Base excision repair (BER) has evolved to preserve the integrity of DNA following cellular oxidative stress and in response to exogenous insults. The pathway is a coordinated, sequential process involving 30 proteins or more in which single strand breaks are generated as intermediates during the repair process. While deficiencies in BER activity can lead to high mutation rates and tumorigenesis, cancer cells often rely on increased BER activity to tolerate oxidative stress. Targeting BER has been an attractive strategy to overwhelm cancer cells with DNA damage, improve the efficacy of radiotherapy and/or chemotherapy, or form part of a lethal combination with a cancer specific mutation/loss of function. We provide an update on the progress of inhibitors to enzymes involved in BER, and some of the challenges faced with targeting the BER pathway.

Overview of the BER pathway

Base excision repair is a highly conserved mechanism dealing with oxidative damage generated by respiration, natural hydrolysis and alkylation reactions that occur in each cell, many thousands of times a day [1]. In humans at least 30 proteins are involved in both short patch repair (SP-BER), the removal of a single non-bulky damaged base; and long patch repair (LP-BER), where 2–8 nucleotides are synthesised to displace the damaged area (Figure 1). The first step of the BER pathway is the recognition and removal of base damage by damage-specific DNA glycosylases [2]. Humans have 11 DNA glycosylases that can be subdivided into three groups (Table 1): (1) Monofunctional enzymes which excise the damaged base leaving an apurinic/apyrimidinic (AP) site and an intact phosphodiester backbone; (2) Bifunctional glycosylases that remove the base and cleave the phosphodiester bond on the 3′ side of the damaged base creating an 3′-α,β-unsaturated aldehyde (β-elimination); (3) Nei-like DNA glycolysases (NEIL) that can catalyse a β/δ-elimination reaction where the phosphodiester bond is cleaved either side of the removed lesion.

Overview of SP-BER and LP-BER

Figure 1
Overview of SP-BER and LP-BER

(A) Base damage (red flag) is recognised by one of 11 damage-specific DNA glycosylases that are monofunctional (removes base creating an AP site); bifunctional (removes base and cleaves phosphodiester backbone 3′ to the lesion) or Nei-like (cleaves phosphodiester bond either side of lesion). APE1 incises the AP site, or removes the 3′-α,β-unsaturated aldehyde remaining from bifunctional DNA glycosylase action. PNKP is required to remove 3′-phosphate termini following Nei-like DNA base excision. At this stage, PARP1 recognises SSB/gap intermediates protecting these from degradation, and facilitates repair through protein recruitment. Polβ through its lyase activity excises the 5′-dRP moiety, and simultaneously fills the one nucleotide gap (green nucleotide). XRCC1-LigIIIα complex interacts with Polβ and repairs the remaining nick in the DNA, thus completing SP-BER. (B) When the 5′-DNA end is not amenable to Polβ, LP-BER is employed. A polymerase switch to Polδ/ε stimulates strand displacement and creates a 2–8 nucleotide 5′-flap. FEN1 cleaves the flap and LigI ligates the subsequent nick, both of which are stimulated by the PCNA clamp slider. Each enzymatic step of the pathway can be targeted by small molecule inhibitors, as indicated by the red cross.

Figure 1
Overview of SP-BER and LP-BER

(A) Base damage (red flag) is recognised by one of 11 damage-specific DNA glycosylases that are monofunctional (removes base creating an AP site); bifunctional (removes base and cleaves phosphodiester backbone 3′ to the lesion) or Nei-like (cleaves phosphodiester bond either side of lesion). APE1 incises the AP site, or removes the 3′-α,β-unsaturated aldehyde remaining from bifunctional DNA glycosylase action. PNKP is required to remove 3′-phosphate termini following Nei-like DNA base excision. At this stage, PARP1 recognises SSB/gap intermediates protecting these from degradation, and facilitates repair through protein recruitment. Polβ through its lyase activity excises the 5′-dRP moiety, and simultaneously fills the one nucleotide gap (green nucleotide). XRCC1-LigIIIα complex interacts with Polβ and repairs the remaining nick in the DNA, thus completing SP-BER. (B) When the 5′-DNA end is not amenable to Polβ, LP-BER is employed. A polymerase switch to Polδ/ε stimulates strand displacement and creates a 2–8 nucleotide 5′-flap. FEN1 cleaves the flap and LigI ligates the subsequent nick, both of which are stimulated by the PCNA clamp slider. Each enzymatic step of the pathway can be targeted by small molecule inhibitors, as indicated by the red cross.

Close modal
Table 1
DNA glycosylases: substrates, inhibitors and synthetic lethal partners
DNA glycosylaseSubstrateInhibitorSynthetic lethal partner
Monofunctional 
UNG
Uracil DNA glycosylase 
Uracil  APOBEC3B 
SMUG1
Single-strand selective monofunctional uracil DNA glycosylase 
Uracil, 5-formyluracil, 5-hydroxyuracil and 5-hydroxymethyluracil   
TDG
G/T Mismatch specific thymine DNA glycosylase 
5-Formylcytosine and 5-carboxylcytosine in CpG. G:T and U:T mismatches   
MBD4
Methyl-CpG binding domain protein 4 
G:T mismatches within methylated and unmethylated CpG sites. Uracil or 5-fluorouracil in G:U mismatches   
MPG
N-methylpurine DNA glycosylase 
3-Methyladenine and 7-methylguanine   
MUTYH
Adenine DNA glycosylase 
7,8-Dihydro-8-oxoguanine (8-oxoG):adenine   
Bifunctional 
NTH1
Endonuclease III-like protein 1 
Oxidised pyrimidines, thymine glycol
8-OxoG 
  
OGG18-oxoguanine DNA glycosylase 1 8-OxoG and formamidopyrimidine(Fapy)G O8SU0268TH5487 MMR deficiency 
Endonuclease VIII-like 
NEIL1 Thymine glycol, Fapy and 5-hydroxyuracil 2TX FANCG 
NEIL2 5-Hydroxyuracil   
NEIL3 Spiroiminodihydantoin and guanidinohydantoin   
DNA glycosylaseSubstrateInhibitorSynthetic lethal partner
Monofunctional 
UNG
Uracil DNA glycosylase 
Uracil  APOBEC3B 
SMUG1
Single-strand selective monofunctional uracil DNA glycosylase 
Uracil, 5-formyluracil, 5-hydroxyuracil and 5-hydroxymethyluracil   
TDG
G/T Mismatch specific thymine DNA glycosylase 
5-Formylcytosine and 5-carboxylcytosine in CpG. G:T and U:T mismatches   
MBD4
Methyl-CpG binding domain protein 4 
G:T mismatches within methylated and unmethylated CpG sites. Uracil or 5-fluorouracil in G:U mismatches   
MPG
N-methylpurine DNA glycosylase 
3-Methyladenine and 7-methylguanine   
MUTYH
Adenine DNA glycosylase 
7,8-Dihydro-8-oxoguanine (8-oxoG):adenine   
Bifunctional 
NTH1
Endonuclease III-like protein 1 
Oxidised pyrimidines, thymine glycol
8-OxoG 
  
OGG18-oxoguanine DNA glycosylase 1 8-OxoG and formamidopyrimidine(Fapy)G O8SU0268TH5487 MMR deficiency 
Endonuclease VIII-like 
NEIL1 Thymine glycol, Fapy and 5-hydroxyuracil 2TX FANCG 
NEIL2 5-Hydroxyuracil   
NEIL3 Spiroiminodihydantoin and guanidinohydantoin   

Following monofunctional DNA glycosylase action, AP endonuclease 1 (APE1) recognises the AP site and hydrolyses the DNA backbone forming a single strand break (SSB) with a 5′-deoxyribosephosphate (dRP) and 3′-hydroxyl ends [3,4]. It also acts on the products of bifunctional glycosylases, where hydrolysis expels the 3′-α,β-unsaturated aldehyde and creates a one nucleotide gap product. The product of the NEIL glycosylases contains a 3′-phosphate group which requires polynucleotide kinase phosphatase (PNKP) to generate a 3′-hydroxyl end that is suitable for DNA polymerase action [5]. At this point, on the formation of a gap or SSB, poly(ADP-ribose) polymerase 1 (PARP1) is engaged which protects the strand break [6] and also plays a role in protein recruitment through its associated poly(ADP-ribosyl)ation activity. The major polymerase employed during BER is DNA polymerase β (Polβ), which fills the gap but also catalyses a lyase reaction that removes the 5′-dRP that may be present [7,8]. Finally, DNA ligase IIIα (LigIIIα) in complex with X-ray repair cross-complementing protein 1 (XRCC1) completes the process of SP-BER (Figure 1A) [9,10]. LP-BER is employed when the 5′-DNA ends are not amenable to Polβ action. Here, a polymerase switch occurs and a flap of 2–8 nucleotides are synthesised by DNA polymerases δ/ε (Polδ/ε), which associates with proliferating cell nuclear antigen (PCNA), displacing the damaged strand. Flap endonuclease 1 (FEN1) activity removes the displaced strand leaving a ligatable nick for DNA ligase I (LigI), which associates with the PCNA clamp slider (Figure 1B) [11,12].

