DNA damage signaling response and repair (DDR) is a critical defense mechanism against genomic instability. Impaired DNA repair capacity is an important risk factor for cancer development. On the other hand, up-regulation of DDR mechanisms is a feature of cancer chemotherapy and radiotherapy resistance. Advances in our understanding of DDR and its complex role in cancer has led to several translational DNA repair-targeted investigations culminating in clinically viable precision oncology strategy using poly(ADP-ribose) polymerase (PARP) inhibitors in breast, ovarian, pancreatic, and prostate cancers. While PARP directed synthetic lethality has improved outcomes for many patients, the lack of sustained clinical response and the development of resistance pose significant clinical challenges. Therefore, the search for additional DDR-directed drug targets and novel synthetic lethality approaches is highly desirable and is an area of intense preclinical and clinical investigation. Here, we provide an overview of the mammalian DNA repair pathways and then focus on current state of PARP inhibitors (PARPi) and other emerging DNA repair inhibitors for synthetic lethality in cancer.

DNA damage

DNA repair pathways are essential for maintenance of genomic integrity, loss of which can promote carcinogenesis and influence response to cancer treatments. DNA damage occurs constantly due to endogenous and exogenous causes. Endogenous causes include reactive oxygen species (ROS), spontaneous base modifications, and errors during DNA replication. Exogenous causes include ultraviolet (UV) light, ionizing radiation, and chemicals, including chemotherapeutic agents. Therefore, in order to maintain genomic integrity, both prokaryotes and eukaryotes have evolved highly conserved DNA repair mechanisms to identify and correct DNA damage [1]. Following detection of DNA damage, cells may initiate different pathways, dependent on the type of damage, resulting in either: tolerance of the damage, transcriptional activation, induction of apoptosis (for highly damaged cells), or cell cycle arrest with subsequent repair of the DNA lesion [2,3] (Figure 1).

Cellular response to DNA damage

Figure 1
Cellular response to DNA damage
Figure 1
Cellular response to DNA damage
Close modal

DNA repair pathways

DNA repair in mammalian cells occurs through six key pathways, dependent on the type of DNA lesion, which are briefly outlined here and are more comprehensively reviewed elsewhere [3]. Figure 2 highlights the potential repair pathways for different types of DNA damage. It should be noted that significant cross-over exists between the effector proteins in each pathway.

An overview of mammalian DNA repair pathways

Figure 2
An overview of mammalian DNA repair pathways
Figure 2
An overview of mammalian DNA repair pathways
Close modal

Direct reversal

A small subset of DNA lesions, namely UV and alkylation-induced damage, can be directly reversed in situ in a relatively ‘error-free’ manner, without cleavage of the DNA phosphodiester backbone. UV radiation results in DNA photolesions including cyclobutane pyrimidine dimers (CPDs) and pyrimidine (6–4) pyrimidine photoproducts that are repaired by specific photolyases [4]; this direct repair occurs primarily in prokaryotes, whereas human cells repair these lesions by nucleotide excision repair (NER) [5]. Various alkylating lesions can occur following treatment with alkylating agents, a commonly used class of chemotherapy drug [6]. While some alkylating lesions may be repaired by base excision repair (BER), direct reversal can be performed by the sacrificial enzymes O6-alkylguanine-DNA alkyltransferase (AGT) or methylguanine methyltransferase (MGMT) [7]. Levels of MGMT in tumors, therefore, partially determines the response to alkylating chemotherapy agents [8,9]. An alternative pathway is the oxidative reversal of alkylation damage by AlkB dioxygenases (ABH2 and ABH3) [7,10].

BER

The BER pathway has evolved to detect and repair DNA damage from ROS, spontaneous deamination, and exogenous causes including alkylating agents and ionizing radiation [11]. BER can be further classified into short-patch (single nucleotide) and long-patch (multiple nucleotide) repair but the overall pathway is similar for both [12]. The BER pathway has five key steps: (i) removal of the base at the site of damage by a specific DNA glycosylase creating an apurinic/apyrimidinic (AP) site; (ii) incision of the phosphodiester backbone at the AP site by AP-endonuclease 1 (APE1); (iii) removal of the remaining deoxyribose-phosphate fragment by either DNA polymerase-β (polβ) (short-patch) or flap-endonuclease 1 (FEN1) (long-patch); (iv) insertion of the correct base at the AP site by DNA polymerases, predominantly by polβ (facilitated by X-ray cross-complementing group 1 protein (XRCC1)), but also polδ and ε (mediated by proliferating cell nuclear antigen (PCNA)) and (v) resealing the incised strand by DNA ligases [12,13]. Single-strand DNA breaks (SSBs), such as those made by APE1, cause the activation of poly(ADP-ribose) polymerase 1 (PARP1), which accelerates the repair pathway by recruiting essential BER enzymes such as DNA polymerases, ligases and XRCC1 [14,15]. PARP2 also plays a role in BER; however, neither PARP enzyme is essential for successful BER as it can proceed in their absence [16].

In contrast, PARP enzymes play a vital role in single-strand break repair (SSBR), an important subpathway of BER. SSBs can result from ROS, base deamination, intermediates from BER, defective activity of DNA topoisomerase 1 or replication-associated damage [3,17]. In summary, SSBR has four key steps: (i) SSB detection by PARP1 that adds PAR to itself (autoPARylation) and other repair proteins while also recruiting XRCC1 before dissociating from DNA; (ii) processing of the damaged 3′ or 5′ termini of the damaged strand by APE1, polβ, and polynucleotide kinase/phosphatase (PNKP) in conjunction with XRCC1, or in some cases by FEN1 in conjunction with PCNA; (iii) DNA gap filling with nucleotide insertion, and (iv) DNA ligation.

NER

The NER pathway is responsible for recognizing and repairing DNA lesions that cause a significant distortion of its helical structure [18]. Two subpathways of NER exist: transcription-coupled NER (TC-NER) for actively transcribed DNA and global-genome NER (GG-NER) for nonactively transcribed DNA. In summary, both NER pathways have the following steps: (i) recognition of DNA damage through sensor proteins [19,20]; (ii) recruitment of transcription factor IIH that contains helicases XPB and XPD to unwind DNA surrounding the lesion; (iii) incisions on either side of the lesion are made by the endonucleases XPG and ERCC1-XPF complex to produce an oligonucleotide product of 25–30 nucleotides in length; and (iv) polε in combination with PCNA and ligase I (in replicating cells) and pol δ and κ with PCNA and ligase IIIα/XRCC1 (in quiescent cells) act to fill and seal the incised gap [2,20]. Germline mutations in components of the NER pathway are a known cause of xeroderma pigmentosum in which patients have a greater than 1000-fold increased risk of nonmelanoma skin cancer [21].

Mismatch repair

The mismatch repair (MMR) pathway is responsible for the recognition and repair of base–base mismatches and insertion/deletion loops (IDLs) that form during DNA replication. Errors that evade the intrinsic proofreading activity of DNA polymerases must be corrected by the MMR pathway [22]. Additionally, MMR is responsible for correction of IDLs within microsatellite DNA and consequently, defective MMR results in ‘microsatellite instability,’ which can in-turn lead to genomic instability. In eukaryotes, MMR is commenced by the MSH2-MSH6 (small mismatches) or MSH2-MSH3 (large mismatches or IDLs) complexes. Both complexes then share a common pathway by first recruiting the MLH1-PMS2 complex which, in conjunction with PCNA, clamps to the DNA lesion [23]. These complexes now work in conjunction with exonuclease-1 (EXO1), polδ and DNA ligase I to excise and reform DNA using the other strand as a template [3]. There is evidence that FEN1 may act as an exonuclease in the MMR pathway, highlighting the cross-over of proteins between different repair pathways [24]. Mutations in the MMR pathway, most notably in MSH2 and MLH1, predispose to cancer and are an identified cause of hereditary nonpolyposis colorectal cancer (HNPCC) [25].

Nonhomologous end-joining

Double-strand breaks (DSBs) may occur as a result of ionizing radiation, ROS, replication errors or in physiological circumstances, such as V(D)J recombination for developing adaptive cell immunity, and certain chemotherapeutics. DSBs are either repaired by the more error-prone (and hence mutagenic) nonhomologous end-joining (NHEJ) or by the more accurate homologous recombination (HR). In brief, NHEJ occurs as follows: (i) recognition of a DSB by the Ku complex; (ii) Ku recruits DNA-dependent protein kinase catalytic subunit (DNA-PKcs) to form a DNA-PK complex; (iii) broken end processing (if necessary) is conducted by endonucleases; (iv) nucleotide insertion is performed by DNA polymerases λ and μ, which bind to Ku via their N-terminal BRCA1 C terminus (BRCT) domains; and (v) ligation occurs via the DNA ligase IV and XRCC4 complex [26,27]. Polymerase activity in NHEJ can be either template-dependent or independent, the latter being a key source of error in the pathway, and more commonly performed by polμ. Although often labelled as ‘error-prone’, DSBs resulting in blunt ends can be repaired with high precision by NHEJ, although such damage does not typically occur with ionizing radiation. Defects in the NHEJ pathway therefore lead to greater sensitivity to ionizing radiation, due to an inability to repair DSBs [28].

Microhomology-mediated end joining (MMEJ) and single-strand annealing (SSA) are mutagenic NHEJ-related pathways that have been reviewed in more detail elsewhere [3,29,30] and are beyond the scope of this review. It is relevant to note that MMEJ may be initiated by PARP1 [29] and is facilitated by polθ [31].

HR

The HR pathway utilizes a homologous template DNA strand to ensure highly accurate repair of DSBs and DNA interstrand cross-links (ICLs). The choice of which DSB repair pathway to follow is dependent on stage of the cell cycle (HR is up-regulated during S and G2 due to availability of a template strand) [32] and type of DSB sustained, with more complex breaks and those occurring during replication preferentially repaired by HR [33]. In summary, HR has the following steps: (i) damage recognition and resection of damaged ends by the MRN complex (Mre11-Rad50-Nbs1) producing single-strand DNA (ssDNA); (ii) coating of ssDNA by Replication Protein A (RPA); (iii) BRCA2-mediated replacement of RPA with RAD51; (iv) RAD51-bound ssDNA searches for and invades the homologous sequence on the sister chromatid; (v) repair synthesis occurs using the template strand by polη; (vi) dissociation of the repaired strand from the template strand followed by (vii) end ligation. These final resolution stages can either occur through synthesis-dependent strand annealing (SDSA) or through formation of Holliday junctions, which may generate cross-over products [33]. Both pathways are reviewed in more detail elsewhere [3,34]. Notably, germline mutations in BRCA2 result in greater susceptibility to breast, ovarian, and other cancers, evidencing the importance of HR in the maintenance of genomic integrity [35].

A subset of HR involves the repair of ICLs; these are recognized and corrected by a range of effector proteins including the Fanconi Anaemia (FA) complex, BRCA1, polν, and other proteins involved in HR including RAD51. While cross-link repair is reviewed in [36], it should be noted that up-regulation of ICL repair proteins may be responsible for resistance to platinum agents, whose primary mechanism of action is DNA damage through creation of ICLs. Consequently, the ICL repair pathway may offer a novel therapeutic target, with the aim of restoring platinum sensitivity.

DNA repair and cancer

If DNA lesions caused by the aforementioned agents remain unrepaired, mutations may arise which, in turn, can promote neoplastic transformation and subsequent carcinogenesis. As noted above, germline mutations in DNA repair proteins are recognized causes of hereditary cancer syndromes. The ‘mutator phenotype’ suggests that an impairment of one or more DNA repair pathways significantly promotes mutagenesis. Based on selection pressures, mutations in tumor-suppressor genes and oncogenes are more favorable to cell survival and growth [37]. Even following carcinogenesis, alterations in certain DNA repair pathway proteins have been shown to correlate with more aggressive tumors and consequently, worse prognosis [38,39].

The underlying mechanism of action of many chemotherapeutic agents and therapeutic ionizing radiation is primarily through initiation of DNA damage, with the aim of inducing cell cycle arrest or apoptosis of cells within the tumor. It therefore follows that intact or even up-regulated DNA repair pathways may contribute to treatment resistance [39]. This complex relationship between DNA repair, carcinogenesis, and therapeutic response is outlined in Figure 3.

DNA repair and cancer, a complex interaction

Figure 3
DNA repair and cancer, a complex interaction
Figure 3
DNA repair and cancer, a complex interaction
Close modal

Based on this rationale, inhibition of DNA repair pathways should potentiate the cytotoxic effects of chemotherapy and radiotherapy by acting as ‘sensitizing agents’ and overcoming resistance. Theoretically, such combinations should improve fractional cell kill and, by extension, clinical effectiveness. However, impairing DNA repair in other rapidly proliferating tissues, such as bone marrow and gastrointestinal mucosa, will result in increased toxicity to these cells and potential life-threatening complications, limiting the utility of such combinatorial therapeutic approaches [40,41]. This is evidenced clinically in trials of MGMT inhibitors and more recently, PARP inhibitors (PARPi) and other DNA repair inhibitors, in combinations with chemotherapy [41–43].

One means of overcoming this challenge is through more localized targeting of the cytotoxic component of combination therapies, which can be achieved with radiotherapy. While preclinical and clinical effectiveness of DNA repair inhibitors and radiotherapy in combination has been demonstrated, concerns remain regarding potential toxicity to normal tissues [44,45].

As a result, there is now a greater emphasis on developing personalized, precision treatments, which can selectively damage tumor cells and minimize toxicity to normal tissues. One means of achieving this is through targeting tumor-specific alterations in DNA repair pathways, thereby exploiting the concept of ‘synthetic lethality.’ In the present review, we will discuss current and evolving synthetic lethality targets in cancer therapy.

