Next Article in Journal
Antitumor Effects of Ral-GTPases Downregulation in Glioblastoma
Previous Article in Journal
Caffeine-Induced Acute and Delayed Responses in Cerebral Metabolism of Control and Schizophrenia-like Wisket Rats
Previous Article in Special Issue
Insights in Post-Translational Modifications: Ubiquitin and SUMO
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

DNA Damage Response Regulation by Histone Ubiquitination

Laboratory of Hygiene and Public Health, Department of Medical Technology, School of Life and Environmental Science, Azabu University, 1-17-71 Fuchinobe, Chuo-ku, Sagamihara 252-5201, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8187; https://doi.org/10.3390/ijms23158187
Submission received: 17 June 2022 / Revised: 19 July 2022 / Accepted: 22 July 2022 / Published: 25 July 2022
(This article belongs to the Special Issue The Role of Post-translational Modifications in Chromatin and Cancer)

Abstract

:
Cells are constantly exposed to numerous genotoxic stresses that induce DNA damage. DNA double-strand breaks (DSBs) are among the most serious damages and should be systematically repaired to preserve genomic integrity. The efficiency of repair is closely associated with chromatin structure, which is regulated by posttranslational modifications of histones, including ubiquitination. Recent evidence shows crosstalk between histone ubiquitination and DNA damage responses, suggesting an integrated model for the systematic regulation of DNA repair. There are two major pathways for DSB repair, viz., nonhomologous end joining and homologous recombination, and the choice of the pathway is partially controlled by posttranslational modifications of histones, including ubiquitination. Histone ubiquitination changes chromatin structure in the vicinity of DSBs and serves as a platform to select and recruit repair proteins; the removal of these modifications by deubiquitinating enzymes suppresses the recruitment of repair proteins and promotes the convergence of repair reactions. This article provides a comprehensive overview of the DNA damage response regulated by histone ubiquitination in response to DSBs.

1. Introduction

Many types of DNA damage occur in cells, of which the most serious is DNA double-strand break (DSB) caused by ionizing radiation, anticancer chemotherapeutic drugs, and endogenous replication fork collapse [1]. Anticancer drugs include cross-linking agents, such as cisplatin and mitomycin C, alkylating agents, such as methyl methanosulfate, and radiomimetic agents, such as bleomycin [2]. Ionizing radiation produces DNA single-strand breaks (SSBs) using water radiolysis-generated radicals [2,3]. Closely opposed SSBs created in complementary DNA strands within a single helical turn lead to DSBs [4]. Since unrepaired DSBs result in the accumulation of chromosomal aberrations and translocations, their repair is essential for cell survival [5]. Two major pathways are involved in DSB repair, viz., nonhomologous end joining (NHEJ) and homologous recombination (HR) [6]. NHEJ can occur throughout the cell cycle and rejoin break ends without sequence homology. It is generally considered to be error-prone, owing to its minimal end processing [7,8]. Contrarily, HR is activated in the S and G2 phases of the cell cycle and when sister chromatids can be used as templates for repair [9]. HR is initiated by the 5′ DNA end resection and the binding of Rad51 to single-stranded DNA. The later stages of HR involve the search for a homologous sequence, formation of a displacement loop (D-loop) by single-stranded DNA invasion, removal of Rad51, and repair synthesis to copy the donor template to restore the genetic information at the break sites [10].
When DSBs occur, the early response is primarily conducted using ataxia-telangiectasia mutated (ATM) protein kinase [11,12]. At DSB sites, ATM phosphorylates many substrates and induces complex downstream posttranslational modifications (PTMs) to promote DNA repair [12,13,14]. Signaling pathways in early responses to DSBs promote the recruitment of 53BP1 and BRCA1 in later responses [8]. Accumulated ATM phosphorylates H2AX, but DNA-dependent protein kinase catalytic subunit (DNA-PKcs), ataxia telangiectasia, and Rad3-related (ATR) also phosphorylate H2AX [15,16,17]. Additionally, increased levels of γH2AX on either side of the DSB further promote chromatin decondensation and DSB repair [18]. Chromatin structure is a key determinant of DSB repair pathway choice. The functional unit of chromatin is the nucleosome, which generally contains 146 base pairs of DNA wrapped around a histone octamer containing four types of histone proteins, viz., H2A, H2B, H3, and H4 [19,20]. Histone proteins have an unstructured amino terminus or C-terminus that extends out from the nucleosome and is known as the histone tail. The histone tail region is enriched in basic amino acids, such as positively charged lysine and arginine that interact with the negatively charged phosphate groups of DNA [21]. The histone tail is subjected to various posttranslational modifications, such as acetylation, methylation, phosphorylation, and ubiquitination. These modifications affect the interaction of histone proteins with DNA or reader proteins. Therefore, histone modifications play a critical role in the signaling network of the DNA damage repair response [22,23]. Upon DNA damage, various posttranslational modifications (methylation, ubiquitination, acetylation, etc.) of histones form the basis for the recruitment of downstream effectors, resulting in chromatin remodeling and regulating enzymatic signaling cascades [24]. This histone signaling network is organized into multiprotein complexes containing modules that are writers, readers, and erasers of chromatin marks, respectively. The modularity of these complexes integrates different histone modifications and allows for highly complex signaling [24].
Histone ubiquitination alters the chromatin structure before and after repair, propagates damage signals, and serves as a platform for the recruitment of repair proteins, which have been extensively reviewed [8,25,26,27,28]. Ubiquitin is a small protein of 76 amino acids with a C-terminal diglycine tail and is highly evolutionarily conserved across species. Covalent attachment of ubiquitin to an ε-amino group of a lysine residue on the substrate proteins, ubiquitination, is catalyzed by E1-activating enzymes, E2-conjugating enzymes, and E3 ubiquitin ligases. The C-terminal carboxyl group of ubiquitin is first activated by E1 in an ATP-dependent process and enters into a thioester linkage with the catalytic cysteine of E1. Then, ubiquitin is transferred to the catalytic cysteine of E2 through a transesterification reaction. Subsequently, E3 ligase is involved in the transfer of the ubiquitin from E2 to the ε-amino group of a substrate lysine, promoting the formation of an isopeptide bond between the C-terminal glycine residue of ubiquitin and lysine residue. The addition of a single ubiquitin molecule to a target protein is termed monoubiquitination, and monoubiquitination to additional lysine residues of the substrate is termed multiubiquitination. Furthermore, ubiquitin itself has seven lysine residues (lysine 6 (K6), K11, K27, K29, K33, K48, and K63) and an N-terminal amine, through which ubiquitin is attached, forming a polyubiquitin chain termed polyubiquitination. Conversely, deubiquitinating enzymes (DUBs) remove ubiquitin modifications from substrate proteins [29].
Ubiquitination of histones H2A and H2B is highly site-selective and is one of the key posttranslational modifications that regulate DNA damage response. RNF168, RING1A/RING1B in the polycomb repressive complex 1 (PRC1), and BRCA1-BARD1 modify H2AK13/K15, K118/119, and K127/129, respectively. Ubiquitination of these different sites on H2A triggers different physiological responses [25]. In the DNA damage response, H2A ubiquitination by RNF168 regulates HR and NHEJ [30,31,32,33,34], and H2A ubiquitination by BRCA1/BARD1 promotes HR [6,35,36]. Ubiquitination of H2A by the PRC1 complex plays the global function of transcriptional repression and may play the same role around the DNA damage response [37]. The RNF20/RNF40 heterodimer specifically ubiquitinates K120 of H2B and functions in the activation of damage checkpoints and repair initiation [25,36].
These specific ubiquitinations of histones form a crucial component of the regulatory network facilitating the DNA damage response, and recent advances have begun to elucidate the molecular basis of the specificity and mechanisms involved in repair pathway selection. In this review, we focus on the role of histone ubiquitination in the DNA damage response and maintenance of genome stability through the integration of multiple signaling entities and selection of repair pathways.