BER dependence following radiotherapy and chemotherapy

Anti-neoplastic drugs inhibit mitosis and many of them do so through alterations to DNA, which normally would be repaired by cells, but overwhelms rapidly dividing cells to trigger cell death. Such classes of chemotherapy agent include: nucleoside analogues that become incorporated into DNA (e.g. 5-fluorouracil; 5-FU) [13]; antifolates that inhibit the synthesis of deoxythymidine triphosphate and so increase uracil incorporation (e.g. pemetrexed) [14]; demethylating agents that cause DNA damage by trapping DNA methyltransferases (e.g. decitabine 5-aza-2′-deoxycytidine; 5-azadC) [15]; platinum drugs (e.g. cisplatin) [16,17]; and alkylating agents that produce DNA adducts (e.g. temozolomide; TMZ). Thus, several chemotherapy agents produce DNA modifications that rely on BER for removal and cytotoxicity in cancer cells. Current external (photon and particle beam) and internal radiotherapy approaches also generate a large proportion of DNA damage that is a target for BER. Ionising radiation emanating from these sources either directly or indirectly creates a mixture of base damage, oxidative damage and SSBs that are recognised by the BER pathway. The amount of DNA damage caused by chemotherapy and/or radiotherapy should overwhelm the cancer cell’s capacity for repair for therapeutic effectiveness.

One strategy for cancer therapy is to target BER enzymes with inhibitors, and in combination with radiotherapy and/or chemotherapy, this will create additional damage that exceeds the BER ability of the cancer cells. This is particularly important for specific tumour types that contain BER gene and protein overexpression. Another desirable strategy is where targeting BER enzymes can lead to specific killing of cancer cells via a synthetic lethal partnership, and where a tumorigenic mutation becomes reliant on BER to survive. In this section, we review the progress made with developing and characterising inhibitors to several BER enzymes using some of these strategies (and summarised in Tables 1 and 2).

Table 2
Targeting BER intermediates and enzymes
TargetInhibitorSynthetic lethal partner
Apurinic/apyrimidinic site MX (TRC102)  
BER enzymes 
APE1   
lyase activity CRT0044876 PTEN 
 API3 BRCA1 
  BRCA2 
  ATM 
Redox function E3330  
 Gossypol/AT101  
PNKP A12B4C3 PTEN 
  SHP1 
FEN1 NSC-281680 BRCA1 
 SC13 BRCA2 
 FENi#2 CDC4 
  MRE11 
Polβ NSC666715  
 Pro13  
 Natamycin  
LigIIIα/LigI L67  
LigI L82  
PARP1/PARP2 Olaparib BRCA1 
 Veliparib BRCA2 
 Talazoparib FANCG 
 Niraparib  
 Rucaparib  
PARG PDD00017272 BRCA1 
 PDD00017273 BRCA2 
 COH34 PALB2 
 JA2-4/JA2131 FAM175A 
  BARD1 
TargetInhibitorSynthetic lethal partner
Apurinic/apyrimidinic site MX (TRC102)  
BER enzymes 
APE1   
lyase activity CRT0044876 PTEN 
 API3 BRCA1 
  BRCA2 
  ATM 
Redox function E3330  
 Gossypol/AT101  
PNKP A12B4C3 PTEN 
  SHP1 
FEN1 NSC-281680 BRCA1 
 SC13 BRCA2 
 FENi#2 CDC4 
  MRE11 
Polβ NSC666715  
 Pro13  
 Natamycin  
LigIIIα/LigI L67  
LigI L82  
PARP1/PARP2 Olaparib BRCA1 
 Veliparib BRCA2 
 Talazoparib FANCG 
 Niraparib  
 Rucaparib  
PARG PDD00017272 BRCA1 
 PDD00017273 BRCA2 
 COH34 PALB2 
 JA2-4/JA2131 FAM175A 
  BARD1 

UNG inhibitors

In humans, there are two isoforms of uracil DNA glycosylase (UNG) that differ in localisation, as UNG1 is mitochondrial whereas UNG2 is nuclear, and these enzymes recognise U:A and U:G pairs in double-stranded DNA. When there are high levels of uracil incorporation in DNA, UNG activity is toxic as repeated attempts to excise the lesion result in an increase in strand breaks. The folate analogue pemetrexed, which inhibits thymidylate synthase to decrease dTTP levels and thus increases uracil misincorporation, is particularly effective in UNG-deficient colon and lung cancer cell lines [14]. Indeed, drug resistance can be induced by up-regulation of UNG expression, and pemetrexed sensitivity restored by using methoxyamine (MX), another BER inhibitor [14,18]. Thus, inhibiting BER could prevent the development of tumours resistant to folate analogues.

Targeting UNG has been suggested to cause synthetic lethality in the many cancer cells that have high APOBEC3B levels (e.g. bladder, cervix, lung, breast, and head and neck cancers) [19]. APOBEC3B is a cytosine deaminase that converts cytosine to uracil and causes an accumulation of C to T signature mutations in cancer genomes. A knockout of UNG has been demonstrated to kill APOBEC3B expressing cells due to an accumulation of uracil lesions in a mechanism dependent on non-canonical mismatch repair (MMR) [20]. Consequently, there is a need to develop specific small molecule inhibitors to UNG. Progress has been made from alkylated uracils that attach to the enzyme active site with sub-micromolar IC50 [21], to identifying small molecule inhibitors using uracil substrate fragment-linked to a library of aldehyde tethers [21,22]. Disappointingly, successful potency in cancer cell lines has not been reported with the UNG inhibitors and their action has only been demonstrated in cell free systems.

OGG1 inhibitors

The bifunctional N-glycosylase/DNA lyase, 8-oxoguanine DNA glycosylase 1 (OGG1) recognises and removes 7,8-dihydro-8-oxoguanine (8-oxoG) and 2,6-diamino-4-hydroxy-5-N-methylformamidopyrimidine (Fapy) from DNA. The potential for OGG1 inhibitors for use as a monotherapy in cancer treatment has been shown by the increased sensitivity of cells from patients with MMR deficiency that accumulate high levels of 8-oxoG [23], identifying a synthetic lethal relationship between MMR and BER.

Several efforts for developing small molecule inhibitors to OGG1 have been made. A hydrazide compound (O8) inhibited OGG1 with an IC50 of 0.22–0.35 μM, and had lesser effect on NEIL1 and NTH1 glycosylases that have overlapping substrate specificity [24]. Another inhibitor, SU0268, had a lower IC50 (59 nM), good permeability and low cytotoxicity on normal cells where an increase in genomic 8-oxoG was demonstrated [25]. A dual inhibitor (SU0383) was subsequently developed that would also inhibit MutT human homolog-1 (MTH1; aka NUDT) whose major substrate is 8-oxoG nucleotides [26]. By inhibiting both enzymes that clear genomic 8-oxoG and oxidized bases from the nucleotide pool, the increased oxidation load would tip the cancer cells into apoptosis, in the same way as proposed by MTH1 inhibitor alone [27,28]. The effect of these OGG1 inhibitors in cellular and animal cancer models is currently being pursued.

A small molecule OGG1 inhibitor, TH5487, with a different mechanism has recently been developed that inhibits the binding of OGG1 to 8-oxoG rather than catalysis [29]. TH5487 is a promising anti-inflammatory drug as it prevents the transcription of inflammatory response genes through deficiencies in 8-oxoG repair [29]. However, the utilisation of TH5487 within cancer cell models has yet to be reported.

NEIL1 inhibitors

Identification of other DNA glycosylases as potential drug targets was investigated using an siRNA screening approach which looked for increased sensitivity to uracil incorporation by various folate and nucleotide analogues. NEIL1 (and OGG1) siRNA-mediated depletion increased the cytotoxicity of thymidylate synthetase inhibitors in U2OS cells [30]. In the same study, depletion of NTH1, MPG, SMUG1 and TDG were only weakly synergistic. Furthermore, siRNA knockdown of NEIL1 (and also PARP1) was synthetic lethal with FANCG loss (involved in DNA interstrand cross-link repair), sporadic mutations of which occur in several cancers [31]. The development of inhibitors to NEIL1 has led to identification of purine analogues, specifically derivatives of 2-thioxanthine (2TX), that are irreversible inhibitors to NEIL1 and effective in vitro [32,33]. However, the development of NEIL1 inhibitors is still in its infancy and we await results using appropriate cancer cell and animal models.

APE1 and AP site inhibition

APE1 is an essential enzyme in BER that recognises and incises AP sites creating SSBs, but it also has 3′-phosphodiesterase activity that can remove terminal lesions such as phosphoglycolate that are formed following certain chemotherapy treatments and ionising radiation [34,35]. Thus inhibiting APE1 function is an attractive strategy to increase efficacy of radiotherapy, chemotherapy drugs and overcoming drug resistance. MX (or TRC102) does not inhibit APE1 directly, but can bind and modify AP sites making them refractory to APE1 binding. MX can also block the lyase activity of bifunctional glycosylases. MX has completed phase I clinical trials for use as a chemosensitizing agent with the antifolate pemetrexed in solid tumours [36] and with fludarabine for lymphomas, the outcome being that the combinations were well tolerated [37]. Other phase I and phase II studies are ongoing.