Synthetic lethality

Synthetic lethality refers to the situation in which a loss of function of either one of two genes does not result in cell death (and may even confer a survival advantage), whereas loss of both genes results in cell death [46] (Figure 4A–D). DNA repair inhibitors can theoretically capitalize on this principle. Tumors often harbor mutations in one or more DNA repair pathways, leading to a reliance on alternative, functioning pathways. Therefore, inhibition of an important alternative pathway can lead to a nonviable accumulation of unrepaired DNA damage (from constant endogenous damage, chemotherapy, or radiotherapy), and subsequent apoptosis. Normal cells possess an intact pathway to repair such damage, leading to selective killing of cancer cells.

Synthetic lethality

Figure 4
Synthetic lethality
Figure 4
Synthetic lethality
Close modal

PARP and BRCA

The discovery that PARP1 inhibition can lead to selective killing of BRCA-mutant cells has formed the cornerstone of evidence for synthetic lethality strategies. As outlined above, PARP plays an essential role in SSBR. Some PARPi, such as olaparib, rucaparib, niraparib, and talazoparib (chemical structures shown in Figure 4E), act by ‘trapping’ PARP1 at its binding site on DNA, thus inhibiting autoPARylation and PARP1 dissociation. The trapped PARP1 protein is cytotoxic (rather than the unrepaired SSB), as it causes collapse of the replication fork leading to a DSB [47]. In ‘normal’ cells, these breaks are repaired by the HR pathway, while in BRCA-deficient cells they remain unrepaired and accumulate, eventually resulting in apoptosis. However, these PARPi vary in their ‘trapping’ efficacy while veliparib, a different PARPi, acts primarily through inhibition of autoPARylation rather than ‘trapping’ [46]. This therefore highlights the heterogeneity of currently available PARPi and may partially explain their differing toxicity profiles [48].

Following promising preclinical evidence supporting this synthetically lethal interaction [49,50], PARPi have been investigated in a number of clinical trials for treating BRCA-mutated cancers. It is estimated that 5–10% of breast and ovarian cancers carry either germline or somatic BRCA mutations [51]. A summary of phase III trials of PARPi for BRCA-mutated cancers is presented in Table 1 [52–61]. These trials in BRCA-mutated breast cancers have demonstrated that olaparib [59], talazoparib [58], and veliparib [54], either as monotherapy or in combination with chemotherapy, significantly improved median progression-free survival (mPFS) as compared with standard chemotherapy regimens alone. These findings are further supported by preceding phase II trials that have also demonstrated acceptable toxicity profiles [62,63]. Consequently, olaparib and talazoparib are licensed in the U.S.A. for the treatment of BRCA-mutated, HER2-negative breast cancers, although approval is still awaited in the U.K. PARPi have been most extensively investigated in the treatment of BRCA-mutated ovarian cancers as both combination therapies and as monotherapy. All published phase III trials have demonstrated significant benefits in mPFS in favor of PARPi, with acceptable toxicity [52,53,55,60]. Based on this evidence, olaparib and niraparib have been approved by the FDA, and NICE in the U.K., for the treatment of BRCA-mutated, advanced ovarian cancer. In addition, a phase III trial of olaparib for the treatment of BRCA-mutated advanced pancreatic cancer demonstrated significantly greater mPFS for PARPi as compared with placebo (7.4 vs 3.8 months; P=0.004), resulting in its FDA approval for this indication [57].

Table 1
Published phase III trials of PARPi for BRCA-mutated cancers
Study titleYear publishedCancerAuthorPARPiComparator armSample sizemPFSOther relevant results
Rucaparib versus standard-of-care chemotherapy in patients with relapsed ovarian cancer and a deleterious BRCA1 or BRCA2 mutation (ARIEL4): an international, open-label, randomised, phase 3 trial [502022 Ovarian Kristeleit et al. Rucaparib (600 mg BD) Chemotherapy 349 Rucaparib 7.4 months vs chemotherapy 5.7 months (P=0.001) NA 
Maintenance olaparib for patients with newly diagnosed advanced ovarian cancer and a BRCA mutation (SOLO1/GOG 3004): 5-year follow-up of a randomised, double-blind, placebo-controlled, phase 3 trial [512021 Ovarian Banerjee et al. Olaparib (300 mg BD) Placebo 391 Olaparib 56.0 months vs placebo 13.8 months NA 
Veliparib with carboplatin and paclitaxel in BRCA-mutated advanced breast cancer (BROCADE3): a randomised, double-blind, placebo-controlled, phase 3 trial [522020 Breast Diéras et al. Veliparib (120 mg BD) plus chemotherapy Placebo plus chemotherapy 513 Veliparib plus chemotherapy 14.5 months vs placebo plus chemotherapy 12.6 months (P=0.0016) NA 
Olaparib versus nonplatinum chemotherapy in patients with platinum-sensitive relapsed ovarian cancer and a germline BRCA1/2 mutation (SOLO3): a randomized phase III trial [532020 Ovarian Penson et al. Olaparib (300 mg BD) Physician’s choice single-agent nonplatinum chemotherapy 266 Olaparib 13.4 months vs physician choice chemotherapy 9.2 months (P=0.013) ORR significantly higher for olaparib (72.2% vs 51.4%) 
Veliparib with first-line chemotherapy and as maintenance therapy in ovarian cancer [542019 Ovarian Coleman et al. Veliparib (150 mg OD) plus chemotherapy followed by veliparib maintenance Chemotherapy plus placebo, chemotherapy plus veliparib followed by placebo maintenance 1140 Veliparib maintenance 23.5 months vs chemotherapy plus placebo 17.3 months (P<0.001) In BRCA-positive group, mPFS: 34.7 months for veliparib maintenance vs 22.0 months for placebo. In HR defect group, mPFS: 31.9 months for veliparib maintenance vs 20.5 months for placebo (P<0.001 for both) 
Maintenance olaparib for germline BRCA-mutated metastatic pancreatic cancer [552019 Pancreatic Golan et al. Olaparib (300 mg BD) Placebo 154 Olaparib 7.4 months vs placebo 3.8 months (P=0.004) Median overall survival: olaparib 18.9 months vs placebo 18.1 months (P=0.68) 
Talazoparib in patients with advanced breast cancer and a germline BRCA mutation [562018 Breast Litton et al. Talozaparib (1 mg OD) Chemotherapy 431 Talozaparib 8.6 months vs chemotherapy 5.6 months (P<0.001) ORR: talozaparib 62.6% vs chemotherapy 27.2% 
Olaparib for metastatic breast cancer in patients with a germline BRCA mutation [572017 Breast Robson et al. Olaparib (300 mg BD) Chemotherapy 302 Olaparib 7.0 months vs chemotherapy 4.2 months (P<0.001) ORR: olaparib 59.9% vs chemotherapy 28.8% 
Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): a double-blind, randomised, placebo-controlled, phase 3 trial [582017 Ovarian Pujade-Lauraine et al. Olaparib (300 mg BD) Placebo 295 Olaparib plus bevacizumab 19.1 months vs placebo 5.5 months (P<0.0001) NA 
Niraparib maintenance therapy in platinum-sensitive, recurrent ovarian cancer [592016 Ovarian Mirza et al. Niraparib (300 mg OD) Placebo 553 In gBRCA cohort: 21.0 months for niraparib vs 5.5 months placebo. In non-BRCA group: 9.3 months for niraparib vs 3.9 months placebo (P<0.001 for both) For non-BRCA group with HR deficiency, mPFS: 12.9 months with niraparib vs 3.8 months with placebo (P<0.001) 
Study titleYear publishedCancerAuthorPARPiComparator armSample sizemPFSOther relevant results
Rucaparib versus standard-of-care chemotherapy in patients with relapsed ovarian cancer and a deleterious BRCA1 or BRCA2 mutation (ARIEL4): an international, open-label, randomised, phase 3 trial [502022 Ovarian Kristeleit et al. Rucaparib (600 mg BD) Chemotherapy 349 Rucaparib 7.4 months vs chemotherapy 5.7 months (P=0.001) NA 
Maintenance olaparib for patients with newly diagnosed advanced ovarian cancer and a BRCA mutation (SOLO1/GOG 3004): 5-year follow-up of a randomised, double-blind, placebo-controlled, phase 3 trial [512021 Ovarian Banerjee et al. Olaparib (300 mg BD) Placebo 391 Olaparib 56.0 months vs placebo 13.8 months NA 
Veliparib with carboplatin and paclitaxel in BRCA-mutated advanced breast cancer (BROCADE3): a randomised, double-blind, placebo-controlled, phase 3 trial [522020 Breast Diéras et al. Veliparib (120 mg BD) plus chemotherapy Placebo plus chemotherapy 513 Veliparib plus chemotherapy 14.5 months vs placebo plus chemotherapy 12.6 months (P=0.0016) NA 
Olaparib versus nonplatinum chemotherapy in patients with platinum-sensitive relapsed ovarian cancer and a germline BRCA1/2 mutation (SOLO3): a randomized phase III trial [532020 Ovarian Penson et al. Olaparib (300 mg BD) Physician’s choice single-agent nonplatinum chemotherapy 266 Olaparib 13.4 months vs physician choice chemotherapy 9.2 months (P=0.013) ORR significantly higher for olaparib (72.2% vs 51.4%) 
Veliparib with first-line chemotherapy and as maintenance therapy in ovarian cancer [542019 Ovarian Coleman et al. Veliparib (150 mg OD) plus chemotherapy followed by veliparib maintenance Chemotherapy plus placebo, chemotherapy plus veliparib followed by placebo maintenance 1140 Veliparib maintenance 23.5 months vs chemotherapy plus placebo 17.3 months (P<0.001) In BRCA-positive group, mPFS: 34.7 months for veliparib maintenance vs 22.0 months for placebo. In HR defect group, mPFS: 31.9 months for veliparib maintenance vs 20.5 months for placebo (P<0.001 for both) 
Maintenance olaparib for germline BRCA-mutated metastatic pancreatic cancer [552019 Pancreatic Golan et al. Olaparib (300 mg BD) Placebo 154 Olaparib 7.4 months vs placebo 3.8 months (P=0.004) Median overall survival: olaparib 18.9 months vs placebo 18.1 months (P=0.68) 
Talazoparib in patients with advanced breast cancer and a germline BRCA mutation [562018 Breast Litton et al. Talozaparib (1 mg OD) Chemotherapy 431 Talozaparib 8.6 months vs chemotherapy 5.6 months (P<0.001) ORR: talozaparib 62.6% vs chemotherapy 27.2% 
Olaparib for metastatic breast cancer in patients with a germline BRCA mutation [572017 Breast Robson et al. Olaparib (300 mg BD) Chemotherapy 302 Olaparib 7.0 months vs chemotherapy 4.2 months (P<0.001) ORR: olaparib 59.9% vs chemotherapy 28.8% 
Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): a double-blind, randomised, placebo-controlled, phase 3 trial [582017 Ovarian Pujade-Lauraine et al. Olaparib (300 mg BD) Placebo 295 Olaparib plus bevacizumab 19.1 months vs placebo 5.5 months (P<0.0001) NA 
Niraparib maintenance therapy in platinum-sensitive, recurrent ovarian cancer [592016 Ovarian Mirza et al. Niraparib (300 mg OD) Placebo 553 In gBRCA cohort: 21.0 months for niraparib vs 5.5 months placebo. In non-BRCA group: 9.3 months for niraparib vs 3.9 months placebo (P<0.001 for both) For non-BRCA group with HR deficiency, mPFS: 12.9 months with niraparib vs 3.8 months with placebo (P<0.001) 

Further to these promising findings, PARPi are also effective preclinically in tumors that do not possess a BRCA mutation but are deficient in another component of the HR pathway, thereby exploiting an alternative synthetically lethal pairing [64]. Such HR-deficient tumors are described as demonstrating ‘BRCAness’ with commonly affected genes including ATM, ATR, PALB2, CHEK1 (encoding CHK1), and CHEK2 [65]. Somatic mutations in these genes, in keeping with the mutator phenotype, are widespread across many cancer types. These cancers, as well as those harboring a BRCA mutation, typically possess a characteristic series of mutations due to an over-reliance on more error-prone repair pathways. Such mutational scars can be identified using mutational signature profiles to assist in distinguishing those most suitable for PARPi therapy. In clinical trials, PARPi have been shown to significantly improve mPFS in HR-deficient ovarian cancers [56,61] with promising results seen in HR-deficient advanced prostate [66,67] and urothelial [68] cancers. Beyond this, Coleman et al. demonstrated that addition of veliparib to chemotherapy in advanced ovarian cancer significantly improved mPFS as compared with chemotherapy and placebo (23.5 vs 17.3 months; P<0.001) even in patients without BRCA or known HR repair mutations [56].

However, developing resistance to PARPi is thought to occur in 40–70% of patients over the course of their treatment [69]. The most notable mechanism, and the only one as yet confirmed in vivo, is a frameshift mutation in BRCA2 that restores the open-reading frame of the gene, thereby restoring its function without a complete reversal back to wild-type [70,71]. This c.6174delT frameshift mutation will result in restored HR repair capabilities and prevention of synthetic lethality in the presence of PARPi [69]. Other mechanisms of resistance occur through an inhibition of NHEJ, thereby forcing cells to repair DSBs via the HR pathway. First, p53-binding protein 1 (53BP1) is a key repair protein that, in BRCA-mutant cells, typically inhibits HR and drives excessive NHEJ and consequent apoptosis. In normal situations, 53BP1 is removed by BRCA1 to allow for repair by the more accurate HR pathway. As a result, loss of 53BP1, as often occurs in BRCA-mutated or triple-negative breast cancers, results in partial restoration of the HR pathway and resistance to PARPi. Evidence suggests that 53BP1 levels may act as a prognostic biomarker by predicting response to both chemotherapy agents and PARPi [69]. Second, the REV7 protein acts in a similar manner, but downstream of 53BP1, by promoting DSB repair via the NHEJ pathway. Consequently, reduced levels of REV7 promote HR repair, hence conferring PARPi resistance and worse patient outcomes [69,72]. More recent evidence suggests that PARPi resistance may occur through restoring fork stabilization via loss of proteins such as PTIP, RADX, SMARCAL1, and FANCD2; this is reviewed more extensively in [69]. Another preclinical mechanism for PARPi resistance is increased expression of ABCB1 that encodes a drug transporter, resulting in increased efflux of PARPi [73]. Finally, mutations or reduced levels of PARP1 itself may lead to resistance to PARPi due to a loss of the cytotoxic ‘trapped’ PARP1 at sites of SSBs [69].