2. Histone Ubiquitination Regulates DNA Repair

2.1. H2AK13/15 Ubiquitination by RNF8-RNF168 and Deubiquitination

Histone ubiquitination by multiple E3 ligases contributes to appropriate DSB repair, and deubiquitination by multiple DUBs regulates these reactions (Table 1). There are also several proteins with ubiquitin-binding domains, which are the reader proteins that recognize multiple histone ubiquitination (Table 1). RNF168, an E3 ubiquitin ligase, consists of a RING finger domain and two ubiquitin interaction motifs, termed MIU for motif interacting with ubiquitin, that selectively bind to ubiquitin chains. The RING finger domain of RNF168 is critical for its ubiquitin E3 ligase activity, and the two MIU domains bind to ubiquitylated H2A. RNF168 functions in the monoubiquitination and K63 polyubiquitination of H2A/H2AX K13/15 that occurs after DSBs (Figure 1). RNF168 catalyzes H2A/H2AX monoubiquitination, whereas RNF8, a RING finger ubiquitin ligase, alone is insufficient to induce H2A/H2AX ubiquitination. RNF168 is responsible for the monoubiquitination of H2AK13/15, and RNF8 is efficient in extending its monoubiquitination to form a K63-linked ubiquitin chain [38]. RNF8 assembles at DSBs through the interaction of its FHA domain with the phosphorylated adaptor protein MDC1 at DSB and recruits RNF168 for K63 polyubiquitination and other downstream effectors [38,39,40,41]. There is extensive evidence on how RNF8 manages the recruitment of RNF168. It was first demonstrated that RNF8 and UBC13 (also known as UBE2N), an E2 ubiquitin-conjugating enzyme, mediate the K63-linked ubiquitinate H1-linker histone, promoting the recruitment of RNF168 that monoubiquitinates H2A at lysine 13/15 [42]. It was also demonstrated that ataxia telangiectasia mutated (ATM)-induced phosphorylation of L3MBTL2 induces the interaction of MDC1 at DSB, which is subsequently ubiquitylated by RNF8. Ubiquitylated L3MBTL2 facilitates RNF168 recruitment to the damage sites and promotes H2A polyubiquitylation [43]. These data suggest that RNF8-induced ubiquitination of more than one protein mediates the recruitment of RNF168 to the damage site [26,27].
RNF168-induced monoubiquitination of H2AK15 is directly bound by the ubiquitin-dependent recruitment (UDR) motif of tumor protein 53 (TP53)-binding protein 1 (53BP1), which results in its accumulation at the DSB sites [44] (Figure 2). 53BP1 is a key regulator of NHEJ to inhibit DNA end resection, the initiation step of HR [30]. Furthermore, 53BP1 binds lysine 20 dimethylation on histone H4 (H4K20me2) through a tandem Tudor domain [45]. The two domains of 53BP1 are essential for its accumulation at the damage sites [30]. Alternatively, the NuA4/TIP60 (histone acetyltransferase KAT5) acetyltransferase complex also binds H4K20me2, and its binding competes for 53BP1 binding [46]. The Tip60 acetyltransferase complex is recruited to the DSB to acetylate H4, H2A, and H2AX, promoting chromatin relaxation [47,48,49]. In addition, Tip60-dependent H4 acetylation attenuates 53BP1 binding to H4K20me2 through the Tudor domain [8,24,50]. Moreover, the Tip60 complex can acetylate H2AK15 and its acetylation mutually exclusive ubiquitination by RNF168 [46]. Furthermore, Tip60 deficiency suppresses HR in a 53BP1-dependent manner. These results indicate that Tip60 competitively inhibits the recruitment of 53BP1 to chromatin, thereby suppressing 53BP1-dependent repair. Meanwhile, the tumor suppressor protein BRCA1-associated RING domain protein 1 (BARD1) interacts with the monoubiquitin of H2AK13/K15 through a BRCT domain UDR motif (BUDR) [31,32,33,34,35,51] (Figure 2). In addition to its binding to H2AK15ub, BARD1 interacts with lysine 20 of histone H4 in its unmethylated form (H4K20me0), but not with H4K20me2, through an ankyrin repeat domain [52]. Consequently, 53BP1 and BARD1–BRCA1 compete for the same damage-induced histone marker, and the BARD1–BRCA1 complex can specifically inhibit 53BP1 binding through dual recognition of H2AK15ub and H4K20me0. The RNF8/RNF168-induced K63 polyubiquitination of H2AK15 interacts with RAP80 throughits ubiquitin-interacting motif to promote the recruitment of the BRCA1-A complex. The BRCA1-A complex contains BRCA1/BARD1, RAP80, DUBs BRCC36, and BRCC45 [53,54]. RAP80 targets the complex to the site of DNA damage, where the complex limits DNA end resection to prevent excessive HR [22,55] (Figure 2).
Like other posttranslational modifications, such as phosphorylation, ubiquitination is reversible, and deubiquitination can be accomplished by specific DUBs to prevent excessive ubiquitylation. DUBs, including ubiquitin-specific proteases, ubiquitin C-terminal hydrolases, ovarian tumor proteases (OTUs), and Machado–Joseph Disease protease family (MJDs), are responsible for the cleavage of ubiquitin chains from substrate proteins [56]. Numerous DUBs have been demonstrated to be involved in the DNA damage response (Table 1). USP3 [57], USP16 [58], USP29 [59], USP44 [59], USP49 [60], USP51 [61], A20 [62], BRCC36 [63], and OTUB1 [64] have been suggested to inhibit DSB-induced histone H2A ubiquitination. Overexpression of USP3 and USP44 suppresses the localization of 53BP1 to the damage site. This suggests that these DUBs negatively regulate the recruitment of downstream factors after RNF8/RNF168-generated H2A ubiquitination [45,59]. The recruitment of USP44 to the damage site was initiated in an RNF8- or RNF168-dependent manner and therefore may facilitate the hydrolysis of ubiquitination induced using RNF8/RNF168 [59]. USP3-deficient mice exhibited short lifespan, increased tumorigenesis, and spontaneous genomic damage in MEFs [65]. USP49 overexpression suppressed ubiquitylation of γH2AX and DSB-induced foci formation of 53BP1. This resulted in hypersensitivity to DNA-damaging anticancer treatments [60]. Overexpression of USP51 suppresses the formation of ionizing radiation (IR)-induced 53BP1 and BRCA1 but not RNF168 foci, suggesting that USP51 functions downstream from RNF168 in the DNA damage response [61]. OTUB1 binds to UBC13 (UBE2N), an E2 ubiquitin-conjugating enzyme for RNF168, and restricts its interaction with RNF168, suppressing the DSB responses by inhibiting RNF168-induced polyubiquitylation independently of its catalytic activity [64]. Furthermore, USP11-mediated histone deubiquitination, together with NurD-mediated histone deacetylation, promotes the termination of the DNA damage repair response and results in the reconstruction of chromatin structure [66]. Further research is necessary to elucidate the function of these DUBs.
Table 1. Histone lysine ubiquitination enzymes and deubiquitination enzymes involved in DNA damage response.
Table 1. Histone lysine ubiquitination enzymes and deubiquitination enzymes involved in DNA damage response.
HistoneTargetE3DUBReaderFunctionRefs.
H2AK13/15RNF8/RNF168USP3, USP29 USP44, USP51,53BP1, BARD1Recruitment of 53BP1 or BARD1 to DSBs[31,32,33,34,35,38,39,40,41,42,51,52,55,56,57,58,61,62,63,64,65,66]
K127/129BRCA1/BARD1USP48SMARCAD1Recruitment of SMARCAD1 to promote end resection for HR[35,67,68]
K118/K119RING1A/RING1BBAP1, USP16RYBP, JARID2, ZRF1Transcriptional repression[69,70,71,72,73,74,75,76,77,78,79]
H2BK120RNF20/RNF40USP22, USP11DOT1LTranscriptional regulation, Recruitment of damage proteins[12,80,80,81,82,83,84,85,86,87]
H2AXK13/15RNF8/RNF168USP3, USP49,53BP1Recruitment of 53BP1 to DSBs[38,41,60]
H3K14/18/23UHRF1USP3, USP29 USP44, USP51,DNMT1Maintaining DNA methylation during DNA replication[88,89,90]
CUL4/DDB/ROC1USP4853BP1, BARD1Recruitment XPC to the damaged foci after UV irradiation[91]
H4K91BBAPBAP1, USP16SMARCAD1Modulation of 53BP1 foci formation[92]
CUL4/DDB/ROC1USP22, USP11RYBP, JARID2, ZRF1Recruitment XPC to the damaged foci after UV irradiation[91]
H1 RNF8USP3, USP29 USP44, USP51, Recruitment of RNF168 to DSBs[42]

