Skip to main content
Molecular and Cellular Biology logoLink to Molecular and Cellular Biology
. 2023 Jul 13;43(8):401–425. doi: 10.1080/10985549.2023.2224199

Implications of Translesion DNA Synthesis Polymerases on Genomic Stability and Human Health

Jegadheeswari Venkadakrishnan a,*, Ganesh Lahane a,*, Arti Dhar a, Wei Xiao b, Krishna Moorthi Bhat c, Tej K Pandita d,, Audesh Bhat e,
PMCID: PMC10448981  PMID: 37439479

Abstract

Replication fork arrest-induced DNA double strand breaks (DSBs) caused by lesions are effectively suppressed in cells due to the presence of a specialized mechanism, commonly referred to as DNA damage tolerance (DDT). In eukaryotic cells, DDT is facilitated through translesion DNA synthesis (TLS) carried out by a set of DNA polymerases known as TLS polymerases. Another parallel mechanism, referred to as homology-directed DDT, is error-free and involves either template switching or fork reversal. The significance of the DDT pathway is well established. Several diseases have been attributed to defects in the TLS pathway, caused either by mutations in the TLS polymerase genes or dysregulation. In the event of a replication fork encountering a DNA lesion, cells switch from high-fidelity replicative polymerases to low-fidelity TLS polymerases, which are associated with genomic instability linked with several human diseases including, cancer. The role of TLS polymerases in chemoresistance has been recognized in recent years. In addition to their roles in the DDT pathway, understanding noncanonical functions of TLS polymerases is also a key to unraveling their importance in maintaining genomic stability. Here we summarize the current understanding of TLS pathway in DDT and its implication for human health.

Keywords: DNA damage, genomic instability, DNA damage tolerance, translesion synthesis, TLS polymerases, human diseases

INTRODUCTION

DNA damage in cells is a continuous process resulting in the creation of DNA lesions via exogenous and endogenous factors (Fig. 1). Exogenous sources include chemical compounds, ionizing radiation (IR), ultraviolet (UV), chemotherapeutic-induced DNA damage, tobacco smoke carcinogens, and alcohol, among many others, whereas endogenous sources include reactive oxygen species (ROS), spontaneous deamination, DNA methylation, apurinic or apyrimidinic (AP) sites, and active enzymatic DNA cleavage.1 Error-free replication and transmission of DNA are needed to maintain genomic stability, which translates the necessary information for life.2,3 Cell death may occur due to an imbalance in DNA integrity, leading to premature aging and cancer. However, all organisms have a DNA damage repair mechanism that correctly senses and repairs DNA lesions and move their genome to the next generation with high fidelity.2,4 The mammalian genome on average gets approximately 105 spontaneous lesions per day,5 most of which are removed by DNA repair mechanisms; however, when the DNA template is missing and the DNA is unwound at replication forks, DNA lesions persist through S-phase. To avoid replication fork collapse and DNA breaks, cells undergo a lesion bypass process via DNA damage tolerance (DDT) pathways and restart the DNA replication process (Fig. 1). DNA damage tolerance can occur mainly through error-free template switching (TS) and error-prone translesion synthesis (TLS).6 Template switching utilizes the newly synthesized nascent sister chromatid as a template for lesion bypass, while TLS employs specialized DNA polymerases to replicate across the otherwise replication-blocking lesion. These pathways are tightly regulated to avoid more severe consequences arising from stalled replication forks.7

FIG 1.

FIG 1

The endogenous, exogenous, and spontaneous decay-induced DNA damage.

Replicative DNA polymerases with stringent requirements can replicate the entire genome with high fidelity but cannot use a template with replication-blocking lesions. When encountered with DNA damage, the replication fork stalls, resulting in long stretches of single-stranded DNA (ssDNA) attached to the replication protein A (RPA) [8]. This leads to the initiation of Ataxia telangiectasia mutated (ATM) and Rad3-related (ATR) checkpoint and initiation of lesion bypass by TLS through monoubiquitinated proliferating cell nuclear antigen (PCNA).8–11 Translesion DNA synthesis uses specialized DNA polymerases to perform lesion bypass using damaged DNA as a template for nucleotide insertion opposite to lesions. Generally, TLS polymerases do not have 3′–5′ exonucleolytic proofreading activity, which helps TLS to avoid additional kinetic barriers. Moreover, they have a more significant active site that helps accommodate bulky DNA adducts, facilitating replication by directly bypassing the lesion.12

To date, it is reported that multiple processes such as transcription, post-translational modifications (PTMs), protein-protein interactions (PPIs), and noncoding RNAs tightly control the critical events of TLS in mammalian cells. Furthermore, this involves polymerase switching and PCNA monoubiquitination (PCNA-mUb). Such tight regulation not only ensures the recruitment of lesion-appropriate TLS polymerase to the damaged site to salvage replication, but also blocks access of these mutagenic DNA polymerases to the undamaged DNA. In addition, TLS is involved in the initiation, progression as well as chemoresistance of many cancers, signifying its importance as a potential target for cancer treatment.

This review summarizes the type of TLS polymerases and their structural properties, TLS polymerase-induced DNA damage, and DNA damage tolerance pathways, which will facilitate the development of potential TLS inhibitors for sensitizing tumor cells to chemotherapy. We will also discuss the noncanonical functions of TLS polymerases and TLS polymerase-associated human diseases.

TYPE OF TLS POLYMERASES AND THEIR STRUCTURAL PROPERTIES

TLS polymerases are a particular class of DNA polymerases capable of utilizing damaged DNA as a template for DNA synthesis.13,14 TLS polymerases are found across all forms of life and consist of the following: Dpo4 (dpo4) and Dbh (dbh) in archaea, Pol II (polB), Pol IV (dinB), and Pol V (umuDC) in prokaryotes, and Rev1 (REV1), Pol κ (POLK/DINB1), Pol η (RAD30/POLH), Pol ζ [Rev3 (REV3L) and Rev7 (REV7/MAD2BL2) subunits], Pol ι (POLI/RAD30B), Pol λ (POLL), Pol β (POLB; not a classical TLS polymerase), Pol θ (POLQ), and the newly identified primase and polymerase enzyme PrimPol (PRIMPOL/CCDC111) in eukaryotes [15–17]. Depending on the protein sequence, eukaryotic domain organization of human TLS polymerases is shown in Fig. 2. TLS polymerases are classified into A, B, X, Y, and PrimPol families.15–17 The majority of the TLS polymerases, including Dpo4, Dbh, Pol IV, Pol V, Rev1, Pol κ, Pol η, and Polι, belong to the Y family DNA polymerases.18 The eukaryotic Pol ζ and prokaryotic Pol II belong to the B family, which also includes high-fidelity replicative DNA polymerases. Unlike other B family polymerases, Pol ζ is a low-fidelity polymerase capable of carrying out TLS and has a unique property to extend from primer-template pairs that are mismatched and/or deformed, including those that are opposite to DNA lesions.19 Pol λ and Pol β belong to the X family whereas Pol θ belongs to the A family.

FIG 2.

FIG 2

Human TLS polymerases and their protein domain.

In general, all Y family polymerases feature a predefined active site where deoxyribonucleoside triphosphate (dNTPs) and pyrophosphate may smoothly migrate in and out due to a tiny finger and thumb domain.20 In Pol κ the fifth domain, known as N-clasp is located at the N-terminus.21 Through interaction with the catalytic core, the little finger, and the DNA duplex region, the Pol κ N-clasp engages with DNA and stabilizes the ternary complex polymerase coupled with DNA and dNTP.22 Despite having 870 amino acid residues, human Pol κ utilizes a primer extension strategy by truncating proteins that keep residues 19–526 active enough to maintain wild-type-like polymerization.21 Until recently, it was difficult to determine the nature of interactions between N-clasp residues and other components of the complex from Pol κ crystal structures; however, a 2.0-Å resolution crystal structure demonstrated structural traits and interactions that were essential for our understanding of this function.23 In a recent study, the reconstruction of structural details of Pol κ in interaction with PCNA and DNA using cryo-electron microscopy (Cryo-EM) mapping revealed the flexible nature of Pol κ C-terminal region (aa residues 535–870), as this region remained invisible in the Cryo-EM map.24,25 Likewise, the N-clasp residues 21–45 also appeared invisible on the map, suggesting a partly flexible nature of the N-clasp as well.25 Besides this, the Cryo-EM mapping also revealed that the clash between the inner rim of the PCNA and the DNA duplex emerging from the catalytic core of Pol κ is avoided by bending the DNA by ∼30°.25 Similar to Pol κ, the three dimensional structure of Rev1 contains a huge gap that separates the little finger (LF) domain from the ring-shaped catalytic core that surrounds the DNA.26 Rev1 structurally has a detachable LF and an N-terminal extension known as the N-digit that connects the LF to the catalytic core.24,25 It has been proposed that two substantial inserts in the human Rev1's finger and palm domains, which are specific to its catalytic core, have a significant role in conferring Rev1 the unique property of carrying out protein-protein interaction during TLS.27

Human Pol ι is an 80-kDa protein with 715 amino acids (GenBank Accession number AF140501.1); however, some isoforms have 740 amino acids, and the commonly found isoform has only 415 amino acids.28 The POLI gene, which is found on human chromosome 18q21.2, has several organismal homologs and is evolutionarily conserved.29 Pol ι structurally resembles the highly conserved Y family polymerase right-hand configuration, with fingers (aa 38–98), thumb (aa 225–288), and palm (aa 25–37 and 99–224) forming the N-terminally located catalytic active site.30 The crystal structure of Pol ι in a ternary complex with DNA substrates and an incoming nucleotide provides important new information about the unique base selectivity, which is characterized by a 105-fold change in fidelity depending on the template base.31,32 For the catalytic activity, all DNA polymerases require a divalent cation. Mg2+ is often thought to be the physiological cofactor for replicative DNA polymerases in vivo.33 The catalytic affinity of pol ι for the metal ions of Mn2+ was higher than that of Mg2+ during TLS.34 Also, Mn2+ achieves more optimal octahedral coordination geometry than Mg2+, with lower values in average coordination distance geometry within the catalytic metal A-site, and crystal structures of Pol ι ternary complex include the primer terminus 3-OH and a nonhydrolyzable dCTP analog opposite G.34 Pol ι appears to contain an intrinsic 5’-deoxyribose phosphate (dRP) lyase activity;35 however, dRP lyase of Pol ι does not shield cells Pol β lacking dRP lyase activity from methylation-induced cytotoxicity, consistent with the exclusion of pol ι from BER of certain lesions.36

Human Pol η is encoded by POLH located on chromosome 6p21.1.37–39 It contains 713 amino acids and plays a vital role in bypassing thymine-thymine (TT) cyclobutane pyrimidine dimers (CPDs) caused by UV irradiation in an error-free manner. It does so by inserting correct nucleotides across the lesion.38,40 While the carboxy-terminus of Pol η is made up of two PCNA interaction protein (PIP) domains and a ubiquitin-binding zinc finger domain (UBZ), the N-terminus of Pol η contains protein-arginine deiminase (PAD) and catalytic domains (Fig. 2).38,40 Because the Pol η active site is bigger than that of other Y-family polymerases, it can accommodate two nucleotides of the CPDs and allow Pol η to synthesize DNA over large adducts.20 This is obvious from the structure of yeast Pol η, which was simulated with a cis-syn TT dimer in the template DNA and incoming deoxyadenosine triphosphate (dATP) (Table 1).41,42 Pol η often causes both clustered and unclustered mutations at A:T nucleotide pairings in numerous cancer types.43 Some carcinogens promote the activity of error-prone noncanonical mismatch repair (ncMMR) pathway involving Pol η, leading to an increased relative mutation rate in the H3K36me3-marked region.44,45 ncMMR is mainly independent of DNA replication, lacks strand directionality, induces PCNA monoubiquitylation, and increases the recruitment of the error-prone polymerase to chromatin,44,45 implying that some environmental factors cause cancer by redistributing mutations to targeted genome regions rather than raising the overall mutation rate.

TABLE 1.

Types of human TLS polymerases, their clinical role, and bypass lesions

TLS polymerase Organism; gene name Mechanisms Clinical role/preclinical roles Bypassed lesion
Pol η Human DNA polymerase; POLH gene Nucleotide insertion and extension (extends up to 2 nucleotides beyond the lesion) is carried out by the enlarged active site of the Pol η polymerase catalytic domain (residues 1–432) [20].
The Pol η’s wide active site stably accommodates the two covalently bonded thymines (T-T dimmer). Unlike other Y family polymerases, Pol η operates by stabilizing the DNA's B-form conformation through a strong electrostatic interaction with four consecutive DNA template phosphates [20, 51].
In the TT bypass, Pol η functions with a biased fidelity, bypassing 5’ T with a low error rate than the 3’ T [52].
Humans who lack active Pol η due to mutation(s) suffer a type of UV-sensitive condition known as Xeroderma pigmentosum variant (XPV), which increases the risk for skin cancer [53, 54].
Pol η expression levels are negatively associated with the survival rate of people with non-small-cell lung cancer, metastatic gastric adenocarcinoma, and head and neck squamous cell cancer [55].
Cisplatin-induced guanine-guanine intra-strand adducts, 8-ooGuanine Abasic sites,
UV-induced lesions, particularly T-T cyclobutane pyrimidine dimers, N-2-acetylaminofluorene (AAF)-modified guanine [35, 120, 121].
Pol ι Human DNA
Polymerase;
POLI gene
Pol ι can efficiently incorporate deoxythymidine triphosphate (dTTP) opposite dA but due to its unusual active site, it promotes Hoogsteen pairing instead (3–10 fold increased dG-dT mispairing) [12, 56].
It may be that Pol ι can avoid N1-methylated deoxyadenosine by using the dA(syn): dTTP (anti) combination [57].
The human cell cycle constrained by UV-induced DNA damage is improved by elevated levels of pol ι triggered by p53 [58]. 5-hydroxyuracil (5-OHU), N2-guanine adduct, 5-hydroxycytosine (5-OHC), thymine-thymine 6–4 Photoproducts, 5,6-dihydrouracil (5,6-DHU), 8-oxoGuanine [127, 128].
Pol κ Human DNA polymerase; POLK gene Polycyclic aromatic hydrocarbon adducts, such as benzo[a]pyrene diol epoxide (BPDE) covalently bound to N2 of guanine in the minor groove, are specifically avoided by Pol κ [59]. Pol κ deficient mice have a spontaneous mutator phenotype, and the mutation rate rises as dietary cholesterol rises, suggesting that Pol κ can avoid naturally occurring steroid adducts [60, 61]. Benzo[a]pyrene-guanine adducts (BP-G),
Thymine glycol,8-oxo-d-guanine [14, 116–119].
Rev1 Human DNA polymerase;
REV1 gene
When the template is abasic, deoxycytidylic (dCMP) transferase activity of Rev1 is highest, followed by dG and dA [50], and later, it was confirmed that Rev1 bypasses abasic lesions in vivo [62]. In somatic hypermutation, Rev1 adds deoxycytidine residues opposite the abasic sites [52].
In human colon carcinoma cells, loss of p53 increases Rev1 expression [58].
8-oxoguanine (8-oxoG), 1, N 6-ethenoadenine adducts, UV-induced lesions, Trans-anti-benzo[a]pyrene-N 2-dG [123, 124].
Pol ζ Human DNA Polymerase; REV3L and REV7 genes Pol ζ was thought to integrate mismatched dNTPs and circumvent various DNA lesions, but later on, its main role was identified in the primer extension step of TLS [10, 63, 64].
Rev7 can cause biallelic mutation in the TLS and the direct binding of CHAMP1 to REV7 lowers the concentration of the Shieldin complex, increasing double-strand break end excision (PMID: 36044844).
Increased DNA damage tolerance and decreased spontaneous tumorigenesis are two ways that Pol ζ improves genomic stability [64].
Impairment of Rev1 precursors has been found to cause bone marrow failure in Fanconi anemia (PMID: 27500492).
cyclobutane pyrimidine dimers, Extender polymerase for numerous lesions, thymine-thymine 6–4Photoproducts [125, 126].
Pol θ Human DNA Polymerase; POLQ gene Pol θ has a role in DSB repair by error-prone end-joining, referred to as alt-EJ or MMEJ [65]. Additionally, Pol θ can expand from mismatched bases opposite large lesions such as 6–4 photoproducts and thymine glycol [66, 67]. Pol θ is overexpressed in various types of cancers [65] which is related to the lymphoid tissue and mainly found associated with lung, stomach, and colon cancer [58–60]. Additionally, in women with an unknown breast cancer gene (BRCA) 1/2 profile, a mutation in the promoter region of POLQ (c.-1060A > G) was found strongly linked to both hereditary breast and ovarian cancer [68]. Thymine glycols, UV-induced lesions, 1, N 6-ethenoadenine adducts [116, 122].
Pol λ Human DNA polymerase; POLL gene PCNA and Pol λ engage in physical and functional interaction without changing the rate of nucleotide incorporation. This interaction stabilizes the binding of Pol λ to the primer template and improves the processivity of DNA synthesis [69–71]. The R438W hPol λ mutation causes chromosomal aberrations, impairs NHEJ-mediated DSB repair, and increases mutation frequency when expressed ectopically in mammalian cells, a potential cause for cancer due to chromosome level genomic instability [72]. 1,2-dihydro-2-oxoadenine (2-OH-A) (PMID: 17666409).
PrimPol Human DNA Polymerase; PRIMPOL gene PrimPol can avoid templated oxidative lesions such as 8-oxo-guanine (8-oxoG), abasic sites, cyclobutane pyrimidine dimer (CPD), and pyrimidine pyrimidone (6–4) photoproducts (6–4 PPs) [73, 74]. Human cells are vulnerable to UVC-induced cell death only in the absence of functional Pol η, as in XPV patients [75]. Apyrimidinic/apurinic site [67].

