Open access peer-reviewed chapter

The Five Families of DNA Repair Proteins and their Functionally Relevant Ubiquitination

Written By

Niko Moses and Xiaohong Mary Zhang

Submitted: 21 November 2016 Reviewed: 06 October 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.71537

From the Edited Volume

Ubiquitination Governing DNA Repair - Implications in Health and Disease

Edited by Effrossyni Boutou and Horst-Werner Stürzbecher

Chapter metrics overview

1,108 Chapter Downloads

View Full Metrics

Abstract

The process of DNA repair, be it a response to replication dysfunction or genotoxic insult, is critical for the resolution of strand errors and the avoidance of DNA mismatches that could result in various molecular pathologies, including carcinogenic development. Here, we will describe the five main mechanisms by which DNA avoids mutation, namely the processes of base excision repair, mismatch repair, nucleotide excision repair, homologous recombination, and nonhomologous end joining. In particular, we will dissect the functional significance of various posttranslational modifications of the essential proteins within these pathways, including but not limited to ubiquitination, acetylation, and phosphorylation.

Keywords

  • base excision repair (BER)
  • mismatch repair (MMR)
  • nucleotide excision repair (NER)
  • homologous recombination (HR)
  • nonhomologous end joining (NHEJ)
  • posttranslational modification

1. Introduction

The mammalian genome is under constant barrage by exogenous and endogenous insult that can beget damage and instability. Exogenous insults include exposure to UV radiation and chemical carcinogens found in the environment, while endogenous factors include ROS produced by cellular metabolism, spontaneous chemical reactions like base deamination and mistakes made during the replicative process. It is critical to the survival of the organism that each cell have the ability to resolve the damage induced by this wide variety of insults, and that the machinery responsible for responding to damage must be equally diverse.

There are five main mechanisms responsible for repairing damaged DNA, and their conservation from bacteria all the way to humans exemplifies their critical role in the maintenance of an organism’s genome. These mechanisms consist of base excision repair, nucleotide excision repair, mismatch repair, homologous recombination, and nonhomologous end joining.

Advertisement

2. Ubiquitination and the proteasome degradation pathway

The ubiquitin proteasome pathway (UPP) is a mechanism used for the maintenance of proper levels of cellular proteins and the destruction of old or misfolded proteins by targeting them for degradation. This targeting comes in the form of ubiquitination, the process of covalently linking a polyubiquitin chain to the protein that is recognized and bound by the 26S proteasome, which degrades the protein and releases the ubiquitin. Ubiquitin is a highly conserved 76-amino acid protein that serves as the subunit of the polyubiquitin chain. Ubiquitin is covalently linked to its target in a three-step cascade conducted by a ubiquitin-activating enzyme (E1), a ubiquitin-conjugating enzyme (E2), and a ubiquitin-ligating enzyme (E3) [1]. Apart from protein degradation, ubiquitination can also mediate protein-protein interaction.

Advertisement

3. Detection and repair within the DNA strand

3.1. Base excision repair

Base excision repair enzymes are responsible for correcting lesions induced by a wide variety of both endogenous and exogenous insults, including sites of base loss, nonbulky base lesions, and DNA single-strand breaks (SSBs) [2]. DNA glycosylases are responsible for the first step of base excision repair (BER) by initially detecting the damage and excising the base via hydrolyzing the N-glycosylic bond linking the DNA base to the sugar phosphate backbone. This process generates an abasic site (AP site) that AP endonuclease 1 recognizes and acts upon by cleaving the phosphodiester bond 5′ to the AP site, leaving a SSB with a 5′-sugar phosphate. A DNA repair complex composed of DNA pol β, XRCC1, and DNA ligase IIIα can recognize this SSB and remove the 5′-sugar phosphate through its AP lyase activity, and add a single nucleotide to the 3′-end through its DNA polymerase activity. The damage is finally resolved when Lig 3 seals the DNA ends together, thus completing what is referred to as short patch BER, the process by which human cells conduct the majority of their BER [3, 4].

At the moment, much of the work focusing on ubiquitination of the proteins involved in BER has been pursued by Dianov et al. [5]. This group has been able to demonstrate that, under normal conditions, BER components are targeted for destruction by the E3 ubiquitin ligase CHIP. When DNA damage occurs in cells, the BER components undergo stabilization to increase their ability to correct the damage. Specifically, pol β, XRCC1, and DNA ligase III are polyubiquitinated by CHIP and Mule when not bound to chromatin, and thus targeted for degradation [6].

DNA pol λ can also be targeted by posttranslational modifications. Pol λ contains four distinct phosphorylation sites, but phosphorylation of the Thr553 has the strongest impact on the stability of the protein. Pol λ can be phosphorylated on all of these sites by the Cdk2/cyclin A complex, but its levels of phosphorylation are reduced when it is interacting with proliferating cell nuclear antigen (PCNA) [7], a sliding clamp that associates with DNA polymerases and ensures accurate and possessive DNA synthesis [8]. Increased phosphorylation of Thr553 on pol λ positively correlates with its protein levels in the cell, likely due to the fact that phosphorylation at this site protects pol λ from ubiquitination and degradation. This stabilization occurs in the late S and G2 phase. Pol λ is closely related to pol β, and indeed both polymerases can be ubiquitinated by CHIP and Mule. It is thought that pol λ is needed in late S and G2, specifically whenever oxidative DNA damage presents at this phase and induces 8-oxo-G lesions [9, 10]. When these lesions occur, it is Mule that is responsible for regulating protein levels of pol λ. When Mule is able to ubiquitinate pol λ, this action targets both pol λ for degradation and decreases its enzymatic activity. Mule is responsible for the monoubiquitination of pol β, which can be further polyubiquitinated by CHIP and targeted for degradation [11]. While pol β is continuously expressed in unstressed cells, it is almost immediately targeted by Mule and CHIP in the absence of damage. Upon DNA damage detection, alternative reading frame (ARF) begins to accumulate and eventually inhibits Mule activity [12], allowing for pol β accumulation and activation of further BER proteins. Once the lesion(s) have been resolved ARF levels drop, Mule activity is restored, and pol β will once again be ubiquitinated and degraded. ARF is a BER protein frequently mutated in cancer cells; it functions by responding to DNA damage by directly inhibiting Mule, as well as regulating p53. The amount of ARF produce in response to DNA damage is dependent on the extent of the damage; and by inhibiting Mule activity, it allows p53 to halt replication while pol β complexes conduct repair [13]. Without ARF, both Mule and Mdm2 repress p53 activity. ARF is a 482 kDa protein belonging to the homologous to E6-AP carboxyl terminus (HECT) family of E3 ubiquitin ligases [14], named such due to their ubiquitous presence of a C-terminal HECT domain of ~350 amino acids that house their E3 catalytic activity. The HECT domain of Mule contains two sub-domains connected by a flexible linker allowing these domains to undergo ubiquitin chain transfer [15]. Aside from allowing for p53 accumulation and activation of the DNA damage response in damaged cells, ARF also plays a p53-independent role in tumor suppression due to its ability to induce proliferation delay in cells lacking functional p53 and p21 [16, 17].

