase IIalpha. Cell 2008, 133, 103–115. [Google Scholar] [CrossRef] [Green Version]
  16. Fousteri, M. Cockayne Syndrome A and B Proteins Differentially Regulaterecruitment of Chromatin Remodeling and Repair Factors to Stalled RNApolymerase II in vivo. Mol. Cell 2006, 23, 471–482. [Google Scholar] [CrossRef] [PubMed]
  17. Belmont, A.S. Mitotic Chromosome Structure and Condensation. Curr. Opin. Cell Biol. 2006, 18, 632–638. [Google Scholar] [CrossRef] [PubMed]
  18. Okuda, M.; Nakazawa, Y.; Guo, C.; Ogi, T.; Nishimura, Y. Common TFIIH Recruitment Mechanism in Global Genome and Transcription-Coupled Repair Subpathways. Nucleic Acids Res. 2017, 45, 13043–13055. [Google Scholar] [CrossRef]
  19. Reardon, J.T.; Sancar, A. Recognition and Repair of the Cyclobutane Thymine Dimer, a Major Cause of Skin Cancers, by the Human Excision Nuclease. Genes Dev. 2003, 17, 2539–2551. [Google Scholar] [CrossRef] [Green Version]
  20. Li, C.-L. Tripartite DNA Lesion Recognition and Verification by XPC, TFIIH, and XPA in Nucleotide Excision Repair. Mol. Cell 2015, 59, 1025–1034. [Google Scholar] [CrossRef] [Green Version]
  21. Cai, Y.; Patel, D.J.; Broyde, S.; Geacintov, N.E. Base sequence context effects on nucleotide excision repair. J. Nucleic Acids 2010, 2010, 174252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kropachev, K.; Kolbanovskii, M.; Cai, Y.; Rodriguez, F.; Kolbanovskii, A.; Liu, Y.; Zhang, L.; Amin, S.; Patel, D.; Broyde, S.; et al. The Sequence Dependence of Human Nucleotide Excision Repair Efficiencies of Benzo[a]Pyrene-Derived DNA Lesions: Insights into the Structural Factors That Favor Dual Incisions. J. Mol. Biol. 2009, 386, 1193–1203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Mocquet, V.; Kropachev, K.; Kolbanovskiy, M.; Kolbanovskiy, A.; Tapias, A.; Cai, Y.; Broyde, S.; Geacintov, N.E.; Egly, J.M. The Human DNA Repair Factor XPC-HR23B Distinguishes Stereoisomeric Benzo[a]Pyrenyl-DNA Lesions. EMBO J. 2007, 26, 2923–2932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Mu, H.; Geacintov, N.E.; Min, J.H.; Zhang, Y.; Broyde, S. Nucleotide Excision Repair Lesion-Recognition Protein Rad4 Captures a Pre-Flipped Partner Base in a Benzo[a] Pyrene-Derived DNA Lesion: How Structure Impacts the Binding Pathway. Chem. Res. Toxicol. 2017, 30, 1344–1354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Akita, M.; Tak, Y.S.; Shimura, T.; Matsumoto, S.; Okuda-Shimizu, Y.; Shimizu, Y.; Nishi, R.; Saitoh, H.; Iwai, S.; Mori, T.; et al. SUMOylation of Xeroderma Pigmentosum Group C Protein Regulates DNA Damage Recognition during Nucleotide Excision Repair. Sci. Rep. 2015, 5, 10984. [Google Scholar] [CrossRef] [Green Version]
  26. Cuijk, L.; Vermeulen, W.; Marteijn, J.A. Ubiquitin at work: The ubiquitousregulation of the damage recognition step of NER. Exp. Cell Res. 2014, 329, 101–109. [Google Scholar] [CrossRef]
  27. Sugasawa, K.; Akagi, J.; Nishi, R.; Iwai, S.; Hanaoka, F. Two-Step Recognition of DNA Damage for Mammalian Nucleotide Excision Repair: Directional Binding of the XPC Complex and DNA Strand Scanning. Mol. Cell 2009, 36, 642–653. [Google Scholar] [CrossRef]
  28. Singh, A. TFIIH Subunit Alterations Causing Xeroderma Pigmentosumand Trichothiodystrophy Specifically Disturb Several Steps Duringtranscription. Am. J. Hum. Genet. 2015, 96, 194–207. [Google Scholar] [CrossRef] [Green Version]
  29. Tsodikov, O.V.; Ivanov, D.; Orelli, B.; Staresincic, L.; Shoshani, I.; Oberman, R.; Scharer, O.D.; Wagner, G.; Ellenberger, T. Structural Basis for the Recruitment of ERCC1-XPF to Nucleotide Excision Repair Complexes by XPA. EMBO J. 2007, 26, 4768–4776. [Google Scholar] [CrossRef] [Green Version]
  30. Sijbers, A.M.; Laat, W.L.; Ariza, R.R.; Biggerstaff, M.; Wei, Y.F.; Moggs, J.G.; Carter, K.C.; Shell, B.K.; Evans, E.; Jong, M.C.; et al. Xeroderma Pigmentosum Group F Caused by a Defect in a Structure- Specific DNA Repair Endonuclease. Cell 1996, 86, 811–822. [Google Scholar] [CrossRef] [Green Version]
  31. Araujo, S.J.; Nigg, E.A.; Wood, R.D. Strong Functional Interactions of TFIIH with XPC and XPG in Human DNA Nucleotide Excision Repair, without a Preassembled Repairosome. Mol. Cell. Biol. 2001, 21, 2281–2291. [Google Scholar] [CrossRef] [Green Version]
  32. Ito, S.; Kuraoka, I.; Chymkowitch, P.; Compe, E.; Takedachi, A.; Ishigami, C.; Coin, F.; Egly, J.M.; Tanaka, K. XPG stabilizes TFIIH Allowing Transactivation of Nuclear Receptors: Implications for Cockayne Syndrome in XP-G/CS Patients. Mol. Cell 2007, 26, 231–243. [Google Scholar] [CrossRef] [PubMed]
  33. Fagbemi, A.F.; Orelli, B.; Schärer, O.D. Regulation of Endonuclease Activity in Human Nucleotide Excision Repair. DNA Repair 2011, 10, 722–729. [Google Scholar] [CrossRef] [Green Version]
  34. Li, S.; Smerdon, M.J. Rpb4 and Rpb9 Mediate Subpathways of Transcription-Coupled DNA Repair in Saccharomyces Cerevisiae. EMBO J. 2002, 21, 5921–5929. [Google Scholar] [CrossRef] [Green Version]
  35. Lee, S.-K.; Yu, S.-L.; Prakash, L.; Prakash, S. Yeast RAD26, a Homolog of the Human CSB Gene, Functions Independently of Nucleotide Excision Repair and Base Excision Repair in Promoting Transcription through Damaged Bases. Mol. Cell. Biol. 2002, 22, 4383–4389. [Google Scholar] [CrossRef] [Green Version]
  36. Yasuhira, S.; Yasui, A.; Morimyo, M. Transcription Dependence and the Roles of Two Excision Repair Pathways for UV Damage in Fission Yeast Schizosaccharomyces Pombe. J. Biol. Chem. 1999, 274, 26822–26827. [Google Scholar] [CrossRef] [Green Version]
  37. Guzder, S.N.; Sung, P.; Prakash, L.; Prakash, S. The DNA-Dependent ATPase Activity of Yeast Nucleotide Excision Repair Factor 4 and Its Role in DNA Damage Recognition. J. Biol. Chem. 1998, 273, 6292–6296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Yu, S.; Smirnova, J.B.; Friedberg, E.C.; Stillman, B.; Akiyama, M. ABF1-binding sites promote efficient global genome nucleotide excision repair. J. Biol. Chem. 2009, 284, 966–973. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. De Laat, W.L.; Jaspers, N.G.; Hoeijmakers, J.H. Molecular mechanism of nucleotide excision repair. Genes Dev. 1999, 13, 768–785. [Google Scholar] [CrossRef] [Green Version]
  40. Sugasawa, K.; Okamoto, T.; Shimizu, Y.; Masutani, C.; Iwai, S.; Hanaoka, F. A multistep damage recognition mechanism for global genomic nucleotide excision repair. Genes Dev. 2001, 15, 507–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Fukumoto, Y.; Hiyama, H.; Yokoi, M.; Nakaseko, Y.; Yanagida, M.; Hanaoka, F. Two Budding Yeast RAD4 Homologs in Fission Yeast Play Different Roles in the Repair of UV-Induced DNA Damage. DNA Repair 2002, 1, 833–845. [Google Scholar] [CrossRef]
  42. Marti, T.M.; Kunz, C.; Fleck, O. Repair of Damaged and Mismatched DNA by the XPC Homologues Rhp41 and Rhp42 of Fission Yeast. Genetics 2003, 164, 457–467. [Google Scholar] [CrossRef] [PubMed]
  43. Verhage, R.; Zeeman, A.M.; Groot, N.D.; Gleig, F.; Bang, D.D. The RAD7 and RAD16 Genes, Which Are Essential for Pyrimidine Dimer Removal from the Silent Mating Type Loci, Are Also Required for Repair of the Nontranscribed Strand of an Active Gene in Saccharomyces Cerevisiae. Mol. Cell. Biol. 1994, 14, 6135–6142. [Google Scholar] [PubMed]
  44. Bardwell, A.J.; Bardwell, L.; Tomkinson, A.E.; Friedberg, E.C. Specific Cleavage of Model Recombination and Repair Intermediates by the Yeast Rad1-Rad10 DNA Endonuclease. Science 1994, 265, 2082–2085. [Google Scholar] [CrossRef]
  45. Carr, A.M.; Schmidt, H.; Kirchoff, S.; Muriel, W.J.; Sheldrick, K.S.; Griffiths, D.J.; Basmacioglu, C.N.; Subramani, S.; Clegg, M.; Nasim, A.; et al. The Rad16 Gene of Schizosaccharomyces Pombe: A Homolog of the RAD1 Gene of Saccharomyces CereÍisiae. Mol. Cell. Biol. 1994, 14, 2029–2040. [Google Scholar]