Several APE1 inhibitors have been generated and explored, although problems with permeability or poor potentiation have been reported (reviewed by [38]). For example, the inhibitor CRT0044876 with an IC50 of ∼3 µM reduced survival of HT1080 fibrosarcoma cells in combination with methylmethanesulfonate (MMS) and TMZ, but not to ionizing radiation [39]. The APE1 inhibitor compound III (API3) with an IC50 of 2–12 µM also sensitized HeLa cells effectively to MMS and TMZ, and was tolerated in mice [40]. However, a more suitable inhibitor is still needed for use in cancer therapy. Interestingly, APE1 inhibition has been found to be synthetically lethal in PTEN deficient melanoma cancer cell lines [41] and also with DSB repair proteins, specifically in BRCA1/2 and ATM-deficient cell lines [42]. Additionally, the interaction with nucleophosmin (NPM1) appears to be important in the response to platinum-based drugs as high expression levels of APE1 and NPM1 predict poor response to treatment. Thus, targeting APE1 activity, or its interaction with NPM1, might sensitize certain cancers expressing higher levels of these proteins [43]. These strategies, however, require further investigation.

APE1 is also known as Redox factor 1 (Ref1), and can respond to altered oxidative states by its cysteine-rich redox domain where it activates the DNA binding of certain transcription factors (e.g. NF-κB, p53, STAT3 and HIF-1α) [44]. Inhibitors have been identified to target the redox function of APE1 (reviewed by [38]). For example, E3330 (an NF-κB inhibitor) and Gossypol/AT101 (BCL2 inhibitors) can bind APE1 making it less redox active [45,46]. These agents can induce cytotoxicity as single agents [47] or in combination [46,48] in lung, lymphoma, prostate, adrenocortical and glioblastoma cancers, and more than 20 clinical trials are currently investigating their suitability.

PNKP inhibitors

Polynucleotide kinase phosphatase (PNKP) has dual 5′-kinase and 3′-phosphatase activities on SSBs and DSBs. Imidopiperidines have been identified as non-competitive inhibitors to the DNA–PNKP complex specifically inhibiting the phosphatase reaction [49,50]. One such inhibitor, A12B4C3, with an IC50 of ∼10 µM acts as a radiosensitizer in response to densely ionising carbon ion irradiation in prostate cancer cells [51], and Auger-emitting radioimmunotherapy in human myeloid leukaemia cells [52]. Synthetic lethal partnerships of PNKP with PTEN in colon cancer cell lines, and with the protein tyrosine phosphatase SHP-1 in T-cell lymphoma cell line have been identified, and where the combination of deficiencies in these proteins with A12B4C3 is effective in cell killing [53,54]. Interestingly, methods of delivering inhibitors of PNKP directly to the tumour through micelle encapsulation of the drug have been described, which are capable of radiosensitizing colon cancer cells [55]. This could potentially avoid any sensitisation of normal cells in proximity to the tumour being treated.

FEN1 inhibitors

The small molecule inhibitor of FEN1, NSC-281680 with an IC50 of 1.2 µM, sensitized MMR-proficient and deficient colon cancer cells to TMZ [56]. Whereas the FEN1 inhibitor, SC13 with an IC50 of 4.2 µM, supressed growth of breast cancer cell lines and also sensitised cells to cisplatin, 5-FU and TMZ [57]. Several synthetic lethal partners for FEN1 have been identified, including MRE11 and CDC4-deficient colorectal cancers [58], and BRCA1/2-deficient cells [59], which both respond to small molecule inhibitors of FEN1. Recently, FEN1 expression was found to be a predictive marker for resistance to tamoxifen in ERα-positive breast cancers, and that a novel FEN1 inhibitor (FENi#2) reduced breast cancer cell proliferation in vitro, even in tamoxifen resistant cell lines [60]. Thus, FEN1 inhibitors appear to show great potential.

DNA polymerase β inhibitors

Polβ is a key player in BER and is therefore an attractive target for chemosensitization of cancer cells. Many inhibitors to Polβ have been developed since the 1990s, although unfortunately most are non-specific (e.g. also targeting other DNA polymerases) or not potent or soluble enough to enter the cell or to be used clinically (see review [61]). However, a small molecule inhibitor of Polβ, NSC666715 (with an IC50 of ∼4 µM), designed by in silico molecular docking blocks the strand-displacement activity of Polβ in LP-BER leading to AP site accumulation and S-phase cell cycle arrest in colorectal cancers [62]. NSC666715 also appeared to potentiate the effects of TMZ in inducing cellular senescence in these cell lines. More recently, Pro-13, an irreversible inhibitor of Polβ (and Polλ) with an IC50 of 0.4 µM demonstrated little cytotoxicity in HeLa cells, but had a large synergistic effect in combination with MMS [63]. Natamycin, an antibiotic/anti-fungal agent, has been shown to inhibit the strand displacement activity of Polβ (at 2–5 nM), and at higher (µM) concentrations inhibited both Polβ and LigI, and consequently reduced proliferation of androgen-depleted prostate cancer cell lines [64].

An alternative strategy to directly inhibiting Polβ is through targeting protein stability. We have described that the deubiquitylating enzyme ubiquitin specific protease 47 (USP47), controls the cellular protein levels of Polβ through ubiquitylation-dependent degradation [65]. An siRNA knockdown of USP47 led to reduced Polβ protein levels and increased sensitivity of HeLa cells to MMS and hydrogen peroxide. Importantly, USP47 shares structural similarity to USP7, which plays a major role in stabilizing the p53 tumour suppressor protein and where inhibitors against USP7 are actively being sought. Indeed, inhibitors to both enzymes have been reported [66,67]. However, improvements in drug potency, solubility and stability are now required prior to detailed examination of USP7/USP47 inhibitors in impacting on BER and cancer cell survival.

DNA ligase inhibitors

The ligIII gene encodes a mitochondrial form of LigIIIα which is essential for cell survival, whereas the nuclear form is in complex with XRCC1 and is dispensable as LigI can compensate for its cellular role [68–70]. As with the DNA polymerases, obtaining inhibitors that are specific and potent has been difficult (comprehensively reviewed by [71]). Structure-based design of inhibitors have produced a series of compounds that are specific to LigI (L82, L82-G17), LigI and LigIIIα (L67) or all DNA ligases (L189) with IC50 of ∼10 μM [72–74]. L82 was cytostatic, whereas L67 and L189 were cytotoxic in MCF7, HeLa and HCT116 cells. Interestingly, at subtoxic levels L67 and L189 were found to increase the sensitivity of MCF7 breast cancer cell lines to MMS or ionising radiation, but had no impact on sensitisation of normal breast cell lines. LigI is often elevated in cancer cells due to hyper proliferation, and also the levels of other DNA ligases may be dysregulated that may explain this apparent selectivity of the inhibitors [75]. Thus, ligase inhibitors have an important place in functional studies and cancer therapy, though not necessarily as a result of BER targeting.

PARP inhibitors

PARP1 (and its associated backup enzyme PARP2) recognise gaps and SSBs in the DNA backbone and catalyse the addition of ADP-ribose units to itself or other proteins from cellular NAD+ (Figure 2). PARP inhibitors (PARPi) have been developed to inhibit PARP1, PARP2 (and PARP3) with nanomolar IC50 by binding to the NAD+ binding site in the catalytic domain, but can vary in selectivity among other PARP family members [76]. It should be noted that PARPi would also affect the response of PARP1 to DSBs and of PARP3 activity on recognition of both SSBs and DSBs [77–79]. The formation of poly(ADP-ribose) can relax chromatin structure and is recognised by a variety of poly(ADP-ribose) binding motifs located in a number of DNA repair factors and chromatin remodelling factors (reviewed by [80]). Importantly, XRCC1 is localised to DNA damage through its interaction with poly(ADP-ribose) and DNA mediated through sites on either side of its BRCT1 domain [81]. XRCC1 acts as a scaffold protein and provides a platform for Polβ, LigIIIα, PNKP, aprataxin and APLF binding. Thus, PARPi can disrupt the coordination of DNA repair proteins, chromatin accessibility and chromatin remodelling.

Functions of PARP1 and PARG during BER

Figure 2
Functions of PARP1 and PARG during BER

PARP1 (pink) recognises SSB/gap intermediates and uses NAD+ (red hexagon) to synthesise linear and branched chains of ADP-ribose units (red circles) to itself and/or other proteins. The BRCT1 domain of XRCC1 can bind poly(ADP-ribose), and make DNA contacts, allowing access for Polβ and LigIIIα to repair the break. Accumulation of PARP1 poly(ADP-ribosyl)ation causes PARP1 to be released through electrostatic repulsion. PARG subsequently cleaves the poly(ADP-ribose) chains allowing PARP1 to bind to additional SSB/gaps. The terminal ADP-ribose unit is refractory to PARG action, so ADP-ribose hydrolases (ARH) are needed for complete removal. PARPi (red cross) can lead to a trapped PARP–DNA complex that interferes with DNA replication.