3.2. Nucleotide excision repair

Nucleotide excision repair is a process undertaken in both prokaryotes and eukaryotes to enzymatically remove bulky, helix-distorting base adducts from DNA. This process is the predominant method of DNA repair in mammals, especially when resolving damage induced by ultraviolet light from the sun. About 30 proteins are involved in eukaryotic nucleotide excision repair (NER), including nine major proteins identified by their mutation in humans and the development of UV-hypersensitivity as a result. Seven of these proteins, when mutated, lead to the development of Xeroderma pigmentosum syndrome (XPA to XPG) and two lead to the development of Cockayne’s syndrome (CSA and CSB) [18]. Additional players in the process of NER include excision repair cross-complementing 1 (ERCC1), replication protein A (RPA), and Rad23 homologs (HR23A and HR23B) [19, 20]. The Rad23 homologs share redundancy with the function of Rad23 in yeast during the recognition of the lesion in NER. Upon initial recognition of a lesion, eukaryotic NER can continue by either the process of global genome NER (GG-NER) or transcription-coupled NER (TC-NER). GG-NER removes DNA from untranscribed regions of DNA; XPC-HR23B and UV-DDB (damaged DNA-binding protein) can recognize UV damage and recruit XPA to this resultant lesion [21]. In TC-NER, RNA polymerase II recognizes the lesion when it is a mediating transcription but finds its progress blocked by the break. This stalling of RNA pol II is recognized by CSA and CSB, which will localize to the lesion and load XPA on the site to initiate NER. After initial lesion recognition, GG-NER and TC-NER follow the same pathway to resolve the damage. XPA further recruits XPB (5′ to 3′ helicase) and XPD (3′ to 5′ helicase) to unwind the DNA at the damage site and allow for incisions on the 3′ and 5′ sides of the gap to be made by XPG and XPF-ERCC1 endonucleases, respectively [22, 23, 24].

The first association between the ubiquitin proteasome pathway (UPP) and nucleotide excision repair (NER) was due to the identification of the ubiquitin-like domain present at the N-terminus of Rad23 [25, 26] that can serve as a ubiquitin receptor, similar to the subunit of the 26S proteasome Rpn1 [27]. Both can recognize polyubiquitinated chains and transport the target proteins to the proteasome [28]. The Rad23 ubiquitin-like domain is required for sufficient NER activity, and deletion of this domain can result in UV radiation sensitivity [29]. Russell et al. demonstrated that complete inhibition of the proteasome does not affect NER, while specifically targeting 19S activity does. Further, 19S influence on NER is mediated by the Ubl domain of Rad23, suggesting to them that 19S may be acting as a molecular chaperone in the context of NER by altering the conformation of certain NER proteins [29, 30]. Rad23 avoids proteasomal degradation due to its ubiquitin-like domain via a C-terminal ubiquitin-associated (UBA) domain [31] and can impart this protection on its binding partner XPC in a mechanism that will be detailed later.

XPE-deficient cells lack the ability for UV-damaged DNA-binding component (DDB), composed of DDB1 (p127) and DDB2 (p48), to bind DNA. DDB2 and CSA are present in separate but nearly identical molecular complexes, both associated by their interaction with DDB1 [32]. Both complexes contain CUL4A and ROC1, both ubiquitin ligase subunits, as well as the constitutive photomorphogenesis 9 (COP9) signalosome (CSN). When the complex is devoid of CSN, they are able to display robust ubiquitin ligase activity. After UV exposure, CSN rapidly dissociates from the DDB2 complex and CUL4A is modified by NEDD8 (via neddylation and polyubiquitination) [33], leading to ubiquitin ligase activity from the complex. This complex ubiquitinates XPC (which is bound to HR23, the specifics of this complex detailed later), allowing both of these complexes to bind the damaged DNA. DDB2 itself is also polyubiquitinated, causing it to dissociate from the complex, and get degraded by the proteasome. Ubiquitinated XPC and HR23 remain on the DNA, where they activate the process of NER. The CSA complex is not as well characterized as the DDB2 complex. What is known is that unlike the DDB2 complex, UV-induced damage stimulates the rapid association of CSN with CSA, suppressing all ubiquitin ligase activity from the complex. A target of DDB2 complex ubiquitination is XPC, which is required for GG-NER at the damage site [34]. In undamaged cells, XPC exists in a heterotrimeric complex with either mammalian homolog of Rad23, HR23A, or HR23B. XPC is normally bound to HR23B, but in its absence HR23A is sufficient [35, 36]. This XPC complex recognizes physical aberrations in the structure of DNA rather than the lesions themselves, and is recruited after ubiquitination by the DDB2 complex. The ubiquitination appears to be protective, as XPC is not a target of proteosomal degradation and is further stabilized through its interaction with the ubiquitin-associated (UBA) domain of the HR23 protein it is bound to [37, 38]. It is of great interest that both DDB2 and XPC are ubiquitinated and this ubiquitination yields drastically different outcomes, yet there are still parts of this mechanism that have not been defined. The specific ubiquitination sites on these two proteins have not been mapped, and the factors that specifically interact with these two proteins upon ubiquitination have yet to be defined.

Returning to the CSA complex, after UV exposure, it rapidly associates with CSN and its ubiquitin ligase activity is suppressed. This action has implications on the function of RNA polymerase II, which stalls on DNA strands during transcription when it encounters a break or adduct (any transcriptional blockade) and signals for the assembly of TC-NER machinery. Reports have indicated that UV exposure can activate CSA- and CSB-dependent polyubiquitination of RNA pol II [39], an observation that contrasts with the previously discussed reports of CSN inhibiting the CSA complex’s ubiquitin ligase activity. Groisman et al. have suggested that CSN can differentially regulate the activity of DDB2 and CSA complexes, and that its interaction with CSA may not in fact be inhibitory [32]. It is also possible that there is an additional member of the CSA complex, or a separate complex is mediating ubiquitination of RNA pol II. Svejstrup et al. have argued that RNA pol II ubiquitination is conducted by a Rad26-Def complex [40]. Def1 is a protein discovered in yeast that complexes with Rad26 on chromatin, and when this protein is deleted in yeast, these cells are unable to degrade stalled DNA pol II in response to DNA lesions [40]. RNA pol II stalling has been reported to induce ubiquitination and degradation of Rpb1, the largest RNA pol II subunit, in a Def1-dependent manner [41]. When RNA pol II is polyubiquitinated after UV-induced damage (an additional E3 ligase is BRCA1/BARD1 of the homologous recombination pathway), it is either degraded or bypasses the transcriptional block, allowing mRNA synthesis to continue [42] and the damage is to be resolved later by GG-NER [43].