2.2. H2AK127/129 Ubiquitination and Repair Pathway Choice

The tumor suppressor protein, breast and ovarian cancer predisposition protein-1 (BRCA1), promotes distinct steps of DSB repair by HR and protects DNA replication forks. Cancers originating from germline mutations in the BRCA1 gene cannot be repaired by HR and are sensitive to exogenous DNA-damaging agents such as PARP1 inhibitors (PARPi) or platinum [93]. BRCA1 counteracts the activity of the 53BP1–RIF1–Shieldin complex, protecting DSB ends from 5′-end resection, and activates the resection of DNA ends. DNA end resection is a critical step in the HR repair pathway. The BRCA1-induced resection of DSBs provides 3′ single-stranded DNA (ssDNA) overhangs and promotes RAD51 filament formation [94]. However, the precise molecular mechanism by which BRCA1 activates resection against the 53BP1 complex is unclear. BRCA1 may physically reposit 53BP1 from the DSB site [67,95] or recruit phosphatases that dephosphorylate 53BP1, resulting in the loss of binding between 53BP1 and RIF1 [96]. As BRCA1 is a component of several different protein complexes, it exerts its molecular function in HR owing to its interaction with proteins and the formation of different multiprotein complexes that are composed of different constituent factors and localize to the damage site in different ways [97]. The above-described formation of the BRCA1–RAP80 complex is governed by the interaction between RAP80 and RNF168/RNF8 catalyzed by K63-linked ubiquitin chains on chromatin surrounding the DSBs [53]. The BRCA1-A complex, including RAP80, was shown to restrict HR rather than promote DNA end resection during the S/G2 phase of the cell cycle [54,55]. It has been demonstrated that RAP80 binds ubiquitinated H2B following DNA damage; however, its significance in DNA repair remains unclear [98]. BRCA1 has several functional domains, including the RING domain, BRCT repeats, a coiled-coil (CC) domain, and a central unstructured region encoded by exon11. BRCA1 and BARD1 interact through their respective N-terminal RING domains. BRCA1 is recruited to the DSB sites through the interaction with BARD1 in the monoubiqitination of H2AK13/K15 [31,32,33,34,35,51]. BRCA1 and BARD1 form a heterodimer and exhibit E3 ubiquitin ligase activity due to the RING domain at the N-terminus of these proteins, and they may promote nucleolytic resection through its interaction with the CtBP-interacting protein (CtIP) in an MRN (RAD50/NBS1/MRE11)-dependent manner [97,99]. The BRCT repeat of BRCA1 forms a phosphopeptide-binding region that facilitates interaction with proteins such as CtIP, ABRAXAS, and BACH1. Recently, nucleosomal histone H2A has been identified as a substrate for the BRCA1–BARD1-dependent E3 ligase activity in DNA repair. The heterodimeric RING domains are sufficient for promoting the ubiquitylation of lysine residues 125, 127, and 129 of H2A, and also the ubiquitylation of lysine residue 123 of the histone variant macroH2A1 [35,68]. The ligase activity of BRCA1–BARD1 contributes to the function of BRCA1 in DNA end resection, which is required before the formation of the RAD51 nucleofilament in HR repair. 53BP1 and its effector protein, human REV7 or Artemis, suppress DNA resection in the absence of BRCA1 [100]. The recruitment of BRCA1 to the DSB sites is associated with 53BP1 removal from DSBs and initiates long-range DNA end resection and HR [95,101]. It was demonstrated that the ubiquitination of H2A K125/K127/K129 recruits the SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily A, containing DEAD/H box1 (SMARCAD1) through its ubiquitin-binding CUE domains. SMARCAD1 was found to be required for chromatin remodeling, 53BP1 repositioning, DNA end resection, and HR [67]. BRCA1–BARD1-induced H2A ubiquitination and subsequent SMARCAD1-dependent 53BP1 repositioning are critical regulators of DNA repair [67]. These findings emphasize the multifaceted and complex function of BRCA1 in the regulation of DNA end resection and promotion of RAD51 loading, as well as the importance of BRCA1 in the pathway selection for DNA damage repair. BRCA1 also directly binds to PALB2 through their respective CC domains, forming a macrocomplex comprising BRCA1–PALB2–BRCA2–RAD51 that directs RAD51-loading onto ssDNA [94,102]. However, BRCA1-independent loading of PALB2 occurs through the activity of ATR and RNF168, and the importance of its interaction remains controversial [103,104,105,106]. USP48 has been identified as an H2A DUB, specific for the BARD1–BRCA1-induced ubiquitination of H2A K125/K127/K129 but not for H2AK119 or H2AK13/15 [107]. Overexpression of USP48 shortens the length of DNA end resection. However, USP48 loss increases DNA end resection and RAD51 foci formation as in the case of 53BP1 depletion, suggesting that USP48 antagonizes BRCA1 function and restricts RAD51 accumulation for promoting genome stability [107].

2.3. H2AK119 Monoubiquitination Promotes Transcriptional Regulation in the DSB Response

In the DNA damage response, histone H2A is ubiquitinated at different sites and thus plays different physiological roles. After the occurrence of DSBs, ubiquitylation at lysine 125/127/129 by BRCA1/BARD1 E3 ubiquitin ligases promotes HR repair [67]. Moreover, studies have reported that ubiquitination of lysine 13/15 by RNF168 is a possible basis for both the NHEJ and HR pathways [31,32,33,38]. H2A is also monoubiquitylated on lysine K119 (H2AK119ub1) [69], which induces transcription repression. This is promoted by the polycomb-group protein (PcG)-containing complex, especially polycomb repressive complex 1 (PRC1) [37,70]. The core component of the PRC1 complex, E3 ubiquitin ligase RING1A/RING1B, stimulates the monoubiquitination of H2AK119 [27,71,72,73]. H2AK119ub1 deposition is associated with repressed genes, particularly PcG-target genes, and colocalized with H3K27 trimethylation (H3K27me3), which is critical for transcription repression catalyzed by PcG repressive complex 2 (PRC2) [108,109,110]. After DNA damage, both the PRC1 and PRC2 subunits are recruited to the damage site, suggesting a role for H2AK119ub and H3K27me3 in transcription silencing in the damaged chromatin [111,112,113]. The repression of transcription in neighboring DSBs prevents conflict between transcription machineries and DNA repair [114,115,116]. Recent studies have suggested that BMI1 and RING1A/RING1B heterodimer facilitate DSB repair, which is dependent on their E3 ligase activity toward H2AK119; moreover, they contribute to γ-H2AX foci formation and recruitment of damage response factors [74,75]. In addition, BMI-mediated H2AK119 ubiquitylation induces DNA end resection, allowing the recruitment of downstream factors, such as replication protein A (RPA), BRCA1, and RAD51, at DNA damage sites, as well as the progression of HR. H2AK119ub could recruit the resection factor CtIP, interacting with ubiquitin moieties at the DSB site [36,76,112].
BAP1 is a deubiquitinase of the ubiquitin carboxyl-terminal of the hydrolase family that regulates gene expression and other cell functions through the deubiquitination of histone H2AK119ub [77]. It binds with ASXL1 to form the polycomb repressive deubiquitinase (PR-DUB) complex and deubiquitinates H2A [77]. Upon DNA damage, BAP is phosphorylated and recruited to the DSB sites in an ATM-dependent manner [78]. BAP1 deficiency inhibits efficient HR repair and increases radiation sensitivity, suggesting that the deubiquitination of H2AK119ub promotes HR [78]. Another DUB, USP16, also deubiquitinates H2AK119ub and promotes the reversal of transcription silencing [79].

2.4. H2B Ubiquitination and the DNA Damage Response

Ubiquitination of H2AK13/15 has been primarily implicated in the DNA damage response, and monoubiquitination of H2AK119 is involved in transcriptional activity silencing, whereas monoubiquitination of H2B promotes transcriptional elongation and activation. This monoubiquitination regulates replication, DNA damage response, cellular proliferation, and developmental plasticity [30,35,117,118]. In mammals, the E3 ligase RNF20/RNF40 heterodimer catalyzes the monoubiquitination of H2B at lysine 120 (H2B K120ub1) [80,81,82] and the deubiquitination modules of the SAGA complex, composed of ATXN7 (homolog of Sgf73), ATXN7L3 (homolog of Sgf11), ENY2 (homolog of Sus1), and USP22 deubiquitinate H2B [83,84]. Ubiquitination and deubiquitination of H2B are implicated in transcriptional regulation and DNA damage response [85,86]. The level of H2BK120ub1 increases following exposure to IR or treatment with radiomimetic drugs, suggesting the role of H2BK120ub in DSB repair [80,81,82]. Upon DNA damage, the ATM kinase phosphorylates the Ser172 and Ser553 residues of RNF20 and Ser114 of RNF40, facilitating the recruitment of the RNF20/RNF40 heterodimer to the DSB site and the catalysis of H2Bub1 [12,81,82]. Like the phosphorylated histone H2AX (γ-H2AX) by ATM, H2Bub1 accumulates in the vicinity of DSBs [82]. This reaction is not involved in the very early step of accumulation of damage sensor proteins, such as 53BP1 and MDC1; however, it is required for the timely recruitment of the NHEJ and HR factors, such as XRCC4, KU80, RPA, RAD51, RAP80, and BRCA1. In this reaction, H2BK120ub1 induces chromatin opening and serves as a platform for both the NHEJ and HR proteins, promoting optimal DSB repair through both pathways [80,81]. It has been reported that the depletion of the RNF20/RNF40 heterodimer resulted in the defects of class switch recombination (CSR), indicating that the heterodimer is required for the distal end joining of DSBs, i.e., efficient NHEJ [82]. It has also been reported that the RNF20/RNF40 heterodimer is required for DSB repair by HR. Altogether, RNF20 and RNF40 function in the DNA damage response proximal to a choice of either the NHEJ or HR pathway [28]. RNF20 is also required for the induction of the chromatin remodeling factor SNF2H to the DSB site, suggesting a role for H2BK120ub1 in chromatin remodeling at the DSB site [80,118].
H2B lysine 120 (H2BK120) is either acetylated or ubiquitinated. The conversion between the ubiquitination and acetylation of H2AK120 is performed by the PCAF and SAGA complexes [82]. USP22, a deubiquitinase of the SAGA complex, removes ubiquitin from H2BK120ub during the repair of programmed DSBs in B cells [87]. USP22-induced deubiquitination is critical for CSR, activation-induced cytidine deaminase, and IR-induced DSB repair, and the H2Bub level was found to be increased in USP22-deficient splenic B cells [87]. USP11 also deubiquitinates H2BK120 and H2AK119. USP11 has been demonstrated to be associated with the NuRD complex and related to efficient DNA repair, inducing chromatin condensation and genome stability [66].