Figure 2
Functions of PARP1 and PARG during BER

PARP1 (pink) recognises SSB/gap intermediates and uses NAD+ (red hexagon) to synthesise linear and branched chains of ADP-ribose units (red circles) to itself and/or other proteins. The BRCT1 domain of XRCC1 can bind poly(ADP-ribose), and make DNA contacts, allowing access for Polβ and LigIIIα to repair the break. Accumulation of PARP1 poly(ADP-ribosyl)ation causes PARP1 to be released through electrostatic repulsion. PARG subsequently cleaves the poly(ADP-ribose) chains allowing PARP1 to bind to additional SSB/gaps. The terminal ADP-ribose unit is refractory to PARG action, so ADP-ribose hydrolases (ARH) are needed for complete removal. PARPi (red cross) can lead to a trapped PARP–DNA complex that interferes with DNA replication.

Close modal

PARP1 was first demonstrated as a synthetic lethal partner to BRCA1 and BRCA2 mutations in breast cancer cells that are unable to effectively perform homologous recombination (HR) [82,83]. PARPi (olaparib, rucaparib, niraparib and talazoparib) have since been approved for use as monotherapy agents for BRCA-mutated breast, ovarian cancer, fallopian tube and peritoneal cancers. Many other synthetic lethal partners for PARPi have been identified and not just limited to HR, increasing the potential use of PARPi as a synthetic lethal agent [84]. Interestingly, there is evidence that PARPi can enhance the radiosensitivity of cell lines from different tumour types that appear to display proficient-HR mechanisms, including head and neck cancer, and glioblastoma [85–87]. Indeed, PARPi are currently in a large number of clinical trials for combination therapies as radio/chemosensitizers for a variety of other cancer types (e.g. prostate, gastric, haematological, lung, brain, head and neck, colorectal and advanced solid tumours). However, there is some debate as to whether sensitization occurs via BER or other PARP1-dependent functions [87–89].

The cytotoxicity of the inhibitors differs considerably (talazoparib >olaparib>veliparib) and depends on specificity, potency of catalytical inhibition, pharmacodynamic/kinetic properties and ‘trapping’ ability [90]. In the absence of auto(ADP-ribosyl)ation, PARP1 and PARP2 inhibited by talazoparib or olaparib remain bound to DNA in trapped complexes that cause PARP retention on chromatin (Figure 2) [91,92], whereas some inhibitors cause an allosteric change that releases the PARP from DNA (e.g. veliparib) [90]. An emerging problem with PARPi is the development of drug resistance largely through restoration of HR by secondary mutations in BRCA proteins or proteins that favour HR pathway choice (reviewed in [93]). PARP1 mutations and down-regulation also occur in PARPi-resistant cells [94]. Alterations in miRNA expression can produce similar outcomes on HR/PARP activity. Interestingly, poly(ADP-ribose) glycohydrolase (PARG) loss can also cause drug resistance presumably by increasing poly(ADP-ribosyl)ation levels [95]. Thus, improving cytotoxicity through drug design and overcoming drug resistance will improve the efficacy of PARPi use in the future.

PARG inhibitors

PARG is an essential protein required for the breakdown of poly(ADP-ribose) chains, and the recycling of PARP1 (and PARP2) for co-ordinating additional BER activity (Figure 2). Like PARPi, PARG loss is synthetic lethal with BRCA2 mutations as well as with other partners involved in HR, including BRCA1, PALB2, FAM175A and BARD1 [96,97]. The specific PARG inhibitor (PARGi), PDD00017273, led to an increase in the number of stalled replication forks requiring HR for repair, and ultimately enhanced death of MCF7 cells. The PARGi was also similarly effective as PARPi in radiosensitising MCF7 cells, but the mechanism of sensitization differed as this occurred through altering mitosis [98].

PARG activity can also regulate the activity of a number of transcription factors [99]. For example, PARG activity can increase the expression of androgen receptor by removing inhibitory poly(ADP-ribosyl)ation from the transcriptional coactivator KDM4D. PARGi, PDD00017272, enhanced the effect of androgen ablation on prostate cancer cell lines, by further reducing androgen receptor signalling in addition to increased cytotoxic breaks arising from the inhibition of BER [100]. The development of other PARGi (e.g. COH34, JA2-4 and JA2131) to exacerbate replicative stress are proving promising as synthetic lethal agents, chemosensitizers and for re-sensitizing PARPi-resistant cells [101,102].

BER is a critical cellular DNA repair pathway responding to DNA base damage and SSBs, and we describe here some studies where BER inhibitors have shown promise as radio/chemosensitizers in several cancers or form synthetic lethal partnerships with common cancer mutations. PARPi in particular have achieved the greatest success as approved monotherapy agents and are also in clinical trials as radio/chemosensitizers. The AP site inhibitor MX and APE1/BCL2 redox inhibitors Gossypol/AT101, are also currently in clinical trials but other inhibitors appear to have fallen short thus far. So, why has BER inhibitor development and clinical use been so challenging?

A large number of the BER proteins are embryonic lethal in knockout mice (e.g. APE1, Polβ, LigIII, LigI and FEN1) suggesting that inhibitors to these proteins might be toxic to normal cells. The DNA glycosylases, on the other hand, are not embryonic lethal in mice (with the exception of TDG) as there is a degree of redundancy among these enzymes, which can therefore diminish the impact of any targeted drug. In addition, backup repair pathways exist for several DNA lesions processed by BER, namely, HR, MMR, nucleotide incision repair and nucleotide excision repair [103–106], which can also potentially reduce inhibitor efficacy. The development of specific inhibitors to many of the BER enzymes has been arduous because they belong to families of functionally diverse but structurally similar enzymes. Nevertheless, new strategies for designing inhibitors based on structural data and in-depth molecular mechanisms have made great advances in recent years, and which should be continually explored.

An impressive advancement of immunotherapy as an approach for effective cancer treatment has been made in recent years. Interestingly, recent reports suggest that PD-L1 expression is negatively correlated with BER gene expression, including OGG1 and APE1 [107], and that anti-PD-1 therapy in combination with ionising radiation is stimulated by PARPi in colorectal cancer models [108]. These intriguing findings have opened up new exciting therapeutic opportunities, and which nevertheless support that there should be ongoing research into targeting the BER both as a monotherapy but also as a combinatorial therapy for cancer treatment.

  • SP-BER and LP-BER are vital in excising damage to bases and repairing SSB in DNA, thereby reducing mutagenesis.

  • DNA damage caused by radiotherapy and many chemotherapeutics is required to exceed BER capacity for effectiveness.

  • Targeting BER enzymes can increase radio/chemosensitivity, re-sensitize drug resistant cancers or form part of a synthetic lethal combination with cancer mutations.

  • PARP inhibitors have proved successful in clinical trials, inhibitors that affect APE1 functions have progressed to phase II/III clinical trials, and several other BER enzymes remain promising targets.

The authors declare that there are no competing interests associated with the manuscript.

G.J.G. is the recipient of the North West Cancer Research Career Development Fellowship [grant number CDF2019.05]. J.L.P. is currently funded by North West Cancer Research [grant numbers CR1145 and CR1197].

Open access for this article was enabled by the participation of University of Liverpool in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.

