Open access peer-reviewed chapter

Regulation of Oxidized Base Repair in Human Chromatin by Posttranslational Modification

Written By

Shiladitya Sengupta, Chunying Yang, Bradley J. Eckelmann, Muralidhar L. Hegde and Sankar Mitra

Submitted: 20 June 2018 Reviewed: 12 October 2018 Published: 11 December 2018

DOI: 10.5772/intechopen.81979

From the Edited Volume

DNA Repair- An Update

Edited by Maddalena Mognato

Chapter metrics overview

1,127 Chapter Downloads

View Full Metrics

Abstract

Base excision repair (BER) is the major pathway for the repair of oxidized bases and apurinic/apyrimidinic (abasic; AP) sites produced by reaction with reactive oxygen/nitrogen species (ROS/RNS). These metabolites are generated spontaneously by endogenous cellular processes and also by environmental agents. Because most of these lesions are promutagenic, linked to diverse disease-associated somatic mutations, as well as heritable single nucleotide polymorphisms (SNPs) in the normal human population, their prompt repair is warranted. Impairment of repair leading to mutation, a hallmark of cancer, underscores the essentiality of BER for maintaining genome integrity in humans and other mammals. In mammals, repair of oxidized bases and other BER substrates is initiated by DNA glycosylases (DGs), which excise the damaged bases and cleave the DNA strands at the resulting AP sites, followed by sequential end processing, gap-filling DNA synthesis, and ligation. In vitro BER performed with naked DNA substrates has been extensively studied, which delineates its basic mechanistic steps and subpathways. However, recent interest is directed to unraveling BER in cell chromatin, including its regulation via posttranslational modifications (PTMs), which occurs possibly in concert with nucleosome remodeling. Emerging reports on various PTMs of BER enzymes indicate that the PTMs, while dispensable for the enzymatic activity, regulate overall repair by modulating interactions with other repair proteins and chromatin factors, assembly of BER complexes, as well as turnover of the proteins, and may ultimately dictate the cellular phenotype. Here, we discuss recent advances in the BER field by reviewing the PTMs and how they regulate BER in chromatin.

Keywords

  • oxidative stress
  • base oxidation
  • base excision repair
  • posttranslational modifications
  • acetylation
  • phosphorylation
  • SUMOylation
  • methylation
  • chromatin

1. Introduction

DNA, the genetic repository of all cellular functions, is packaged with histones into chromatin consisting of nucleosome units. One hundred forty-seven base pair (bp) segments in DNA wrap ~1.65 times in a left-handed superhelical turn around a histone octamer consisting of two histone H2A-H2B dimers and a H3-H4 tetramer, which form the nucleosome core; the adjacent nucleosomes are separated by some 50 bp unfolded, linker DNA bound to histone H1 or H5. Organization of DNA into chromatin enables the compaction required to accommodate large eukaryotic genomes inside the cell nucleus. This compaction renders DNA inaccessible to any DNA transaction machinery. Replication and transcription are tightly coordinated with specific interactions of their complexes with DNA [1, 2].

The integrity of DNA is under constant threat, naturally from endogenous sources, as well as by environmental factors in the form of a chemical addition, an alteration in the nitrogen base structure, thereby creating an abnormal nucleotide, or a break in one or both strands of DNA [3, 4, 5, 6, 7, 8]. Cellular metabolic processes including mitochondrial respiration and hydrolytic reactions generate reactive molecules, such as reactive oxygen species (ROS), reactive nitrogen species (RNS), and alkylating agents. Some chemical bonds in DNA are susceptible to spontaneous hydrolysis. About 70,000 lesions are generated per cell, per day in humans. Single-strand breaks (SSBs), as well as a plethora of oxidized bases, are formed during oxidative genome damage. In addition, deamination, depurination, depyrimidination, double-strand breaks (DSBs), propano-, etheno-, and malondialdehyde-derived DNA adducts, base propenals, and alkylated bases are also formed endogenously. Environmental factors such as UV rays, ionizing radiation (IR), heat, and chemicals from tobacco smoke and industrial sources pose additional risks to DNA.

Advertisement

2. Oxidative genome damage and oxidized bases

For aerobic organisms, oxygen acts like a double-edged sword; while it is absolutely essential for life, it is also a threat to the life, recognized as the “Oxygen Paradox” [9, 10, 11]. ROS, which include the superoxide anion (O2•−), hydrogen peroxide (H2O2), singlet oxygen (1O2), and the hydroxyl radical (OH), along with RNS, for example, peroxynitrite (ONOO) react with all biological molecules including DNA. The hydroxyl radical having the highest reduction potential is mainly generated from Fenton reaction between reduced redox active metal ions (Fe2+, Cu+) and H2O2 [12], as well as by the IR-induced radiolysis of water [13]. A wide variety of cellular antioxidant defense mechanisms including both redox-buffering enzymatic and nonenzymatic systems have evolved, for example, superoxide dismutases, catalases, glutathione peroxidases, peroxiredoxins, and glutaredoxins; these counteract the detrimental effect of oxidative stress to the biological molecules, and an imbalance in their homeostasis leads to increased damage to the biomolecules [14].

A plethora of oxidized base lesions are generated mostly from guanine (G) in DNA, which has the lowest redox potential among the natural bases. Other lesions including 2-deoxyribose modifications, SSBs, DSBs, and protein-DNA cross-links are also ROS reaction products in DNA [10, 14, 15, 16, 17]. Nearly 100 such lesions have been identified; however, because of the lack of sensitivity of the techniques used to identify the lesions and inherent instability of some of them, the total number formed in the genome under a pro-oxidant environment is likely to be much higher [18].

The most commonly formed oxidized base lesion is 7,8-dihydro-8-oxoguanine (8-oxoguanine, 8-oxoG), which was discovered by Kasai and Nishimura in 1983 and coined as 8-hydroxyguanine [19, 20, 21]. All the nucleobases are also ionized by IR and by high intensity 266-ns laser photolysis. The DNA bases undergo one-electron oxidation (one electron ionization potential of G<A<C~T). 8-oxoG is generated at a much higher level (>5-fold) than the combined level of other one-electron base oxidation products. Singlet oxygen (1O2), the major ROS in UVA-mediated oxidation of DNA, specifically targets G and 2-deoxyribose moiety [22, 23, 24]. Other major oxidized base lesions are 5-hydroxy-6-hydrothymine, thymine glycol (TG), cytosine glycol (CG), 5-hydroxycytosine (5-OHC), uracil glycol (UG), 5-hydroxyuracil (5-OHU), 8-hydroxyadenine, and 2-hydroxyadenine [14, 17]. Hypochlorous acid (HOCl), generated by myeloperoxidase in neutrophils during inflammation, chlorinates both DNA and RNA bases [25, 26], and the main products are 5-chlorocytosine, 8-chloroadenine, and 8-chloroguanine. A summary of commonly formed oxidized bases detected in cellular DNA is shown in Table 1 [16]. Apart from ROS-induced generation of oxidized bases, 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC), and 5-carboxylcytosine (5caC) are formed enzymatically during transcriptional reprogramming involving oxidative demethylation of 5-methyl C (5mC), mostly localized in promoter CpG islands, induced by TET dioxygenases [27, 28]. However, enzymatically generated 5-methyl C oxidation products are produced >2-fold higher than that from direct oxidative damage to DNA [29, 30]. Additionally, tandem base lesions are produced by radicals generated from OH or one-electron oxidation reactions. Examples include the addition of either 5-(uracilyl)methyl radicals or 6-hydroxy-5,6-dihydrocytosin-5-yl radicals to 5′-adjacent guanine moieties in the DNA of cells exposed to H2O2 [31, 32] and formation of a guanine-thymine cross-link upon initial formation of guanine radical cation [33, 34]. One-electron oxidation also leads to DNA-protein cross-links. UVA irradiation of 6-thioguanine-containing DNA forms DNA-protein cross-links in human cells [33, 35].

Table 1.

Common oxidized bases detected in DNA.

Advertisement

3. Fate of oxidized bases and accumulation of mutations

ROS-induced oxidized base lesions and AP sites if left unrepaired are replicated by replicative or DNA translesion synthesis (TLS) polymerases [36]. Their misreplication generates mutations, a hallmark of cancer genomes, which account for two-thirds of single base pair substitutions [37, 38, 39, 40]. Furthermore, single nucleotide polymorphisms (SNPs), observed in normal human genomes, also likely result from such spontaneous single base pair substitutions. U and 5-OHU, the spontaneous and ROS-induced oxidative deamination product of C, respectively, preferably pair with A during replication, resulting in GC → AT transition mutation; 8-oxoG, the predominant oxidized base lesion mispairs with A, leading to GC → TA transversion mutation [41, 42]. In response to continuous assault by both endogenous and environmental factors, cellular defense mechanisms including diverse DNA repair pathways have evolved in all organisms to correct these base modifications and maintain genomic integrity.

Advertisement

4. Base excision repair of oxidized bases

Base excision repair (BER) is responsible for repairing most oxidized base lesions, AP sites, and DNA SSBs. The basic mechanism of BER first elucidated in Escherichia coli is broadly conserved across all organisms, as highlighted in several reviews [43, 44, 45, 46]. BER requiring only four or five enzymes in the basic reaction steps is initiated with excision of the damaged base by a monofunctional DNA glycosylase (DG), for example, uracil-DNA glycosylase (UDG) or 3-methyladenine-DNA glycosylase, generating an abasic apurinic/apyrimidinic (AP) site due to hydrolysis of the N-glycosidic bond of the damaged base. The AP endonuclease (APE1 in mammalian cells) cleaves the resulting AP site in the second step and generates 3′ OH and 5′ deoxyribose phosphate (dRP) termini. The DNA polymerase in the third step fills in the single nucleotide gap. In mammalian cells, DNA polymerase β (Pol β) also has intrinsic dRP lyase activity, which cleaves the dRP residue and generates 5′ phosphate; the resulting nick after incorporation of the correct base is sealed by DNA ligase III (Lig III) complexed with XRCC1 in the final step.

The BER initiating DGs for oxidized bases, on the other hand, are bifunctional with intrinsic AP lyase activity. The bifunctional oxidized base-specific DGs further process the AP site via β or βδ lyase reaction. The Nth family of DGs, OGG1, and NTH1, via β eliminations generates 3′ phospho α,β-unsaturated aldehyde (3′ PUA; formally named 3′ phospho 4-hydroxylpentenal) and 5′ phosphate at the strand break. NTH1 prefers oxidized pyrimidines as substrates, and 8-oxoG and ring opened guanine, that is, formamidopyrimidine (Fapy-G), are preferred substrates for OGG1. The Fpg/Nei family DGs NEIL1, NEIL2, NEIL3, discovered by us and others [47, 48, 49, 50, 51] catalyze βδ elimination and remove the deoxyribose residue to produce a 3′ phosphate and 5′ phosphate at the strand break. NEILs prefer modified pyrimidine substrates, NEIL1 having preference for ring-opened purines, for example, Fapy-A and Fapy-G. The activity and substrate specificity of NEILs depend on the DNA structure, and NEILs have significant 5-OHU excision activity with single-stranded or bubble, forked DNA. In contrast, OGG1 and NTH1 prefer double-stranded DNA substrates. Usually, the base excision and lyase reactions act in a concerted sequence. However, due to weak lyase activity of OGG1, intact AP sites are the major product after OGG1-catalyzed cleavage of 8-oxoG [52, 53]. All these bifunctional DGs have broad and overlapping substrate range and possess backup activity for many base lesions. This accounts for the fact that only few DGs have been discovered so far for much larger number of oxidized bases and for the nonessentiality of individual DGs.

The 3′ phosphate generated by the NEILs by βδ elimination is a poor substrate for mammalian APE1 and is processed by polynucleotide kinase phosphatase (PNKP) [54, 55, 56, 57]. Thus, for oxidized bases, the DGs actually define the subsequent steps. APE1 is responsible for processing the β elimination product of OGG1 and NTH1, whereas PNKP is required for generating 3′-OH termini from 3′ phosphate, a βδ elimination product of NEILs. Furthermore, AP sites and 3′ PUA generated by other DNA glycosylases can also be processed through a NEIL-PNKP-dependent pathway [53, 57]. This alternative repair route provides the functional redundancy in mammalian BER for genome safeguarding against a plethora of endogenous and induced oxidative damages.