Overcoming PARPi resistance has led to a need to identify new synthetically lethal pairings and novel targets for DNA repair inhibition. Recent evidence suggests that potential targets include proteins involved in cell-cycle checkpoints and DNA repair pathways such as ATR, ATM, and WEE1.

ATR/CHK1

The ataxia telangiectasia and Rad3-related (ATR)-checkpoint kinase 1 (CHK1) pathway is a key component of the DNA damage response. Following end resection of DSBs (as mediated by ATM) or at stalled replication forks, ssDNA is coated by RPA as discussed above. This bound RPA subsequently recruits ATR and ATR interacting protein (ATRIP) that in turn activates CHK1. CHK1 then acts to inhibit CDK2 during the S-phase by causing degradation of CDC25A. The reduced activity of CDK2 results in activation of the intra-S and G2/M phase cell-cycle checkpoints, allowing the cell to initiate DSB repair. Furthermore, downstream components of the ATR-CHK1 pathway play a key role in suppressing the replication stress response, which is triggered by stalled replication forks [74].

The ATR/CHK1 pathway has been identified as a potential synthetic lethality target. Cells that lack a functional G1 checkpoint, as often occur in tumors with p53 or retinoblastoma mutations, may be particularly sensitive to ATR/CHK1 inhibition. In such cells, ATR/CHK1 inhibition will result in loss of the G1, intra-S and G2/M checkpoints with premature progression to mitosis leading to a ‘mitotic catastrophe’ and cell death [74]. As discussed above, 53BP1 mutations may lead to PARPi resistance and ATR inhibition may therefore offer a means of overcoming this. Similarly, chemotherapy agents trigger a replication stress response as a result of DNA damage; inhibition of ATR will prevent suppression of this response and in tumor cells that overexpress oncogenes, this can be synthetically lethal [74].

ATR inhibitors (ATRi) currently in use are predominantly small-molecule inhibitors that preclinically have been shown to sensitize cells to ionizing radiation and chemotherapy, as well as inducing synthetic lethality in p53- and ataxia telangiectasia mutated (ATM)-deficient cell lines. Furthermore, ATR/CHK1 inhibition sensitized cells to PARPi, thereby rationalizing combination strategies [75].

Following these promising in vitro results, four ATRi have been used in clinical trials, namely M6620 (berzosertib, IV), AZD6738 (PO), BAY1895344 (PO), and most recently RP-3500 (PO). It should be noted that ATRi, particularly in combination with other chemotherapy agents, carry a significant toxicity profile with over a third of patients across two trials experiencing grade 3 or 4 adverse events, predominantly cytopenias [76,77]. M6620 has shown some promise as a monotherapy with pre- and post-treatment tumor biopsies demonstrating a reduction in CHK1 phosphorylation (a biomarker of ATR activity) [78]. In phase I trials, ATRi have been shown to act synergistically with chemotherapy agents such as platinum agents [76–78], gemcitabine [79], and topotecan [80]. The phase I Patriot study investigating the combination of ATR inhibition with ionizing radiation is ongoing, although there is positive preclinical evidence [74]. To date, there are no published phase II clinical trials evaluating ATRi as they remain in the early stages of development; ongoing phase II trials of ATRi are summarized in Table 2.

Table 2
Ongoing phase II trials of ATRi
Study titleNCT trial numberYear commencedCancer(s)ATRiComparatorEstimated/ actual enrollmentSpecific biomarkersEstimated study completion date
ATARI: ATr inhibitor in combination with olaparib in gynaecological cancers with ARId1A loss NCT04065269 2019 Gynecological cancers (including ovarian and endometrial) AZD6738 plus olaparib AZD6738 alone 105 ARID1A March 2023 
Phase II trial of AZD6738 alone and in combination with olaparib in patients with selected solid tumor malignancies NCT03682289 2019 Advanced solid tumors (excluding ovarian cancer) AZD6738 plus olaparib AZD6738 alone 59 BAF250a and ATM December 2023 
Phase 1/2a study of the safety, pharmacokinetics, pharmacodynamics and preliminary clinical activity of RP-3500 alone or in combination with talazoparib or gemcitabine in advanced solid tumors with ATR inhibitor sensitizing mutations (TRESR study) NCT04497116 2020 Advanced solid tumors RP-3500 RP-3500 plus talazoparib plus gemcitabine 451 Not specified March 2024 
A multi-center phase II study testing the activity of olaparib and AZD6738 (ATR inhibitor) in metastatic castration-resistant prostate cancer NCT03787680 2019 Prostate AZD6738 plus olaparib Nil 49 NA November 2026 
Phase 1b/2 study of ATR inhibiTor RP-3500 and PARP inhibitor combinations in patients with molecularly selected cancers (ATTACC) NCT04972110 2021 Advanced solid tumors RP-3500 plus niraparib or olaparib Nil 108 Not specified November 2023 
Phase 2 study of M6620 (VX-970) in combination with gemcitabine versus gemcitabine alone in subjects with platinum-resistant recurrent ovarian or primary peritoneal fallopian tube cancer NCT02595892 2016 Ovarian M6620 (berzosertib) plus gemicitabine Gemcitabine alone 70 Nil March 2023 
A phase I/II trial of lurbinectedin with berzosertib, an ATR kinase inhibitor in small cell cancers and high grade neuroendocrine cancers NCT04802174 2021 Small-cell lung cancer and high-grade neuroendocrine tumors M6620 (berzosertib) plus lurbinectedin Nil 75 Nil December 2026 
A phase I/II trial of topotecan with VX-970 (M6620), an ATR kinase inhibitor in small-cell cancers NCT02487095 2015 Small-cell cancers (lung and extrapulmonary) M6620 (berzosertib) plus topotecan Nil 62 Nil October 2025 
A phase I/IIa, open-label, multi-center study to assess the safety, tolerability, pharmacokinetics and preliminary efficacy of the ATR kinase inhibitor ART0380 administered orally as monotherapy and in combination to patients with advanced or metastatic solid tumors NCT04657068 2021 Advanced solid tumors with ATM mutation or ovarian cancer ATR0380 plus gemcitabine or irinotecan ART0380 alone 232 ATM December 2023 
A phase 1/ 2 study of BAY 1895344 (Elimusertib, NSC#810486) in pediatric patients with relapsed or refractory solid tumors NCT05071209 2021 Relapsed solid tumors Elimusertib Nil 23 ATM, ATRX, BRCA1, BRCA2, CDK12, CHEK1, CHEK2, FANCA, MSH2, MRE11, PALB2, PARP1, POLD1, RAD51, or XRCC2 June 2024 
National lung matrix trial: multi-drug, genetic marker-directed, non-comparative, multi-centre, multi-arm phase II trial in non-small cell lung cancer NCT02664935 2015 Non-small-cell lung cancer AZD6738 plus durvalumab AZD4547, vistusertib, palbociclib, crizotinib, selumitinib plus docetaxel, AZD5363, osimertinib, durvalumab, sitravatinib 423 KRAS, SKT11/LKB1 October 2022 
Combination ATR and PARP inhibitor (CAPRI) trial with AZD 6738 and olaparib in recurrent ovarian cancer NCT03462342 2018 Recurrent ovarian cancer AZD6738 plus olaparib Nil 86 BRCA or HRD mutations December 2022 
Study titleNCT trial numberYear commencedCancer(s)ATRiComparatorEstimated/ actual enrollmentSpecific biomarkersEstimated study completion date
ATARI: ATr inhibitor in combination with olaparib in gynaecological cancers with ARId1A loss NCT04065269 2019 Gynecological cancers (including ovarian and endometrial) AZD6738 plus olaparib AZD6738 alone 105 ARID1A March 2023 
Phase II trial of AZD6738 alone and in combination with olaparib in patients with selected solid tumor malignancies NCT03682289 2019 Advanced solid tumors (excluding ovarian cancer) AZD6738 plus olaparib AZD6738 alone 59 BAF250a and ATM December 2023 
Phase 1/2a study of the safety, pharmacokinetics, pharmacodynamics and preliminary clinical activity of RP-3500 alone or in combination with talazoparib or gemcitabine in advanced solid tumors with ATR inhibitor sensitizing mutations (TRESR study) NCT04497116 2020 Advanced solid tumors RP-3500 RP-3500 plus talazoparib plus gemcitabine 451 Not specified March 2024 
A multi-center phase II study testing the activity of olaparib and AZD6738 (ATR inhibitor) in metastatic castration-resistant prostate cancer NCT03787680 2019 Prostate AZD6738 plus olaparib Nil 49 NA November 2026 
Phase 1b/2 study of ATR inhibiTor RP-3500 and PARP inhibitor combinations in patients with molecularly selected cancers (ATTACC) NCT04972110 2021 Advanced solid tumors RP-3500 plus niraparib or olaparib Nil 108 Not specified November 2023 
Phase 2 study of M6620 (VX-970) in combination with gemcitabine versus gemcitabine alone in subjects with platinum-resistant recurrent ovarian or primary peritoneal fallopian tube cancer NCT02595892 2016 Ovarian M6620 (berzosertib) plus gemicitabine Gemcitabine alone 70 Nil March 2023 
A phase I/II trial of lurbinectedin with berzosertib, an ATR kinase inhibitor in small cell cancers and high grade neuroendocrine cancers NCT04802174 2021 Small-cell lung cancer and high-grade neuroendocrine tumors M6620 (berzosertib) plus lurbinectedin Nil 75 Nil December 2026 
A phase I/II trial of topotecan with VX-970 (M6620), an ATR kinase inhibitor in small-cell cancers NCT02487095 2015 Small-cell cancers (lung and extrapulmonary) M6620 (berzosertib) plus topotecan Nil 62 Nil October 2025 
A phase I/IIa, open-label, multi-center study to assess the safety, tolerability, pharmacokinetics and preliminary efficacy of the ATR kinase inhibitor ART0380 administered orally as monotherapy and in combination to patients with advanced or metastatic solid tumors NCT04657068 2021 Advanced solid tumors with ATM mutation or ovarian cancer ATR0380 plus gemcitabine or irinotecan ART0380 alone 232 ATM December 2023 
A phase 1/ 2 study of BAY 1895344 (Elimusertib, NSC#810486) in pediatric patients with relapsed or refractory solid tumors NCT05071209 2021 Relapsed solid tumors Elimusertib Nil 23 ATM, ATRX, BRCA1, BRCA2, CDK12, CHEK1, CHEK2, FANCA, MSH2, MRE11, PALB2, PARP1, POLD1, RAD51, or XRCC2 June 2024 
National lung matrix trial: multi-drug, genetic marker-directed, non-comparative, multi-centre, multi-arm phase II trial in non-small cell lung cancer NCT02664935 2015 Non-small-cell lung cancer AZD6738 plus durvalumab AZD4547, vistusertib, palbociclib, crizotinib, selumitinib plus docetaxel, AZD5363, osimertinib, durvalumab, sitravatinib 423 KRAS, SKT11/LKB1 October 2022 
Combination ATR and PARP inhibitor (CAPRI) trial with AZD 6738 and olaparib in recurrent ovarian cancer NCT03462342 2018 Recurrent ovarian cancer AZD6738 plus olaparib Nil 86 BRCA or HRD mutations December 2022 

Preclinical evidence suggests that ATR inhibition may lead to down-regulation of programmed death-ligand 1 (PDL-1), thereby sensitizing tumors to immune-cell mediated killing. Consequently, a phase I study evaluated the safety and effectiveness of AZD6738 combined with the PDL-1 inhibitor durvulumab; the combination was well tolerated and promisingly, there was one potential complete response and one partial response in the cohort [76].

On the basis of in vitro evidence suggesting that ATRi can sensitize cancer cells to PARPi treatment [81], a number of trials testing this combination have commenced. Yap et al. assessed the combination of AZD6738 and olaparib in a phase I trial in which two patients with triple-negative breast cancer (TNBC) achieved partial responses.

Mechanisms of resistance to ATRi continue to be investigated; Lloyd et al. identified using CRISPR–Cas9 genome-wide screening that loss of cyclin C or CDK8 can lead to ATRi resistance through suppression of the replication stress response [82]. Further elucidation of these resistance mechanisms, particularly in in vivo models, is essential to identify patients mostly likely to respond and therefore develop more successful targeted, precision therapies.

Selective CKH1 inhibitors, such as MK-8776, have shown positive results in preclinical studies as monotherapy [83], with chemotherapy agents [84] and with ionizing radiation [85]. MK-8776 has been well tolerated in trials with gemcitabine and cytarabine for solid tumors (phase I) [86] and refractory acute leukemias (phase II) [87], respectively. There was promising antitumor activity in solid tumors [86], while for hematological malignancies, the addition of MK-8776 to cytarabine had no significant benefit [87]. Prexasertib, a second-generation CHK1 inhibitor with some anti-CHK2 activity, has been investigated in phase II trials for ovarian cancer [88] and TNBC [89] with encouraging results, although severe neutropenia was common in both. Further trials investigating CHK1 inhibitors across various tumor types are ongoing. Intriguingly, recent evidence suggests CHK1 inhibitors may induce BRCAness in cells, thereby sensitizing BRCA wild-type but p53-deficient cells to olaparib. This rationalizes combinations of CHK1 inhibitors and PARPi and hence warrants further investigation in clinical trials [90].