UV radiation has often been used to elucidate the mechanisms of NER components, as helix-distorting damage (cyclobutane pyrimidine dimers, 6-4 photoproducts) is repaired by NER [44]. These studies have also revealed the posttranslational modifications necessary for the functional relevance of these proteins. UV radiation experiments were responsible for the initial observation that genes encoding certain components of the UPP influenced the ability of cells to survive after being irradiated, and the researchers interpreted this data in a manner that highlights the proteolytic activity for the proteasome in NER [45, 46, 47]. After these initial observations, Rad23 was investigated and determined not to be targeted for ubiquitination, and Rad4 (yeast homolog of XPC) became the next potential target for ubiquitination. This focus was based on the observation that Rad4 overexpression can increase NER activity [47]. Further studies in human cells revealed that XPC also accumulated after DNA damage, and like their yeast counterparts, increased NER activity [37]. This accumulation was correlated to hHR23 in mouse cells, and it was found that Rad23 could use its UBA domains to stabilize Rad4/XPC by acting in trans [48] as well as controlling its own turnover by acting in cis [49]. The C-terminal tail of H2A is a target for posttranslational modification, with as much as 5–15% of H2A being monoubiquitinated in mammals [50]. Ubiquitinated H2A is associated with condensed DNA and gene repression, and Ring2 is the predominant E3 ubiquitin ligase responsible for this modification [51, 52]. UV-induced DNA damage can induce monoubiquitination of H2A in close proximity to the lesions [53] in a manner very similar to its phosphorylation. Both of these histone modifications occur in UV treated, non-S-phase cells and are dependent on functional NER, and ATR signaling is required for the tail modification to occur [54, 55]. Ubiquitination of XPC, DDB2, and PCNA can still occur in NER-deficient XP-A cell lines, but H2A ubiquitination relies upon NER-sufficiency. Ubc13 and RNF8 are responsible for perpetuating sustained H2A ubiquitination so that NER can occur, but do not initially ubiquitinate H2A [56].

3.3. Mismatch repair

The DNA mismatch repair pathway is responsible for correcting mispaired nucleotides and insertion/deletion loops (IDLs) that are a consequence of replication, recombination, and repair errors [57].

The role of ubiquitination in the process of mismatch repair is relatively uncharacterized, compared with the rest of the DNA repair pathways detailed in this chapter. However, research conducted in our lab has identified that the stability of MutS protein homolog 2 (MSH2), an essential DNA mismatch repair protein, is regulated through ubiquitination by histone deacetylase 6 [58]. Ubiquitination of MutSα was first reported by Lautier et al. [59], although the enzyme responsible remained undetermined until our 2014 publication [58]. MSH2 forms two heterodimers, MSH2-MSH6 (MutSα) and MSH2-MSH3 (MutSβ). MutSα recognizes single base mismatches and 1-2 nucleotide insertions and deletions [60] while MutSβ recognizes bulky DNA adducts and larger insertions and deletions [61]. MutSα specifically recognizes DNA lesions induced by a wide variety of DNA-damaging agents (6-thioguanine, cisplatin, doxorubicin, etoposide) [62]. In the absence of MutSα, cells display resistance to these DNA-damaging agents and do not undergo apoptosis as a result of a futile repair cycle [63, 64]. Elucidation of the mechanism of MSH2 stability in cells is critical to the field of mismatch repair, as the initiation of MMR is controlled by the binding of MutSα and MutSβ to the mispair. These proteins subsequently signal the downstream effectors of MMR; MutLα (MLH1-PMS2), PCNA, and RPA, which can further lead to the recruitment of excision protein exonuclease 1 (EXO1). EXO1 excises the mismatched base, forming a gap that is filled by polymerase δ and a nick that is resolved by DNA ligase 1. When MSH2 is acetylated, it cannot be ubiquitinated, and thus is retained and is able to form MutSα and Mutsβ complexes. MSH2 turnover can be induced by HDAC6 activity, which subsequently deacetylates and ubiquinates MSH2 to target it for proteosomal degradation. This action is possible because of the E3 ubiquitin ligase activity HDAC6 possesses in its DAC1 domain (HDAC6 has two active sites: DAC1 and DAC2). HDAC6 can target MSH2 even when it is in its heterodimeric complex; MSH2 deacetylation causes it to dissociate from its stabilizing partner MSH6 [65], and as a free monomer MSH2 can be ubiquitinated [58]. MSH2 can be acetylated at four lysine residues (K845, K847, K871, and K892), and all of these sites can also be ubiquitinated. MSH2 can be protected from ubiquitination and degradation by protein kinase C (PKC), which can phosphorylate the MutSα complex [66].

Further research out of our lab has indicated that ubiquitin-specific peptidase 10 (USP10) also plays a role in MSH2 stability, but rather than targeting it for degradation like HDAC6, USP10 is responsible for stabilizing MSH2 by deubiquitinating it [67]. USP10 has recently been identified as a regulator of p53 in response to DNA damage in a tumor development context [68, 69, 70]; ATM phosphorylation of USP10 induces its translocation to the nucleus, where it stabilizes p53. However, we now know that USP10 can work in opposition to HDAC6 by interacting with the N-terminal region of MSH2, while HDAC6 interacts with the C-terminal region. Under stress conditions (IR, carcinogen treatment), USP10 phosphorylation is increased [68] suggesting enhanced translocation to the nucleus where it may increase stabilization of the MutSα complex.

MMR can respond to endogenous insult to genomic integrity as well as exogenous. Oxidative DNA damage, for example, can induce MutSα-dependent PCNA ubiquitination, a process dependent on the PCNA E3-ubiquitin ligase RAD18 [71] in a process of noncanonical MMR (ncMMR) described by Jiricny et al. [72]. Briefly, ncMMR is mostly independent of DNA replication, lacks strand directionality, and could potentially play a role in genomic instability. This type of MMR occurs outside of S-phase when the dNTP pool is limited and replicative polymerases are not present, and the activity of MutLα in this situation can result in nicks in either strand of the DNA. This noncanonical MMR activation can itself promote ubiquitination of PCNA, which is directly responsible for recruiting pol-η (an error-prone polymerase) to chromatin [72] in the absence of higher fidelity polymerases. ncMMR is currently considered a stress response to genotoxic agents that contribute to genomic instability.

Advertisement

4. Repair of DNA strand breakage

While base and nucleotide damage can occur both by mistakes of the replicative machinery and chemical carcinogens, more robust insults to genome stability can induce single-strand and double-strand DNA breaks. These breaks can be caused by chemical carcinogens operating by different mechanisms than the ones previously mentioned, as well as ionizing radiation.