Pol ζ has a pentameric ring-like design with a catalytic subunit Rev3, two noncatalytic Rev7 subunits, and auxiliary Pol31 and Pol32 subunits, forming a continuous daisy chain of protein-protein interactions [38,69]. REV7 was first recognized as a gene responsible for UV-induced mutagenesis in Saccharomyces cerevisiae46 and later findings referred to the Rev3-Rev7 complex as DNA polymerase ζ since it is the sixth eukaryotic DNA polymerase (Table 1).19 The human mitotic checkpoint protein hMAD2 shares 23% identity and 53% similarity with hRev7, which also shares 23% identity and 54% similarity with ScRev7. Although no deleterious nucleotide alteration(s) have been detected in hREV3 or hREV7 genes in primary human tumors or human tumor cell lines, hREV7 is mapped to chromosome 1p36 in a location that often suffers from loss of heterozygosity in human cancers and hREV3 is located in the common fragile site, a hotspot of chromosomal rearrangements.19,47–49 The major role of Rev7 is to connect between Rev3 and Rev1, while a new structural study has shown possible interactions between Rev7 and Pol32 (POLD3 in mammals), a regulatory component shared by Pol ζ and the replicative polymerase δ.49,50

PrimPol is emerging as an important factor in the tolerance to DNA damage, notably in vertebrate and human cells. PrimPol is a 560 amino acid long protein in humans and is found in vertebrates, plants, and lower eukaryotes with a varied number of amino acid residues.51–53 The PrimPol nucleotidyltransferase activity, necessary for its primase and polymerase activities, is regulated by two highly conserved modules, ModN (residues 35–105) and ModC (residues 108–200 and 261–348) in the catalytic archaeo-eukaryotic primase (AEP) core.54 Changes to the catalytic residues Asp114, Glu116, or Asp280 of ModC abolish both primase and polymerase activities of PrimPol.54 The DNA binding affinity of ModN is diminished when critical residues are modified, resulting in a decreased enzymatic activity.55,56 A zinc finger (ZnF) domain, an RPA binding domain, and an N-terminal helix that stabilizes PrimPol’s interaction with ssDNA are all present in PrimPol in addition to the AEP core.54,55 PrimPol has a low processivity while synthesizing DNA, infrequently adding more than four nucleotides to an intact template. The ZnF domain of PrimPol, which is linked to the catalytic core by a 140-amino-acid flexible linker and regulates the amount of nucleotide incorporation, is one of the mechanisms responsible for this poor processivity.56,57

TLS POLYMERASE-INDUCED MUTAGENESIS

Correct replication by DNA polymerases is a critical step for normal biological processes and the maintenance of genetic integrity. Various human diseases have been associated with mutations in genes involved in the DDT pathway, and importantly the TLS polymerase genes. For instance, mutations in POLB result in lymphomas in mouse and human breast, lung, colorectal, and prostate carcinomas.58 Likewise, mutations in POLG cause progressive external ophthalmoplegia.59 POLD1 and POLE encode human DNA polymerases δ and ε, respectively, and their mutations cause hereditary colorectal cancer (doi: 10.1186/s13046-022-02422-1). Xeroderma pigmentosum variant (XPV), a well-characterized syndrome associated with a defective TLS pathway, is caused by a mutation in POLH.60,61 As evident from several studies,4,5,12,16 the importance of DNA polymerases in maintaining genomic stability and, therefore, human health is beyond doubt.

Numerous studies have suggested that TLS polymerases exhibit decreased replication fidelity due to their lack of 3'>5' proofreading activity and the increased use of altered DNA as templates. Compared to TLS polymerases, replicative DNA polymerases exhibit error rates in the range of one nucleotide for every 106 to 108 replicated bases.62 In contrast, TLS polymerases when replicating undamaged DNA exhibit error rates in the range of one base for every 10–10,000 bases.41,62–65 Moreover, unlike replicative polymerases, TLS polymerases do not proceed with an induced fit model upon nucleotide binding,65,66 allowing bypass of replication-blocking lesions. Furthermore, some TLS polymerases, such as Rev1 and DNA Pol ι, do not use the Watson–Crick canonical base pairing.42,67 Hence, TLS polymerase activities are largely mutagenic.

TLS POLYMERASES AND DNA DAMAGE TOLERANCE

DNA regularly comes into contact with endogenous and exogenous substances that cause DNA damage. The presence of unpaired DNA lesions in the template can arrest the replication fork movement that leads to genomic instability via fork collapse and DSB formation. To complete the replication and avoid strand breaks, DDT plays an essential role in allowing replication in the presence of template lesions.68 There are two main DDT pathways: (i) TLS, which involves the recruitment of TLS DNA polymerase to bypass the lesion, which can occur either directly at the site of replication arrest or behind the arrest by repriming DNA synthesis at the daughter strand gaps (DSGs) (Fig. 3); and (ii) homology-directed damage tolerance, which is accomplished by fork reversal (FR) and/or TS (Fig. 3).69 Monoubiquitination of the clamp protein PCNA in response to DNA damage results in the recruitment of TLS polymerases, which are capable of bypassing the lesion. At the site of an arrested fork, lesion bypass takes place directly or during gap filling following replication restart away from the lesion site.26,70 However, despite certain exceptions mentioned below, TLS polymerase-mediated lesion bypass tends to misincorporate bases resulting in increased mutations.15,70 The error-prone characteristics of TLS polymerases have been implicated in both carcinogenesis and resistance to chemotherapy.16,71,72 K63-linked polyubiquitination of monoubiquitinated PCNA initiates FR and/or TS to facilitate a temporary switch from the lesion-containing strand to a newly synthesized complementary sister chromatid strand of the homologous chromosome. Because the undamaged template is being copied, FR and TS are considered error-free lesion bypass pathways [98]. A structure resembling a chicken foot is created by fork reversal, allowing the replisome on the halted nascent strand to reach the homologous sister template.73,74 Different from FR, TS arises after repriming at DSGs generated at lesion locations behind the replication fork.75,76 Template switch is characterized by strand invasion, in which the newly synthesized strand serves as a template for nascent strand synthesis, which avoids the lesion from replication machinery.77,78

FIG 3.

FIG 3

Schematic diagram representing main DNA damage tolerance pathways. Yellow color indicates the TLS pathways; Teal color indicates the homology-based damage tolerance in eukaryotic cells.

Mechanism of DNA damage tolerance initiation

The TS and TLS pathways share the same “substrate” at the DNA lesion site; stalling of replicative DNA polymerase along with the ongoing helicase activity at the replication fork results in the formation of ssDNA on the template strand to which the RPA binds. The ssDNA-RPA complex brings in the ATR interacting protein (ATRIP), which activates the ATR-dependent replication checkpoint.77,79 Simultaneously, the INO80 ATPase, a chromatin-remodeling protein, attaches to the stalled replication fork.80–82 Moreover, Rad6 is recruited by Rad18, resulting in the E2-E3 ubiquitinase complex formation at the lesion site that monoubiquitinates PCNA at K164.67,83–86 Apart from this, monoubiquitination may take place through different E3 ligases, such as ring finger protein domain 8 (RNF8) and ubiquitin-conjugating enzyme H5c (UbcH5c).87 At this point, the two DDT pathways separate, one with monoubiquitinated PCNA leading to TLS induction, and the second to further polyubiquitinate PCNA, resulting in homology-directed DDT.

Translesion synthesis

Subsequent to PCNA-K164 monoubiquitination and dissociation of the replicative polymerases, one or more TLS polymerases get recruited depending on the nature of the replication-blocking lesion (Fig. 4). For Y-family TLS polymerases, the conserved N-terminal domain consists of an active site that catalyzes lesion bypass, whereases the variable C-terminal region helps in the enrolment of protein to stalled forks.88,89 They can attach directly to K164-ubiquitinated PCNA by ubiquitin-binding motifs (UBM) found in Pol ι and Rev1 or UBZ present in Pol η and κ.89 TLS polymerases binding to PCNA is also facilitated through the BRCA1C-terminus (BRCT) domain located in the N-terminus of Rev1 and PIP boxes on Pol ι, η, and κ.50,89,90 The ubiquitin-binding domain (UBD) and PIP box mutations impair the damage-induced interaction of TLS polymerase with monoubiquitinated PCNA and their recruitment at the replication forks site [110, 111]. Generally, translesion synthesis is accomplished by the sequential action of two TLS polymerases. A distributive Y-family TLS polymerase inserts nucleotide(s) across the lesion site and then another TLS polymerase continues to extend the replicated DNA strand before it is replaced by replicative polymerase.88,90,91 In S. cerevisiae, during the REV1 mediated bypass, the lesion on the template strand is flipped into an extra-helical position and gets stabilized inside a hydrophobic pocket of Rev1, where it stays during the integration of incoming cytosine.92 Rev1 incorporates a single dCTP opposite to the lesion site.51,54,55 The R324 residue in the Rev1 side chain shifts the DNA lesion and acts as a substitutive template in Watson–Crick base pairing for incoming cytosine.92 Hydrogen bonding between R324 and cytosine is broken after the coupling of hydrolysis of pyrophosphate and phosphodiester bond. Rev1 is then separated from the DNA, and the lesion is reinserted in a double helix.92 Y family DNA polymerases lack 3′≥5′ proofreading exonuclease activity and have open active sites that can accommodate the altered bases.15,20,93 For example, in UV-induced cis-syn thymine-thymine CPD lesion bypass, the two covalently connected thymine bases can fit in Pol η active site.20,93 The protein domain in the β-strand in the LF provides a molecular splint that leads to the stabilization of newly synthesized dsDNA into a structure of B-form. It prevents CPD-induced framework shift and duplex distortion, which helps efficient and correct Pol η associated bypass of thymine-thymine CPDs.20,88,93

FIG 4.

FIG 4

Schematic diagram representing translesion DNA synthesis.

The accuracy of TLS polymerase in lesion bypass depends on the lesion type, and bypassing it seems to be at the cost of replication fidelity.94 For instance, Pol θ often inserts the correct base during replication across a 1,N6-ethenodeoxyadenosine lesion in human cells.95 In contrast, Pol θ also plays an essential role in the error-prone bypass of PP lesions and UV-induced cis-syn CPDs.26,95 The overall accuracy with which a TLS polymerase bypasses the lesion depends on many factors, such as the polymerases-lesion affinity, the biochemical properties of the individual TLS polymerases, and the sequence context of the lesion14,15,96 (Table 1).

Homology-directed DDT

As mentioned above, TS and FR are the two homology-directed DDT pathways that bypass the lesion in an error-free manner during S-phase (Fig. 3). The choice of error-free DDT pathways during lesion bypass is initiated by the polyubiquitination of the previously Rad6-Rad18-mediated monoubiquitinated PCNA, a process required in the signaling of TLS-mediated lesion bypass.97 In yeast, the polyubiquitination step requires the formation of a Mms2-Ubc13-Rad5 complex, whereas, in mammals, the polyubiquitination step is carried out by the complex of Ubc13-MMS2 and one of the two mammalian Rad5 homologs – SNF2 histone linker PHD RING helicase (SHPRH) and helicase-like transcription factor (HLTF).97,98 The direct binding of helicase protein SMARCAL1 to ssDNA removes the bound RPA, resulting in the remodeling of stalled replication fork to a “chicken-foot” structure in the FR pathway.99,100 Post RPA removal, translocase zinc-finger RANBP2-type containing 3 (ZRANB3) initiates further FR.69,101,102 The RAD51, BRCA1, and BRCA2 proteins help in the stabilization of reversed fork through which binding to the lagging and leading strand prevents the exonucleolytic degradation mediated by MRE11.103–105 Protein-DNA complex and the Fanconi anemia complementation group M (FANCM) helicase, when combined, lead to the formation of a four-way junction.106,107 After the successful lesion bypass, the original three-way junction formed by the regression of the reversed fork forms a four-way junction. This is catalyzed by RECQ-like helicase (WRN), DNA replication helicase/nuclease 2 (DNA2), and RecQ-like helicase (RECQ1).69,108,109

In the TS pathway, recruitment of exonuclease-1 (Exo1) is facilitated through the attachment of the 9-1-1 complex to the 5’ end of the gap on the nascent DNA strand.7,68 The Rad51-ssDNA presynaptic filament formed above the ssDNA portion of the template strand gets stabilized by Rad55/Rad57.110,111 The nucleofilament is disrupted by ATP-dependent DNA helicase Srs2 and restricts the action of Rad55/Rad57; the stability of Rad51-ssDNA presynaptic filament is determined by the balance between these processes.75,112,113 In sister chromatids, the Rad52 and Rad54 nucleofilament induce homology search and strand invasion.75,114 Following complementary base-pairing between the homologous template and the invading strand, DNA Pol δ gets recruited, which continues DNA replication and synthesizes the D-loop and later a sister-chromatid junction (SCJ).68,114,115 The Srs2 negatively correlates with the D-loop formation.78,116 Lastly, DNA topoisomerase 3 or RecQ-mediated genome instability protein 1 (Rmi1) or the slow growth suppressor 1 complex separates SCJ and regenerates the regular double-helical DNA structure.68,117,118

Regulation of DDT pathway choice

The selection of TLS or homology-directed DDT pathways is highly dependent on the type of PCNA ubiquitination.15,69,119 TLS pathway takes place when PCNA monoubiquitination occurs; however, polyubiquitination of PCNA leads to the initiation of homology-directed DDT. Regulation of ubiquitinated PCNA level is achieved via the enhanced level of genomic instability 1 (Elg1), ubiquitin-specific protease 7 (USP7), and USP1/upstream activation factor (UAF1) complex.119–122 After UV irradiation, USP1 proceeds toward inactivation through autocleavage, thereby upregulating the level of modified PCNA.123 It has been proposed that the type of PCNA ubiquitination depends on how long the replication arrest stays. Therefore, in promoting a switch to homology-directed DDT, extended replication occurs, resulting in the polyubiquitination of PCNA molecules that remain bound at the arrest site.119 The homology-directed DDT in mammals is first activated by the recruitment of HLTF along with the RAD6/RAD18 complex, leading to the instant polyubiquitination of PCNA.119 Apart from ubiquitination, PCNA experiences other related modifications. A previous study suggests that SUMOylation of PCNA at K164 mediated by protein inhibitors of STAT 1 and 4 (PIAS1 and PIAS4) encourages TS instead of TLS.124 When the TLS process completes, modification happens in monoubiquitinated PCNA via the accumulation of interferon-stimulated gene 15 (ISG15) molecules, resulting in USP10 recruitment and PCNA deubiquitination.125,126 Hence, a deep understanding of the connection between the choice of DNA damage tolerance and the PCNA modification pathway is a vital area for future study. Yeast RAD5 also interacts with REV1 to be involved in TLS and controls pathway choice.127

Regulation of TLS

Interaction between a RAD18 E3 ligase and PCNA accessory proteins with TLS polymerases maintains the regulation of TLS.9 However, there is some uncertainty about the role of terminal nucleotidyltransferase 4 A (TENT4A; formerly referred to as PAPD7), a noncanonical poly(A) polymerase in the TLS regulation. It was recently identified as another potential TLS regulator via its action on mRNA stability by controlling the poly(A) tail and/or Pol η and RAD18 translation.128 As discussed above, PCNA monoubiquitination by the RAD6/RAD18 complex is the key signal that allows the recruitment of TLS polymerases to the site of DNA damage. In addition, indirect regulation of RAD18 by TENT4A was also reported in the same study involving tumor suppressor gene CYLD and the long non-coding antisense RNA PAXIP1-AS2, suggesting the involvement of CYLD and PAXIP1-AS2 in the TLS regulation as well.128 In prokaryotes, the access of TLS polymerase Pol IV to the DNA was recently shown to be regulated by MutS, a mismatch repair protein that blocks the interaction of Pol IV with the clamp processivity factor, thus limiting the mutagenic impact of Pol IV during replication under normal conditions.129 Other factors involved in TLS regulation include HLTF, SHPRH, TIMELESS, SIVA1, CLASPIN, CHK1, and SprT-like domain at the N-terminus (SPARTAN), all having a crucial role in the monoubiquitination of PCNA.130–133 The differential role of HLTF and SHPRH in the recruitment of lesion-specific TLS polymerase Pol η (UV induced) and Pol κ (MMS induced), respectively132 highlights the complex yet essential facets of TLS regulation.