ATM/CHK2

ATM/CHK2 signaling also plays an essential role in the DNA damage signaling response and repair (DDR), in particular the recognition and repair of DSBs. As discussed above, DSBs are recognized by the MRN complex; this complex subsequently activates ATM-CHK2 kinase resulting in phosphorylation of p53 and cell-cycle arrest at the G1/S checkpoint. Furthermore, ATM mediates end processing of DSBs resulting in RPA coating of ssDNA and therefore activation of the ATR/CHK1 pathways as described above. ATM and both ATR and p53 demonstrate a synthetically lethal relationship as the loss of three key cell-cycle checkpoints results in mitotic catastrophe. Although ATM is frequently mutated across cancer types, functional ATM deficiency due to hypermethylation of its promoter region is more common [91]. Preclinical evidence suggests that ATM activation may contribute to chemotherapeutic resistance [92]. This therefore rationalizes the development of ATM inhibitors (ATMi; and downstream CHK2 inhibitors) both as a means of overcoming chemotherapy resistance, sensitizing cells to ionizing radiation and as a synthetic lethality strategy in p53-deficient tumors.

In the laboratory setting, a wide range of ATMi have been tested. These studies have demonstrated that ATMi can sensitize cells to chemotherapy and ionizing radiation, although they appear to lack utility as a monotherapy. Phosphate and tensin homolog (PTEN) plays an important role in the DDR. Multiple preclinical studies across tumor types have demonstrated that the ATMi KU-60019 in combination with cisplatin is synthetically lethal in PTEN-deficient cells [93,94].

While these studies demonstrate the potential benefits of ATM inhibition, the development of ATMi for clinical studies remains in its infancy. Three ATMi (AZD0156, KU-60019, and AZD1390) are being investigated in clinical trials with results expected in the coming years. AZD0156 is being investigated as monotherapy, in conjunction with other chemotherapy agents, and as a combination therapy with olaparib for the treatment of advanced solid tumors (NCT02588105). AZD1390 is being assessed in combination with ionizing radiation for the treatment of glioblastoma multiforme (NCT034236280). Finally, KU-60019 is being assessed in combination with silimitasertib (a casein kinase II inhibitor involved in the PI3K/AKT pathway) for the treatment of renal cell cancers (NCT03571438) [95]. Further preclinical research to identify potential mechanisms of resistance to ATMi will be a vital aspect of their ongoing development.

As a downstream target of ATM, CHK2 inhibitors can also induce mitotic catastrophe in a similar manner to CHK1 inhibition. In preclinical studies, two selective CHK2 inhibitors have been investigated: PV1019 (NIH) and CCT241533 (ICR). The former has been shown to act synergistically with chemotherapy and radiotherapy [96], while the latter can potentiate the activity of PARPi [97]. While these selective inhibitors have not, as yet, been investigated in clinical trials, the nonselective CHK1/2 inhibitor AZD7762 has undergone phase I trials with promising results. However, AZD7762 was associated with significant, dose-limiting cardiotoxicity [98] and two other phase I trials with the drug were suspended (NCT00937664 and NCT00473616). Additional research into the toxicity profile of CHK inhibitors is necessary in order to determine whether the observed cardiotoxicity is a wider problem with the whole drug class.

WEE1

WEE1 kinase plays an essential role at the G2/M cell-cycle checkpoint. The enzyme acts by phosphorylating, and thereby inhibiting, CDK1/cyclin complexes and hence preventing cell-cycle progression to mitosis [99]. In addition to controlling cell-cycle progression and maintaining genomic integrity, WEE1 also plays a role in epigenetic modulation through suppression of histone transcription in late S-phase [100]. WEE1 expression has been shown to be both up- and down-regulated across different cancer types, with both associated with poor prognosis [101]. In those tumors with high WEE1 expression, it is likely that they are dependent on an intact G2/M checkpoint for survival, possibly due to inactivation of the G1/S checkpoint following a loss-of-function p53 mutation. Therefore, in such tumors, inhibition of WEE1 kinase in combination with DNA-damaging agents may result in mitotic catastrophe through accumulation of mutations and premature mitosis. A number of potent small-molecule WEE1 kinase inhibitors (WEE1i) have been identified through drug screening and used in preclinical and clinical trials [101]. The most developed of these is AZD1775, also known as adavosertib.

In the preclinical setting, AZD1775 has been shown to act synergistically with a range of chemotherapy agents in p53-deficient tumors and with ionizing radiation or cisplatin in medulloblastoma cells, irrespective of p53 phenotype [101]. WEEi may also work in conjunction with other DNA repair inhibitors; for example, the addition of WEE1i to ATRi therapy resensitized ATR-resistant cells through forced premature entry into mitosis [102]. Evidence for WEE1i monotherapy is weak however, and clinical trials predominantly evaluate WEE1i in combination with other therapeutic agents [101].

A phase I trial with AZD1775 in combination with various chemotherapy agents for patients with advanced solid tumors demonstrated encouraging efficacy, with higher response rates observed in p53-mutated patients [103]. A further phase II trial evaluated AZD1775 with carboplatin in the treatment of advanced platinum-resistant ovarian cancer; the ORR was 31.9% with a mPFS of 5.5 months. However, treatment-related toxicities, including gastrointestinal symptoms and cytopenias, were common and 12.8% of patients discontinued AZD1775 [104]. A phase II trial comparing AZD1775 in combination with gemcitabine to placebo showed significantly improved overall survival from 7.2 to 11.5 months (P=0.022), although hematological toxicity remained an issue [105]. Given the preclinical evidence suggesting WEE1i may be of most benefit in p53-mutated cancers, AZD1775 was assessed in combination with carboplatin and paclitaxcel for the treatment of p53-mutated, platinum-sensitive ovarian cancer. While there was no significant difference in response rates, mPFS was significantly greater in the AZD1775 arm as compared with placebo (9.9 vs 8.0 months; P=0.030) [106]. There is further promising evidence for triple combinations of AZD1775 with chemoradiotherapy regimens for both pancreatic and head and neck cancers [107].

Following concerns regarding the toxicity profile of WEE1i and chemotherapy agents, it has been suggested that WEE1i may be better tolerated in combination with precision anticancer therapies such as DNA repair inhibitors or immunotherapy. A phase Ib trial assessed the combination of AZD1775 and olaparib in 119 patients; despite demonstrating good efficacy, hematological toxicity was again common [108]. A randomized phase II trial in 273 metastatic TNBC patients found no significant differences in response rates or PFS with the addition of AZD1775 to olaparib alone (NCT03330847). A phase I trial found that the combination of AZD1775 with the PDL-1-inhibitor durvalumab, had an acceptable safety profile and evidence of antitumor activity [109]. A summary of clinical trials with WEE1i can be found in Table 3 [101–115]. Given these encouraging results and improved safety profiles, further trials of these novel combination therapies are warranted.

Table 3
Completed trials of WEE1 inhibitors
Study titleNCT trial numberYear completedCancer(s)WEE1 inhibitorComparatorSample sizeRelevant results
A phase I trial of WEE1 inhibition with chemotherapy and radiotherapy as adjuvant treatment, and a window of opportunity trial with cisplatin in patients with head and neck cancer NCT03028766 2021 Head and neck cancer AZD1775 plus cisplatin plus radiotherapy AZD1775 plus cisplatin 58 Awaiting publication of results 
Phase Ib trial of dose-escalating AZD1775 in combination with concurrent radiation and cisplatin for intermediate and high risk head and neck squamous cell carcinoma (HNSCC) [108NCT02585973 2021 Head and neck cancer (SCC) AZD1775 plus cisplatin plus radiotherapy Nil 12 ORR: 100% at 3 months, mPFS and median overall survival were 90% 
A phase Ib study combining irinotecan with AZD1775, a selective WEE1 inhibitor, in RAS (KRAS or NRAS) or BRAF mutated metastatic colorectal cancer patients who have progressed on first-line therapy NCT02906059 2020 Colorectal AZD1775 plus irinotecan Nil Awaiting publication of results 
A phase II study of cisplatin + AZD1775 in metastatic triple-negative breast cancer and evaluation of pCDC2 as a biomarker of target response NCT03012477 2020 Breast AZD1775 plus cisplatin Nil 34 ORR: 26% (95% CI: 13–44%). mPFS: 4.9 months (95% CI: 2.3–5.7) 
A biomarker-enriched, randomized phase II trial of adavosertib (AZD1775) plus paclitaxel and carboplatin for women with platinum-sensitive TP53-mutant ovarian cancer [104NCT01357161 2020 Ovarian AZD1775 plus paclitaxcel and carboplatin Placebo plus paclitaxcel plus carboplatin 121 Adavosertib improved ePFS: 7.9 vs 7.3 months (P<0.2) 
Open-label, multicenter, phase I study to assess safety and tolerability of adavosertib plus durvalumab in patients with advanced solid tumors [107NCT02617277 2019 Solid tumors AZD1775 plus durvalumab Nil 54 Disease control rate was 36% 
Adavosertib with chemotherapy (CT) in patients (pts) with platinum-resistant ovarian cancer (PPROC): an open-label, four-arm, phase II study [102]. NCT02272790 2019 Ovarian AZD1775 plus chemotherapy Other chemotherapy regimens 94 ORR in combination with cisplatin was 67% with mPFS 10.1 months 
A randomized double-blind placebo-controlled phase II trial comparing gemcitabine monotherapy to gemcitabine in combination with adavosertib in women with recurrent, platinum resistant epithelial ovarian cancer: a trial of the Princess Margaret, California, Chicago and Mayo Phase II Consortia [103]. NCT02151292 2019 Ovarian AZD1775 plus gemcitabine Gemcitabine alone 124 mPFS greater with AZD1775 (3.0–4.6 months, P=0.015) and overall survival (7.2–11.5 months, P=0.022). Greater response rate with AZD1775 
Phase Ib study of adavosertib in combination with olaparib in patients with refractory solid tumors: dose escalation [106NCT02511795 2019 Solid tumors AZD1775 plus olaparib Nil 119 ORR: 11.1%; disease control rate: 55.7% 
VIOLETTE: a randomized phase II study to assess the DNA damage response inhibitors AZD6738 or AZD1775 in combination with olaparib (Ola) versus Ola monotherapy in patients (pts) with metastatic, triple-negative breast cancer (TNBC). NCT03330847 2019 Triple-negative breast cancer AZD1775 plus olaparib Olaparib alone 273 No significant difference in mPFS or ORR 
A phase 2 study of WEE1 inhibition with AZD1775 alone or combined with cytarabine in patients with advanced acute myeloid leukemia and myelodysplastic syndrome NCT02666950 2018 Acute myeloid leukemia and myelodysplastic syndrome AZD1775 AZD1775 plus cytarabine No responses seen 
A phase I clinical trial of AZD1775 in combination with neoadjuvant weekly docetaxel and cisplatin prior to surgery in squamous cell carcinoma of the head and neck (HNSCC) [109NCT02508246 2018 Head and neck cancer (SCC) AZD1775 plus cisplatin plus docetaxcel Nil 10 Seven patients (70%) had a response. Two complete responses and four pathological responses 
Dose escalation trial of the Wee1 inhibitor AZD1775, in combination with gemcitabine (+ radiation) for patients with unresectable adenocarcinoma of the pancreas [110NCT02037230 2018 Pancreatic AZD1775 plus gemicibabine plus radiotherapy Nil 34 Median overall survival: 21.7 months (90% CI: 16.7–24.8); mPFS: 9.4 months (90% CI: 8.0–9.9) 
Phase II, single-arm study of AZD1775 monotherapy in relapsed small cell lung cancer patients [111NCT02593019 2018 Small-cell lung cancer AZD1775 Nil No objective responses, stable disease in three (42.9%) 
A phase Ib, dose finding study evaluating AZD1775 in monotherapy, in combination with carboplatin and paclitaxel, and in combination with only carboplatin in adult asian patients with advanced solid tumours [112NCT02341456 2018 Solid tumors AZD1775 AZD1775 plus cisplatin or paclitaxcel 19 Partial response in two patients (16.7%) 
Phase I study evaluating WEE1 inhibitor AZD1775 as monotherapy and in combination with gemcitabine, cisplatin, or carboplatin in patients with advanced solid tumors [101NCT00648648 2016 Solid tumors AZD1775 AZD1775 plus chemotherapy 173 Partial response in 17 patients (10%). Stable disease in 94 (53%) 
A phase I study of single-agent AZD1775 (MK-1775), a Wee1 inhibitor, in patients with advanced refractory solid tumors [113NCT01748825 2015 Solid tumors AZD1775 Nil 25 Two partial responses (8%) in BRCA cohort 
A phase II study of AZD1775 plus pemetrexed and carboplatin followed by a randomised comparison of pemetrexed and carboplatin with or without AZD1775 in patients with previously untreated stage IV non-squamous non-small-cell lung cancer NCT02087241 2015 Non-small-cell lung cancer AZD1775 plus pemetrexed plus carboplatin Placebo plus pemetrexed plus carboplatin 14 (terminated) ORR 35.7% (5 responses from 14 patients before trial terminated) 
Study titleNCT trial numberYear completedCancer(s)WEE1 inhibitorComparatorSample sizeRelevant results
A phase I trial of WEE1 inhibition with chemotherapy and radiotherapy as adjuvant treatment, and a window of opportunity trial with cisplatin in patients with head and neck cancer NCT03028766 2021 Head and neck cancer AZD1775 plus cisplatin plus radiotherapy AZD1775 plus cisplatin 58 Awaiting publication of results 
Phase Ib trial of dose-escalating AZD1775 in combination with concurrent radiation and cisplatin for intermediate and high risk head and neck squamous cell carcinoma (HNSCC) [108NCT02585973 2021 Head and neck cancer (SCC) AZD1775 plus cisplatin plus radiotherapy Nil 12 ORR: 100% at 3 months, mPFS and median overall survival were 90% 
A phase Ib study combining irinotecan with AZD1775, a selective WEE1 inhibitor, in RAS (KRAS or NRAS) or BRAF mutated metastatic colorectal cancer patients who have progressed on first-line therapy NCT02906059 2020 Colorectal AZD1775 plus irinotecan Nil Awaiting publication of results 
A phase II study of cisplatin + AZD1775 in metastatic triple-negative breast cancer and evaluation of pCDC2 as a biomarker of target response NCT03012477 2020 Breast AZD1775 plus cisplatin Nil 34 ORR: 26% (95% CI: 13–44%). mPFS: 4.9 months (95% CI: 2.3–5.7) 
A biomarker-enriched, randomized phase II trial of adavosertib (AZD1775) plus paclitaxel and carboplatin for women with platinum-sensitive TP53-mutant ovarian cancer [104NCT01357161 2020 Ovarian AZD1775 plus paclitaxcel and carboplatin Placebo plus paclitaxcel plus carboplatin 121 Adavosertib improved ePFS: 7.9 vs 7.3 months (P<0.2) 
Open-label, multicenter, phase I study to assess safety and tolerability of adavosertib plus durvalumab in patients with advanced solid tumors [107NCT02617277 2019 Solid tumors AZD1775 plus durvalumab Nil 54 Disease control rate was 36% 
Adavosertib with chemotherapy (CT) in patients (pts) with platinum-resistant ovarian cancer (PPROC): an open-label, four-arm, phase II study [102]. NCT02272790 2019 Ovarian AZD1775 plus chemotherapy Other chemotherapy regimens 94 ORR in combination with cisplatin was 67% with mPFS 10.1 months 
A randomized double-blind placebo-controlled phase II trial comparing gemcitabine monotherapy to gemcitabine in combination with adavosertib in women with recurrent, platinum resistant epithelial ovarian cancer: a trial of the Princess Margaret, California, Chicago and Mayo Phase II Consortia [103]. NCT02151292 2019 Ovarian AZD1775 plus gemcitabine Gemcitabine alone 124 mPFS greater with AZD1775 (3.0–4.6 months, P=0.015) and overall survival (7.2–11.5 months, P=0.022). Greater response rate with AZD1775 
Phase Ib study of adavosertib in combination with olaparib in patients with refractory solid tumors: dose escalation [106NCT02511795 2019 Solid tumors AZD1775 plus olaparib Nil 119 ORR: 11.1%; disease control rate: 55.7% 
VIOLETTE: a randomized phase II study to assess the DNA damage response inhibitors AZD6738 or AZD1775 in combination with olaparib (Ola) versus Ola monotherapy in patients (pts) with metastatic, triple-negative breast cancer (TNBC). NCT03330847 2019 Triple-negative breast cancer AZD1775 plus olaparib Olaparib alone 273 No significant difference in mPFS or ORR 
A phase 2 study of WEE1 inhibition with AZD1775 alone or combined with cytarabine in patients with advanced acute myeloid leukemia and myelodysplastic syndrome NCT02666950 2018 Acute myeloid leukemia and myelodysplastic syndrome AZD1775 AZD1775 plus cytarabine No responses seen 
A phase I clinical trial of AZD1775 in combination with neoadjuvant weekly docetaxel and cisplatin prior to surgery in squamous cell carcinoma of the head and neck (HNSCC) [109NCT02508246 2018 Head and neck cancer (SCC) AZD1775 plus cisplatin plus docetaxcel Nil 10 Seven patients (70%) had a response. Two complete responses and four pathological responses 
Dose escalation trial of the Wee1 inhibitor AZD1775, in combination with gemcitabine (+ radiation) for patients with unresectable adenocarcinoma of the pancreas [110NCT02037230 2018 Pancreatic AZD1775 plus gemicibabine plus radiotherapy Nil 34 Median overall survival: 21.7 months (90% CI: 16.7–24.8); mPFS: 9.4 months (90% CI: 8.0–9.9) 
Phase II, single-arm study of AZD1775 monotherapy in relapsed small cell lung cancer patients [111NCT02593019 2018 Small-cell lung cancer AZD1775 Nil No objective responses, stable disease in three (42.9%) 
A phase Ib, dose finding study evaluating AZD1775 in monotherapy, in combination with carboplatin and paclitaxel, and in combination with only carboplatin in adult asian patients with advanced solid tumours [112NCT02341456 2018 Solid tumors AZD1775 AZD1775 plus cisplatin or paclitaxcel 19 Partial response in two patients (16.7%) 
Phase I study evaluating WEE1 inhibitor AZD1775 as monotherapy and in combination with gemcitabine, cisplatin, or carboplatin in patients with advanced solid tumors [101NCT00648648 2016 Solid tumors AZD1775 AZD1775 plus chemotherapy 173 Partial response in 17 patients (10%). Stable disease in 94 (53%) 
A phase I study of single-agent AZD1775 (MK-1775), a Wee1 inhibitor, in patients with advanced refractory solid tumors [113NCT01748825 2015 Solid tumors AZD1775 Nil 25 Two partial responses (8%) in BRCA cohort 
A phase II study of AZD1775 plus pemetrexed and carboplatin followed by a randomised comparison of pemetrexed and carboplatin with or without AZD1775 in patients with previously untreated stage IV non-squamous non-small-cell lung cancer NCT02087241 2015 Non-small-cell lung cancer AZD1775 plus pemetrexed plus carboplatin Placebo plus pemetrexed plus carboplatin 14 (terminated) ORR 35.7% (5 responses from 14 patients before trial terminated) 