4.1. Homologous recombination

Homologous recombination (HR) is a major DNA repair pathway in which a sister strand of DNA is used to accurately repair DSBs. DSBs generally occur in euchromatin (as heterochromatin is relatively protected in its condensed state), and must be sensed, identified, and stabilized so that repair machinery can be recruited to the site without further damage occurring. The initial sensing of these ends occurs via the joint effort of ATM, and to a lesser extent, the MRN complex. ATM is a resident protein of the nucleus, existing in its inactive dimerized form, but upon the detection of a lesion it can activate itself via autophoshorylation. ATM can recognize large-scale changes in the chromatin structure [73], RNF8- and CHFR-mediated chromatin relaxation by histone ubiquitination [74], and R-loops (RNA/DNA hybrids) at lesions blocking the transcriptional machinery [75]. Thus, begins the ATM signaling cascade, recruiting a wide variety of DNA damage response elements and break responders, as well as the proteins that modify these responders to activate or enhance their function. To open the damage site to this massive recruitment effort, ATM phosphorylates the methyltransferase MMSET to methylate the surrounding histones and promote 53BP binding [76, 77]. ATM can also phosphorylate MDC1, which leads to the recruitment of ubiquitin ligase RNF8 via its FAA domain [78], which subsequently ubiquitinates histones H2A and H2AX, and promotes the retention of the factors recruited by ATM until the damage has been fully resolved [79]. MDC1, once initially activated by ATM, can bind ATM as well as the MRN complex, thus stabilizing these critical responders at the site of damage and amplifying their continued colocalization with the breaks [80, 81]. Ubiquitinated H2A and H2AX in the presence of RNF8 can recruit a second ubiquitin ligase, RNF168, which amplifies the ubiquitination signal at these histones and ensures that BRCA1, Rap80, Rad18, and 53BP1 localize to the site of damage [82, 83].

BRCA1 is a crucial responder to DNA damage that plays roles in cell cycle checkpoints, DNA cross link repair, and replication fork stability at the sites of DNA damage. Mutations in this gene severely limit its function and force cells to repair their DSBs via the error-prone process of NHEJ, which can predispose individuals to developing breast or ovarian cancer. BRCA1 can recruit RAD51 to the sites of DSBs and is necessary for the cell to repair the damage via homologous recombination and subsequent progress through the G2/M checkpoint [84, 85]. BRCA1 can also form a complex with BRCA2, which contributes to DNA break resolution. One of the proteins that can recruit BRCA1 to the DSB site is Rap80, which directs BRCA1 to K63-linked ubiquitin chains present on postreplication repair effector and sliding clamp PCNA [86]. These ubiquitin chains are generated by RING type E3 ubiquitin ligases RNF8 and RNF168 previously recruited by ATM action [79, 87]. Depletion of RAP80 has been demonstrated to increase the frequency of HR in reporter cells, and these cells eventually developed large chromosomal rearrangements.

BRCA1 itself can also serve as an E3 ubiquitin ligase by forming an obligate RING heterodimer with binding partner BARD1 [88], and this dimerization is required for BRCA1 to exert its tumor suppressor function. BRCA1’s RING domain is adjacent to a large sequence of α helices that interact with a similar α helix sequence on BARD1 [89], while the RING domain is left free to interact with E2 enzymes and exert its ubiquitin ligase activities on target proteins [90]. BRCA1-BARD1 is a type I dimeric RING E3 ubiquitin ligase, but is missing a conserved positive residue for these E3 ligases that is required for its binding activity, so this residue must be supplied by their binding partner [91]. The BRCA1-BARD1 heterodimer can target histones (H2A and H2AX), RNA polII, TFIIE, NPM1, CtIP, γ-tubulin, ER-α, and claspin [88]. BRCA1-BARD1 can also interact with 53BP1, and its ligase activity is thought to relocate 53BP1 to the periphery of the damage foci to allow for damage proteins like RPA and RAD51 to localize. E3 ligase-defective cells demonstrate reduced, but not entirely eliminated RPA and RAD51 foci in S-phase cells after being hit with a dose of IR [92]. However, in their normal S-phase counterparts, BRCA1 can counter the 53BP1-mediated stall on resection and allows HR to occur [93] by removing 53BP1 to the periphery and allowing RPA foci to form at the damage site [94]. These observations are thought to be mediated by the human homolog of the yeast SWI/SNF-like chromatin remodeler Fun30, SMARDCAD1, which is recruited by BRCA1-BARD1 to interact with BP531 and remove it from the vicinity of the break [95, 96]. 53BP1 appears to serve as a regulator of end resection and DSB resolution based on its associations with factors implicated in transcriptional silencing as well as its previously discussed functions. It can control the length of the resected ends in HR, and serves to prevent aberrant resection that can lead to RAD52-mediated ssDNA annealing and subsequent chromosomal rearrangements [97]. Further evidence for the interplay between BRCA1 and 53BP1 comes from mouse studies, where researchers found that lacking BRCA1 exon two (but expressing a RING-less BRCA1) is an embryonic lethal condition that can be rescued if the deletion occurs in a 53BP−/− embryo, suggesting that murine embryos lacking RING die because of the presence of 53BP [98]. BRCA1 and BARD1 interact with cyclin-dependent kinase 9 (CDK9) via their RING finger and BRCT domains, and localize to γ-H2AX foci indicative of damage to induce the process of HR over NHEJ [99].

Neddylation is a form of posttranslational modification similar to ubiquitination that has also been implicated in the process of double-strand break repair. Neural precursor cell expressed developmentally down-regulated 8 (NEDD8) is a ubiquitin-like protein involved in regulating cell growth, viability, and development [100]. Neddylation can serve as yet another layer of regulation in the function of DNA repair in damaged cells, and targeting this process has demonstrated some efficacy in preclinical models. Given the ubiquitous nature of BRCA1 in HR, it makes sense that this protein is a target of neddylation. In order for a cell to undergo HR, it must recognize the damage and be in the correct stage of the cell cycle (in this case, S/G2) so that a sister chromatid is present for the repair machinery to use as a template. This process of choice can be mediated by BRCA1 in complex with CtIP (RBBP8) in a number of different ways. For instance, CtIP must be phosphorylated on serine residue 327 for the cell to undergo HR, otherwise repair will be conducted via the error-prone process of microhomology-mediated end joining [101]. If this complex undergoes RNF111/UBE2M-mediated neddylation, the complex is rendered unable to perform its 5’→3′ nucleolytic end resection at the DSB, and without the ssDNA overhang tails HR cannot occur [102]. The COP9 signalosome is an additional mediator of the choice between types of DSB repair mechanisms [103]. COP9, the constitutive photomorphogenesis 9 signalosome, has significant homology with the 19S lid complex of the proteome and functions by deneddylating cullin-RING ubiquitin ligases, which may subsequently coordinate CRL-mediated ubiquitination of downstream protein targets [104]. COP9 is recruited to sites of DNA damage in a neddylation-dependent mechanism, and once there mediates deep end resection of the breaks, the first step of HR.

Targeting the process of neddylation as a preclinical strategy to sensitize tumors to chemotherapy is an avenue that has just recently began to garner attention. In a model of nonsmall cell lung cancer, neddylation inhibitor MLN4924 was able to inhibit the recruitment of members of the BRCA1 complex to sites of DNA damage. Examining expression of NEDD8, BRCA1, and PARP via Kaplan–Meier survival analysis revealed that high expression of these three factors correlated with a poor overall survival [105].

4.2. Nonhomologous end joining

The first step of nonhomologous end joining is the detection of the DSB by the Ku70/80 heterodimer, a 150 kDa Ku forms a ring-like structure that surrounds a single-strand of DNA with its central channel, and threads the broken DNA ends through this channel [106]. Because this protein can only accommodate one strand of DNA, in order for DNA replication to continue after resolution of the DSBs, Ku70/Ku80 must be removed [107]. The E3 ubiquitin ligase RING finger protein 8 (RNF8) has been found to down-regulate Ku80 at sites of DNA damage. Depletion of RNF8 leads to prolonged retention of Ku80 at damage sites and impairs NHEJ [108].