BER, in the simplistic model, generates a 1-nucleotide gap after excision of the damaged base and has been termed single nucleotide BER (SN-BER) or short-patch BER (SP-BER). In contrast, long-patch BER (LP-BER) involves repair synthesis of two to eight deoxynucleotides. The 5′ blocking group after oxidation of AP sites cannot be removed by Pol β via its dRP lyase activity. Instead it is removed by 5′-flap endonuclease 1 (FEN-1), which is normally required for removing the 5′ RNA primers from Okazaki fragments during DNA replication. Thus, the subsequent steps of LP-BER are identical to that of DNA replication, utilizing DNA replication machinery, involving DNA polymerases δ/ε (Pol δ/ε) and DNA ligase I (Lig I). These enzymes including FEN-1 are recruited by the sliding clamp PCNA, loaded by replication factor-C (RFC), as in replication [58]. Thus, the choice of LP-BER vs. SN-BER depends on the 5′-terminus at the base cleavage site. With unaltered aldehyde group in deoxyribose, Pol β could carry out SN-BER by excising the 5′-dRP. LP-BER becomes necessary for repairing the oxidized AP sites, which cannot be processed by the 5′ end cleaning lyase activity of Pol β. The nuclear replicative Pol δ/ε lack dRP lyase activity and thus repair synthesis by these enzymes have to follow the LP-BER subpathway. Because Pol β-depleted cells are resistant to oxidative stress, Pol δ/ε can substitute for DNA Pol β and carry out the preferred LP-BER. The BER subpathways are schematically shown in Figure 1, adapted from [44].

Figure 1.

A schematic representation of oxidized base-specific BER subpathways. The damaged base is represented as . BER is initiated by the DGs: OGG1, NTH1, NEILs, and converge to common steps for end cleaning, followed by repair synthesis and ligation. See text for details.

Advertisement

5. Prereplicative BER of oxidized bases

The genomic integrity is particularly vulnerable during replication. Transient single-stranded (ss) DNA serving as a template during DNA replication after unwinding of the duplex genome is particularly vulnerable to ROS, which induces oxidized bases, sugar fragments, as well as strand breaks. Most oxidized bases do not stall replicative DNA polymerases, but they mispair during replication, thereby causing mutations. In contrast, bulky lesions, which stall replicative polymerases, block replisomes so as to allow repair. However, blocked replication may also lead to fork collapse, causing significant alteration in genomic stability. Furthermore, oxidized deoxynucleotides may be incorporated into the progeny strand during replication. If left unrepaired, these mutations could accumulate in progeny cells, a recipe for pathologies linked to genomic instability, including cancer, accelerated aging, and degenerative brain diseases [59, 60]. Repair of oxidative lesions, which are generated at much higher abundance than the bulky adducts in the replicating genome, is thus critical to maintain genomic fidelity. Mammalian cells have developed multiple ways to faithfully repair such base damages via prereplicative repair in the template strand and postreplicative repair in the progeny strand, immediately after replicative synthesis. Both the pathways involve an intricate collaboration of specific repair machinery with the replication proteins, likely via formation of dynamic “preformed” “repair-replication complexes” at the replication fork [61, 62].

Repair of most mutagenic base lesions except 8-oxoG, for example, 5-OHU, TG, 5-OHC, Fapy-A, 8-oxoA, and UG must be carried out prior to replication in order to prevent mutation fixation. How such lesions, which do not block replicative Pol δ, are flagged for prereplicative repair without causing DSBs was unclear. Our recent study showed that the mammalian DG NEIL1 binds to the oxidized lesion sites in ss DNA substrates in vitro to facilitate fork regression and participates in prereplicative repair of the damaged base in the reannealed duplex DNA [61, 62]. We compared the function of NEIL1 in stalling the replication fork at the damage sites for the prereplicative repair to the function of a “cow catcher” attached to the front of early steam locomotives that served to push aside animals or debris from the track ahead of the train’s traversal, in a simplistic analogy to this exquisitely orchestrated process [63]. The key features of this “cow catcher” model are the ability of NEIL1 to recognize base lesions in ss DNA templates and its nonproductive binding to lesions in ss DNA, which, while preventing lethal DSB formation, causes the stalling of the replication fork. Subsequent fork reversal allows base lesion repair in the reannealed duplex. High expression and activation of NEIL1 in replicating cells, together with its stable physical and functional association with proteins in the DNA replication complex [48, 64, 65, 66], are consistent with this surveillance role of NEIL1. The human genome during each cell division may be at higher risk for oxidative damage whose repair would prevent accumulation of mutations in the daughter cells. Thus NEIL1’s prereplicative BER function appears to be critical for preventing mutations and maintaining genome fidelity during cell division.

Advertisement

6. Posttranslational modifications of BER proteins

In vitro BER studies, carried out during the last couple of decades, are straightforward, mainly documenting functions of the repair proteins; however, in the complex cellular environment, the pathways are tightly regulated by interactions among the partner proteins in multiprotein complexes, which in turn also dictates the stability of the complexes. The stability and subcellular localization of these proteins are regulated by site-specific posttranslational modifications (PTMs), primarily involving acetylation, methylation, phosphorylation, SUMOylation, ubiquitination, and PARylation. Thus PTMs are at the root of major regulatory processes, by bestowing novel biochemical properties to the modified proteins, including changes in enzymatic activity, subcellular localization, interaction partners, protein stability, and DNA binding. Although purified recombinant BER proteins without any PTMs are proficient in their enzymatic activities, in cellulo BER is significantly affected by these PTMs. In this section, we discuss all the major PTMs of BER proteins identified so far.

The hallmark of mammalian DGs and early BER proteins is the presence of nonconserved, intrinsically disordered appendages at the N or C terminus, which are absent in their bacterial orthologs. Some examples are the N-terminal extension in human NTH1 absent in the E. coli Nth, C-terminal extension in human NEIL1 which is lacking in E. coli Nei, N-terminal extension in human APE1 lacking in E. coli Xth [44, 65, 67, 68]. Although the unfolded sequence generally exists at the N or C terminus, this could also exist internally as in Human NEIL2, where it may serve as a linker of the two domains. Analogous to the situation of histones H3 and H4, where mostly all PTMs occur in the disordered N-terminal tail [69, 70], PTMs in many early BER proteins are clustered in their disordered domains. See Table 2 for the major BER PTMs known so far.

Functional class BER protein PTM and identified site BER activity Protein stability Reference
DNA glycosylases Uracil DNA glycosylase (UNG) Phosphorylation; T6, S23, T60, S64, T126 + [127, 128, 213]
SUMOylation + [214]
Ubiquitination [127, 215, 216]
Single-strand-selective monofunctional uracil DNA glycosylase 1 (SMUG1) Ubiquitination [215, 216]
Methyl CpG-binding domain protein 4, DNA glycosylase (MBD4) Phosphorylation; S156,S262 + [217]
Thymine DNA glycosylase (TDG) Acetylation; K94, K95, K98 [129, 130]
Phosphorylation; S93, S96, S99 + [129, 130]
SUMOylation;
K330
K341

+
[131, 218, 219, 220, 221]
Ubiquitination [222, 223]
MutY DNA glycosylase homolog (MYH) Phosphorylation; S524 + [224, 225]
Ubiquitination;
C-terminal K between aa 475–535
[226]
8-Oxo guanine DNA glycosylase 1 (OGG1) Acetylation; K338, K341 + [203]
Phosphorylation; S326 + [227, 228]
Ubiquitination [229]
Nei-like DNA glycosylase 1 (NEIL1) Acetylation; K296, K297, K298 + + [161]
Phosphorylation; S61, S207, Y263, S269, S306 [230, 231, 232]
Nei-like DNA glycosylase 2 NEIL2 Acetylation; K49, K153 [233]
N-methylpurine DNA glycosylase (MPG) Acetylation + [234]
Phosphorylation; S172 + [235]
End processors Apurinic/apyrimidinic endonuclease 1 (APE1) Acetylation; K6, K7, K27, K31, K32, K35 + + [135, 136, 236, 237, 238]
Phosphorylation; T233 [239, 240]
Ubiquitination; K6, K7, K24, K25, K27, K31, K32, K35 [137, 138, 241]
Polynucleotide kinase phosphatase (PNKP) Phosphorylation; S114, S126 + + [139, 140, 141]
Ubiquitination; K414, K417, K484 [139]
Flap endonuclease-1 (FEN-1) Acetylation; K354, K355, K377, K380 [242]
Phosphorylation; S187 [144, 145]
Methylation; R192 + + [146]
SUMOylation; K168 [143]
DNA polymerases DNA polymerase β (Pol β) Acetylation; K72 [147]
Methylation; R137, R83, R152 + [148, 149]
Ubiquitination; K41, K61, K81 [150, 151]
DNA polymerases δ (Pol δ) Phosphorylation; S458 of p68 subunit [243]
DNA ligases DNA ligase IIIα (Lig IIIα) Phosphorylation; S123 [244]
Ubiquitination [150, 245]
Accessory proteins X-Ray repair cross-complementing 1 (XRCC1) Phosphorylation; S518, T519, T523, C-terminal linker, T284, S371 + + [246, 247, 248, 249, 250, 251, 252]
SUMOylation [152, 253]
Ubiquitination; BRCA1 C terminus (BRCT II) motif on the C-terminal end [150, 245, 252]
Poly(ADP-ribose) polymerase 1 (PARP-1) Acetylation; K498, K505, K508, K521, K524 [254]
Phosphorylation; S372, T373 + [255]
SUMOylation; K203, K482, and K486 [132, 133]
Ubiquitination [132, 245, 256]

Table 2.

PTMs of BER proteins.

6.1. Acetylation

Acetylation of histones was discovered back in 1963 after the Nobel prize-winning discovery of acetyl CoA [71, 72, 73, 74], and acetylation of histones at the ε-amino group of Lys residues in their disordered N-terminal region was shown to suppress their abilities to inhibit transcription [75]. Following these pioneering discoveries that linked histone acetylation to chromatin decondensation and transcriptional activation [76, 77, 78], diverse acetylation modifiers were identified and characterized. These include various histone acetyltransferases (HATs) such as E1a-binding protein p300 (p300), CREB-binding protein (CBP), ortholog of yeast transcription regulator Gcn5, TAF(II)250 subunit of transcription factor IID, several members of the MYST family (MOZ, YBF2/SAS3, SAS2, and TIP60) and p300/CBP associated factor (PCAF). Histone deacetylases (HDACs) were subsequently discovered as “erasers,” which include distinct members, HDACs1–11 and SIRTs in different transcriptional repressor complexes SIN3, NURD, etc., which regulate acetylation/deacetylation cycle in cells [79, 80, 81]. These discoveries set the stage for epigenetic regulation of gene expression. Simultaneously, the concept of “reader” proteins [80, 82] that specifically recognize acetylated Lys residues through their bromodomains was introduced in addition to the “writers” (HATs) and “erasers” (HDACs). Although the first discovered nonhistone protein acetylation dated back in 1997 for the tumor suppressor TP53 [83], the overwhelming numbers of nonhistone protein acetylation, particularly in large macromolecular complexes involved in chromatin remodeling, DNA repair, cell cycle, etc., were appreciated much later, after 2006, from mass spectrometric-based proteomic approaches, and provided the global scenario of “cellular acetylome” [81, 84, 85, 86].

6.2. Phosphorylation

Although enzymatic phosphorylation of proteins was discovered in 1954 [87], phosphorylated protein was known much earlier, based on identification of phosphate in vitellin [88], followed by detection of phosphoserine in this protein [89]. During the 1950s, ATP was discovered to be required for phosphorylation when the phosphate group was found to be covalently attached to specific serine/threonine residues [90, 91]. Subsequently, various kinases that phosphorylate serine/threonine and later tyrosine residues were characterized for their ability to modulate protein functions [91, 92, 93]. As with acetylation, phosphorylation induces conformational changes in the protein that stimulates its enzymatic activity and modulates protein-protein interactions [92, 94, 95]. Although the initial studies in protein phosphorylation were focused on cellular communications and signal transduction pathways, eventually the critical role of protein kinases and the relevance of phosphorylation/dephosphorylation events in DNA damage response (DDR) are extensively acknowledged, and mass spectrometry-based global screening approaches enabled identification of diverse phosphorylation targets [96, 97].

6.3. Ubiquitination and SUMOylation

Proteins are also posttranslationally modified via isopeptide bond formation with small proteins, which leads to nonlinear polypeptides [98, 99]. Ubiquitin is the first-discovered and well-characterized member of this growing family of small peptide modifiers, which covalently modify diverse proteins involved in chromatin organization, gene expression, signal transduction, DDR, DNA repair, and protein degradation [100, 101, 102]. Ubiquitin signals are generated by an enzymatic cascade involving E1 ubiquitin-activating enzymes, E2 ubiquitin-conjugating enzymes, and E3 ubiquitin ligases. Ubiquitination is a highly dynamic process with deubiquitinases (DUBs) involved in this signaling, and growing evidence indicates the involvement of ubiquitination/deubiquitination in BER, as shown in Table 2.