8-oxoG

7,8-Dihydro-8-oxoguanine

APE1

AP endonuclease 1

AP

apurinic/apyrimidinic

BER

base excision repair

dRP

5′-deoxyribosephosphate

Fapy

2,6-diamino-4-hydroxy-5-N-methylformamidopyrimidine

FEN1

Flap endonuclease 1

HR

homologous recombination

LigI

DNA ligase I

LigIIIα

DNA ligase IIIα

MMR

mismatch repair

MMS

methylmethanesulfonate

MX

methoxyamine

NEIL

endonuclease VIII-like

NPM1

nucleophosmin

OGG1

8-oxoguanine DNA glycosylase

PARG

poly(ADP-ribose) glycohydrolase

PARP1

poly(ADP-ribose) polymerase 1

PARPi

PARP inhibitor

PNKP

polynucleotide kinase phosphatase

Polβ

DNA polymerase β

Ref1

Redox factor 1

SSB

single strand break

TMZ

Temozolomide

UNG

uracil DNA glycosylase

XRCC1

X-ray repair cross-complementing protein 1

1.
Lindahl
T.
(
1993
)
Instability and decay of the primary structure of DNA
.
Nature
362
,
709
715
[PubMed]
2.
Jacobs
A.L.
and
Schar
P.
(
2012
)
DNA glycosylases: in DNA repair and beyond
.
Chromosoma
121
,
1
20
[PubMed]
3.
Robson
C.N.
and
Hickson
I.D.
(
1991
)
Isolation of cDNA clones encoding a human apurinic/apyrimidinic endonuclease that corrects DNA repair and mutagenesis defects in E. coli xth (exonuclease III) mutants
.
Nucleic Acids Res.
19
,
5519
5523
[PubMed]
4.
Demple
B.
,
Herman
T.
and
Chen
D.S.
(
1991
)
Cloning and expression of APE, the cDNA encoding the major human apurinic endonuclease: definition of a family of DNA repair enzymes
.
Proc. Natl. Acad. Sci. U.S.A.
88
,
11450
11454
[PubMed]
5.
Wiederhold
L.
,
Leppard
J.B.
,
Kedar
P.
,
Karimi-Busheri
F.
,
Rasouli-Nia
A.
,
Weinfeld
M.
et al.
(
2004
)
AP endonuclease-independent DNA base excision repair in human cells
.
Mol. Cell
15
,
209
220
[PubMed]
6.
Woodhouse
B.C.
,
Dianova
I.I.
,
Parsons
J.L.
and
Dianov
G.L.
(
2008
)
Poly(ADP-ribose) polymerase-1 modulates DNA repair capacity and prevents formation of DNA double strand breaks
.
DNA Repair (Amst.)
7
,
932
940
[PubMed]
7.
Matsumoto
Y.
and
Kim
K.
(
1995
)
Excision of deoxyribose phosphate residues by DNA polymerase beta during DNA repair
.
Science
269
,
699
702
[PubMed]
8.
Sobol
R.W.
,
Horton
J.K.
,
Kuhn
R.
,
Gu
H.
,
Singhal
R.K.
,
Prasad
R.
et al.
(
1996
)
Requirement of mammalian DNA polymerase-beta in base-excision repair
.
Nature
379
,
183
186
[PubMed]
9.
Nash
R.A.
,
Caldecott
K.W.
,
Barnes
D.E.
and
Lindahl
T.
(
1997
)
XRCC1 protein interacts with one of two distinct forms of DNA ligase III
.
Biochemistry
36
,
5207
5211
[PubMed]
10.
Cappelli
E.
,
Taylor
R.
,
Cevasco
M.
,
Abbondandolo
A.
,
Caldecott
K.
and
Frosina
G.
(
1997
)
Involvement of XRCC1 and DNA ligase III gene products in DNA base excision repair
.
J. Biol. Chem.
272
,
23970
23975
[PubMed]
11.
Frosina
G.
,
Fortini
P.
,
Rossi
O.
,
Carrozzino
F.
,
Raspaglio
G.
,
Cox
L.S.
et al.
(
1996
)
Two pathways for base excision repair in mammalian cells
.
J. Biol. Chem.
271
,
9573
9578
[PubMed]
12.
Podlutsky
A.J.
,
Dianova
I.I.
,
Podust
V.N.
,
Bohr
V.A.
and
Dianov
G.L.
(
2001
)
Human DNA polymerase beta initiates DNA synthesis during long-patch repair of reduced AP sites in DNA
.
EMBO J.
20
,
1477
1482
[PubMed]
13.
Longley
D.B.
,
Harkin
D.P.
and
Johnston
P.G.
(
2003
)
5-fluorouracil: mechanisms of action and clinical strategies
.
Nat. Rev. Cancer
3
,
330
338
[PubMed]
14.
Weeks
L.D.
,
Fu
P.
and
Gerson
S.L.
(
2013
)
Uracil-DNA glycosylase expression determines human lung cancer cell sensitivity to pemetrexed
.
Mol. Cancer Ther.
12
,
2248
2260
[PubMed]
15.
Orta
M.L.
,
Hoglund
A.
,
Calderon-Montano
J.M.
,
Dominguez
I.
,
Burgos-Moron
E.
,
Visnes
T.
et al.
(
2014
)
The PARP inhibitor Olaparib disrupts base excision repair of 5-aza-2′-deoxycytidine lesions
.
Nucleic Acids Res.
42
,
9108
9120
[PubMed]
16.
Kothandapani
A.
,
Dangeti
V.S.
,
Brown
A.R.
,
Banze
L.A.
,
Wang
X.H.
,
Sobol
R.W.
et al.
(
2011
)
Novel role of base excision repair in mediating cisplatin cytotoxicity
.
J. Biol. Chem.
286
,
14564
14574
[PubMed]
17.
Caiola
E.
,
Salles
D.
,
Frapolli
R.
,
Lupi
M.
,
Rotella
G.
,
Ronchi
A.
et al.
(
2015
)
Base excision repair-mediated resistance to cisplatin in KRAS(G12C) mutant NSCLC cells
.
Oncotarget
6
,
30072
30087
[PubMed]
18.
Bulgar
A.D.
,
Weeks
L.D.
,
Miao
Y.
,
Yang
S.
,
Xu
Y.
,
Guo
C.
et al.
(
2012
)
Removal of uracil by uracil DNA glycosylase limits pemetrexed cytotoxicity: overriding the limit with methoxyamine to inhibit base excision repair
.
Cell Death Dis.
3
,
e252
[PubMed]
19.
Burns
M.B.
,
Temiz
N.A.
and
Harris
R.S.
(
2013
)
Evidence for APOBEC3B mutagenesis in multiple human cancers
.
Nat. Genet.
45
,
977
983
[PubMed]
20.
Serebrenik
A.A.
,
Starrett
G.J.
,
Leenen
S.
,
Jarvis
M.C.
,
Shaban
N.M.
,
Salamango
D.J.
et al.
(
2019
)
The deaminase APOBEC3B triggers the death of cells lacking uracil DNA glycosylase
.
Proc. Natl. Acad. Sci. U.S.A.
116
,
22158
22163
[PubMed]
21.
Jiang
Y.L.
,
Krosky
D.J.
,
Seiple
L.
and
Stivers
J.T.
(
2005
)
Uracil-directed ligand tethering: an efficient strategy for uracil DNA glycosylase (UNG) inhibitor development
.
J. Am. Chem. Soc.
127
,
17412
17420
[PubMed]
22.
Chung
S.
,
Parker
J.B.
,
Bianchet
M.
,
Amzel
L.M.
and
Stivers
J.T.
(
2009
)
Impact of linker strain and flexibility in the design of a fragment-based inhibitor
.
Nat. Chem. Biol.
5
,
407
413
[PubMed]
23.
Martin
S.A.
,
Lord
C.J.
and
Ashworth
A.
(
2010
)
Therapeutic targeting of the DNA mismatch repair pathway
.
Clin. Cancer Res.
16
,
5107
5113
[PubMed]
24.
Donley
N.
,
Jaruga
P.
,
Coskun
E.
,
Dizdaroglu
M.
,
McCullough
A.K.
and
Lloyd
R.S.
(
2015
)
Small Molecule Inhibitors of 8-Oxoguanine DNA Glycosylase-1 (OGG1)
.
ACS Chem. Biol.
10
,
2334
2343
[PubMed]
25.
Tahara
Y.K.
,
Auld
D.
,
Ji
D.
,
Beharry
A.A.
,
Kietrys
A.M.
,
Wilson
D.L.
et al.
(
2018
)
Potent and Selective Inhibitors of 8-Oxoguanine DNA Glycosylase
.
J. Am. Chem. Soc.
140
,
2105
2114
[PubMed]
26.
Tahara
Y.K.
,
Kietrys
A.M.
,
Hebenbrock
M.
,
Lee
Y.
,
Wilson
D.L.
and
Kool
E.T.
(
2019
)
Dual Inhibitors of 8-Oxoguanine Surveillance by OGG1 and NUDT1
.
ACS Chem. Biol.
14
,
2606
2615
[PubMed]
27.
Gad
H.
,
Koolmeister
T.
,
Jemth
A.S.
,
Eshtad
S.
,
Jacques
S.A.
,
Strom
C.E.
et al.
(
2014
)
MTH1 inhibition eradicates cancer by preventing sanitation of the dNTP pool
.
Nature
508
,
215
221
[PubMed]
28.
Huber
K.V.
,
Salah
E.
,
Radic
B.
,
Gridling
M.
,
Elkins
J.M.
,
Stukalov
A.
et al.
(
2014
)
Stereospecific targeting of MTH1 by (S)-crizotinib as an anticancer strategy