DNA-PK plays a central role in NHEJ of DNA DSBs largely during the G1 phase of the cell cycle as well as in V(D)J recombination [109, 110]. A poorly characterized ringer finger protein RNF144A has been reported as an E3 ubiquitin ligase for DNA-PK catalytic subunit (DNA-PKcs). RNF144A induces ubiquitination of DNA-PKcs in vitro and in vivo and promotes its degradation. Depletion of RNF144A results in an increased level of DNA-PKcs and resistance to DNA damaging agents [111]. Overall, there is no doubt that ubiquitination – either by regulating protein degradation or protein-protein interaction- plays a critical role in all five DNA repair families. Futures studies to better understand the role of ubiquitination, ubiquitin-like modifications, and enzymes responsible for these modifications in DNA repair pathways will be warranted.

References

  1. 1. Weissman AM. Themes and variations on ubiquitylation, nature reviews. Molecular Cell Biology. 2001;2:169-178
  2. 2. Dianov GL, Sleeth KM, Dianova II, Allinson SL. Repair of abasic sites in DNA. Mutation Research. 2003;531:157-163
  3. 3. Dianov G, Price A, Lindahl T. Generation of single-nucleotide repair patches following excision of uracil residues from DNA. Molecular and Cellular Biology. 1992;12:1605-1612
  4. 4. Dianov G, Bischoff C, Piotrowski J, Bohr VA. Repair pathways for processing of 8-oxoguanine in DNA by mammalian cell extracts. The Journal of Biological Chemistry. 1998;273:33811-33816
  5. 5. Khoronenkova SV, Dianov GL. The emerging role of mule and ARF in the regulation of base excision repair. FEBS Letters. 2011;585:2831-2835
  6. 6. 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
  7. 7. Wimmer U, Ferrari E, Hunziker P, Hubscher U. Control of DNA polymerase lambda stability by phosphorylation and ubiquitination during the cell cycle. EMBO Reports. 2008;9:1027-1033
  8. 8. Hoege C, Pfander B, Moldovan GL, Pyrowolakis G, Jentsch S. RAD6-dependent DNA repair is linked to modification of PCNA by ubiquitin and SUMO. Nature. 2002;419:135-141
  9. 9. van Loon B, Hubscher U. An 8-oxo-guanine repair pathway coordinated by MUTYH glycosylase and DNA polymerase lambda. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:18201-18206
  10. 10. Maga G, Villani G, Crespan E, Wimmer U, Ferrari E, Bertocci B, Hubscher U. 8-oxo-guanine bypass by human DNA polymerases in the presence of auxiliary proteins. Nature. 2007;447:606-608
  11. 11. Parsons JL, Tait PS, Finch D, Dianova II, Edelmann MJ, Khoronenkova SV, Kessler BM, Sharma RA, McKenna WG, Dianov GL. Ubiquitin ligase ARF-BP1/Mule modulates base excision repair. The EMBO Journal. 2009;28:3207-3215
  12. 12. Lee C, Smith BA, Bandyopadhyay K, Gjerset RA. DNA damage disrupts the p14ARF-B23(nucleophosmin) interaction and triggers a transient subnuclear redistribution of p14ARF. Cancer Research. 2005;65:9834-9842
  13. 13. Khan S, Guevara C, Fujii G, Parry D. p14ARF is a component of the p53 response following ionizing irradiation of normal human fibroblasts. Oncogene. 2004;23:6040-6046
  14. 14. Chen D, Kon N, Li M, Zhang W, Qin J, Gu W. ARF-BP1/mule is a critical mediator of the ARF tumor suppressor. Cell. 2005;121:1071-1083
  15. 15. Pandya RK, Partridge JR, Love KR, Schwartz TU, Ploegh HL. A structural element within the HUWE1 HECT domain modulates self-ubiquitination and substrate ubiquitination activities. The Journal of Biological Chemistry. 2010;285:5664-5673
  16. 16. Eymin B, Leduc C, Coll JL, Brambilla E, Gazzeri S. p14ARF induces G2 arrest and apoptosis independently of p53 leading to regression of tumours established in nude mice. Oncogene. 2003;22:1822-1835
  17. 17. Normand G, Hemmati P.G, Verdoodt B, von Haefen C, Wendt J, Guner D, May E, Dorken B, Daniel PT. p14ARF induces G2 cell cycle arrest in p53- and p21-deficient cells by down-regulating p34cdc2 kinase activity. The Journal of Biological Chemistry. 2005;280:7118-7130
  18. 18. Berneburg M, Lehmann AR. Xeroderma pigmentosum and related disorders: Defects in DNA repair and transcription. Advances in Genetics. 2001;43:71-102
  19. 19. Wood RD. Nucleotide excision repair in mammalian cells. The Journal of Biological Chemistry. 1997;272:23465-23468
  20. 20. Hoeijmakers JH. Genome maintenance mechanisms for preventing cancer. Nature. 2001;411:366-374
  21. 21. Volker M, Mone MJ, Karmakar P, van Hoffen A, Schul W, Vermeulen W, Hoeijmakers JH, van Driel R, van Zeeland AA, Mullenders LH. Sequential assembly of the nucleotide excision repair factors in vivo. Molecular Cell. 2001;8:213-224
  22. 22. Coin F, Oksenych V, Egly JM. Distinct roles for the XPB/p52 and XPD/p44 subcomplexes of TFIIH in damaged DNA opening during nucleotide excision repair. Molecular Cell. 2007;26:245-256
  23. 23. Hanawalt PC. Subpathways of nucleotide excision repair and their regulation. Oncogene. 2002;21:8949-8956
  24. 24. Sijbers AM, de Laat WL, Ariza R.R, Biggerstaff M, Wei YF, Moggs JG, Carter KC, Shell BK,Evans E, de Jong MC, Rademakers S, de Rooij J, Jaspers NG, Hoeijmakers JH, Wood RD.Xeroderma pigmentosum group F caused by a defect in a structure-specific DNA repair endonuclease. Cell. 1996;86:811-822
  25. 25. Watkins JF, Sung P, Prakash L, Prakash S. The Saccharomyces cerevisiae DNA repair gene RAD23 encodes a nuclear protein containing a ubiquitin-like domain required for biological function. Molecular and Cellular Biology. 1993;13:7757-7765
  26. 26. Madura K, Varshavsky A. Degradation of G alpha by the N-end rule pathway. Science (New York, N.Y.). 1994;265:1454-1458
  27. 27. Lambertson D, Chen L, Madura K. Pleiotropic defects caused by loss of the proteasome-interacting factors Rad23 and Rpn10 of Saccharomyces cerevisiae. Genetics. 1999;153:69-79