2.5. Ubiquitination of H3 and H4 in the DNA Damage Response

Histones H3 and H4 are also ubiquitinated; however, their role in DNA damage response remains unclear. The well-characterized ubiquitination of H3 is H3K14/18/23 [88,89,90]. During DNA replication, the RING domain of ubiquitin-like with PHD and RING finger domain 1 (UHRF1) ubiquitinate histone H3 tail at K14/18/23, which are then recognized by DNMT1 to methylate hemimethylated DNA. DNA interstrand crosslinking (ICL) is a major type of DNA damage during DNA replication, and the FA pathway is mainly responsible for ICL repair. The major function of the FA pathway is presumably the regulation of SLX4/FANCP, the ICL lesion-processing nuclease, by the monoubiquitination of FANCD2/FANCI [119,120,121,122,123]. Intriguingly, UHRF1 functions in ICL repair by binding to ICL lesions that are FA pathway-independent barriers to DNA replication and recruiting ICL repair nucleases [9]. Hence, UHRF1 functions not only in maintaining DNA methylation but also in the DNA damage response. Furthermore, a ubiquitin ligase complex, comprising CUL4A, CUL4B, DDB1, DDB2, and ROC1 (RBX1), catalyzes H3 and H4 ubiquitylation in response to UV irradiation [91]. In addition to these H3 or H4 modifications [91], H4K91 ubiquitination has been found to occur after DNA damage [92], although its detailed mechanism requires further investigation.

3. Conclusions

Eukaryotic cells respond to various genotoxic stresses, and histones and their variants proximal or distal to the DNA damage site undergo posttranslational modifications to regulate chromatin structure. Histone modifications are closely associated with the initiation, progression, and convergence of the DNA damage repair response. The function of histone modifications in DNA damage repair depends on reader proteins that directly bind to modified residues (Table 1). Histone methylation readers recognize methylated residues through several different domains, such as chromodomain (CD), malignant brain tumor (MBT), Tudor, and plant homeodomain (PHD). Similarly, there are several proteins with ubiquitin-binding domains that are reader proteins and recognize histone ubiquitination following DNA damage. Several reader proteins recognize the ubiquitination of chromatin at several different sites, and the antagonistic recognition of ubiquitination of the same chromatin by different reader proteins also appears to select the optimal repair pathway for the damage [22,23]. Histone modification in the DNA damage repair reaction is also recognized by several proteins, among which the best-known is 53BP1, a central protein in the NHEJ pathway. As reviewed in this article, 53BP1 binds to H2AK15ub through a UDR motif and transfers it to the DSB site [30,44]. 53BP1 also directly binds to H4K20me2 in the tandem Tudor domain [45]. Contrarily, BARD1 binds to H2AK15ub through BUDR. BARD1 does not bind to H4K20me2 but binds to unmethylated H4K20me0 [31,32,51]. BARD1 also binds to BRCA1 and performs H2A ubiquitylation (K125/127/129), recruits SMARCAD1 to the DSB sites, and reassociates DNA ends by 53BP1 repositioning [67]. Therefore, BARD1 plays a central role in the HR pathway by binding to BRCA1 and antagonizing NHEJ in the repair response, indicating that the two reader proteins antagonistically recognize a common histone modification [67].
Histone modification has been shown to play a vital role in the selection of DSB repair pathways; however, it also plays an essential role in the interaction with transcription and replication. Ubiquitination of histone lysine residues regulates not only the DNA damage response but also various biological phenomena, such as DNA methylation maintenance and gene expression regulation. During DNA replication, UHRF1 catalyzes multiple monoubiquitination of H3 tail, which is then recognized by DNMT1 to methylate hemimethylated DNA [88,89,90]. Moreover, the regulation of transcription is closely related to DNA damage repair. Upon DNA damage, the chromatin proximal to the damage site is regulated by transcriptional repression to prevent conflict between transcription and repair [111,112,113]. Of these regulatory mechanisms, PRC1 and PRC2 are well known to be recruited to the damage site and induce H2AK119ub and H3K27me3, respectively [124]. Consequently, transcription is repressed in the chromatin proximal to the damage site, and binding to H2AK119ub recruits the DNA end resection factor CtIP to promote HR repair [76]. Therefore, histone modification, which functions in the DNA damage repair response, is also associated with transcription and replication. Nevertheless, only a few DNA damage response proteins have been shown to bind to histone modification. Further biochemical and structural analyses would demonstrate that many more proteins are histone modification readers. In addition, more detailed characterization of the chromatin structure at the damage site will provide a chronological understanding of the DNA damage response.

