Next Article in Journal
Molecular Background of Toxic-Substances-Induced Morphological Alterations in the Umbilical Cord Vessels and Fetal Red Blood Cells
Next Article in Special Issue
Clinical Impact of Next-Generation Sequencing Multi-Gene Panel Highlighting the Landscape of Germline Alterations in Ovarian Cancer Patients
Previous Article in Journal
Mitochondrial Dysfunction and Induction of Apoptosis in Hepatocellular Carcinoma and Cholangiocarcinoma Cell Lines by Thymoquinone
Previous Article in Special Issue
DNA Damage Response (DDR) and DNA Repair
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

DNA Damage Response in Cancer Therapy and Resistance: Challenges and Opportunities

by
Dana Jurkovicova
1,
Christiana M. Neophytou
2,
Ana Čipak Gašparović
3 and
Ana Cristina Gonçalves
4,5,6,*
1
Department of Genetics, Cancer Research Institute, Biomedical Research Center, v.v.i. of the Slovak Academy of Sciences, 845 05 Bratislava, Slovakia
2
Department of Life Sciences, European University Cyprus, Nicosia 2404, Cyprus
3
Division of Molecular Medicine, Ruđer Bošković Institute, HR-10000 Zagreb, Croatia
4
Laboratory of Oncobiology and Hematology (LOH), University Clinic of Hematology, Faculty of Medicine University of Coimbra (FMUC), University of Coimbra, 3000-548 Coimbra, Portugal
5
Group of Environment Genetics and Oncobiology (CIMAGO), Coimbra Institute for Clinical and Biomedical Research (iCBR), Faculty of Medicine University of Coimbra (FMUC), University of Coimbra, 3000-548 Coimbra, Portugal
6
Center for Innovative Biomedicine and Biotechnology (CIBB), 3004-504 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(23), 14672; https://doi.org/10.3390/ijms232314672
Submission received: 17 October 2022 / Revised: 20 November 2022 / Accepted: 21 November 2022 / Published: 24 November 2022
(This article belongs to the Special Issue DNA Damage Response (DDR) and DNA Repair 2.0)

Abstract

:
Resistance to chemo- and radiotherapy is a common event among cancer patients and a reason why new cancer therapies and therapeutic strategies need to be in continuous investigation and development. DNA damage response (DDR) comprises several pathways that eliminate DNA damage to maintain genomic stability and integrity, but different types of cancers are associated with DDR machinery defects. Many improvements have been made in recent years, providing several drugs and therapeutic strategies for cancer patients, including those targeting the DDR pathways. Currently, poly (ADP-ribose) polymerase inhibitors (PARP inhibitors) are the DDR inhibitors (DDRi) approved for several cancers, including breast, ovarian, pancreatic, and prostate cancer. However, PARPi resistance is a growing issue in clinical settings that increases disease relapse and aggravate patients’ prognosis. Additionally, resistance to other DDRi is also being found and investigated. The resistance mechanisms to DDRi include reversion mutations, epigenetic modification, stabilization of the replication fork, and increased drug efflux. This review highlights the DDR pathways in cancer therapy, its role in the resistance to conventional treatments, and its exploitation for anticancer treatment. Biomarkers of treatment response, combination strategies with other anticancer agents, resistance mechanisms, and liabilities of treatment with DDR inhibitors are also discussed.

1. Introduction

DNA and other biological molecules are susceptible to damage, but DNA damage can have much more complex consequences due to its function. Unlike other molecules, which are synthesized and degraded depending on their necessity, DNA is constantly present and replicating when the cell is in division. Therefore, there is a great need for a reparation system to maintain DNA integrity. Estimations suggest that every day, about 105 lesions occur in the cell [1]. In addition to exogenous threats, such as irradiation, chemical pollutants, and chemical agents, endogenous processes increase reactive oxygen species (ROS), damaging DNA directly or indirectly. DNA damage includes single- and double-strand brakes, inter-and intra-strand links, abasic sites, bulky adducts, and base changes, such as 8-deoxyguanosine [1]. Due to the diversity of DNA damage, there are different repair mechanisms implicating many proteins. The activation of different repair mechanisms with the primary goal to restore DNA integrity is collectively known as the DNA damage response (DDR). These proteins/genes were historically identified in Fanconi’s anemia (FA), a rare genetic disorder characterized by bone marrow failure, skeletal malformation, and increased cancer incidence. Mutations in this rare disease include genes in the FA pathway that are fundamental genes involved in DNA damage repair [2]. If DNA damage is not repaired or misrepaired, genomic instability and mutations will be established, which are among the hallmarks of cancer [3]. The DDR plays a relevant role not only in cancer development but also in cancer treatment. Defects in DDR genes are known as cancer drivers, and cells with deficient DDR show a higher sensitivity to DNA-damaging agents [4]. In this review, we summarize recent evidence of DDR pathways in cancer therapy, its role in the resistance to conventional treatments, and its exploitation for anticancer treatment. Since DDR is involved in cancer development and is a molecular target of cancer treatment, biomarkers of treatment response, combination strategies with other anticancer agents, resistance mechanisms, and liabilities of treatment with DDR inhibitors are also discussed.

2. The DNA Damage Response

The DNA damage response pathways are composed of an intricate system of sensors, transductors, and effectors involved in DNA repair and cell cycle checkpoint control that manage the execution of DNA replication and cell proliferation. The wide variety of DNA lesion types requires multiple and different DNA repair mechanisms. To repair single-strand breaks (SSBs), the mismatch repair (MMR), base excision repair (BER), and nucleotide excision repair (NER) are activated, while homologous recombination (HR) and non-homologous end joining (NHEJ) pathways repair the double-strand breaks (DSBs) [5]. DDR also includes damage tolerance processes and the consequent signaling control of cell decisions on senescence or death. In parallel, DDR affects epigenetics and gene expression regulation preferentially related to the induction of apoptosis [6,7].

2.1. Single-Strand Break Repair

The need for MMR occurs during DNA replication, where polymerases are prone to mistakes. Therefore, a mismatch repair system repairs wrongly matched bases to obtain replicative fidelity. The mismatch repair system in humans includes several proteins: DNA mismatch repair protein Mlh1 (MLH1), DNA mismatch repair protein Msh2 (MSH2), DNA mismatch repair protein Mlh3 (MLH3), DNA mismatch repair protein Msh6 (MSH6), DNA mismatch repair endonuclease PMS2 (PMS2), and DNA mismatch repair protein Msh3 (MSH3) [8,9]. These proteins act as heterodimers MSH2 with MSH6 or MSH3 (MutSα or MutSβ complexes, respectively) and MLH1 with PMS2 or MLH3 (MutLα, MutLβ, or MutLγ complexes, respectively) [9]. In addition to these protein complexes, proliferating cell nuclear antigen (PCNA) with the help of MSH2/MSH3 and MSH2/MSH6 complexes recognizes and binds to the mispaired region [10]. Furthermore, single-strand DNA-binding protein RPA (replication protein A) and EXOI (exonuclease I) both contribute to MMR by protecting the gap from excision, while EXOI is needed for the repair of the break located either 5′ or 3′ to the mispair [11]. The evidence strongly suggests that polymerase δ (pol δ) is required for MMR [12], and DNA ligase I is needed for the final step in MMR [13].
NER is the mechanism of bulky adduct reparation. The offset of this mechanism can be initiated by global genome NER or transcription-coupled NER [14]. Like other repair systems, there are two steps involved: recognition of the damage and the reparation step. Bulky adducts cause DNA distortions recognized by XPC-RAD23B, but only if the nucleotide opposing the lesion is not missing [15]. Once XPC-RAD23B recognizes the bulky adduct, a small bubble is formed, and TFIIH, a complex of 10 proteins, is recruited [16]. Then, XPB translocase and XPD helicase open the bubble even more (22 to 25 base pairs) [17], which allows binding of XPA and RPA. RPA protects the undamaged strand while XPA verifies the damage and further recruits XPF-ERCC1 endonuclease that cleaves the 5′ end, while XPF cleaves the 3′ end [18]. The resulting gap is filled by DNA polymerase δ and ε, replication factor C (RFC), PCNA, and RPA [19].
Like NER, BER can be activated during transcription and initiated by global genome BER. The damaged base is recognized and removed by DNA glycosylase [20]. Several DNA glycosylases recognize damaged bases. Monofunctional DNA glycosylases, such as alkyl adenine DNA glycosylase or uracil DNA glycosylase, create abasic sites, while bifunctional DNA glycosylases, such as 8-oxoguanine DNA glycosylase (OGG1), NEIL1−3, additionally cleave the 3′ site on the abasic sugar [21,22]. The excision of the damaged base initiates the repair process, where the excision site is recognized by apurinic/apyrimidinic endonuclease (APE) [23]. If monofunctional DNA glycosylase acts on the damaged nucleotide, APE1 cleaves the site, while after bifunctional glycosylase, APE2 is the endonuclease responsible for the cleavage [23]. Replacement of the damaged/missing nucleotide is filled by the action of polymerase β or, if the gap is bigger than one nucleotide, with polymerase δ or ε. The ends are ligated by DNA ligase I (LIG1) or DNA ligase III (LIG3) [24]. Alternative pathways include poly(ADP-ribose) polymerase 1 (PARP1), which acts as a sensor for the damaged nucleotide [25]. PARP1, a protein located in the nucleus in high abundance, takes part in the BER system [26]. Following SSB or DSB DNA damage, PARP1 swiftly localizes to the damaged site, and its enzymatic activity is increased 10- to 500-fold. This activity leads to the synthesis of poly-ADP-ribose (PAR) chains after damage within 15 to 30 s [26]. PARP1 transfers 50–200 residues of PAR to itself and its substrates, including enzymes such as DNA polymerases, topoisomerases, and DNA ligase-2, as well as histones, high-mobility-group proteins, and transcription factors [27]. The modification of these proteins by poly (ADP-ribosyl)ation (PARylation) allows PARP1 to control not only cellular repair, DNA replication, and transcription but also protein degradation, organization of the cytoskeleton, and other cellular functions [27]. The extent of PARP1 activation after DNA damage controls whether cells will live or die. Caspase-3 and -7 mediate cleavage of PARP1 into a ∼25-kDa N-terminal and a ∼89-kDa C-terminal fragment, which are among the hallmarks of apoptosis. Cleaved PARP1 cannot participate in DNA repair during apoptosis and allows cells to commit to the apoptotic pathway [28,29]. Nevertheless, PARP1, if excessively engaged, induces cytotoxicity. Therefore, the action of DNA repair protein XRCC1 (XRCC1) is needed since XRCC1 binds PARP1 and DNA ligase II, forming a complex, which controls the activity of PARP1 and prevents its toxicity [30]. ROS and the dysregulation in the activity of DNA topoisomerase 1 (TOP1) create SSBs, resulting in abortive TOP1–DNA complexes, which are removed by tyrosyl-DNA phosphodiesterase 1 (TDP1), a target of PARP1 [31]. TDP1 thereby enhances the recruitment of other proteins involved in the repair [32]. Furthermore, TDP1 and PARP1 recruit XRCC1, another substrate for PARP1 [31]. PARylation of XRCC1 recruits polymerase β and DNA ligase III, which finalize the repair [31].

2.2. Double-Strand Breaks Repair

Double-strand breaks are repaired by at least five different mechanisms: canonical non-homologous end joining (cNHEJ), HR, alternative non-homologous end joining (Alt-NHEJ), single-strand annealing (SSA), and break-induced replication (BIR) [33]. The first step is the recognition of the break by scaffolding proteins 53BP1 and BRCA1, after which the break is repaired by one of the mentioned mechanisms [33]. In addition, PARP1 can also detect double-strand breaks [34]. However, the repair pathway choice is regulated by resectosome, a protein complex responsible for DNA end resection [35]. Resectosome comprises a helicase, a nuclease, and other regulating proteins, while the length of the resected DNA determines the pathway [36]. Canonical NHEJ joins two ends of the break without homology check [36], while HR uses the sister chromatid as a template for repair [37], thereby dictating the part of the cell cycle where each of these two mechanisms can function and limiting the HR to S/G2 transition. Due to the template, HR is mostly error-free in its outcome as the newly synthesized chromatid or the homologous chromosome is used as a template for DNA repair [38,39]. Still, in some cases, it can cause genetic instability and rearrangements [37].
Both c-NHEJ and HR repair start with binding Ku70-Ku80 heterodimer to a double-strand break [40]. The binding of Ku70-Ku80 recruits other factors: DNA-dependent protein kinase catalytic subunit (DNA-PKcs), DNA ligase IV (LIG4), and the associated scaffolding factors of DNA repair protein XRCC4 (XRCC4), XRCC4-like factor (XLF), and paralogue of XRCC4 and XLF (PAXX), which bring the two ends closer, enabling end processing by Artemis and DNA polymerases λ and μ [37]. The HR mechanism is usually associated with cancer and BRCA1 and BRCA2 mutations, which are linked to hereditary breast and ovarian cancer [41]. However, new evidence suggests that the basis of hypersensitivity of BRCA-deficient tumors is not double-strand breaks induced by chemotherapy but rather single-strand breaks [42]. The HR process includes numerous steps, the first one being recognition by two kinases, ATM and ATR, which phosphorylate targets: CHEK2, P53, BRCA1, and H2AX. BRCA1 serves as a scaffold that recruits other proteins [43]. Recognition is followed by DNA resection with MRN complex together with EXO1 and the recQ-like DNA helicase (BLM) heterodimer [43]. The endonuclease activity causes displacement of Ku70-Ku80, and binding of RPA occurs to the single-stranded-DNA [37]. Next, BRCA2 working with PALB2 loads RAD51 to the single-stranded DNA and mediates strand exchange using the sister chromatid as a template [37,43,44].