  46. McCready, S.J.; Burkhill, H.; Evans, S. The Saccharomyces CereÍisiae RAD2 Gene Complements a Schizosaccharomyces Pombe Repair Mutation. Curr. Genet. 1989, 15, 27–30. [Google Scholar] [CrossRef] [PubMed]
  47. Beard, W.A.; Horton, J.K.; Prasad, R.; Wilson, S.H. Eukaryotic Base Excision Repair: New Approaches Shine Light on Mechanism. Annu. Rev. Biochem. 2019, 88, 137–162. [Google Scholar] [CrossRef]
  48. Chastain, P.D.; Nakamura, J.; Rao, S.; Chu, H.; Ibrahim, J.G. Abasic Sites Preferentially Form at Regions Undergoing DNA Replication. FASEB J. 2010, 24, 3674–3680. [Google Scholar]
  49. Nakamura, J.; Swenberg, J.A. Endogenous Apurinic/Apyrimidinic Sites in Genomic DNA of Mammalian Tissues. Cancer Res. 1999, 59, 2522–2526. [Google Scholar]
  50. Hegde, M.L.; Hazra, T.K.; Mitra, S. Functions of Disordered Regions in Mammalian Early Base Excision Repair Proteins. Cell. Mol. Life Sci. 2010, 67, 3573–3587. [Google Scholar] [CrossRef] [Green Version]
  51. Svilar, D.; Goellner, E.M.; Almeida, K.H.; Sobol, R.W. Base Excision Repair and Lesion-Dependent Subpathways for Repair of Oxidative DNA Damage. Antioxid. Redox Signal. 2011, 14, 2491–2507. [Google Scholar] [CrossRef] [Green Version]
  52. Jacobs, A.L.; Schar, P. DNA glycosylases: In DNA repair and beyond. Chromosoma 2012, 121, 1–20. [Google Scholar] [CrossRef] [Green Version]
  53. Huffman, J.L.; Sundheim, O.; Tainer, J.A. DNA Base Damage Recognition and Removal: New Twists and Grooves. Mutat. Res. 2005, 577, 55–76. [Google Scholar] [CrossRef] [PubMed]
  54. Sassa, A.; Beard, W.A.; Prasad, R.; Wilson, S.H. DNA sequence context effects on the glycosylase activity of human 8-oxoguanine DNA glycosylase. J. Biol. Chem. 2012, 287, 36702–36710. [Google Scholar] [CrossRef] [Green Version]
  55. Wong, I.; Lundquist, A.J.; Bernards, A.S.; Mosbaugh, D.W. Presteady-state analysis of a single catalytic turnover by Escherichia coli uracil-DNA glycosylase reveals a pinch-pull-push mechanism. J. Biol. Chem. 2002, 277, 19424–19432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Robertson, A.B.; Klungland, A.; Rognes, T.; Leiros, I. DNA Repair in Mammalian Cells: Base Excision Repair: The Long and Short of It. Cell. Mol. Life Sci. 2009, 66, 981–993. [Google Scholar] [CrossRef] [PubMed]
  57. Beard, W.A.; Wilson, S.H. Structure and mechanism of DNA polymerase b. Chem. Rev. 2006, 106, 361–382. [Google Scholar] [CrossRef] [PubMed]
  58. Prasad, R.; Batra, V.K.; Yang, X.-P.; Krahn, J.M.; Pedersen, L.C. Structural Insight into the DNA Polymerase Βdeoxyribose Phosphate Lyase Mechanism. DNA Repair 2005, 4, 1347–1357. [Google Scholar] [CrossRef]
  59. Fortini, P.; Parlanti, E.; Sidorkina, O.M.; Laval, J.; Dogliotti, E. The Type of DNA Glycosylase Determines the Base Excision Repair Pathway in Mammalian Cells. J. Biol. Chem. 1999, 274, 15230–15236. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, Y.; Beard, W.A.; Shock, D.D.; Prasad, R.; Hou, E.W.; Wilson, S.H. DNA Polymerase β and Flap Endonuclease 1 Enzymatic Specificities Sustain DNA Synthesis for Long Patch Base Excision Repair. J. Biol. Chem. 2005, 280, 3665–3674. [Google Scholar] [CrossRef] [Green Version]
  61. Elder, R.T.; Zhu, X.; Priet, S. A Fission Yeast Homologue of the Human Uracil-DNA-Glycosylase and Their Roles in Causing DNA Damage after Overexpression. Biochem. Biophys. Res. Commun. 2003, 306, 693–700. [Google Scholar] [CrossRef]
  62. Hardeland, U.; Bentele, M.; Jiricny, J.; Schär, P. The Versatile Thymine DNA-glycosylase: A Comparative Characterization of the Human, Drosophila and Fission Yeast Orthologs. Nucleic Acids Res. 2003, 31, 2261–2271. [Google Scholar] [CrossRef]
  63. Dong, L.; Mi, R.; Glass, R.A.; Barry, J.N.; Cao, W. Repair of Deaminated Base Damage by Schizosaccharomyces Pombe Thymine DNA Glycosylase. DNA Repair 2008, 7, 1962–1972. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, J.; Derfler, B.; Samson, L. Saccharomyces Cerevisiae 3-Methyladenine DNA Glycosylase Has Homology to the AlkA Glycosylase of E. Coli and Is Induced in Response to DNA Alkylation Damage. EMBO J. 1990, 9, 4569–4575. [Google Scholar] [CrossRef]
  65. Berdal, K.G.; Johansen, R.F.; Seeberg, E. Release of Normal Bases from Intact DNA by a Native DNA Repair Enzyme. EMBO J. 1998, 17, 363–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kanamitsu, K.; Tanihigashi, H.; Tanita, Y.; Inatani, S.; Ikeda, S. Involvement of 3-Methyladenine DNA Glycosylases Mag1p and Mag2p in Base Excision Repair of Methyl Methanesulfonate-Damaged DNA in the Fission Yeast Schizosaccharomyces Pombe. Genes Genet. Syst. 2007, 82, 489–494. [Google Scholar] [CrossRef] [Green Version]
  67. Memisoglu, A.; Samson, L. Contribution of Base Excision Repair, Nucleotide Excision Repair, and DNA Recombination to Alkylation Resistance of the Fission Yeast Schizosaccharomyces Pombe. J. Bacteriol. 2000, 182, 2104–2112. [Google Scholar] [CrossRef] [Green Version]
  68. Alseth, I.; Osman, F.; Korvald, H. Biochemical Characterization and DNA Repair Pathway Interactions of Mag1-Mediated Base Excision Repair in Schizosaccharomyces Pombe. Nucleic Acids Res. 2005, 33, 1123–1131. [Google Scholar] [CrossRef] [Green Version]
  69. Alseth, I.; Korvald, H.; Osman, F.; Seeberg, E.; Bjørås, M. A General Role of the DNA Glycosylase Nth1 in the Abasic Sites Cleavage Step of Base Excision Repair in Schizosaccharomyces Pombe. Nucleic Acids Res. 2004, 32, 5119–5125. [Google Scholar] [CrossRef] [Green Version]
  70. Ribar, B.; Izumi, T.; Mitra, S. The Major Role of Human AP-Endonuclease Homolog Apn2 in Repair of Abasic Sites in Schizosaccharomyces Pombe. Nucleic Acids Res. 2004, 32, 115–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Sugimoto, T.; Igawa, E.; Tanihigashi, H.; Matsubara, M.; Ide, H.; Ikeda, S. Roles of Base Excision Repair Enzymes Nth1p and Apn2p from Schizosaccharomyces Pombe in Processing Alkylation and Oxidative DNA Damage. DNA Repair 2005, 4, 1270–1280. [Google Scholar] [CrossRef] [PubMed]
  72. Alseth, I.; Eide, L.; Pirovano, M.; Rognes, T.; Seeberg, E. The Saccharomyces Cerevisiae Homologues of Endonuclease III from Escherichia Coli, Ntg1 and Ntg2, Are Both Required for Efficient Repair of Spontaneous and Induced Oxidative DNA Damage in Yeast. Mol. Cell. Biol. 1999, 19, 3779–3787. [Google Scholar] [CrossRef] [Green Version]
  73. Senturker, S.; Kemp, P.A.V.D.; You, H.J.; Doetsch, P.W.; Dizdaroglu, M. Substrate Specificities of the Ntg1 and Ntg2 Proteins of Saccharomyces Cerevisiae for Oxidized DNA Bases Are Not Identical. Nucleic Acids Res. 1998, 26, 5270–5276. [Google Scholar] [CrossRef] [Green Version]
  74. Girard, P.M.; Guibourt, N.; Boiteux, S. The Ogg1 Protein of Saccharomyces Cerevisiae: A 7,8-Dihydro-8-Oxoguanine DNA Glycosylase/AP Lyase Whose Lysine 241 Is a Critical Residue for Catalytic Activity. Nucleic Acids Res. 1997, 25, 3204–3211. [Google Scholar] [CrossRef] [Green Version]
  75. Fleck, O. DNA repair pathway. In The Molecular Biology of Schizosaccharomyces Pombe; Egel, R., Ed.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 101–115. [Google Scholar]
  76. Karahalil, B.; Roldán-Arjona, T.; Dizdaroglu, M. Substrate Specificity of Schizosaccharomyces Pombe Nth Protein for Products of Oxidative DNA Damage. Biochemistry 1998, 37, 590–595. [Google Scholar] [CrossRef]
  77. Lu, A.-L.; Fawcett, W.P. Characterization of the Recombinant MutY Homolog, an Adenine DNA Glycosylase, from Yeast Schizosaccharomyces Pombe. J. Biol. Chem. 1998, 273, 25098–25105. [Google Scholar] [CrossRef] [Green Version]