TLS polymerases regulate TLS pathways individually by undergoing post-translational modifications like SUMOylation, ubiquitination, and phosphorylation. Apart from ubiquitination, Pol η is also regulated through SUMOylation.134 Protein inhibitors of STAT1-mediated SUMOylation at K163 targets Pol η to the difficult-to-replicate genomic regions, including fragile sites in the absence of any exogenous DNA damage.135 After the lesion bypass is over, multiple lysine residues of Pol η are SUMOylated, thereby preventing it from interacting with ubiquitinated PCNA, resulting in SUMO-targeted ubiquitin ligase (STUbL) induced Pol η ubiquitination and its elimination from damage sites.135

Pol η is phosphorylated in the C-terminus involving multiple sites by protein kinases CDK2, PKC, and ATR. After DNA damage, ATR phosphorylates Pol η on S601, leaving it from sequestration by Pol δ-interacting protein of 38 kDa (PDIP38) and allowing Pol η to bind to the monoubiquitinated PCNA.131,136 This correlates with ATR activation via replication arrest-induced ssDNA, leading to TLS polymerase enrolment at the arrested replication fork.136 Furthermore, Pol η gets phosphorylated by PKC on S587 and T617, and by CDK2 on S687, leading to its stabilization in late S- and G2/M-phases.137,138

Regulation of TLS can also occur at the transcriptional level. After DNA damage, the expression of Pol η relies on p53,139 whereas the expression of Pol κ is governed by an aryl hydrocarbon receptor (AhR).140 A recent study reported that the function of Pumilio 1 (PUM1), an RNA-binding protein, is negatively correlated with TLS, as activation of TLS in response to DNA damage depends on the inhibition of PUM1-mediated mRNA decay.141

Regulation of homology-directed DDT

During fork reversal, the interplay between fork-protective and fork-degradative factors significantly contributes to the regulation of homology-directed repair.106 The MRE11 exonuclease degrades the nascent DNA strand at the reversed fork, which is protected by RAD51, BRCA1, and BRCA2.69,103,104 Subsequently, the DNA2-mediated fork degradation and structure-specific endonuclease subunit (SLX4)-facilitated fork cleavage is prevented by WRN helicase interacting protein 1 (WRNIP1).142,143 Progression of the fork slows down due to the interaction of ZRANB3 with polyubiquitinated PCNA, resulting in fork reversal by the ZRANB3 translocase activity.101,144 Furthermore, properly regulated DNA translocase SMARCAL1 is necessary to facilitate stalled fork repair and restart, and to ensure no aberrant fork progression takes place.144 This is promoted by the ATR-mediated phosphorylation of SMARCAL1 on S652, thereby restricting its fork regression activity and controlling its recruitment to the stalled fork site.142,145 Fork reversal and fork restart are modified by poly(ADP-ribose) polymerase 1 (PARP1) through the inhibition of RECQ1 helicase; further, it extends RECQ1-facilitated reversed forks to the three-way junction regression.74,108,142 Template switching is regulated at multiple points, including PCNA polyubiquitination, the formation of RAD51-ssDNA presynaptic filament, and SCJ formation.67 In human cells, SHPRH, HLTF, and INO80 chromatin remodeling proteins are essential for PCNA polyubiquitination.146 In the remodeling of chromatin, INO80 assists the assembly of K63-linked polyubiquitin chains to PCNA through SHRPH and HLTF.69,82,147,148 The Rad51-ssDNA presynaptic filament is disrupted by Srs2 and later participates in homology search;68,149–151 however, high mobility group protein 1 (HMO1), (EXO1, and INO80 promote the SCJ formation.68,82,151,152

NONCANONICAL FUNCTIONS OF TLS POLYMERASE

As discussed above, each TLS polymerase has certain known functions as part of recognized mechanisms. However, new functions and pathways mediated by TLS polymerases have emerged recently, linking these highly specialized polymerases with numerous cellular functions, including somatic hypermutation (SHM), DNA repair, and maintenance of genomic integrity in the absence of exogenous lesions, spindle checkpoint, cell cycle regulation, and fragile site maintenance.

SHM is a mutagenesis process that takes place in secondary lymphoid organs and produces B cells with the greatest affinity receptors for a particular antigen. This process has been linked to several TLS polymerases with Pol η and REV1 playing a major role.153 The frequency of mutations during SHM is a million times higher than the rest of the genome in the hypermutable region of the immunoglobulins around 2 kB in size.154 These mutations are mostly base substitutions that contribute to the diversity of antibodies. A well-defined SHM spectrum of C/G transversions, C/G transitions, and A/T mutation around the initial lesion has been reported by five mutagenic DNA damage response mechanisms, starting with the U as the original lesion: (1) replication opposite template U generates transitions at C/G; (2) UNG2-dependent TLS; (3) a hybrid pathway consisting of UNG2-dependent TLS and ncMMR; (4) ncMMR; and (5) PCNA ubiquitination and UNG2-dependent mutations at A/T [194]. Other TLS polymerases, such as Pol ζ [195,196], Pol θ,155–158 Pol ι,159,160 and Pol κ,159 have also been linked to SHM. However, it appears that these polymerases play a more supporting role. For instance, it was reported that Pol κ only contributed to the production of mutations when Pol η was completely absent.159

As Pol ζ can replicate beyond non-B DNA types that prevent the replicative polymerases from working, it also plays a crucial role in the genome maintenance. In mice, the absence of Pol ζ catalytic subunit Rev3L is embryonically lethal, and primary cells exhibit large-scale genomic instability as a result of an increase in spontaneous chromosomal translocations.13 Intriguingly, the codisruption of REV3L and POLH in chicken DT40 cells can correct this REV3L mutant-associated genomic instability and chromosomal aberrations, suggesting that Pol ζ and Pol η work together and that Pol η generates a toxic intermediate that requires Pol ζ to get resolved.161 In general, Rev1 and Pol ζ are both missing the ubiquitous hydrophobic motif that many proteins utilize to link with PCNA at its interdomain connecting loop. Pol ζ and Rev1 are distinctive to PCNA-interacting proteins that use the new binding surface close to the intermolecular interface of PCNA. The reported novel form of Rev1-PCNA binding raises the possibility of a mechanism by which Rev1 acquires a catalytically inactive conformation at the replication fork.162 Breakage at both inverted repeats and GAA/TTC repeats is exacerbated by DNA replication flaws. Increased fragility is linked to higher mutation levels in reporter genes located up to 8 kB away from both sides of the repeats. The existence of inverted or GAA/TTC repeats, as well as the activity of Pol ζ, influences increased mutagenesis.163

Fragile motifs cause it to create extended single-stranded areas in the disrupted chromosome, invasion of the undamaged sister chromatid for repair, and incorrect DNA synthesis using Pol ζ.164 The TLS-independent role of Rev3 in maintaining common fragile sites (CFSs) and the Rev3-independent, therefore TLS-independent roles of Rev7 in spindle formation, mitotic checkpoint, NER pathway, and many more roles of Pol ζ subunits again highlight the significance of these polymerases in the TLS-independent cellular functions.165–168

A previous study revealed that Pol η recruits DHX9 helicase to promote replication across guanine quadruplex structures.169 The DNA replication protein DHX9 was discovered to interact with the WRN helicase and PCNA, promoting the WRN activity in unraveling Okazaki fragment-like nucleic acid structures.170,171 Through its interaction with monoubiquitinated PCNA, Pol η is brought into DNA replication forks that are halted.9 Similarly, when replication stress occurs at CFSs, Pol η makes it easier for CFS loci to replicate, indicating its broader role. When Pol η is absent, CFS DNA sequences that match certain pause sites may result in non-B DNA structures.172

The involvement of Pol λ in TLS opposite photoproducts strongly argues that Pol λ would similarly insert a proper nucleotide from where polymerase would continue synthesis in the absence of numerous such additional DNA distorting lesions. Complex formation with Pol λ, such as Pol λ/Pol ζ during the error-free TLS opposite Tg and εdA damage, may allow Pol ζ to act in an error-free manner opposite additional DNA lesions when its structuring role is required.173 One study reported that the incorporation of A, G, or C opposite dA in mouse embryonic fibroblasts or human follicular lymphoma (HF) cells is coordinated by Tyr-2387 and Tyr-2391 via a Hoogsteen base pairing mechanism, as opposed to the actual role of Pol θ, which uses an AP-like mode to misincorporate A against εdA. Most significantly, only εdA adopting syn conformation at the Pol θ active site could account for 92% incorporation of T opposite εdA.174 Besides this, Pol θ-mediated end-joining (TMEJ) has been linked to chromosomal repair. TMEJ is an end-joining pathway uniquely capable of repairing substrates when conventionally established NHEJ is inefficient. The consequences of Pol θ deficiency indicate that Pol θ/TMEJ is necessary for the majority of what was previously characterized as MMEJ or Alt-NHEJ. A deficiency of Pol θ has only a moderate effect on the activity of other repair pathways (NHEJ and HR) in Cas9-induced DSBs.169–171

UBE2V2, the human homolog of MMS2, was reported to play a modest or redundant role in the Rad18-dependent DDT. However, UBE2V2 has been implicated in the abrogation of UV-induced gene conversion following antisense suppression of the transcript in HFs. To ascertain the underlying mechanism, it was demonstrated that DT40 cells do not exhibit considerable hypersensitivity to DNA damage or the enhanced sister chromatid exchange as seen in vertebrate Rad18 mutants.175 In addition, despite the crucial role Rad18 plays in DNA damage-induced TLS, neither UBE2V2 nor Rad18 in DT40 cells exhibited abnormally high levels of immunoglobulin gene conversion, and unlike Rev1, Rad18 in DT40 cells does not exhibit a defect in the nontemplated immunoglobulin gene mutation. Together, the data revealed that immunoglobulin diversification in DT40 cells is not dependent on the PCNA ubiquitination signaling.175 A study has shown that localization of Pol η to the chromatin is reduced in FANCD2- and Rad51-deficient cells and that Pol η knockdown alone leads to increased hydroxyurea (HU) sensitivity in the cells.176 The DNA damage response processes the “U” lesion in a mutagenic manner, producing a full range of base substitutions that define SHM at and near the initial “U” lesion.

Pol µ lengthens primers with 3'-terminal nucleotides opposite the abasic site. Most notably, this extension occurs through a mechanism of nucleotidyl transferase activity that is independent of the template sequence. This is not attributable to simple terminal nucleotidyl transferase activity, because, under typical conditions, Pol µ cannot add dNTPs to a blunt end duplex oligonucleotide or an oligo(dT)29 primer. Pol µ serves a dual mechanism of DNA-synthesizing enzyme that can function as a traditional DNA polymerase or a noncanonical, template-dependent but sequence-independent nucleotidyl transferase. This is the first time such a DNA-synthesizing enzyme has been described.177

Although TLS polymerases possess high accuracy for certain lesions, they have poor fidelity on undamaged DNA, suggesting their tight regulation in vivo to avoid mutagenesis. When dysregulated, some TLS polymerases confer a hypermutability state178–181 and alter replication fork progression.182–184 The loss of activity of some TLS polymerase may have a stronger effect on multicellular organisms than their upregulation.38,185,186 The early embryonic lethality in mice due to loss of Pol ζ,187–189 but not due to deletion of Pol κ or Pol ι highlights the distinctive roles some TLS have acquired which are not attributed to their TLS function. There might be a substantial genetic redundancy between TLS polymerases. For instance, mutant phenotype has been observed only upon simultaneous loss of more than one TLS polymerase. The majority of mutations are harmful, and organisms have a protective mechanism that keeps the mutational rate negligible.190 Even an increase in genetic discrepancy in the population may act beneficial under harmful or adverse conditions, as it sometimes results in a better variant that can withstand stress conditions.191,192 Furthermore, in response to stress, mutations that occur by TLS polymerase can serve as an important factor in the evolution of genetic variability. Consistent with this possibility, TLS polymerase induces adaptive mutagenesis in bacteria upon cellular stress.191,192 Moreover, the mutagenic ability of TLS polymerase has been hampered for SHM in higher eukaryotes, resulting in mutation in variable regions of antibodies synthesized by B cell lymphocytes.193 Thus, even though the strong deleterious mutagenic effect of TLS polymerases is rather disadvantageous, it may also provide many advantages to the cells.

TLS POLYMERASES AND HUMAN HEALTH

TLS activity mediated mutagenesis and human diseases

Selection of the wrong error-prone polymerase during lesion bypass or mutation(s) in the TLS polymerase genes that compromises their function often results in mutagenesis, a cause of many human diseases, including cancer. Several studies have shown that the levels of Pol ι are considerably higher in esophageal squamous cell carcinoma (ESCC) patients with lymph node metastasis than those without lymph node metastasis. The Kaplan–Meier method demonstrated a negative relationship between Pol ι expression and patient prognosis. In ESCC tissues, the expression levels of matrix metalloproteinase-2 (MMP-2) and matrix metalloproteinase-9 (MMP-9) were found to be favorably linked with Pol ι expression.194 Besides ESCC, Pol ι overexpression was correlated with breast cancer cells,195 and bladder cancer.196

In a recent study, Fanconi anemia cell lines were shown to upregulate Pol ι and rely on Y-family DNA polymerase for survival.197 The inactivation/impairment of alternative end joining (alt-EJ) by Pol θ depletion slows the development of pancreatic cancer and increases the longevity of experimental mice, but it does not prevent the formation of pancreatic ductal adenocarcinoma (PDAC), which results in full-blown PDAC with widespread metastases.198 In the initial S phase, poly(ADP-ribose) polymerase inhibitor (PARPi) causes ssDNA gaps behind replication forks in a PrimPol-dependent manner. Even though cells engage postreplication repair mechanisms to patch the gaps, gap repair cannot be completed in the presence of PARPi until the following S phase, resulting in DSBs in a trans-cell cycle manner. Poly(ADP-ribose) polymerase inhibitor has the unusual capacity to produce DSBs progressively in a trans-cell cycle way.199 BRCA1/2 loss exacerbates DSB accumulation spanning numerous cell cycles while also disrupting repair, making BRCA1/2-deficient cells particularly vulnerable to PARPi. This can be the potential basis for targeting cancer through PARPi.199 In primary hematopoietic cells, silencing of REV7 reduced progenitor activity, indicating that DNA repair deficiency is a major concern of bone marrow failure in Fanconi anemia.200 The overexpression of error-prone Pol θ which is involved in the DSB repair causes homologous recombination deficiency tumors.201 Pol θ-inserted errors are susceptible to error-prone microhomology-mediated end-joining of the DSB, which explains the genetic profile of BRCA-altered cancers.202 Most cancer types accumulate a unique mutation signature generated predominantly by the error-prone polymerases and other error-prone repair activities203 and to date more than 30 single-base substitution (SBS) mutation signatures have been found in various cancer types with four SBS signatures (SBS2, 5, 13 and 9) attribute to TLS polymerases.204 Apolipoprotein B MRNA editing catalytic polypeptides (APOBEC) family of proteins deaminate cytosine to uracil and its subsequent removal creates AP sites. The insertion of cytosines across these AP sites by Rev1 results in C > T or C > G mutations which are commonly observed in breast and bladder cancers.205–207 Pol η-dependent somatic hypermutation (SHM) activity is believed to generate the SBS9 mutation signature as seen in chronic lymphocytic leukemia (CLL) and malignant B-cell lymphomas and supported by an animal study;208 however, direct experimental evidence in support of this hypothesis is still missing.

The identification of cancer/tests antigen (CTA) melanoma antigen A4 (MAGE-A4) as a RAD18 stabilizer which in turn promotes PCNA monoubiquitination that can induce mutagenesis in cancer cells by favoring recruitment of TLS polymerases to the replisome.209,210 RAD18 is also seen overexpressed in the majority of the cancer types,204 thus hinting towards a link between the TLS pathway and cancer. An increased level of E3 ubiquitin ligase MDM2 seen in many cancer types causes depletion of Pol η, thereby increasing cancer predisposition similar to XPV syndrome.211 Besides these, another TLS influencer RNF168, an E3 ubiquitin ligase is frequently overexpressed in cancer cells.212 Although not experimentally validated, this evidence strongly supports the potential role of the TLS pathway in the initiation and progression of cancer, thus making the TLS polymerase a potential therapeutic target. In addition, by looking at the catastrophic connection between TLS and other DNA repair mechanisms could also help look forward to a therapeutic approach.

Apart from their role in cancer, TLS polymerases also have an extended role(s) in other comorbidities, such as the preservation of stem cells, and aging.213 It was found that Pol κ deficiency increases spontaneous mutations during aging in an organ-specific manner, whereas Pol η deficiency causes UV-induced mutagenesis in the skin.214,215 The premature aging of hematopoietic stem cells (HSCs), skewing of differentiation toward the myeloid/erythroid-associated multipotent progenitor 2 (MPP2) lineage at the expense of MPP4 lineage, and significantly higher DNA damage in lineage, Sca-1+, cKit+ (LSK) cells are all observed in DDT deficient PCNAK164R/K164R mice.216 Increased DNA damage in HSCs and skewing toward myeloid/erythroid progenitor lineages is an indicator of accelerated aging.216,217 Furthermore, ATR is a crucial enzyme of the DNA damage response involved in maintaining fork stability and controlling replication stress.218 In addition to greater dwarfism, higher replication stress, and hastened aging, the homozygous ATR-deficient mice do not show increased malignancy.219 In conclusion, abnormalities in DDT tolerance pathways can increase DNA damage and replication stress, which can speed up aging. It will be interesting to find out whether DDT has any impact on how quickly other tissues and their stem cells age. Studies reported that depletion of Pol κ affects the gap-filling DNA synthesis phase of the NER process in dorsal root ganglion (DRG) neurons, delaying the removal of cisplatin adducts and, consequently, the restart of RNA synthesis. While the expression of other DNA polymerases remained unaltered, Pol κ was shown to have a transcriptional upregulation.220 DNA Pol κ is thought to be a key component in cellular survival under harmful circumstances because it participates in the bypass synthesis of various DNA damage.14 Pol κ levels were decreased using siRNA to evaluate the effects of Pol κ deficiency on DRG neurons that had been treated with cisplatin. As evidenced by a lower incorporation of thymidine analog into nuclear DNA, Pol κ -targeting siRNA reduced the level of cisplatin-induced DNA repair synthesis and nuclear Pol κ immunoreactivity in DRG neurons.220 Moreover, cisplatin-induced global transcriptional inhibition in DRG neurons was significantly worsened by Pol κ deletion.220 Thus, Pol κ may be essential for repairing neuronal DNA damage and maintaining neuronal function and its loss might make DRG neurons more susceptible to genotoxic assaults. Since DRG neurons do not have strong blood–brain barrier protection, it is conceivable that Pol κ gives these neurons a stronger capability to deal with DNA damage, as the basal levels of Pol κ are higher in DRG neurons than in the cortical neurons.220,221 Therefore, Pol κ may be essential for the repair of neuronal DNA damage.