2.3. Epigenetic Control in DNA Damage Response

To properly understand how cells control complex DDR, epigenetics and miRNA regulations cannot be omitted. Epigenomic alterations are known to significantly affect gene expression and overall tumor heterogeneity. Therefore, it is not surprising that DNA repair processes are also affected by epigenetic chromatin regulation. Histone deacetylases (HDACs) are important players in chromatin preparation to promote DSBs repair through HR and NHEJ. For example, PARP1 recruits the nucleosome remodeling deacetylation (NuRD) complex by attaching a PAR chain signal essential for DSB repair [45]. In fact, PARylation inhibition stops chromatin relaxation at DNA damage sites, suggesting that chromatin relaxation is PARylation-dependent [46]. On the other hand, HDAC1 and HDAC2 deacetylases stimulate the RNF8/RNF168-dependent ubiquitination at DSB promoting NHEJ repair, while histones H4 and H2 acetylation/deacetylation, at specific sites, switch DNA repair from NHEJ to HR via 53BP1 binding regulation at the DSB site [45,47]. DNA methylation is a common and stable epigenetic mechanism of gene inactivation, and in cancer cells, DDR components also show changes in their gene promoter methylation status [48]. For example, hypermethylation of OGG1 genes was observed in thyroid cancer [49], MLH1 gene in oral squamous cell carcinoma [50], neck squamous cell carcinoma [51], non-small cell lung cancer (NSCLC) [52], acute myeloid leukemia (AML) [53], gastric cancer [54], ovarian cancer [55], and BRCA1 gene in breast cancer [56], bladder cancer [57], NSCLC [58], and gastric cancer [59]. Additionally, the methylation status of some DDR genes has been used as diagnostic, prognostic, and therapy response biomarkers in various cancer types. MLH1 methylation has been indicated not only as a diagnostic biomarker and an indicator of good prognosis in several cancers, including colorectal, ovarian, and breast cancers, but also as a therapy response biomarker associated with platinum compounds, temozolomide, and epirubicin resistance and with methotrexate sensitivity [60].
MicroRNAs (miRNAs) are other plausible players important for DDR modulation. MiRNAs regulate multiple processes of tumorigenesis, post-transcriptionally controlling expression of components of DNA damage repair and other mechanisms defining response to treatment and overall outcome and survival of cancer patients. As regulatory elements, miRNAs can modulate cancer cell sensitivity toward DNA-damaging agents by regulating the expression level of DNA repair genes. Therefore, miRNAs represent promising therapeutical tools modifying treatment response, mainly in highly resistant cancers, such as breast cancer [61,62,63,64,65]. In DDR, miRNAs play a significant regulatory role as transcriptional and post-transcriptional regulators of DNA damage sensors, signal transducers, and effector genes. For example, miRNAs can directly target genes involved in cell cycle regulation, e.g., miR-34 targeting cyclins, miR-93 targeting E2F1 and CCND1 [66,67], or miR-125b and miR-34a controlling expression of TP53 [61]. DNA repair checkpoints are targets of, for example, miR-15, miR-195, and CHK1 and WEE1 are targets of miR-497 [68,69,70]. MiR-191 was shown to target CHK2 in osteosarcoma cells [71]; miR-124 targeting PARP1 [72]; miR-181a/b targeting ATM [73]; miR-182 and miR-218 targeting BRCA1 [74,75]; and miR-155, miR-103/miR-107, and miR-221/222 targeting DDR gene RAD51 [61,76]. MiR-494 and miR-99b were upregulated after γ-irradiation and directly inhibited a protein complex consisting of MRE11, RAD50, and NBS1 (MRN complex) crucial for DSBs repair in human endothelial cells [77]. In aggressive triple-negative breast cancer (TNBC), miR-155 [78] and miR-21 [79] act as typical oncomiRs, and miR-205 [80,81], miR-200c/miR-141 [82], let-7 [83], and miR-221/222 [84] were shown to be typical tumor suppressor miRNAs. MiR-19a-3p, miR-218-5p, and miR-874-3p were shown to directly target RAD51 and BRCA2 in HR and XRCC5 (KU80) and PRKDC of the NHEJ pathway, affecting recombination repair. XRCC5 can also be targeted by miR-526b and miR-623 to induce apoptosis when overexpressed in NSCLC and breast cancer cells [85,86]. Today, there is no doubt that miRNAs importantly regulate the expression of components of DNA repair pathways. The exploitation of DDR gene/miRNA interactions and the possibility of their easy inhibition with antagomiRs or reintroduction using miRNA mimics open a novel field for clinical utilization in terms of new potential biomarkers and new therapeutic tools.

3. DNA Damage Response Inhibitors in Cancer Therapy

The DDR mechanisms are involved in the control of core processes of cell fate, survival, and genome maintenance. In cancer, accumulated genetic defects compromise the cell response to physiological growth control and promote uncontrolled division and evasion of apoptosis. Vogelstein and collaborators identified approximately 130,000 different mutations in more than 3000 individual drivers of tumorigenesis, including both oncogenes and tumor suppressors. From these, about 330 genes were identified as drivers involved in the regulation of cell survival, genome maintenance, and overall DDR. These genes are valuable targets for new approaches to cancer treatment [87].
The effect of primary anticancer therapies, including ionizing radiation and different chemotherapeutic agents that damage both nuclear (nDNA) and mitochondrial (mtDNA) DNA, are expected to drive the cancer cell, directly or indirectly, towards death. Cisplatin represents such a highly efficient DNA-damaging agent with substantial anticancer effects. Despite the discovery of its cytotoxic effects and its first Food and Drug Administration (FDA) approval for treating testicular cancer in 1978, cisplatin is still used as first-line chemotherapy in numerous solid tumors. The cell response to cisplatin is complex and includes mechanisms regulating its entry, exit, accumulation, and detoxification and mechanisms modulating DNA repair, cell survival, and the tumor microenvironment [88].

3.1. DNA Damage Response Inhibitors in Single-Agent Approaches

Cancer cells show a higher level of endogenous DNA damage and increased replication stress than normal cells and usually have one or more DDR pathways disabled. Such deficiencies are attractive points for novel cancer treatments development, mostly those exploiting synthetic lethality concepts [89,90]. In principle, synthetic lethality in cancer treatment is based on targeting and inhibiting the DDR pathway that is left functional. If one DDR pathway is compromised and not functional, e.g., due to mutations in one or more DNA repair genes, the cancer cell attempts to restore the DNA damage utilizing the backup repair mechanism. However, if this backup mechanism is pharmacologically targeted, the cancer cell has no functional DNA repair pathway and is doomed. In 2014, in both Europe and the USA, Olaparib—a PARP inhibitor (PARPi)—was the first DDR inhibitor (DDRi) approved for cancer treatment [89]. Shortly after, in 2017, two other PARP inhibitors, Rucaparib and Niraparib, were FDA-approved for use in cancer patients with BRCA mutation as well as for non-carriers of BRCA mutations to treat primary peritoneal cancer, fallopian tube, or recurrent epithelial ovarian cancer resistant to cisplatin chemotherapy [91,92]. Later, Talazoparib was approved by the FDA to treat patients with germline BRCA (gBRCA) mutations or metastatic breast cancer [93,94]. In 2020, the FDA approved Veliparib for use in combination with gamma-ray radiotherapy or chemotherapy for advanced lung squamous cell carcinoma [95] and in combination with paclitaxel/carboplatin for the treatment of recurrent ovarian, breast, and lung cancer patients [96,97,98]. Veliparib and another PARPi Iniparib in combination with gemcitabine/carboplatin for breast and lung cancer treatment reached phase III trials [99].
In clinical practice, the conventional treatment of women with OC is based on debulking surgery and selecting first-line platinum-based chemotherapy, followed by second-line platinum chemotherapy in case of relapse. Further management of OC patients diverges with the decision for maintenance treatment or active surveillance watching until the third relapse [100]. If OC patients harbor BRCA mutation, targeted therapy with PARPi may replace chemotherapy to maintain the response, delay disease progression, and prolong the period between treatment cycles [91,101]. Recently, Olaparib has been approved for maintenance treatment in the first-line setting for women with a BRCA mutation [102]. Similarly, rucaparib was approved in the treatment setting for patients with relapsed BRCA-mutated platinum-sensitive OC [103] and has been shown to be beneficial in both the maintenance as well as treatment settings. The substantial benefit of PARPis in the first-line setting has been demonstrated in randomized phase III trials (SOLO-1, PAOLA-1, PRIMA, VELIA) [104,105,106,107]. However, besides the significant contribution of PARPis to better treatment management of OC patients, the side effects of PARPis on patients’ quality of life must be thoroughly monitored.
PARP is the best-known element of the DDR. Functional PARP identifies single-strand breaks and utilizes nicotinamide adenine dinucleotide (NAD+) to form poly ADP-ribose chains that open chromatin to allow DNA repair proteins to access the DNA [108]. PARPi prevents the formation of PAR chains and keeps PARP on the DNA at SSBs. Consequently, formed PARP–DNA complexes stall or collapse the replication fork and generate DSBs. DSBs can be repaired by HR, but if HR is missing or defective in cancer cells, as in BRCA1 mutation-carrying cancers, the cell must use error-prone NHEJ, leading to genomic instability and cancer cell death [108,109,110]. PARPs are involved in the repair of SSBs through BER and DSBs, through HR, NHEJ, and alt-NHEJ (or microhomology-mediated end joining, MMEJ). Along with successful validation on patients carrying BRCA1 and BRCA2 mutations [111], positive effects of PARPi were also observed in patients without BRCA mutations with high-grade serous or poorly differentiated ovarian carcinoma or TNBC [112]. Accordingly, PARP inhibition as a therapeutic approach successfully expanded to other cancers, including pancreatic, endometrial, prostate, urothelial, colorectal, lung, and glioblastoma [113]. Overall, DDR involves more than 450 proteins [89,114,115], and several are being investigated as potential novel therapeutic targets. The most promising include DNA damage sensors (MLH1), damage signaling molecules (ataxia telangiectasia mutated (ATM), ATM- and RAD3-related (ATR), CHK1, CHK2, DNA-dependent protein kinase, catalytic subunit (DNA-PKcs), and WEE1), or effector proteins for DNA repair (POLQ, RAD51, or PARG) [113]. Several DDRi are currently in preclinical and clinical trials (Table 1). Clinical trials have been initiated to test their targeting with single agents or in combination therapy. In parallel, different technologies are being explored to screen for synthetic lethal combinations, including small interfering RNA (siRNA) or exploiting CRISPR-Cas9-based strategies to target them for anticancer treatment purposes.