  78. Doi, T.; Yonekura, S.-I.; Tano, K.; Yasuhira, S.; Yonei, S.; Zhang, Q.-M. The Shizosaccharomyces Pombe Homolog (SpMYH) of the Escherichia Coli MutY Is Required for Removal of Guanine from 8-Oxoguanine/Guanine Mispairs to Prevent G:C to C:G Transversions. J. Radiat. Res. 2005, 46, 205–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Chang, D.-Y.; Lu, A.-L. Interaction of Checkpoint Proteins Hus1/Rad1/Rad9 with DNA Base Excision Repair Enzyme MutY Homolog in Fission Yeast, Schizosaccharomyces Pombe. J. Biol. Chem. 2005, 280, 408–417. [Google Scholar] [CrossRef] [Green Version]
  80. Jansson, K.; Warringer, J.; Farewell, A. The Tumor Suppressor Homolog in Fission Yeast, Myh1+, Displays a Strong Interaction with the Checkpoint Gene Rad1+. Mutat. Res. 2008, 644, 48–55. [Google Scholar] [CrossRef]
  81. Popoff, S.C.; Spira, A.I.; Johnson, A.W.; Demple, B. Yeast structural gene (APN1) for the major apurinic endonuclease: Homology to Escherichia coli endonuclease IV. Proc. Natl. Acad. Sci. USA 1990, 87, 4193–4197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Acevedo-Torres, K.; Fonseca-Williams, S.; Ayala-Torres, S.; Torres-Ramos, C.A. Requirement of the Saccharomyces Cerevisiae APN1 Gene for the Repair of Mitochondrial DNA Alkylation Damage. Environ. Mol. Mutagen. 2009, 50, 317–327. [Google Scholar] [CrossRef] [Green Version]
  83. Tanihigashi, H.; Yamada, A.; Igawa, E.; Ikeda, S. The Role of Schizosaccharomyces Pombe DNA Repair Enzymes Apn1p and Uve1p in the Base Excision Repair of Apurinic/Apyrimidinic Sites. Biochem. Biophys. Res. Commun. 2006, 347, 889–894. [Google Scholar] [CrossRef] [PubMed]
  84. Alleva, J.L.; Zuo, S.; Hurwitz, J.; Doetsch, P.W. In Vitro Reconstitution of the Schizosaccharomyces Pombe Alternative Excision Repair Pathway. Biochemistry 2000, 39, 2659–2666. [Google Scholar] [CrossRef] [PubMed]
  85. Yasuhira, S.; Yasui, A. Alternative Excision Repair Pathway of UV-Damaged DNA in Schizosaccharomyces Pombe Operates Both in Nucleus and in Mitochondria. J. Biol. Chem. 2000, 275, 11824–11828. [Google Scholar] [CrossRef] [Green Version]
  86. Kunkel, T.A.; Erie, D.A. Eukaryotic Mismatch Repair in Relation to DNA Replication. Annu. Rev. Genet. 2015, 49, 291–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lamers, M.H.; Perrakis, A.; Enzlin, J.H.; Winterwerp, H.H.K.; Wind, N.D. The Crystal Structure of DNA Mismatch Repair Protein MutS Binding to a G:T Mismatch. Nature 2000, 407, 711–717. [Google Scholar] [CrossRef]
  88. Obmolova, G.; Ban, C.; Hsieh, P.; Yang, W. Crystal Structures of Mismatch Repair Protein MutS and Its Complex with a Substrate DNA. Nature 2000, 407, 703–710. [Google Scholar] [CrossRef] [PubMed]
  89. Iyer, R.R.; Pluciennik, A.; Burdett, V.; Modrich, P.L. DNA mismatch repair: Functions and mechanisms. Chem. Rev. 2006, 106, 302–323. [Google Scholar] [CrossRef]
  90. Cerretelli, G.; Ager, A.; Arends, M.J.; Frayling, I.M. Molecular Pathology of Lynch Syndrome. J. Pathol. 2020, 250, 518–531. [Google Scholar] [CrossRef] [Green Version]
  91. Surtees, J.A.; Argueso, J.L.; Alani, E. Mismatch Repair Proteins: Key Regulators of Genetic Recombination. CGR 2004, 107, 146–159. [Google Scholar] [CrossRef]
  92. Kawasoe, Y.; Tsurimoto, T.; Nakagawa, T.; Masukata, H.; Takahashi, T.S. MutSalpha Maintains the Mismatch Repair Capability by Inhibiting PCNA Unloading. eLife 2016, 5, e15155. [Google Scholar] [CrossRef]
  93. Erdeniz, N.; Nguyen, M.; Deschenes, S.M.; Liskay, R.M. Mutations Affecting a Putative MutLa Endonuclease Motif Impact Multiple Mismatch Repair Functions. DNA Repair 2007, 6, 1463–1470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhang, Y.; Yuan, F.; Presnell, S.R.; Tian, K.; Gao, Y.; Tomkinson, A.E.; Gu, L.; Li, G.M. Reconstitution of 5’-Directed Human Mismatch Repair in a Purified System. Cell 2005, 122, 693–705. [Google Scholar] [CrossRef] [Green Version]
  95. Kadyrov, F.A.; Genschel, J.; Fang, Y.; Penland, E.; Edelmann, W.; Modrich, P. A possible mechanism for exonuclease 1-independent eukaryotic mismatch repair. Proc. Natl. Acad. Sci. USA 2009, 106, 8495–8500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Harrington, J.M.; Kolodner, R.D. Saccharomyces Cerevisiae Msh2-Msh3 Acts in Repair of Base-Base Mispairs. Mol. Cell. Biol. 2007, 27, 6546–6554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Detloff, P.; Sieber, J.; Petes, T.D. Repair of Specific Base Pair Mismatches Formed during Meiotic Recombination in the Yeast Saccharomyces Cerevisiae. Mol. Cell. Biol. 1991, 11, 737–745. [Google Scholar]
  98. Strand, M.; Earley, M.C.; Crouse, G.F.; Petes, T.D. Mutations in the MSH3 Gene Preferentially Lead to Deletions within Tracts of Simple Repetitive DNA in Saccharomyces Cerevisiae. Proc. Natl. Acad. Sci. USA 1995, 92, 10418–10421. [Google Scholar] [CrossRef] [Green Version]
  99. Harfe, B.D.; Minesinger, B.K.; Jinks-Robertson, S. Discrete in vivo roles for the MutL homologs Mlh2p and Mlh3p in the removal of frameshift intermediates in budding yeast. Curr. Biol. 2002, 10, 145–148. [Google Scholar] [CrossRef] [Green Version]
  100. Tornier, C.; Bessone, S.; Varlet, I.; Rudolph, C.; Darmon, M.; Fleck, O. Requirement for Msh6, but Not for Swi4 (Msh3), in Msh2-Dependent Repair of Base-Base Mismatches and Mononucleotide Loops in Schizosaccharomyces Pombe. Genetics 2001, 158, 65–75. [Google Scholar] [CrossRef] [PubMed]
  101. Saparbaev, M.; Prakash, L.; Prakash, S. Requirement of Mismatch Repair Genes MSH2 and MSH3 in the RAD1-RAD10 Pathway of Mitotic Recombination in Saccharomyces Cerevisiae. Genetics 1996, 142, 727–736. [Google Scholar] [CrossRef]
  102. Rudolph, C.; Kunz, C.; Parisi, S.; Lehmann, E.; Hartsuiker, E.; Fartmann, B.; Kramer, W.; Kohli, J.; Fleck, O. The Msh2 Gene OfSchizosaccharomyces Pombe Is Involved in Mismatch Repair, Mating-Type Switching, and Meiotic Chromosome Organization. Mol. Cell. Biol. 1999, 19, 241–250. [Google Scholar] [CrossRef] [Green Version]
  103. Ross-MacDonald, P.; Roeder, G.S. Mutation of a Meiosis-Specific MutS Homolog Decreases Crossing over but Not Mismatch Correction. Cell 1994, 79, 1069–1080. [Google Scholar] [CrossRef]
  104. Hollingsworth, N.M.; Ponte, L.; Halsey, C. MSH5, a Novel MutS Homolog, Facilitates Meiotic Reciprocal Recombination between Homologs in Saccharomyces Cerevisiae but Not Mismatch Repair. Genes Dev. 1995, 9, 1728–1739. [Google Scholar] [CrossRef] [Green Version]
  105. Zakharyevich, K.; Tang, S.; Ma, Y.; Hunter, N. Delineation of Joint Molecule Resolution Pathways in Meiosis Identifies a Crossover-Specific Resolvase. Cell 2012, 149, 334–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wang, T.F.; Kleckner, N.; Hunter, N. Functional Specificity of MutL Homologs in Yeast: Evidence for Three Mlh1- Based Heterocomplexes with Distinct Roles during Meiosis in Recombination and Mismatch Correction. Proc. Natl. Acad. Sci. USA 1999, 96, 13914–13919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Tran, P.T.; Erdeniz, N.; Symington, L.S.; Liskay, R.M. EXO1—A Multi-Tasking Eukaryotic Nuclease. DNA Repair 2004, 3, 1549–1559. [Google Scholar] [CrossRef] [PubMed]
  108. Johnson, R.E.; Kovvali, G.K.; Prakash, L.; Prakash, S. Requirement of the Yeast RTH1 59 to 39 Exonuclease for the Stability of Simple Repetitive DNA. Science 1995, 269, 238–240. [Google Scholar] [CrossRef]
  109. Mansour, A.A.; Tornier, C.; Lehmann, E.; Darmon, M.; Fleck, O. Control of GT Repeat Stability in Schizosaccharomyces Pombe by Mismatch Repair Factors. Genetics 2001, 158, 77–85. [Google Scholar] [CrossRef]
  110. Schär, P.; Kohli, J. Marker Effects of G to C Transversions on Intragenic Recombination and Mismatch Repair in Schizosaccharomyces Pombe. Genetics 1993, 133, 825–835. [Google Scholar] [CrossRef]
  111. Fleck, O.; Lehmann, E.; Schär, P.; Kohli, J. Involvement of Nucleotide-Excision Repair in Msh2 Pms1-Independent Mismatch Repair. Nat. Genet. 1999, 21, 314–317. [Google Scholar] [CrossRef]