TLS polymerase gene mutations and human diseases

The link between TLS polymerase gene mutation(s) and human diseases dates back to the time when XPV syndrome was linked with a point mutation in the POLH. As mentioned earlier the elevated rates of UV-induced mutagenesis in XPV patients are ascribed to their high incidence of skin cancer. The mutated Pol η hampers the lesion bypass accuracy of UV-induced TT-dimers, causing mutations at TT sites and the development of skin cancer222 by activating compensatory, albeit error-prone TLS of CPD lesions by other “inserter” Y-family DNA polymerases. Since then several other mutations and SNPs have been identified in the TLS polymerase genes, with the majority showing significant association with cancer.16 The biallelic inactivating mutation in REV7 that codes for a mutant REV7-V85E protein was recently reported in an infant with acute marrow failure.200 When exposed to DNA crosslinking agents, cells derived from the patient showed an extended Fanconi anemia phenotype comprised of increased chromosomal breaks and G2/M accumulation, the buildup of γH2AX and 53BP1 foci, and higher p53/p21 activation.200 In the Han Chinese population, rs10077427 SNP of the POLK gene and rs462779 SNP of the REV3 gene were found significantly associated with breast cancer and colorectal cancer, respectively.223,224 Our group recently reported a significant association of three REV3 SNPs; rs1002481, rs462779, and rs465646 with nonsquamous cell lung cancer in the North Indian population.225 A study comparing individuals with a POLK C-C (rs5744533-rs5744724) haplotype to those with a POLK C-G haplotype revealed that the latter had a lower chance of developing lung cancer. The heterozygous condition of REV1 rs3087386 and rs3792136 were also an independent predictive factor for lung cancer survival, with hazard ratio (HR) values of 1.54 (95% CI: 1.12–2.12) and 1.44 (95% CI: 1.06–1.97), respectively.226 Interestingly, while REV7 mutations were connected to Fanconi anemia, mutations in the catalytic subunit REV3 in humans are autosomal dominantly responsible for the Mobius syndrome.227 Baring few studies, such as the association of Y89D mutation with myopia in a Chinese population,55 the primase and polymerase abnormalities caused by the PRIMPOL Y89D mutation in the DT40 B-cell lymphoma,228 and the Y100H mutation detected in various cancers.229 there is limited evidence to support the role of PRIMPOL in human diseases. The Y100H mutation promotes the incorporation of ribonucleotides (NTPs) at the expense of dNTPs and is believed to facilitate the survival of cells at an early stage of tumorigenesis.229

TLS activity and therapeutic drug resistance

The primary objective of TLS is to replicate preceding DNA damage brought on by first-line genotoxic drugs, which leads to decreasing effectiveness and the development of chemoresistance. Different tumor types exhibit different levels of relative expression of the Y family polymerase Pol κ. In colorectal and stomach cancers, for instance, the expression of Pol κ was found to decrease230,231 whereas, in lung and brain cancer samples the levels were found significantly elevated.232–234 Surprisingly, two-thirds of head and neck mucosal-derived squamous cell carcinoma (HNSCC) specimens had high expression of Pol η, in contrast to a decreased level in colorectal, lung, and stomach malignancies.235,236 While Pol η is unquestionably crucial for the emergence of platinum drug resistance, Rev1, and Pol ζ are also crucial for the pathways behind chemoresistance and mutagenesis in malignancies treated with genotoxic chemicals. In the case of cisplatin resistance, Rad18- and PCNA-mediated ubiquitination aids in coordinating REV1, Pol η, and Pol ζ bypass of intrastrand cross-links, while REV1 and Pol ζ work independently of PCNA-Ub to allow repair of interstrand cross-links brought on by cisplatin.237 Furthermore, the pace at which tumor cells develop cisplatin resistance is slowed down by the reduction of either REV1 or Pol ζ.238 In a murine model of Burkitt’s lymphoma, REV1 expression was downregulated, leading to a decrease in mutations and the development of cyclophosphamide resistance, tumor regression, and an improvement in overall survival.239 Additional research on TLS and chemoresistance has revealed a potential application for TLS inhibition in combination therapy with various other DNA-damaging agents. In the TGCT models mentioned above, REV7 depletion increased susceptibility to the DNA intercalator doxorubicin and the alkylating agent mitomycin C.240 The acquired resistance and tumor relapse that are frequently observed after first therapy are ultimately caused by cancer cells that survive first-line genotoxic medicines by showing enhanced mutation rates. TLS is thought to be crucial for this process, enabling malignancies to resist DNA-damaging chemicals.93,241

Strategies to enhance cancer therapy

TLS DNA polymerases are attractive targets for improving the efficacy of chemotherapy, and some DNA polymerase mutations, polymorphisms, and activity levels can serve as useful prognostic indicators for determining the best course of treatment for a variety of oncological conditions. Undoubtedly, a viable strategy for improving the efficiency and lowering cytotoxicity of chemotherapeutic medications is to find novel, precise, and efficient inhibitors of TLS polymerases and crucial protein-protein interactions of the translesome.242 Studies are now focusing on the development of TLS polymerase inhibitors which can hamper the exacerbation of the diseases including chemoresistance by targeting the TLS polymerases and their pathways. One study reported the identification of JH-RE-06, a small molecule that inhibits the recruitment of mutagenic Pol ζ.243 Interestingly, JH-RE-06 concentrates on a region of essentially featureless surface of REV1 where it interacts with the REV7 subunit of Pol ζ. When JH-RE-06 binds to REV1, REV1 dimerization is induced, which prevents the interaction between REV1 and REV7, thus failing to recruit Pol ζ to the DNA. In cultured human and mouse cell lines, JH-RE-06 suppresses mutagenic TLS and increases cisplatin-induced toxicity.244 The development of TLS inhibitors as a new class of chemotherapy adjuvants was made possible by the co-administration of JH-RE-06 and cisplatin, which inhibited the growth of xenograft human melanoma in mice.243 In another study, a small molecule inhibitor of REV1 UBM2 prevented the DDT with the treated cells showing delayed removal of UV-induced CPDs from nuclei, reduced UV-induced mutations of the HPRT gene, and reduced ability of cells exposed to cyclophosphamide or cisplatin to proliferate through clonogenic processes.245 Novobiocin, an aminocoumarin antibiotic also known as cathomycin or albamycin, has been demonstrated to bind directly to the Pol θ ATPase domain, reduce its ATPase function, and mimic Pol θ depletion.246 In genetically engineered mouse models, xenograft, and patient-derived xenograft models, novobiocin kills HR-deficient breast and ovarian cancers.246

The predetermined active sites of TLS polymerases that can readily accommodate a wide range of DNA lesions may overcome the damage caused by chemotherapy drugs. As mentioned earlier, TLS involves two-step processes with Y family polymerase proceeded by B family DNA polymerases in cancerous cells.247,248 Hence, by using short hairpin RNA or other target protein depletion strategies, it is possible to enhance the therapeutic effects of the chemotherapeutic drug treatment through the depletion of TLS polymerases.239,249–251 This further helps by not allowing drug resistance and reoccurrence of the cancerous cells. Based on the capacity to skip the damaged site it can lead to the prevention of collapsing of the replication fork leading to apoptosis. The wrong nucleotide incorporation can either lead to mutation postcell division or help cells to survive; hence, TLS is a key process by which cancer cells develop resistance to genotoxic treatment. Recent research reveals that DNA replication-associated single-stranded DNA (ssDNA) gaps cause PARPi toxicity in HR-deficient (HRD) cells. This susceptibility raises the possibility that the TLS pathway may serve as a novel therapeutic target for HRD tumors. In addition, the development of small molecule inhibitors of the TLS polymerase machinery has been identified as a therapeutic approach with promise for improving the efficacy of first-line chemotherapy and reducing cellular acquired resistance to DNA damaging drugs.

FUTURE PERSPECTIVES

The cumulative evidence suggests that inhibition of DDT, particularly the error-prone TLS pathway, can be used as a chemotherapeutic strategy for conventional treatment-resistant cancer cells. Current research on understanding the molecular mechanism of DNA-damage tolerance will create ideas and opportunities for further advancement in this area. Under the circumstances, an in-depth investigation is needed to understand the pathway choice and factors that influence it. In the TLS pathway, it is important to understand the individual role of TLS polymerases in bypassing the lesion in normal and disease cells. This will provide additional information for designing specific inhibitors and targeting TLS polymerases and accessory proteins.

Furthermore, agents that will directly block the protein function or interactions that are crucial for DDT can be effective therapeutic drugs. Tremendous attention has been given to the inhibition of TLS. However, with the recent understanding of genetics, fork reversal biochemistry, and template switching, there is an opportunity to identify inhibitors targeting these pathways as well. Additional challenges include the clinical utility of synthesized molecules and the development of target-specific inhibitory molecules that produce a cellular phenotypic response. Ongoing and future research gives a potential hope to target DDT pathways as a therapeutic approach for various diseases and associated complications.

CONCLUSION

In response to DNA damage, cells depend on the highly conserved DNA damage tolerance pathways, including translesion synthesis to prevent replication fork collapse and carry on DNA replication unhindered across the lesion-containing sites or at stalled replication forks. However, inequities in these pathways in neoplastic cells cause substantial mutagenesis, which may lead to chemoresistance and enhanced cancer cell survival. Inhibiting DDT pathways, especially the error-prone TLS pathway could therefore possibly sensitize cancer cells to existing chemotherapeutic agents and may overcome chemoresistance. Therefore, the design and characterization of DDT inhibitors might eventually lead to the development of a new class of cancer therapeutics that have the potential to improve the treatment efficacy of existing chemotherapeutic medicines, especially in chemoresistant cancers. In addition, understanding the noncanonical functions of TLS polymerases in human cells is also going to be an area of active research since many essential functions have already been attributed to TLS polymerases.

Funding Statement

A.B acknowledges the financial support from Indian Council of Medical Research (grant nos. 5/10/15/CAR-SMVDU/2018-RBMCH and 6719/2020-DDl/BMS) and A.D acknowledges the financial support from Indian Council of Medical Research (grant no. 6719/2020-DDl/BMS).

AUTHORS’ CONTRIBUTIONS

Conceptualization, A.B., T.K.P.; methodology, A.B., A.D., J.V., and G.L.; writing—original draft preparation, J.V., and G.L.; writing – review and editing, A.B., A.D., T.K.P., K.M.B and W.X.; visualization, A.B., and A.D.; supervision, A.B.; project administration, A.B. and A.D. All authors have read and agreed to the published version of the manuscript.

DATA AVAILABILITY STATEMENT

Data sharing not applicable – no new data generated.