3.2. DNA Damage Response Inhibitors in Combinatory Therapies

To gain maximum efficiency of anticancer treatments, combinations of individual agents and strategies are tested and validated. Concerning the interplay between epigenetics and DNA repair, the most explored therapeutic approach is the combination of epigenetic inhibitors with chemotherapeutic agents. Additionally, several clinical trials have been initiated to evaluate the efficacy of combined administration of HDAC and PARP inhibitors. Such combinations have been examined in pre-clinical models and effectively kill prostate, ovarian, and breast cancer cells [116,117,118,119]. Similarly, combined administration of DNMT1 and PARP in AML and breast cancer showed a synergistic activity [120]. Chromatin remodeling inhibitors also prevent HR repair and sensitize different cancers cells towards DNA-damaging agents [121], while HDAC, DNMT, and LSD1 inhibitors restore chemosensitivity in different solid tumors [122]. A synergism of combinations has been identified between radiotherapy and epigenetic inhibitors [123,124,125,126,127]. However, due to high toxicity and limited patient benefits, these approaches still require more investigation and clinical validation. In recent years, much attention has been paid to investigating the therapeutical effect of combinations of PARP inhibitors with antiangiogenic therapy. Multi-kinase inhibitors targeting VEGFR, PDGFR, and FGFR were shown to sensitize tumor cells to PARP inhibitors via induction of hypoxia and triggering HR defects [128].
Combination regimens between two DDRi have also been investigated. The synergic effect of PARPi in combination with ATR inhibitor (ATRi) was reported in HR-deficient HGSOC in vivo [129]. The therapeutic combinatory approach of non-toxic concentrations of a CHK1 inhibitor (CHK1i; PF-00477736) with a WEE1 inhibitor (WEE1i; MK-1775) showed a synergistic effect in breast, ovarian, colon, and prostate cancer cell lines in a P53 status-independent manner [130]. The combinations of the CHK1is (PF-00477736 or AZD-7762) with the WEE1i (AZD-1775) showed a synergistic effect in all the diffuse large B-cell lymphoma cell lines independently of the molecular subtype and MYC status [131]. The combination of CHK1i with WEE1i also showed a strong synergy in mantle cell lymphoma [132], lung, prostate, and erythroleukemia [133]. The synergistic effect of DDRi in B-cell lymphomas was also observed in combinations of ATRi with CHK1i and ATRi with WEE1i [134]. Moreover, the combination of ATR (VE-821) and CHK1 inhibitors (AZD7762) induced replication fork arrest, ssDNA accumulation, replication collapse, and synergistic cell death in osteosarcoma, breast, and lung cancer cells in vitro and in vivo [135]. The co-administration of Olaparib and AZD1775 (WEE1 inhibitor) demonstrated a synergistic antiproliferative effect in TNBC cell lines and significantly inhibited tumor growth in a xenograft model of BC [136]. Preclinical studies showed that ATR inhibitor synergizes with WEE1i in TNBC [18,137]. This therapeutic association reduced cell proliferation and induced cell death in several BC cell lines [137,138] and tumor remission, increased survival, and inhibited metastasis in orthotopic BC xenografts mouse models [138]. Bukhari et al. also demonstrated the therapeutic potential of this association in mammospheres, reporting similar sensitivities to the combined treatment in cancer stem cells [138]. Kim and colleagues demonstrated, using acquired and de novo PARPi and platinum-resistant models, that PARPi (AZD2281) in combination with ATRi (AZD6738) synergistically decreased cell viability and colony formation using doses with minimal off-target effects [139]. This combination also induced tumor regression and a significant increase in overall survival in HGSOCs patient-derived xenograft (PDX) models.
The metabolic vulnerability of cancer cells is a highly relevant dimension that can be exploited for therapeutical targeting and potential overcoming therapy resistance. Mammalian target of rapamycin (mTOR) kinase, mTORC1, and mTORC2 complexes are considered critical drivers of cancer drug resistance that integrate signaling pathways driving cell metabolism and growth [140,141]. Both mTOR complexes belong to effectors of the most oncogenic drivers, including RAS-driven MAPK and PI3K-AKT pathways. Sustained mTOR signaling contributes to resistance to therapeutics targeted against the driving oncogenes [142] or chemotherapy resistance, for example, by inducing the FA DNA repair pathway [143] and modulating other proteins that are essential in chromosomal integrity and DNA damage response [144,145]. Deregulation of mTOR has been found in various human cancers [146], including resistant ones such as TNBCs [147,148]. Inhibitors of mTOR are therefore considered a valuable addition to chemotherapy or targeted cancer therapy, either as an option for relapsed patients or as a frontline combination therapy to prevent or delay the development of resistance due to sustained mTOR signaling [142,149].
Similarly, using DDR inhibitors and/or radiation as sensitizers provide new potential to increase immunotherapy efficacy. Immunotherapy attracts much attention and is considered a breakthrough in the field of cancer treatment. Individual DNA repair pathways’ defects were associated with immune checkpoint blockade response. DNA damage induced in cancer cells upon radiation or chemotherapy leads to the release of chromosome fragments or small pieces of DNA that activate an immune response. When DDRi (e.g., PAPRi) are used, more DNA fragments are released, making tumor cells more immunogenic and more sensitive to immunotherapy. For example, defects in MMR result in neoantigen generation [150] that is associated with better anti-PD-1/PD-L1 immunotherapy outcomes [151]. The benefits of multiple combination therapies involving immune checkpoint inhibitors with DDRi are undergoing clinical trials.
A novel, although limited, field of anticancer approaches is opened via targeting mtDNA repair pathways. The capacity of mtDNA repair significantly contributes to therapeutical cancer cell response. BER is the main repair pathway used by mitochondria to repair mainly ROS-induced lesions. mtDNA carries many mutations that usually correlate with cancer progression [152]. Defective mtDNA repair pathways or downregulated mtDNA repair-associated proteins, such as mitochondrial transcription factor A (mtTFA) and POLγ, together with administered DDR inhibitors, can result in higher sensitivity of cancer cells to radio- or chemotherapy [153,154].

4. Biomarkers of DNA Damage Response Inhibitors

As mentioned previously, a significant number of new DDR inhibitors and DDR-based therapeutic strategies have arisen recently. The remarkable clinical success of PARP inhibitors in patients with BRCA1 and BRCA2 gene mutations showed that the clinical utility of DDRi relies on establishing response biomarkers that select patients who will benefit from these therapies. PARPi has shown efficacy in HR-deficient cancers, including those with RAD51C, RAD51D, and PALB2 mutations [155] and with a “BRCAness” phenotype [156], but not all HR alterations have the same impact on the efficacy of these inhibitors [157,158]. The “BRCAness” phenotype is defined by the lack of BRCA1/2 mutations in tumors with similar molecular phenotypes. This phenotype can result from mutations and epigenetic modifications of HR-related genes that cause homologous recombination deficiency (HRD), such as RAD51C, RAD51D, ATM, BARD1, PALB2, BRIP1, and MRE11 mutations and BRCA1 hypermethylation [155,156,159,160,161,162]. The sensitivity to platinum-based chemotherapy is considered a surrogate biomarker of “BRCAness” phenotype to PARPi, and FDA approved this biomarker as a response biomarker for Olaparib therapy in maintenance settings [163]. However, not all patients who respond to platinum-based therapy will respond to PARPi, and some patients resistant to these conventional therapies will respond to PARPi [164]. Moreover, gene alterations, mutations, or functional loss of proteins involved in DDR mechanisms result in defective ATR, CHEK1, CHEK2, DSS1, MRE11A/NBS1, Fanconi anemia complementation group (FANC family of genes), EMSY, XRCC2, XRCC3, or PTEN, predispose patients to the success of PARP inhibitors for cancer treatment [165,166,167,168].
Throughout the years, several HRD assays have been developed to try to identify patients who will benefit from DDRi. These tests include the mutational status of DDR genes that identify specific causes of HRD, “genomic scars” or mutational signatures that identify HRD cancers, and functional assays that provide a readout of HRD or homologous recombination proficiency [169]. HRD cancers are expected to have genomic instability, and patients with these features are identified as “BRCAness”. For example, tumors with BRCA1/2 mutations were associated with loss of heterozygosity (LOH), large genomic deletions, large-scale transitions (LST), and telomeric allelic imbalance (TAI) [170,171,172,173,174], while microsatellite instability (MSI) is characteristic of MMR deficiency [175]. A combination of these genomic scars, LOH, LST, and TAI, robustly predicted the “BRCAness” phenotype and the sensibility of PARPi and was the basis of myChoise HDR (Myriad Genetics) and FoundationOne CDx (Foundation Medicine) commercial assays [158,176]. Additionally, a genomic mutational signature, “signature 3”, was significantly associated with BRCA1/2 mutations [177]. The reduced nuclear RAD51 foci have been associated with BRCA1/2 mutations and with PARPi responses. However, currently, the functional assays of HRD, as an estimation of nuclear RAD51 amount being the most used system, have insufficient evidence to establish their clinical value to predict PARPi response [169]. Several other genomic alterations have been proposed as potential biomarkers of DDRi response. For example, decreased CHK1 phosphorylation, increased expression of γH2AX, and increased replication fork instability are associated with ATR inhibitors response, while P53 deficiency and replication stress promoting genomic charges, including CCNE1 and MYC amplification, are associated with WEE1 inhibitors response [158]. KRAS mutations and the overexpression of CCNE1, CCND2, and MYC genes induce hypersensitivity to ATR inhibitors in cancer cell lines [178,179,180]. Moreover, tumor cells with loss of H3K36me3 due to mutation in SETD2, the gene that encodes a histone lysine methyltransferase, or mutation in histone H3 itself showed HR, NHEJ, and MMR impairment and sensitivity to WEE1, CHK, or ATR inhibitors [181]. These tumors were also sensitive to PARP and ATR inhibitors [158].
According to the European Society for Medical Oncology (ESMO) Translational Research and Precision Medicine Working Group, the most useful predictive biomarkers for HRD and indicate the PARPi benefit in the clinic are single-gene aberrations and/or genomic scars [169,182]. These tests reflect HRD phenotype and aim to identify patients who may benefit from PARPi. The germline and tumor (incorporating germline and somatic) BRCA mutation testing exhibit adequate clinical validity by consistently identifying the subgroup of OC patients who benefit more from PARPi therapy and remain the gold-standard predictive biomarker for PARPi. Additionally, HRD tests using genomic scars incorporating scores of allelic imbalances (GIS or LOH) are also reasonable since this test identifies a subgroup of BRCA wild-type, platinum-sensitive cancers that will benefit from PARPi therapy in some settings [169]. However, HRD biomarkers able to evaluate cancer evolution and provide a real-time read-out of homologous recombination proficiency still need to be developed and optimized [169,183], and further studies are required to clinically validate the existing ones.
Liquid biopsy approaches may also facilitate the selection of therapy and predict chemoresistance in cancer patients by identifying mutations in genes implicated in DNA repair mechanisms. In HER+ breast cancer patients, circulating tumor DNA (ctDNA) profiling identified ERBB2, TP53, EGFR, NF1, and SETD2 mutations contributing to trastuzumab resistance; in the same retrospective study, genetic aberrations in TP53, PIK3CA, and DNA damage repair genes were found in HER2-negative BC patients resistant to chemotherapy [184]. The components of the DNA repair machinery may also be utilized as biomarkers for evaluating tumor mutational burden (TMB). In a retrospective study of NSCLC, next-generation sequencing (NGS) was applied to different specimens, including small biopsy and cytology specimens; genomic alterations were found in genes implicated in DNA repair, including TP53 and BRCA2 [185]. Liquid biopsies have also been used to identify BRCA1/2 reversions in several pathologies [186]. One example is the identification of BRCA1/2 reversions in circulating free DNA (cfDNA) in HGOC patients treated with rucaparib [186]. In this study, 18% of platinum-resistant and 13% of platinum-refractory patients had BRCA1/2 reversions pretreatment in comparison with 2% of platinum-sensitive patients. This study supports the potential clinical use of liquid biopsies prior to initiating PARPi therapy since this test allowed the identification of patients who benefit from rucaparib therapy (patients without pretreatment cfDNA BRCA1/2 reversions had 2-fold higher PFS on rucaparib; 9.0 versus 1.8 months; p < 0.001) [187]. Another advantage of liquid biopsies is the detection of clonal heterogeneity of reversion events [186]. Lin et al. described the detection of eight BRCA1 mutation reversions in cfDNA, but only one of them was detected in the tumor biopsy [187]. The development of liquid biopsy approaches to detect specific aberrations in genes involved in DDR would provide a non-invasive and efficient means to improve treatment selection and disease outcome.