  28. 28. Elsasser S, Chandler-Militello D, Muller B, Hanna J, Finley D. Rad23 and Rpn10 serve as alternative ubiquitin receptors for the proteasome. The Journal of Biological Chemistry. 2004;279:26817-26822
  29. 29. Russell SJ, Reed SH, Huang W, Friedberg EC, Johnston SA. The 19S regulatory complex of the proteasome functions independently of proteolysis in nucleotide excision repair. Molecular Cell. 1999;3:687-695
  30. 30. Gillette TG, Huang W, Russell SJ, Reed SH, Johnston SA, Friedberg EC. The 19S complex of the proteasome regulates nucleotide excision repair in yeast. Genes & Development. 2001;15:1528-1539
  31. 31. Elsasser S, Gali RR, Schwickart M, Larsen CN, Leggett DS, Muller B, Feng MT, Tubing F, Dittmar GA, Finley D. Proteasome subunit Rpn1 binds ubiquitin-like protein domains. Nature Cell Biology. 2002;4:725-730
  32. 32. Groisman R, Polanowska J, Kuraoka I, Sawada J, Saijo M, Drapkin R, Kisselev AF, Tanaka K, Nakatani Y. The ubiquitin ligase activity in the DDB2 and CSA complexes is differentially regulated by the COP9 signalosome in response to DNA damage. Cell. 2003;113:357-367
  33. 33. Cope GA, Suh GS, Aravind L, Schwarz SE, Zipursky SL, Koonin EV, Deshaies RJ. Role of predicted metalloprotease motif of Jab1/Csn5 in cleavage of Nedd8 from Cul1. Science (New York, N.Y.). 2002;298:608-611
  34. 34. Sugasawa K, Okuda Y, Saijo M, Nishi R, Matsuda N, Chu G, Mori T, Iwai S, Tanaka K, Tanaka K, Hanaoka F. UV-induced ubiquitylation of XPC protein mediated by UV-DDB-ubiquitin ligase complex. Cell. 2005;121:387-400
  35. 35. Masutani C, Sugasawa K, Yanagisawa J, Sonoyama T, Ui M, Enomoto T, Takio K, Tanaka K, van der Spek PJ, Bootsma D, et al. Purification and cloning of a nucleotide excision repair complex involving the xeroderma pigmentosum group C protein and a human homologue of yeast RAD23. The EMBO Journal. 1994;13:1831-1843
  36. 36. Shivji MK, Eker AP, Wood RD. DNA repair defect in xeroderma pigmentosum group C and complementing factor from HeLa cells. The Journal of Biological Chemistry. 1994;269:22749-22757
  37. 37. Ng JM, Vermeulen W, van der Horst GT, Bergink S, Sugasawa K, Vrieling H, Hoeijmakers JH. A novel regulation mechanism of DNA repair by damage-induced and RAD23-dependent stabilization of xeroderma pigmentosum group C protein. Genes & Development. 2003;17:1630-1645
  38. 38. Okuda Y, Nishi R, Ng JM, Vermeulen W, van der Horst GT, Mori T, Hoeijmakers JH, Hanaoka F, Sugasawa K. Relative levels of the two mammalian Rad23 homologs determine composition and stability of the xeroderma pigmentosum group C protein complex. DNA Repair. 2004;3:1285-1295
  39. 39. Bregman DB, Halaban R, van Gool AJ, Henning KA, Friedberg EC, Warren SL. UV-induced ubiquitination of RNA polymerase II: A novel modification deficient in Cockayne syndrome cells. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:11586-11590
  40. 40. Woudstra EC, Gilbert C, Fellows J, Jansen L, Brouwer J, Erdjument-Bromage H, Tempst P, Svejstrup JQ. A Rad26-Def1 complex coordinates repair and RNA pol II proteolysis in response to DNA damage. Nature. 2002;415:929-933
  41. 41. Huibregtse JM, Yang JC, Beaudenon SL. The large subunit of RNA polymerase II is a substrate of the Rsp5 ubiquitin-protein ligase. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:3656-3661
  42. 42. Kleiman FE, Wu-Baer F, Fonseca D, Kaneko S, Baer R, Manley JL. BRCA1/BARD1 inhibition of mRNA 3′ processing involves targeted degradation of RNA polymerase II. Genes & Development. 2005;19:1227-1237
  43. 43. Svejstrup JQ. Rescue of arrested RNA polymerase II complexes. Journal of Cell Science. 2003;116:447-451
  44. 44. de Laat WL, Jaspers NG, Hoeijmakers JH. Molecular mechanism of nucleotide excision repair. Genes & Development. 1999;13:768-785
  45. 45. Sweder K, Madura K. Regulation of repair by the 26S proteasome. Journal of Biomedicine & Biotechnology. 2002;2:94-105
  46. 46. Lommel L, Ortolan T, Chen L, Madura K, Sweder KS. Proteolysis of a nucleotide excision repair protein by the 26S proteasome. Current Genetics. 2002;42:9-20
  47. 47. Lommel L, Chen L, Madura K, Sweder K. The 26S proteasome negatively regulates the level of overall genomic nucleotide excision repair. Nucleic Acids Research. 2000;28:4839-4845
  48. 48. Chen L, Shinde U, Ortolan TG, Madura K. Ubiquitin-associated (UBA) domains in Rad23 bind ubiquitin and promote inhibition of multi-ubiquitin chain assembly. EMBO Reports. 2001;2:933-938
  49. 49. Heessen S, Masucci MG, Dantuma NP. The UBA2 domain functions as an intrinsic stabilization signal that protects Rad23 from proteasomal degradation. Molecular Cell. 2005;18:225-235
  50. 50. Zhang Y. Transcriptional regulation by histone ubiquitination and deubiquitination. Genes & Development. 2003;17:2733-2740
  51. 51. de Napoles M, Mermoud JE, Wakao R, Tang YA, Endoh M, Appanah R, Nesterova TB, Silva J, Otte AP, Vidal M, Koseki H, Brockdorff N. Polycomb group proteins Ring1A/B link ubiquitylation of histone H2A to heritable gene silencing and X inactivation. Developmental Cell. 2004;7:663-676
  52. 52. Levinger L, Varshavsky A. Selective arrangement of ubiquitinated and D1 protein-containing nucleosomes within the drosophila genome. Cell. 1982;28:375-385
  53. 53. Bergink S, Salomons FA, Hoogstraten D, Groothuis TA, de Waard H, Wu J, Yuan L, Citterio E, Houtsmuller AB, Neefjes J, Hoeijmakers JH, Vermeulen W, Dantuma NP. DNA damage triggers nucleotide excision repair-dependent monoubiquitylation of histone H2A. Genes & Development. 2006;20:1343-1352
  54. 54. O'Driscoll M, Ruiz-Perez VL, Woods CG, Jeggo PA, Goodship JA. A splicing mutation affecting expression of ataxia-telangiectasia and Rad3-related protein (ATR) results in Seckel syndrome. Nature Genetics. 2003;33:497-501