.
Nature
508
,
222
227
[PubMed]
29.
Visnes
T.
,
Cazares-Korner
A.
,
Hao
W.
,
Wallner
O.
,
Masuyer
G.
,
Loseva
O.
et al.
(
2018
)
Small-molecule inhibitor of OGG1 suppresses proinflammatory gene expression and inflammation
.
Science
362
,
834
839
[PubMed]
30.
Taricani
L.
,
Shanahan
F.
,
Pierce
R.H.
,
Guzi
T.J.
and
Parry
D.
(
2010
)
Phenotypic enhancement of thymidylate synthetase pathway inhibitors following ablation of Neil1 DNA glycosylase/lyase
.
Cell Cycle
9
,
4876
4883
[PubMed]
31.
Kennedy
R.D.
,
Chen
C.C.
,
Stuckert
P.
,
Archila
E.M.
,
De la Vega
M.A.
,
Moreau
L.A.
et al.
(
2007
)
Fanconi anemia pathway-deficient tumor cells are hypersensitive to inhibition of ataxia telangiectasia mutated
.
J. Clin. Invest.
117
,
1440
1449
[PubMed]
32.
Rieux
C.
,
Goffinont
S.
,
Coste
F.
,
Tber
Z.
,
Cros
J.
,
Roy
V.
et al.
(
2020
)
Thiopurine Derivative-Induced Fpg/Nei DNA Glycosylase Inhibition: Structural, Dynamic and Functional Insights
.
Int. J. Mol. Sci.
21
,
2058
[PubMed]
33.
Jacobs
A.C.
,
Calkins
M.J.
,
Jadhav
A.
,
Dorjsuren
D.
,
Maloney
D.
,
Simeonov
A.
et al.
(
2013
)
Inhibition of DNA glycosylases via small molecule purine analogs
.
PLoS ONE
8
,
e81667
[PubMed]
34.
Parsons
J.L.
,
Dianova
I.I.
and
Dianov
G.L.
(
2004
)
APE1 is the major 3′-phosphoglycolate activity in human cell extracts
.
Nucleic Acids Res.
32
,
3531
3536
[PubMed]
35.
Parsons
J.L.
,
Dianova
I.I.
and
Dianov
G.L.
(
2005
)
APE1-dependent repair of DNA single-strand breaks containing 3′-end 8-oxoguanine
.
Nucleic Acids Res.
33
,
2204
2209
[PubMed]
36.
Gordon
M.S.
,
Rosen
L.S.
,
Mendelson
D.
,
Ramanathan
R.K.
,
Goldman
J.
,
Liu
L.
et al.
(
2013
)
A phase 1 study of TRC102, an inhibitor of base excision repair, and pemetrexed in patients with advanced solid tumors
.
Invest. New Drugs
31
,
714
723
[PubMed]
37.
Caimi
P.F.
,
Cooper
B.W.
,
William
B.M.
,
Dowlati
A.
,
Barr
P.M.
,
Fu
P.
et al.
(
2017
)
Phase I clinical trial of the base excision repair inhibitor methoxyamine in combination with fludarabine for patients with advanced hematologic malignancies
.
Oncotarget
8
,
79864
79875
[PubMed]
38.
Laev
S.S.
,
Salakhutdinov
N.F.
and
Lavrik
O.I.
(
2017
)
Inhibitors of nuclease and redox activity of apurinic/apyrimidinic endonuclease 1/redox effector factor 1 (APE1/Ref-1)
.
Bioorg. Med. Chem.
25
,
2531
2544
[PubMed]
39.
Madhusudan
S.
,
Smart
F.
,
Shrimpton
P.
,
Parsons
J.L.
,
Gardiner
L.
,
Houlbrook
S.
et al.
(
2005
)
Isolation of a small molecule inhibitor of DNA base excision repair
.
Nucleic. Acids. Res.
33
,
4711
4724
[PubMed]
40.
Rai
G.
,
Vyjayanti
V.N.
,
Dorjsuren
D.
,
Simeonov
A.
,
Jadhav
A.
,
Wilson
D.M.
III
et al.
(
2012
)
Synthesis, biological evaluation, and structure-activity relationships of a novel class of apurinic/apyrimidinic endonuclease 1 inhibitors
.
J. Med. Chem.
55
,
3101
3112
[PubMed]
41.
Abbotts
R.
,
Jewell
R.
,
Nsengimana
J.
,
Maloney
D.J.
,
Simeonov
A.
,
Seedhouse
C.
et al.
(
2014
)
Targeting human apurinic/apyrimidinic endonuclease 1 (APE1) in phosphatase and tensin homolog (PTEN) deficient melanoma cells for personalized therapy
.
Oncotarget
5
,
3273
3286
[PubMed]
42.
Sultana
R.
,
McNeill
D.R.
,
Abbotts
R.
,
Mohammed
M.Z.
,
Zdzienicka
M.Z.
,
Qutob
H.
et al.
(
2012
)
Synthetic lethal targeting of DNA double-strand break repair deficient cells by human apurinic/apyrimidinic endonuclease inhibitors
.
Int. J. Cancer
131
,
2433
2444
[PubMed]
43.
Malfatti
M.C.
,
Gerratana
L.
,
Dalla
E.
,
Isola
M.
,
Damante
G.
,
Di Loreto
C.
et al.
(
2019
)
APE1 and NPM1 protect cancer cells from platinum compounds cytotoxicity and their expression pattern has a prognostic value in TNBC
.
J. Exp. Clin. Cancer Res.
38
,
309
[PubMed]
44.
Tell
G.
,
Quadrifoglio
F.
,
Tiribelli
C.
and
Kelley
M.R.
(
2009
)
The many functions of APE1/Ref-1: not only a DNA repair enzyme
.
Antioxid. Redox Signal.
11
,
601
620
[PubMed]
45.
Zhang
J.
,
Luo
M.
,
Marasco
D.
,
Logsdon
D.
,
LaFavers
K.A.
,
Chen
Q.
et al.
(
2013
)
Inhibition of apurinic/apyrimidinic endonuclease I's redox activity revisited
.
Biochemistry
52
,
2955
2966
[PubMed]
46.
Wei
X.
,
Duan
W.
,
Li
Y.
,
Zhang
S.
,
Xin
X.
,
Sun
L.
et al.
(
2016
)
AT101 exerts a synergetic efficacy in gastric cancer patients with 5-FU based treatment through promoting apoptosis and autophagy
.
Oncotarget
7
,
34430
34441
[PubMed]
47.
Luo
M.
,
Delaplane
S.
,
Jiang
A.
,
Reed
A.
,
He
Y.
,
Fishel
M.
et al.
(
2008
)
Role of the multifunctional DNA repair and redox signaling protein Ape1/Ref-1 in cancer and endothelial cells: small-molecule inhibition of the redox function of Ape1
.
Antioxid. Redox Signal.
10
,
1853
1867
[PubMed]
48.
Ren
T.
,
Shan
J.
,
Li
M.
,
Qing
Y.
,
Qian
C.
,
Wang
G.
et al.
(
2015
)
Small-molecule BH3 mimetic and pan-Bcl-2 inhibitor AT-101 enhances the antitumor efficacy of cisplatin through inhibition of APE1 repair and redox activity in non-small-cell lung cancer
.
Drug Des. Devel. Ther.
9
,
2887
2910
[PubMed]
49.
Freschauf
G.K.
,
Karimi-Busheri
F.
,
Ulaczyk-Lesanko
A.
,
Mereniuk
T.R.
,
Ahrens
A.
,
Koshy
J.M.
et al.
(
2009
)
Identification of a small molecule inhibitor of the human DNA repair enzyme polynucleotide kinase/phosphatase
.
Cancer Res.
69
,
7739
7746
[PubMed]
50.
Freschauf
G.K.
,
Mani
R.S.
,
Mereniuk
T.R.
,
Fanta
M.
,
Virgen
C.A.
,
Dianov
G.L.
et al.
(
2010
)
Mechanism of action of an imidopiperidine inhibitor of human polynucleotide kinase/phosphatase
.
J. Biol. Chem.
285
,
2351
2360
[PubMed]
51.
Srivastava
P.
,
Sarma
A.
and
Chaturvedi
C.M.
(
2018
)
Targeting DNA repair with PNKP inhibition sensitizes radioresistant prostate cancer cells to high LET radiation
.
PLoS ONE
13
,
e0190516
[PubMed]
52.
Zereshkian
A.
,
Leyton
J.V.
,
Cai
Z.
,
Bergstrom
D.
,
Weinfeld
M.
and
Reilly
R.M.
(
2014
)
The human polynucleotide kinase/phosphatase (hPNKP) inhibitor A12B4C3 radiosensitizes human myeloid leukemia cells to Auger electron-emitting anti-CD123 (1)(1)(1)In-NLS-7G3 radioimmunoconjugates
.
Nucl. Med. Biol.
41
,
377
383
[PubMed]
53.
Mereniuk
T.R.
,
El Gendy
M.A.
,
Mendes-Pereira
A.M.
,
Lord
C.J.
,
Ghosh
S.
,
Foley
E.
et al.
(
2013
)
Synthetic lethal targeting of PTEN-deficient cancer cells using selective disruption of polynucleotide kinase/phosphatase
.
Mol. Cancer Ther.
12
,
2135
2144
[PubMed]
54.
Mereniuk
T.R.
,
Maranchuk
R.A.
,
Schindler
A.
,
Penner-Chea
J.
,
Freschauf
G.K.
,
Hegazy
S.
et al.
(
2012