5. DNA Damage Response as a Mechanism of Cancer Therapy Resistance

Chemotherapy and radiotherapy rely on the cytotoxic DNA-damaging effects that for proliferating cancer cells, already burdened by genomic instability and defective DNA repair pathways, represent an induction of unrepairable genome-wide DNA damage leading to apoptosis. Therefore, chemotherapy and radiotherapy are still used as a first-line approach for many unresectable or metastatic malignancies. However, due to the large capacity of cancer cells to resist anticancer agents and adapt, the DDR is dysregulated and can lead cancer cells to genotoxic hypersensitivity or resistance development. Defective DDR allows for tumor heterogeneity development by preselecting subclones with intrinsic or acquired resistance, driving cancer progression and tumor relapse [188,189,190]. For example, in the case of cisplatin, despite its consistent rate of initial responses in multiple solid tumors [191], the treatment often results in the development of chemoresistance and therapeutic failure. Platinum salts such as cisplatin cause DNA inter- and intrastrand crosslinks, with DNA lesions repaired by a combination of NER and HR pathways. Higher expression and activity of DNA damage repair enzymes are observed in cisplatin-resistant tumor cells, and NER inhibition enhanced their sensitivity to cisplatin [192,193]. High-grade serous ovarian cancers with germline or somatic mutations in BRCA1 or BRCA2 genes and hypermethylation of the BRCA1 gene and NSCLC without ERCC1 are sensitive to platinum compounds [194,195].
Temozolomide (TMZ) is the standard treatment in glioblastoma that acts mainly through O6-methylguanine (O6-meG) lesions. Other lesions caused by TMZ, such as N3-meA and N7-meG DNA adducts, are easily repaired by BER enzymes such as O-6-methylguanine-DNA methyltransferase (MGMT) [196]. The MGMT gene promoter hypermethylation has been associated with longer survival in glioblastoma patients treated with TMZ [197,198]. The MGMT enzyme removes alkyl groups from guanine at the O6 position, reducing the effect of TMZ and suggesting that MGMT activity is likely a biomarker of alkylating agents’ sensitivity [199]. On the other hand, Oldrini and colleagues (2021) reported that MGMT genomic rearrangements carried by a subset of recurrent gliomas led to MGMT overexpression and TMZ resistance in vitro and in vivo, independently from changes in its promoter methylation [200].
Radiotherapy response is also modulated by DDR machinery. In TNBC patients, low expression of 53BP1, an NHEJ pathway protein, is associated with radioresistance [201]. Glioblastoma patients with nuclear PTEN phosphorylation show reduced sensitivity to radiation by enhancing DNA repair [202]. Overexpression or activation of the BER pathway is observed in radioresistant cells. Glioma cell lines with higher endogenous APE1 endonuclease are more radioresistant, and the APE1 ectopic expression increases radioresistance [203]. Moreover, radioresistant cancer cells and biopsies from radioresistant cancer patients show low expression of GADD45α in cervical cancer [204]. Cervical cancer patients with low expression of Ku80 respond better to radiotherapy, and hypopharyngeal squamous cell carcinoma patients with low Ku70 or XRCC4 proteins have better sensitivity to chemoradiotherapy [205,206]. The residual carcinoma from patients with cervical cancer after radiotherapy showed increased expression of DNA-PKcs, Ku70, and Ku86 genes, components of NHEJ pathways, compared to the counterpart primary tumors [207]. HR is also involved in radioresistance in cancer cells. For example, overexpression of BRCA1, BRCA2, RAD51, and RPA1 was observed in hypopharyngeal and nasopharyngeal carcinoma cells resistant to radiotherapy [208,209].

6. DNA Damage Response Inhibitors to Overcome Therapy Resistance

Tumor cells exhibit several therapy resistance mechanisms, but probably the most relevant ones are the inter- and intrapatient heterogeneity and intratumor heterogeneity [184]. Several strategies could be implemented to circumvent drug resistance, including adjustment of drug dose, optimization of therapies sequence, and targeting bypass mechanisms or alternative molecular targets by combinatorial approaches. The DDRi are potential alternative strategies to overcome cancer resistance (Table 2), and their combination with radiation, cytotoxic, or targeted agents can maximize the benefits of DDR targeted therapies.

6.1. Inhibition of PARP

In the absence of functional BRCA, targeting PARP is effective as monotherapy [210] and also sensitizes cancer cells to other drugs. In Palbociclib-resistant breast cancer cells, PARP inhibition combined with a STAT3 specific inhibitor re-sensitized cells to the agent, suggesting that concurrent targeting of DDR mechanisms and the IL6/STAT3 pathway could effectively treat acquired resistance to Palbociclib [211]. Epigenetic modifying agents have also been combined with PARP to improve therapeutic effectiveness. Specifically, the combination of DNA methyltransferase Gadecitabine and PARPi Tlazoparib were found to synergize in PARPi-resistant breast and ovarian cancer cells irrespective of BRCA status [212]. PARP inhibitors have been approved to treat high-grade serous ovarian cancer (HGSOC); however, women often develop resistance to treatment. The combination of antiangiogenic agent Cediranib with PARP inhibitor Olaparib was evaluated with varied results in a clinical trial of women with HGSOC who had developed resistance to therapy [213]. In HGSOC cell lines and PDX animal models, the checkpoint kinase 1 (CHK1) inhibitor Prexasertib showed efficacy as monotherapy but also sensitized cells to PARP inhibition [214]. PARP inhibition via Olaparib was able to reverse Sorafenib resistance in hepatocellular carcinoma by suppressing DDR mechanisms. In addition, Olaparib caused CHD1L-mediated chromatin condensation in the promoter region of transcription factors that promote cancer pluripotency [215].
Increased DNA repair in multiple myeloma is one of the main reasons for treatment resistance. Although not the standard of care for multiple myeloma, melphalan (MEL) is still used in combinatory strategies to treat this disease. To study whether PARP inhibition would reverse resistance to MEL, the agent was combined with several PARP inhibitors (Veliparib, Olaparib, and Iraparib) in multiple myeloma cell lines. The combination of MEL and PARPi in MEL-resistant cells showed an enhanced effect. However, MEL resistance, possibly caused by HR and NHEJ pathways, was not completely reversed by PARPi [216]. To discover the therapeutic potential of PARPi in combination treatment with other agents, a systems approach was developed by performing reverse-phase protein arrays to characterize adaptive responses to different therapies. The results indicate that the combination of PARPi with MEK/ERK, WEE1/ATR, and PI3K/AKT/mTOR inhibitors would show efficacy; this was also evident in various preclinical models. Based on this approach, several combinational therapies using PARPi are being assessed in clinical trials [217]. As mentioned before, TMZ, such as N3-meA and N7-meG DNA adducts, are easily repaired by BER [196]. To block BER and sensitize cells to TMZ, the agent may potentially be combined with PARP inhibitors that sustain N3-meA and N7-meG TMZ-induced lesions and improve the drug’s efficacy. A BER-independent function of PARP inhibitor Veliparib has also been shown to re-sensitize MMR deficient cells to TMZ [218]. In addition, knock-down of HR-involved proteins such as BRCA1 or RAD51 has been shown to improve the efficacy of TMZ [219].

6.2. Inhibition of ATM–ATR Complexes and Downstream Effectors

Resistant cancer cells may also be sensitized to treatment following exposure to ATM inhibitors. Kinases ATM, ATR, and their downstream effector kinases CHK1 and CHK2 are activated in response to DNA damage, leading to cell cycle arrest. The ATM–CHK2 axis is involved in G1 checkpoint control, whereas ATR–CHK1 controls the S and G2 checkpoints. Both ATM and ATR can convey their effects through P53, either directly or via activation of checkpoint kinase 2 (CHK2). P53 induces the CDK2 inhibitor P21 preventing damaged cells from entering the S phase [220].
Agent 2-morpholin-4-yl-6-thianthren-1-yl-pyran-4-one (KU-55933) acts as an inhibitor of ATM; specifically, it blocks phosphorylation of ATM and inhibits its downstream targets. KU55933 sensitized radioresistant breast cancer cells to ionizing radiation (IR). Specifically, breast cancer cells with a defective disabled homolog 2-interacting protein (DAB2IP) are often aggressive and resistant to radiation. KU-55933 improved the efficacy of IR against siDAB2IP breast cancer cells by targeting ATM and impairing DNA repair mechanisms [221]. Improved ATMi analogs with enhanced bioavailability, such as KU-60019 and AZ32, were also found to radio-sensitize human cancer glioma cells [222,223]. Similarly, VE-822, an ATR inhibitor, decreased the viability of pancreatic cancer cells following exposure to irradiation or to gemcitabine both in vitro and in vivo. The effect of VE-822 was achieved through dysregulation of cell cycle checkpoints and maintenance of DNA damage. Notably, the agent did not display cytotoxicity against normal cells. VE-821, another ATR inhibitor, successfully sensitized bone and ovarian cancer cells to radiation in vitro, forcing irradiated cells to divide into daughter cells and decreased survival selectively in cancer cells [224]. A detailed review describing the different approaches to sensitize cancer cells to radiation therapy by targeting DNA damage response components was recently published [225].
VX-970, a small-molecule ATR inhibitor, is currently being tested with promising results in many clinical trials in combination with chemotherapeutic drugs against resistant and aggressive cancers [226]. VX-970 also displayed radio-sensitizing effects in TNBC cells and PDX models. Specifically, the agent inhibited the ATR–CHK1–CDC25a axis signaling, sustained DNA double-strand breaks, and reduced colony formation following radiotherapy in TNBC cells. These effects were selective to cancer cells compared to normal epithelial breast cells [227]. A combination of CHK1 inhibitor PF-00477736 with Ibrutinib showed synergistic effects in vitro in several mantle cell lymphoma (MCL) cell lines. Ibrutinib is a Bruton’s tyrosine kinase (BTK) inhibitor that has been approved for refractory MCL. The study showed that in MCL cells resistant to Ibrutinib, the combination with CHK1 inhibitor led to enhanced effects [228]. The ATR inhibitor NU60 induces G2/M arrest and impairs homologous recombination, leading to increased sensitivity of breast cancer cells to DNA-damaging agents, such as cisplatin, and PARP inhibitors [229].
The APC tumor suppressor gene is inactive in 70% of sporadic breast cancers; APC-deficient tumors resemble the aggressive TNBC subtype. APC deficiency decreases sensitivity to doxorubicin (DOX), which is attributed to the inactivation of ATM, CHK1, and CHK2 and increased DNA repair in the presence of DOX. Concurrent inhibition of ATM and DNA-PK enhanced DOX-induced apoptosis in resistant cells [230]. These findings support that inhibition of the ATM–ATR–CHK axis is a promising approach to enhance radiation or chemotherapy therapeutic efficacy [231]. Importantly, synthetic lethality with ATM, ATR, and DNA-PK inhibitors is being evaluated to target HR-proficient cells [232,233].

6.3. Inhibition of WEE1

A recent review highlights the potential of WEE1 inhibition in radio- and chemosensitization [234]. WEE1 is a protein kinase mainly localized in the nucleus. It negatively regulates the G2/M transition following the detection of DSB [235,236]. It affects the CDK1–cyclin B complex by phosphorylating and inactivating Cyclin B on Tyr15, causing cell cycle arrest at G2. When errors happen during replication, this mechanism blocks the cell cycle to allow for repair; downregulation of WEE1, either by decreased synthesis or through proteolytic degradation, promotes entry into mitosis [237,238]. The role of WEE1 as a gatekeeper for the G2/M transition suggests that it acts as a tumor suppressor gene; however, WEE1 was overexpressed in patients with hepatocellular carcinoma, medulloblastoma, and glioma, and its levels were further elevated after exposure to chemotherapy in patients with ovarian cancer [239,240,241,242]. Overexpression of WEE1 in melanoma cells has been correlated with proliferation markers, including Ki-67 [243]. Therefore, it is postulated that the expression of WEE1 allows cancer cells to repair DNA damage following chemo- or radiotherapy, develop resistance, and continue to proliferate. In addition, cancer stem cells adopt high WEE1 expression as a protection mechanism against therapeutic agents [244]. It is well established that cancer stem cells convey resistance to DNA-damaging treatments; their percentage increases in the tumor cell population following the progressive deterioration of non-stem cells.
WEE1 inhibition represents an attractive approach for radio- and chemotherapy potentiation. Several pharmacological inhibitors belonging to different chemical classes have been developed against WEE1 and are described in a recent review [245]. AZD1775 is a WEE1 inhibitor currently in clinical trials, combined with DNA damage agents or radiotherapy. AZD1775 has been found to have a radio-sensitizing effect in pancreatic cancer, pontine gliomas, and glioblastoma [246,247,248]. Some studies suggest that AZD1775′s ability to sensitize cells to therapy is effective only in TP53-deficient tumors [249,250,251]. The combination of AZD1775 with cisplatin sensitized squamous cell carcinoma of the head and neck (HNSCC) cells to the latter in in vitro and in vivo models; importantly, HNSCC cells carrying high-risk TP53 mutations became sensitive to cisplatin treatment by the selective WEE1 kinase inhibitor [252]. In conclusion, inhibition of WEE1 may sensitize cells to DNA damage therapy; although P53 has been reported to affect the effectiveness of this approach, other studies support that WEE1 inhibition sensitizes cancer cells to chemotherapeutics independently of P53 function [253].

6.4. Inhibition of DNA-Dependent Protein Kinase to Re-Sensitize Cells

DNA-PK belongs to the PI3K-related protein kinase (PIKK) superfamily. It participates in NHEJ to repair DSBs in DNA [254]. DNA-PK may play a role in resistance to chemotherapy and radiotherapy [255,256]. Several inhibitors, including small molecules, have been developed to target DNA-PK, block the DSB repair pathway, and sensitize cells to therapy [199,257]. The small molecule DNA-PK inhibitor, PI-103 or NU7441, combined with the third-generation epidermal growth factor receptor (EGFR) tyrosine kinase inhibitor (TKI) Osimertinib, led to synergistic effects in TKI-resistant lung cancer cells. The enhanced effect was attributed to the prolongation of DNA damage and cell cycle arrest [258]. The DNA-PKcs inhibitor NU7441 blocked glioma stem cell tumorsphere formation in vitro. In addition, in human-derived glioblastoma xenograft mice, the inhibitor blocked tumor growth and sensitized cancer cells to radiotherapy [259].