Author Contributions

Conceptualization, writing (original draft preparation, review, and editing), N.M.; review and editing; M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported in part by Grant-in-Aid for Scientific Research (C) from the Japan Society for the Promotion of Science (Grant No. 22K06604 to N.M.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roos, W.P.; Thomas, A.D.; Kaina, B. DNA damage and the balance between survival and death in cancer biology. Nat. Rev. Cancer 2016, 16, 20–33. [Google Scholar] [CrossRef]
  2. Mehta, A.; Haber, J.E. Sources of DNA Double-Strand Breaks and Models of Recombinational DNA Repair. Cold Spring Harb. Perspect. Biol. 2014, 6, a016428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ward, J.F. The complexity of DNA damage: Relevance to biological consequences. Int. J. Radiat. Biol. 1994, 66, 427–432. [Google Scholar] [CrossRef] [PubMed]
  4. Milligan, J.R.; Ng, J.Y.; Wu, C.C.; Aguilera, J.A.; Fahey, R.C.; Ward, J.F. DNA repair by thiols in air shows two radicals make a double-strand break. Radiat. Res. 1995, 143, 273–280. [Google Scholar] [CrossRef] [PubMed]
  5. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Chapman, J.R.; Taylor, M.R.G.; Boulton, S.J. Playing the end game: DNA double-strand break repair pathway choice. Mol. Cell 2012, 47, 497–510. [Google Scholar] [CrossRef] [Green Version]
  7. Rothkamm, K.; Krüger, I.; Thompson, L.H.; Löbrich, M. Pathways of DNA double-strand break repair during the mammalian cell cycle. Mol. Cell Biol. 2003, 23, 5706–5715. [Google Scholar] [CrossRef] [Green Version]
  8. Panier, S.; Boulton, S.J. Double-strand break repair: 53BP1 comes into focus. Nat. Rev. Mol. Cell Biol. 2014, 15, 7–18. [Google Scholar] [CrossRef]
  9. van Gent, D.C.; Hoeijmakers, J.H.; Kanaar, R. Chromosomal stability and the DNA double-stranded break connection. Nat. Rev. Genet. 2001, 2, 196–206. [Google Scholar] [CrossRef]
  10. Renkawitz, J.; Lademann, C.A.; Jentsch, S. Mechanisms and principles of homology search during recombination. Nat. Rev. Mol. Cell Biol. 2014, 15, 369–383. [Google Scholar] [CrossRef]
  11. Maréchal, A.; Zou, L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 2013, 5, a012716. [Google Scholar] [CrossRef] [PubMed]
  12. Matsuoka, S.; Ballif, B.A.; Smogorzewska, A.; McDonald, E.R.; Hurov, K.E.; Luo, J.; Bakalarski, C.E.; Zhao, Z.; Solimini, N.; Lerenthal, Y.; et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 2007, 316, 1160–1166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Yu, J.; Qin, B.; Lou, Z. Ubiquitin and ubiquitin-like molecules in DNA double strand break repair. Cell Biosci. 2020, 10, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mu, J.-J.; Wang, Y.; Luo, H.; Leng, M.; Zhang, J.; Yang, T.; Besusso, D.; Jung, S.Y.; Qin, J. A proteomic analysis of ataxia telangiectasia-mutated (ATM)/ATM-Rad3-related (ATR) substrates identifies the ubiquitin-proteasome system as a regulator for DNA damage checkpoints. J. Biol. Chem. 2007, 282, 17330–17334. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, H.; Wang, M.; Wang, H.; Böcker, W.; Iliakis, G. Complex H2AX phosphorylation patterns by multiple kinases including ATM and DNA-PK in human cells exposed to ionizing radiation and treated with kinase inhibitors. J. Cell Physiol. 2005, 202, 492–502. [Google Scholar] [CrossRef]
  16. Stiff, T.; O’Driscoll, M.; Rief, N.; Iwabuchi, K.; Löbrich, M.; Jeggo, P.A. ATM and DNA-PK function redundantly to phosphorylate H2AX after exposure to ionizing radiation. Cancer Res. 2004, 64, 2390–2396. [Google Scholar] [CrossRef] [Green Version]
  17. Caron, P.; Choudjaye, J.; Clouaire, T.; Bugler, B.; Daburon, V.; Aguirrebengoa, M.; Mangeat, T.; Iacovoni, J.S.; Álvarez-Quilón, A.; Cortés-Ledesma, F.; et al. Non-redundant Functions of ATM and DNA-PKcs in Response to DNA Double-Strand Breaks. Cell Rep. 2015, 13, 1598–1609. [Google Scholar] [CrossRef] [Green Version]
  18. Kieffer, S.R.; Lowndes, N.F. Immediate-Early, Early, and Late Responses to DNA Double Stranded Breaks. Front. Genet. 2022, 13, 793884. [Google Scholar] [CrossRef]
  19. Kornberg, R.D. Structure of Chromatin. Annu. Rev. Biochem. 1977, 46, 931–954. [Google Scholar] [CrossRef]
  20. Luger, K.; Mäder, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef]
  21. Bannister, A.J.; Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 2011, 21, 381–395. [Google Scholar] [CrossRef] [PubMed]
  22. Smerdon, M.J. DNA repair and the role of chromatin structure. Curr. Opin. Cell Biol. 1991, 3, 422–428. [Google Scholar] [CrossRef]
  23. Soria, G.; Polo, S.E.; Almouzni, G. Prime, repair, restore: The active role of chromatin in the DNA damage response. Mol. Cell 2012, 46, 722–734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dantuma, N.P.; van Attikum, H. Spatiotemporal regulation of posttranslational modifications in the DNA damage response. EMBO J. 2016, 35, 6–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Jackson, S.P.; Durocher, D. Regulation of DNA damage responses by ubiquitin and SUMO. Mol Cell. 2013, 49, 795–807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tang, M.; Li, S.; Chen, J. Ubiquitylation in DNA double-strand break repair. DNA Repair. 2021, 103, 103129. [Google Scholar] [CrossRef]
  27. Uckelmann, M.; Sixma, T.K. Histone ubiquitination in the DNA damage response. DNA Repair. 2017, 56, 92–101. [Google Scholar] [CrossRef]
  28. Aquila, L.; Atanassov, B.S. Regulation of Histone Ubiquitination in Response to DNA Double Strand Breaks. Cells 2020, 9, 1699. [Google Scholar] [CrossRef]
  29. Clague, M.J.; Urbé, S.; Komander, D. Breaking the chains: Deubiquitylating enzyme specificity begets function. Nat. Rev. Mol. Cell Biol. 2019, 20, 338–352. [Google Scholar] [CrossRef]
  30. Fradet-Turcotte, A.; Canny, M.D.; Escribano-Díaz, C.; Orthwein, A.; Leung, C.C.Y.; Huang, H.; Landry, M.-C.; Kitevski-LeBlanc, J.; Noordermeer, S.M.; Sicheri, F.; et al. 53BP1 is a reader of the DNA-damage-induced H2A Lys 15 ubiquitin mark. Nature 2013, 499, 50–54. [Google Scholar] [CrossRef] [Green Version]
  31. Becker, J.R.; Clifford, G.; Bonnet, C.; Groth, A.; Wilson, M.D.; Chapman, J.R. BARD1 reads H2A lysine 15 ubiquitination to direct homologous recombination. Nature 2021, 596, 433–437. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, Q.; Botuyan, M.V.; Zhao, D.; Cui, G.; Mer, E.; Mer, G. Mechanisms of BRCA1–BARD1 nucleosome recognition and ubiquitylation. Nature 2021, 596, 438–443. [Google Scholar] [CrossRef] [PubMed]
  33. Krais, J.J.; Wang, Y.; Patel, P.; Basu, J.; Bernhardy, A.J.; Johnson, N. RNF168-mediated localization of BARD1 recruits the BRCA1-PALB2 complex to DNA damage. Nat. Commun. 2021, 12, 5016. [Google Scholar] [CrossRef] [PubMed]
  34. Witus, S.R.; Burrell, A.L.; Farrell, D.P.; Kang, J.; Wang, M.; Hansen, J.M.; Pravat, A.; Tuttle, L.M.; Stewart, M.D.; Brzovic, P.S.; et al. BRCA1/BARD1 site-specific ubiquitylation of nucleosomal H2A is directed by BARD1. Nat. Struct. Mol. Biol. 2021, 28, 268–277. [Google Scholar] [CrossRef]
  35. Kalb, R.; Mallery, D.L.; Larkin, C.; Huang, J.T.J.; Hiom, K. BRCA1 Is a Histone-H2A-Specific Ubiquitin Ligase. Cell Rep. 2014, 8, 999–1005. [Google Scholar] [CrossRef] [Green Version]
  36. Schwertman, P.; Bekker-Jensen, S.; Mailand, N. Regulation of DNA double-strand break repair by ubiquitin and ubiquitin-like modifiers. Nat. Rev. Mol. Cell Biol. 2016, 17, 379–394. [Google Scholar] [CrossRef]
  37. Simon, J.A.; Kingston, R.E. Occupying chromatin: Polycomb mechanisms for getting to genomic targets, stopping transcriptional traffic, and staying put. Mol. Cell 2013, 49, 808–824. [Google Scholar] [CrossRef] [Green Version]
  38. Mattiroli, F.; Vissers, J.H.A.; van Dijk, W.J.; Ikpa, P.; Citterio, E.; Vermeulen, W.; Marteijn, J.A.; Sixma, T.K. RNF168 ubiquitinates K13-15 on H2A/H2AX to drive DNA damage signaling. Cell 2012, 150, 1182–1195. [Google Scholar] [CrossRef] [Green Version]
  39. Mailand, N.; Bekker-Jensen, S.; Faustrup, H.; Melander, F.; Bartek, J.; Lukas, C.; Lukas, J. RNF8 Ubiquitylates Histones at DNA Double-Strand Breaks and Promotes Assembly of Repair Proteins. Cell 2007, 131, 887–900. [Google Scholar] [CrossRef] [Green Version]