  112. Hohl, M.; Christensen, O.; Kunz, C.; Naegeli, H.; Fleck, O. Binding and Repair of Mismatched DNA Mediated by Rhp14, the Fission Yeast Homologue of Human XPA. J. Biol. Chem. 2001, 276, 30766–30772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Kunz, C.; Fleck, O. Role of the DNA Repair Nucleases Rad13, Rad2 and Uve1 of Schizosaccharomyces Pombe in Mismatch Correction11Edited by S. Reed. J. Mol. Biol. 2001, 313, 241–253. [Google Scholar] [CrossRef]
  114. Coïc, E.; Gluck, L.; Fabre, F. Evidence for Short-Patch Mismatch Repair in Saccharomyces Cerevisiae. EMBO J. 2000, 19, 3408–3417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Harfe, B.D.; Jinks-Robertson, S. Dna Mismatch Repair and Genetic Instability. Annu. Rev. Genet. 2000, 34, 359–399. [Google Scholar] [CrossRef] [PubMed]
  116. Carpenter, A.T. Mismatch Repair, Gene Conversion, and Crossing-over in Two Recombination-Defective Mutants of Drosophila Melanogaster. Proc. Natl. Acad. Sci. USA 1982, 79, 5961–5965. [Google Scholar] [CrossRef] [Green Version]
  117. Bhui-Kaur, A.; Goodman, M.F.; Tower, J. DNA Mismatch Repair Catalyzed by Extracts of Mitotic, Postmitotic, and Senescent Drosophila Tissues and Involvement of Mei-9 Gene Function for Full Activity. Mol. Cell. Biol. 1998, 18, 1436–1443. [Google Scholar] [CrossRef] [Green Version]
  118. Petit, C.; Sancar, A. Nucleotide Excision Repair: From E. Coli to Man. Biochimie 1999, 81, 15–25. [Google Scholar] [CrossRef]
  119. Vilenchik, M.M.; Knudson, A.G. Endogenous DNA Double-Strand Breaks: Production, Fidelity of Repair, and Induction of Cancer. Proc. Natl. Acad. Sci. USA 2003, 100, 12871–12876. [Google Scholar] [CrossRef] [Green Version]
  120. Cannan, W.J.; Pederson, D.S. Mechanisms and Consequences of Double-Strand DNA Break Formation in Chromatin. J. Cell. Physiol. 2016, 231, 3–14. [Google Scholar] [CrossRef] [Green Version]
  121. Her, J.; Bunting, S.F. How Cells Ensure Correct Repair of DNA Double-Strand Breaks. J. Biol. Chem. 2018, 293, 10502–10511. [Google Scholar] [CrossRef] [Green Version]
  122. Karanam, K.; Kafri, R.; Loewer, A.; Lahav, G. Quantitative Live Cell Imaging Reveals a Gradual Shift between DNA Repair Mechanisms and a Maximal Use of HR in Mid S Phase. Mol. Cell. 2012, 47, 320–329. [Google Scholar] [CrossRef] [Green Version]
  123. Mao, Z.; Bozzella, M.; Seluanov, A.; Gorbunova, V. Comparison of Nonhomologous End Joining and Homologous Recombination in Human Cells. DNA Repair 2008, 7, 1765–1771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Pierce, A.J.; Hu, P.; Han, M.; Ellis, N.; Jasin, M. Protein Modulates Homologous Repair of Double-Strand Breaks in Mammalian Cells. Genes Dev. 2001, 15, 3237–3242. [Google Scholar] [CrossRef] [Green Version]
  125. Ahrabi, S.; Sarkar, S.; Pfister, S.X.; Pirovano, G.; Higgins, G.S.; Porter, A.C.; Humphrey, T.C. A Role for Human Homologous Recombination Factors in Suppressing Microhomology-Mediated End Joining. Nucleic Acids Res. 2016, 44, 5743–5757. [Google Scholar] [CrossRef] [Green Version]
  126. Bétermier, M.; Bertrand, P.; Lopez, B.S. Is Non-Homologous End-Joining Really an Inherently Error-Prone Process? PLoS Genet. 2014, 10, 1004086. [Google Scholar] [CrossRef] [Green Version]
  127. Rass, E.; Grabarz, A.; Plo, I.; Gautier, J.; Bertrand, P.; Lopez, B.S. Role of Mre11 in Chromosomal Nonhomologous End Joining in Mammalian Cells. Nat. Struct. Mol. Biol. 2009, 16, 819–824. [Google Scholar] [CrossRef]
  128. So, A.; Le Guen, T.; Lopez, B.S.; Guirouilh-Barbat, J. Genomic rearrangements induced by unscheduled DNA double strand breaks in somatic mammalian cells. FEBS J. 2017, 284, 2324–2344. [Google Scholar] [CrossRef] [Green Version]
  129. Escribano-Diaz, C. A Cell Cycle-Dependent Regulatory Circuit Composed of 53BP1 RIF1 and BRCA1 CtIP Controls DNA Repair Pathway Choice. Mol. Cell 2013, 49, 872–883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Yun, M.H.; Hiom, K. CtIP-BRCA1 Modulates the Choice of DNA Double-Strand-Break Repair Pathway throughout the Cell Cycle. Nature 2009, 459, 460–463. [Google Scholar] [CrossRef] [PubMed]
  131. Johnson, R.D.; Jasin, M. Sister Chromatid Gene Conversion Is a Prominent Double-Strand Break Repair Pathway in Mammalian Cells. EMBO J. 2000, 19, 3398–3407. [Google Scholar] [CrossRef]
  132. Howard, S.M.; Yanez, D.A.; Stark, J.M. DNA Damage Response Factors from Diverse Pathways, Including DNA Crosslink Repair, Mediate Alternative End Joining. PLoS Genet. 2015, 11, 1004943. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Bhargava, R.; Onyango, D.O.; Stark, J.M. Regulation of Single Strand Annealing and Its Role in Genome Maintenance. Trends Genet. 2016, 32, 566–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Krejci, L.; Altmannova, V.; Spirek, M.; Zhao, X. Homologous Recombination and Its Regulation. Nucleic Acids Res. 2012, 40, 5795–5818. [Google Scholar] [CrossRef]
  135. Tkach, J.M.; Yimit, A.; Lee, A.Y.; Riffle, M.; Costanzo, M.; Jaschob, D.; Hendry, J.A.; Ou, J.; Moffat, J.; Boone, C.; et al. Dissecting DNA Damage Response Pathways by Analysing Protein Localization and Abundance Changes during DNA Replication Stress. Nat. Cell Biol. 2012, 14, 966–976. [Google Scholar] [CrossRef] [Green Version]
  136. Aylon, Y.; Kupiec, M. DSB Repair: The Yeast Paradigm. DNA Repair 2004, 3, 797–815. [Google Scholar] [CrossRef] [PubMed]
  137. Wu, D.; Topper, L.M.; Wilson, T.E. Recruitment and dissociation of nonhomologous end joining proteins at a DNA double-strand break in Saccharomyces cerevisiae. Genetics 2008, 178, 1237–1249. [Google Scholar] [CrossRef] [Green Version]
  138. Lobachev, K.; Vitriol, E.; Stemple, J.; Resnick, M.A.; Bloom, K. Chromosome Fragmentation after Induction of a Double-Strand Break Is an Active Process Prevented by the RMX Repair Complex. Curr. Biol. 2004, 14, 2107–2112. [Google Scholar] [CrossRef] [Green Version]
  139. 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]
  140. Clerici, M.; Mantiero, D.; Guerini, I.; Lucchini, G.; Longhese, M.P. The Yku70-Yku80 Complex Contributes to Regulate Double-Strand Break Processing and Checkpoint Activation during the Cell Cycle. EMBO Rep. 2008, 9, 810–818. [Google Scholar] [CrossRef] [Green Version]
  141. Zhang, Y.; Shim, E.Y.; Davis, M.; Lee, S.E. Regulation of Repair Choice: Cdk1 Suppresses Recruitment of End Joining Factors at DNA Breaks. DNA Repair 2009, 8, 1235–1241. [Google Scholar] [CrossRef] [Green Version]
  142. Manolis, K.G.; Nimmo, E.R.; Hartsuiker, E. Novel Functional Requirements for Non-Homologous DNA Endjoining in Schizosaccharomyces Pombe. EMBO J. 2001, 20, 210–221. [Google Scholar] [CrossRef] [Green Version]
  143. Muris, D.F.; Vreeken, K.; Schmidt, H. Homologous Recombination in the Fission Yeast Schizosaccharomyces Pombe: Different Requirements for the Rhp51+, Rhp54+ and Rad22+ Genes. Curr. Genet. 1997, 31, 248–254. [Google Scholar] [CrossRef] [PubMed]
  144. Prudden, J.; Evans, J.S.; Hussey, S.P. Pathway Utilization in Response to a Site-Specific DNA Double-Strand Break in Fission Yeast. EMBO J. 2003, 22, 1419–1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Marchetti, M.A.; Kumar, S.; Hartsuiker, E. 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]
  146. Chahwan, C.; Nakamura, T.M.; Sivakumar, S.; Russell, P.; Rhind, N. The Fission Yeast Rad32(Mre11)–Rad50–Nbs1 Complex Is Required for the S-Phase DNA Damage Checkpoint. Mol. Cell Biol. 2003, 23, 6564–6573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Thacker, J. The RAD51 Gene Family, Genetic Instability and Cancer. Cancer Lett. 2005, 219, 125–135. [Google Scholar] [CrossRef]
  148. Schmidt, H.; Kapitza-Fecke, P.; Stephen, E.R.; Gutz, H. Some of the Swi Genes of Schizosaccharomyces Pombe Also Have a Function in the Repair of Radiation Damage. Curr. Genet. 1989, 16, 89–94. [Google Scholar] [CrossRef]