7. Mechanisms of Resistance to DNA Damage Response Inhibitors

As with most target therapies, some patients are primarily resistant to DDRi and others eventually develop acquired resistance (Figure 1), with the latter being more frequent in patients with advanced disease [260]. Since PARPi are the only DDRi approved for clinical use, most known resistance mechanisms are associated with these inhibitors.

7.1. Resistance to PARP Inhibitors

The mechanisms of resistance to PARPi can be credited to several factors, including restoration of the mechanisms controlled by BRCA, such as HR repair and/or stabilization of replication forks [261,262]. Like other systemic chemotherapies, cancer cell develops PARPi resistance via several different mechanisms: (i) increased expression of multidrug resistance pumps (MDRs), enhancing the efflux of the PARPi out of the cell [263], (ii) reduced PARP1 binding affinity to DNA due to mutations and functional alterations of the PARP1 protein and/or disrupted PARylation [264,265], or (iii) restored HR and/or replication fork stabilization [266,267,268,269].
In BRCA-deficient tumors, the most frequent acquired resistance mechanism to PARPi is the re-establishment of BRCA1 or BRCA2 functionality by secondary intragenic mutations; specifically, genetic alterations may reinstate the open reading frame (ORF), leading to the expression of functional BRCA [270]. In addition, restoration of the wild-type BRCA protein may occur via a secondary mutation that reverses the inherited mutation or by the demethylation of the BRCA1 promoter; both these events may lead to restoration of the wild-type BRCA protein [160,271]. Specific mutations in the BRCA gene, including the BRCA1-C61G mutation, may also confer PARPi and cisplatin resistance [272]. Another possible way of HR restoration is due to loss of the shieldin complex, consisting of REV7, c20orf196 (SHLD1), FAM35a (SHLD2), and FLJ26957 (SHLD3), which normally prevents DSB resection and facilitates NHEJ. However, if lost, shieldin can promote PARPi resistance even in the absence of BRCA [273,274]. ATPase TRIP13 inactivates the shieldin complex, triggering the 5′ to 3′ resection of double-strand breaks and promoting HR [275]. In many BRCA-deficient tumors, TRIP13 is upregulated, contributing to the intrinsic PARPi resistance. Inhibiting the ATPase domain of TRIP13 can stabilize the shieldin complex to promote NHEJ, block HR and overcome intrinsic PARPi resistance. Inhibiting TRIP13 might be useful to treat BRCA-deficient tumors with intrinsic but also acquired PARPi resistance [275]. Secondary mutations restoring BRCA function were found in patients with germline BRCA mutation-associated ovarian and breast cancer upon acquired resistance to PARPi and/or cisplatin [276]. Reversion mutations of BRCA1 can also exhibit the MMEJ signature, pointing to the potential involvement of POLQ in driving resistance [277]. Consequently, inhibitors of POLQ can suppress PARPi resistance in HR and NHEJ-deficient cancers [277]. A detailed review of the mechanisms of BRCA re-activation in PARPi resistant cells was recently published [261].
In addition, BRCA1/2-deficient cancer cells may develop PARPi resistance by protecting their replication forks; they achieve this by blocking the recruitment of nucleases, MRE11 or MUS81, to the stalled fork, thereby resulting in fork protection [267,268]. These studies indicate that PARPi resistance is achieved without restoring HR repair. Furthermore, other mechanisms of resistance to PARPi have been reported, such as downregulation of PARP levels and increased levels of the P-glycoprotein efflux pump [260,278]. Overexpression of ABCB1 has been reported in PARPi-resistant human ovarian cancer cells; administration of MDR1 inhibitors such as Verapamil and Elacridar reversed resistance to PARPi [279]. It has been reported that most clinical PARP inhibitors induce cytotoxicity by trapping PARP1 at sites of DNA damage [280]. Resistance to PARP inhibitors can emerge through point mutations in PARP1 that alter PARP1 trapping, highlighting the importance of PARP1 intramolecular interactions in PARPi-mediated cytotoxicity [264].

7.2. Cell Cycle Regulators in DNA Damage Response Inhibitors in Resistant Cells

Several reports suggest that the silencing of cyclins may confer resistance to DDR inhibitors. There are two major classes of G1 cyclins that regulate cell cycle progression during the G1 phase: cyclin D, which cooperates with either CDK4 or CDK6, and cyclin E, which binds CDK2. During cell cycle progression in the early G1 phase, cyclin D–CDK4/6 complexes phosphorylate the Rb protein. Complete phosphorylation of the Rb protein is achieved at the end of the G1 phase by the CDK2/cyclin E complex. Fully phosphorylated Rb protein is inactive and releases the E2F factor, allowing the expression of S phase genes, leading the cells through the G1/S checkpoint. In response to DNA damage, p21 levels are increased; p21 binds both to the cyclin and the CDK subunits of the CDK/cyclin complex and disrupts the interaction between CDK and its substrates, blocking cell cycle progression [281].
Downregulation of cyclin D has been shown to confer resistance to CHK1 inhibition. CHK1, a serine/threonine kinase that acts as an ATM–ATR effector, is activated following exogenous DNA damage, including nicks caused by chemotherapeutic drugs. CHK1 activates the S and G2 checkpoints by controlling different mechanisms of DNA repair, including activation of homologous recombination repair or apoptosis if DNA damage is too severe [282,283]. In a MCL cell line that was made resistant to the CHK1 inhibitor PF-00477736, the re-expression of cyclin D1 partially re-sensitized cells to the agent. This suggests that low levels of cyclin D1 confer resistance to CHK1 inhibitors and that re-establishment of this protein may re-sensitize cells [284].
Cell division cycle 25A (CDC25A) is a dual-specificity phosphatase implicated in cell cycle control by inhibiting CDK phosphorylation and causing the formation of cyclin–CDK complexes. Following DNA damage, CDC25A is degraded, leading to cell cycle arrest. CDC25A is overexpressed in cancer and promotes tumorigenesis [285]; interestingly, a genome-wide CRISPR screen showed that the absence of CDC25A leads to ATR inhibitor resistance. Loss of CDC25A led to cell cycle arrest in cells treated with ATR inhibitor, diminishing the DNA damage caused by ATR inhibitors might otherwise generate; resistance was reversed using a WEE1 inhibitor that forced mitotic entry [286].
Dysfunctional apoptosis is one of the hallmarks of cancer. The increased levels and/or activity of anti-apoptotic proteins and, concurrently, the inactivation of pro-apoptotic molecules convey resistance to many anticancer drugs [287]. P53, a key molecule controlling cell cycle fate following DNA damage, is silenced in most human cancers. However, restoration of P53 following inhibition of MDM2 by Nutlin conveyed resistance to the cytotoxic effects of WEE1 inhibitor AZD1775 [288].

7.3. Activation of Alternative DNA Repair Pathways

Several studies report that the activation of alternative pathways to repair DNA damage is responsible for the observed resistance to DDR inhibitors. In P53-deficient cells, the induction of DSBs using a radiomimetic agent and DNA-PK inhibition led to an increased DSB burden in the S-phase; however, a subset of the cell population exhibited resistance to this combination therapy, which was caused by the recruitment of DNA polymerase theta (Pol θ or POLQ). Pol θ mediated end joining repair to improve cell viability following therapy-induced DNA damage. Concurrent inhibition of Pol θ and DNA-PK sensitized p53-deficient breast cancer cells to therapy [289].

8. Liabilities upon Treatment with DDR Inhibitors

Target therapies, including PARPis, contribute to important therapeutic breakthroughs in oncology, improving the quality of life and increasing the life expectancy of cancer patients. As mentioned, PARPis were demonstrated to be clinically effective, with acceptable tolerability and safety, in a specific range of solid tumors, which led to FDA and European Medicines Agency (EMA) approval of Olaparib, Rucaparib, Niraparib, and Talazoparib [31,290]. However, a consolidated body of evidence from studies of PARPi in patients has identified several adverse events and specific indications for their prevention, monitoring, and management [291,292,293,294]. PARPi display several on- and off-target toxicities, with hematological and gastrointestinal toxicities among the most common adverse events. Pneumonitis and therapy-related myeloid neoplasias (t-MN), such as AML and myelodysplastic syndromes (MDS), have been reported with PARPi, but despite their rare frequency, they are potentially life-threatening, often fatal, and deserve particular attention due to their severity [291]. The t-MN is typically a late complication of some chemo- and radiotherapy, and the subtype and latency period are usually treatment-dependent [295].
The link between PARPi and the development of t-MN is not fully understood. The pretreatment presence of clonal hematopoiesis of indeterminate potential (CHIP) with TP53 mutations [296], a hematopoietic cell population with one or more somatic mutations/copy number alterations that can expand with time and under positive clonal selection pressures [297], have been proposed as a possible explanation. Kwan et al. also analyzed the risk of t-MN development in patients with HR gene alterations and found a higher prevalence in patients with high-grade ovarian cancer that harbored a deleterious mutation in BRCA1, BRCA2, RAD51C, or RAD51D (4.1%) compared to those with mutation-containing cancers (1.0%) and without mutations (1.0%) [296]. Mutations of DDR genes (e.g., TP53, PPM1D, and CHEK2) involved in CHIP occur with increased frequency in cancer patients exposed to platinum compounds/topoisomerase II inhibitors or radiation therapy [296,298]. Additionally, previous treatments with platinum and alkylating agents may increase the risk of t-MN development in BRCA-associated high-grade ovarian cancer patients treated with PARPi as maintenance therapy [299]. PARPi may potentiate t-MN in patients with preexisting CHIP by selecting clones with DDR gene mutations that improve the competitive fitness of the cells under these conditions [31,299]. Oliveira et al. (2022) performed a comprehensive analysis of the pathologic and genetic characteristics of PARPi-related t-MN patients, showing that these patients have complex karyotypes and frequently have pathogenic TP53 mutations [300].
Most data available about t-MN arise from gynecologic cancer patients treated with Olaparib, with an estimated frequency of t-MN development between 1% in the PAOLA-1 study [105] and 8% in the SOLO-2 trial [301]. A recent study by Morice et al. (2021) evaluated the safety profile of 31 randomized controlled trials with PARPis as one arm in different tumor types and settings [302]. In this systematic review, PARP inhibitors significantly increased the risk of AML and MDS in comparison with placebo treatment (Peto OR 2.63 [95% CI 1.13–6.14], p = 0.026); the incidence of these t-MN across PARPi groups of 0.73% and placebo groups was 0.47%, with a median latency between first PARPi and the t-MN onset of 17.8 months [302]. The risk of t-MN development was small but more than doubled, even after controlling for prior platinum-based chemotherapy. In a meta-analysis, Nitecki et al. (2021) did not find an increased incidence of t-MN in PARPi-treated patients compared to control treatments [303]. However, patients who received a PARPi as frontline treatment and those who received fewer than two prior lines of chemotherapy showed a higher risk of t-MN [303]. In a pharmacovigilance analysis of the FDA adverse event reporting system, Ma et al. (2021) verified a dramatic increase in PARPi related t-MN from 2015 to 2019 and found a higher reporting of t-MN in patients treated with PARPi (reporting odds ratio (ROR) 16.47, 95% CI 14.72–18.44), with RORs (95% CI) of 48.03 (42.21–54.64) for Olaparib, 6.58 (5.03–8.61) for Niraparib, and 2.23 (1.32–3.77) for Rucaparib [304]. Current studies showed several limitations, including the cross-over between control and PARPi arms. This scenery may overestimate the incidence of t-MN in control/placebo arms since the subsequent therapies are not regularly reported [303].
Clinicians need to be aware of these late but potentially fatal adverse events, especially in the front-line maintenance settings, and pharmacovigilance and mechanistic studies should be implemented to improve the understanding of the risk factors that predispose to t-MN. Identifying biomarkers that discriminate patients at high risk of t-MN development upon PARPi treatment from those who benefit from frontline PARPi will improve treatment outcomes and prevent undesired adverse events.