)
Genetic screening for synthetic lethal partners of polynucleotide kinase/phosphatase: potential for targeting SHP-1-depleted cancers
.
Cancer Res.
72
,
5934
5944
[PubMed]
55.
Shire
Z.
,
Vakili
M.R.
,
Morgan
T.D.R.
,
Hall
D.G.
,
Lavasanifar
A.
and
Weinfeld
M.
(
2018
)
Nanoencapsulation of Novel Inhibitors of PNKP for Selective Sensitization to Ionizing Radiation and Irinotecan and Induction of Synthetic Lethality
.
Mol. Pharm.
15
,
2316
2326
[PubMed]
56.
Panda
H.
,
Jaiswal
A.S.
,
Corsino
P.E.
,
Armas
M.L.
,
Law
B.K.
and
Narayan
S.
(
2009
)
Amino acid Asp181 of 5′-flap endonuclease 1 is a useful target for chemotherapeutic development
.
Biochemistry
48
,
9952
9958
[PubMed]
57.
He
L.
,
Zhang
Y.
,
Sun
H.
,
Jiang
F.
,
Yang
H.
,
Wu
H.
et al.
(
2016
)
Targeting DNA Flap Endonuclease 1 to Impede Breast Cancer Progression
.
EBioMedicine
14
,
32
43
[PubMed]
58.
van Pel
D.M.
,
Barrett
I.J.
,
Shimizu
Y.
,
Sajesh
B.V.
,
Guppy
B.J.
,
Pfeifer
T.
et al.
(
2013
)
An evolutionarily conserved synthetic lethal interaction network identifies FEN1 as a broad-spectrum target for anticancer therapeutic development
.
PLos Genet.
9
,
e1003254
[PubMed]
59.
Mengwasser
K.E.
,
Adeyemi
R.O.
,
Leng
Y.
,
Choi
M.Y.
,
Clairmont
C.
,
D'Andrea
A.D.
et al.
(
2019
)
Genetic Screens Reveal FEN1 and APEX2 as BRCA2 Synthetic Lethal Targets
.
Mol. Cell
73
,
885e6
899e6
60.
Flach
K.D.
,
Periyasamy
M.
,
Jadhav
A.
,
Dorjsuren
D.
,
Siefert
J.C.
,
Hickey
T.E.
et al.
(
2020
)
Endonuclease FEN1 coregulates ERalpha activity and provides a novel drug interface in tamoxifen resistant breast cancer
.
Cancer Res.
80
,
1914
1926
[PubMed]
61.
Barakat
K.H.
,
Gajewski
M.M.
and
Tuszynski
J.A.
(
2012
)
DNA polymerase beta (pol beta) inhibitors: a comprehensive overview
.
Drug Discov. Today
17
,
913
920
[PubMed]
62.
Jaiswal
A.S.
,
Panda
H.
,
Law
B.K.
,
Sharma
J.
,
Jani
J.
,
Hromas
R.
et al.
(
2015
)
NSC666715 and Its Analogs Inhibit Strand-Displacement Activity of DNA Polymerase beta and Potentiate Temozolomide-Induced DNA Damage, Senescence and Apoptosis in Colorectal Cancer Cells
.
PLoS ONE
10
,
e0123808
[PubMed]
63.
Paul
R.
,
Banerjee
S.
and
Greenberg
M.M.
(
2017
)
Synergistic Effects of an Irreversible DNA Polymerase Inhibitor and DNA Damaging Agents on HeLa Cells
.
ACS Chem. Biol.
12
,
1576
1583
[PubMed]
64.
Vasquez
J.L.
,
Lai
Y.
,
Annamalai
T.
,
Jiang
Z.
,
Zhang
M.
,
Lei
R.
et al.
(
2020
)
Inhibition of base excision repair by natamycin suppresses prostate cancer cell proliferation
.
Biochimie
168
,
241
250
[PubMed]
65.
Parsons
J.L.
,
Dianova
I.I.
,
Khoronenkova
S.V.
,
Edelmann
M.J.
,
Kessler
B.M.
and
Dianov
G.L.
(
2011
)
USP47 is a deubiquitylating enzyme that regulates base excision repair by controlling steady-state levels of DNA polymerase beta
.
Mol. Cell
41
,
609
615
[PubMed]
66.
Weinstock
J.
,
Wu
J.
,
Cao
P.
,
Kingsbury
W.D.
,
McDermott
J.L.
,
Kodrasov
M.P.
et al.
(
2012
)
Selective Dual Inhibitors of the Cancer-Related Deubiquitylating Proteases USP7 and USP47
.
ACS Med. Chem. Lett.
3
,
789
792
[PubMed]
67.
Altun
M.
,
Kramer
H.B.
,
Willems
L.I.
,
McDermott
J.L.
,
Leach
C.A.
,
Goldenberg
S.J.
et al.
(
2011
)
Activity-based chemical proteomics accelerates inhibitor development for deubiquitylating enzymes
.
Chem. Biol.
18
,
1401
1412
[PubMed]
68.
Simsek
D.
,
Furda
A.
,
Gao
Y.
,
Artus
J.
,
Brunet
E.
,
Hadjantonakis
A.K.
et al.
(
2011
)
Crucial role for DNA ligase III in mitochondria but not in Xrcc1-dependent repair
.
Nature
471
,
245
248
[PubMed]
69.
Gao
Y.
,
Katyal
S.
,
Lee
Y.
,
Zhao
J.
,
Rehg
J.E.
,
Russell
H.R.
et al.
(
2011
)
DNA ligase III is critical for mtDNA integrity but not Xrcc1-mediated nuclear DNA repair
.
Nature
471
,
240
244
[PubMed]
70.
Sallmyr
A.
,
Matsumoto
Y.
,
Roginskaya
V.
,
Van Houten
B.
and
Tomkinson
A.E.
(
2016
)
Inhibiting Mitochondrial DNA Ligase IIIalpha Activates Caspase 1-Dependent Apoptosis in Cancer Cells
.
Cancer Res.
76
,
5431
5441
[PubMed]
71.
Singh
D.K.
,
Krishna
S.
,
Chandra
S.
,
Shameem
M.
,
Deshmukh
A.L.
and
Banerjee
D.
(
2014
)
Human DNA ligases: a comprehensive new look for cancer therapy
.
Med. Res. Rev.
34
,
567
595
[PubMed]
72.
Chen
X.
,
Zhong
S.
,
Zhu
X.
,
Dziegielewska
B.
,
Ellenberger
T.
,
Wilson
G.M.
et al.
(
2008
)
Rational design of human DNA ligase inhibitors that target cellular DNA replication and repair
.
Cancer Res.
68
,
3169
3177
[PubMed]
73.
Srivastava
M.
,
Nambiar
M.
,
Sharma
S.
,
Karki
S.S.
,
Goldsmith
G.
,
Hegde
M.
et al.
(
2012
)
An inhibitor of nonhomologous end-joining abrogates double-strand break repair and impedes cancer progression
.
Cell
151
,
1474
1487
[PubMed]
74.
Howes
T.R.L.
,
Sallmyr
A.
,
Brooks
R.
,
Greco
G.E.
,
Jones
D.E.
,
Matsumoto
Y.
et al.
(
2017
)
Structure-activity relationships among DNA ligase inhibitors: Characterization of a selective uncompetitive DNA ligase I inhibitor
.
DNA Repair (Amst.)
60
,
29
39
[PubMed]
75.
Tomkinson
A.E.
,
Naila
T.
and
Khattri Bhandari
S.
(
2020
)
Altered DNA ligase activity in human disease
.
Mutagenesis
35
,
51
60
[PubMed]
76.
Wahlberg
E.
,
Karlberg
T.
,
Kouznetsova
E.
,
Markova
N.
,
Macchiarulo
A.
,
Thorsell
A.G.
et al.
(
2012
)
Family-wide chemical profiling and structural analysis of PARP and tankyrase inhibitors
.
Nat. Biotechnol.
30
,
283
288
[PubMed]
77.
Grundy
G.J.
,
Polo
L.M.
,
Zeng
Z.
,
Rulten
S.L.
,
Hoch
N.C.
,
Paomephan
P.
et al.
(
2016
)
PARP3 is a sensor of nicked nucleosomes and monoribosylates histone H2B(Glu2)
.
Nat. Commun.
7
,
12404
[PubMed]
78.
Langelier
M.F.
,
Riccio
A.A.
and
Pascal
J.M.
(
2014
)
PARP-2 and PARP-3 are selectively activated by 5′ phosphorylated DNA breaks through an allosteric regulatory mechanism shared with PARP-1
.
Nucleic Acids Res.
42
,
7762
7775
[PubMed]
79.
Rulten
S.L.
,
Fisher
A.E.
,
Robert
I.
,
Zuma
M.C.
,
Rouleau
M.
,
Ju
L.
et al.
(
2011
)
PARP-3 and APLF function together to accelerate nonhomologous end-joining
.
Mol. Cell
41
,
33
45
[PubMed]
80.
Ray Chaudhuri
A.
and
Nussenzweig
A.
(
2017
)
The multifaceted roles of PARP1 in DNA repair and chromatin remodelling
.
Nat. Rev. Mol. Cell Biol.
18
,
610
621
[PubMed]
81.
Polo
L.M.
,
Xu
Y.
,
Hornyak
P.
,
Garces
F.
,
Zeng
Z.
,
Hailstone
R.
et al.
(
2019
)
Efficient Single-Strand Break Repair Requires Binding to Both Poly(ADP-Ribose) and DNA by the Central BRCT Domain of XRCC1
.
Cell Rep.
26
,
573e5
581e5
82.
Bryant
H.E.
,
Schultz
N.
,
Thomas
H.D.
,
Parker