Potential resistance mechanisms to WEE1i continue to be investigated, although suggested mechanisms include restoration of the G1/S cell-cycle checkpoint or up-regulation of other survival pathways in order to avoid mitotic catastrophe. One means of overcoming the former includes concurrent use of CDK4/6 inhibitors to remove the G1 checkpoint; this combination has been shown to act synergistically in the preclinical setting [116].

BER targets

As discussed above, BER plays an essential role in DNA repair. Up-regulation of BER is thought to contribute to chemoresistance, rationalizing the pathway as a pharmacological target. Further to this, the BER pathway may be a source of novel synthetic lethality targets as HR-deficient cells would lose the means of repairing both single- and double-strand breaks.

One emerging BER target is thought to be APE1; small-molecule inhibitors of the enzyme have been shown to be synthetically lethal in vitro to BRCA- and ATM-deficient cell lines [117]. APE1 is often overexpressed and associated with worse prognosis in NSCLC. In NSCLC cell lines, APE1 inhibition induced apoptosis, overcame chemotherapy resistance, and impeded cancer progression in a mouse model [118]. The APE1 inhibitor APX3330 was well tolerated in an early phase I trial, most commonly causing grade 1 fatigue, and demonstrated antitumor activity [119].

Similarly to APE1, FEN1 is often overexpressed in tumors and is particularly associated with development of chemoresistance. FEN1 inhibition was assessed in ovarian cancer cell lines and was demonstrated to potentiate cisplatin cytotoxicity as well as being synthetically lethal to BRCA2-deficient cells. In a similar manner to PARPi, resistance arose following restoration of BRCA2 function [120].

XRCC1 plays an integral role in the BER, SSBR, and back-up NHEJ pathways. Loss of XRCC1 has been shown to correlate with more aggressive cancers and worse prognosis. Intriguingly, PARP, ATM, ATR, WEE1, Mre11, and DNA-PKcs inhibitors have all been found to be synthetically lethal to XRCC1-deficient cells, highlighting a novel therapeutic avenue in these aggressive tumors [121–124].

DNA polymerases

DNA polymerases, such as polβ and polθ, are integral components of DNA repair pathways. Polβ is vital for BER and hence maintaining genomic integrity. In ovarian cancer, high polβ expression was associated with worse patient outcomes while in vitro polβ depletion led to increased platinum sensitivity [125]. Furthermore, polβ inhibition has been shown to be synthetically lethal to both BRCA1- [126] and BRCA2-deficient cell lines [125]. Despite these promising findings, polβ inhibitors are yet to enter clinical trials possibly due to challenges in identifying suitably potent and specific inhibitors for in vivo use.

Polθ predominantly acts to repair DSBs through MMEJ, although recent evidence suggests that it may possess additional functions such as DNA cross-link repair or within the BER pathway. It is commonly overexpressed in many cancers, typically correlating with other HR defects and worse patient outcomes. Furthermore, overexpression of polθ has been shown to contribute to resistance to DNA-damaging agents such as radiotherapy, chemotherapy agents, and PARPi [31]. Following the discovery of a selective polθ inhibitor and using knockout models, it was identified that inhibition or loss of polθ can induce synthetic lethality in BRCA- and HR-deficient cells [31,127]. It has been suggested that polθ inhibitors may be of benefit in combination with other DNA repair inhibitors, such as PARPi and ATRi, as well as standard chemotherapy agents. For instance, loss of 53BP1 is thought to be a mechanism of PARPi resistance yet 53BP1 and polθ have been shown to be a synthetically lethal pairing. This therefore rationalizes combination strategies of PARP and polθ inhibitors as a means of preventing resistance. The polθ inhibitor ART4215 is the first to enter clinical trials and is being assessed for safety, tolerability, and preliminary efficacy in patients with advanced solid tumors, as both monotherapy and in combination with talazoparib or niraparib (NCT04991480).

Other PARPi

The PARP family of proteins currently contains 17 members with wide-ranging cellular functions, including within DNA repair and mitosis. While currently licensed PARPi predominantly target PARP1 to PARP3, the other members of the PARP family may offer an avenue to novel therapies. The cellular functions and significance of each of these family members is reviewed in [128] but of particular clinical significance are PARP6 and PARP7. PARP6 inhibition has been shown to cause multipolar spindle (MPS) formation and centrosomal defects which in turn, caused cancer cell apoptosis both in vitro and in vivo. Additionally, PARP6 was identified to act on CHK1 and inhibition therefore prevents CHK1 modification, resulting in defective mitotic signaling [129]. PARP7 has a variety of roles but importantly loss-of-function results in increased microtubule stability leading to reduced mitotic rate as well as slowing of migration of ovarian cancer cells [128]. A selective PARP7 inhibitor, RBN-2397, has demonstrated good preclinical efficacy in lung cancer xenografts [130] and is now being evaluated in a phase I trial for treating advanced solid tumors (NCT04053673).

DDR is a critical defense mechanism against genomic instability. Our current understanding of the DDR process has led to several translational investigations culminating in clinically viable precision oncology strategies. This is best exemplified by the current clinical use of PARPi in BRCA germline-deficient breast or ovarian cancers and platinum-sensitive sporadic epithelial ovarian cancers. However, response rate to PARPi is about 50% and progression-free survival is only about 7 months. Therefore, development of intrinsic and acquired resistance remains a clinical challenge. The development of biomarkers of response to PARP and other DDR inhibitor therapies remains an area of unmet clinical need. The importance and current development of validated, predictive biomarkers in relation to DDR inhibitors is reviewed in [131]. More recently to address these challenges, several new potential drugs such as those targeting ATM, ATR, WEE1, and others have emerged. These next-generation DNA repair inhibitors either as monotherapy or in combination with PAPR inhibitors could potentially improve outcomes but will need to be tested in phase III randomized trials in the future. Preclinically, several novel DNA repair targets are under evaluation. Finally, discovery of additional synthetic lethality interaction partners focused on DDR remains an area of intense investigation and will help advances in precision medicine strategies for cancer patients.

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

Sanat Kulkarni: Conceptualization, Resources, Data curation, Software, Formal Analysis, Validation, Investigation, Visualization, Methodology, Writing—original draft, Writing—review & editing. Juliette Brownlie: Conceptualization, Writing—original draft, Project administration, Writing—review & editing. Jennie N Jeyapalan: Conceptualization, Writing—original draft, Project administration, Writing—review & editing. Nigel P Mongan: Conceptualization, Writing—original draft, Project administration, Writing—review & editing. Emad A Rakha: Conceptualization, Writing—original draft, Project administration, Writing—review & editing. Srinivasan Madhusudan: Conceptualization, Resources, Data curation, Formal Analysis, Supervision, Validation, Investigation, Methodology, Writing—original draft, Project administration, Writing—review & editing.