9. Conclusions and Perspectives

Cancer cells display several defects in DDR pathways, offering a chance to explore these deficiencies clinically. DDR-based cancer treatments and combinatory regimens provide potential therapeutic approaches that exploit deficiency DDR pathways via synthetic lethality strategies. Despite the success of PARPi in HR-deficient cancers, such as breast, ovarian, and pancreatic cancers, several patients present serious toxicities or developed resistance to DDRi. A variety of DDRi resistance mechanisms have already been identified in preclinical models and patients, but clinical data are still scarce, and this remains an open field of research. Another challenge in DDR-based cancer treatments is the identification of genetic and functional biomarkers that define the patients who will be most suitable, suffer fewer side effects and toxicity, and benefit more from these therapeutic options. Moreover, although the higher benefits of DDRi are observed in patients with impaired DDR machinery, patients with proficient cancers can also benefit from these therapeutic approaches. Thus, further investigation is warranted to identify differential strategies for these patients, including combinatory approaches with targeted therapies such as immunotherapies (e.g., immune checkpoint inhibitors and non-specific immunotherapies), anti-angiogenic agents (e.g., VEGF inhibitors), and metabolic drugs (e.g., IDH inhibitors), among others. Another possible strategy is to combine different DDRi (e.g., PARPi with ATM, ATR, WEE1, or CHK1/2 inhibitors). Currently, PARPi is the maintenance therapy of choice for some cancers, such as ovarian, fallopian tube, primary perineal, and pancreatic cancer, showing manageable toxicity profiles. However, it should be highlighted that PARPi treatment increases the risk of AML and MDS development. This is a rare but frequently fatal event, and prescribing clinicians should remain vigilant about this complication. Additional research, including long-term pharmacovigilance studies, is needed to identify toxicity-predisposing factors and susceptibility biomarkers to further refine and personalize DDRi treatment and prevent t-MN development in front-line and maintenance settings. Furthermore, a better understanding of the molecular mechanisms of resistance to DDRi and the development of strategies to prevent or delay the acquisition of resistance are needed.