REFERENCES

  • 1.Saini N, Giacobone CK, Klimczak LJ, Papas BN, Burkholder AB, Li JL, Fargo DC, Bai R, Gerrish K, Innes CL, et al. UV-exposure, endogenous DNA damage, and DNA replication errors shape the spectra of genome changes in human skin. PLoS Genet. 2021;17:e1009302. doi: 10.1371/journal.pgen.1009302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ciccia A, Elledge SJ.. The DNA damage response: making it safe to play with knives. Mol Cell. 2010;40:179–204. doi: 10.1016/j.molcel.2010.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Saha P, Mandal T, Talukdar AD, Kumar D, Kumar S, Tripathi PP, Wang QE, Srivastava AK.. DNA polymerase eta: a potential pharmacological target for cancer therapy. J Cell Physiol. 2021;236:4106–4120. doi: 10.1002/jcp.30155. [DOI] [PubMed] [Google Scholar]
  • 4.Jackson SP, Bartek J.. The DNA-damage response in human biology and disease. Nature. 2009;461:1071–1078. doi: 10.1038/nature08467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Lindahl T, Barnes DE.. Repair of endogenous DNA damage. Cold Spring Harb Symp Quant Biol. 2000;65:127–133. doi: 10.1101/sqb.2000.65.127. [DOI] [PubMed] [Google Scholar]
  • 6.Zhao L, Washington MT.. Translesion synthesis: insights into the selection and switching of DNA polymerases. Genes. 2017;8:24. doi: 10.3390/genes8010024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Lehmann AR, Niimi A, Ogi T, Brown S, Sabbioneda S, Wing JF, Kannouche PL, Green CM.. Translesion synthesis: Y-family polymerases and the polymerase switch. DNA Repair. 2007;6:891–899. doi: 10.1016/j.dnarep.2007.02.003. [DOI] [PubMed] [Google Scholar]
  • 8.Lupardus PJ, Byun T, Yee MC, Hekmat-Nejad M, Cimprich KA.. A requirement for replication in activation of the ATR-dependent DNA damage checkpoint. Genes Dev. 2002;16:2327–2332. doi: 10.1101/gad.1013502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Watanabe K, Tateishi S, Kawasuji M, Tsurimoto T, Inoue H, Yamaizumi M.. Rad18 guides poleta to replication stalling sites through physical interaction and PCNA monoubiquitination. Embo J. 2004;23:3886–3896. doi: 10.1038/sj.emboj.7600383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Byun TS, Pacek M, Yee MC, Walter JC, Cimprich KA.. Functional uncoupling of MCM helicase and DNA polymerase activities activates the ATR-dependent checkpoint. Genes Dev. 2005;19:1040–1052. doi: 10.1101/gad.1301205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Lee YS, Gregory MT, Yang W.. Human Pol zeta purified with accessory subunits is active in translesion DNA synthesis and complements Pol eta in cisplatin bypass. Proc Natl Acad Sci U S A. 2014;111:2954–2959. doi: 10.1073/pnas.1324001111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Sale JE. Translesion DNA synthesis and mutagenesis in eukaryotes. Cold Spring Harb Perspect Biol. 2013;5:a012708. doi: 10.1101/cshperspect.a012708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Gan GN, Wittschieben JP, Wittschieben BO, Wood RD.. DNA polymerase zeta (pol zeta) in higher eukaryotes. Cell Res. 2008;18:174–183. doi: 10.1038/cr.2007.117. [DOI] [PubMed] [Google Scholar]
  • 14.Chatterjee N, Walker GC.. Mechanisms of DNA damage, repair, and mutagenesis. Environ Mol Mutagen. 2017;58:235–263. doi: 10.1002/em.22087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.McCulloch SD, Kunkel TA.. The fidelity of DNA synthesis by eukaryotic replicative and translesion synthesis polymerases. Cell Res. 2008;18:148–161. doi: 10.1038/cr.2008.4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Makridakis NM, Reichardt JK.. Translesion DNA polymerases and cancer. Front Genet. 2012;3:174. doi: 10.3389/fgene.2012.00174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Goodman MF, Woodgate R.. Translesion DNA polymerases. Cold Spring Harb Perspect Biol. 2013;5:a010363. doi: 10.1101/cshperspect.a010363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Ohmori H, Friedberg EC, Fuchs RP, Goodman MF, Hanaoka F, Hinkle D, Kunkel TA, Lawrence CW, Livneh Z, Nohmi T, et al. The Y-family of DNA polymerases. Mol Cell. 2001;8:7–8. doi: 10.1016/s1097-2765(01)00278-7. [DOI] [PubMed] [Google Scholar]
  • 19.Nelson JR, Lawrence CW, Hinkle DC.. Thymine-thymine dimer bypass by yeast DNA polymerase zeta. Science. 1996;272:1646–1649. doi: 10.1126/science.272.5268.1646. [DOI] [PubMed] [Google Scholar]
  • 20.Biertumpfel C, Zhao Y, Kondo Y, Ramon-Maiques S, Gregory M, Lee JY, Masutani C, Lehmann AR, Hanaoka F, Yang W.. Structure and mechanism of human DNA polymerase eta. Nature. 2010;465:1044–1048. doi: 10.1038/nature09196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Uljon SN, Johnson RE, Edwards TA, Prakash S, Prakash L, Aggarwal AK.. Crystal structure of the catalytic core of human DNA polymerase kappa. Structure. 2004;12:1395–1404. doi: 10.1016/j.str.2004.05.011. [DOI] [PubMed] [Google Scholar]
  • 22.Lone S, Townson SA, Uljon SN, Johnson RE, Brahma A, Nair DT, Prakash S, Prakash L, Aggarwal AK.. Human DNA polymerase kappa encircles DNA: implications for mismatch extension and lesion bypass. Mol Cell. 2007;25:601–614. doi: 10.1016/j.molcel.2007.01.018. [DOI] [PubMed] [Google Scholar]
  • 23.Jha V, Ling H.. 2.0 A resolution crystal structure of human polkappa reveals a new catalytic function of N-clasp in DNA replication. Sci Rep. 2018;8:15125. doi: 10.1038/s41598-018-33371-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Ohmori H, Hanafusa T, Ohashi E, Vaziri C.. Separate roles of structured and unstructured regions of Y-family DNA polymerases. Adv Protein Chem Struct Biol. 2009;78:99–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Lancey C, Tehseen M, Bakshi S, Percival M, Takahashi M, Sobhy MA, Raducanu VS, Blair K, Muskett FW, Ragan TJ, et al. Cryo-EM structure of human Pol kappa bound to DNA and mono-ubiquitylated PCNA. Nat Commun. 2021;12:6095. doi: 10.1038/s41467-021-26251-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Yang W. An overview of Y-Family DNA polymerases and a case study of human DNA polymerase eta. Biochemistry. 2014;53:2793–2803. doi: 10.1021/bi500019s. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Swan MK, Johnson RE, Prakash L, Prakash S, Aggarwal AK.. Structure of the human Rev1-DNA-dNTP ternary complex. J Mol Biol. 2009;390:699–709. doi: 10.1016/j.jmb.2009.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Frank EG, McLenigan MP, McDonald JP, Huston D, Mead S, Woodgate R.. DNA polymerase iota: the long and the short of it! DNA Repair. 2017;58:47–51. doi: 10.1016/j.dnarep.2017.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.McIntyre J. Polymerase iota - an odd sibling among Y family polymerases. DNA Repair. 2020;86:102753. doi: 10.1016/j.dnarep.2019.102753. [DOI] [PubMed] [Google Scholar]
  • 30.Yang W, Woodgate R.. What a difference a decade makes: insights into translesion DNA synthesis. Proc Natl Acad Sci U S A. 2007;104:15591–15598. doi: 10.1073/pnas.0704219104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Nair DT, Johnson RE, Prakash S, Prakash L, Aggarwal AK.. Replication by human DNA polymerase-iota occurs by Hoogsteen base-pairing. Nature. 2004;430:377–380. doi: 10.1038/nature02692. [DOI] [PubMed] [Google Scholar]
  • 32.Nair DT, Johnson RE, Prakash L, Prakash S, Aggarwal AK.. An incoming nucleotide imposes an anti to syn conformational change on the templating purine in the human DNA polymerase-iota active site. Structure. 2006;14:749–755. doi: 10.1016/j.str.2006.01.010. [DOI] [PubMed] [Google Scholar]
  • 33.Frank EG, Woodgate R.. Increased catalytic activity and altered fidelity of human DNA polymerase iota in the presence of manganese. J Biol Chem. 2007;282:24689–24696. doi: 10.1074/jbc.M702159200. [DOI] [PubMed] [Google Scholar]
  • 34.Choi JY, Patra A, Yeom M, Lee YS, Zhang Q, Egli M, Guengerich FP.. Kinetic and structural impact of metal ions and genetic variations on human DNA polymerase iota. J Biol Chem. 2016;291:21063–21073. doi: 10.1074/jbc.M116.748285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bebenek K, Tissier A, Frank EG, McDonald JP, Prasad R, Wilson SH, Woodgate R, Kunkel TA.. 5'-Deoxyribose phosphate lyase activity of human DNA polymerase iota in vitro. Science. 2001;291:2156–2159. doi: 10.1126/science.1058386. [DOI] [PubMed] [Google Scholar]
  • 36.Sobol RW, Prasad R, Evenski A, Baker A, Yang XP, Horton JK, Wilson SH.. The lyase activity of the DNA repair protein beta-polymerase protects from DNA-damage-induced cytotoxicity. Nature. 2000;405:807–810. doi: 10.1038/35015598. [DOI] [PubMed] [Google Scholar]
  • 37.Burtis KC, Harris PV.. A possible functional role for a new class of eukaryotic DNA polymerases. Curr Biol. 1997;7:R743–744. doi: 10.1016/s0960-9822(06)00391-5. [DOI] [PubMed] [Google Scholar]
  • 38.Masutani C, Kusumoto R, Yamada A, Dohmae N, Yokoi M, Yuasa M, Araki M, Iwai S, Takio K, Hanaoka F.. The XPV (xeroderma pigmentosum variant) gene encodes human DNA polymerase eta. Nature. 1999;399:700–704. doi: 10.1038/21447. [DOI] [PubMed] [Google Scholar]
  • 39.Burgers PM, Koonin EV, Bruford E, Blanco L, Burtis KC, Christman MF, Copeland WC, Friedberg EC, Hanaoka F, Hinkle DC, et al. Eukaryotic DNA polymerases: proposal for a revised nomenclature. J Biol Chem. 2001;276:43487–43490. doi: 10.1074/jbc.R100056200. [DOI] [PubMed] [Google Scholar]
  • 40.Thakur M, Wernick M, Collins C, Limoli CL, Crowley E, Cleaver JE.. DNA polymerase eta undergoes alternative splicing, protects against UV sensitivity and apoptosis, and suppresses Mre11-dependent recombination. Genes Chromosomes Cancer. 2001;32:222–235. doi: 10.1002/gcc.1186. [DOI] [PubMed] [Google Scholar]
  • 41.Trincao J, Johnson RE, Escalante CR, Prakash S, Prakash L, Aggarwal AK.. Structure of the catalytic core of S. cerevisiae DNA polymerase eta: implications for translesion DNA synthesis. Mol Cell. 2001;8:417–426. doi: 10.1016/s1097-2765(01)00306-9. [DOI] [PubMed] [Google Scholar]
  • 42.Prakash S, Johnson RE, Prakash L.. Eukaryotic translesion synthesis DNA polymerases: specificity of structure and function. Annu Rev Biochem. 2005;74:317–353. doi: 10.1146/annurev.biochem.74.082803.133250. [DOI] [PubMed] [Google Scholar]
  • 43.Supek F, Lehner B.. Clustered mutation signatures reveal that error-prone DNA repair targets mutations to active genes. Cell. 2017;170:534–547 e523. doi: 10.1016/j.cell.2017.07.003. [DOI] [PubMed] [Google Scholar]
  • 44.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. Mol Cell. 2011;43:649–662. doi: 10.1016/j.molcel.2011.06.023. [DOI] [PubMed] [Google Scholar]
  • 45.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. Mol Cell. 2012;47:669–680. doi: 10.1016/j.molcel.2012.07.006. [DOI] [PubMed] [Google Scholar]
  • 46.Lawrence CW, Das G, Christensen RB.. REV7, a new gene concerned with UV mutagenesis in yeast. Mol Gen Genet. 1985;200:80–85. doi: 10.1007/BF00383316. [DOI] [PubMed] [Google Scholar]
  • 47.Murakumo Y, Roth T, Ishii H, Rasio D, Numata S, Croce CM, Fishel R.. A human REV7 homolog that interacts with the polymerase zeta catalytic subunit hREV3 and the spindle assembly checkpoint protein hMAD2. J Biol Chem. 2000;275:4391–4397. doi: 10.1074/jbc.275.6.4391. [DOI] [PubMed] [Google Scholar]
  • 48.Durkin SG, Glover TW.. Chromosome fragile sites. Annu Rev Genet. 2007;41:169–192. doi: 10.1146/annurev.genet.41.042007.165900. [DOI] [PubMed] [Google Scholar]
  • 49. Malik R, Kopylov M, Gomez-Llorente Y, Jain R, Johnson RE, Prakash L, Prakash S, Ubarretxena-Belandia I, Aggarwal AK.. Structure and mechanism of B-family DNA polymerase zeta specialized for translesion DNA synthesis. Nat Struct Mol Biol. 2020;27:913–924. doi: 10.1038/s41594-020-0476-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Pustovalova Y, Maciejewski MW, Korzhnev DM.. NMR mapping of PCNA interaction with translesion synthesis DNA polymerase Rev1 mediated by Rev1-BRCT domain. J Mol Biol. 2013;425:3091–3105. doi: 10.1016/j.jmb.2013.05.029. [DOI] [PubMed] [Google Scholar]
  • 51.Iyer LM, Koonin EV, Leipe DD, Aravind L.. Origin and evolution of the archaeo-eukaryotic primase superfamily and related palm-domain proteins: structural insights and new members. Nucleic Acids Res. 2005;33:3875–3896. doi: 10.1093/nar/gki702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Bianchi J, Rudd SG, Jozwiakowski SK, Bailey LJ, Soura V, Taylor E, Stevanovic I, Green AJ, Stracker TH, Lindsay HD, et al. PrimPol bypasses UV photoproducts during eukaryotic chromosomal DNA replication. Mol Cell. 2013;52:566–573. doi: 10.1016/j.molcel.2013.10.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Garcia-Gomez S, Reyes A, Martinez-Jimenez MI, Chocron ES, Mouron S, Terrados G, Powell C, Salido E, Mendez J, Holt IJ, et al. PrimPol, an archaic primase/polymerase operating in human cells. Mol Cell. 2013;52:541–553. doi: 10.1016/j.molcel.2013.09.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Rechkoblit O, Gupta YK, Malik R, Rajashankar KR, Johnson RE, Prakash L, Prakash S, Aggarwal AK.. Structure and mechanism of human PrimPol, a DNA polymerase with primase activity. Sci Adv. 2016;2:e1601317. doi: 10.1126/sciadv.1601317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Keen BA, Bailey LJ, Jozwiakowski SK, Doherty AJ.. Human PrimPol mutation associated with high myopia has a DNA replication defect. Nucleic Acids Res. 2014;42:12102–12111. doi: 10.1093/nar/gku879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Keen BA, Jozwiakowski SK, Bailey LJ, Bianchi J, Doherty AJ.. Molecular dissection of the domain architecture and catalytic activities of human PrimPol. Nucleic Acids Res. 2014;42:5830–5845. doi: 10.1093/nar/gku214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Guilliam TA, Doherty AJ.. PrimPol-Prime time to reprime. Genes. 2017;8:20. doi: 10.3390/genes8010020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Lang T, Maitra M, Starcevic D, Li SX, Sweasy JB.. A DNA polymerase beta mutant from colon cancer cells induces mutations. Proc Natl Acad Sci U S A. 2004;101:6074–6079. doi: 10.1073/pnas.0308571101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Van Goethem G, Schwartz M, Lofgren A, Dermaut B, Van Broeckhoven C, Vissing J.. Novel POLG mutations in progressive external ophthalmoplegia mimicking mitochondrial neurogastrointestinal encephalomyopathy. Eur J Hum Genet. 2003;11:547–549. doi: 10.1038/sj.ejhg.5201002. [DOI] [PubMed] [Google Scholar]
  • 60.Johnson RE, Kondratick CM, Prakash S, Prakash L.. hRAD30 mutations in the variant form of xeroderma pigmentosum. Science. 1999;285:263–265. doi: 10.1126/science.285.5425.263. [DOI] [PubMed] [Google Scholar]
  • 61.Ikeda M, Furukohri A, Philippin G, Loechler E, Akiyama MT, Katayama T, Fuchs RP, Maki H.. DNA polymerase IV mediates efficient and quick recovery of replication forks stalled at N2-dG adducts. Nucleic Acids Res. 2014;42:8461–8472. doi: 10.1093/nar/gku547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Kunkel TA. DNA replication fidelity. J Biol Chem. 2004;279:16895–16898. doi: 10.1074/jbc.R400006200. [DOI] [PubMed] [Google Scholar]
  • 63.Friedberg EC, Wagner R, Radman M.. Specialized DNA polymerases, cellular survival, and the genesis of mutations. Science. 2002;296:1627–1630. doi: 10.1126/science.1070236. [DOI] [PubMed] [Google Scholar]
  • 64.Goodman MF. Error-prone repair DNA polymerases in prokaryotes and eukaryotes. Annu Rev Biochem. 2002;71:17–50. doi: 10.1146/annurev.biochem.71.083101.124707. [DOI] [PubMed] [Google Scholar]
  • 65.Yang W. Portraits of a Y-family DNA polymerase. FEBS Lett. 2005;579:868–872. doi: 10.1016/j.febslet.2004.11.047. [DOI] [PubMed] [Google Scholar]
  • 66.Yang W. Damage repair DNA polymerases Y. Curr Opin Struct Biol. 2003;13:23–30. doi: 10.1016/s0959-440x(02)00003-9. [DOI] [PubMed] [Google Scholar]
  • 67.Nair DT, Johnson RE, Prakash L, Prakash S, Aggarwal AK.. Rev1 employs a novel mechanism of DNA synthesis using a protein template. Science. 2005;309:2219–2222. doi: 10.1126/science.1116336. [DOI] [PubMed] [Google Scholar]
  • 68.Bi X. Mechanism of DNA damage tolerance. World J Biol Chem. 2015;6:48–56. doi: 10.4331/wjbc.v6.i3.48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Pilzecker B, Buoninfante OA, Jacobs H.. DNA damage tolerance in stem cells, ageing, mutagenesis, disease and cancer therapy. Nucleic Acids Res. 2019;47:7163–7181. doi: 10.1093/nar/gkz531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Makarova AV, Burgers PM.. Eukaryotic DNA polymerase zeta. DNA Repair. 2015;29:47–55. doi: 10.1016/j.dnarep.2015.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Basu A, Broyde S, Iwai S, Kisker C.. DNA damage, mutagenesis, and DNA repair. J Nucleic Acids. 2010;2010:182894. doi: 10.4061/2010/182894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Knobel PA, Marti TM.. Translesion DNA synthesis in the context of cancer research. Cancer Cell Int. 2011;11:39. doi: 10.1186/1475-2867-11-39. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73. Sogo JM, Lopes M, Foiani M.. Fork reversal and ssDNA accumulation at stalled replication forks owing to checkpoint defects. Science. 2002;297:599–602. doi: 10.1126/science.1074023. [DOI] [PubMed] [Google Scholar]