Small ubiquitin-related modifier (SUMO), containing 100 amino acid (aa) residues protein, is ubiquitin-like polypeptide, which is conjugated to substrates in a manner similar to ubiquitination [102, 103]. The SUMO paralogs are synthesized as precursor proteins that are cleaved by a family of SUMO isopeptidases [104]. Mature SUMO is subsequently activated by a heterodimeric E1-activating enzyme Aos1/Uba2 (SAE1/SAE2) forming a thioester bond between its catalytic cysteine and the C-terminal carboxyl group of mature SUMO. Then SUMO is transferred to the catalytic cysteine of the E2-conjugating enzyme Ubc9. In contrast to the ubiquitin system where dozens of E2 enzymes have been identified, Ubc9 is the only known SUMO E2 conjugating enzyme. Finally, an isopeptide bond is formed between SUMO and the substrate by E3 ligases. A consensus SUMO acceptor site has been identified consisting of the sequence ΨKXE, where Ψ is a large hydrophobic amino acid and K is the site of SUMO conjugation [105]. There are at least four SUMO paralogs in humans, SUMO1, SUMO2, SUMO3, and SUMO4, which have more than 1000 protein targets. SUMOylation is highly dynamic and can be reversed by the action of deSUMOylating enzymes (SENPs). SUMOylation regulates protein-protein interactions involving SUMO-interacting motifs (SIMs), and it targets a group of proteins in the same pathway to facilitate association of multiprotein complexes for transcription, nuclear transport, chromatin assembly and modification, chromosome segregation, DNA damage repair, replication, and cell signaling [106, 107].

6.4. PARylation

Poly ADP-ribosylation (PARylation), a crucial PTM that appears rapidly at DNA damage sites, is catalyzed by poly(ADP-ribose) polymerases (PARPs). The human PARP family contains 17 members among which only PARP1, 2, and 3 are involved in DDR [108, 109, 110, 111]. PARPs covalently attach the ADP-ribose unit via an ester bond to the carboxyl group of glutamate or aspartate and sometimes also attach to cysteine or lysine of the target proteins [112, 113, 114]. PARPs successively transfer ADP-ribose units from NAD+ to produce PAR chains containing up to 200 ADP-ribose units; however, in many cases, only single mono ADP-ribose moiety is transferred to the target proteins. Strand breaks in DNA activate PARP1, the founding and predominant member of the PARP family; the primary substrate of PARP1 is itself. Many proteins in the DDR pathways as well as the damage processing enzymes interact with PARP1 and/or are PARylated [112, 115]. In cells, PARylation/dePARylation is tightly and dynamically regulated; the PAR polymers are degraded by PAR glycohydrolase (PARG), possessing both exoglycosidic and endoglycosidic activities, and release free ADP-ribose moieties [116, 117, 118]. ADP-ribosyl-acceptor hydrolase (ARH) also exhibits PAR-degrading activity, although it has only exoglycosidase activity [119, 120]. Retention of PAR chains in cells triggers apoptotic cell death [121]. Although PARP1 interacts with the SSBR sensor XRCC1, as well as with other BER/SSBR proteins, and enables early recruitment of XRCC1 to the DNA lesions [122, 123, 124, 125, 126], there is no convincing evidence for PARylation of BER/SSBR proteins.

6.5. Cross-talks between different PTMs, their regulation, and effect on BER

Proteins employ diverse PTMs sequentially or concurrently to expand their repertoire of functions, thereby impacting global cellular signaling. The best example is the disordered N-terminal tail of histone H3, which has multiple sites for acetylation, methylation, phosphorylation, ubiquitination, and SUMOylation [69, 70]. These PTMS could act synergistically or via reciprocal exclusion to modulate chromatin organization, thus affecting the transcriptome. The same Lys residues (K9, K27) in H3 are targets for both acetylation (marker of active chromatin) and methylation; however, monomethylation of these residues are markers of active chromatin, while di- and trimethylation are associated with repression. Recent evidence on BER enzymes, summarized below, suggests that specific modification at one site can dramatically influence another modification at a different site, which may critically impact BER activity.

Cyclin-dependent kinase (CDK)-mediated phosphorylation of UNG2 (nuclear UDG) in S phase signals its ubiquitination-dependent degradation, and CDK-inhibitor roscovitine prevents such degradation [127, 128]. This suggests that phosphorylation-induced conformational change in UNG2 is a prerequisite for ubiquitination.

In the case of TDG, acetylation inhibits its repair activity by two distinct mechanisms. TDG acetylation at K94, K95, and K98 by p300/CBP suppresses BER by preventing APE1 recruitment to the damage site [129]. Protein kinase C (PKC)-mediated phosphorylation at S93, S96, and S99, close to the acetylation sites, may promote repair by sterically blocking repair-inhibitory acetylation of adjacent lysine residues [130]. On the other hand, SUMOylation at K341 inhibits TDG’s interaction with CBP, preventing its acetylation and thereby promoting BER [131].

PARP1, SUMOylated at K203 and K486, is a target for ubiquitination and degradation, which is believed to be the mechanism for its turnover [132]. In contrast, PARP1’s SUMOylation at K482 does not degrade the protein, rather stimulates PARylation of chromatin-associated proteins [133]. On the other hand, acetylation of PARP1, which stimulates its transactivation function, is inhibited by K486 SUMOylation. Thus, K486 SUMOylation restrains PARP1’s transactivation function [134].

While acetylation of APE1 enhances its stability in chromatin and enzymatic activity [135, 136], CDK5-mediated phosphorylation enhances its ubiquitination and degradation [137, 138]. Thus, it is possible that phosphorylation and acetylation are mutually exclusive, acetylation stabilizing the protein, and phosphorylation guiding to its degradation.

In the case of PNKP, ATM-dependent phosphorylation was shown to prevent ubiquitination and hence its degradation. Thus, in response to oxidative stress, ATM phosphorylates and stabilizes PNKP in order to activate a coordinated DDR pathway [139, 140, 141]. Furthermore, PNKP interacts with the deubiquitination enzyme ataxin-3 (ATXN3), which enhances its stability and phosphatase activity [142].

Phosphorylation of FEN-1 by CDK1 at S187 was shown to promote SUMOylation at K168, which enhanced its polyubiquitination-dependent degradation [143]. Phosphorylation inhibits FEN-1’s flap endonuclease activity [144, 145], which cross-talks with methylation, a lesser studied PTM of BER proteins. Methylation by arginine methyltransferase 5 at R192 prevents this phosphorylation and thus is proposed to be essential for the repair activities of FEN-1 [146]. Thus, in response to oxidative stress in cycling cells, methylation of FEN-1 could be a critical requirement for LP-BER.

Acetylation of Pol β at K72 inhibits its dRP lyase activity [147], and this could account for acetylation-induced inhibition of enzymatic activity and switch from SN-BER to LP-BER. Methylation of Pol β at R137 has no effect on dRP lyase or DNA polymerase activities but inhibits its interaction with PCNA [148] and could thus be predicted to inhibit LP-BER. In contrast, R83 and R152 methylation enhanced Pol β’s DNA binding and increased processivity [149]. Cellular Pol β level appears to be maintained by two ubiquitin E3 ligases, Mule and CHIP. DNA Pol β is monoubiquitinated by Mule, which in turn is recognized and polyubiquitinated by CHIP in undamaged cells. In response to oxidative stress, it is deubiquitinated, thus ensuring its stability and oxidized base damage repair [150, 151].

A recent study shows how PARylation stimulates SUMOylation [152]. In response to DNA strand breaks induced by alkylating agent methylmethanesulfonate (MMS), PARP1 is activated and synthesizes PAR chains; this promotes recruitment of SUMO E3 TOPORS to XRCC1, which facilitates XRCC1 SUMOylation. XRCC1 SUMOylation recruits Pol β at the damaged sites and thus ensures completion of BER.

Advertisement

7. Does chromatin organization affect BER? Understanding BER at the chromatin context

BER, as studied in vitro with naked DNA substrates, involves sequential enzymatic steps in which each enzyme utilizes the product of the previous step as the substrate. This observation inspired the prevailing dogma that the sequential steps in BER involves the hand-off process where the product of one step is handed over to the enzyme in the next step [153, 154]. Later steps generate intermediate product lesions that are more toxic than the original lesions. The BER intermediates such as AP sites and SSBs, which are highly mutagenic, interfere with replication and transcription, and hence the entire BER steps must be coordinated once the repair is initiated [155, 156, 157, 158]. Cumulating evidence suggests that the BER proteins act in concert beyond simply recognizing and acting upon the product of the previous step, by being present at the site of the original lesion [43, 52, 61, 62, 64, 65, 125, 159, 160]. This is the basis for the emerging paradigm of “preformed BER complexes,” named, “BERosomes” in mammalian cells. Being an integral part of complexes, it may be easier for the BER intermediates to be handed over to the next enzyme, which likely undergoes allosteric changes after binding to its substrate. Recent studies in our and collaborators’ labs suggest that these “BERosomes” are constitutively chromatin-bound to ensure prompt repair in the event of any threat [62, 135, 161]. Simultaneously, recent interests in the BER field have evolved toward deciphering the role of different chromatin factors and the underlying chromatin remodeling in oxidized base repair.

Several in vitro studies showed reduced BER activity with reconstituted core nucleosome particles, where every step during repair of diverse lesions was found to be inhibited by histones [162, 163, 164, 165, 166, 167, 168, 169, 170]. Overall BER efficiency is strongly inhibited by the presence of nucleosomes, which interfere with the interaction between the repair proteins and their substrate lesions, thereby compromising physical interaction and catalysis. Because oxidized bases perturb the DNA structure only mildly [170], whether chromatin remodeling occurs during BER was questionable. But, as BER efficiently occurs in cells, the results from these in vitro experiments imply that chromatin rearrangement occurs at oxidized DNA damage sites in cells, as was shown in the case of repair of DSBs, UV ray-mediated damages, and mismatched base pairs [171, 172, 173].

An inverse correlation exists in cells between BER and chromatin compaction. ROS induces assembly of BER complexes preferentially on open chromatin regions [174], as we have also observed that the BER complexes are constitutively present on actively transcribing sequences [175]. Interestingly, BER is involved during active CpG demethylation in promoters, mediated by TET dioxygenase(s) during transcriptional activation [176, 177, 178, 179, 180]. The TET proteins oxidize 5mC to 5hmC, 5fC, and 5caC; 5fC and 5caC are the TDG substrates. Thus, this coordination between CpG DNA demethylation, an epigenetic process essential for chromatin decondensation during transcriptional activation, and base damage repair supports our notion that “open-chromatin prefers BER activity across the genomic landscape” and highlights a regulatory link between epigenetics, chromatin remodeling, and BER.

Various ATP-dependent chromatin remodeling (ACR) complexes, which play significant roles in protein/DNA and protein/protein interactions in chromatin and regulate transcription, DNA repair processes such as DSB repair (DSBR), nucleotide excision repair (NER), and cross-link repair, also affect BER. ACR complexes utilize the energy of ATP hydrolysis to restructure nucleosomes on chromatin [181, 182, 183], thereby affecting gene expression profile and DNA repair. Four structurally related, but functionally distinct, ACR complex families were identified: SWI/SNF (switching defective/sucrose nonfermenting; most extensively studied), ISWI (imitation switch), CHD (chromodomain, helicase, DNA binding), and INO80 (inositol requiring 80). In vitro BER studies with reconstituted nucleosomes showed enhanced repair activity in the presence of purified SWI/SNF or ISW1/ISW2 complexes [184, 185, 186]. There are some indirect evidences of ACR during BER in yeast and mammalian cells. Depletion of STH1 (ATPase subunit of RSC, a member of SWI/SNF family) causes genome-wide BER inhibition and thus emphasizes a link between chromatin organization and BER [187]. In a recent study, depletion of ALC1/CHD1L, another member of SWI/SNF, compromises chromatin relaxation, associated with BER inhibition and increased sensitivity to MMS and H2O2 in chicken cells [188]. On the contrary, INO80 deficiency in MMS-sensitive yeast cells has no effect on genome-wide BER [189]. K56 acetylation in histone H3 is increased in chromatin of both yeast and mammalian cells following MMS treatment, which generates alkylated base substrates for BER, [190, 191]. H3K56Ac was also found to be enriched at DSBR sites and responsible for SWI/SNF complex recruitment during transcription [192]. Thus, it would be interesting to examine if any specific PTM(s) would target ACR after oxidized base damage and illuminate the phenomenon of ACR during BER. In any event, additional studies are required to test if ACR plays a role in enabling BER in condensed chromatin. It would be also of interest to explore if the BER proteins possess inherent chromatin remodeling activities, similar to the NER proteins, which have SWI/SNF domains [193, 194, 195]. Though no known BER proteins have SWI/SNF domains, the XRCC1-Lig IIIα complex could disrupt nucleosomes in vitro and enable BER completion [166].