  149. Suto, K.; Nagata, A.; Murakami, H.; Okayama, H. A double-strand break repair component is essential for S phase completion in fission yeast cell cycling. Mol. Biol. Cell 1999, 10, 3331–3343. [Google Scholar] [CrossRef] [Green Version]
  150. Van den Bosch, M.; Vreeken, K.; Zonneveld, J.B.; Brandsma, J.A.; Lombaerts, M.; Murray, J.M.; Lohman, P.H.; Pastink, A. Characterization of RAD52 Homologs in the Fission Yeast Schizosaccharomyces Pombe. Mutat. Res. 2001, 461, 311–323. [Google Scholar] [CrossRef]
  151. Nimonkar, A.V.; Genschel, J.; Kinoshita, E.; Polaczek, P.; Campbell, J.L.; Wyman, C.; Modrich, P.; Kowalczykowski, S.C. BLM-DNA2-RPA-MRN and EXO1-BLM-RPA-MRN Constitute Two DNA End Resection Machineries for Human DNA Break Repair. Genes Dev. 2011, 25, 350–362. [Google Scholar] [CrossRef] [Green Version]
  152. Cox, M.M. Motoring along with the Bacterial RecA Protein. Nat. Rev. Mol. Cell. Biol. 2007, 8, 127–138. [Google Scholar] [CrossRef]
  153. San Filippo, J.; Sung, P.; Klein, H. Mechanism of Eukaryotic Homologous Recombination. Annu. Rev. Biochem. 2008, 77, 229–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Klein, H.L.; Symington, L.S. Sgs1—The Maestro of Recombination. Cell 2012, 149, 257–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Mimitou, E.P.; Symington, L.S. Sae2, Exo1 and Sgs1 Collaborate in DNA Double-Strand Break Processing. Nature 2008, 455, 770–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Tomita, K.; Matsuura, A.; Caspari, T.; Carr, A.M.; Akamatsu, Y.; Iwasaki, H.; Mizuno, K.; Ohta, K.; Uritani, M.; Ushimaru, T.; et al. Competition between the Rad50 Complex and the Ku Heterodimer Reveals a Role for Exo1 in Processing Double-Strand Breaks but Not Telomeres. Mol. Cell. Biol. 2003, 23, 5186–5197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Ferretti, L.P.; Lafranchi, L.; Sartori, A.A. Controlling DNA-End Resection: A New Task for CDKs. Front. Genet. 2013, 4, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Sugiyama, T.; Zaitseva, E.M.; Kowalczykowski, S.C. A Single-Stranded DNA-Binding Protein Is Needed for Efficient Presynaptic Complex Formation by the Saccharomyces Cerevisiae Rad51 Protein. J. Biol. Chem. 1997, 272, 7940–7945. [Google Scholar] [CrossRef] [Green Version]
  159. Parker, A.E.; Clyne, R.K.; Carr, A.M.; Kelly, T.J. The Schizosaccharomyces Pombe Rad11+ Gene Encodes the Large Subunit of Replication Protein A. Mol. Cell. Biol. 1997, 17, 2381–2390. [Google Scholar] [CrossRef] [Green Version]
  160. Krejci, L.; Song, B.; Bussen, W.; Rothstein, R.; Mortensen, U.H.; Sung, P. Interaction with Rad51 Is Indispensable for Recombination Mediator Function of Rad52. J. Biol. Chem. 2002, 277, 40132–40141. [Google Scholar] [CrossRef] [Green Version]
  161. Hays, S.L.; Firmenich, A.A.; Berg, P. Complex formation in yeast double-strand break repair: Participation of Rad51, Rad52, Rad55, and Rad57 proteins. Proc. Natl. Acad. Sci. USA 1995, 92, 6925–6929. [Google Scholar] [CrossRef] [Green Version]
  162. Muris, D.F.; Vreeken, K.; Carr, A.M. Isolation of the Schizosaccharomyces Pombe RAD54 Homologue, Rhp54+, a Gene Involved in the Repair of Radiation Damage and Replication Fidelity. J. Cell Sci. 1996, 109, 73–81. [Google Scholar] [CrossRef] [PubMed]
  163. Sauvageau, S.; Stasiak, A.Z.; Banville, I. Fission Yeast Rad51 and Dmc1, Two Efficient DNA Recombinases Forming Helical Nucleoprotein Filaments. Mol. Cell Biol. 2005, 25, 4377–4387. [Google Scholar] [CrossRef] [Green Version]
  164. Aylon, Y.; Liefshitz, B.; Bitan-Banin, G.; Kupiec, M. Molecular Dissection of Mitotic Recombination in the Yeast Saccharomyces Cerevisiae. Mol. Cell Biol. 2003, 23, 1403–1417. [Google Scholar] [CrossRef] [Green Version]
  165. Bzymek, M.; Thayer, N.H.; Oh, S.D.; Kleckner, N.; Hunter, N. Double Holliday Junctions Are Intermediates of DNA Break Repair. Nature 2010, 464, 937–941. [Google Scholar] [CrossRef] [Green Version]
  166. Grishchuk, A.L.; Kohli, J. Five RecA-like Proteins of Schizosaccharomyces Pombe Are Involved in Meiotic Recombination. Genetics 2003, 165, 1031–1043. [Google Scholar] [CrossRef]
  167. Segurado, M.; Gómez, M.; Antequera, F. Increased Recombination Intermediates and Homologous Integration Hot Spots at DNA Replication Origins. Mol. Cell 2002, 10, 907–916. [Google Scholar] [CrossRef]
  168. Lambert, S.; Watson, A.; Sheedy, D.M.; Martin, B.; Carr, A.M. Gross Chromosomal Rearrangements and Elevated Recombination at an Inducible Site-Specific Replication Fork Barrier. Cell 2005, 121, 689–702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Nemoz, C.; Ropars, V.; Frit, P.; Gontier, A.; Drevet, P.; Yu, J.; Guerois, R.; Pitois, A.; Comte, A.; Deltiel, C.; et al. XLF and APLF Bind Ku80 at Two Remote Sites to Ensure DNA Repair by Non-Homologous End Joining. Nat. Struct. Mol. Biol. 2018, 25, 971–980. [Google Scholar] [CrossRef] [PubMed]
  170. Meek, K.; Dang, V.; Lees-Miller, S.P. DNA-PK: The Means to Justify the Ends? Adv. Immunol. 2008, 99, 33–58. [Google Scholar]
  171. Sibanda, B.L.; Chirgadze, D.Y.; Ascher, D.B.; Blundell, T.L. DNA-PKcs Structure Suggests an Allosteric Mechanism Modulating DNA Double-Strand Break Repair. Science 2017, 355, 520–524. [Google Scholar] [CrossRef] [Green Version]
  172. Ma, Y.; Pannicke, U.; Schwarz, K.; Lieber, M.R. Hairpin Opening and Overhang Processing by an Artemis/DNA-Dependent Protein Kinase Complex in Nonhomologous End Joining and V(D)J Recombination. Cell 2002, 108, 781–794. [Google Scholar] [CrossRef] [Green Version]
  173. Hammel, M.; Rey, M.; Yu, Y.; Mani, R.S.; Classen, S.; Liu, M.; Pique, M.E.; Fang, S.; Mahaney, B.L.; Weinfeld, M.; et al. XRCC4 Protein Interactions with XRCC4-like Factor (XLF) Create an Extended Grooved Scaffold for DNA Ligation and Double Strand Break Repair. J. Biol. Chem. 2011, 286, 32638–32650. [Google Scholar] [CrossRef] [Green Version]
  174. Ochi, T.; Blackford, A.N.; Coates, J.; Jhujh, S.; Mehmood, S.; Tamura, N.; Travers, J.; Wu, Q.; Draviam, V.M.; Robinson, C.V. DNA Repair. PAXX, a Paralog of XRCC4 and XLF, Interacts with Ku to Promote DNA Double-Strand Break Repair. Science 2015, 347, 185–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Frank-Vaillant, M.; Marcand, S. NHEJ Regulation by Mating Type Is Exercised through a Novel Protein, Lif2p, Essential to the Ligase IV Pathway. Genes Dev. 2001, 15, 3005–3012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Kegel, A.; Sjöstrand, J.O.O.; Åström, S.U. Nej1p, a Cell Type-Specific Regulator of Nonhomologous End Joining in Yeast. Curr. Biol. 2001, 11, 1611–1617. [Google Scholar] [CrossRef] [Green Version]
  177. Valencia, M.; Bentele, M.; Vaze, M.B.; Herrmann, G.; Kraus, E.; Lee, S.E.; Schär, P.; Haber, J.E. NEJ1 Controls Non-Homologous End Joining in Saccharomyces Cerevisiae. Nature 2001, 414, 666–669. [Google Scholar] [CrossRef] [PubMed]
  178. Dudášová, Z.; Dudáš, A.; Chovanec, M. Non-Homologous End-Joining Factors of Saccharomyces Cerevisiae. FEMS Microbiol. Rev. 2004, 28, 581–601. [Google Scholar] [CrossRef] [Green Version]
  179. Daley, J.M.; Niu, H.; Miller, A.S.; Sung, P. Biochemical Mechanism of DSB End Resection and Its Regulation. DNA Repair 2015, 32, 66–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Callebaut, I.; Malivert, L.; Fischer, A. Cernunnos Interacts with the XRCC4–DNA-Ligase IV Complex and Is Homologous to the Yeast Nonhomologous End-Joining Factor Nej1. J. Biol. Chem. 2006, 281, 13857–13860. [Google Scholar] [CrossRef] [Green Version]
  181. Li, Y.; Wang, J.; Zhou, G.; Lajeunesse, M.; Le, N.; Stawicki, B.N.; Corcino, Y.L.; Berkner, K.L.; Runge, K.W. Nonhomologous End-Joining with Minimal Sequence Loss Is Promoted by the Mre11-Rad50-Nbs1-Ctp1 Complex in Schizosaccharomyces Pombe. Genetics 2017, 206, 481–496. [Google Scholar] [CrossRef] [Green Version]