Author Contributions

D.J., C.M.N., A.Č.G. and A.C.G. contributed to the study conception, design, and first draft of the manuscript. D.J., C.M.N., A.Č.G. and A.C.G. commented on previous versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The Foundation for Science and Technology, Portugal, supports A.C.G. (UID/NEU/04539/2019, UIDB/04539/2020, UIDP/04539/2020). D.J. was supported by VEGA Grant Agency of the Slovak Republic (grant no. 2/0056/21), the Slovak Research and Development Agency (grant no. APVV-19-0286), the Ministry of Education, Science Research and Sport of Slovak Republic (grant no. MVTS 34097104). This article is based on work from COST Action STRATAGEM, CA17104, supported by COST (European Cooperation in Science and Technology) (www.cost.eu) (accessed on 30 August 2021).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hoeijmakers, J.H.J. DNA Damage, Aging, and Cancer. N. Engl. J. Med. 2009, 361, 1475–1485. [Google Scholar] [CrossRef]
  2. Ishiai, M. Regulation of the Fanconi Anemia DNA Repair Pathway by Phosphorylation and Monoubiquitination. Genes 2021, 12, 1763. [Google Scholar] [CrossRef]
  3. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Barnieh, F.M.; Loadman, P.M.; Falconer, R.A. Progress towards a clinically-successful ATR inhibitor for cancer therapy. Curr. Res. Pharmacol. Drug Discov. 2021, 2, 100017. [Google Scholar] [CrossRef] [PubMed]
  5. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Giglia-Mari, G.; Zotter, A.; Vermeulen, W. DNA damage response. Cold Spring Harb. Perspect. Biol. 2011, 3, a000745. [Google Scholar] [CrossRef] [Green Version]
  7. Tian, H.; Gao, Z.; Li, H.; Zhang, B.; Wang, G.; Zhang, Q.; Pei, D.; Zheng, J. DNA damage response—A double-edged sword in cancer prevention and cancer therapy. Cancer Lett. 2015, 358, 8–16. [Google Scholar] [CrossRef] [PubMed]
  8. Pilié, P.G.; Tang, C.; Mills, G.B.; Yap, T.A. State-of-the-art strategies for targeting the DNA damage response in cancer. Nat. Rev. Clin. Oncol. 2018, 16, 81–104. [Google Scholar] [CrossRef] [PubMed]
  9. Baretti, M.; Le, D.T. DNA mismatch repair in cancer. Pharmacol. Ther. 2018, 189, 45–62. [Google Scholar] [CrossRef] [PubMed]
  10. Lau, P.J.; Kolodner, R.D. Transfer of the MSH2·MSH6 Complex from Proliferating Cell Nuclear Antigen to Mispaired Bases in DNA. J. Biol. Chem. 2003, 278, 14–17. [Google Scholar] [CrossRef] [PubMed]
  11. Genschel, J.; Modrich, P. Mechanism of 5′-Directed Excision in Human Mismatch Repair. Mol. Cell 2003, 12, 1077–1086. [Google Scholar] [CrossRef]
  12. Longley, M.J.; Pierce, A.J.; Modrich, P. DNA Polymerase δ Is Required for Human Mismatch Repair in Vitro. J. Biol. Chem. 1997, 272, 10917–10921. [Google Scholar] [CrossRef] [Green Version]
  13. Zhang, Y.; Yuan, F.; Presnell, S.R.; Tian, K.; Gao, Y.; Tomkinson, A.E.; Gu, L.; Li, G.-M. Reconstitution of 5′-Directed Human Mismatch Repair in a Purified System. Cell 2005, 122, 693–705. [Google Scholar] [CrossRef] [Green Version]
  14. Schärer, O.D. Nucleotide Excision Repair in Eukaryotes. Cold Spring Harb. Perspect. Biol. 2013, 5, a012609. [Google Scholar] [CrossRef] [Green Version]
  15. Feher, K.M.; Kolbanovskiy, A.; Durandin, A.; Shim, Y.; Min, J.-H.; Lee, Y.C.; Shafirovich, V.; Mu, H.; Broyde, S.; Geacintov, N.E. The DNA damage-sensing NER repair factor XPC-RAD23B does not recognize bulky DNA lesions with a missing nucleotide opposite the lesion. DNA Repair 2020, 96, 102985. [Google Scholar] [CrossRef]
  16. Tsutakawa, S.E.; Tsai, C.-L.; Yan, C.; Bralić, A.; Chazin, W.J.; Hamdan, S.M.; Schärer, O.D.; Ivanov, I.; Tainer, J.A. Envisioning how the prototypic molecular machine TFIIH functions in transcription initiation and DNA repair. DNA Repair 2020, 96, 102972. [Google Scholar] [CrossRef]
  17. Coin, F.; Oksenych, V.; Egly, J.-M. Distinct Roles for the XPB/p52 and XPD/p44 Subcomplexes of TFIIH in Damaged DNA Opening during Nucleotide Excision Repair. Mol. Cell 2007, 26, 245–256. [Google Scholar] [CrossRef]
  18. Staresincic, L.; Fagbemi, A.F.; Enzlin, J.H.; Gourdin, A.M.; Wijgers, N.; Dunand-Sauthier, I.; Giglia-Mari, G.; Clarkson, S.G.; Vermeulen, W.; Schärer, O.D. Coordination of dual incision and repair synthesis in human nucleotide excision repair. EMBO J. 2009, 28, 1111–1120. [Google Scholar] [CrossRef] [Green Version]
  19. Shivji, M.K.K.; Podust, V.N.; Huebscher, U.; Wood, R.D. Nucleotide Excision Repair DNA Synthesis by DNA Polymerase epsilon in the Presence of PCNA, RFC, and RPA. Biochemistry 1995, 34, 5011–5017. [Google Scholar] [CrossRef]
  20. Jacobs, A.L.; Schär, P. DNA glycosylases: In DNA repair and beyond. Chromosoma 2011, 121, 1–20. [Google Scholar] [CrossRef]
  21. Caldecott, K.W. Mammalian DNA base excision repair: Dancing in the moonlight. DNA Repair 2020, 93, 102921. [Google Scholar] [CrossRef] [PubMed]
  22. Zhou, J.; Fleming, A.M.; Averill, A.M.; Burrows, C.J.; Wallace, S.S. The NEIL glycosylases remove oxidized guanine lesions from telomeric and promoter quadruplex DNA structures. Nucleic Acids Res. 2015, 43, 4039–4054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Chaudhari, S.; Ware, A.P.; Jayaram, P.; Gorthi, S.P.; El-Khamisy, S.F.; Satyamoorthy, K. Apurinic/Apyrimidinic Endonuclease 2 (APE2): An ancillary enzyme for contextual base excision repair mechanisms to preserve genome stability. Biochimie 2021, 190, 70–90. [Google Scholar] [CrossRef] [PubMed]
  24. Chakraborty, A.; Tapryal, N.; Islam, A.; Mitra, S.; Hazra, T. Transcription coupled base excision repair in mammalian cells: So little is known and so much to uncover. DNA Repair 2021, 107, 103204. [Google Scholar] [CrossRef]
  25. Caldecott, K. Protein ADP-ribosylation and the cellular response to DNA strand breaks. DNA Repair 2014, 19, 108–113. [Google Scholar] [CrossRef]
  26. Rouleau, M.; Patel, A.; Hendzel, M.J.; Kaufmann, S.H.; Poirier, G.G. PARP inhibition: PARP1 and beyond. Nat. Rev. Cancer 2010, 10, 293–301. [Google Scholar] [CrossRef] [Green Version]
  27. Hong, S.J.; Dawson, T.M.; Dawson, V.L. Nuclear and mitochondrial conversations in cell death: PARP-1 and AIF signaling. Trends Pharmacol. Sci. 2004, 25, 259–264. [Google Scholar] [CrossRef]
  28. Soldani, C.; Scovassi, A.I. Poly(ADP-ribose) polymerase-1 cleavage during apoptosis: An update. Apoptosis 2002, 7, 321–328. [Google Scholar] [CrossRef]
  29. Kaufmann, S.H.; Desnoyers, S.; Ottaviano, Y.; Davidson, N.E.; Poirier, G.G. Specific proteolytic cleavage of poly(ADP-ribose) polymerase: An early marker of chemotherapy-induced apoptosis. Cancer Res. 1993, 53, 3976–3985. [Google Scholar]
  30. Demin, A.A.; Hirota, K.; Tsuda, M.; Adamowicz, M.; Hailstone, R.; Brazina, J.; Gittens, W.; Kalasova, I.; Shao, Z.; Zha, S.; et al. XRCC1 prevents toxic PARP1 trapping during DNA base excision repair. Mol. Cell 2021, 81, 3018–3030.e5. [Google Scholar] [CrossRef]
  31. Padella, A.; Di Rorà, A.G.L.; Marconi, G.; Ghetti, M.; Martinelli, G.; Simonetti, G. Targeting PARP proteins in acute leukemia: DNA damage response inhibition and therapeutic strategies. J. Hematol. Oncol. 2022, 15, 10. [Google Scholar] [CrossRef] [PubMed]
  32. Das, B.B.; Huang, S.Y.; Murai, J.; Rehman, I.; Amé, J.C.; Sengupta, S.; Das, S.K.; Majumdar, P.; Zhang, H.; Biard, D.; et al. PARP1–TDP1 coupling for the repair of topoisomerase I–induced DNA damage. Nucleic Acids Res. 2014, 42, 4435–4449. [Google Scholar] [CrossRef] [PubMed]
  33. Kieffer, S.R.; Lowndes, N.F. Immediate-Early, Early, and Late Responses to DNA Double Stranded Breaks. Front. Genet. 2022, 13, 793884. [Google Scholar] [CrossRef] [PubMed]
  34. Caron, M.-C.; Sharma, A.K.; O’Sullivan, J.; Myler, L.R.; Ferreira, M.T.; Rodrigue, A.; Coulombe, Y.; Ethier, C.; Gagné, J.-P.; Langelier, M.-F.; et al. Poly(ADP-ribose) polymerase-1 antagonizes DNA resection at double-strand breaks. Nat. Commun. 2019, 10, 2954. [Google Scholar] [CrossRef] [Green Version]
  35. Ronato, D.; Mersaoui, S.Y.; Busatto, F.F.; Affar, E.B.; Richard, S.; Masson, J.-Y. Limiting the DNA Double-Strand Break Resectosome for Genome Protection. Trends Biochem. Sci. 2020, 45, 779–793. [Google Scholar] [CrossRef]
  36. Lieber, M.R. The Mechanism of Double-Strand DNA Break Repair by the Nonhomologous DNA End-Joining Pathway. Annu. Rev. Biochem. 2010, 79, 181–211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Scully, R.; Panday, A.; Elango, R.; Willis, N.A. DNA double-strand break repair-pathway choice in somatic mammalian cells. Nat. Rev. Mol. Cell Biol. 2019, 20, 698–714. [Google Scholar] [CrossRef]
  38. Branzei, D.; Szakal, B. DNA helicases in homologous recombination repair. Curr. Opin. Genet. Dev. 2021, 71, 27–33. [Google Scholar] [CrossRef]
  39. Branzei, D.; Szakal, B. DNA damage tolerance by recombination: Molecular pathways and DNA structures. DNA Repair 2016, 44, 68–75. [Google Scholar] [CrossRef]
  40. Koike, M. Dimerization, Translocation and Localization of Ku70 and Ku80 Proteins. J. Radiat. Res. 2002, 43, 223–236. [Google Scholar] [CrossRef] [Green Version]
  41. Bowcock, A. Molecular cloning of BRCA1: A gene for early onset familial breast and ovarian cancer. Breast Cancer Res. Treat. 1993, 28, 121–135. [Google Scholar] [CrossRef] [PubMed]
  42. Panzarino, N.J.; Krais, J.J.; Cong, K.; Peng, M.; Mosqueda, M.; Nayak, S.U.; Bond, S.M.; Calvo, J.A.; Doshi, M.B.; Bere, M.; et al. Replication Gaps Underlie BRCA Deficiency and Therapy Response. Cancer Res. 2021, 81, 1388–1397. [Google Scholar] [CrossRef] [PubMed]
  43. Walsh, C.S. Two decades beyond BRCA1/2: Homologous recombination, hereditary cancer risk and a target for ovarian cancer therapy. Gynecol. Oncol. 2015, 137, 343–350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Krejci, L.; Altmannova, V.; Spirek, M.; Zhao, X. Homologous recombination and its regulation. Nucleic Acids Res. 2012, 40, 5795–5818. [Google Scholar] [CrossRef] [PubMed]
  45. Fernandez, A.; O’Leary, C.; O’Byrne, K.J.; Burgess, J.; Richard, D.J.; Suraweera, A. Epigenetic Mechanisms in DNA Double Strand Break Repair: A Clinical Review. Front. Mol. Biosci. 2021, 8, 685440. [Google Scholar] [CrossRef]
  46. Sellou, H.; Lebeaupin, T.; Chapuis, C.; Smith, R.; Hegele, A.; Singh, H.R.; Kozlowski, M.; Bultmann, S.; Ladurner, A.G.; Timinszky, G.; et al. The poly(ADP-ribose)-dependent chromatin remodeler Alc1 induces local chromatin relaxation upon DNA damage. Mol. Biol. Cell 2016, 27, 3791–3799. [Google Scholar] [CrossRef]
  47. Karakaidos, P.; Karagiannis, D.; Rampias, T. Resolving DNA Damage: Epigenetic Regulation of DNA Repair. Molecules 2020, 25, 2496. [Google Scholar] [CrossRef]
  48. Lahtz, C.; Pfeifer, G.P. Epigenetic changes of DNA repair genes in cancer. J. Mol. Cell Biol. 2011, 3, 51–58. [Google Scholar] [CrossRef] [Green Version]
  49. Guan, H.; Ji, M.; Hou, P.; Liu, Z.; Wang, C.; Shan, Z.; Teng, W.; Xing, M. Hypermethylation of the DNA mismatch repair gene hMLH1 and Its association with lymph node metastasis and T1799A BRAF mutation in patients with papillary thyroid cancer. Cancer 2008, 113, 247–255. [Google Scholar] [CrossRef]
  50. Czerninski, R.; Krichevsky, S.; Ashhab, Y.; Gazit, D.; Patel, V.; Ben-Yehuda, D. Promoter hypermethylation of mismatch repair genes, hMLH1 and hMSH2 in oral squamous cell carcinoma. Oral Dis. 2009, 15, 206–213. [Google Scholar] [CrossRef] [PubMed]
  51. Liu, K.; Huang, H.; Mukunyadzi, P.; Suen, J.Y.; Hanna, E.; Fan, C.-Y. Promoter Hypermethylation: An Important Epigenetic Mechanism for hMLH1 Gene Inactivation in Head and Neck Squamous Cell Carcinoma. Otolaryngol. Neck Surg. 2002, 126, 548–553. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.-C.; Lu, Y.-P.; Tseng, R.-C.; Lin, R.-K.; Chang, J.-W.; Chen, J.-T.; Shih, C.-M.; Chen, C.-Y. Inactivation of hMLH1 and hMSH2 by promoter methylation in primary non-small cell lung tumors and matched sputum samples. J. Clin. Investig. 2003, 111, 887–895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Seedhouse, C.H.; Das-Gupta, E.P.; Russell, N.H. Methylation of the hMLH1 promoter and its association with microsatellite instability in acute myeloid leukemia. Leukemia 2003, 17, 83–88. [Google Scholar] [CrossRef] [Green Version]
  54. Fleisher, A.S.; Esteller, M.; Tamura, G.; Rashid, A.; Stine, O.C.; Yin, J.; Zou, T.-T.; Abraham, J.M.; Kong, D.; Nishizuka, S.; et al. Hypermethylation of the hMLH1 gene promoter is associated with microsatellite instability in early human gastric neoplasia. Oncogene 2001, 20, 329–335. [Google Scholar] [CrossRef] [Green Version]
  55. Gras, E.; Cortes, J.; Diez, O.; Alonso, C.; Matias-Guiu, X.; Baiget, M.; Prat, J. Loss of heterozygosity on chromosome 13q12-q14, BRCA-2 mutations and lack of BRCA-2 promoter hypermethylation in sporadic epithelial ovarian tumors. Cancer 2001, 92, 787–795. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, L.; Long, X. Association of BRCA1 promoter methylation with sporadic breast cancers: Evidence from 40 studies. Sci. Rep. 2015, 5, 17869. [Google Scholar] [CrossRef] [Green Version]
  57. Yu, J.; Zhu, T.; Wang, Z.; Zhang, H.; Qian, Z.; Xu, H.; Gao, B.; Wang, W.; Gu, L.; Meng, J.; et al. A Novel Set of DNA Methylation Markers in Urine Sediments for Sensitive/Specific Detection of Bladder Cancer. Clin. Cancer Res. 2007, 13, 7296–7304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Lee, M.-N.; Tseng, R.-C.; Hsu, H.-S.; Chen, J.-Y.; Tzao, C.; Ho, W.L.; Wang, Y.-C. Epigenetic Inactivation of the Chromosomal Stability Control Genes BRCA1, BRCA2, and XRCC5 in Non–Small Cell Lung Cancer. Clin. Cancer Res. 2007, 13, 832–838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Bernal, C.; Vargas, M.; Ossandón, F.; Santibáñez, E.; Urrutia, J.; Luengo, V.; Zavala, L.F.; Backhouse, C.; Palma, M.; Argandoña, J.; et al. DNA methylation profile in diffuse type gastric cancer: Evidence for hypermethylation of the BRCA1 promoter region in early-onset gastric carcinogenesis. Biol. Res. 2008, 41, 303–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Gao, D.; Herman, J.G.; Guo, M. The clinical value of aberrant epigenetic changes of DNA damage repair genes in human cancer. Oncotarget 2016, 7, 37331–37346. [Google Scholar] [CrossRef] [PubMed]
  61. Plantamura, I.; Cosentino, G.; Cataldo, A. MicroRNAs and DNA-Damaging Drugs in Breast Cancer: Strength in Numbers. Front. Oncol. 2018, 8, 352. [Google Scholar] [CrossRef] [PubMed]