K.M.
,
Flower
D.
,
Lopez
E.
et al.
(
2005
)
Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase
.
Nature
434
,
913
917
[PubMed]
83.
Farmer
H.
,
McCabe
N.
,
Lord
C.J.
,
Tutt
A.N.
,
Johnson
D.A.
,
Richardson
T.B.
et al.
(
2005
)
Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy
.
Nature
434
,
917
921
[PubMed]
84.
Zimmermann
M.
,
Murina
O.
,
Reijns
M.A.M.
,
Agathanggelou
A.
,
Challis
R.
,
Tarnauskaite
Z.
et al.
(
2018
)
CRISPR screens identify genomic ribonucleotides as a source of PARP-trapping lesions
.
Nature
559
,
285
289
[PubMed]
85.
Nickson
C.M.
,
Moori
P.
,
Carter
R.J.
,
Rubbi
C.P.
and
Parsons
J.L.
(
2017
)
Misregulation of DNA damage repair pathways in HPV-positive head and neck squamous cell carcinoma contributes to cellular radiosensitivity
.
Oncotarget
8
,
29963
29975
[PubMed]
86.
Glorieux
M.
,
Dok
R.
and
Nuyts
S.
(
2017
)
Novel DNA targeted therapies for head and neck cancers: clinical potential and biomarkers
.
Oncotarget
8
,
81662
81678
[PubMed]
87.
Dungey
F.A.
,
Loser
D.A.
and
Chalmers
A.J.
(
2008
)
Replication-dependent radiosensitization of human glioma cells by inhibition of poly(ADP-Ribose) polymerase: mechanisms and therapeutic potential
.
Int. J. Radiat. Oncol. Biol. Phys.
72
,
1188
1197
[PubMed]
88.
Hanzlikova
H.
,
Kalasova
I.
,
Demin
A.A.
,
Pennicott
L.E.
,
Cihlarova
Z.
and
Caldecott
K.W.
(
2018
)
The Importance of Poly(ADP-Ribose) Polymerase as a Sensor of Unligated Okazaki Fragments during DNA Replication
.
Mol. Cell
71
,
319e3
331e3
89.
Higuchi
F.
,
Nagashima
H.
,
Ning
J.
,
Koerner
M.V.A.
,
Wakimoto
H.
and
Cahill
D.P.
(
2020
)
Restoration of Temozolomide Sensitivity by PARP Inhibitors in Mismatch Repair Deficient Glioblastoma is Independent of Base Excision Repair
.
Clin. Cancer Res.
26
,
1690
1699
[PubMed]
90.
Zandarashvili
L.
,
Langelier
M.F.
,
Velagapudi
U.K.
,
Hancock
M.A.
,
Steffen
J.D.
,
Billur
R.
et al.
(
2020
)
Structural basis for allosteric PARP-1 retention on DNA breaks
.
Science
368
,
eaxax6367
[PubMed]
91.
Murai
J.
,
Huang
S.Y.
,
Das
B.B.
,
Renaud
A.
,
Zhang
Y.
,
Doroshow
J.H.
et al.
(
2012
)
Trapping of PARP1 and PARP2 by Clinical PARP Inhibitors
.
Cancer Res.
72
,
5588
5599
[PubMed]
92.
Murai
J.
,
Huang
S.Y.
,
Renaud
A.
,
Zhang
Y.
,
Ji
J.
,
Takeda
S.
et al.
(
2014
)
Stereospecific PARP trapping by BMN 673 and comparison with olaparib and rucaparib
.
Mol. Cancer Ther.
13
,
433
443
[PubMed]
93.
Wang
Y.Q.
,
Wang
P.Y.
,
Wang
Y.T.
,
Yang
G.F.
,
Zhang
A.
and
Miao
Z.H.
(
2016
)
An Update on Poly(ADP-ribose)polymerase-1 (PARP-1) Inhibitors: Opportunities and Challenges in Cancer Therapy
.
J. Med. Chem.
59
,
9575
9598
[PubMed]
94.
Pettitt
S.J.
,
Krastev
D.B.
,
Brandsma
I.
,
Drean
A.
,
Song
F.
,
Aleksandrov
R.
et al.
(
2018
)
Genome-wide and high-density CRISPR-Cas9 screens identify point mutations in PARP1 causing PARP inhibitor resistance
.
Nat. Commun.
9
,
1849
[PubMed]
95.
Gogola
E.
,
Duarte
A.A.
,
de Ruiter
J.R.
,
Wiegant
W.W.
,
Schmid
J.A.
,
de Bruijn
R.
et al.
(
2018
)
Selective Loss of PARG Restores PARylation and Counteracts PARP Inhibitor-Mediated Synthetic Lethality
.
Cancer Cell
33
,
1078e12
1093e12
96.
Fathers
C.
,
Drayton
R.M.
,
Solovieva
S.
and
Bryant
H.E.
(
2012
)
Inhibition of poly(ADP-ribose) glycohydrolase (PARG) specifically kills BRCA2-deficient tumor cells
.
Cell Cycle
11
,
990
997
[PubMed]
97.
Gravells
P.
,
Grant
E.
,
Smith
K.M.
,
James
D.I.
and
Bryant
H.E.
(
2017
)
Specific killing of DNA damage-response deficient cells with inhibitors of poly(ADP-ribose) glycohydrolase
.
DNA Repair (Amst.)
52
,
81
91
[PubMed]
98.
Gravells
P.
,
Neale
J.
,
Grant
E.
,
Nathubhai
A.
,
Smith
K.M.
,
James
D.I.
et al.
(
2018
)
Radiosensitization with an inhibitor of poly(ADP-ribose) glycohydrolase: A comparison with the PARP1/2/3 inhibitor olaparib
.
DNA Repair (Amst.)
61
,
25
36
[PubMed]
99.
Le May
N.
,
Iltis
I.
,
Ame
J.C.
,
Zhovmer
A.
,
Biard
D.
,
Egly
J.M.
et al.
(
2012
)
Poly (ADP-ribose) glycohydrolase regulates retinoic acid receptor-mediated gene expression
.
Mol. Cell
48
,
785
798
[PubMed]
100.
Zhang
M.
,
Lai
Y.
,
Vasquez
J.L.
,
James
D.I.
,
Smith
K.M.
,
Waddell
I.D.
et al.
(
2020
)
Androgen Receptor and Poly(ADP-ribose) Glycohydrolase Inhibition Increases Efficiency of Androgen Ablation in Prostate Cancer Cells
.
Sci. Rep.
10
,
3836
[PubMed]
101.
Houl
J.H.
,
Ye
Z.
,
Brosey
C.A.
,
Balapiti-Modarage
L.P.F.
,
Namjoshi
S.
,
Bacolla
A.
et al.
(
2019
)
Selective small molecule PARG inhibitor causes replication fork stalling and cancer cell death
.
Nat. Commun.
10
,
5654
[PubMed]
102.
Chen
S.H.
and
Yu
X.
(
2019
)
Targeting dePARylation selectively suppresses DNA repair-defective and PARP inhibitor-resistant malignancies
.
Sci. Adv.
5
,
eaav4340
[PubMed]
103.
Huehls
A.M.
,
Huntoon
C.J.
,
Joshi
P.M.
,
Baehr
C.A.
,
Wagner
J.M.
,
Wang
X.
et al.
(
2016
)
Genomically Incorporated 5-Fluorouracil that Escapes UNG-Initiated Base Excision Repair Blocks DNA Replication and Activates Homologous Recombination
.
Mol. Pharmacol.
89
,
53
62
[PubMed]
104.
Fu
D.
,
Calvo
J.A.
and
Samson
L.D.
(
2012
)
Balancing repair and tolerance of DNA damage caused by alkylating agents
.
Nat. Rev. Cancer
12
,
104
120
[PubMed]
105.
Prorok
P.
,
Alili
D.
,
Saint-Pierre
C.
,
Gasparutto
D.
,
Zharkov
D.O.
,
Ishchenko
A.A.
et al.
(
2013
)
Uracil in duplex DNA is a substrate for the nucleotide incision repair pathway in human cells
.
Proc. Natl. Acad. Sci. U.S.A.
110
,
E3695
E3703
[PubMed]
106.
Kitsera
N.
,
Rodriguez-Alvarez
M.
,
Emmert
S.
,
Carell
T.
and
Khobta
A.
(
2019
)
Nucleotide excision repair of abasic DNA lesions
.
Nucleic Acids Res.
47
,
8537
8547
[PubMed]
107.
Permata
T.B.M.
,
Hagiwara
Y.
,
Sato
H.
,
Yasuhara
T.
,
Oike
T.
,
Gondhowiardjo
S.
et al.
(
2019
)
Base excision repair regulates PD-L1 expression in cancer cells
.
Oncogene
38
,
4452
4466
[PubMed]
108.
Seyedin
S.N.
,
Hasibuzzaman
M.M.
,
Pham
V.
,
Petronek
M.S.
,
Callaghan
C.
,
Kalen
A.L.
et al.
(
2020
)
Combination Therapy with Radiation and PARP Inhibition Enhances Responsiveness to Anti-PD-1 Therapy in Colorectal Tumor Models
.
Int. J. Radiat. Oncol. Biol. Phys.
[PubMed]
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY). Open access for this article was enabled by the participation of University of Liverpool in an all-inclusive Read & Publish pilot with Portland Press and the Biochemical Society under a transformative agreement with JISC.