Poly-ADP-ribosylation of histones by PARP1 after genome damage adds negative charge on histones and disrupts histone-DNA interactions, thereby promoting chromatin decondensation and enhancing interaction between the proteins involved in DNA transactions and DNA [111, 196, 197, 198]. This could increase DNA accessibility to the BER proteins. Although PARP1’s role in regulating transcription is well established, this would link chromatin remodeling to BER.

Nucleosomes pose obstruction to all DNA transactions and are likely disassembled to allow DNA replication, repair, and transcription, followed by their reassembly, which utilizes both parental histones and newly synthesized histones. Such replication-coupled nucleosome assembly in the S phase or replication-independent, transcription-coupled assembly throughout the cell cycle involves histone chaperones functioning at multiple steps of nucleosome formation [172, 199, 200]. Replication-coupled nucleosome assembly is aided by the chromatin assembly factor (CAF-1) and Rtt106 with the help of antisilencing function 1A (ASF1A) protein. Histone cell cycle regulator (HIRA) protein, along with Daxx, mediates replication-independent nucleosome assembly. While exploring chromatin-bound BER complexes, we serendipitously discovered CHAF1A (the largest subunit of CAF-1, along with other subunits CHAF1B and RBBP4), ASF1A, and various H3/H4 variants in the immunoprecipitation complex of NEIL1 or acetylated NEIL1 (201; unpublished). This underscores the importance of the diverse chromatin components in preformed “BERosomes,” which could regulate oxidize base repair in chromatin. We showed that ROS-induced oxidized base lesions caused transient dissociation of CHAF1A, ASF1A, and histones from the BER complexes and were restored back after repair completion. The repair activities of NEIL1 and OGG1, as well as complete cellular BER, were found to be inhibited by CAF-1, as well as the CHAF1A monomer [201]. So, we propose a hypothesis of temporal regulation of BER by the histone chaperones, whose dissociation from BER complexes is essential to initiate BER [201]. This has been illustrated in Figure 2.

Figure 2.

A schematic showing chromatin-bound BER complexes with histones and histone chaperones. ROS-induced damage causes transient dissociation of histones and histone chaperones to initiate BER, which are restored back after repair completion.

Recently, we discovered acetylation of NEIL1 at the disordered C-terminal K296-K298 by p300, which enhances its activity, and found that acetylated NEIL1 (AcNEIL1) could be detected only in the chromatin fraction and not in the soluble nuclear fraction [161]. Although the nonacetylable NEIL1 3KRmutant (Lys296–298 substituted with Arg) translocates to the nucleus and binds to chromatin, presumably due to retention of positive charges as in the WT enzyme, it forms less stable BER complexes with the histones, histone chaperones, and downstream BER proteins. Thus, as proposed earlier [65], the positive charge cluster in the disordered C-terminal region is required for NEIL1’s nonspecific DNA binding, after which acetylation occurs on the chromatin. Hydrophobic interaction of NEIL1 after acetylation-mediated charge neutralization probably stabilizes NEIL1’s complexes with nucleosome components and downstream BER proteins. Consequently, cells with acetylable NEIL1 exhibit enhanced BER efficiency and are less sensitive to oxidative stress. It is thus likely that unmodified NEIL1 binds to chromatin nonspecifically, and acetylation specifically at the promoter regions of actively transcribing genes by enhanced p300 activity actually stabilizes NEIL1’s (and possibly other DG’s) BERosomes on these preferred chromatin regions (Figure 3), which warrants further investigation.

Figure 3.

An illustrative view of “open” chromatin regions, containing bound “BERosomes” with histones, histone chaperones, PARPs, TETs, etc., for preferential repair of these transcriptionally active regions.

In a separate study, while investigating how APE1 repairs AP sites in cells, our collaborator’s lab found that acetylated APE1 (AcAPE1), like AcNEIL1, is exclusively and stably chromatin-bound throughout the cell cycle [135]. APE1 undergoes acetylation after binding to AP sites in chromatin, which enhances its enzymatic activity. In the absence of APE1 acetylation, cells accumulated AP sites and exhibited higher sensitivity to DNA damaging agents. We predict that other BER proteins OGG1 and MPG, whose repair activity is enhanced by acetylation, are similarly stabilized in chromatin-bound state.

Advertisement

8. Future perspectives

The genome-wide impact of various PTMs in the cross-talks among BER proteins, which dictates the overall repair efficiency, thus preserving genomic integrity against genotoxic insults from both endogenous and external oxidative stress, has not been investigated. In this NextGen era, holistic, whole-genome scanning approaches, although a daunting challenge, make it likely to map individual PTMs of BER proteins, the kinetics of their formation and removal, and their correlation with both intrinsic and ROS-induced BER efficiency across the genomic landscape. Because histone PTMs have been well established in chromatin remodeling, it is also important to explore how specific histone PTMs interfere with the BER PTMs.

The Access-Repair-Restore model [182, 202] provides an accepted view of DNA repair in chromatin, where chromatin remodeling is essential for the DNA repair machineries to get access to the damaged DNA. For BER, it is still not clear how chromatin remodeling and the associated histone PTMs initiate BER. The BER complexes constitutively bind to “open” chromatin regions, and chromatin remodeling could assist specific enzyme-substrate binding and enzyme catalysis needed to initiate and propagate BER. Moreover, although chromatin remodeling has been found to enable BER, the enhanced repair activity may be simply due to ROS-induced stimulation of BER genes’ expression or their specific PTM (acetylation), as has been shown by us [203, 204, 205], along with enhanced substrate binding in “open” chromatin. This may underestimate the contribution of ACR complexes at oxidized base lesion sites to enhance BER. Alternatively, in cells, chromatin remodeling-stimulated BER could be linked to replication and transcription, similar to transcription-coupled NER, which always occurs on “open” chromatin [206]. Indeed, repair of oxidized bases preferentially occurs in the transcribed strand [175], which could be assisted by Cockayne syndrome protein B (CSB), a NER factor, in transcription-coupled but NER-independent fashion [207]. Because BER/SSBR proteins such as PARP1 and APE1 are emerging as potential therapeutic targets [208, 209, 210, 211, 212], understanding if and how chromatin remodeling impacts BER activity is crucial to manipulating BER for effective modulation of repair activity in cancer cells. This would provide better efficacy and specificity in cancer therapy.

Advertisement

Acknowledgments

This work was supported by NIH R01 CA158910 (SM), R01 GM105090 (SM), P01 CA092584 (SM), R01 NS088645 (MLH), and HMRI laboratory start-up funds to SM and MLH.

Advertisement

Conflict of interest statement

The authors declare that there are no conflicts of interest associated with this study.