  • 74.Ray Chaudhuri A, Hashimoto Y, Herrador R, Neelsen KJ, Fachinetti D, Bermejo R, Cocito A, Costanzo V, Lopes M.. Topoisomerase I poisoning results in PARP-mediated replication fork reversal. Nat Struct Mol Biol. 2012;19:417–423. doi: 10.1038/nsmb.2258. [DOI] [PubMed] [Google Scholar]
  • 75.Giannattasio M, Zwicky K, Follonier C, Foiani M, Lopes M, Branzei D.. Visualization of recombination-mediated damage bypass by template switching. Nat Struct Mol Biol. 2014;21:884–892. doi: 10.1038/nsmb.2888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Fumasoni M, Zwicky K, Vanoli F, Lopes M, Branzei D.. Error-free DNA damage tolerance and sister chromatid proximity during DNA replication rely on the Polalpha/Primase/Ctf4 Complex. Mol Cell. 2015;57:812–823. doi: 10.1016/j.molcel.2014.12.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Binz SK, Sheehan AM, Wold MS.. Replication protein A phosphorylation and the cellular response to DNA damage. DNA Repair. 2004;3:1015–1024. doi: 10.1016/j.dnarep.2004.03.028. [DOI] [PubMed] [Google Scholar]
  • 78.Branzei D, Psakhye I.. DNA damage tolerance. Curr Opin Cell Biol. 2016;40:137–144. doi: 10.1016/j.ceb.2016.03.015. [DOI] [PubMed] [Google Scholar]
  • 79.Unsal-Kacmaz K, Makhov AM, Griffith JD, Sancar A.. Preferential binding of ATR protein to UV-damaged DNA. Proc Natl Acad Sci U S A. 2002;99:6673–6678. doi: 10.1073/pnas.102167799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Papamichos-Chronakis M, Peterson CL.. The Ino80 chromatin-remodeling enzyme regulates replisome function and stability. Nat Struct Mol Biol. 2008;15:338–345. doi: 10.1038/nsmb.1413. [DOI] [PubMed] [Google Scholar]
  • 81.Shimada K, Oma Y, Schleker T, Kugou K, Ohta K, Harata M, Gasser SM.. Ino80 chromatin remodeling complex promotes recovery of stalled replication forks. Curr Biol. 2008;18:566–575. doi: 10.1016/j.cub.2008.03.049. [DOI] [PubMed] [Google Scholar]
  • 82.Falbo KB, Alabert C, Katou Y, Wu S, Han J, Wehr T, Xiao J, He X, Zhang Z, Shi Y, et al. Involvement of a chromatin remodeling complex in damage tolerance during DNA replication. Nat Struct Mol Biol. 2009;16:1167–1172. doi: 10.1038/nsmb.1686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Bailly V, Lauder S, Prakash S, Prakash L.. Yeast DNA repair proteins Rad6 and Rad18 form a heterodimer that has ubiquitin conjugating, DNA binding, and ATP hydrolytic activities. J Biol Chem. 1997;272:23360–23365. doi: 10.1074/jbc.272.37.23360. [DOI] [PubMed] [Google Scholar]
  • 84.Ma X, Dong L, Liu X, Ou K, Yang. POLE/POLD1 mutation and tumor immunotherapy. J Exp Coin Cancer Res. 2022;216. doi: 10.1186/s13046-022-024221-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Yoon JH, Prakash S, Prakash L.. Requirement of Rad18 protein for replication through DNA lesions in mouse and human cells. Proc Natl Acad Sci U S A. 2012;109:7799–7804. doi: 10.1073/pnas.1204105109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hedglin M, Benkovic SJ.. Regulation of Rad6/Rad18 activity during DNA damage tolerance. Annu Rev Biophys. 2015;44:207–228. doi: 10.1146/annurev-biophys-060414-033841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Zhang S, Chea J, Meng X, Zhou Y, Lee EY, Lee MY.. PCNA is ubiquitinated by RNF8. Cell Cycle. 2008;7:3399–3404. doi: 10.4161/cc.7.21.6949. [DOI] [PubMed] [Google Scholar]
  • 88. Yang W, Gao Y.. Translesion and repair DNA polymerases: diverse structure and mechanism. Annu Rev Biochem. 2018;87:239–261. doi: 10.1146/annurev-biochem-062917-012405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Ma X, Tang TS, Guo C.. Regulation of translesion DNA synthesis in mammalian cells. Environ Mol Mutagen. 2020;61:680–692. doi: 10.1002/em.22359. [DOI] [PubMed] [Google Scholar]
  • 90.Guo C, Sonoda E, Tang TS, Parker JL, Bielen AB, Takeda S, Ulrich HD, Friedberg EC.. REV1 protein interacts with PCNA: significance of the REV1 BRCT domain in vitro and in vivo. Mol Cell. 2006;23:265–271. doi: 10.1016/j.molcel.2006.05.038. [DOI] [PubMed] [Google Scholar]
  • 91.Quinet A, Lerner LK, Martins DJ, Menck CFM.. Filling gaps in translesion DNA synthesis in human cells. Mutat Res Genet Toxicol Environ Mutagen. 2018;836:127–142. doi: 10.1016/j.mrgentox.2018.02.004. [DOI] [PubMed] [Google Scholar]
  • 92.Weaver TM, Cortez LM, Khoang TH, Washington MT, Agarwal PK, Freudenthal BD.. Visualizing Rev1 catalyze protein-template DNA synthesis. Proc Natl Acad Sci U S A. 2020;117:25494–25504. doi: 10.1073/pnas.2010484117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Sale JE, Lehmann AR, Woodgate R.. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nat Rev Mol Cell Biol. 2012;13:141–152. doi: 10.1038/nrm3289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Temprine K, Campbell NR, Huang R, Langdon EM, Simon-Vermot T, Mehta K, Clapp A, Chipman M, White RM.. Regulation of the error-prone DNA polymerase Polkappa by oncogenic signaling and its contribution to drug resistance. Sci Signal. 2020;13:eaau1453. doi: 10.1126/scisignal.aau1453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Yoon JH, McArthur MJ, Park J, Basu D, Wakamiya M, Prakash L, Prakash S.. Error-prone replication through UV lesions by DNA polymerase theta protects against skin cancers. Cell. 2019;176:1295–1309 e1215. doi: 10.1016/j.cell.2019.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Wang Z, Xiao W.. Distinct requirements for budding yeast Rev1 and Poleta in translesion DNA synthesis across different types of DNA damage. Curr Genet. 2020;66:1019–1028. doi: 10.1007/s00294-020-01092-w. [DOI] [PubMed] [Google Scholar]
  • 97.Xu X, Blackwell S, Lin A, Li F, Qin Z, Xiao W.. Error-free DNA-damage tolerance in Saccharomyces cerevisiae. Mutat Res Rev Mutat Res. 2015;764:43–50. doi: 10.1016/j.mrrev.2015.02.001. [DOI] [PubMed] [Google Scholar]
  • 98.Motegi A, Liaw HJ, Lee KY, Roest HP, Maas A, Wu X, Moinova H, Markowitz SD, Ding H, Hoeijmakers JH, et al. Polyubiquitination of proliferating cell nuclear antigen by HLTF and SHPRH prevents genomic instability from stalled replication forks. Proc Natl Acad Sci U S A. 2008;105:12411–12416. doi: 10.1073/pnas.0805685105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Yusufzai T, Kadonaga JT.. HARP is an ATP-driven annealing helicase. Science. 2008;322:748–750. doi: 10.1126/science.1161233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Poole LA, Cortez D.. Functions of SMARCAL1, ZRANB3, and HLTF in maintaining genome stability. Crit Rev Biochem Mol Biol. 2017;52:696–714. doi: 10.1080/10409238.2017.1380597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Vujanovic M, Krietsch J, Raso MC, Terraneo N, Zellweger R, Schmid JA, Taglialatela A, Huang JW, Holland CL, Zwicky K, et al. Replication fork slowing and reversal upon DNA damage require PCNA polyubiquitination and ZRANB3 DNA translocase activity. Mol Cell. 2017;67:882–890 e885. doi: 10.1016/j.molcel.2017.08.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Tian T, Bu M, Chen X, Ding L, Yang Y, Han J, Feng XH, Xu P, Liu T, Ying S, et al. The ZATT-TOP2A-PICH axis drives extensive replication fork reversal to promote genome stability. Mol Cell. 2021;81:198–211 e196. doi: 10.1016/j.molcel.2020.11.007. [DOI] [PubMed] [Google Scholar]
  • 103.Kolinjivadi AM, Sannino V, De Antoni A, Zadorozhny K, Kilkenny M, Techer H, Baldi G, Shen R, Ciccia A, Pellegrini L, et al. Smarcal1-mediated fork reversal triggers Mre11-dependent degradation of nascent DNA in the absence of Brca2 and stable Rad51 nucleofilaments. Mol Cell. 2017;67:867–881 e867. doi: 10.1016/j.molcel.2017.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Lemacon D, Jackson J, Quinet A, Brickner JR, Li S, Yazinski S, You Z, Ira G, Zou L, Mosammaparast N, et al. MRE11 and EXO1 nucleases degrade reversed forks and elicit MUS81-dependent fork rescue in BRCA2-deficient cells. Nat Commun. 2017;8:860. doi: 10.1038/s41467-017-01180-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Mason JM, Chan YL, Weichselbaum RW, Bishop DK.. Non-enzymatic roles of human RAD51 at stalled replication forks. Nat Commun. 2019;10:4410. doi: 10.1038/s41467-019-12297-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gari K, Decaillet C, Delannoy M, Wu L, Constantinou A.. Remodeling of DNA replication structures by the branch point translocase FANCM. Proc Natl Acad Sci U S A. 2008;105:16107–16112. doi: 10.1073/pnas.0804777105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Gari K, Decaillet C, Stasiak AZ, Stasiak A, Constantinou A.. The Fanconi anemia protein FANCM can promote branch migration of Holliday junctions and replication forks. Mol Cell. 2008;29:141–148. doi: 10.1016/j.molcel.2007.11.032. [DOI] [PubMed] [Google Scholar]
  • 108.Berti M, Ray Chaudhuri A, Thangavel S, Gomathinayagam S, Kenig S, Vujanovic M, Odreman F, Glatter T, Graziano S, Mendoza-Maldonado R, et al. Human RECQ1 promotes restart of replication forks reversed by DNA topoisomerase I inhibition. Nat Struct Mol Biol. 2013;20:347–354. doi: 10.1038/nsmb.2501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Quinet A, Lemacon D, Vindigni A.. Replication fork reversal: players and guardians. Mol Cell. 2017;68:830–833. doi: 10.1016/j.molcel.2017.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Liu J, Renault L, Veaute X, Fabre F, Stahlberg H, Heyer WD.. Rad51 paralogues Rad55-Rad57 balance the antirecombinase Srs2 in Rad51 filament formation. Nature. 2011;479:245–248. doi: 10.1038/nature10522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Godin S, Wier A, Kabbinavar F, Bratton-Palmer DS, Ghodke H, Van Houten B, VanDemark AP, Bernstein KA.. The Shu complex interacts with Rad51 through the Rad51 paralogues Rad55-Rad57 to mediate error-free recombination. Nucleic Acids Res. 2013;41:4525–4534. doi: 10.1093/nar/gkt138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112. Bernstein KA, Reid RJ, Sunjevaric I, Demuth K, Burgess RC, Rothstein R.. The Shu complex, which contains Rad51 paralogues, promotes DNA repair through inhibition of the Srs2 anti-recombinase. Mol Biol Cell. 2011;22:1599–1607. doi: 10.1091/mbc.E10-08-0691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Sasanuma H, Tawaramoto MS, Lao JP, Hosaka H, Sanda E, Suzuki M, Yamashita E, Hunter N, Shinohara M, Nakagawa A, et al. A new protein complex promoting the assembly of Rad51 filaments. Nat Commun. 2013;4:1676. doi: 10.1038/ncomms2678. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Vanoli F, Fumasoni M, Szakal B, Maloisel L, Branzei D.. Replication and recombination factors contributing to recombination-dependent bypass of DNA lesions by template switch. PLoS Genet. 2010;6:e1001205. doi: 10.1371/journal.pgen.1001205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.San Filippo J, Sung P, Klein H.. Mechanism of eukaryotic homologous recombination. Annu Rev Biochem. 2008;77:229–257. doi: 10.1146/annurev.biochem.77.061306.125255. [DOI] [PubMed] [Google Scholar]
  • 116.Burkovics P, Sebesta M, Sisakova A, Plault N, Szukacsov V, Robert T, Pinter L, Marini V, Kolesar P, Haracska L, et al. Srs2 mediates PCNA-SUMO-dependent inhibition of DNA repair synthesis. Embo J. 2013;32:742–755. doi: 10.1038/emboj.2013.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Bernstein KA, Shor E, Sunjevaric I, Fumasoni M, Burgess RC, Foiani M, Branzei D, Rothstein R.. Sgs1 function in the repair of DNA replication intermediates is separable from its role in homologous recombinational repair. Embo J. 2009;28:915–925. doi: 10.1038/emboj.2009.28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Cejka P, Plank JL, Dombrowski CC, Kowalczykowski SC.. Decatenation of DNA by the S. cerevisiae Sgs1-Top3-Rmi1 and RPA complex: a mechanism for disentangling chromosomes. Mol Cell. 2012;47:886–896. doi: 10.1016/j.molcel.2012.06.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119. Masuda Y, Masutani C.. Spatiotemporal regulation of PCNA ubiquitination in damage tolerance pathways. Crit Rev Biochem Mol Biol. 2019;54:418–442. doi: 10.1080/10409238.2019.1687420. [DOI] [PubMed] [Google Scholar]
  • 120.Cohn MA, Kowal P, Yang K, Haas W, Huang TT, Gygi SP, D'Andrea AD.. A UAF1-containing multisubunit protein complex regulates the Fanconi anemia pathway. Mol Cell. 2007;28:786–797. doi: 10.1016/j.molcel.2007.09.031. [DOI] [PubMed] [Google Scholar]
  • 121.Lee KY, Yang K, Cohn MA, Sikdar N, D'Andrea AD, Myung K.. Human ELG1 regulates the level of ubiquitinated proliferating cell nuclear antigen (PCNA) through Its interactions with PCNA and USP1. J Biol Chem. 2010;285:10362–10369. doi: 10.1074/jbc.M109.092544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Kashiwaba S, Kanao R, Masuda Y, Kusumoto-Matsuo R, Hanaoka F, Masutani C.. USP7 Is a suppressor of PCNA ubiquitination and oxidative-stress-induced mutagenesis in human cells. Cell Rep. 2015;13:2072–2080. doi: 10.1016/j.celrep.2015.11.014. [DOI] [PubMed] [Google Scholar]
  • 123.Huang TT, Nijman SM, Mirchandani KD, Galardy PJ, Cohn MA, Haas W, Gygi SP, Ploegh HL, Bernards R, D'Andrea AD.. Regulation of monoubiquitinated PCNA by DUB autocleavage. Nat Cell Biol. 2006;8:339–347. doi: 10.1038/ncb1378. [DOI] [PubMed] [Google Scholar]
  • 124.Mohiuddin M, Evans TJ, Rahman MM, Keka IS, Tsuda M, Sasanuma H, Takeda S.. SUMOylation of PCNA by PIAS1 and PIAS4 promotes template switch in the chicken and human B cell lines. Proc Natl Acad Sci U S A. 2018;115:12793–12798. doi: 10.1073/pnas.1716349115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125. Park JM, Yang SW, Yu KR, Ka SH, Lee SW, Seol JH, Jeon YJ, Chung CH.. Modification of PCNA by ISG15 plays a crucial role in termination of error-prone translesion DNA synthesis. Mol Cell. 2014;54:626–638. doi: 10.1016/j.molcel.2014.03.031. [DOI] [PubMed] [Google Scholar]