  40. Kolas, N.K.; Chapman, J.R.; Nakada, S.; Ylanko, J.; Chahwan, R.; Sweeney, F.D.; Panier, S.; Mendez, M.; Wildenhain, J.; Thomson, T.M.; et al. Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase. Science 2007, 318, 1637–1640. [Google Scholar] [CrossRef] [Green Version]
  41. Doil, C.; Mailand, N.; Bekker-Jensen, S.; Menard, P.; Larsen, D.H.; Pepperkok, R.; Ellenberg, J.; Panier, S.; Durocher, D.; Bartek, J.; et al. RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins. Cell 2009, 136, 435–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Thorslund, T.; Ripplinger, A.; Hoffmann, S.; Wild, T.; Uckelmann, M.; Villumsen, B.; Narita, T.; Sixma, T.K.; Choudhary, C.; Bekker-Jensen, S.; et al. Histone H1 couples initiation and amplification of ubiquitin signalling after DNA damage. Nature 2015, 527, 389–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Nowsheen, S.; Aziz, K.; Aziz, A.; Deng, M.; Qin, B.; Luo, K.; Jeganathan, K.B.; Zhang, H.; Liu, T.; Yu, J.; et al. L3MBTL2 orchestrates ubiquitin signalling by dictating the sequential recruitment of RNF8 and RNF168 after DNA damage. Nat. Cell Biol. 2018, 20, 455–464. [Google Scholar] [CrossRef] [PubMed]
  44. Hu, Q.; Botuyan, M.V.; Cui, G.; Zhao, D.; Mer, G. Mechanisms of Ubiquitin-Nucleosome Recognition and Regulation of 53BP1 Chromatin Recruitment by RNF168/169 and RAD18. Mol. Cell 2017, 66, 473–487.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Botuyan, M.V.; Lee, J.; Ward, I.M.; Kim, J.-E.; Thompson, J.R.; Chen, J.; Mer, G. Structural basis for the methylation state-specific recognition of histone H4–K20 by 53BP1 and Crb2 in DNA repair. Cell 2006, 127, 1361–1373. [Google Scholar] [CrossRef] [Green Version]
  46. Jacquet, K.; Fradet-Turcotte, A.; Avvakumov, N.; Lambert, J.-P.; Roques, C.; Pandita, R.K.; Paquet, E.; Herst, P.; Gingras, A.-C.; Pandita, T.K.; et al. The TIP60 complex regulates bivalent chromatin recognition by 53BP1 through direct H4K20me binding and H2AK15 acetylation. Mol. Cell 2016, 62, 409–421. [Google Scholar] [CrossRef] [Green Version]
  47. Murr, R.; Loizou, J.I.; Yang, Y.-G.; Cuenin, C.; Li, H.; Wang, Z.-Q.; Herceg, Z. Histone acetylation by Trrap-Tip60 modulates loading of repair proteins and repair of DNA double-strand breaks. Nat. Cell Biol. 2006, 8, 91–99. [Google Scholar] [CrossRef]
  48. Bird, A.W.; Yu, D.Y.; Pray-Grant, M.G.; Qiu, Q.; Harmon, K.E.; Megee, P.C.; Grant, P.A.; Smith, M.M.; Christman, M.F. Acetylation of histone H4 by Esa1 is required for DNA double-strand break repair. Nature 2002, 419, 411–415. [Google Scholar] [CrossRef]
  49. Downs, J.A.; Allard, S.; Jobin-Robitaille, O.; Javaheri, A.; Auger, A.; Bouchard, N.; Kron, S.J.; Jackson, S.P.; Côté, J. Binding of chromatin-modifying activities to phosphorylated histone H2A at DNA damage sites. Mol. Cell 2004, 16, 979–990. [Google Scholar] [CrossRef]
  50. Tang, J.; Cho, N.W.; Cui, G.; Manion, E.M.; Shanbhag, N.M.; Botuyan, M.V.; Mer, G.; Greenberg, R.A. Acetylation limits 53BP1 association with damaged chromatin to promote homologous recombination. Nat. Struct. Mol. Biol. 2013, 20, 317–325. [Google Scholar] [CrossRef]
  51. Dai, L.; Dai, Y.; Han, J.; Huang, Y.; Wang, L.; Huang, J.; Zhou, Z. Structural insight into BRCA1-BARD1 complex recruitment to damaged chromatin. Mol. Cell 2021, 81, 2765–2777.e6. [Google Scholar] [CrossRef] [PubMed]
  52. Nakamura, K.; Saredi, G.; Becker, J.R.; Foster, B.M.; Nguyen, N.V.; Beyer, T.E.; Cesa, L.C.; Faull, P.A.; Lukauskas, S.; Frimurer, T.; et al. H4K20me0 recognition by BRCA1-BARD1 directs homologous recombination to sister chromatids. Nat. Cell Biol. 2019, 21, 311–318. [Google Scholar] [CrossRef]
  53. Shao, G.; Patterson-Fortin, J.; Messick, T.E.; Feng, D.; Shanbhag, N.; Wang, Y.; Greenberg, R.A. MERIT40 controls BRCA1–Rap80 complex integrity and recruitment to DNA double-strand breaks. Genes Dev. 2009, 23, 740–754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hu, Y.; Scully, R.; Sobhian, B.; Xie, A.; Shestakova, E.; Livingston, D.M. RAP80-directed tuning of BRCA1 homologous recombination function at ionizing radiation-induced nuclear foci. Genes Dev. 2011, 25, 685–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Coleman, K.A.; Greenberg, R.A. The BRCA1-RAP80 Complex Regulates DNA Repair Mechanism Utilization by Restricting End Resection. J. Biol. Chem. 2011, 286, 13669–13680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ronau, J.A.; Beckmann, J.F.; Hochstrasser, M. Substrate specificity of the ubiquitin and Ubl proteases. Cell Res. 2016, 26, 441–456. [Google Scholar] [CrossRef] [Green Version]
  57. Sharma, N.; Zhu, Q.; Wani, G.; He, J.; Wang, Q.; Wani, A.A. USP3 counteracts RNF168 via deubiquitinating H2A and γH2AX at lysine 13 and 15. Cell Cycle 2014, 13, 106–114. [Google Scholar] [CrossRef] [Green Version]
  58. Zhang, Z.; Yang, H.; Wang, H. The histone H2A deubiquitinase USP16 interacts with HERC2 and fine-tunes cellular response to DNA damage. J. Biol. Chem. 2014, 289, 32883–32894. [Google Scholar] [CrossRef] [Green Version]
  59. Mosbech, A.; Lukas, C.; Bekker-Jensen, S.; Mailand, N. The Deubiquitylating Enzyme USP44 Counteracts the DNA Double-strand Break Response Mediated by the RNF8 and RNF168 Ubiquitin Ligases*. J. Biol. Chem. 2013, 288, 16579–16587. [Google Scholar] [CrossRef] [Green Version]
  60. Matsui, M.; Kajita, S.; Tsuchiya, Y.; Torii, W.; Tamekuni, S.; Nishi, R. USP49 is a novel deubiquitylating enzyme for γ H2AX in DNA double-strand break repair. Gene 2022, 833, 146599. [Google Scholar] [CrossRef]
  61. Wang, Z.; Zhang, H.; Liu, J.; Cheruiyot, A.; Lee, J.-H.; Ordog, T.; Lou, Z.; You, Z.; Zhang, Z. USP51 deubiquitylates H2AK13,15ub and regulates DNA damage response. Genes Dev. 2016, 30, 946–959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Yang, C.; Zang, W.; Tang, Z.; Ji, Y.; Xu, R.; Yang, Y.; Luo, A.; Hu, B.; Zhang, Z.; Liu, Z.; et al. A20/TNFAIP3 Regulates the DNA Damage Response and Mediates Tumor Cell Resistance to DNA-Damaging Therapy. Cancer Res. 2018, 78, 1069–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Feng, L.; Wang, J.; Chen, J. The Lys63-specific deubiquitinating enzyme BRCC36 is regulated by two scaffold proteins localizing in different subcellular compartments. J. Biol. Chem. 2010, 285, 30982–30988. [Google Scholar] [CrossRef] [Green Version]
  64. Nakada, S.; Tai, I.; Panier, S.; Al-Hakim, A.; Iemura, S.-I.; Juang, Y.-C.; O’Donnell, L.; Kumakubo, A.; Munro, M.; Sicheri, F.; et al. Non-canonical inhibition of DNA damage-dependent ubiquitination by OTUB1. Nature 2010, 466, 941–946. [Google Scholar] [CrossRef] [PubMed]
  65. Lancini, C.; van den Berk, P.C.M.; Vissers, J.H.A.; Gargiulo, G.; Song, J.-Y.; Hulsman, D.; Serresi, M.; Tanger, E.; Blom, M.; Vens, C.; et al. Tight regulation of ubiquitin-mediated DNA damage response by USP3 preserves the functional integrity of hematopoietic stem cells. J. Exp. Med. 2014, 211, 1759–1777. [Google Scholar] [CrossRef] [PubMed]
  66. Ting, X.; Xia, L.; Yang, J.; He, L.; Si, W.; Shang, Y.; Sun, L. USP11 acts as a histone deubiquitinase functioning in chromatin reorganization during DNA repair. Nucleic Acids Res. 2019, 47, 9721–9740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Densham, R.M.; Garvin, A.J.; Stone, H.R.; Strachan, J.; Baldock, R.A.; Daza-Martin, M.; Fletcher, A.; Blair-Reid, S.; Beesley, J.; Johal, B.; et al. Human BRCA1-BARD1 ubiquitin ligase activity counteracts chromatin barriers to DNA resection. Nat. Struct. Mol. Biol. 2016, 23, 647–655. [Google Scholar] [CrossRef] [Green Version]
  68. Kim, B.-J.; Chan, D.W.; Jung, S.Y.; Chen, Y.; Qin, J.; Wang, Y. The Histone Variant MacroH2A1 Is a BRCA1 Ubiquitin Ligase Substrate. Cell Rep. 2017, 19, 1758–1766. [Google Scholar] [CrossRef] [Green Version]
  69. Wang, H.; Wang, L.; Erdjument-Bromage, H.; Vidal, M.; Tempst, P.; Jones, R.S.; Zhang, Y. Role of histone H2A ubiquitination in Polycomb silencing. Nature 2004, 431, 873–878. [Google Scholar] [CrossRef]