References

  1. 1. Talbert PB, Henikoff S. Histone variants on the move: Substrates for chromatin dynamics. Nature Reviews Molecular Cell Biology. 2016;18:115
  2. 2. Venkatesh S, Workman JL. Histone exchange, chromatin structure and the regulation of transcription. Nature Reviews Molecular Cell Biology. 2015;16:178
  3. 3. Lindahl T, Barnes DE. Repair of endogenous DNA damage. Cold Spring Harbor Symposia on Quantitative Biology. 2000;65:127-133
  4. 4. Tubbs A, Nussenzweig A. Endogenous DNA damage as a source of genomic instability in cancer. Cell. 2017;168:644-656
  5. 5. De Bont R, van Larebeke N. Endogenous DNA damage in humans: A review of quantitative data. Mutagenesis. 2004;19:169-185
  6. 6. Dizdaroglu M. Oxidatively induced DNA damage and its repair in cancer. Mutation Research. Reviews in Mutation Research. 2015;763:212-245
  7. 7. Dizdaroglu M, Coskun E, Jaruga P. Measurement of oxidatively induced DNA damage and its repair, by mass spectrometric techniques. Free Radical Research. 2015;49:525-548
  8. 8. Lan L, Nakajima S, Oohata Y, Takao M, Okano S, Masutani M, et al. In situ analysis of repair processes for oxidative DNA damage in mammalian cells. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:13738-13743
  9. 9. Cadet J, Davies KJA. Oxidative DNA damage & repair: An introduction. Free Radical Biology & Medicine. 2017;107:2-12
  10. 10. Cadet J, Douki T, Ravanat JL. Oxidatively generated base damage to cellular DNA. Free Radical Biology & Medicine. 2010;49:9-21
  11. 11. Davies KJ. Oxidative stress: The paradox of aerobic life. Biochemical Society Symposium. 1995;61:1-31
  12. 12. Winterbourn CC. Reconciling the chemistry and biology of reactive oxygen species. Nature Chemical Biology. 2008;4:278-286
  13. 13. Reisz JA, Bansal N, Qian J, Zhao W, Furdui CM. Effects of ionizing radiation on biological molecules—Mechanisms of damage and emerging methods of detection. Antioxidants & Redox Signaling. 2014;21:260-292
  14. 14. Storr SJ, Woolston CM, Zhang Y, Martin SG. Redox environment, free radical, and oxidative DNA damage. Antioxidants & Redox Signaling. 2013;18:2399-2408
  15. 15. Cadet J, Ravanat JL, TavernaPorro M, Menoni H, Angelov D. Oxidatively generated complex DNA damage: Tandem and clustered lesions. Cancer Letters. 2012;327:5-15
  16. 16. Cadet J, Wagner JR. DNA base damage by reactive oxygen species, oxidizing agents, and UV radiation. Cold Spring Harbor Perspectives in Biology. 2013;5:a012559
  17. 17. Evans MD, Dizdaroglu M, Cooke MS. Oxidative DNA damage and disease: Induction, repair and significance. Mutation Research. 2004;567:1-61
  18. 18. Cadet J, Douki T, Ravanat JL. Measurement of oxidatively generated base damage in cellular DNA. Mutation Research. 2011;711:3-12
  19. 19. Kasai H, Nishimura S. Hydroxylation of the C-8 position of deoxyguanosine by reducing agents in the presence of oxygen. Nucleic Acids Symposium Series. 1983;12:165-167
  20. 20. Nishimura S. 8-Hydroxyguanine: From its discovery in 1983 to the present status. Proceedings of the Japan Academy Series B, Physical and Biological Sciences. 2006;82:127-141
  21. 21. Nishimura S. 8-Hydroxyguanine: A base for discovery. DNA Repair (Amst). 2011;10:1078-1083
  22. 22. Ravanat JL, Di Mascio P, Martinez GR, Medeiros MH. Singlet oxygen induces oxidation of cellular DNA. The Journal of Biological Chemistry. 2001;276:40601-40604
  23. 23. Ravanat JL, Douki T, Cadet J. Direct and indirect effects of UV radiation on DNA and its components. Journal of Photochemistry and Photobiology. B. 2001;63:88-102
  24. 24. Ravanat JL, Sauvaigo S, Caillat S, Martinez GR, Medeiros MH, Di Mascio P, et al. Singlet oxygen-mediated damage to cellular DNA determined by the comet assay associated with DNA repair enzymes. Biological Chemistry. 2004;385:17-20
  25. 25. Malle E, Furtmuller PG, Sattler W, Obinger C. Myeloperoxidase: A target for new drug development? British Journal of Pharmacology. 2007;152:838-854
  26. 26. Masuda M, Suzuki T, Friesen MD, Ravanat JL, Cadet J, Pignatelli B, et al. Chlorination of guanosine and other nucleosides by hypochlorous acid and myeloperoxidase of activated human neutrophils. Catalysis by nicotine and trimethylamine. Journal of Biological Chemitry. 2001;276:40486-40496
  27. 27. An J, Rao A, Ko M. TET family dioxygenases and DNA demethylation in stem cells and cancers. Experimental & Molecular Medicine. 2017;49:e323
  28. 28. Wu X, Zhang Y. TET-mediated active DNA demethylation: Mechanism, function and beyond. Nature Reviews. 2017;18:517-534
  29. 29. Munzel M, Globisch D, Bruckl T, Wagner M, Welzmiller V, Michalakis S, et al. Quantification of the sixth DNA base hydroxymethylcytosine in the brain. Angewandte Chemie (International ed.). 2010;49:5375-5377
  30. 30. Pfaffeneder T, Hackner B, Truss M, Munzel M, Muller M, Deiml CA, et al. The discovery of 5-formylcytosine in embryonic stem cell DNA. Angewandte Chemie (International ed.). 2011;50:7008-7012
  31. 31. Hong IS, Carter KN, Sato K, Greenberg MM. Characterization and mechanism of formation of tandem lesions in DNA by a nucleobase peroxyl radical. Journal of the American Chemical Society. 2007;129:4089-4098
  32. 32. Jiang Y, Hong H, Cao H, Wang Y. In vivo formation and in vitro replication of a guanine-thymine intrastrand cross-link lesion. Biochemistry. 2007;46:12757-12763
  33. 33. Ding S, Kropachev K, Cai Y, Kolbanovskiy M, Durandina SA, Liu Z, et al. Structural, energetic and dynamic properties of guanine(C8)-thymine(N3) cross-links in DNA provide insights on susceptibility to nucleotide excision repair. Nucleic Acids Research. 2012;40:2506-2517
  34. 34. Yun BH, Geacintov NE, Shafirovich V. Generation of guanine-thymidine cross-links in DNA by peroxynitrite/carbon dioxide. Chemical Research in Toxicology. 2011;24:1144-1152
  35. 35. Gueranger Q, Kia A, Frith D, Karran P. Crosslinking of DNA repair and replication proteins to DNA in cells treated with 6-thioguanine and UVA. Nucleic Acids Research. 2011;39:5057-5066
  36. 36. Prakash S, Johnson RE, Prakash L. Eukaryotic translesion synthesis DNA polymerases: Specificity of structure and function. Annual Review of Biochemistry. 2005;74:317-353
  37. 37. Bacolla A, Cooper DN, Vasquez KM. Mechanisms of base substitution mutagenesis in cancer genomes. Genes. 2014;5:108-146
  38. 38. Helleday T, Eshtad S, Nik-Zainal S. Mechanisms underlying mutational signatures in human cancers. Nature Reviews. 2014;15:585-598
  39. 39. Lawrence MS, Stojanov P, Polak P, Kryukov GV, Cibulskis K, Sivachenko A, et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature. 2013;499:214-218
  40. 40. Tomasetti C, Li L, Vogelstein B. Stem cell divisions, somatic mutations, cancer etiology, and cancer prevention. Science. 2017;355:1330-1334
  41. 41. Hsu GW, Ober M, Carell T, Beese LS. Error-prone replication of oxidatively damaged DNA by a high-fidelity DNA polymerase. Nature. 2004;431:217-221
  42. 42. Shibutani S, Takeshita M, Grollman AP. Insertion of specific bases during DNA synthesis past the oxidation-damaged base 8-oxodG. Nature. 1991;349:431-434
  43. 43. Dutta A, Yang C, Sengupta S, Mitra S, Hegde ML. New paradigms in the repair of oxidative damage in human genome: Mechanisms ensuring repair of mutagenic base lesions during replication and involvement of accessory proteins. Cellular and Molecular Life Sciences. 2015;72:1679-1698
  44. 44. Hegde ML, Hazra TK, Mitra S. Early steps in the DNA base excision/single-strand interruption repair pathway in mammalian cells. Cell Research. 2008;18:27-47
  45. 45. Mitra S, Hazra TK, Roy R, Ikeda S, Biswas T, Lock J, et al. Complexities of DNA base excision repair in mammalian cells. Molecules and Cells. 1997;7:305-312
  46. 46. Mitra S, Izumi T, Boldogh I, Bhakat KK, Hill JW, Hazra TK. Choreography of oxidative damage repair in mammalian genomes. Free Radical Biology & Medicine. 2002;33:15-28
  47. 47. Bandaru V, Sunkara S, Wallace SS, Bond JP. A novel human DNA glycosylase that removes oxidative DNA damage and is homologous to Escherichia coli endonuclease VIII. DNA Repair (Amst). 2002;1:517-529
  48. 48. Hazra TK, Izumi T, Boldogh I, Imhoff B, Kow YW, Jaruga P, et al. Identification and characterization of a human DNA glycosylase for repair of modified bases in oxidatively damaged DNA. Proceedings of the National Academy of Sciences of the United States of America. 2002;99:3523-3528
  49. 49. Hazra TK, Kow YW, Hatahet Z, Imhoff B, Boldogh I, Mokkapati SK, et al. Identification and characterization of a novel human DNA glycosylase for repair of cytosine-derived lesions. The Journal of Biological Chemistry. 2002;277:30417-30420
  50. 50. Morland I, Rolseth V, Luna L, Rognes T, Bjoras M, Seeberg E. Human DNA glycosylases of the bacterial Fpg/MutM superfamily: An alternative pathway for the repair of 8-oxoguanine and other oxidation products in DNA. Nucleic Acids Research. 2002;30:4926-4936
  51. 51. Takao M, Kanno S, Kobayashi K, Zhang QM, Yonei S, van der Horst GT, et al. A back-up glycosylase in Nth1 knock-out mice is a functional Nei (endonuclease VIII) homologue. The Journal of Biological Chemistry. 2002;277:42205-42213
  52. 52. Hill JW, Hazra TK, Izumi T, Mitra S. Stimulation of human 8-oxoguanine-DNA glycosylase by AP-endonuclease: Potential coordination of the initial steps in base excision repair. Nucleic Acids Research. 2001;29:430-438
  53. 53. Mokkapati SK, Wiederhold L, Hazra TK, Mitra S. Stimulation of DNA glycosylase activity of OGG1 by NEIL1: Functional collaboration between two human DNA glycosylases. Biochemistry. 2004;43:11596-11604
  54. 54. Das A, Wiederhold L, Leppard JB, Kedar P, Prasad R, Wang H, et al. NEIL2-initiated, APE-independent repair of oxidized bases in DNA: Evidence for a repair complex in human cells. DNA Repair (Amst). 2006;5:1439-1448
  55. 55. Meijer M, Karimi-Busheri F, Huang TY, Weinfeld M, Young D. Pnk1, a DNA kinase/phosphatase required for normal response to DNA damage by gamma-radiation or camptothecin in Schizosaccharomyces pombe. The Journal of Biological Chemistry. 2002;277:4050-4055
  56. 56. Whitehouse CJ, Taylor RM, Thistlethwaite A, Zhang H, Karimi-Busheri F, Lasko DD, et al. XRCC1 stimulates human polynucleotide kinase activity at damaged DNA termini and accelerates DNA single-strand break repair. Cell. 2001;104:107-117
  57. 57. Wiederhold L, Leppard JB, Kedar P, Karimi-Busheri F, Rasouli-Nia A, Weinfeld M, et al. AP endonuclease-independent DNA base excision repair in human cells. Molecular Cell. 2004;15:209-220
  58. 58. Levin DS, McKenna AE, Motycka TA, Matsumoto Y, Tomkinson AE. Interaction between PCNA and DNA ligase I is critical for joining of Okazaki fragments and long-patch base-excision repair. Current Biology. 2000;10:919-922
  59. 59. Maynard S, Fang EF, Scheibye-Knudsen M, Croteau DL, Bohr VA. DNA damage, DNA repair, aging, and neurodegeneration. Cold Spring Harbor Perspectives in Medicine. 2015;5:a025130
  60. 60. Roos WP, Thomas AD, Kaina B. DNA damage and the balance between survival and death in cancer biology. Nature Reviews Cancer. 2016;16:20-33
  61. 61. Hegde ML, Hegde PM, Bellot LJ, Mandal SM, Hazra TK, Li GM, et al. Prereplicative repair of oxidized bases in the human genome is mediated by NEIL1 DNA glycosylase together with replication proteins. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:E3090-E3099
  62. 62. Hegde PM, Dutta A, Sengupta S, Mitra J, Adhikari S, Tomkinson AE, et al. The C-terminal domain (CTD) of human DNA glycosylase NEIL1 is required for forming BERosome repair complex with DNA replication proteins at the replicating genome: Dominant negative function of the CTD. The Journal of Biological Chemistry. 2015;290:20919-20933
  63. 63. Rangaswamy S, Pandey A, Mitra S, Hegde ML. Pre-replicative repair of oxidized bases maintains fidelity in mammalian genomes: The cowcatcher role of NEIL1 DNA glycosylase. Genes. 2017;8:E175
  64. 64. Hegde ML, Theriot CA, Das A, Hegde PM, Guo Z, Gary RK, et al. Physical and functional interaction between human oxidized base-specific DNA glycosylase NEIL1 and flap endonuclease 1. The Journal of Biological Chemistry. 2008;283:27028-27037
  65. 65. Hegde ML, Tsutakawa SE, Hegde PM, Holthauzen LM, Li J, Oezguen N, et al. The disordered C-terminal domain of human DNA glycosylase NEIL1 contributes to its stability via intramolecular interactions. Journal of Molecular Biology. 2013;425:2359-2371