53BP1

p53-binding protein 1

AGT

O6-alkylguanine-DNA alkyltransferase

AP

apurinic/apyrimidinic

APE1

AP-endonuclease 1

ATM

ataxia telangiectasia mutated

ATMi

ATM inhibitor

ATR

ataxia telangiectasia and Rad3-related

ATRi

ATR inhibitor

ATRIP

ATR-interacting protein

BER

base excision repair

BRCT

BRCA1 C terminus

CHK1

checkpoint kinase 1

CPD

cyclobutane pyrimidine dimer

DDR

DNA damage signaling response and repair

DNA-PKcs

DNA-dependent protein kinase catalytic subunit

DSB

double-strand break

FA

fanconi anemia

FEN1

flap-endonuclease 1

GG-NER

global-genome NER

HNPCC

hereditary nonpolyposis colorectal cancer

HR

homologous recombination

ICL

interstrand cross-link

IDL

insertion/deletion loop

MGMT

methylguanine methyltransferase

MMEJ

microhomology-mediated end joining

MMR

mismatch repair

mPFS

medial progression-free survival

MPS

multipolar spindle

NER

nucleotide excision repair

PARP

poly(ADP-ribose) polymerase

PARPi

PARP inhibitor

PCNA

proliferating cell nuclear antigen

PDL-1

programmed death-ligand 1

PNKP

polynucleotide kinase/phosphatase

polβ

polymerase-β

PTEN

phosphate and tensin homolog

ROS

reactive oxygen species

RPA

replication protein A

SDSA

synthesis-dependent strand annealing

SSA

single-strand annealing

SSB

single-strand DNA break

SSBR

single-strand break repair

ssDNA

single-strand DNA

TC-NER

transcription-coupled NER

TNBC

triple-negative breast cancer

UV

ultraviolet

WEE1i

WEE1 kinase inhibitor

1.
Hiom
K.
(
2003
)
DNA repair: bacteria join in
.
Curr. Biol.
13
,
R28
R30
[PubMed]
2.
Friedberg
E.C.
(
2001
)
How nucleotide excision repair protects against cancer
.
Nat. Rev. Cancer
1
,
22
33
[PubMed]
3.
Abbotts
R.
,
Thompson
N.
and
Madhusudan
S.
(
2014
)
DNA repair in cancer: emerging targets for personalized therapy
.
Cancer Manag. Res.
6
,
77
92
[PubMed]
4.
Yi
C.
and
He
C.
(
2013
)
DNA repair by reversal of DNA damage
.
Cold Spring Harb. Perspect. Biol.
5
,
a012575
[PubMed]
5.
Ramirez-Gamboa
D.
,
Diaz-Zamorano
A.L.
,
Melendez-Sanchez
E.R.
,
Reyes-Pardo
H.
,
Villasenor-Zepeda
K.R.
,
Lopez-Arellanes
M.E.
et al.
(
2022
)
Photolyase production and current applications: a review
.
Molecules
27
,
5998
6015
6.
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]
7.
Mishina
Y.
,
Duguid
E.M.
and
He
C.
(
2006
)
Direct reversal of DNA alkylation damage
.
Chem. Rev.
106
,
215
232
[PubMed]
8.
Esteller
M.
,
Garcia-Foncillas
J.
,
Andion
E.
,
Goodman
S.N.
,
Hidalgo
O.F.
,
Vanaclocha
V.
et al.
(
2000
)
Inactivation of the DNA-repair gene MGMT and the clinical response of gliomas to alkylating agents
.
N. Engl. J. Med.
343
,
1350
1354
[PubMed]
9.
Qi
Z.
and
Tan
H.
(
2020
)
Association between MGMT status and response to alkylating agents in patients with neuroendocrine neoplasms: a systematic review and meta-analysis
.
Biosci. Rep.
40
,
1
8
10.
Duncan
T.
,
Trewick
S.C.
,
Koivisto
P.
,
Bates
P.A.
,
Lindahl
T.
and
Sedgwick
B.
(
2002
)
Reversal of DNA alkylation damage by two human dioxygenases
.
Proc. Natl. Acad. Sci. U.S.A.
99
,
16660
16665
[PubMed]
11.
Seeberg
E.
,
Eide
L.
and
Bjoras
M.
(
1995
)
The base excision repair pathway
.
Trends Biochem. Sci.
20
,
391
397
[PubMed]
12.
Robertson
A.B.
,
Klungland
A.
,
Rognes
T.
and
Leiros
I.
(
2009
)
DNA repair in mammalian cells: base excision repair: the long and short of it
.
Cell. Mol. Life Sci.
66
,
981
993
[PubMed]
13.
Fortini
P.
,
Pascucci
B.
,
Parlanti
E.
,
D'Errico
M.
,
Simonelli
V.
and
Dogliotti
E.
(
2003
)
The base excision repair: mechanisms and its relevance for cancer susceptibility
.
Biochimie
85
,
1053
1071
[PubMed]
14.
Ko
H.L.
and
Ren
E.C.
(
2012
)
Functional aspects of PARP1 in DNA repair and transcription
.
Biomolecules
2
,
524
548
[PubMed]
15.
Demin
A.A.
,
Hirota
K.
,
Tsuda
M.
,
Adamowicz
M.
,
Hailstone
R.
,
Brazina
J.
et al.
(
2021
)
XRCC1 prevents toxic PARP1 trapping during DNA base excision repair
.
Mol. Cell.
81
,
3018e5
3030e5
16.
Ronson
G.E.
,
Piberger
A.L.
,
Higgs
M.R.
,
Olsen
A.L.
,
Stewart
G.S.
,
McHugh
P.J.
et al.
(
2018
)
PARP1 and PARP2 stabilise replication forks at base excision repair intermediates through Fbh1-dependent Rad51 regulation
.
Nat. Commun.
9
,
746
[PubMed]
17.
Caldecott
K.W.
(
2014
)
DNA single-strand break repair
.
Exp. Cell. Res.
329
,
2
8
[PubMed]
18.
de Laat
W.L.
,
Jaspers
N.G.
and
Hoeijmakers
J.H.
(
1999
)
Molecular mechanism of nucleotide excision repair
.
Genes Dev.
13
,
768
785
[PubMed]
19.
Shuck
S.C.
,
Short
E.A.
and
Turchi
J.J.
(
2008
)
Eukaryotic nucleotide excision repair: from understanding mechanisms to influencing biology
.
Cell Res.
18
,
64
72
[PubMed]
20.
Spivak
G.
(
2015
)
Nucleotide excision repair in humans
.
DNA Repair (Amst.)
36
,
13
18
[PubMed]
21.
Copeland
N.E.
,
Hanke
C.W.
and
Michalak
J.A.
(
1997
)
The molecular basis of xeroderma pigmentosum
.
Dermatol. Surg.
23
,
447
455
[PubMed]
22.
Bebenek
A.
and
Ziuzia-Graczyk
I.
(
2018
)
Fidelity of DNA replication-a matter of proofreading
.
Curr. Genet.
64
,
985
996
[PubMed]
23.
Kunkel
T.A.
and
Erie
D.A.
(
2005
)
DNA mismatch repair
.
Annu. Rev. Biochem.
74
,
681
710
[PubMed]
24.
Kolodner
R.D.
and
Marsischky
G.T.
(
1999
)
Eukaryotic DNA mismatch repair
.
Curr. Opin. Genet. Dev.
9
,
89
96
[PubMed]
25.
Papadopoulos
N.
and
Lindblom
A.
(
1997
)
Molecular basis of HNPCC: mutations of MMR genes
.
Hum. Mutat.
10
,
89
99
[PubMed]
26.
Wang
C.
and
Lees-Miller
S.P.
(
2013
)
Detection and repair of ionizing radiation-induced DNA double strand breaks: new developments in nonhomologous end joining
.
Int. J. Radiat. Oncol. Biol. Phys.
86
,
440
449
[PubMed]
27.
Chang
H.H.Y.
,
Pannunzio
N.R.
,
Adachi
N.
and
Lieber
M.R.
(
2017
)
Non-homologous DNA end joining and alternative pathways to double-strand break repair
.
Nat. Rev. Mol. Cell Biol.
18
,
495
506
[PubMed]
28.
Adachi
N.
,
Ishino
T.
,
Ishii
Y.
,
Takeda
S.
and
Koyama
H.
(
2001
)
DNA ligase IV-deficient cells are more resistant to ionizing radiation in the absence of Ku70: Implications for DNA double-strand break repair
.
Proc. Natl. Acad. Sci. U.S.A.
98
,
12109
12113
[PubMed]
29.
Sfeir
A.
and
Symington
L.S.
(
2015
)
Microhomology-mediated end joining: a back-up survival mechanism or dedicated pathway?
Trends Biochem. Sci.
40
,
701
714
[PubMed]
30.
Bhargava
R.
,
Onyango
D.O.
and
Stark
J.M.
(
2016
)
Regulation of single-strand annealing and its role in genome maintenance
.
Trends Genet.
32
,
566
575
[PubMed]
31.
Chen
X.S.
and
Pomerantz
R.T.
(
2021
)
DNA polymerase theta: a cancer drug target with reverse transcriptase activity
.
Genes (Basel)
12
,
1146
1159
32.
Shrivastav
M.
,
De Haro
L.P.
and
Nickoloff
J.A.
(
2008
)
Regulation of DNA double-strand break repair pathway choice
.
Cell Res.
18
,
134
147
[PubMed]
33.
Brandsma
I.
and
Gent
D.C.
(
2012
)
Pathway choice in DNA double strand break repair: observations of a balancing act
.
Genome Integr
3
,
9
[PubMed]
34.
Li
X.
and
Heyer
W.D.
(
2008
)
Homologous recombination in DNA repair and DNA damage tolerance
.
Cell Res.
18
,
99
113
[PubMed]
35.
Chen
S.
and
Parmigiani
G.
(
2007
)
Meta-analysis of BRCA1 and BRCA2 penetrance
.
J. Clin. Oncol.
25
,
1329
1333
[PubMed]
36.
Deans
A.J.
and
West
S.C.
(
2011
)
DNA interstrand crosslink repair and cancer
.
Nat. Rev. Cancer
11
,
467
480
[PubMed]
37.
Loeb
L.A.
,
Bielas
J.H.
and
Beckman
R.A.
(
2008
)
Cancers exhibit a mutator phenotype: clinical implications
.
Cancer Res.
68
,
3551
3557
,
discussion 7
[PubMed]
38.
Abdel-Fatah
T.M.
,
Russell
R.
,
Agarwal
D.
,
Moseley
P.
,
Abayomi
M.A.
,
Perry
C.
et al.
(
2014
)
DNA polymerase beta deficiency is linked to aggressive breast cancer: a comprehensive analysis of gene copy number, mRNA and protein expression in multiple cohorts
.
Mol. Oncol.
8
,
520
532
[PubMed]
39.
Gachechiladze
M.
,
Skarda
J.
,
Bouchalova
K.
,
Soltermann
A.
and
Joerger
M.
(
2020
)
Predictive and prognostic value of DNA damage response associated kinases in solid tumors
.
Front. Oncol.
10
,
581217
[PubMed]
40.
Madhusudan
S.
and
Hickson
I.D.
(
2005
)
DNA repair inhibition: a selective tumour targeting strategy
.
Trends Mol. Med.
11
,
503
511
[PubMed]
41.
Martorana
F.
,
Da Silva
L.A.
,
Sessa
C.
and
Colombo
I.
(
2022
)
Everything comes with a price: the toxicity profile of DNA-damage response targeting agents
.
Cancers (Basel)
14
,
953
975
42.
Gerson
S.L.
(
2004
)
MGMT: its role in cancer aetiology and cancer therapeutics
.
Nat. Rev. Cancer
4
,
296
307
[PubMed]
43.
Herath
N.I.
,
Berthault
N.
,
Thierry
S.
,
Jdey
W.
,
Lienafa
M.C.
,
Bono
F.
et al.
(
2019
)
Preclinical studies comparing efficacy and toxicity of DNA repair inhibitors, olaparib, and AsiDNA, in the treatment of carboplatin-resistant tumors
.
Front Oncol.
9
,
1097
[PubMed]
44.
O'Connor
M.J.
(
2015
)
Targeting the DNA damage response in cancer
.
Mol. Cell.
60
,
547
560
[PubMed]
45.
Ferreira
S.
and
Dutreix
M.
(
2019
)
DNA repair inhibitors to enhance radiotherapy: progresses and limitations
.
Cancer Radiother.
23
,
883
890
[PubMed]
46.
Lord
C.J.
and
Ashworth
A.
(
2017
)
PARP inhibitors: synthetic lethality in the clinic
.
Science
355
,
1152
1158
[PubMed]
47.
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]
48.
Valabrega
G.
,
Scotto
G.
,
Tuninetti
V.
,
Pani
A.
and
Scaglione
F.
(
2021
)
Differences in PARP inhibitors for the treatment of ovarian cancer: mechanisms of action, pharmacology, safety, and efficacy
.
Int. J. Mol. Sci.
22
,
4203
4219
[PubMed]
49.
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]
50.
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]
51.
Narod
S.A.
and
Foulkes
W.D.
(
2004
)
BRCA1 and BRCA2: 1994 and beyond
.
Nat. Rev. Cancer
4
,
665
676
[PubMed]
52.
Kristeleit
R.
,
Lisyanskaya
A.
,
Fedenko
A.
,
Dvorkin
M.
,
de Melo
A.C.
,
Shparyk
Y.
et al.
(
2022
)