  70. Yuan, G.; Ma, B.; Yuan, W.; Zhang, Z.; Chen, P.; Ding, X.; Feng, L.; Shen, X.; Chen, S.; Li, G.; et al. Histone H2A Ubiquitination Inhibits the Enzymatic Activity of H3 Lysine 36 Methyltransferases*. J. Biol. Chem. 2013, 288, 30832–30842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Cao, R.; Tsukada, Y.; Zhang, Y. Role of Bmi-1 and Ring1A in H2A Ubiquitylation and Hox Gene Silencing. Mol. Cell 2005, 20, 845–854. [Google Scholar] [CrossRef] [PubMed]
  72. de Napoles, M.; Mermoud, J.E.; Wakao, R.; Tang, Y.A.; Endoh, M.; Appanah, R.; Nesterova, T.B.; Silva, J.; Otte, A.P.; Vidal, M.; et al. Polycomb group proteins Ring1A/B link ubiquitylation of histone H2A to heritable gene silencing and X inactivation. Dev. Cell. 2004, 7, 663–676. [Google Scholar] [CrossRef] [PubMed]
  73. Wei, J.; Zhai, L.; Xu, J.; Wang, H. Role of Bmi1 in H2A Ubiquitylation and Hox Gene Silencing*. J. Biol. Chem. 2006, 281, 22537–22544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Pan, M.-R.; Peng, G.; Hung, W.-C.; Lin, S.-Y. Monoubiquitination of H2AX protein regulates DNA damage response signaling. J. Biol. Chem. 2011, 286, 28599–28607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ginjala, V.; Nacerddine, K.; Kulkarni, A.; Oza, J.; Hill, S.J.; Yao, M.; Citterio, E.; van Lohuizen, M.; Ganesan, S. BMI1 is recruited to DNA breaks and contributes to DNA damage-induced H2A ubiquitination and repair. Mol. Cell Biol. 2011, 31, 1972–1982. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Fitieh, A.; Locke, A.J.; Mashayekhi, F.; Khaliqdina, F.; Sharma, A.K.; Ismail, I.H. BMI-1 regulates DNA end resection and homologous recombination repair. Cell Rep. 2022, 38, 110536. [Google Scholar] [CrossRef]
  77. Scheuermann, J.C.; Alonso, A.G.D.; Oktaba, K.; Ly-Hartig, N.; McGinty, R.K.; Fraterman, S.; Wilm, M.; Muir, T.W.; Müller, J. Histone H2A deubiquitinase activity of the Polycomb repressive complex PR-DUB. Nature 2010, 465, 243–247. [Google Scholar] [CrossRef] [Green Version]
  78. Ismail, I.H.; Davidson, R.; Gagné, J.-P.; Xu, Z.Z.; Poirier, G.G.; Hendzel, M.J. Germline mutations in BAP1 impair its function in DNA double-strand break repair. Cancer Res. 2014, 74, 4282–4294. [Google Scholar] [CrossRef] [Green Version]
  79. Shanbhag, N.M.; Rafalska-Metcalf, I.U.; Balane-Bolivar, C.; Janicki, S.M.; Greenberg, R.A. ATM-dependent chromatin changes silence transcription in cis to DNA double-strand breaks. Cell 2010, 141, 970–981. [Google Scholar] [CrossRef] [Green Version]
  80. Nakamura, K.; Kato, A.; Kobayashi, J.; Yanagihara, H.; Sakamoto, S.; Oliveira, D.V.N.P.; Shimada, M.; Tauchi, H.; Suzuki, H.; Tashiro, S.; et al. Regulation of Homologous Recombination by RNF20-Dependent H2B Ubiquitination. Mol. Cell 2011, 41, 515–528. [Google Scholar] [CrossRef]
  81. Moyal, L.; Lerenthal, Y.; Gana-Weisz, M.; Mass, G.; So, S.; Wang, S.-Y.; Eppink, B.; Chung, Y.M.; Shalev, G.; Shema, E.; et al. Requirement of ATM-Dependent Monoubiquitylation of Histone H2B for Timely Repair of DNA Double-Strand Breaks. Mol. Cell 2011, 41, 529–542. [Google Scholar] [CrossRef] [PubMed]
  82. So, C.C.; Ramachandran, S.; Martin, A. E3 Ubiquitin Ligases RNF20 and RNF40 Are Required for Double-Stranded Break (DSB) Repair: Evidence for Monoubiquitination of Histone H2B Lysine 120 as a Novel Axis of DSB Signaling and Repair. Mol. Cell. Biol. 2019, 39, e00488-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Zhao, Y.; Lang, G.; Ito, S.; Bonnet, J.; Metzger, E.; Sawatsubashi, S.; Suzuki, E.; le Guezennec, X.; Stunnenberg, H.G.; Krasnov, A.; et al. A TFTC/STAGA module mediates histone H2A and H2B deubiquitination, coactivates nuclear receptors, and counteracts heterochromatin silencing. Mol. Cell 2008, 29, 92–101. [Google Scholar] [CrossRef]
  84. Lang, G.; Bonnet, J.; Umlauf, D.; Karmodiya, K.; Koffler, J.; Stierle, M.; Devys, D.; Tora, L. The Tightly Controlled Deubiquitination Activity of the Human SAGA Complex Differentially Modifies Distinct Gene Regulatory Elements. Mol. Cell Biol. 2011, 31, 3734–3744. [Google Scholar] [CrossRef] [Green Version]
  85. Henry, K.W.; Wyce, A.; Lo, W.-S.; Duggan, L.J.; Emre, N.C.T.; Kao, C.-F.; Pillus, L.; Shilatifard, A.; Osley, M.A.; Berger, S.L. Transcriptional activation via sequential histone H2B ubiquitylation and deubiquitylation, mediated by SAGA-associated Ubp8. Genes Dev. 2003, 17, 2648–2663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Bonnet, J.; Wang, C.-Y.; Baptista, T.; Vincent, S.D.; Hsiao, W.-C.; Stierle, M.; Kao, C.-F.; Tora, L.; Devys, D. The SAGA coactivator complex acts on the whole transcribed genome and is required for RNA polymerase II transcription. Genes Dev. 2014, 28, 1999–2012. [Google Scholar] [CrossRef] [Green Version]
  87. Li, C.; Irrazabal, T.; So, C.C.; Berru, M.; Du, L.; Lam, E.; Ling, A.K.; Gommerman, J.L.; Pan-Hammarström, Q.; Martin, A. The H2B deubiquitinase Usp22 promotes antibody class switch recombination by facilitating non-homologous end joining. Nat. Commun. 2018, 9, 1006. [Google Scholar] [CrossRef] [PubMed]
  88. Ishiyama, S.; Nishiyama, A.; Saeki, Y.; Moritsugu, K.; Morimoto, D.; Yamaguchi, L.; Arai, N.; Matsumura, R.; Kawakami, T.; Mishima, Y.; et al. Structure of the Dnmt1 Reader Module Complexed with a Unique Two-Mono-Ubiquitin Mark on Histone H3 Reveals the Basis for DNA Methylation Maintenance. Mol. Cell 2017, 68, 350–360.e7. [Google Scholar] [CrossRef] [Green Version]
  89. Qin, W.; Wolf, P.; Liu, N.; Link, S.; Smets, M.; Mastra, F.L.; Forné, I.; Pichler, G.; Hörl, D.; Fellinger, K.; et al. DNA methylation requires a DNMT1 ubiquitin interacting motif (UIM) and histone ubiquitination. Cell Res. 2015, 25, 911–929. [Google Scholar] [CrossRef] [Green Version]
  90. Li, T.; Wang, L.; Du, Y.; Xie, S.; Yang, X.; Lian, F.; Zhou, Z.; Qian, C. Structural and mechanistic insights into UHRF1-mediated DNMT1 activation in the maintenance DNA methylation. Nucleic Acids Res. 2018, 46, 3218–3231. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, H.; Zhai, L.; Xu, J.; Joo, H.-Y.; Jackson, S.; Erdjument-Bromage, H.; Tempst, P.; Xiong, Y.; Zhang, Y. Histone H3 and H4 Ubiquitylation by the CUL4-DDB-ROC1 Ubiquitin Ligase Facilitates Cellular Response to DNA Damage. Mol. Cell 2006, 22, 383–394. [Google Scholar] [CrossRef] [PubMed]
  92. Yan, Q.; Dutt, S.; Xu, R.; Graves, K.; Juszczynski, P.; Manis, J.P.; Shipp, M.A. BBAP monoubiquitylates histone H4 at lysine 91 and selectively modulates the DNA damage response. Mol. Cell 2009, 36, 110–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.J.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef] [PubMed]
  94. Roy, R.; Chun, J.; Powell, S.N. BRCA1 and BRCA2: Different roles in a common pathway of genome protection. Nat. Rev. Cancer 2011, 12, 68–78. [Google Scholar] [CrossRef] [Green Version]
  95. Chapman, J.R.; Sossick, A.J.; Boulton, S.J.; Jackson, S.P. BRCA1-associated exclusion of 53BP1 from DNA damage sites underlies temporal control of DNA repair. J. Cell Sci. 2012, 125, 3529–3534. [Google Scholar] [CrossRef] [Green Version]
  96. Isono, M.; Niimi, A.; Oike, T.; Hagiwara, Y.; Sato, H.; Sekine, R.; Yoshida, Y.; Isobe, S.-Y.; Obuse, C.; Nishi, R.; et al. BRCA1 Directs the Repair Pathway to Homologous Recombination by Promoting 53BP1 Dephosphorylation. Cell Rep. 2017, 18, 520–532. [Google Scholar] [CrossRef] [Green Version]
  97. Savage, K.I.; Harkin, D.P. BRCA1, a ‘complex’ protein involved in the maintenance of genomic stability. FEBS J. 2015, 282, 630–646. [Google Scholar] [CrossRef]
  98. Wu, J.; Huen, M.S.Y.; Lu, L.-Y.; Ye, L.; Dou, Y.; Ljungman, M.; Chen, J.; Yu, X. Histone ubiquitination associates with BRCA1-dependent DNA damage response. Mol. Cell Biol. 2009, 29, 849–860. [Google Scholar] [CrossRef] [Green Version]
  99. Wu, W.; Nishikawa, H.; Fukuda, T.; Vittal, V.; Asano, M.; Miyoshi, Y.; Klevit, R.E.; Ohta, T. Interaction of BARD1 and HP1 is required for BRCA1 retention at sites of DNA damage. Cancer Res 2015, 75, 1311–1321. [Google Scholar] [CrossRef] [Green Version]
  100. Boersma, V.; Moatti, N.; Segura-Bayona, S.; Peuscher, M.H.; van der Torre, J.; Wevers, B.A.; Orthwein, A.; Durocher, D.; Jacobs, J.J.L. MAD2L2 controls DNA repair at telomeres and DNA breaks by inhibiting 5’ end resection. Nature 2015, 521, 537–540. [Google Scholar] [CrossRef] [Green Version]