  66. 66. Theriot CA, Hegde ML, Hazra TK, Mitra S. RPA physically interacts with the human DNA glycosylase NEIL1 to regulate excision of oxidative DNA base damage in primer-template structures. DNA Repair. 2010;9:643-652
  67. 67. Hegde ML, Hazra TK, Mitra S. Functions of disordered regions in mammalian early base excision repair proteins. Cellular and Molecular Life Sciences. 2010;67:3573-3587
  68. 68. Hegde ML, Izumi T, Mitra S. Oxidized base damage and single-strand break repair in mammalian genomes: Role of disordered regions and posttranslational modifications in early enzymes. Progress in Molecular Biology and Translational Science. 2012;110:123-153
  69. 69. Zhang C, Liu Y. Retrieving quantitative information of histone PTMs by mass spectrometry. Methods in Enzymology. 2017;586:165-191
  70. 70. Zhao Y, Garcia BA. Comprehensive catalog of currently documented histone modifications. Cold Spring Harbor Perspectives in Biology. 2015;7:a025064
  71. 71. Bloch K, Borek E. Biological acetylation of natural amino acids. The Journal of Biological Chemistry. 1946;164:483
  72. 72. Lipmann F. Development of the acetylation problem, a personal account. Science. 1954;120:855-865
  73. 73. Lipmann F, Kaplan NO, et al. Coenzyme for acetylation, a pantothenic acid derivative. The Journal of Biological Chemistry. 1947;167:869
  74. 74. Phillips DM. The presence of acetyl groups of histones. The Biochemical Journal. 1963;87:258-263
  75. 75. Allfrey VG, Faulkner R, Mirsky AE. Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proceedings of the National Academy of Sciences of the United States of America. 1964;51:786-794
  76. 76. Hebbes TR, Thorne AW, Crane-Robinson C. A direct link between core histone acetylation and transcriptionally active chromatin. The EMBO Journal. 1988;7:1395-1402
  77. 77. Turner BM, Birley AJ, Lavender J. Histone H4 isoforms acetylated at specific lysine residues define individual chromosomes and chromatin domains in Drosophila polytene nuclei. Cell. 1992;69:375-384
  78. 78. Turner BM, Fellows G. Specific antibodies reveal ordered and cell-cycle-related use of histone-H4 acetylation sites in mammalian cells. European Journal of Biochemistry/FEBS. 1989;179:131-139
  79. 79. Marmorstein R, Trievel RC. Histone modifying enzymes: Structures, mechanisms, and specificities. Biochimica et Biophysica Acta. 2009;1789:58-68
  80. 80. Marmorstein R, Zhou MM. Writers and readers of histone acetylation: Structure, mechanism, and inhibition. Cold Spring Harbor Perspectives in Biology. 2014;6:a018762
  81. 81. Verdin E, Ott M. 50 years of protein acetylation: From gene regulation to epigenetics, metabolism and beyond. Nature Reviews. 2015;16:258-264
  82. 82. Dhalluin C, Carlson JE, Zeng L, He C, Aggarwal AK, Zhou MM. Structure and ligand of a histone acetyltransferase bromodomain. Nature. 1999;399:491-496
  83. 83. Gu W, Roeder RG. Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell. 1997;90:595-606
  84. 84. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, et al. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science. 2009;325:834-840
  85. 85. Choudhary C, Weinert BT, Nishida Y, Verdin E, Mann M. The growing landscape of lysine acetylation links metabolism and cell signalling. Nature Reviews. 2014;15:536-550
  86. 86. Kim SC, Sprung R, Chen Y, Xu Y, Ball H, Pei J, et al. Substrate and functional diversity of lysine acetylation revealed by a proteomics survey. Molecular Cell. 2006;23:607-618
  87. 87. Burnett G, Kennedy EP. The enzymatic phosphorylation of proteins. The Journal of Biological Chemistry. 1954;211:969-980
  88. 88. Levene PA, Alsberg CL. The cleavage products of vitellin. The Journal of Biological Chemistry. 1906;2:127-133
  89. 89. Levene PA, Schormüller A. Serinephosphoric acid obtained on hydrolysis of vitellinic acid. The Journal of Biological Chemistry. 1933;103:537-542
  90. 90. Fischer EH, Krebs EG. Conversion of phosphorylase b to phosphorylase a in muscle extracts. The Journal of Biological Chemistry. 1955;216:121-132
  91. 91. Lecture KEGN. Protein phosphorylation and cellular regulation I. Bioscience Reports. 1993;13:127-142
  92. 92. Barford D, Johnson LN. The allosteric transition of glycogen phosphorylase. Nature. 1989;340:609-616
  93. 93. Walsh DA, Perkins JP, Krebs EG. An adenosine 3′,5′-monophosphate-dependant protein kinase from rabbit skeletal muscle. The Journal of Biological Chemistry. 1968;243:3763-3765
  94. 94. Pawson T. Specificity in signal transduction: From phosphotyrosine-SH2 domain interactions to complex cellular systems. Cell. 2004;116:191-203
  95. 95. Pawson T, Scott JD. Protein phosphorylation in signaling—50 years and counting. Trends in Biochemical Sciences. 2005;30:286-290
  96. 96. Bensimon A, Schmidt A, Ziv Y, Elkon R, Wang SY, Chen DJ, et al. ATM-dependent and -independent dynamics of the nuclear phosphoproteome after DNA damage. Science Signaling. 2010;3:rs3
  97. 97. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER 3rd, Hurov KE, Luo J, et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science. 2007;316:1160-1166
  98. 98. Psakhye I, Jentsch S. Protein group modification and synergy in the SUMO pathway as exemplified in DNA repair. Cell. 2012;151:807-820
  99. 99. Welchman RL, Gordon C, Mayer RJ. Ubiquitin and ubiquitin-like proteins as multifunctional signals. Nature Reviews. 2005;6:599-609
  100. 100. Glickman MH, Ciechanover A. The ubiquitin-proteasome proteolytic pathway: Destruction for the sake of construction. Physiological Reviews. 2002;82:373-428
  101. 101. Hershko A, Ciechanover A. The ubiquitin system for protein degradation. Annual Review of Biochemistry. 1992;61:761-807
  102. 102. Pickart CM. Ubiquitin enters the new millennium. Molecular Cell. 2001;8:499-504
  103. 103. Kim KI, Baek SH, Chung CH. Versatile protein tag, SUMO: Its enzymology and biological function. Journal of Cellular Physiology. 2002;191:257-268
  104. 104. Mukhopadhyay D, Dasso M. Modification in reverse: The SUMO proteases. Trends in Biochemical Sciences. 2007;32:286-295
  105. 105. Rodriguez MS, Dargemont C, Hay RT. SUMO-1 conjugation in vivo requires both a consensus modification motif and nuclear targeting. The Journal of Biological Chemistry. 2001;276:12654-12659
  106. 106. Geiss-Friedlander R, Melchior F. Concepts in sumoylation: A decade on. Nature Reviews. 2007;8:947-956
  107. 107. Kerscher O. SUMO junction-what's your function? New insights through SUMO-interacting motifs. EMBO Reports. 2007;8:550-555
  108. 108. Gibson BA, Kraus WL. New insights into the molecular and cellular functions of poly(ADP-ribose) and PARPs. Nature Reviews. 2012;13:411-424
  109. 109. Juarez-Salinas H, Levi V, Jacobson EL, Jacobson MK. Poly(ADP-ribose) has a branched structure in vivo. The Journal of Biological Chemistry. 1982;257:607-609
  110. 110. Hottiger MO. Nuclear ADP-ribosylation and its role in chromatin plasticity, cell differentiation, and epigenetics. Annual Review of Biochemistry. 2015;84:227-263
  111. 111. Rouleau M, Patel A, Hendzel MJ, Kaufmann SH, Poirier GG. PARP inhibition: PARP1 and beyond. Nature Reviews. Cancer. 2010;10:293-301
  112. 112. D'Amours D, Desnoyers S, D'Silva I, Poirier GG. Poly(ADP-ribosyl)ation reactions in the regulation of nuclear functions. The Biochemical Journal. 1999;342(Pt 2):249-268
  113. 113. Tallis M, Morra R, Barkauskaite E, Ahel I. Poly(ADP-ribosyl)ation in regulation of chromatin structure and the DNA damage response. Chromosoma. 2014;123:79-90
  114. 114. Vyas S, Chang P. New PARP targets for cancer therapy. Nature Reviews Cancer. 2014;14:502-509
  115. 115. Daniels CM, Ong SE, Leung AK. The promise of proteomics for the study of ADP-ribosylation. Molecular Cell. 2015;58:911-924
  116. 116. Dunstan MS, Barkauskaite E, Lafite P, Knezevic CE, Brassington A, Ahel M, et al. Structure and mechanism of a canonical poly(ADP-ribose) glycohydrolase. Nature Communications. 2012;3:878
  117. 117. Kim IK, Kiefer JR, Ho CM, Stegeman RA, Classen S, Tainer JA, et al. Structure of mammalian poly(ADP-ribose) glycohydrolase reveals a flexible tyrosine clasp as a substrate-binding element. Nature Structural & Molecular Biology. 2012;19:653-656
  118. 118. Ueda K, Oka J, Naruniya S, Miyakawa N, Hayaishi O. Poly ADP-ribose glycohydrolase from rat liver nuclei, a novel enzyme degrading the polymer. Biochemical and Biophysical Research Communications. 1972;46:516-523
  119. 119. Mueller-Dieckmann C, Kernstock S, Lisurek M, von Kries JP, Haag F, Weiss MS, et al. The structure of human ADP-ribosylhydrolase 3 (ARH3) provides insights into the reversibility of protein ADP-ribosylation. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:15026-15031
  120. 120. Niere M, Mashimo M, Agledal L, Dolle C, Kasamatsu A, Kato J, et al. ADP-ribosylhydrolase 3 (ARH3), not poly(ADP-ribose) glycohydrolase (PARG) isoforms, is responsible for degradation of mitochondrial matrix-associated poly(ADP-ribose). The Journal of Biological Chemistry. 2012;287:16088-16102
  121. 121. Wang Y, Kim NS, Haince JF, Kang HC, David KK, Andrabi SA, et al. Poly(ADP-ribose) (PAR) binding to apoptosis-inducing factor is critical for PAR polymerase-1-dependent cell death (parthanatos). Science Signaling. 2011;4:ra20
  122. 122. Caldecott KW. Single-strand break repair and genetic disease. Nature Reviews. 2008;9:619-631
  123. 123. Dantzer F, de La Rubia G, Menissier-De Murcia J, Hostomsky Z, de Murcia G, Schreiber V. Base excision repair is impaired in mammalian cells lacking poly(ADP-ribose) polymerase-1. Biochemistry. 2000;39:7559-7569
  124. 124. Harris JL, Jakob B, Taucher-Scholz G, Dianov GL, Becherel OJ, Lavin MF. Aprataxin, poly-ADP ribose polymerase 1 (PARP-1) and apurinic endonuclease 1 (APE1) function together to protect the genome against oxidative damage. Human Molecular Genetics. 2009;18:4102-4117
  125. 125. Hegde ML, Hegde PM, Arijit D, Boldogh I, Mitra S. Human DNA glycosylase NEIL1's interactions with downstream repair proteins is critical for efficient repair of oxidized DNA base damage and enhanced cell survival. Biomolecules. 2012;2:564-578
  126. 126. Liu L, Kong M, Gassman NR, Freudenthal BD, Prasad R, Zhen S, et al. PARP1 changes from three-dimensional DNA damage searching to one-dimensional diffusion after auto-PARylation or in the presence of APE1. Nucleic Acids Research. 2017;45:12834-12847
  127. 127. Fischer JA, Muller-Weeks S, Caradonna S. Proteolytic degradation of the nuclear isoform of uracil-DNA glycosylase occurs during the S phase of the cell cycle. DNA Repair (Amst). 2004;3:505-513
  128. 128. Hagen L, Kavli B, Sousa MM, Torseth K, Liabakk NB, Sundheim O, et al. Cell cycle-specific UNG2 phosphorylations regulate protein turnover, activity and association with RPA. The EMBO Journal. 2008;27:51-61
  129. 129. Tini M, Benecke A, Um SJ, Torchia J, Evans RM, Chambon P. Association of CBP/p300 acetylase and thymine DNA glycosylase links DNA repair and transcription. Molecular Cell. 2002;9:265-277
  130. 130. Mohan RD, Litchfield DW, Torchia J, Tini M. Opposing regulatory roles of phosphorylation and acetylation in DNA mispair processing by thymine DNA glycosylase. Nucleic Acids Research. 2010;38:1135-1148
  131. 131. Mohan RD, Rao A, Gagliardi J, Tini M. SUMO-1-dependent allosteric regulation of thymine DNA glycosylase alters subnuclear localization and CBP/p300 recruitment. Molecular and Cellular Biology. 2007;27:229-243
  132. 132. Martin N, Schwamborn K, Schreiber V, Werner A, Guillier C, Zhang XD, et al. PARP-1 transcriptional activity is regulated by sumoylation upon heat shock. The EMBO Journal. 2009;28:3534-3548
  133. 133. Ryu H, Al-Ani G, Deckert K, Kirkpatrick D, Gygi SP, Dasso M, et al. PIASy mediates SUMO-2/3 conjugation of poly(ADP-ribose) polymerase 1 (PARP1) on mitotic chromosomes. The Journal of Biological Chemistry. 2010;285:14415-14423
  134. 134. Messner S, Schuermann D, Altmeyer M, Kassner I, Schmidt D, Schar P, et al. Sumoylation of poly(ADP-ribose) polymerase 1 inhibits its acetylation and restrains transcriptional coactivator function. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2009;23:3978-3989
  135. 135. Roychoudhury S, Nath S, Song H, Hegde ML, Bellot LJ, Mantha AK, et al. Human apurinic/apyrimidinic endonuclease (APE1) is acetylated at DNA damage sites in chromatin, and acetylation modulates its DNA repair activity. Molecular and Cellular Biology. 2017;37:16
  136. 136. Sengupta S, Mantha AK, Song H, Roychoudhury S, Nath S, Ray S, et al. Elevated level of acetylation of APE1 in tumor cells modulates DNA damage repair. Oncotarget. 2016;7:75197-75209