  • 126.Leung W, Baxley RM, Moldovan GL, Bielinsky AK.. Mechanisms of DNA damage tolerance: post-translational regulation of PCNA. Genes. 2018;10:10. doi: 10.3390/genes10010010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Xu X, Lin A, Zhou C, Blackwell SR, Zhang Y, Wang Z, Feng Q, Guan R, Hanna MD, Chen Z, et al. Involvement of budding yeast Rad5 in translesion DNA synthesis through physical interaction with Rev1. Nucleic Acids Res. 2016;44:5231–5245. doi: 10.1093/nar/gkw183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Swain U, Friedlander G, Sehrawat U, Sarusi-Portuguez A, Rotkopf R, Ebert C, Paz-Elizur T, Dikstein R, Carell T, Geacintov NE, et al. TENT4A non-canonical poly(A) polymerase regulates DNA-damage tolerance via multiple pathways that are mutated in endometrial cancer. Int J Mol Sci. 2021;22:6957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Margara LM, Fernandez MM, Malchiodi EL, Argarana CE, Monti MR.. MutS regulates access of the error-prone DNA polymerase Pol IV to replication sites: a novel mechanism for maintaining replication fidelity. Nucleic Acids Res. 2016;44:7700–7713. doi: 10.1093/nar/gkw494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Yang XH, Shiotani B, Classon M, Zou L.. Chk1 and Claspin potentiate PCNA ubiquitination. Genes Dev. 2008;22:1147–1152. doi: 10.1101/gad.1632808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Gohler T, Sabbioneda S, Green CM, Lehmann AR.. ATR-mediated phosphorylation of DNA polymerase eta is needed for efficient recovery from UV damage. J Cell Biol. 2011;192:219–227. doi: 10.1083/jcb.201008076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Lin JR, Zeman MK, Chen JY, Yee MC, Cimprich KA.. SHPRH and HLTF act in a damage-specific manner to coordinate different forms of postreplication repair and prevent mutagenesis. Mol Cell. 2011;42:237–249. doi: 10.1016/j.molcel.2011.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Juhasz S, Balogh D, Hajdu I, Burkovics P, Villamil MA, Zhuang Z, Haracska L.. Characterization of human Spartan/C1orf124, an ubiquitin-PCNA interacting regulator of DNA damage tolerance. Nucleic Acids Res. 2012;40:10795–10808. doi: 10.1093/nar/gks850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Despras E, Sittewelle M, Pouvelle C, Delrieu N, Cordonnier AM, Kannouche PL.. Rad18-dependent SUMOylation of human specialized DNA polymerase eta is required to prevent under-replicated DNA. Nat Commun. 2016;7:13326. doi: 10.1038/ncomms13326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Guerillon C, Smedegaard S, Hendriks IA, Nielsen ML, Mailand N.. Multisite SUMOylation restrains DNA polymerase eta interactions with DNA damage sites. J Biol Chem. 2020;295:8350–8362. doi: 10.1074/jbc.RA120.013780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Peddu C, Zhang S, Zhao H, Wong A, Lee EYC, Lee M, Zhang Z.. Phosphorylation alters the properties of pol eta: implications for translesion synthesis. iScience. 2018;6:52–67. doi: 10.1016/j.isci.2018.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137. Chen YW, Cleaver JE, Hatahet Z, Honkanen RE, Chang JY, Yen Y, Chou KM.. Human DNA polymerase eta activity and translocation is regulated by phosphorylation. Proc Natl Acad Sci U S A. 2008;105:16578–16583. doi: 10.1073/pnas.0808589105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Bertoletti F, Cea V, Liang CC, Lanati T, Maffia A, Avarello MDM, Cipolla L, Lehmann AR, Cohn MA, Sabbioneda S.. Phosphorylation regulates human poleta stability and damage bypass throughout the cell cycle. Nucleic Acids Res. 2017;45:9441–9454. doi: 10.1093/nar/gkx619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139. Lerner LK, Francisco G, Soltys DT, Rocha CR, Quinet A, Vessoni AT, Castro LP, David TI, Bustos SO, Strauss BE, et al. Predominant role of DNA polymerase eta and p53-dependent translesion synthesis in the survival of ultraviolet-irradiated human cells. Nucleic Acids Res. 2017;45:1270–1280. doi: 10.1093/nar/gkw1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Bostian AC, Maddukuri L, Reed MR, Savenka T, Hartman JH, Davis L, Pouncey DL, Miller GP, Eoff RL.. Kynurenine signaling increases DNA polymerase kappa expression and promotes genomic instability in glioblastoma cells. Chem Res Toxicol. 2016;29:101–108. doi: 10.1021/acs.chemrestox.5b00452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Yamada T, Imamachi N, Imamura K, Taniue K, Kawamura T, Suzuki Y, Nagahama M, Akimitsu N.. Systematic analysis of targets of pumilio-mediated mRNA decay reveals that PUM1 repression by DNA damage activates translesion synthesis. Cell Rep. 2020;31:107542. doi: 10.1016/j.celrep.2020.107542. [DOI] [PubMed] [Google Scholar]
  • 142.Neelsen KJ, Lopes M.. Replication fork reversal in eukaryotes: from dead end to dynamic response. Nat Rev Mol Cell Biol. 2015;16:207–220. doi: 10.1038/nrm3935. [DOI] [PubMed] [Google Scholar]
  • 143. Leuzzi G, Marabitti V, Pichierri P, Franchitto A.. WRNIP1 protects stalled forks from degradation and promotes fork restart after replication stress. Embo J. 2016;35:1437–1451. doi: 10.15252/embj.201593265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Ciccia A, Nimonkar AV, Hu Y, Hajdu I, Achar YJ, Izhar L, Petit SA, Adamson B, Yoon JC, Kowalczykowski SC, et al. Polyubiquitinated PCNA recruits the ZRANB3 translocase to maintain genomic integrity after replication stress. Mol Cell. 2012;47:396–409. doi: 10.1016/j.molcel.2012.05.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Couch FB, Bansbach CE, Driscoll R, Luzwick JW, Glick GG, Betous R, Carroll CM, Jung SY, Qin J, Cimprich KA, et al. ATR phosphorylates SMARCAL1 to prevent replication fork collapse. Genes Dev. 2013;27:1610–1623. doi: 10.1101/gad.214080.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Unk I, Hajdu I, Blastyak A, Haracska L.. Role of yeast Rad5 and its human orthologs, HLTF and SHPRH in DNA damage tolerance. DNA Repair. 2010;9:257–267. doi: 10.1016/j.dnarep.2009.12.013. [DOI] [PubMed] [Google Scholar]
  • 147.Unk I, Hajdu I, Fatyol K, Szakal B, Blastyak A, Bermudez V, Hurwitz J, Prakash L, Prakash S, Haracska L.. Human SHPRH is a ubiquitin ligase for Mms2-Ubc13-dependent polyubiquitylation of proliferating cell nuclear antigen. Proc Natl Acad Sci U S A. 2006;103:18107–18112. doi: 10.1073/pnas.0608595103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.MacKay C, Toth R, Rouse J.. Biochemical characterisation of the SWI/SNF family member HLTF. Biochem Biophys Res Commun. 2009;390:187–191. doi: 10.1016/j.bbrc.2009.08.151. [DOI] [PubMed] [Google Scholar]
  • 149.Veaute X, Jeusset J, Soustelle C, Kowalczykowski SC, Le Cam E, Fabre F.. The Srs2 helicase prevents recombination by disrupting Rad51 nucleoprotein filaments. Nature. 2003;423:309–312. doi: 10.1038/nature01585. [DOI] [PubMed] [Google Scholar]
  • 150.Antony E, Tomko EJ, Xiao Q, Krejci L, Lohman TM, Ellenberger T.. Srs2 disassembles Rad51 filaments by a protein-protein interaction triggering ATP turnover and dissociation of Rad51 from DNA. Mol Cell. 2009;35:105–115. doi: 10.1016/j.molcel.2009.05.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Ler AAL, Carty MP.. DNA damage tolerance pathways in human cells: a potential therapeutic target. Front Oncol. 2021;11:822500. doi: 10.3389/fonc.2021.822500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Gonzalez-Huici V, Szakal B, Urulangodi M, Psakhye I, Castellucci F, Menolfi D, Rajakumara E, Fumasoni M, Bermejo R, Jentsch S, et al. DNA bending facilitates the error-free DNA damage tolerance pathway and upholds genome integrity. Embo J. 2014;33:327–340. doi: 10.1002/embj.201387425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Weill JC, Reynaud CA.. DNA polymerases in adaptive immunity. Nat Rev Immunol. 2008;8:302–312. doi: 10.1038/nri2281. [DOI] [PubMed] [Google Scholar]
  • 154.Saribasak H, Gearhart PJ.. Does DNA repair occur during somatic hypermutation? Semin Immunol. 2012;24:287–292. doi: 10.1016/j.smim.2012.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Masuda K, Ouchida R, Takeuchi A, Saito T, Koseki H, Kawamura K, Tagawa M, Tokuhisa T, Azuma T, J OW.. DNA polymerase theta contributes to the generation of C/G mutations during somatic hypermutation of Ig genes. Proc Natl Acad Sci U S A. 2005;102:13986–13991. doi: 10.1073/pnas.0505636102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Masuda K, Ouchida R, Hikida M, Nakayama M, Ohara O, Kurosaki T, J OW.. Absence of DNA polymerase theta results in decreased somatic hypermutation frequency and altered mutation patterns in Ig genes. DNA Repair. 2006;5:1384–1391. doi: 10.1016/j.dnarep.2006.06.006. [DOI] [PubMed] [Google Scholar]
  • 157.Masuda K, Ouchida R, Hikida M, Kurosaki T, Yokoi M, Masutani C, Seki M, Wood RD, Hanaoka F, J OW.. DNA polymerases eta and theta function in the same genetic pathway to generate mutations at A/T during somatic hypermutation of Ig genes. J Biol Chem. 2007;282:17387–17394. doi: 10.1074/jbc.M611849200. [DOI] [PubMed] [Google Scholar]
  • 158.Zan H, Shima N, Xu Z, Al-Qahtani A, Evinger Iii AJ, Zhong Y, Schimenti JC, Casali P.. The translesion DNA polymerase theta plays a dominant role in immunoglobulin gene somatic hypermutation. Embo J. 2005;24:3757–3769. doi: 10.1038/sj.emboj.7600833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Faili A, Stary A, Delbos F, Weller S, Aoufouchi S, Sarasin A, Weill JC, Reynaud CA.. A backup role of DNA polymerase kappa in Ig gene hypermutation only takes place in the complete absence of DNA polymerase eta. J Immunol. 2009;182:6353–6359. doi: 10.4049/jimmunol.0900177. [DOI] [PubMed] [Google Scholar]
  • 160. Maul RW, MacCarthy T, Frank EG, Donigan KA, McLenigan MP, Yang W, Saribasak H, Huston DE, Lange SS, Woodgate R, et al. DNA polymerase iota functions in the generation of tandem mutations during somatic hypermutation of antibody genes. J Exp Med. 2016;213:1675–1683. doi: 10.1084/jem.20151227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Hirota K, Sonoda E, Kawamoto T, Motegi A, Masutani C, Hanaoka F, Szuts D, Iwai S, Sale JE, Lehmann A, et al. Simultaneous disruption of two DNA polymerases, Poleta and Polzeta, in Avian DT40 cells unmasks the role of Poleta in cellular response to various DNA lesions. PLoS Genet. 2010;6:e1001151. doi: 10.1371/journal.pgen.1001151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Sharma NM, Kochenova OV, Shcherbakova PV.. The non-canonical protein binding site at the monomer-monomer interface of yeast proliferating cell nuclear antigen (PCNA) regulates the Rev1-PCNA interaction and Polzeta/Rev1-dependent translesion DNA synthesis. J Biol Chem. 2011;286:33557–33566. doi: 10.1074/jbc.M110.206680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Shah KA, Shishkin AA, Voineagu I, Pavlov YI, Shcherbakova PV, Mirkin SM.. Role of DNA polymerases in repeat-mediated genome instability. Cell Rep. 2012;2:1088–1095. doi: 10.1016/j.celrep.2012.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Saini N, Zhang Y, Nishida Y, Sheng Z, Choudhury S, Mieczkowski P, Lobachev KS.. Fragile DNA motifs trigger mutagenesis at distant chromosomal loci in saccharomyces cerevisiae. PLoS Genet. 2013;9:e1003551. doi: 10.1371/journal.pgen.1003551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Bhat A, Andersen PL, Qin Z, Xiao W.. Rev3, the catalytic subunit of Polzeta, is required for maintaining fragile site stability in human cells. Nucleic Acids Res. 2013;41:2328–2339. doi: 10.1093/nar/gks1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Bhat A, Wu Z, Maher VM, McCormick JJ, Xiao W.. Rev7/Mad2B plays a critical role in the assembly of a functional mitotic spindle. Cell Cycle. 2015;14:3929–3938. doi: 10.1080/15384101.2015.1120922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Bhat A, Qin Z, Wang G, Chen W, Xiao W.. Rev7, the regulatory subunit of Polzeta, undergoes UV-induced and Cul4-dependent degradation. Febs J. 2017;284:1790–1803. doi: 10.1111/febs.14088. [DOI] [PubMed] [Google Scholar]
  • 168.de Krijger I, Boersma V, Jacobs JJL.. REV7: jack of many trades. Trends Cell Biol. 2021;31:686–701. doi: 10.1016/j.tcb.2021.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Tang F, Wang Y, Gao Z, Guo S, Wang Y.. Polymerase eta recruits DHX9 helicase to promote replication across guanine quadruplex structures. J Am Chem Soc. 2022;144:14016–14020. doi: 10.1021/jacs.2c05312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Loor G, Zhang SJ, Zhang P, Toomey NL, Lee MY.. Identification of DNA replication and cell cycle proteins that interact with PCNA. Nucleic Acids Res. 1997;25:5041–5046. doi: 10.1093/nar/25.24.5041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Chakraborty P, Grosse F.. WRN helicase unwinds Okazaki fragment-like hybrids in a reaction stimulated by the human DHX9 helicase. Nucleic Acids Res. 2010;38:4722–4730. doi: 10.1093/nar/gkq240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Twayana S, Bacolla A, Barreto-Galvez A, De-Paula RB, Drosopoulos WC, Kosiyatrakul ST, Bouhassira EE, Tainer JA, Madireddy A, Schildkraut CL.. Translesion polymerase eta both facilitates DNA replication and promotes increased human genetic variation at common fragile sites. Proc Natl Acad Sci U S A. 2021;118:e2106477118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Yoon JH, Basu D, Sellamuthu K, Johnson RE, Prakash S, Prakash L.. A novel role of DNA polymerase lambda in translesion synthesis in conjunction with DNA polymerase zeta. Life Sci Alliance. 2021;4:e202000900. doi: 10.26508/lsa.202000900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Yoon JH, Johnson RE, Prakash L, Prakash S.. Genetic evidence for reconfiguration of DNA polymerase theta active site for error-free translesion synthesis in human cells. J Biol Chem. 2020;295:5918–5927. doi: 10.1074/jbc.RA120.012816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Simpson LJ, Sale JE.. UBE2V2 (MMS2) is not required for effective immunoglobulin gene conversion or DNA damage tolerance in DT40. DNA Repair. 2005;4:503–510. doi: 10.1016/j.dnarep.2004.12.002. [DOI] [PubMed] [Google Scholar]
  • 176.Chen X, Bosques L, Sung P, Kupfer GM.. A novel role for non-ubiquitinated FANCD2 in response to hydroxyurea-induced DNA damage. Oncogene. 2016;35:22–34. doi: 10.1038/onc.2015.68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Covo S, Blanco L, Livneh Z.. Lesion bypass by human DNA polymerase mu reveals a template-dependent, sequence-independent nucleotidyl transferase activity. J Biol Chem. 2004;279:859–865. doi: 10.1074/jbc.M310447200. [DOI] [PubMed] [Google Scholar]
  • 178.Kim SR, Maenhaut-Michel G, Yamada M, Yamamoto Y, Matsui K, Sofuni T, Nohmi T, Ohmori H.. Multiple pathways for SOS-induced mutagenesis in Escherichia coli: an overexpression of dinB/dinP results in strongly enhancing mutagenesis in the absence of any exogenous treatment to damage DNA. Proc Natl Acad Sci U S A. 1997;94:13792–13797. doi: 10.1073/pnas.94.25.13792. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Ogi T, Kato T, Jr., Kato T, Ohmori H.. Mutation enhancement by DINB1, a mammalian homologue of the Escherichia coli mutagenesis protein dinB. Genes Cells. 1999;4:607–618. doi: 10.1046/j.1365-2443.1999.00289.x. [DOI] [PubMed] [Google Scholar]
  • 180.Bergoglio V, Bavoux C, Verbiest V, Hoffmann JS, Cazaux C.. Localisation of human DNA polymerase kappa to replication foci. J Cell Sci. 2002;115:4413–4418. doi: 10.1242/jcs.00162. [DOI] [PubMed] [Google Scholar]
  • 181.Bavoux C, Hoffmann JS, Cazaux C.. Adaptation to DNA damage and stimulation of genetic instability: the double-edged sword mammalian DNA polymerase kappa. Biochimie. 2005;87:637–646. doi: 10.1016/j.biochi.2005.02.007. [DOI] [PubMed] [Google Scholar]
  • 182.Lenne-Samuel N, Wagner J, Etienne H, Fuchs RP.. The processivity factor beta controls DNA polymerase IV traffic during spontaneous mutagenesis and translesion synthesis in vivo. EMBO Rep. 2002;3:45–49. doi: 10.1093/embo-reports/kvf007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Fujii S, Fuchs RP.. Defining the position of the switches between replicative and bypass DNA polymerases. Embo J. 2004;23:4342–4352. doi: 10.1038/sj.emboj.7600438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Pillaire MJ, Betous R, Conti C, Czaplicki J, Pasero P, Bensimon A, Cazaux C, Hoffmann JS.. Upregulation of error-prone DNA polymerases beta and kappa slows down fork progression without activating the replication checkpoint. Cell Cycle. 2007;6:471–477. doi: 10.4161/cc.6.4.3857. [DOI] [PubMed] [Google Scholar]
  • 185.Kawamoto T, Araki K, Sonoda E, Yamashita YM, Harada K, Kikuchi K, Masutani C, Hanaoka F, Nozaki K, Hashimoto N, et al. Dual roles for DNA polymerase eta in homologous DNA recombination and translesion DNA synthesis. Mol Cell. 2005;20:793–799. doi: 10.1016/j.molcel.2005.10.016. [DOI] [PubMed] [Google Scholar]
  • 186.Masuda Y, Kamiya K.. Role of single-stranded DNA in targeting REV1 to primer termini. J Biol Chem. 2006;281:24314–24321. doi: 10.1074/jbc.M602967200. [DOI] [PubMed] [Google Scholar]
  • 187.Bemark M, Khamlichi AA, Davies SL, Neuberger MS.. Disruption of mouse polymerase zeta (Rev3) leads to embryonic lethality and impairs blastocyst development in vitro. Curr Biol. 2000;10:1213–1216. doi: 10.1016/s0960-9822(00)00724-7. [DOI] [PubMed] [Google Scholar]
  • 188. Esposito G, Godindagger I, Klein U, Yaspo ML, Cumano A, Rajewsky K.. Disruption of the Rev3l-encoded catalytic subunit of polymerase zeta in mice results in early embryonic lethality. Curr Biol. 2000;10:1221–1224. doi: 10.1016/s0960-9822(00)00726-0. [DOI] [PubMed] [Google Scholar]
  • 189.Van Sloun PP, Varlet I, Sonneveld E, Boei JJ, Romeijn RJ, Eeken JC, De Wind N.. Involvement of mouse Rev3 in tolerance of endogenous and exogenous DNA damage. Mol Cell Biol. 2002;22:2159–2169. doi: 10.1128/MCB.22.7.2159-2169.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Drake JW. A constant rate of spontaneous mutation in DNA-based microbes. Proc Natl Acad Sci U S A. 1991;88:7160–7164. doi: 10.1073/pnas.88.16.7160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Foster PL. Stress-induced mutagenesis in bacteria. Crit Rev Biochem Mol Biol. 2007;42:373–397. doi: 10.1080/10409230701648494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Galhardo RS, Hastings PJ, Rosenberg SM.. Mutation as a stress response and the regulation of evolvability. Crit Rev Biochem Mol Biol. 2007;42:399–435. doi: 10.1080/10409230701648502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Casali P, Pal Z, Xu Z, Zan H.. DNA repair in antibody somatic hypermutation. Trends Immunol. 2006;27:313–321. doi: 10.1016/j.it.2006.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194. Wyatt DW, Feng W, Conlin MP, Yousefzadeh MJ, Roberts SA, Mieczkowski P, Wood RD, Gupta GP, Ramsden DA.. Essential roles for polymerase theta-mediated end joining in the repair of chromosome breaks. Mol Cell. 2016;63:662–673. doi: 10.1016/j.molcel.2016.06.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Yang J, Chen Z, Liu Y, Hickey RJ, Malkas LH.. Altered DNA polymerase iota expression in breast cancer cells leads to a reduction in DNA replication fidelity and a higher rate of mutagenesis. Cancer Res. 2004;64:5597–5607. doi: 10.1158/0008-5472.CAN-04-0603. [DOI] [PubMed] [Google Scholar]