  55. 55. Shroff R, Arbel-Eden A, Pilch D, Ira G, Bonner WM, Petrini JH, Haber JE, Lichten M.Distribution and dynamics of chromatin modification induced by a defined DNA double-strand break. Current Biology: CB. 2004;14:1703-1711
  56. 56. Marteijn JA, Bekker-Jensen S, Mailand N, Lans H, Schwertman P, Gourdin AM, Dantuma NP,Lukas J, Vermeulen W. Nucleotide excision repair-induced H2A ubiquitination is dependent on MDC1 and RNF8 and reveals a universal DNA damage response. The Journal of Cell Biology. 2009;186:835-847
  57. 57. Schroering AG, Edelbrock MA, Richards TJ, Williams KJ. The cell cycle and DNA mismatch repair. Experimental Cell Research. 2007;313:292-304
  58. 58. Zhang M, Xiang S, Joo HY, Wang L, Williams KA, Liu W, Hu C, Tong D, Haakenson J, Wang C, Zhang S, Pavlovicz RE, Jones A, Schmidt KH, Tang J, Dong H, Shan B, Fang B, Radhakrishnan R, Glazer PM, Matthias P, Koomen J, Seto E, Bepler G, Nicosia SV, Chen J, Li C, Gu L, Li GM, Bai W, Wang H, Zhang X. HDAC6 deacetylates and ubiquitinates MSH2 to maintain proper levels of MutSalpha. Molecular Cell. 2014;55:31-46
  59. 59. Hernandez-Pigeon H, Laurent G, Humbert O, Salles B, Lautier D. Degradation of mismatch repair hMutSalpha heterodimer by the ubiquitin-proteasome pathway. FEBS Letters. 2004;562:40-44
  60. 60. Drummond JT, Li GM, Longley MJ, Modrich P. Isolation of an hMSH2-p160 heterodimer that restores DNA mismatch repair to tumor cells. Science (New York, N.Y.). 1995;268:1909-1912
  61. 61. Genschel J, Littman SJ, Drummond JT, Modrich P. Isolation of MutSbeta from human cells and comparison of the mismatch repair specificities of MutSbeta and MutSalpha. The Journal of Biological Chemistry. 1998;273:19895-19901
  62. 62. Fink D, Aebi S, Howell SB. The role of DNA mismatch repair in drug resistance. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 1998;4:1-6
  63. 63. Li GM. The role of mismatch repair in DNA damage-induced apoptosis. Oncology Research. 1999;11:393-400
  64. 64. Li GM. Mechanisms and functions of DNA mismatch repair. Cell Research. 2008;18:85-98
  65. 65. Chang DK, Ricciardiello L, Goel A, Chang CL, Boland CR. Steady-state regulation of the human DNA mismatch repair system. The Journal of Biological Chemistry. 2000;275:18424-18431
  66. 66. Humbert O, Hermine T, Hernandez H, Bouget T, Selves J, Laurent G, Salles B, Lautier D.Implication of protein kinase C in the regulation of DNA mismatch repair protein expression and function. The Journal of Biological Chemistry. 2002;277:18061-18068
  67. 67. Zhang M, Hu C, Tong D, Xiang S, Williams K, Bai W, Li GM, Bepler G, Zhang X. Ubiquitin-specific peptidase 10 (USP10) Deubiquitinates and stabilizes MutS homolog 2 (MSH2) to regulate cellular sensitivity to DNA damage. The Journal of Biological Chemistry. 2016;291:10783-10791
  68. 68. Yuan J, Luo K, Zhang L, Cheville JC, Lou Z. USP10 regulates p53 localization and stability by deubiquitinating p53. Cell. 2010;140:384-396
  69. 69. Jochemsen AG, Shiloh Y. USP10: Friend and foe. Cell. 2010;140:308-310
  70. 70. Liu J, Xia H, Kim M, Xu L, Li Y, Zhang L, Cai Y, Norberg HV, Zhang T, Furuya T, Jin M, Zhu Z, Wang H, Yu J, Li Y, Hao Y, Choi A, Ke H, Ma D, Yuan J. Beclin1 controls the levels of p53 by regulating the deubiquitination activity of USP10 and USP13. Cell. 2011;147:223-234
  71. 71. Zlatanou A, Despras E, Braz-Petta T, Boubakour-Azzouz I, Pouvelle C, Stewart GS, Nakajima S, Yasui A, Ishchenko AA, Kannouche PL. The hMsh2-hMsh6 complex acts in concert with monoubiquitinated PCNA and pol eta in response to oxidative DNA damage in human cells. Molecular Cell. 2011;43:649-662
  72. 72. Pena-Diaz J, Bregenhorn S, Ghodgaonkar M, Follonier C, Artola-Boran M, Castor D, Lopes M, Sartori AA, Jiricny J. Noncanonical mismatch repair as a source of genomic instability in human cells. Molecular Cell. 2012;47:669-680
  73. 73. Bakkenist CJ, Kastan MB. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature. 2003;421:499-506
  74. 74. Wu J, Chen Y, Lu LY, Wu Y, Paulsen MT, Ljungman M, Ferguson DO, Yu X. Chfr and RNF8 synergistically regulate ATM activation. Nature Structural & Molecular Biology. 2011;18:761-768
  75. 75. Tresini M, Warmerdam DO, Kolovos P, Snijder L, Vrouwe MG, Demmers JA, van IJcKen WF, Grosveld FG, Medema RH, Hoeijmakers JH, Mullenders LH, Vermeulen W, Marteijn JA. The core spliceosome as target and effector of non-canonical ATM signalling. Nature. 2015;523:53-58
  76. 76. Huyen Y, Zgheib O, Ditullio RA Jr, Gorgoulis VG, Zacharatos P, Petty TJ, Sheston EA, Mellert HS, Stavridi ES, Halazonetis TD. Methylated lysine 79 of histone H3 targets 53BP1 to DNA double-strand breaks. Nature. 2004;432:406-411
  77. 77. Pei H, Zhang L, Luo K, Qin Y, Chesi M, Fei F, Bergsagel PL, Wang L, You Z, Lou Z. MMSET regulates histone H4K20 methylation and 53BP1 accumulation at DNA damage sites. Nature. 2011;470:124-128
  78. 78. Bekker-Jensen S, Mailand N. The ubiquitin- and SUMO-dependent signaling response to DNA double-strand breaks. FEBS Letters. 2011;585:2914-2919
  79. 79. 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
  80. 80. Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA, Celeste A, Manis JP, van Deursen J, Nussenzweig A, Paull TT, Alt FW, Chen J. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Molecular Cell. 2006;21:187-200
  81. 81. Stiff T, O'Driscoll M, Rief N, Iwabuchi K, Lobrich M, Jeggo PA. ATM and DNA-PK function redundantly to phosphorylate H2AX after exposure to ionizing radiation. Cancer Research. 2004;64:2390-2396
  82. 82. Wilson MD, Benlekbir S, Fradet-Turcotte A, Sherker A, Julien JP, McEwan A, Noordermeer SM, Sicheri F, Rubinstein JL, Durocher D. The structural basis of modified nucleosome recognition by 53BP1. Nature. 2016;536:100-103