  137. 137. Busso CS, Iwakuma T, Izumi T. Ubiquitination of mammalian AP endonuclease (APE1) regulated by the p53-MDM2 signaling pathway. Oncogene. 2009;28:1616-1625
  138. 138. Meisenberg C, Tait PS, Dianova II, Wright K, Edelmann MJ, Ternette N, et al. Ubiquitin ligase UBR3 regulates cellular levels of the essential DNA repair protein APE1 and is required for genome stability. Nucleic Acids Research. 2012;40:701-711
  139. 139. Parsons JL, Khoronenkova SV, Dianova II, Ternette N, Kessler BM, Datta PK, et al. Phosphorylation of PNKP by ATM prevents its proteasomal degradation and enhances resistance to oxidative stress. Nucleic Acids Research. 2012;40:11404-11415
  140. 140. Segal-Raz H, Mass G, Baranes-Bachar K, Lerenthal Y, Wang SY, Chung YM, et al. ATM-mediated phosphorylation of polynucleotide kinase/phosphatase is required for effective DNA double-strand break repair. EMBO Reports. 2011;12:713-719
  141. 141. Zolner AE, Abdou I, Ye R, Mani RS, Fanta M, Yu Y, et al. Phosphorylation of polynucleotide kinase/phosphatase by DNA-dependent protein kinase and ataxia-telangiectasia mutated regulates its association with sites of DNA damage. Nucleic Acids Research. 2011;39:9224-9237
  142. 142. Chatterjee A, Saha S, Chakraborty A, Silva-Fernandes A, Mandal SM, Neves-Carvalho A, et al. The role of the mammalian DNA end-processing enzyme polynucleotide kinase 3′-phosphatase in spinocerebellar ataxia type 3 pathogenesis. PLoS Genetics. 2015;11:e1004749
  143. 143. Guo Z, Kanjanapangka J, Liu N, Liu S, Liu C, Wu Z, et al. Sequential posttranslational modifications program FEN1 degradation during cell-cycle progression. Molecular Cell. 2012;47:444-456
  144. 144. Guo Z, Qian L, Liu R, Dai H, Zhou M, Zheng L, et al. Nucleolar localization and dynamic roles of flap endonuclease 1 in ribosomal DNA replication and damage repair. Molecular and Cellular Biology. 2008;28:4310-4319
  145. 145. Henneke G, Koundrioukoff S, Hubscher U. Phosphorylation of human Fen1 by cyclin-dependent kinase modulates its role in replication fork regulation. Oncogene. 2003;22:4301-4313
  146. 146. Guo Z, Zheng L, Xu H, Dai H, Zhou M, Pascua MR, et al. Methylation of FEN1 suppresses nearby phosphorylation and facilitates PCNA binding. Nature Chemical Biology. 2010;6:766-773
  147. 147. Hasan S, El-Andaloussi N, Hardeland U, Hassa PO, Burki C, Imhof R, et al. Acetylation regulates the DNA end-trimming activity of DNA polymerase beta. Molecular Cell. 2002;10:1213-1222
  148. 148. El-Andaloussi N, Valovka T, Toueille M, Hassa PO, Gehrig P, Covic M, et al. Methylation of DNA polymerase beta by protein arginine methyltransferase 1 regulates its binding to proliferating cell nuclear antigen. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2007;21:26-34
  149. 149. El-Andaloussi N, Valovka T, Toueille M, Steinacher R, Focke F, Gehrig P, et al. Arginine methylation regulates DNA polymerase beta. Molecular Cell. 2006;22:51-62
  150. 150. Parsons JL, Tait PS, Finch D, Dianova II, Allinson SL, Dianov GL. CHIP-mediated degradation and DNA damage-dependent stabilization regulate base excision repair proteins. Molecular Cell. 2008;29:477-487
  151. 151. Parsons JL, Tait PS, Finch D, Dianova II, Edelmann MJ, Khoronenkova SV, et al. Ubiquitin ligase ARF-BP1/Mule modulates base excision repair. The EMBO Journal. 2009;28:3207-3215
  152. 152. Hu LY, Chang CC, Huang YS, Chou WC, Lin YM, Ho CC, et al. SUMOylation of XRCC1 activated by poly (ADP-ribosyl)ation regulates DNA repair. Human Molecular Genetics. 2018;27:2306-2317
  153. 153. Prasad R, Shock DD, Beard WA, Wilson SH. Substrate channeling in mammalian base excision repair pathways: Passing the baton. The Journal of Biological Chemistry. 2010;285:40479-40488
  154. 154. Wilson SH, Kunkel TA. Passing the baton in base excision repair. Nature Structural Biology. 2000;7:176-178
  155. 155. Clauson CL, Oestreich KJ, Austin JW, Doetsch PW. Abasic sites and strand breaks in DNA cause transcriptional mutagenesis in Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:3657-3662
  156. 156. Fu D, Calvo JA, Samson LD. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nature Reviews Cancer. 2012;12:104-120
  157. 157. Goodman MF. Error-prone repair DNA polymerases in prokaryotes and eukaryotes. Annual Review of Biochemistry. 2002;71:17-50
  158. 158. Tornaletti S, Maeda LS, Hanawalt PC. Transcription arrest at an abasic site in the transcribed strand of template DNA. Chemical Research in Toxicology. 2006;19:1215-1220
  159. 159. Della-Maria J, Hegde ML, McNeill DR, Matsumoto Y, Tsai MS, Ellenberger T, et al. The interaction between polynucleotide kinase phosphatase and the DNA repair protein XRCC1 is critical for repair of DNA alkylation damage and stable association at DNA damage sites. The Journal of Biological Chemistry. 2012;287:39233-39244
  160. 160. Waters TR, Gallinari P, Jiricny J, Swann PF. Human thymine DNA glycosylase binds to apurinic sites in DNA but is displaced by human apurinic endonuclease 1. The Journal of Biological Chemistry. 1999;274:67-74
  161. 161. Sengupta S, Yang C, Hegde ML, Hegde PM, Mitra J, Pandey A, et al. Acetylation of oxidized base repair-initiating NEIL1 DNA glycosylase required for chromatin-bound repair complex formation in the human genome increases cellular resistance to oxidative stress. DNA Repair (Amst). 2018;66-67:1-10
  162. 162. Cole HA, Tabor-Godwin JM, Hayes JJ. Uracil DNA glycosylase activity on nucleosomal DNA depends on rotational orientation of targets. The Journal of Biological Chemistry. 2010;285:2876-2885
  163. 163. Hinz JM. Impact of abasic site orientation within nucleosomes on human APE1 endonuclease activity. Mutation Research. 2014;766-767:19-24
  164. 164. Hinz JM, Mao P, McNeill DR, Wilson DM 3rd. Reduced nuclease activity of apurinic/apyrimidinic endonuclease (APE1) variants on nucleosomes: Identification of access residues. The Journal of Biological Chemistry. 2015;290:21067-21075
  165. 165. Hinz JM, Rodriguez Y, Smerdon MJ. Rotational dynamics of DNA on the nucleosome surface markedly impact accessibility to a DNA repair enzyme. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:4646-4651
  166. 166. Odell ID, Barbour JE, Murphy DL, Della-Maria JA, Sweasy JB, Tomkinson AE, et al. Nucleosome disruption by DNA ligase III-XRCC1 promotes efficient base excision repair. Molecular and Cellular Biology. 2011;31:4623-4632
  167. 167. Odell ID, Newick K, Heintz NH, Wallace SS, Pederson DS. Non-specific DNA binding interferes with the efficient excision of oxidative lesions from chromatin by the human DNA glycosylase, NEIL1. DNA Repair (Amst). 2010;9:134-143
  168. 168. Odell ID, Wallace SS, Pederson DS. Rules of engagement for base excision repair in chromatin. Journal of Cellular Physiology. 2013;228:258-266
  169. 169. Prasad A, Wallace SS, Pederson DS. Initiation of base excision repair of oxidative lesions in nucleosomes by the human, bifunctional DNA glycosylase NTH1. Molecular and Cellular Biology. 2007;27:8442-8453
  170. 170. Rodriguez Y, Smerdon MJ. The structural location of DNA lesions in nucleosome core particles determines accessibility by base excision repair enzymes. The Journal of Biological Chemistry. 2013;288:13863-13875
  171. 171. House NC, Koch MR, Freudenreich CH. Chromatin modifications and DNA repair: Beyond double-strand breaks. Frontiers in Genetics. 2014;5:296
  172. 172. Ransom M, Dennehey BK, Tyler JK. Chaperoning histones during DNA replication and repair. Cell. 2010;140:183-195
  173. 173. Rodriges Blanko E, Kadyrova LY, Kadyrov FA. DNA mismatch repair interacts with CAF-1- and ASF1A-H3-H4-dependent histone (H3-H4)2 tetramer deposition. The Journal of Biological Chemistry. 2016;291:9203-9217
  174. 174. Amouroux R, Campalans A, Epe B, Radicella JP. Oxidative stress triggers the preferential assembly of base excision repair complexes on open chromatin regions. Nucleic Acids Research. 2010;38:2878-2890
  175. 175. Banerjee D, Mandal SM, Das A, Hegde ML, Das S, Bhakat KK, et al. Preferential repair of oxidized base damage in the transcribed genes of mammalian cells. The Journal of Biological Chemistry. 2011;286:6006-6016
  176. 176. He YF, Li BZ, Li Z, Liu P, Wang Y, Tang Q, et al. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science. 2011;333:1303-1307
  177. 177. Ito S, Shen L, Dai Q, Wu SC, Collins LB, Swenberg JA, et al. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science. 2011;333:1300-1303
  178. 178. Kohli RM, Zhang Y. TET enzymes, TDG and the dynamics of DNA demethylation. Nature. 2013;502:472-479
  179. 179. Maiti A, Drohat AC. Thymine DNA glycosylase can rapidly excise 5-formylcytosine and 5-carboxylcytosine: Potential implications for active demethylation of CpG sites. The Journal of Biological Chemistry. 2011;286:35334-35338
  180. 180. Weber AR, Krawczyk C, Robertson AB, Kusnierczyk A, Vagbo CB, Schuermann D, et al. Biochemical reconstitution of TET1-TDG-BER-dependent active DNA demethylation reveals a highly coordinated mechanism. Nature Communications. 2016;7:10806
  181. 181. Clapier CR, Cairns BR. The biology of chromatin remodeling complexes. Annual Review of Biochemistry. 2009;78:273-304
  182. 182. Hinz JM, Czaja W. Facilitation of base excision repair by chromatin remodeling. DNA Repair (Amst). 2015;36:91-97
  183. 183. Narlikar GJ, Sundaramoorthy R, Owen-Hughes T. Mechanisms and functions of ATP-dependent chromatin-remodeling enzymes. Cell. 2013;154:490-503
  184. 184. Menoni H, Gasparutto D, Hamiche A, Cadet J, Dimitrov S, Bouvet P, et al. ATP-dependent chromatin remodeling is required for base excision repair in conventional but not in variant H2A.Bbd nucleosomes. Molecular and Cellular Biology. 2007;27:5949-5956
  185. 185. Menoni H, Shukla MS, Gerson V, Dimitrov S, Angelov D. Base excision repair of 8-oxoG in dinucleosomes. Nucleic Acids Research. 2012;40:692-700
  186. 186. Nakanishi S, Prasad R, Wilson SH, Smerdon M. Different structural states in oligonucleosomes are required for early versus late steps of base excision repair. Nucleic Acids Research. 2007;35:4313-4321
  187. 187. Czaja W, Mao P, Smerdon MJ. Chromatin remodelling complex RSC promotes base excision repair in chromatin of Saccharomyces cerevisiae. DNA Repair (Amst). 2014;16:35-43
  188. 188. Tsuda M, Cho K, Ooka M, Shimizu N, Watanabe R, Yasui A, et al. ALC1/CHD1L, a chromatin-remodeling enzyme, is required for efficient base excision repair. PLoS One. 2017;12:e0188320
  189. 189. Czaja W, Bespalov VA, Hinz JM, Smerdon MJ. Proficient repair in chromatin remodeling defective ino80 mutants of Saccharomyces cerevisiae highlights replication defects as the main contributor to DNA damage sensitivity. DNA Repair (Amst). 2010;9:976-984
  190. 190. Haldar D, Kamakaka RT. Schizosaccharomyces pombe Hst4 functions in DNA damage response by regulating histone H3 K56 acetylation. Eukaryotic Cell. 2008;7:800-813
  191. 191. Vempati RK, Jayani RS, Notani D, Sengupta A, Galande S, Haldar D. p300-mediated acetylation of histone H3 lysine 56 functions in DNA damage response in mammals. The Journal of Biological Chemistry. 2010;285:28553-28564
  192. 192. Xu F, Zhang K, Grunstein M. Acetylation in histone H3 globular domain regulates gene expression in yeast. Cell. 2005;121:375-385
  193. 193. Citterio E, Rademakers S, van der Horst GT, van Gool AJ, Hoeijmakers JH, Vermeulen W. Biochemical and biological characterization of wild-type and ATPase-deficient Cockayne syndrome B repair protein. The Journal of Biological Chemistry. 1998;273:11844-11851
  194. 194. Citterio E, Van Den Boom V, Schnitzler G, Kanaar R, Bonte E, Kingston RE, et al. ATP-dependent chromatin remodeling by the Cockayne syndrome B DNA repair-transcription-coupling factor. Molecular and Cellular Biology. 2000;20:7643-7653
  195. 195. Yu S, Owen-Hughes T, Friedberg EC, Waters R, Reed SH. The yeast Rad7/Rad16/Abf1 complex generates superhelical torsion in DNA that is required for nucleotide excision repair. DNA Repair (Amst). 2004;3:277-287
  196. 196. Kim MY, Mauro S, Gevry N, Lis JT, Kraus WL. NAD+-dependent modulation of chromatin structure and transcription by nucleosome binding properties of PARP-1. Cell. 2004;119:803-814
  197. 197. Kim MY, Zhang T, Kraus WL. Poly(ADP-ribosyl)ation by PARP-1: ‘PAR-laying’ NAD+ into a nuclear signal. Genes & Development. 2005;19:1951-1967