Rucaparib versus standard-of-care chemotherapy in patients with relapsed ovarian cancer and a deleterious BRCA1 or BRCA2 mutation (ARIEL4): an international, open-label, randomised, phase 3 trial
.
Lancet Oncol.
23
,
465
478
[PubMed]
53.
Banerjee
S.
,
Moore
K.N.
,
Colombo
N.
,
Scambia
G.
,
Kim
B.G.
,
Oaknin
A.
et al.
(
2021
)
Maintenance olaparib for patients with newly diagnosed advanced ovarian cancer and a BRCA mutation (SOLO1/GOG 3004): 5-year follow-up of a randomised, double-blind, placebo-controlled, phase 3 trial
.
Lancet Oncol.
22
,
1721
1731
[PubMed]
54.
Dieras
V.
,
Han
H.S.
,
Kaufman
B.
,
Wildiers
H.
,
Friedlander
M.
,
Ayoub
J.P.
et al.
(
2020
)
Veliparib with carboplatin and paclitaxel in BRCA-mutated advanced breast cancer (BROCADE3): a randomised, double-blind, placebo-controlled, phase 3 trial
.
Lancet Oncol.
21
,
1269
1282
[PubMed]
55.
Penson
R.T.
,
Valencia
R.V.
,
Cibula
D.
,
Colombo
N.
,
Leath
C.A.
3rd
,
Bidzinski
M.
et al.
(
2020
)
Olaparib versus nonplatinum chemotherapy in patients with platinum-sensitive relapsed ovarian cancer and a germline BRCA1/2 Mutation (SOLO3): a randomized phase III trial
.
J. Clin. Oncol.
38
,
1164
1174
[PubMed]
56.
Coleman
R.L.
,
Fleming
G.F.
,
Brady
M.F.
,
Swisher
E.M.
,
Steffensen
K.D.
,
Friedlander
M.
et al.
(
2019
)
Veliparib with first-line chemotherapy and as maintenance therapy in ovarian cancer
.
N. Engl. J. Med.
381
,
2403
2415
[PubMed]
57.
Golan
T.
,
Hammel
P.
,
Reni
M.
,
Van Cutsem
E.
,
Macarulla
T.
,
Hall
M.J.
et al.
(
2019
)
Maintenance olaparib for germline BRCA-mutated metastatic pancreatic cancer
.
N. Engl. J. Med.
381
,
317
327
[PubMed]
58.
Litton
J.K.
,
Rugo
H.S.
,
Ettl
J.
,
Hurvitz
S.A.
,
Goncalves
A.
,
Lee
K.H.
et al.
(
2018
)
Talazoparib in patients with advanced breast cancer and a germline BRCA mutation
.
N. Engl. J. Med.
379
,
753
763
[PubMed]
59.
Robson
M.
,
Im
S.A.
,
Senkus
E.
,
Xu
B.
,
Domchek
S.M.
,
Masuda
N.
et al.
(
2017
)
Olaparib for metastatic breast cancer in patients with a germline BRCA mutation
.
N. Engl. J. Med.
377
,
523
533
[PubMed]
60.
Pujade-Lauraine
E.
,
Ledermann
J.A.
,
Selle
F.
,
Gebski
V.
,
Penson
R.T.
,
Oza
A.M.
et al.
(
2017
)
Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): a double-blind, randomised, placebo-controlled, phase 3 trial
.
Lancet Oncol.
18
,
1274
1284
[PubMed]
61.
Mirza
M.R.
,
Monk
B.J.
,
Herrstedt
J.
,
Oza
A.M.
,
Mahner
S.
,
Redondo
A.
et al.
(
2016
)
Niraparib maintenance therapy in platinum-sensitive, recurrent ovarian cancer
.
N. Engl. J. Med.
375
,
2154
2164
[PubMed]
62.
Han
H.S.
,
Dieras
V.
,
Robson
M.
,
Palacova
M.
,
Marcom
P.K.
,
Jager
A.
et al.
(
2018
)
Veliparib with temozolomide or carboplatin/paclitaxel versus placebo with carboplatin/paclitaxel in patients with BRCA1/2 locally recurrent/metastatic breast cancer: randomized phase II study
.
Ann. Oncol.
29
,
154
161
[PubMed]
63.
Fasching
P.A.
,
Jackisch
C.
,
Rhiem
K.
,
Schneeweiss
A
,
Klare
P
,
Hanusch
C
et al.
(
2019
)
GeparOLA: a randomized phase II trial to assess the efficacy of paclitaxel and olaparib in comparison to paclitaxel/carboplatin followed by epirubicin/cyclophosphamide as neoadjuvant chemotherapy in patients (pts) with HER2-negative early breast cancer (BC) and homologous recombination deficiency (HRD)
.
J. Clin. Oncol.
37
,
506
64.
McCabe
N.
,
Turner
N.C.
,
Lord
C.J.
,
Kluzek
K.
,
Bialkowska
A.
,
Swift
S.
et al.
(
2006
)
Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition
.
Cancer Res.
66
,
8109
8115
[PubMed]
65.
Lord
C.J.
and
Ashworth
A.
(
2016
)
BRCAness revisited
.
Nat. Rev. Cancer
16
,
110
120
[PubMed]
66.
Mateo
J.
,
Porta
N.
,
Bianchini
D.
,
McGovern
U.
,
Elliott
T.
,
Jones
R.
et al.
(
2020
)
Olaparib in patients with metastatic castration-resistant prostate cancer with DNA repair gene aberrations (TOPARP-B): a multicentre, open-label, randomised, phase 2 trial
.
Lancet Oncol.
21
,
162
174
[PubMed]
67.
Nizialek
E.
and
Antonarakis
E.S.
(
2020
)
PARP inhibitors in metastatic prostate cancer: evidence to date
.
Cancer Manag Res.
12
,
8105
8114
[PubMed]
68.
Crabb
S.J.
,
Hussain
S.A.
,
Soulis
E.
,
Hinsley
S.
,
Dempsey
L.
,
Trevethan
A
et al.
(
2022
)
A randomized, double blind, biomarker selected, phase II clinical trial of maintenance PARP inhibition following chemotherapy for metastatic urothelial carcinoma (mUC): final analysis of the ATLANTIS rucaparib arm
.
J. Clin. Oncol.
40
,
436
69.
Kim
D.
and
Nam
H.J.
(
2022
)
PARP inhibitors: clinical limitations and recent attempts to overcome them
.
Int. J. Mol. Sci.
23
,
8412
8430
70.
Edwards
S.L.
,
Brough
R.
,
Lord
C.J.
,
Natrajan
R.
,
Vatcheva
R.
,
Levine
D.A.
et al.
(
2008
)
Resistance to therapy caused by intragenic deletion in BRCA2
.
Nature
451
,
1111
1115
[PubMed]
71.
Sakai
W.
,
Swisher
E.M.
,
Karlan
B.Y.
,
Agarwal
M.K.
,
Higgins
J.
,
Friedman
C.
et al.
(
2008
)
Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers
.
Nature
451
,
1116
1120
[PubMed]
72.
Xu
G.
,
Chapman
J.R.
,
Brandsma
I.
,
Yuan
J.
,
Mistrik
M.
,
Bouwman
P.
et al.
(
2015
)
REV7 counteracts DNA double-strand break resection and affects PARP inhibition
.
Nature
521
,
541
544
[PubMed]
73.
Vaidyanathan
A.
,
Sawers
L.
,
Gannon
A.L.
,
Chakravarty
P.
,
Scott
A.L.
,
Bray
S.E.
et al.
(
2016
)
ABCB1 (MDR1) induction defines a common resistance mechanism in paclitaxel- and olaparib-resistant ovarian cancer cells
.
Br. J. Cancer
115
,
431
441
[PubMed]
74.
Mei
L.
,
Zhang
J.
,
He
K.
and
Zhang
J.
(
2019
)
Ataxia telangiectasia and Rad3-related inhibitors and cancer therapy: where we stand
.
J. Hematol. Oncol.
12
,
43
[PubMed]
75.
Qiu
Z.
,
Oleinick
N.L.
and
Zhang
J.
(
2018
)
ATR/CHK1 inhibitors and cancer therapy
.
Radiother. Oncol.
126
,
450
464
[PubMed]
76.
Yap
T.A.
,
Krebs
M.
,
Postel-Vinay
S.
,
Bang
Y.-J.
,
El-Khoueiry
A.
,
Abida
W.
et al.
(
2016
)
Phase I modular study of AZD6738, a novel oral, potent and selective ataxia telangiectasia Rad3-related (ATR) inhibitor in combination (combo) with carboplatin, olaparib or durvalumab in patients (pts) with advanced cancers
.
Eur. J. Cancer
69
,
S2
77.
Telli
M.
,
Lord
S.
,
Dean
E.
,
Abramson
V.
,
Arkenau
H.-T.
,
Murias
C.
et al.
(
2018
)
Abstract OT2-07-07: ATR inhibitor M6620 (formerly VX-970) with cisplatin in metastatic triple-negative breast cancer: Preliminary results from a phase 1 dose expansion cohort (NCT02157792)
.
Cancer Res.
78
,
OT2-07–OT2–
78.
Yap
T.A.
,
O'Carrigan
B.
,
Penney
M.S.
,
Lim
J.S.
,
Brown
J.S.
,
de Miguel Luken
M.J.
et al.
(
2020
)
Phase I trial of first-in-class ATR inhibitor M6620 (VX-970) as monotherapy or in combination with carboplatin in patients with advanced solid tumors
.
J. Clin. Oncol.
38
,
3195
3204
[PubMed]
79.
Middleton
M.R.
,
Dean
E.
,
Evans
T.R.J.
,
Shapiro
G.I.
,
Pollard
J.
,
Hendriks
B.S.
et al.
(
2021
)
Phase 1 study of the ATR inhibitor berzosertib (formerly M6620, VX-970) combined with gemcitabine +/- cisplatin in patients with advanced solid tumours
.
Br. J. Cancer
125
,
510
519
[PubMed]
80.
Thomas
A.
,
Redon
C.E.
,
Sciuto
L.
,
Padiernos
E.
,
Ji
J.
,
Lee
M.J.
et al.
(
2018
)
Phase I study of ATR inhibitor M6620 in combination with topotecan in patients with advanced solid tumors
.
J. Clin. Oncol.
36
,
1594
1602
[PubMed]
81.
Huntoon
C.J.
,
Flatten
K.S.
,
Wahner Hendrickson
A.E.
,
Huehls
A.M.
,
Sutor
S.L.
,
Kaufmann
S.H.
et al.
(
2013
)
ATR inhibition broadly sensitizes ovarian cancer cells to chemotherapy independent of BRCA status
.
Cancer Res.
73
,
3683
3691
[PubMed]
82.
Lloyd
R.L.
,
Urban
V.
,
Munoz-Martinez
F.
,
Ayestaran
I.
,
Thomas
J.C.
,
de Renty
C.
et al.
(
2021
)
Loss of cyclin C or CDK8 provides ATR inhibitor resistance by suppressing transcription-associated replication stress
.
Nucleic. Acids. Res.
49
,
8665
8683
[PubMed]
83.
Guzi
T.J.
,
Paruch
K.
,
Dwyer
M.P.
,
Labroli
M.
,
Shanahan
F.
,
Davis
N.
et al.
(
2011
)
Targeting the replication checkpoint using SCH 900776, a potent and functionally selective CHK1 inhibitor identified via high content screening
.
Mol. Cancer Ther.
10
,
591
602
[PubMed]
84.
Otto
T.
and
Sicinski
P.
(
2017
)
Cell cycle proteins as promising targets in cancer therapy
.
Nat. Rev. Cancer
17
,
93
115
[PubMed]
85.
Zhou
Z.R.
,
Yang
Z.Z.
,
Wang
S.J.
,
Zhang
L.
,
Luo
J.R.
,
Feng
Y.
et al.
(
2017
)
The Chk1 inhibitor MK-8776 increases the radiosensitivity of human triple-negative breast cancer by inhibiting autophagy
.
Acta Pharmacol. Sin.
38
,
513
523
[PubMed]
86.
Daud
A.I.
,
Ashworth
M.T.
,
Strosberg
J.
,
Goldman
J.W.
,
Mendelson
D.
,
Springett
G.
et al.
(
2015
)
Phase I dose-escalation trial of checkpoint kinase 1 inhibitor MK-8776 as monotherapy and in combination with gemcitabine in patients with advanced solid tumors
.
J. Clin. Oncol.
33
,
1060
1066
[PubMed]
87.
Webster
J.A.
,
Tibes
R.
,
Morris
L.
,
Blackford
A.L.
,
Litzow
M.
,
Patnaik
M.
et al.
(
2017
)
Randomized phase II trial of cytosine arabinoside with and without the CHK1 inhibitor MK-8776 in relapsed and refractory acute myeloid leukemia
.
Leuk. Res.
61
,
108
116
[PubMed]
88.
Lee
J.M.
,
Nair
J.
,
Zimmer
A.
,
Lipkowitz
S.
,
Annunziata
C.M.
,
Merino
M.J.
et al.
(
2018
)
Prexasertib, a cell cycle checkpoint kinase 1 and 2 inhibitor, in BRCA wild-type recurrent high-grade serous ovarian cancer: a first-in-class proof-of-concept phase 2 study
.
Lancet Oncol.
19
,
207
215
[PubMed]
89.
Gatti-Mays
M.E.
,
Karzai
F.H.
,
Soltani
S.N.
,
Zimmer
A.
,
Green
J.E.
,
Lee
M.J.
et al.
(
2020
)
A phase II single arm pilot study of the CHK1 inhibitor prexasertib (LY2606368) in BRCA wild-type, advanced triple-negative breast cancer
.
Oncologist
25
,
e1013
e1824
[PubMed]
90.
Zhao
Y.
,
Zhou
K.
,
Xia
X.
,
Guo
Y.
and
Tao
L.
(
2022
)
Chk1 inhibition-induced BRCAness synergizes with olaparib in p53-deficient cancer cells
.
Cell Cycle
8
,
1
13
91.
Jin
M.H.
and
Oh
D.Y.
(
2019
)
ATM in DNA repair in cancer
.
Pharmacol. Ther.
203
,
107391
[PubMed]
92.
Zhang
X.
,
Zhang
Z.
,