  101. Kakarougkas, A.; Ismail, A.; Katsuki, Y.; Freire, R.; Shibata, A.; Jeggo, P.A. Co-operation of BRCA1 and POH1 relieves the barriers posed by 53BP1 and RAP80 to resection. Nucleic Acids Res. 2013, 41, 10298–10311. [Google Scholar] [CrossRef] [PubMed]
  102. Huen, M.S.Y.; Sy, S.M.H.; Chen, J. BRCA1 and its toolbox for the maintenance of genome integrity. Nat. Rev. Mol. Cell Biol. 2010, 11, 138–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Yazinski, S.A.; Comaills, V.; Buisson, R.; Genois, M.-M.; Nguyen, H.D.; Ho, C.K.; Kwan, T.T.; Morris, R.; Lauffer, S.; Nussenzweig, A.; et al. ATR inhibition disrupts rewired homologous recombination and fork protection pathways in PARP inhibitor-resistant BRCA-deficient cancer cells. Genes Dev. 2017, 31, 318–332. [Google Scholar] [CrossRef] [PubMed]
  104. Luijsterburg, M.S.; Typas, D.; Caron, M.C.; Wiegant, W.W.; van den Heuvel, D.; Boonen, R.A.; Couturier, A.M.; Mullenders, L.H.; Masson, J.Y.; van Attikum, H. A PALB2-interacting domain in RNF168 couples homologous recombination to DNA break-induced chromatin ubiquitylation. eLife 2017, 6, e20922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zong, D.; Adam, S.; Wang, Y.; Sasanuma, H.; Callén, E.; Murga, M.; Day, A.; Kruhlak, M.J.; Wong, N.; Munro, M.; et al. BRCA1 Haploinsufficiency Is Masked by RNF168-Mediated Chromatin Ubiquitylation. Mol. Cell 2019, 73, 1267–1281.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Nakada, S.; Yonamine, R.M.; Matsuo, K. RNF8 regulates assembly of RAD51 at DNA double-strand breaks in the absence of BRCA1 and 53BP1. Cancer Res. 2012, 72, 4974–4983. [Google Scholar] [CrossRef] [Green Version]
  107. Uckelmann, M.; Densham, R.M.; Baas, R.; Winterwerp, H.H.K.; Fish, A.; Sixma, T.K.; Morris, J.R. USP48 restrains resection by site-specific cleavage of the BRCA1 ubiquitin mark from H2A. Nat. Commun. 2018, 9, 229. [Google Scholar] [CrossRef] [Green Version]
  108. Lee, H.-G.; Kahn, T.G.; Simcox, A.; Schwartz, Y.B.; Pirrotta, V. Genome-wide activities of Polycomb complexes control pervasive transcription. Genome Res. 2015, 25, 1170–1181. [Google Scholar] [CrossRef] [Green Version]
  109. Tamburri, S.; Lavarone, E.; Fernández-Pérez, D.; Conway, E.; Zanotti, M.; Manganaro, D.; Pasini, D. Histone H2AK119 Mono-Ubiquitination Is Essential for Polycomb-Mediated Transcriptional Repression. Mol. Cell 2020, 77, 840–856.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Schwartz, Y.B.; Pirrotta, V. A new world of Polycombs: Unexpected partnerships and emerging functions. Nat. Rev. Genet. 2013, 14, 853–864. [Google Scholar] [CrossRef]
  111. Campbell, S.; Ismail, I.H.; Young, L.C.; Poirier, G.G.; Hendzel, M.J. Polycomb repressive complex 2 contributes to DNA double-strand break repair. Cell Cycle 2013, 12, 2675–2683. [Google Scholar] [CrossRef]
  112. Ismail, I.H.; Andrin, C.; McDonald, D.; Hendzel, M.J. BMI1-mediated histone ubiquitylation promotes DNA double-strand break repair. J. Cell Biol. 2010, 191, 45–60. [Google Scholar] [CrossRef] [PubMed]
  113. Hong, Z.; Jiang, J.; Lan, L.; Nakajima, S.; Kanno, S.; Koseki, H.; Yasui, A. A polycomb group protein, PHF1, is involved in the response to DNA double-strand breaks in human cell. Nucleic Acids Res. 2008, 36, 2939–2947. [Google Scholar] [CrossRef] [Green Version]
  114. Izhar, L.; Adamson, B.; Ciccia, A.; Lewis, J.; Pontano-Vaites, L.; Leng, Y.; Liang, A.C.; Westbrook, T.F.; Harper, J.W.; Elledge, S.J. A Systematic Analysis of Factors Localized to Damaged Chromatin Reveals PARP-Dependent Recruitment of Transcription Factors. Cell Rep. 2015, 11, 1486–1500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Kakarougkas, A.; Ismail, A.; Chambers, A.L.; Riballo, E.; Herbert, A.D.; Künzel, J.; Löbrich, M.; Jeggo, P.A.; Downs, J.A. Requirement for PBAF in transcriptional repression and repair at DNA breaks in actively transcribed regions of chromatin. Mol. Cell 2014, 55, 723–732. [Google Scholar] [CrossRef] [Green Version]
  116. Sanchez, A.; de Vivo, A.; Uprety, N.; Kim, J.; Stevens, S.M.; Kee, Y. BMI1–UBR5 axis regulates transcriptional repression at damaged chromatin. Proc. Natl. Acad. Sci. USA 2016, 113, 11243–11248. [Google Scholar] [CrossRef] [Green Version]
  117. Minsky, N.; Shema, E.; Field, Y.; Schuster, M.; Segal, E.; Oren, M. Monoubiquitinated H2B is associated with the transcribed region of highly expressed genes in human cells. Nat. Cell Biol. 2008, 10, 483–488. [Google Scholar] [CrossRef] [PubMed]
  118. Shiloh, Y.; Shema, E.; Moyal, L.; Oren, M. RNF20-RNF40: A ubiquitin-driven link between gene expression and the DNA damage response. FEBS Lett. 2011, 585, 2795–2802. [Google Scholar] [CrossRef] [Green Version]
  119. Guervilly, J.-H.; Takedachi, A.; Naim, V.; Scaglione, S.; Chawhan, C.; Lovera, Y.; Despras, E.; Kuraoka, I.; Kannouche, P.; Rosselli, F.; et al. The SLX4 complex is a SUMO E3 ligase that impacts on replication stress outcome and genome stability. Mol. Cell 2015, 57, 123–137. [Google Scholar] [CrossRef] [Green Version]
  120. Muñoz, I.M.; Hain, K.; Déclais, A.-C.; Gardiner, M.; Toh, G.W.; Sanchez-Pulido, L.; Heuckmann, J.M.; Toth, R.; Macartney, T.; Eppink, B.; et al. Coordination of structure-specific nucleases by human SLX4/BTBD12 is required for DNA repair. Mol. Cell 2009, 35, 116–127. [Google Scholar] [CrossRef] [PubMed]
  121. Matsushita, N.; Kitao, H.; Ishiai, M.; Nagashima, N.; Hirano, S.; Okawa, K.; Ohta, T.; Yu, D.S.; McHugh, P.J.; Hickson, I.D.; et al. A FancD2-monoubiquitin fusion reveals hidden functions of Fanconi anemia core complex in DNA repair. Mol. Cell 2005, 19, 841–847. [Google Scholar] [CrossRef] [PubMed]
  122. Ouyang, J.; Garner, E.; Hallet, A.; Nguyen, H.D.; Rickman, K.A.; Gill, G.; Smogorzewska, A.; Zou, L. Noncovalent interactions with SUMO and ubiquitin orchestrate distinct functions of the SLX4 complex in genome maintenance. Mol. Cell 2015, 57, 108–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Tian, Y.; Paramasivam, M.; Ghosal, G.; Chen, D.; Shen, X.; Huang, Y.; Akhter, S.; Legerski, R.; Chen, J.; Seidman, M.M.; et al. UHRF1 contributes to DNA damage repair as a lesion recognition factor and nuclease scaffold. Cell Rep. 2015, 10, 1957–1966. [Google Scholar] [CrossRef] [Green Version]
  124. Sparmann, A.; van Lohuizen, M. Polycomb silencers control cell fate, development and cancer. Nat. Rev. Cancer 2006, 6, 846–856. [Google Scholar] [CrossRef]
Figure 1. Summary of histone lysine ubiquitination and deubiquitination enzymes involved in DNA damage response. The E3 enzymes are listed above the ubiquitination site, whereas the deubiquitinating enzymes are listed below.
Figure 1. Summary of histone lysine ubiquitination and deubiquitination enzymes involved in DNA damage response. The E3 enzymes are listed above the ubiquitination site, whereas the deubiquitinating enzymes are listed below.
Ijms 23 08187 g001
Figure 2. Model of H2AK13/15 ubiquitination and reader proteins in the DNA damage response. RNF168-induced H2AK13/15 ubiquitination leads to the activation of the NHEJ (53BP1) or HR (BRCA1/BARD1) pathway. RNF8-RNF168 induced K63-linked chains to recruit the BRCA1-A complex (RAP80/BRCA1) and restrict the HR pathway.
Figure 2. Model of H2AK13/15 ubiquitination and reader proteins in the DNA damage response. RNF168-induced H2AK13/15 ubiquitination leads to the activation of the NHEJ (53BP1) or HR (BRCA1/BARD1) pathway. RNF8-RNF168 induced K63-linked chains to recruit the BRCA1-A complex (RAP80/BRCA1) and restrict the HR pathway.
Ijms 23 08187 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sekiguchi, M.; Matsushita, N. DNA Damage Response Regulation by Histone Ubiquitination. Int. J. Mol. Sci. 2022, 23, 8187. https://doi.org/10.3390/ijms23158187

AMA Style

Sekiguchi M, Matsushita N. DNA Damage Response Regulation by Histone Ubiquitination. International Journal of Molecular Sciences. 2022; 23(15):8187. https://doi.org/10.3390/ijms23158187

Chicago/Turabian Style

Sekiguchi, Miyu, and Nobuko Matsushita. 2022. "DNA Damage Response Regulation by Histone Ubiquitination" International Journal of Molecular Sciences 23, no. 15: 8187. https://doi.org/10.3390/ijms23158187

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

Article Metrics

Back to TopTop