  198. 198. Rouleau M, Aubin RA, Poirier GG. Poly(ADP-ribosyl)ated chromatin domains: Access granted. Journal of Cell Science. 2004;117:815-825
  199. 199. Groth A, Rocha W, Verreault A, Almouzni G. Chromatin challenges during DNA replication and repair. Cell. 2007;128:721-733
  200. 200. Burgess RJ, Zhang Z. Histone chaperones in nucleosome assembly and human disease. Nature Structural & Molecular Biology. 2013;20:14-22
  201. 201. Yang C, Sengupta S, Hegde PM, Mitra J, Jiang S, Holey B, et al. Regulation of oxidized base damage repair by chromatin assembly factor 1 subunit A. Nucleic Acids Research. 2017;45:739-748
  202. 202. Polo SE, Almouzni G. Chromatin dynamics after DNA damage: The legacy of the access-repair-restore model. DNA Repair (Amst). 2015;36:114-121
  203. 203. Bhakat KK, Mokkapati SK, Boldogh I, Hazra TK, Mitra S. Acetylation of human 8-oxoguanine-DNA glycosylase by p300 and its role in 8-oxoguanine repair in vivo. Molecular and Cellular Biology. 2006;26:1654-1665
  204. 204. Das A, Hazra TK, Boldogh I, Mitra S, Bhakat KK. Induction of the human oxidized base-specific DNA glycosylase NEIL1 by reactive oxygen species. The Journal of Biological Chemistry. 2005;280:35272-35280
  205. 205. Sengupta S, Mantha AK, Mitra S, Bhakat KK. Human AP endonuclease (APE1/Ref-1) and its acetylation regulate YB-1-p300 recruitment and RNA polymerase II loading in the drug-induced activation of multidrug resistance gene MDR1. Oncogene. 2011;30:482-493
  206. 206. Hanawalt PC, Spivak G. Transcription-coupled DNA repair: Two decades of progress and surprises. Nature Reviews. 2008;9:958-970
  207. 207. Menoni H, Wienholz F, Theil AF, Janssens RC, Lans H, Campalans A, et al. The transcription-coupled DNA repair-initiating protein CSB promotes XRCC1 recruitment to oxidative DNA damage. Nucleic Acids Research. 2018;46:7747-7756
  208. 208. Abbotts R, Madhusudan S. Human AP endonuclease 1 (APE1): From mechanistic insights to druggable target in cancer. Cancer Treatment Reviews. 2010;36:425-435
  209. 209. Al-Safi RI, Odde S, Shabaik Y, Neamati N. Small-molecule inhibitors of APE1 DNA repair function: An overview. Current Molecular Pharmacology. 2012;5:14-35
  210. 210. Thakur S, Sarkar B, Cholia RP, Gautam N, Dhiman M, Mantha AK. APE1/Ref-1 as an emerging therapeutic target for various human diseases: Phytochemical modulation of its functions. Experimental & Molecular Medicine. 2014;46:e106
  211. 211. Ferrarotto R, Cardnell R, Su S, Diao L, Eterovic AK, Prieto V, et al. Poly ADP-ribose polymerase-1 as a potential therapeutic target in Merkel cell carcinoma. Head & Neck. 2018;40:1676-1684
  212. 212. Drean A, Lord CJ, Ashworth A. PARP inhibitor combination therapy. Critical Reviews in Oncology/Hematology. 2016;108:73-85
  213. 213. Lu X, Bocangel D, Nannenga B, Yamaguchi H, Appella E, Donehower LA. The p53-induced oncogenic phosphatase PPM1D interacts with uracil DNA glycosylase and suppresses base excision repair. Molecular Cell. 2004;15:621-634
  214. 214. Ma KW, Au SW, Waye MM. Over-expression of SUMO-1 induces the up-regulation of heterogeneous nuclear ribonucleoprotein A2/B1 isoform B1 (hnRNP A2/B1 isoform B1) and uracil DNA glycosylase (UDG) in hepG2 cells. Cell Biochemistry and Function. 2009;27:228-237
  215. 215. Schrofelbauer B, Hakata Y, Landau NR. HIV-1 Vpr function is mediated by interaction with the damage-specific DNA-binding protein DDB1. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:4130-4135
  216. 216. Schrofelbauer B, Yu Q, Zeitlin SG, Landau NR. Human immunodeficiency virus type 1 Vpr induces the degradation of the UNG and SMUG uracil-DNA glycosylases. Journal of Virology. 2005;79:10978-10987
  217. 217. Kim MS, Kondo T, Takada I, Youn MY, Yamamoto Y, Takahashi S, et al. DNA demethylation in hormone-induced transcriptional derepression. Nature. 2009;461:1007-1012
  218. 218. Coey CT, Drohat AC. Defining the impact of sumoylation on substrate binding and catalysis by thymine DNA glycosylase. Nucleic Acids Research. 2018;46:5159-5170
  219. 219. Coey CT, Fitzgerald ME, Maiti A, Reiter KH, Guzzo CM, Matunis MJ, et al. E2-mediated small ubiquitin-like modifier (SUMO) modification of thymine DNA glycosylase is efficient but not selective for the enzyme-product complex. The Journal of Biological Chemistry. 2014;289:15810-15819
  220. 220. Smet-Nocca C, Wieruszeski JM, Leger H, Eilebrecht S, Benecke A. SUMO-1 regulates the conformational dynamics of thymine-DNA glycosylase regulatory domain and competes with its DNA binding activity. BMC Biochemistry. 2011;12:4
  221. 221. Steinacher R, Schar P. Functionality of human thymine DNA glycosylase requires SUMO-regulated changes in protein conformation. Current Biology. 2005;15:616-623
  222. 222. Shibata E, Dar A, Dutta A. CRL4Cdt2 E3 ubiquitin ligase and proliferating cell nuclear antigen (PCNA) cooperate to degrade thymine DNA glycosylase in S phase. The Journal of Biological Chemistry. 2014;289:23056-23064
  223. 223. Slenn TJ, Morris B, Havens CG, Freeman RM Jr, Takahashi TS, Walter JC. Thymine DNA glycosylase is a CRL4Cdt2 substrate. The Journal of Biological Chemistry. 2014;289:23043-23055
  224. 224. Kundu S, Brinkmeyer MK, Eigenheer RA, David SS. Ser 524 is a phosphorylation site in MUTYH and Ser 524 mutations alter 8-oxoguanine (OG): A mismatch recognition. DNA Repair (Amst). 2010;9:1026-1037
  225. 225. Parker AR, O'Meally RN, Sahin F, Su GH, Racke FK, Nelson WG, et al. Defective human MutY phosphorylation exists in colorectal cancer cell lines with wild-type MutY alleles. The Journal of Biological Chemistry. 2003;278:47937-47945
  226. 226. Dorn J, Ferrari E, Imhof R, Ziegler N, Hubscher U. Regulation of human MutYH DNA glycosylase by the E3 ubiquitin ligase mule. The Journal of Biological Chemistry. 2014;289:7049-7058
  227. 227. Dantzer F, Luna L, Bjoras M, Seeberg E. Human OGG1 undergoes serine phosphorylation and associates with the nuclear matrix and mitotic chromatin in vivo. Nucleic Acids Research. 2002;30:2349-2357
  228. 228. Hu J, Imam SZ, Hashiguchi K, de Souza-Pinto NC, Bohr VA. Phosphorylation of human oxoguanine DNA glycosylase (alpha-OGG1) modulates its function. Nucleic Acids Research. 2005;33:3271-3282
  229. 229. Fantini D, Moritz E, Auvre F, Amouroux R, Campalans A, Epe B, et al. Rapid inactivation and proteasome-mediated degradation of OGG1 contribute to the synergistic effect of hyperthermia on genotoxic treatments. DNA Repair (Amst). 2013;12:227-237
  230. 230. Bian Y, Song C, Cheng K, Dong M, Wang F, Huang J, et al. An enzyme assisted RP-RPLC approach for in-depth analysis of human liver phosphoproteome. Journal of Proteomics. 2014;96:253-262
  231. 231. Prakash A, Cao VB, Doublie S. Phosphorylation sites identified in the NEIL1 DNA glycosylase are potential targets for the JNK1 kinase. PLoS One. 2016;11:e0157860
  232. 232. Sui S, Wang J, Yang B, Song L, Zhang J, Chen M, et al. Phosphoproteome analysis of the human Chang liver cells using SCX and a complementary mass spectrometric strategy. Proteomics. 2008;8:2024-2034
  233. 233. Bhakat KK, Hazra TK, Mitra S. Acetylation of the human DNA glycosylase NEIL2 and inhibition of its activity. Nucleic Acids Research. 2004;32:3033-3039
  234. 234. Likhite VS, Cass EI, Anderson SD, Yates JR, Nardulli AM. Interaction of estrogen receptor alpha with 3-methyladenine DNA glycosylase modulates transcription and DNA repair. The Journal of Biological Chemistry. 2004;279:16875-16882
  235. 235. Agnihotri S, Burrell K, Buczkowicz P, Remke M, Golbourn B, Chornenkyy Y, et al. ATM regulates 3-methylpurine-DNA glycosylase and promotes therapeutic resistance to alkylating agents. Cancer Discovery. 2014;4:1198-1213
  236. 236. Bhakat KK, Izumi T, Yang SH, Hazra TK, Mitra S. Role of acetylated human AP-endonuclease (APE1/Ref-1) in regulation of the parathyroid hormone gene. The EMBO Journal. 2003;22:6299-6309
  237. 237. Fantini D, Vascotto C, Marasco D, D'Ambrosio C, Romanello M, Vitagliano L, et al. Critical lysine residues within the overlooked N-terminal domain of human APE1 regulate its biological functions. Nucleic Acids Research. 2010;38:8239-8256
  238. 238. Lirussi L, Antoniali G, Vascotto C, D'Ambrosio C, Poletto M, Romanello M, et al. Nucleolar accumulation of APE1 depends on charged lysine residues that undergo acetylation upon genotoxic stress and modulate its BER activity in cells. Molecular Biology of the Cell. 2012;23:4079-4096
  239. 239. Yacoub A, Kelley MR, Deutsch WA. The DNA repair activity of human redox/repair protein APE/Ref-1 is inactivated by phosphorylation. Cancer Research. 1997;57:5457-5459
  240. 240. Huang E, Qu D, Zhang Y, Venderova K, Haque ME, Rousseaux MW, et al. The role of Cdk5-mediated apurinic/apyrimidinic endonuclease 1 phosphorylation in neuronal death. Nature Cell Biology. 2010;12:563-571
  241. 241. Busso CS, Wedgeworth CM, Izumi T. Ubiquitination of human AP-endonuclease 1 (APE1) enhanced by T233E substitution and by CDK5. Nucleic Acids Research. 2011;39:8017-8028
  242. 242. Hasan S, Stucki M, Hassa PO, Imhof R, Gehrig P, Hunziker P, et al. Regulation of human flap endonuclease-1 activity by acetylation through the transcriptional coactivator p300. Molecular Cell. 2001;7:1221-1231
  243. 243. Rahmeh AA, Zhou Y, Xie B, Li H, Lee EY, Lee MY. Phosphorylation of the p68 subunit of Pol delta acts as a molecular switch to regulate its interaction with PCNA. Biochemistry. 2012;51:416-424
  244. 244. Dong Z, Tomkinson AE. ATM mediates oxidative stress-induced dephosphorylation of DNA ligase IIIalpha. Nucleic Acids Research. 2006;34:5721-5279
  245. 245. Kang HC, Lee YI, Shin JH, Andrabi SA, Chi Z, Gagne JP, et al. Iduna is a poly(ADP-ribose) (PAR)-dependent E3 ubiquitin ligase that regulates DNA damage. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:14103-14108
  246. 246. Chou WC, Wang HC, Wong FH, Ding SL, Wu PE, Shieh SY, et al. Chk2-dependent phosphorylation of XRCC1 in the DNA damage response promotes base excision repair. The EMBO Journal. 2008;27:3140-3150
  247. 247. Levy N, Martz A, Bresson A, Spenlehauer C, de Murcia G, Menissier-de Murcia J. XRCC1 is phosphorylated by DNA-dependent protein kinase in response to DNA damage. Nucleic Acids Research. 2006;34:32-41
  248. 248. Loizou JI, El-Khamisy SF, Zlatanou A, Moore DJ, Chan DW, Qin J, et al. The protein kinase CK2 facilitates repair of chromosomal DNA single-strand breaks. Cell. 2004;117:17-28
  249. 249. Luo H, Chan DW, Yang T, Rodriguez M, Chen BP, Leng M, et al. A new XRCC1-containing complex and its role in cellular survival of methyl methanesulfonate treatment. Molecular and Cellular Biology. 2004;24:8356-8365
  250. 250. Parsons JL, Dianova II, Finch D, Tait PS, Strom CE, Helleday T, et al. XRCC1 phosphorylation by CK2 is required for its stability and efficient DNA repair. DNA Repair (Amst). 2010;9:835-841
  251. 251. Strom CE, Mortusewicz O, Finch D, Parsons JL, Lagerqvist A, Johansson F, et al. CK2 phosphorylation of XRCC1 facilitates dissociation from DNA and single-strand break formation during base excision repair. DNA Repair (Amst). 2011;10:961-969
  252. 252. Wei L, Nakajima S, Hsieh CL, Kanno S, Masutani M, Levine AS, et al. Damage response of XRCC1 at sites of DNA single strand breaks is regulated by phosphorylation and ubiquitylation after degradation of poly(ADP-ribose). Journal of Cell Science. 2013;126:4414-4423
  253. 253. Gocke CB, Yu H, Kang J. Systematic identification and analysis of mammalian small ubiquitin-like modifier substrates. The Journal of Biological Chemistry. 2005;280:5004-5012
  254. 254. Hassa PO, Haenni SS, Buerki C, Meier NI, Lane WS, Owen H, et al. Acetylation of poly(ADP-ribose) polymerase-1 by p300/CREB-binding protein regulates coactivation of NF-kappaB-dependent transcription. The Journal of Biological Chemistry. 2005;280:40450-40464
  255. 255. Kauppinen TM, Chan WY, Suh SW, Wiggins AK, Huang EJ, Swanson RA. Direct phosphorylation and regulation of poly(ADP-ribose) polymerase-1 by extracellular signal-regulated kinases 1/2. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:7136-7141
  256. 256. Kashima L, Idogawa M, Mita H, Shitashige M, Yamada T, Ogi K, et al. CHFR protein regulates mitotic checkpoint by targeting PARP-1 protein for ubiquitination and degradation. The Journal of Biological Chemistry. 2012;287:12975-12984

Written By

Shiladitya Sengupta, Chunying Yang, Bradley J. Eckelmann, Muralidhar L. Hegde and Sankar Mitra

Submitted: 20 June 2018 Reviewed: 12 October 2018 Published: 11 December 2018