  182. Runge, K.W.; Li, Y. A Curious New Role for MRN in Schizosaccharomyces Pombe Non-Homologous End-Joining. Curr. Genet. 2018, 64, 359–364. [Google Scholar] [CrossRef]
  183. Paques, F.; Haber, J.E. Multiple Pathways of Recombination Induced by Double-Strand Breaks in Saccharomyces Cerevisiae. Microbiol. Mol. Biol. Rev. 1999, 63, 349–404. [Google Scholar] [CrossRef] [Green Version]
  184. Zhao, F.; Kim, W.; Kloeber, J.A.; Lou, Z. DNA End Resection and Its Role in DNA Replication and DSB Repair Choice in Mammalian Cells. Exp. Mol. Med. 2020, 52, 1705–1714. [Google Scholar] [CrossRef]
  185. Ma, C.J.; Gibb, B.; Kwon, Y.; Sung, P.; Greene, E.C. Protein Dynamics of Human RPA and RAD51 on SsDNA during Assembly and Disassembly of the RAD51 Filament. Nucleic Acids Res. 2017, 45, 749–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Motycka, T.A.; Bessho, T.; Post, S.M.; Sung, P.; Tomkinson, A.E. Physical and Functional Interaction between the XPF/ERCC1 Endonuclease and HRad52. J. Biol. Chem. 2004, 279, 13634–13639. [Google Scholar] [CrossRef] [Green Version]
  187. Osman, F.; Fortunato, E.A.; Subramani, S. Double-Strand Break-Induced Mitotic Intrachromosomal Recombination in the Fission Yeast Schizosaccharomyces Pombe. Genetics 1996, 142, 341–357. [Google Scholar] [CrossRef] [PubMed]
  188. Shinohara, A.; Ogawa, H.; Matsuda, Y.; Ushio, N.; Ikeo, K.; Ogawa, T. Cloning of Human, Mouse and Fission Yeast Recombination Genes Homologous to RAD51 and RecA. Nat. Genet. 1993, 4, 239–243. [Google Scholar] [CrossRef]
  189. Moore, C.W. Responses of Radiation-Sensitive Mutants of Saccharomyces Cerevisiae to Lethal Effects of Bleomycin. Mutat. Res. 1978, 51, 165–180. [Google Scholar] [CrossRef]
  190. Fabre, F.; Moustacchi, E. Removal of Pyrimidine Dimers in Cells of Schizosaccharomyces Pombe Mutated in Different Repair Pathways. Biochim. Biophys. Acta (BBA) Nucleic Acids Protein Synth. 1973, 312, 617–625. [Google Scholar] [CrossRef]
  191. Avery, A.M.; Kaur, B.; Taylor, J.-S.; Mello, J.A.; Essigmann, J.M.; Doetsch, P.W. Substrate Specificity of Ultraviolet DNA Endonuclease (UVDE/Uve1p) from Schizosaccharomyces Pombe. Nucleic Acids Res. 1999, 27, 2256–2264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Yasui, A.; McCready, S.J. Alternative Repair Pathways for UV-Induced DNA Damage. BioEssays 1998, 20, 291–297. [Google Scholar] [CrossRef]
  193. Kanno, S.; Shigenori, I.; Takao, M.; Yasui, A. Repair of Apurinic/Apyrimidinic Sites by UV Damage Endonuclease; a Repair Protein for UV and Oxidative Damage. Nucleic Acids Res. 1999, 27, 3096–3103. [Google Scholar] [CrossRef] [Green Version]
  194. Schupbach, M. The Isolation and Genetic Characterization of UV-Sensitive Mutants of Schizosaccharomyces Pombe. Mutat. Res. 1971, 11, 361–371. [Google Scholar]
  195. Nasim, A.; Smith, B.P. Dark Repair Inhibitors and Pathways for Repair of Radiation Damage in Schizosaccharomyces Pombe. Molec. Gen. Genet. 1974, 132, 13–22. [Google Scholar] [CrossRef]
  196. Lehmann, A.R.; Walicka, M.; Griffiths, D.J.; Murray, J.M.; Watts, F.Z.; McCready, S.; Carr, A.M. The Rad18 Gene of Schizosaccharomyces Pombe Defines a New Subgroup of the SMC Superfamily Involved in DNA Repair. Mol. Cell. Biol. 1995, 15, 7067–7080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Szanski, P.; Smith, G.R. A Role for Exonuclease I from S. Pombe in Mutation Avoidance and Mismatch Correction. Science 1995, 267, 1166–1169. [Google Scholar]
  198. Resnick, M.A. A Photoreactivationless Mutant of Saccharomyces Cerevisiae. Photochem. Photobiol. 1969, 9, 307–312. [Google Scholar] [CrossRef] [PubMed]
  199. Schild, D.; Johnston, J.; Chang, C.; Mortimer, R.K. Cloning and Mapping of Saccharomyces Cerevisiae Photoreactivation Gene PHR1. Mol. Cell. Biol. 1984, 4, 1864–1870. [Google Scholar]
  200. Sebastian, J.; Kraus, B.; Sancar, G.B. Expression of the Yeast PHR1 Gene Is Induced by DNA-Damaging Agents. Mol. Cell. Biol. 1990, 10, 4630–4637. [Google Scholar] [PubMed] [Green Version]
  201. Sancar, A. Structure and Function of Photolyase and in Vivo Enzymology: 50th Anniversary. J. Biol. Chem. 2008, 283, 32153–32157. [Google Scholar] [CrossRef] [Green Version]
  202. Sassanfar, M.; Samson, L. Identification and preliminary characterization of an O6-methylguanine DNA repair methyltransferase in the yeast Saccharomyces cerevisiae. J. Biol. Chem. 1990, 265, 20–25. [Google Scholar] [CrossRef]
  203. Xiao, W.; Derfler, B.; Chen, J.; Samson, L. Primary Sequence and Biological Functions of a Saccharomyces Cerevisiae O6-Methylguanine/O4-Methylthymine DNA Repair Methyltransferase Gene. EMBO J. 1991, 10, 2179–2186. [Google Scholar] [CrossRef] [PubMed]
  204. Costa, Y.; Speed, R.; Öllinger, R.; Alsheimer, M.; Semple, C.A.; Gautier, P.; Maratou, K.; Novak, I.; Höög, C.; Benavente, R.; et al. Two Novel Proteins Recruited by Synaptonemal Complex Protein 1 (SYCP1) Are at the Centre of Meiosis. J. Cell Sci. 2005, 118, 2755–2762. [Google Scholar] [CrossRef] [Green Version]
  205. Marcon, L.; Boissonneault, G. Transient DNA Strand Breaks During Mouse and Human Spermiogenesis:New Insights in Stage Specificity and Link to Chromatin Remodeling. Biol. Reprod. 2004, 70, 910–918. [Google Scholar] [CrossRef]
  206. Sakkas, D.; Seli, E.; Manicardi, G.C.; Nijs, M.; Ombelet, W.; Bizzaro, D. The Presence of Abnormal Spermatozoa in the Ejaculate: Did Apoptosis Fail? Hum. Fertil. 2004, 7, 99–103. [Google Scholar] [CrossRef]
  207. Agarwal, A.; Saleh, R.A.; Bedaiwy, M.A. Role of Reactive Oxygen Species in the Pathophysiology of Human Reproduction. Fertil. Steril. 2003, 79, 829–843. [Google Scholar] [CrossRef] [Green Version]
  208. Gosalvez, J.; Alvarez, J.G.; Fernandez, J.L.; Gosalbez, A.; Gonzalez-Marin, C.; Lopez-Fernandez, C. Assessment of Single and Double-Stranded DNA Damage Following Short- and Long-Term Incubation Post-Thawing of Cryopreserved Human Spermatozoa. Biol. Reprod. 2011, 85, 501. [Google Scholar] [CrossRef]
  209. Torregrosa, N.; Domínguez-Fandos, D.; Camejo, M.I.; Shirley, C.R.; Meistrich, M.L.; Ballescà, J.L.; Oliva, R. Protamine 2 Precursors, Protamine 1/Protamine 2 Ratio, DNA Integrity and Other Sperm Parameters in Infertile Patients. Hum. Reprod. 2006, 21, 2084–2089. [Google Scholar] [CrossRef]
  210. Akematsu, T.; Fukuda, Y.; Garg, J.; Fillingham, J.S.; Pearlman, R.E.; Loidl, J. Post-Meiotic DNA Double-Strand Breaks Occur in Tetrahymena, and Require Topoisomerase II and Spo11. eLife 2017, 6, e26176. [Google Scholar] [CrossRef] [PubMed]
  211. Aoki, V.W.; Moskovtsev, S.I.; Willis, J.; Liu, L.; Mullen, J.B.M.; Carrell, D.T. DNA Integrity Is Compromised in Protamine-Deficient Human Sperm. J. Androl. 2005, 26, 741–748. [Google Scholar] [CrossRef] [PubMed]
  212. Carrell, D.T.; Emery, B.R.; Hammoud, S. Altered Protamine Expression and Diminished Spermatogenesis: What is the link? Hum. Reprod. Update 2007, 13, 313–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. De Mateo, S.; Gázquez, C.; Guimerà, M.; Balasch, J.; Meistrich, M.L.; Ballescà, J.L.; Oliva, R. Protamine 2 Precursors (Pre-P2), Protamine 1 to Protamine 2 Ratio (P1/P2), and Assisted Reproduction Outcome. Fertil. Steril. 2009, 91, 715–722. [Google Scholar] [CrossRef]
  214. Derijck, A.; van der Heijden, G.; Giele, M.; Philippens, M.; de Boer, P. DNA Double-Strand Break Repair in Parental Chromatin of Mouse Zygotes, the First Cell Cycle as an Origin of de Novo Mutation. Hum. Mol. Genet. 2008, 17, 1922–1937. [Google Scholar] [CrossRef] [Green Version]
  215. Aitken, R.J.; Iuliis, G.N.D.; McLachlan, R.I. Biological and Clinical Significance of DNA Damage in the Male Germ Line. Int. J. Androl. 2009, 32, 46–56. [Google Scholar] [CrossRef]
  216. Leduc, F.; Maquennehan, V.; Nkoma, G.B.; Boissonneault, G. DNA Damage Response During Chromatin Remodeling in Elongating Spermatids of Mice. Biol. Reprod. 2008, 78, 324–332. [Google Scholar] [CrossRef]