Zhang
Q.
,
Zhang
Q.
,
Sun
P.
,
Xiang
R.
et al.
(
2018
)
ZEB1 confers chemotherapeutic resistance to breast cancer by activating ATM
.
Cell Death Dis.
9
,
57
[PubMed]
93.
McCabe
N.
,
Hanna
C.
,
Walker
S.M.
,
Gonda
D.
,
Li
J.
,
Wikstrom
K.
et al.
(
2015
)
Mechanistic rationale to target PTEN-deficient tumor cells with inhibitors of the DNA damage response kinase ATM
.
Cancer Res.
75
,
2159
2165
[PubMed]
94.
Li
K.
,
Yan
H.
,
Guo
W.
,
Tang
M.
,
Zhao
X.
,
Tong
A.
et al.
(
2018
)
ATM inhibition induces synthetic lethality and enhances sensitivity of PTEN-deficient breast cancer cells to cisplatin
.
Exp. Cell. Res.
366
,
24
33
[PubMed]
95.
Lavin
M.F.
and
Yeo
A.J.
(
2020
)
Clinical potential of ATM inhibitors
.
Mutat. Res.
821
,
111695
[PubMed]
96.
Jobson
A.G.
,
Lountos
G.T.
,
Lorenzi
P.L.
,
Llamas
J.
,
Connelly
J.
,
Cerna
D.
et al.
(
2009
)
Cellular inhibition of checkpoint kinase 2 (Chk2) and potentiation of camptothecins and radiation by the novel Chk2 inhibitor PV1019 [7-nitro-1H-indole-2-carboxylic acid {4-[1-(guanidinohydrazone)-ethyl]-phenyl}-amide]
.
J. Pharmacol. Exp. Ther.
331
,
816
826
[PubMed]
97.
Anderson
V.E.
,
Walton
M.I.
,
Eve
P.D.
,
Boxall
K.J.
,
Antoni
L.
,
Caldwell
J.J.
et al.
(
2011
)
CCT241533 is a potent and selective inhibitor of CHK2 that potentiates the cytotoxicity of PARP inhibitors
.
Cancer Res.
71
,
463
472
[PubMed]
98.
Sausville
E.
,
Lorusso
P.
,
Carducci
M.
,
Carter
J.
,
Quinn
M.F.
,
Malburg
L.
et al.
(
2014
)
Phase I dose-escalation study of AZD7762, a checkpoint kinase inhibitor, in combination with gemcitabine in US patients with advanced solid tumors
.
Cancer Chemother. Pharmacol.
73
,
539
549
[PubMed]
99.
Katayama
K.
,
Fujita
N.
and
Tsuruo
T.
(
2005
)
Akt/protein kinase B-dependent phosphorylation and inactivation of WEE1Hu promote cell cycle progression at G2/M transition
.
Mol. Cell. Biol.
25
,
5725
5737
[PubMed]
100.
Lianga
N.
,
Williams
E.C.
,
Kennedy
E.K.
,
Dore
C.
,
Pilon
S.
,
Girard
S.L.
et al.
(
2013
)
A Wee1 checkpoint inhibits anaphase onset
.
J. Cell Biol.
201
,
843
862
[PubMed]
101.
Matheson
C.J.
,
Backos
D.S.
and
Reigan
P.
(
2016
)
Targeting WEE1 kinase in cancer
.
Trends Pharmacol. Sci.
37
,
872
881
[PubMed]
102.
Ruiz
S.
,
Mayor-Ruiz
C.
,
Lafarga
V.
,
Murga
M.
,
Vega-Sendino
M.
,
Ortega
S.
et al.
(
2016
)
A genome-wide CRISPR screen identifies CDC25A as a determinant of sensitivity to ATR inhibitors
.
Mol. Cell.
62
,
307
313
[PubMed]
103.
Leijen
S.
,
van Geel
R.M.J.M.
,
Pavlick
A.C.
,
Tibes
R.
,
Rosen
L.
,
Razak
A.R.A.
et al.
(
2016
)
Phase I study evaluating WEE1 inhibitor AZD1775 as monotherapy and in combination with gemcitabine, cisplatin, or carboplatin in patients with advanced solid tumors
.
J. Clin. Oncol.
34
,
4371
4380
[PubMed]
104.
Moore
K.N.
,
Chambers
S.K.
,
Hamilton
E.P.
,
Chen
L.-m.
,
Oza
A.M.
,
Ghamande
S.A.
et al.
(
2019
)
Adavosertib with chemotherapy (CT) in patients (pts) with platinum-resistant ovarian cancer (PPROC): An open label, four-arm, phase II study
.
J. Clin. Oncol.
37
,
5513
105.
Lheureux
S.
,
Cabanero
M.
,
Cristea
M.C.
,
Mantia-Smaldone
G.
,
Olawaiye
A.
,
Ellard
S.
et al.
(
2019
)
A randomized double-blind placebo-controlled phase II trial comparing gemcitabine monotherapy to gemcitabine in combination with adavosertib in women with recurrent, platinum resistant epithelial ovarian cancer: a trial of the Princess Margaret, California, Chicago and Mayo Phase II Consortia
.
J. Clin. Oncol.
37
,
5518
106.
Oza
A.M.
,
Estevez-Diz
M.
,
Grischke
E.M.
,
Hall
M.
,
Marme
F.
,
Provencher
D.
et al.
(
2020
)
A biomarker-enriched, randomized phase II trial of adavosertib (AZD1775) plus paclitaxel and carboplatin for women with platinum-sensitive tp53-mutant ovarian cancer
.
Clin. Cancer Res.
26
,
4767
4776
[PubMed]
107.
Kong
A.
and
Mehanna
H.
(
2021
)
WEE1 inhibitor: clinical development
.
Curr. Oncol. Rep.
23
,
107
[PubMed]
108.
Hamilton
E.
,
Falchook
G.S.
,
Wang
J.S.
,
Fu
S.
,
Oza
A.
,
Karen
S.
et al.
(
2019
)
Abstract CT025: Phase Ib study of adavosertib in combination with olaparib in patients with refractory solid tumors: dose escalation
.
Cancer Res.
79
,
CT025-CT
[PubMed]
109.
Patel
M.R.
,
Falchook
G.S.
,
Wang
J.S.-Z.
,
Imedio
E.R.
,
Kumar
S.
,
Motlagh
P.
et al.
(
2019
)
Open-label, multicenter, phase I study to assess safety and tolerability of adavosertib plus durvalumab in patients with advanced solid tumors
.
J. Clin. Oncol.
37
,
2562
110.
Chera
B.S.
,
Sheth
S.H.
,
Patel
S.A.
,
Goldin
D.
,
Douglas
K.E.
,
Green
R.L.
et al.
(
2021
)
Phase 1 trial of adavosertib (AZD1775) in combination with concurrent radiation and cisplatin for intermediate-risk and high-risk head and neck squamous cell carcinoma
.
Cancer
127
,
4447
4454
[PubMed]
111.
Mendez
E.
,
Rodriguez
C.P.
,
Kao
M.
,
Harbison
R.A.
,
Martins
R.G.
,
Futran
N.D.
et al.
(
2017
)
A phase I clinical trial of AZD1775 in combination with neoadjuvant weekly cisplatin and docetaxel in borderline resectable head and neck squamous cell carcinoma (HNSCC)
.
J. Clin. Oncol.
35
,
6034
112.
Cuneo
K.C.
,
Morgan
M.A.
,
Sahai
V.
,
Schipper
M.J.
,
Parsels
L.A.
,
Parsels
J.D.
et al.
(
2019
)
Dose escalation trial of the Wee1 inhibitor adavosertib (AZD1775) in combination with gemcitabine and radiation for patients with locally advanced pancreatic cancer
.
J. Clin. Oncol.
37
,
2643
2650
[PubMed]
113.
Park
S.
,
Shim
J.
,
Jung
H.A.
,
Sun
J.-M.
,
Lee
S.-H.
,
Park
W.-Y.
et al.
(
2019
)
Biomarker driven phase II umbrella trial study of AZD1775, AZD2014, AZD2811 monotherapy in relapsed small cell lung cancer
.
J. Clin. Oncol.
37
,
8514
114.
Kato
H.
,
de Souza
P.
,
Kim
S.W.
,
Lickliter
J.D.
,
Naito
Y.
,
Park
K.
et al.
(
2020
)
Safety, pharmacokinetics, and clinical activity of adavosertib in combination with chemotherapy in Asian patients with advanced solid tumors: phase Ib study
.
Target Oncol.
15
,
75
84
[PubMed]
115.
Do
K.
,
Wilsker
D.
,
Ji
J.
,
Zlott
J.
,
Freshwater
T.
,
Kinders
R.J.
et al.
(
2015
)
Phase I study of single-agent AZD1775 (MK-1775), a Wee1 kinase inhibitor, in patients with refractory solid tumors
.
J. Clin. Oncol.
33
,
3409
3415
[PubMed]
116.
Francis
A.M.
,
Alexander
A.
,
Liu
Y.
,
Vijayaraghavan
S.
,
Low
K.H.
,
Yang
D.
et al.
(
2017
)
CDK4/6 inhibitors sensitize Rb-positive sarcoma cells to Wee1 kinase inhibition through reversible cell-cycle arrest
.
Mol. Cancer Ther.
16
,
1751
1764
[PubMed]
117.
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]
118.
Long
K.
,
Gu
L.
,
Li
L.
,
Zhang
Z.
,
Li
E.
,
Zhang
Y.
et al.
(
2021
)
Small-molecule inhibition of APE1 induces apoptosis, pyroptosis, and necroptosis in non-small cell lung cancer
.
Cell Death Dis.
12
,
503
[PubMed]
119.
Shahda
S.
,
Lakhani
N.J.
,
O'Neil
B.
,
Rasco
D.W.
,
Wan
J.
,
Mosley
A.L.
et al.
(
2019
)
A phase I study of the APE1 protein inhibitor APX3330 in patients with advanced solid tumors
.
J. Clin. Oncol.
37
,
3097
120.
Mesquita
K.A.
,
Ali
R.
,
Doherty
R.
,
Toss
M.S.
,
Miligy
I.
,
Alblihy
A.
et al.
(
2021
)
FEN1 blockade for platinum chemo-sensitization and synthetic lethality in epithelial ovarian cancers
.
Cancers (Basel)
13
,
1866
1884
121.
Sultana
R.
,
Abdel-Fatah
T.
,
Abbotts
R.
,
Hawkes
C.
,
Albarakati
N.
,
Seedhouse
C.
et al.
(
2013
)
Targeting XRCC1 deficiency in breast cancer for personalized therapy
.
Cancer Res.
73
,
1621
1634
[PubMed]
122.
Ali
R.
,
Al-Kawaz
A.
,
Toss
M.S.
,
Green
A.R.
,
Miligy
I.M.
,
Mesquita
K.A.
et al.
(
2018
)
Targeting PARP1 in XRCC1-deficient sporadic invasive breast cancer or preinvasive ductal carcinoma in situ induces synthetic lethality and chemoprevention
.
Cancer Res.
78
,
6818
6827
[PubMed]
123.
Ali
R.
,
Alblihy
A.
,
Toss
M.S.
,
Algethami
M.
,
Al Sunni
R.
,
Green
A.R.
et al.
(
2020
)
XRCC1 deficient triple negative breast cancers are sensitive to ATR, ATM and Wee1 inhibitor either alone or in combination with olaparib
.
Ther. Adv. Med. Oncol.
12
,
1758835920974201
[PubMed]
124.
Alblihy
A.
,
Ali
R.
,
Algethami
M.
,
Shoqafi
A.
,
Toss
M.S.
,
Brownlie
J.
et al.
(
2022
)
Targeting Mre11 overcomes platinum resistance and induces synthetic lethality in XRCC1 deficient epithelial ovarian cancers
.
Npj Precision Oncol.
6
,
51
125.
Ali
R.
,
Alblihy
A.
,
Miligy
I.M.
,
Alabdullah
M.L.
,
Alsaleem
M.
,
Toss
M.S.
et al.
(
2021
)
Molecular disruption of DNA polymerase β for platinum sensitisation and synthetic lethality in epithelial ovarian cancers
.
Oncogene
40
,
2496
2508
[PubMed]
126.
Yuhas
S.C.
,
Mishra
A.
,
DeWeese
T.L.
and
Greenberg
M.M.
(
2021
)
Suppression of DNA polymerase beta activity is synthetically lethal in BRCA1-deficient cells
.
ACS Chem. Biol.
16
,
1339
1343
[PubMed]
127.
Zatreanu
D.
,
Robinson
H.M.R.
,
Alkhatib
O.
,
Boursier
M.
,
Finch
H.
,
Geo
L.
et al.
(
2021
)
Poltheta inhibitors elicit BRCA-gene synthetic lethality and target PARP inhibitor resistance
.
Nat. Commun.
12
,
3636
[PubMed]
128.
Richard
I.A.
,
Burgess
J.T.
,
O'Byrne
K.J.
and
Bolderson
E.
(
2021
)
Beyond PARP1: the potential of other members of the poly (ADP-Ribose) polymerase family in DNA repair and cancer therapeutics
.
Front Cell Dev. Biol.
9
,
801200
[PubMed]
129.
Wang
Z.
,
Grosskurth
S.E.
,
Cheung
T.
,
Petteruti
P.
,
Zhang
J.
,
Wang
X.
et al.
(
2018
)
Pharmacological inhibition of PARP6 triggers multipolar spindle formation and elicits therapeutic effects in breast cancer
.
Cancer Res.
78
,
6691
6702
[PubMed]
130.
Gozgit
J.M.
,
Vasbinder
M.M.
,
Abo
R.P.
,
Kunii
K.
,
Kuplast-Barr
K.G.
,
Gui
B.
et al.
(
2021
)
PARP7 negatively regulates the type I interferon response in cancer cells and its inhibition triggers antitumor immunity
.
Cancer Cell.
39
,
1214e10
1226e10
131.
Cleary
J.M.
,
Aguirre
A.J.
,
Shapiro
G.I.
and
D'Andrea
A.D.
(
2020
)
Biomarker-guided development of DNA repair inhibitors
.
Mol. Cell.
78
,
1070
1085
[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).