  • 196.Yuan F, Xu Z, Yang M, Wei Q, Zhang Y, Yu J, Zhi Y, Liu Y, Chen Z, Yang J.. Overexpressed DNA polymerase iota regulated by JNK/c-Jun contributes to hypermutagenesis in bladder cancer. PLoS One. 2013;8:e69317. doi: 10.1371/journal.pone.0069317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Wang R, Lenoir WF, Wang C, Su D, McLaughlin M, Hu Q, Shen X, Tian Y, Klages-Mundt N, Lynn E, et al. DNA polymerase iota compensates for Fanconi anemia pathway deficiency by countering DNA replication stress. Proc Natl Acad Sci U S A. 2020;117:33436–33445. doi: 10.1073/pnas.2008821117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Smolinska A, Singer K, Golchert J, Smyczynska U, Fendler W, Sendler M, van den Brandt J, Singer S, Homuth G, Lerch MM, et al. DNA polymerase theta plays a critical role in pancreatic cancer development and metastasis. Cancers. 2022;14:4077. doi: 10.3390/cancers14174077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Simoneau A, Xiong R, Zou L.. The trans cell cycle effects of PARP inhibitors underlie their selectivity toward BRCA1/2-deficient cells. Genes Dev. 2021;35:1271–1289. doi: 10.1101/gad.348479.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Bluteau D, Masliah-Planchon J, Clairmont C, Rousseau A, Ceccaldi R, Dubois d‘Enghien C, Bluteau O, Cuccuini W, Gachet S, Peffault de Latour R, et al. Biallelic inactivation of REV7 is associated with Fanconi anemia. J Clin Invest. 2016;126:3580–3584. doi: 10.1172/JCI88010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Ceccaldi R, Liu JC, Amunugama R, Hajdu I, Primack B, Petalcorin MI, O'Connor KW, Konstantinopoulos PA, Elledge SJ, Boulton SJ, et al. Homologous-recombination-deficient tumours are dependent on Poltheta-mediated repair. Nature. 2015;518:258–262. doi: 10.1038/nature14184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202. Schimmel J, van Schendel R, den Dunnen JT, Tijsterman M.. Templated insertions: a smoking gun for polymerase theta-mediated end joining. Trends Genet. 2019;35:632–644. doi: 10.1016/j.tig.2019.06.001. [DOI] [PubMed] [Google Scholar]
  • 203. Alexandrov LB, Kim J, Haradhvala NJ, Huang MN, Tian Ng AW, Wu Y, Boot A, Covington KR, Gordenin DA, Bergstrom EN, et al. The repertoire of mutational signatures in human cancer. Nature. 2020;578:94–101. doi: 10.1038/s41586-020-1943-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204. Anand J, Chiou L, Sciandra C, Zhang X, Hong J, Wu D, Zhou P, Vaziri C.. Roles of trans-lesion synthesis (TLS) DNA polymerases in tumorigenesis and cancer therapy. NAR Cancer. 2023;5:zcad005. doi: 10.1093/narcan/zcad005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205. Roberts SA, Sterling J, Thompson C, Harris S, Mav D, Shah R, Klimczak LJ, Kryukov GV, Malc E, Mieczkowski PA, et al. Clustered mutations in yeast and in human cancers can arise from damaged long single-strand DNA regions. Mol Cell. 2012;46:424–435. doi: 10.1016/j.molcel.2012.03.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206. Morganella S, Alexandrov LB, Glodzik D, Zou X, Davies H, Staaf J, Sieuwerts AM, Brinkman AB, Martin S, Ramakrishna M, et al. The topography of mutational processes in breast cancer genomes. Nat Commun. 2016;7:11383. doi: 10.1038/ncomms11383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207. Petljak M, Dananberg A, Chu K, Bergstrom EN, Striepen J, von Morgen P, Chen Y, Shah H, Sale JE, Alexandrov LB, et al. Mechanisms of APOBEC3 mutagenesis in human cancer cells. Nature. 2022;607:799–807. doi: 10.1038/s41586-022-04972-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208. Delbos F, De Smet A, Faili A, Aoufouchi S, Weill JC, Reynaud CA.. Contribution of DNA polymerase eta to immunoglobulin gene hypermutation in the mouse. J Exp Med. 2005;201:1191–1196. doi: 10.1084/jem.20050292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209. Durando M, Tateishi S, Vaziri C.. A non-catalytic role of DNA polymerase eta in recruiting Rad18 and promoting PCNA monoubiquitination at stalled replication forks. Nucleic Acids Res. 2013;41:3079–3093. doi: 10.1093/nar/gkt016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210. Gao Y, Mutter-Rottmayer E, Greenwalt AM, Goldfarb D, Yan F, Yang Y, Martinez-Chacin RC, Pearce KH, Tateishi S, Major MB, et al. A neomorphic cancer cell-specific role of MAGE-A4 in trans-lesion synthesis. Nat Commun. 2016;7:12105. doi: 10.1038/ncomms12105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Jung YS, Qian Y, Chen X.. DNA polymerase eta is targeted by Mdm2 for polyubiquitination and proteasomal degradation in response to ultraviolet irradiation. DNA Repair. 2012;11:177–184. doi: 10.1016/j.dnarep.2011.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Sitz J, Blanchet SA, Gameiro SF, Biquand E, Morgan TM, Galloy M, Dessapt J, Lavoie EG, Blondeau A, Smith BC, et al. Human papillomavirus E7 oncoprotein targets RNF168 to hijack the host DNA damage response. Proc Natl Acad Sci U S A. 2019;116:19552–19562. doi: 10.1073/pnas.1906102116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Pilzecker B, Jacobs H.. Mutating for good: DNA damage responses during somatic hypermutation. Front Immunol. 2019;10:438. doi: 10.3389/fimmu.2019.00438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214. Dumstorf CA, Clark AB, Lin Q, Kissling GE, Yuan T, Kucherlapati R, McGregor WG, Kunkel TA.. Participation of mouse DNA polymerase iota in strand-biased mutagenic bypass of UV photoproducts and suppression of skin cancer. Proc Natl Acad Sci U S A. 2006;103:18083–18088. doi: 10.1073/pnas.0605247103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Stancel JN, McDaniel LD, Velasco S, Richardson J, Guo C, Friedberg EC.. Polk mutant mice have a spontaneous mutator phenotype. DNA Repair. 2009;8:1355–1362. doi: 10.1016/j.dnarep.2009.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216. Pilzecker B, Buoninfante OA, van den Berk P, Lancini C, Song JY, Citterio E, Jacobs H.. DNA damage tolerance in hematopoietic stem and progenitor cells in mice. Proc Natl Acad Sci U S A. 2017;114:E6875–E6883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Young K, Borikar S, Bell R, Kuffler L, Philip V, Trowbridge JJ.. Progressive alterations in multipotent hematopoietic progenitors underlie lymphoid cell loss in aging. J Exp Med. 2016;213:2259–2267. doi: 10.1084/jem.20160168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Zeman MK, Cimprich KA.. Causes and consequences of replication stress. Nat Cell Biol. 2014;16:2–9. doi: 10.1038/ncb2897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Murga M, Bunting S, Montana MF, Soria R, Mulero F, Canamero M, Lee Y, McKinnon PJ, Nussenzweig A, Fernandez-Capetillo O.. A mouse model of ATR-Seckel shows embryonic replicative stress and accelerated aging. Nat Genet. 2009;41:891–898. doi: 10.1038/ng.420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Zhuo M, Gorgun MF, Englander EW.. Translesion synthesis DNA polymerase kappa is indispensable for DNA repair synthesis in cisplatin exposed dorsal root ganglion neurons. Mol Neurobiol. 2018;55:2506–2515. doi: 10.1007/s12035-017-0507-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Weerasuriya A, Mizisin AP.. The blood-nerve barrier: structure and functional significance. Methods Mol Biol. 2011;686:149–173. doi: 10.1007/978-1-60761-938-3_6. [DOI] [PubMed] [Google Scholar]
  • 222.Gratchev A, Strein P, Utikal J, Sergij G.. Molecular genetics of Xeroderma pigmentosum variant. Exp Dermatol. 2003;12:529–536. doi: 10.1034/j.1600-0625.2003.00124.x. [DOI] [PubMed] [Google Scholar]
  • 223. Pan J, Chi P, Lu X, Xu Z.. Genetic polymorphisms in translesion synthesis genes are associated with colorectal cancer risk and metastasis in Han Chinese. Gene. 2012;504:151–155. doi: 10.1016/j.gene.2012.05.042. [DOI] [PubMed] [Google Scholar]
  • 224.Dai ZJ, Liu XH, Ma YF, Kang HF, Jin TB, Dai ZM, Guan HT, Wang M, Liu K, Dai C, et al. Association between single nucleotide polymorphisms in DNA polymerase kappa gene and breast cancer risk in chinese han population: a STROBE-compliant observational study. Medicine. 2016;95:e2466. doi: 10.1097/MD.0000000000002466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Jamwal RS, Mahajan N, Bhat GR, Bhat A, Shah R, Verma S, Sharma M, Sharma B, Qadri RA, Kumar R, et al. REV3L single nucleotide variants lead to increased susceptibility towards non-small cell lung cancer in the population of Jammu and Kashmir. Cancer Epidemiol. 2021;75:102047. doi: 10.1016/j.canep.2021.102047. [DOI] [PubMed] [Google Scholar]
  • 226.Xu HL, Gao XR, Zhang W, Cheng JR, Tan YT, Zheng W, Shu XO, Xiang YB.. Effects of polymorphisms in translesion DNA synthesis genes on lung cancer risk and prognosis in Chinese men. Cancer Epidemiol. 2013;37:917–922. doi: 10.1016/j.canep.2013.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Verzijl HT, van der Zwaag B, Cruysberg JR, Padberg GW.. Mobius syndrome redefined: a syndrome of rhombencephalic maldevelopment. Neurology. 2003;61:327–333. doi: 10.1212/01.wnl.0000076484.91275.cd. [DOI] [PubMed] [Google Scholar]
  • 228.Zhao F, Wu J, Xue A, Su Y, Wang X, Lu X, Zhou Z, Qu J, Zhou X.. Exome sequencing reveals CCDC111 mutation associated with high myopia. Hum Genet. 2013;132:913–921. doi: 10.1007/s00439-013-1303-6. [DOI] [PubMed] [Google Scholar]
  • 229.Díaz-Talavera A, Calvo PA, González-Acosta D, Díaz M, Sastre-Moreno G, Blanco-Franco L, Guerra S, Martínez-Jiménez MI, Méndez J, Blanco L.. A cancer-associated point mutation disables the steric gate of human PrimPol. Sci Rep. 2019;9:1121. doi: 10.1038/s41598-018-37439-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Albertella MR, Lau A, O'Connor MJ.. The overexpression of specialized DNA polymerases in cancer. DNA Repair. 2005;4:583–593. doi: 10.1016/j.dnarep.2005.01.005. [DOI] [PubMed] [Google Scholar]
  • 231.Lemee F, Bavoux C, Pillaire MJ, Bieth A, Machado CR, Pena SD, Guimbaud R, Selves J, Hoffmann JS, Cazaux C.. Characterization of promoter regulatory elements involved in downexpression of the DNA polymerase kappa in colorectal cancer. Oncogene. 2007;26:3387–3394. doi: 10.1038/sj.onc.1210116. [DOI] [PubMed] [Google Scholar]
  • 232.Ow J, Kawamura K, Tada Y, Ohmori H, Kimura H, Sakiyama S, Tagawa M.. DNA polymerase kappa, implicated in spontaneous and DNA damage-induced mutagenesis, is overexpressed in lung cancer. Cancer Res. 2001;61:5366–5369. [PubMed] [Google Scholar]
  • 233.Wang Y, Seimiya M, Kawamura K, Yu L, Ogi T, Takenaga K, Shishikura T, Nakagawara A, Sakiyama S, Tagawa M, et al. Elevated expression of DNA polymerase kappa in human lung cancer is associated with p53 inactivation: negative regulation of POLK promoter activity by p53. Int J Oncol. 2004;25:161–165. [PubMed] [Google Scholar]
  • 234.Wang H, Wu W, Wang HW, Wang S, Chen Y, Zhang X, Yang J, Zhao S, Ding HF, Lu D.. Analysis of specialized DNA polymerases expression in human gliomas: association with prognostic significance. Neuro Oncol. 2010;12:679–686. doi: 10.1093/neuonc/nop074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Pan Q, Fang Y, Xu Y, Zhang K, Hu X.. Down-regulation of DNA polymerases kappa, eta, iota, and zeta in human lung, stomach, and colorectal cancers. Cancer Lett. 2005;217:139–147. doi: 10.1016/j.canlet.2004.07.021. [DOI] [PubMed] [Google Scholar]
  • 236.Zhou W, Chen YW, Liu X, Chu P, Loria S, Wang Y, Yen Y, Chou KM.. Expression of DNA translesion synthesis polymerase eta in head and neck squamous cell cancer predicts resistance to gemcitabine and cisplatin-based chemotherapy. PLoS One. 2013;8:e83978. doi: 10.1371/journal.pone.0083978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Hicks JK, Chute CL, Paulsen MT, Ragland RL, Howlett NG, Gueranger Q, Glover TW, Canman CE.. Differential roles for DNA polymerases eta, zeta, and REV1 in lesion bypass of intrastrand versus interstrand DNA cross-links. Mol Cell Biol. 2010;30:1217–1230. doi: 10.1128/MCB.00993-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Okuda T, Lin X, Trang J, Howell SB.. Suppression of hREV1 expression reduces the rate at which human ovarian carcinoma cells acquire resistance to cisplatin. Mol Pharmacol. 2005;67:1852–1860. doi: 10.1124/mol.104.010579. [DOI] [PubMed] [Google Scholar]
  • 239.Xie K, Doles J, Hemann MT, Walker GC.. Error-prone translesion synthesis mediates acquired chemoresistance. Proc Natl Acad Sci U S A. 2010;107:20792–20797. doi: 10.1073/pnas.1011412107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Sakurai Y, Ichinoe M, Yoshida K, Nakazato Y, Saito S, Satoh M, Nakada N, Sanoyama I, Umezawa A, Numata Y, et al. Inactivation of REV7 enhances chemosensitivity and overcomes acquired chemoresistance in testicular germ cell tumors. Cancer Lett. 2020;489:100–110. doi: 10.1016/j.canlet.2020.06.001. [DOI] [PubMed] [Google Scholar]
  • 241.Ghosal G, Chen J.. DNA damage tolerance: a double-edged sword guarding the genome. Transl Cancer Res. 2013;2:107–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Shilkin ES, Boldinova EO, Stolyarenko AD, Goncharova RI, Chuprov-Netochin RN, Smal MP, Makarova AV.. Translesion DNA synthesis and reinitiation of DNA synthesis in chemotherapy resistance. Biochemistry. 2020;85:869–882. doi: 10.1134/S0006297920080039. [DOI] [PubMed] [Google Scholar]
  • 243.Wojtaszek JL, Chatterjee N, Najeeb J, Ramos A, Lee M, Bian K, Xue JY, Fenton BA, Park H, Li D, et al. A small molecule targeting mutagenic translesion synthesis improves chemotherapy. Cell. 2019;178:152–159 e111. doi: 10.1016/j.cell.2019.05.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Chatterjee N, Whitman MA, Harris CA, Min SM, Jonas O, Lien EC, Luengo A, Vander Heiden MG, Hong J, Zhou P, et al. REV1 inhibitor JH-RE-06 enhances tumor cell response to chemotherapy by triggering senescence hallmarks. Proc Natl Acad Sci U S A. 2020;117:28918–28921. doi: 10.1073/pnas.2016064117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Vanarotti M, Evison BJ, Actis ML, Inoue A, McDonald ET, Shao Y, Heath RJ, Fujii N.. Small-molecules that bind to the ubiquitin-binding motif of REV1 inhibit REV1 interaction with K164-monoubiquitinated PCNA and suppress DNA damage tolerance. Bioorg Med Chem. 2018;26:2345–2353. doi: 10.1016/j.bmc.2018.03.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Zhou J, Gelot C, Pantelidou C, Li A, Yucel H, Davis RE, Farkkila A, Kochupurakkal B, Syed A, Shapiro GI, et al. A first-in-class polymerase theta inhibitor selectively targets homologous-recombination-deficient tumors. Nat Cancer. 2021;2:598–610. doi: 10.1038/s43018-021-00203-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Makarova AV, Kulbachinskiy AV.. Structure of human DNA polymerase iota and the mechanism of DNA synthesis. Biochemistry. 2012;77:547–561. doi: 10.1134/S0006297912060016. [DOI] [PubMed] [Google Scholar]
  • 248.Ignatov AV, Bondarenko KA, Makarova AV.. Non-bulky lesions in human DNA: the ways of formation, repair, and replication. Acta Naturae. 2017;9:12–26. doi: 10.32607/20758251-2017-9-3-12-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Doles J, Oliver TG, Cameron ER, Hsu G, Jacks T, Walker GC, Hemann MT.. Suppression of Rev3, the catalytic subunit of Polzeta, sensitizes drug-resistant lung tumors to chemotherapy. Proc Natl Acad Sci U S A. 2010;107:20786–20791. doi: 10.1073/pnas.1011409107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Xu X, Xie K, Zhang XQ, Pridgen EM, Park GY, Cui DS, Shi J, Wu J, Kantoff PW, Lippard SJ, et al. Enhancing tumor cell response to chemotherapy through nanoparticle-mediated codelivery of siRNA and cisplatin prodrug. Proc Natl Acad Sci U S A. 2013;110:18638–18643. doi: 10.1073/pnas.1303958110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Srivastava AK, Han C, Zhao R, Cui T, Dai Y, Mao C, Zhao W, Zhang X, Yu J, Wang QE.. Enhanced expression of DNA polymerase eta contributes to cisplatin resistance of ovarian cancer stem cells. Proc Natl Acad Sci U S A. 2015;112:4411–4416. doi: 10.1073/pnas.1421365112. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing not applicable – no new data generated.


Articles from Molecular and Cellular Biology are provided here courtesy of Taylor & Francis

RESOURCES