  217. Sun, G.; Guzman, E.; Balasanyan, V.; Conner, C.M.; Wong, K.; Zhou, H.R.; Kosik, K.S.; Montell, D.J. A Molecular Signature for Anastasis, Recovery from the Brink of Apoptotic Cell Death. J. Cell. Biol. 2017, 216, 3355–3368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Shaha, C.; Tripathi, R.; Mishra, D.P. Male Germ Cell Apoptosis: Regulation and Biology. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 1501–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Aitken, R.J.; Smith, T.B.; Jobling, M.S.; Baker, M.A.; De Iuliis, G.N. Oxidative Stress and Male Reproductive Health. Asian J. 2014, 16, 31–38. [Google Scholar] [CrossRef]
  220. Kothari, S.; Thompson, A.; Agarwal, A.; du Plessis, S.S. Free Radicals: Their Beneficial and Detrimental Effects on Sperm Function. Indian J. Exp. Biol. 2010, 48, 425–435. [Google Scholar]
  221. Kemal Duru, N.; Morshedi, M.; Oehninger, S. Effects of Hydrogen Peroxide on DNA and Plasma Membrane Integrity of Human Spermatozoa. Fertil. Steril. 2000, 74, 1200–1207. [Google Scholar] [CrossRef]
  222. Burden, H.P.; Holmes, C.H.; Persad, R.; Whittington, K. Prostasomes—Their Effects on Human Male Reproduction and Fertility. Hum. Reprod. Update 2006, 12, 283–292. [Google Scholar] [CrossRef] [Green Version]
  223. Agarwal, A.; Durairajanayagam, D.; du Plessis, S.S. Utility of Antioxidants during Assisted Reproductive Techniques: An Evidence Based Review. Reprod. Biol. Endocrinol. 2014, 12, 112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Xu, G.; Spivak, G.; Mitchell, D.L.; Mori, T.; McCarrey, J.R.; McMahan, C.A.; Walter, R.B.; Hanawalt, P.C.; Walter, C.A. Nucleotide Excision Repair Activity Varies Among Murine Spermatogenic Cell Types1. Biol. Reprod. 2005, 73, 123–130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Goukassian, D.; Gad, F.; Yaar, M.; Eller, M.S.; Nehal, U.S.; Gilchrest, B.A. Mechanisms and Implications of the Age-Associated Decrease in DNA Repair Capacity. FASEB J. 2000, 14, 1325–1334. [Google Scholar] [CrossRef]
  226. Jansen, J.; Olsen, A.K.; Wiger, R.; Naegeli, H.; de Boer, P.; van der Hoeven, F.; Holme, J.A.; Brunborg, G.; Mullenders, L. Nucleotide Excision Repair in Rat Male Germ Cells: Low Level of Repair in Intact Cells Contrasts with High Dual Incision Activity in Vitro. Nucleic Acids Res. 2001, 29, 1791–1800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Sassone-Corsi, P. Unique Chromatin Remodeling and Transcriptional Regulation in Spermatogenesis. Science 2002, 296, 2176–2178. [Google Scholar] [CrossRef]
  228. Ahmed, E.A.; Scherthan, H.; De Rooij, D.G. DNA Double Strand Break Response and Limited Repair Capacity in Mouse Elongated Spermatids. Int. J. Mol. Sci. 2015, 16, 29923–29935. [Google Scholar] [CrossRef] [Green Version]
  229. Zenzes, M.T. Smoking and Reproduction: Gene Damage to Human Gametes and Embryos. Hum. Reprod. Update 2000, 6, 122–131. [Google Scholar] [CrossRef]
  230. Stringer, J.M.; Winship, A.; Liew, S.H.; Hutt, K. The Capacity of Oocytes for DNA Repair. Cell. Mol. Life Sci. 2018, 75, 2777–2792. [Google Scholar] [CrossRef]
  231. Menezo, Y.J.; Russo, G.; Tosti, E.; Mouatassim, S.E.; Benkhalifa, M. Expression Profile of Genes Coding for DNA Repair in Human Oocytes Using Pangenomic Microarrays, with a Special Focus on ROS Linked Decays. J. Assist. Reprod. Genet. 2007, 24, 513–520. [Google Scholar] [CrossRef] [Green Version]
  232. Jaroudi, S.; Kakourou, G.; Cawood, S.; Doshi, A.; Ranieri, D.M.; Serhal, P.; Harper, J.C.; SenGupta, S.B. Expression Profiling of DNA Repair Genes in Human Oocytes and Blastocysts Using Microarrays. Hum. Reprod. 2009, 24, 2649–2655. [Google Scholar] [CrossRef] [Green Version]
  233. Zeng, F.; Baldwin, D.A.; Schultz, R.M. Transcript Profiling during Preimplantation Mouse Development. Dev. Biol. 2004, 272, 483–496. [Google Scholar] [CrossRef] [Green Version]
  234. Zheng, P.; Schramm, R.D.; Latham, K.E. Developmental Regulation and In Vitro Culture Effects on Expression of DNA Repair and Cell Cycle Checkpoint Control Genes in Rhesus Monkey Oocytes and Embryos1. Biol. Reprod. 2005, 72, 1359–1369. [Google Scholar] [CrossRef] [Green Version]
  235. Martin, J.H.; Bromfield, E.G.; Aitken, R.J.; Lord, T.; Nixon, B. Double Strand Break DNA Repair Occurs via Non-Homologous End-Joining in Mouse MII Oocytes. Sci. Rep. 2018, 8, 9685. [Google Scholar] [CrossRef]
  236. Hamatani, T.; Falco, G.; Carter, M.G.; Akutsu, H.; Stagg, C.A.; Sharov, A.A.; Dudekula, D.B.; VanBuren, V.; Ko, M.S.H. Age-Associated Alteration of Gene Expression Patterns in Mouse Oocytes. Hum. Mol. Genet. 2004, 13, 2263–2278. [Google Scholar] [CrossRef] [Green Version]
  237. Matsuda, Y.; Tobari, I.; Yamagiwa, J.; Utsugi, T.; Okamoto, M.; Nakai, S. Dose-Response Relationship of γ-Ray-Induced Reciprocal Translocations at Low Doses in Spermatogonia of the Crab-Eating Monkey (Macaca Fascicularis). Mutat. Res./Fundam. Mol. Mech. Mutagenesis 1985, 151, 121–127. [Google Scholar] [CrossRef]
  238. Agarwal, A.; Aziz, N.; Rizk, B. Studies on Women’s Health; Humana Press: Totowa, NJ, USA, 2013. [Google Scholar]
  239. Richards, J.S.; Russell, D.L.; Ochsner, S.; Espey, L.L. Ovulation: New Dimensions and New Regulators of the Inflammatory-Like Response. Annu. Rev. Physiol. 2002, 64, 69–92. [Google Scholar] [CrossRef] [PubMed]
  240. Revelli, A.; Piane, L.D.; Casano, S.; Molinari, E.; Massobrio, M.; Rinaudo, P. Follicular Fluid Content and Oocyte Quality: From Single Biochemical Markers to Metabolomics. Reprod. Biol. Endocrinol. 2009, 7, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Krisher, R.L. In Vivo and In Vitro Environmental Effects on Mammalian Oocyte Quality. Annu. Rev. Anim. Biosci. 2013, 1, 393–417. [Google Scholar] [CrossRef] [PubMed]
  242. Du Plessis, S.S.; Makker, K.; Desai, N.R.; Agarwal, A. Impact of Oxidative Stress on IVF. Expert Rev. Obstet. Gynecol. 2008, 3, 539–554. [Google Scholar] [CrossRef]
  243. Borini, A.; Tarozzi, N.; Bizzaro, D.; Bonu, M.A.; Fava, L.; Flamigni, C.; Coticchio, G. Sperm DNA Fragmentation: Paternal Effect on Early Post-Implantation Embryo Development in ART. Hum. Reprod. 2006, 21, 2876–2881. [Google Scholar] [CrossRef] [PubMed]
  244. Braude, P.; Bolton, V.; Moore, S. Human Gene Expression First Occurs between the Four- and Eight-Cell Stages of Preimplantation Development. Nature 1988, 332, 459–461. [Google Scholar] [CrossRef] [PubMed]
  245. Ahmadi, A.; Ng, S. Fertilizing Ability of DNA-damaged Spermatozoa. J. Exp. Zool. 1999, 284, 696–704. [Google Scholar] [CrossRef]
  246. Genescà, A.; Caballín, M.R.; Miró, R.; Benet, J.; Germà, J.R.; Egozcue, J. Repair of Human Sperm Chromosome Aberrations in the Hamster Egg. Hum. Genet. 1992, 89, 181–186. [Google Scholar] [CrossRef] [PubMed]
  247. Gawecka, J.E.; Marh, J.; Ortega, M.; Yamauchi, Y.; Ward, M.A.; Ward, W.S. Mouse Zygotes Respond to Severe Sperm DNA Damage by Delaying Paternal DNA Replication and Embryonic Development. PLoS ONE 2013, 8, e56385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Origin, type, and repair pathways of DNA damages. Various inducers of DNA damage are represented (top line), whereas the type of damage is represented for each (middle line). The circles (bottom line) represent known repair pathways. Darker circles represent pathways found through the eukaryotic domain whereas circles in light grey are those specific to yeast. (Abbreviations: ROS: Reactive Oxygen Species; SSB: Single Strand break; DSB: Double-Strand Break; CPD: Cyclo-Pyrimidine Dimer; NHEJ: Non-Homologous End-Joining; HR: Homologous Recombination; SSA: Single Strand Annealing; Alt-NHEJ: Alternative NHEJ; BER: Base Excision Repair; NER: Nucleotide Excision Repair; UVDER: UV Damage Excision Repair; MMR: Mismatch Repair).