  83. 83. Fradet-Turcotte A, Canny MD, Escribano-Diaz C, Orthwein A, Leung CC, Huang H, Landry MC, Kitevski-LeBlanc J, Noordermeer SM, Sicheri F, Durocher D. 53BP1 is a reader of the DNA-damage-induced H2A Lys 15 ubiquitin mark. Nature. 2013;499:50-54
  84. 84. Xu X, Weaver Z, Linke SP, Li C, Gotay J, Wang XW, Harris CC, Ried T, Deng CX. Centrosome amplification and a defective G2-M cell cycle checkpoint induce genetic instability in BRCA1 exon 11 isoform-deficient cells. Molecular Cell. 1999;3:389-395
  85. 85. Moynahan ME, Chiu JW, Koller BH, Jasin M. Brca1 controls homology-directed DNA repair. Molecular Cell. 1999;4:511-518
  86. 86. Wang B, Matsuoka S, Ballif BA, Zhang D, Smogorzewska A, Gygi SP, Elledge SJ. Abraxas and RAP80 form a BRCA1 protein complex required for the DNA damage response. Science (New York, N.Y.). 2007;316:1194-1198
  87. 87. Huen MS, Grant R, Manke I, Minn K, Yu X, Yaffe MB, Chen J. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell. 2007;131:901-914
  88. 88. Wu W, Koike A, Takeshita T, Ohta T. The ubiquitin E3 ligase activity of BRCA1 and its biological functions. Cell Division. 2008;3:1
  89. 89. Brzovic PS, Rajagopal P, Hoyt DW, King MC, Klevit RE. Structure of a BRCA1-BARD1 heterodimeric RING-RING complex. Nature Structural Biology. 2001;8:833-837
  90. 90. Lorick KL, Jensen JP, Fang S, Ong AM, Hatakeyama S, Weissman AM. RING fingers mediate ubiquitin-conjugating enzyme (E2)-dependent ubiquitination. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:11364-11369
  91. 91. Taherbhoy AM, Huang OW, Cochran AG. BMI1-RING1B is an autoinhibited RING E3 ubiquitin ligase. Nature Communications. 2015;6:7621
  92. 92. Densham RM, Morris JR. The BRCA1 ubiquitin ligase function sets a new trend for remodelling in DNA repair. Nucleus (Austin, Tex.). 2017;8:116-125
  93. 93. Bunting SF, Callen E, Wong N, Chen HT, Polato F, Gunn A, Bothmer A, Feldhahn N, Fernandez-Capetillo O, Cao L, Xu X, Deng CX, Finkel T, Nussenzweig M, Stark JM, Nussenzweig A. 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell. 2010;141:243-254
  94. 94. Chapman JR, Sossick AJ, Boulton SJ, Jackson SP. BRCA1-associated exclusion of 53BP1 from DNA damage sites underlies temporal control of DNA repair. Journal of Cell Science. 2012;125:3529-3534
  95. 95. Costelloe T, Louge R, Tomimatsu N, Mukherjee B, Martini E, Khadaroo B, Dubois K, Wiegant WW, Thierry A, Burma S, van Attikum H, Llorente B. The yeast Fun30 and human SMARCAD1 chromatin remodellers promote DNA end resection. Nature. 2012;489:581-584
  96. 96. Chen X, Cui D, Papusha A, Zhang X, Chu CD, Tang J, Chen K, Pan X, Ira G. The Fun30 nucleosome remodeller promotes resection of DNA double-strand break ends. Nature. 2012;489:576-580
  97. 97. Ochs F, Somyajit K, Altmeyer M, Rask MB, Lukas J, Lukas C. 53BP1 fosters fidelity of homology-directed DNA repair. Nature Structural & Molecular Biology. 2016;23:714-721
  98. 98. Li M, Cole F, Patel DS, Misenko SM, Her J, Malhowski A, Alhamza A, Zheng H, Baer R, Ludwig T, Jasin M, Nussenzweig A, Serrano L, Bunting SF. 53BP1 ablation rescues genomic instability in mice expressing 'RING-less' BRCA1. EMBO Reports. 2016;17:1532-1541
  99. 99. Nepomuceno TC, Fernandes VC, Gomes TT, Carvalho RS, Suarez-Kurtz G, Monteiro AN, Carvalho MA. BRCA1 recruitment to damaged DNA sites is dependent on CDK9. Cell cycle (Georgetown, Tex.). 2017;16:665-672
  100. 100. Xirodimas DP. Novel substrates and functions for the ubiquitin-like molecule NEDD8. Biochemical Society Transactions. 2008;36:802-806
  101. 101. Yun MH, Hiom K. CtIP-BRCA1 modulates the choice of DNA double-strand-break repair pathway throughout the cell cycle. Nature. 2009;459:460-463
  102. 102. Jimeno S, Fernandez-Avila MJ, Cruz-Garcia A, Cepeda-Garcia C, Gomez-Cabello D, Huertas P. Neddylation inhibits CtIP-mediated resection and regulates DNA double strand break repair pathway choice. Nucleic Acids Research. 2015;43:987-999
  103. 103. Meir M, Galanty Y, Kashani L, Blank M, Khosravi R, Fernandez-Avila MJ, Cruz-Garcia A, Star A, Shochot L, Thomas Y, Garrett LJ, Chamovitz DA, Bodine DM, Kurz T, Huertas P, Ziv Y, Shiloh Y. The COP9 signalosome is vital for timely repair of DNA double-strand breaks. Nucleic Acids Research. 2015;43:4517-4530
  104. 104. Lee MH, Zhao R, Phan L, Yeung SC. Roles of COP9 signalosome in cancer. Cell cycle (Georgetown, Tex.). 2011;10:3057-3066
  105. 105. Guo ZP, Hu YC, Xie Y, Jin F, Song ZQ, Liu XD, Ma T, Zhou PK. MLN4924 suppresses the BRCA1 complex and synergizes with PARP inhibition in NSCLC cells. Biochemical and Biophysical Research Communications. 2017;483:223-229
  106. 106. Walker JR, Corpina RA, Goldberg J. Structure of the Ku heterodimer bound to DNA and its implications for double-strand break repair. Nature. 2001;412:607-614
  107. 107. Postow L. Destroying the ring: Freeing DNA from Ku with ubiquitin. FEBS Letters. 2011;585:2876-2882
  108. 108. Feng L, Chen J. The E3 ligase RNF8 regulates KU80 removal and NHEJ repair. Nature Structural & Molecular Biology. 2012;19:201-206
  109. 109. Davis AJ, Chen BP, Chen DJ. DNA-PK: A dynamic enzyme in a versatile DSB repair pathway. DNA Repair. 2014;17:21-29
  110. 110. Zeman MK, Cimprich KA. Causes and consequences of replication stress. Nature Cell Biology. 2014;16:2-9
  111. 111. Ho SR, Mahanic CS, Lee YJ, Lin WC. RNF144A, an E3 ubiquitin ligase for DNA-PKcs, promotes apoptosis during DNA damage. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:E2646-2655

Written By

Niko Moses and Xiaohong Mary Zhang

Submitted: 21 November 2016 Reviewed: 06 October 2017 Published: 20 December 2017