Figure 1. Origin, type, and repair pathways of DNA damages. Various inducers of DNA damage are represented (top line), whereas the type of damage is represented for each (middle line). The circles (bottom line) represent known repair pathways. Darker circles represent pathways found through the eukaryotic domain whereas circles in light grey are those specific to yeast. (Abbreviations: ROS: Reactive Oxygen Species; SSB: Single Strand break; DSB: Double-Strand Break; CPD: Cyclo-Pyrimidine Dimer; NHEJ: Non-Homologous End-Joining; HR: Homologous Recombination; SSA: Single Strand Annealing; Alt-NHEJ: Alternative NHEJ; BER: Base Excision Repair; NER: Nucleotide Excision Repair; UVDER: UV Damage Excision Repair; MMR: Mismatch Repair).
Ijms 22 12418 g001
Figure 2. DNA base repair pathways. Simplified schematic representation of Nucleotide Excision Repair (NER), Base Excision Repair (BER), and Mismatch Repair (MMR). Transcription Coupled NER (TC-NER) starts with RNA pol II stalling in front of a lesion. Then CSB recruits CSA and UVSSA-USP7 allowing the TFIIH docking. Global Genome NER (GG-NER) is initiated by XPC lesion recognition. Through a cascade of events, TFIIH is recruited. Both NER pathways converge when TFIIH is recruited. Thanks to its helicases, XPB/XPD, and XPA, TFIIH creates a DNA repair bubble exposing the lesion. Then ERCC1-XPF and XPG cleave DNA around the lesion. The gap is then filled by DNA polymerase. In BER, glycosylase targets the lesion and cleaves base-deoxyribose glycosidic bounds. When Glycosylase is released from DNA, APE1 can now bind to cleave the faulty deoxyribose. DNA polymerase fills the gap and displaces some of the damaged strands. The resulting 3′ tail is then removed by FEN1. In MMR the mismatch is targeted by MutS, then recruiting MutL. Then an endonuclease cleaves DNA, and a resection occurs. Finally, polymerase fills the gap and could create a 3′ overhang tail.
Figure 2. DNA base repair pathways. Simplified schematic representation of Nucleotide Excision Repair (NER), Base Excision Repair (BER), and Mismatch Repair (MMR). Transcription Coupled NER (TC-NER) starts with RNA pol II stalling in front of a lesion. Then CSB recruits CSA and UVSSA-USP7 allowing the TFIIH docking. Global Genome NER (GG-NER) is initiated by XPC lesion recognition. Through a cascade of events, TFIIH is recruited. Both NER pathways converge when TFIIH is recruited. Thanks to its helicases, XPB/XPD, and XPA, TFIIH creates a DNA repair bubble exposing the lesion. Then ERCC1-XPF and XPG cleave DNA around the lesion. The gap is then filled by DNA polymerase. In BER, glycosylase targets the lesion and cleaves base-deoxyribose glycosidic bounds. When Glycosylase is released from DNA, APE1 can now bind to cleave the faulty deoxyribose. DNA polymerase fills the gap and displaces some of the damaged strands. The resulting 3′ tail is then removed by FEN1. In MMR the mismatch is targeted by MutS, then recruiting MutL. Then an endonuclease cleaves DNA, and a resection occurs. Finally, polymerase fills the gap and could create a 3′ overhang tail.
Ijms 22 12418 g002
Figure 3. Double-Strand Breaks (DSBs) repair pathways. The repair of a DSB begins with DNA ends resection or direct ligation through Non-Homologous End-Joining (NHEJ). If resection does occur, Homologous Recombination (HR), Single Strand Annealing (SSA), and Alternative-NHEJ (Alt-NHEJ) compete to repair the DSB. NHEJ begins with Ku dimers binding to DNA ends and recruiting of DNA-Pkc. Artemis endonuclease is then recruited in order to clean dirty DNA ends, allowing the direct ligation by Ligase IV. Resection is carried out by exonuclease I (EXO1) but initiated by the MRN complex, assisted by CtIP. Single strand DNA is then protected by RPA. In HR, Rad52 will displace some RPA to allow the formation of Rad51 filament. Rad51 filament will find a homologous sequence and start strand invasion. DNA polymerase then fills gaps by using homologous chromatid as a template. Resolution of Holliday Junction leads to Chromosomal Exchange (CO) or Non-Chromosomal Exchange (NCO). SSA also uses Rad52 and Rad51, but the filaments target homologies surrounding the DSB. When hybridization occurs, unpaired 5′ overhangs are removed by ERCC1/XPF. Finally, polymerase fills the gap. Alt-NHEJ simply uses thermodynamics for annealing close to the DSB. Then DNA polymerase θ fills the gap.
Figure 3. Double-Strand Breaks (DSBs) repair pathways. The repair of a DSB begins with DNA ends resection or direct ligation through Non-Homologous End-Joining (NHEJ). If resection does occur, Homologous Recombination (HR), Single Strand Annealing (SSA), and Alternative-NHEJ (Alt-NHEJ) compete to repair the DSB. NHEJ begins with Ku dimers binding to DNA ends and recruiting of DNA-Pkc. Artemis endonuclease is then recruited in order to clean dirty DNA ends, allowing the direct ligation by Ligase IV. Resection is carried out by exonuclease I (EXO1) but initiated by the MRN complex, assisted by CtIP. Single strand DNA is then protected by RPA. In HR, Rad52 will displace some RPA to allow the formation of Rad51 filament. Rad51 filament will find a homologous sequence and start strand invasion. DNA polymerase then fills gaps by using homologous chromatid as a template. Resolution of Holliday Junction leads to Chromosomal Exchange (CO) or Non-Chromosomal Exchange (NCO). SSA also uses Rad52 and Rad51, but the filaments target homologies surrounding the DSB. When hybridization occurs, unpaired 5′ overhangs are removed by ERCC1/XPF. Finally, polymerase fills the gap. Alt-NHEJ simply uses thermodynamics for annealing close to the DSB. Then DNA polymerase θ fills the gap.
Ijms 22 12418 g003
Table 1. Table of homologous proteins.
Table 1. Table of homologous proteins.
H. sapiensS. cerevisiaeS. pombeFunctions
TC NERRNA pol IIRNA pol IIRNA pol IImRNA synthesis
CSBRad 26Rhp 26Detection of RNA pol II arrest
-Rpd 9-Part of RNA pol II, detection of arrest
CSARad 28Ckn1TC-NER protein hub
UVSSA-USP7--Ubiquitine ligase
GG NERXPCRad4rhp41, rhp42Scan to detect damages
Rad23bRad23Rhp 23XPC partner
TFIIHTFIIHTFIIHTranscription factor, check for damage
XPBRad 25Ptr 8Core units of TFIIH (helicases)
XPDRad 3Rad 15
XPARad 14Rph 14TFIIH helicases stimulator
ERCCI-XPFRad 10Swi 10Endonuclease complex, cleaves 5′ to damage
XPFRad 1Rad 16
XPGRad 2Rad 13Endonuclease, cleaves 3′ to damage
BERUNGUng 1Ung 1Uracil DNA Glycosylase (UDG)
TDG-Thp 1UDG, T:G, U:G mismatches, 5-fluorouracil, 3,N4-ethanolcytosine, 5-hydroxyuracil, Xanthine, Oxanine, Hypoxanthine DNA glycosylase
-Mag 1Mag 1Alkylation DNA Glycosylase
-Mag 2Mag 2
OGG 1Ogg 1-8-oxoG, fapy-G, 7,8-dihydro8oxoG DNA Glycosylase
NTHL 1Ntg1/Ntg2Nth 18-oxoG, 8-hydroxycytosine, thimidineglycol, 8-hydroxyuracil DNA Glycosylase, AP site β lyase
MUTYH-Myh 1Glycosylase of adenine mismatches
APE1Apn 2Apn 2AP endonucleases
-Apn 1Apn 1
FEN 1Rad 27Rad 2Structure specific endonuclease, cleaves at ssDNA-dsDNA transition
MMRMutS αMutS αMutS αMismatch detection
MutS βMutS βMutS βLong insertion deletion loops detection, involved in recombination
MutS γMutS γMutS γHolliday Junction resolvase
MutL αMutL αMutL αMutSα and MutSβ interactor, weakly endonuclease
MutL βMutL β-
MutL γMutL γ-MutSγ interactor
EXO 1EXO 1EXO 1Exonuclease
PCNAPCNAPCNADNA polymerase helicase
UVDER--UVE 1UV damage endonucleases
--UVDE
Photolyase-Phr1-CPD specific photolyase
-Mgt1-O6-MeG, O4-MeT methyl transferase
HRMre11/Rad50/Nbs1Mre11/Rad50/Xrs1Mre11/Rad50/Nbs1MRN/X complex, DNA ends detection, resection initiation
CtIPSae2Ctp 1Endonuclease
RPARPARad 11Single strand DNA binding protein
Rad 52Rad 52Rad 22Displace RPA to form Rad51 filament
Rad51Rad 51Rhp 51Globular protein forming filaments for strand invasion
NHEJKu 70/80YKu 70/80PKu 70/80DNA ends detection and protection
DNA-Pkc--Protein kinase
Artemis--Endonuclease, DNA ends processing
XRCC4Nej 1-DNA ends bridging by filaments formation
XLFLif 1Xlfl
DNA ligase IVDnl 4Lig 4DNA ends ligation
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mourrain, L.; Boissonneault, G. DNA Repair in Haploid Context. Int. J. Mol. Sci. 2021, 22, 12418. https://doi.org/10.3390/ijms222212418

AMA Style

Mourrain L, Boissonneault G. DNA Repair in Haploid Context. International Journal of Molecular Sciences. 2021; 22(22):12418. https://doi.org/10.3390/ijms222212418

Chicago/Turabian Style

Mourrain, Loïs, and Guylain Boissonneault. 2021. "DNA Repair in Haploid Context" International Journal of Molecular Sciences 22, no. 22: 12418. https://doi.org/10.3390/ijms222212418

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