  62. Petri, B.; Klinge, C.M. Regulation of breast cancer metastasis signaling by miRNAs. Cancer Metastasis Rev. 2020, 39, 837–886. [Google Scholar] [CrossRef] [PubMed]
  63. Fridrichova, I.; Zmetakova, I. MicroRNAs Contribute to Breast Cancer Invasiveness. Cells 2019, 8, 1361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Ding, L.; Gu, H.; Xiong, X.; Ao, H.; Cao, J.; Lin, W.; Yu, M.; Lin, J.; Cui, Q. MicroRNAs Involved in Carcinogenesis, Prognosis, Therapeutic Resistance, and Applications in Human Triple-Negative Breast Cancer. Cells 2019, 8, 1492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Kandettu, A.; Radhakrishnan, R.; Chakrabarty, S.; Sriharikrishnaa, S.; Kabekkodu, S.P. The emerging role of miRNA clusters in breast cancer progression. Biochim. Biophys. Acta Rev. Cancer 2020, 1874, 188413. [Google Scholar] [CrossRef] [PubMed]
  66. Kastl, L.; Brown, I.; Schofield, A.C. miRNA-34a is associated with docetaxel resistance in human breast cancer cells. Breast Cancer Res. Treat. 2011, 131, 445–454. [Google Scholar] [CrossRef] [PubMed]
  67. Bao, C.; Chen, J.; Chen, D.; Lu, Y.; Lou, W.; Ding, B.; Xu, L.; Fan, W. MiR-93 suppresses tumorigenesis and enhances chemosensitivity of breast cancer via dual targeting E2F1 and CCND1. Cell Death Dis. 2020, 11, 618. [Google Scholar] [CrossRef] [PubMed]
  68. Mei, Z.; Su, T.; Ye, J.; Yang, C.; Zhang, S.; Xie, C. The miR-15 Family Enhances the Radiosensitivity of Breast Cancer Cells by Targeting G2 Checkpoints. Radiat. Res. 2015, 183, 196–207. [Google Scholar] [CrossRef] [PubMed]
  69. Xie, Y.; Wei, R.-R.; Huang, G.-L.; Zhang, M.-Y.; Yuan, Y.-F.; Wang, H.-Y. Checkpoint kinase 1 is negatively regulated by miR-497 in hepatocellular carcinoma. Med. Oncol. 2014, 31, 844. [Google Scholar] [CrossRef] [PubMed]
  70. Liu, B.; Qu, J.; Xu, F.; Guo, Y.; Wang, Y.; Yu, H.; Qian, B. MiR-195 suppresses non-small cell lung cancer by targeting CHEK1. Oncotarget 2015, 6, 9445–9456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Huang, Y.-Z.; Zhang, J.; Shao, H.-Y.; Chen, J.-P.; Zhao, H.-Y. MicroRNA-191 promotes osteosarcoma cells proliferation by targeting checkpoint kinase 2. Tumor Biol. 2015, 36, 6095–6101. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, S.-M.; Chou, W.-C.; Hu, L.-Y.; Hsiung, C.-N.; Chu, H.-W.; Huang, Y.-L.; Hsu, H.-M.; Yu, J.-C.; Shen, C.-Y. The Effect of MicroRNA-124 Overexpression on Anti-Tumor Drug Sensitivity. PLoS ONE 2015, 10, e0128472. [Google Scholar] [CrossRef] [PubMed]
  73. Bisso, A.; Faleschini, M.; Zampa, F.; Capaci, V.; De Santa, J.; Santarpia, L.; Piazza, S.; Cappelletti, V.; Daidone, M.G.; Agami, R.; et al. Oncogenic miR-181a/b affect the DNA damage response in aggressive breast cancer. Cell Cycle 2013, 12, 1679–1687. [Google Scholar] [CrossRef] [PubMed]
  74. Moskwa, P.; Buffa, F.M.; Pan, Y.; Panchakshari, R.; Gottipati, P.; Muschel, R.J.; Beech, J.; Kulshrestha, R.; Abdelmohsen, K.; Weinstock, D.M.; et al. miR-182-Mediated Downregulation of BRCA1 Impacts DNA Repair and Sensitivity to PARP Inhibitors. Mol. Cell 2011, 41, 210–220. [Google Scholar] [CrossRef] [PubMed]
  75. He, X.; Xiao, X.; Dong, L.; Wan, N.; Zhou, Z.; Deng, H.; Zhang, X. MiR-218 regulates cisplatin chemosensitivity in breast cancer by targeting BRCA1. Tumor Biol. 2014, 36, 2065–2075. [Google Scholar] [CrossRef]
  76. Huang, J.-W.; Wang, Y.; Dhillon, K.K.; Calses, P.; Villegas, E.; Mitchell, P.S.; Tewari, M.; Kemp, C.J.; Taniguchi, T. Systematic Screen Identifies miRNAs That Target RAD51 and RAD51D to Enhance Chemosensitivity. Mol. Cancer Res. 2013, 11, 1564–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Espinosa-Diez, C.; Wilson, R.; Chatterjee, N.; Hudson, C.; Ruhl, R.; Hipfinger, C.; Helms, E.; Khan, O.F.; Anderson, D.G.; Anand, S. MicroRNA regulation of the MRN complex impacts DNA damage, cellular senescence, and angiogenic signaling. Cell Death Dis. 2018, 9, 632. [Google Scholar] [CrossRef]
  78. Chernyy, V.; Pustylnyak, V.; Kozlov, V.; Gulyaeva, L. Increased expression of miR-155 and miR-222 is associated with lymph node positive status. J. Cancer 2018, 9, 135–140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Fang, H.; Xie, J.; Zhang, M.; Zhao, Z.; Wan, Y.; Yao, Y. miRNA-21 promotes proliferation and invasion of triple-negative breast cancer cells through targeting PTEN. Am. J. Transl. Res. 2017, 9, 953–961. [Google Scholar] [PubMed]
  80. Piasecka, D.; Braun, M.; Kordek, R.; Sadej, R.; Romanska, H. MicroRNAs in regulation of triple-negative breast cancer progression. J. Cancer Res. Clin. Oncol. 2018, 144, 1401–1411. [Google Scholar] [CrossRef] [Green Version]
  81. Xiao, Y.; Li, Y.; Tao, H.; Humphries, B.; Li, A.; Jiang, Y.; Yang, C.; Luo, R.; Wang, Z. Integrin α5 down-regulation by miR-205 suppresses triple negative breast cancer stemness and metastasis by inhibiting the Src/Vav2/Rac1 pathway. Cancer Lett. 2018, 433, 199–209. [Google Scholar] [CrossRef] [PubMed]
  82. Damiano, V.; Brisotto, G.; Borgna, S.; di Gennaro, A.; Armellin, M.; Perin, T.; Guardascione, M.; Maestro, R.; Santarosa, M. Epigenetic silencing of miR-200c in breast cancer is associated with aggressiveness and is modulated by ZEB1. Genes Chromosom. Cancer 2016, 56, 147–158. [Google Scholar] [CrossRef] [PubMed]
  83. D’Ippolito, E.; Iorio, M.V. MicroRNAs and Triple Negative Breast Cancer. Int. J. Mol. Sci. 2013, 14, 22202–22220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Piva, R.; Spandidos, D.A.; Gambari, R. From microRNA functions to microRNA therapeutics: Novel targets and novel drugs in breast cancer research and treatment (Review). Int. J. Oncol. 2013, 43, 985–994. [Google Scholar] [CrossRef] [Green Version]
  85. Li, Q.; Liu, J.; Jia, Y.; Li, T.; Zhang, M. miR-623 suppresses cell proliferation, migration and invasion through direct inhibition of XRCC5 in breast cancer. Aging 2020, 12, 10246–10258. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, Z.-Y.; Fu, S.-L.; Xu, S.-Q.; Zhou, X.; Liu, X.-S.; Xu, Y.-J.; Zhao, J.-P.; Wei, S. By downregulating Ku80, hsa-miR-526b suppresses non-small cell lung cancer. Oncotarget 2014, 6, 1462–1477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Vogelstein, B.; Papadopoulos, N.; Velculescu, V.E.; Zhou, S.; Diaz, L.A., Jr.; Kinzler, K.W. Cancer Genome Landscapes. Science 2013, 339, 1546–1558. [Google Scholar] [CrossRef]
  88. Huang, D.; Savage, S.R.; Calinawan, A.P.; Lin, C.; Zhang, B.; Wang, P.; Starr, T.K.; Birrer, M.J.; Paulovich, A.G. A highly annotated database of genes associated with platinum resistance in cancer. Oncogene 2021, 40, 6395–6405. [Google Scholar] [CrossRef]
  89. O’Connor, M.J. Targeting the DNA Damage Response in Cancer. Mol. Cell 2015, 60, 547–560. [Google Scholar] [CrossRef] [Green Version]
  90. Reinhardt, H.C.; Jiang, H.; Hemann, M.T.; Yaffe, M.B. Exploiting synthetic lethal interactions for targeted cancer therapy. Cell Cycle 2009, 8, 3112–3119. [Google Scholar] [CrossRef] [Green Version]
  91. Coleman, R.L.; Oza, A.M.; Lorusso, D.; Aghajanian, C.; Oaknin, A.; Dean, A.; Colombo, N.; Weberpals, J.I.; Clamp, A.; Scambia, G.; et al. Rucaparib maintenance treatment for recurrent ovarian carcinoma after response to platinum therapy (ARIEL3): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet 2017, 390, 1949–1961. [Google Scholar] [CrossRef] [PubMed]
  92. Heo, Y.-A.; Duggan, S.T. Niraparib: A Review in Ovarian Cancer. Target. Oncol. 2018, 13, 533–539. [Google Scholar] [CrossRef] [PubMed]
  93. Hoy, S.M. Talazoparib: First Global Approval. Drugs 2018, 78, 1939–1946. [Google Scholar] [CrossRef] [PubMed]
  94. Litton, J.K.; Rugo, H.S.; Ettl, J.; Hurvitz, S.A.; Gonçalves, A.; Lee, K.-H.; Fehrenbacher, L.; Yerushalmi, R.; Mina, L.A.; Martin, M.; et al. Talazoparib in Patients with Advanced Breast Cancer and a Germline BRCA Mutation. N. Engl. J. Med. 2018, 379, 753–763. [Google Scholar] [CrossRef]
  95. Boussios, S.; Karihtala, P.; Moschetta, M.; Abson, C.; Karathanasi, A.; Zakynthinakis-Kyriakou, N.; Ryan, J.E.; Sheriff, M.; Rassy, E.; Pavlidis, N. Veliparib in ovarian cancer: A new synthetically lethal therapeutic approach. Investig. New Drugs 2019, 38, 181–193. [Google Scholar] [CrossRef] [PubMed]
  96. Nishio, S.; Takekuma, M.; Takeuchi, S.; Kawano, K.; Tsuda, N.; Tasaki, K.; Takahashi, N.; Abe, M.; Tanaka, A.; Nagasawa, T.; et al. Phase 1 study of veliparib with carboplatin and weekly paclitaxel in Japanese patients with newly diagnosed ovarian cancer. Cancer Sci. 2017, 108, 2213–2220. [Google Scholar] [CrossRef] [Green Version]
  97. Isakoff, S.J.; Puhalla, S.; Domchek, S.M.; Friedlander, M.; Kaufman, B.; Robson, M.; Telli, M.L.; Diéras, V.; Han, H.S.; Garber, E.J.; et al. A randomized Phase II study of veliparib with temozolomide or carboplatin/paclitaxel versus placebo with carboplatin/paclitaxel in BRCA1/2 metastatic breast cancer: Design and rationale. Futur. Oncol. 2017, 13, 307–320. [Google Scholar] [CrossRef] [Green Version]
  98. Ramalingam, S.S.; Blais, N.; Mazieres, J.; Reck, M.; Jones, C.M.; Juhasz, E.; Urban, L.; Orlov, S.; Barlesi, F.; Kio, E.; et al. Randomized, Placebo-Controlled, Phase II Study of Veliparib in Combination with Carboplatin and Paclitaxel for Advanced/Metastatic Non–Small Cell Lung Cancer. Clin. Cancer Res. 2017, 23, 1937–1944. [Google Scholar] [CrossRef] [Green Version]
  99. Klinakis, A.; Karagiannis, D.; Rampias, T. Targeting DNA repair in cancer: Current state and novel approaches. Cell Mol. Life Sci. 2020, 77, 677–703. [Google Scholar] [CrossRef]
  100. Tookman, L.; Krell, J.; Nkolobe, B.; Burley, L.; McNeish, I. Practical guidance for the management of side effects during rucaparib therapy in a multidisciplinary UK setting. Ther. Adv. Med. Oncol. 2020, 12, 1758835920921980. [Google Scholar] [CrossRef]
  101. Markman, M. Maintenance chemotherapy in the management of epithelial ovarian cancer. Cancer Metastasis Rev. 2015, 34, 11–17. [Google Scholar] [CrossRef] [PubMed]
  102. Electronic Medicines Compendium (EMC). Lynparza. Datapharm. Available online: https://www.medicines.org.uk/emc/search?q=%22olaparib%22 (accessed on 3 February 2020).
  103. Electronic Medicines Compendium (EMC). Rubraca. Datapharm. Available online: https://www.medicines.org.uk/emc/search?q=%22Rubraca%22 (accessed on 3 February 2020).
  104. Moore, K.; Colombo, N.; Scambia, G.; Kim, B.-G.; Oaknin, A.; Friedlander, M.; Lisyanskaya, A.; Floquet, A.; Leary, A.; Sonke, G.S.; et al. Maintenance Olaparib in Patients with Newly Diagnosed Advanced Ovarian Cancer. N. Engl. J. Med. 2018, 379, 2495–2505. [Google Scholar] [CrossRef] [PubMed]
  105. Ray-Coquard, I.; Pautier, P.; Pignata, S.; Pérol, D.; González-Martín, A.; Berger, R.; Fujiwara, K.; Vergote, I.; Colombo, N.; Mäenpää, J.; et al. Olaparib plus Bevacizumab as First-Line Maintenance in Ovarian Cancer. N. Engl. J. Med. 2019, 381, 2416–2428. [Google Scholar] [CrossRef] [PubMed]
  106. González-Martín, A.; Pothuri, B.; Vergote, I.; DePont Christensen, R.; Graybill, W.; Mirza, M.R.; McCormick, C.; Lorusso, D.; Hoskins, P.; Freyer, G.; et al. Niraparib in Patients with Newly Diagnosed Advanced Ovarian Cancer. N. Engl. J. Med. 2019, 381, 2391–2402. [Google Scholar] [CrossRef] [Green Version]
  107. Coleman, R.L.; Fleming, G.F.; Brady, M.F.; Swisher, E.M.; Steffensen, K.D.; Friedlander, M.; Okamoto, A.; Moore, K.N.; Efrat Ben-Baruch, N.; Werner, T.L.; et al. Veliparib with First-Line Chemotherapy and as Maintenance Therapy in Ovarian Cancer. N. Engl. J. Med. 2019, 381, 2403–2415. [Google Scholar] [CrossRef]
  108. Pommier, Y.; O’Connor, M.J.; de Bono, J. Laying a trap to kill cancer cells: PARP inhibitors and their mechanisms of action. Sci. Transl. Med. 2016, 8, 362ps17. [Google Scholar] [CrossRef]
  109. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.J.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef]
  110. Bryant, H.E.; Schultz, N.; Thomas, H.D.; Parker, K.M.; Flower, D.; Lopez, E.; Kyle, S.; Meuth, M.; Curtin, N.J.; Helleday, T. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 2005, 434, 913–917. [Google Scholar] [CrossRef]
  111. Fong, P.C.; Boss, D.S.; Yap, T.A.; Tutt, A.; Wu, P.; Mergui-Roelvink, M.; Mortimer, P.; Swaisland, H.; Lau, A.; O’Connor, M.J.; et al. Inhibition of Poly(ADP-Ribose) Polymerase in Tumors from BRCA Mutation Carriers. N. Engl. J. Med. 2009, 361, 123–134. [Google Scholar] [CrossRef] [Green Version]
  112. Gelmon, K.A.; Tischkowitz, M.; Mackay, H.; Swenerton, K.; Robidoux, A.; Tonkin, K.; Hirte, H.; Huntsman, D.; Clemons, M.; Gilks, B.; et al. Olaparib in patients with recurrent high-grade serous or poorly differentiated ovarian carcinoma or triple-negative breast cancer: A phase 2, multicentre, open-label, non-randomised study. Lancet Oncol. 2011, 12, 852–861. [Google Scholar] [CrossRef]
  113. Gourley, C.; Balmaña, J.; Ledermann, J.A.; Serra, V.; Dent, R.; Loibl, S.; Pujade-Lauraine, E.; Boulton, S.J. Moving from Poly (ADP-Ribose) Polymerase Inhibition to Targeting DNA Repair and DNA Damage Response in Cancer Therapy. J. Clin. Oncol. 2019, 37, 2257–2269. [Google Scholar] [CrossRef] [PubMed]
  114. Pearl, L.H.; Schierz, A.C.; Ward, S.E.; Al-Lazikani, B.; Pearl, F.M. Therapeutic opportunities within the DNA damage response. Nat. Rev. Cancer 2015, 15, 166–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. London, R.E. The structural basis of XRCC1-mediated DNA repair. DNA Repair 2015, 30, 90–103. [Google Scholar] [CrossRef] [Green Version]
  116. Ha, K.; Fiskus, W.; Choi, D.S.; Bhaskara, S.; Cerchietti, L.; Devaraj, S.G.T.; Shah, B.; Sharma, S.; Chang, J.C.; Melnick, A.M.; et al. Histone deacetylase inhibitor treatment induces ‘BRCAness’ and synergistic lethality with PARP inhibitor and cisplatin against human triple negative breast cancer cells. Oncotarget 2014, 5, 5637–5650. [Google Scholar] [CrossRef] [Green Version]
  117. Konstantinopoulos, P.A.; Wilson, A.J.; Saskowski, J.; Wass, E.; Khabele, D. Suberoylanilide hydroxamic acid (SAHA) enhances olaparib activity by targeting homologous recombination DNA repair in ovarian cancer. Gynecol. Oncol. 2014, 133, 599–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Marijon, H.; Lee, D.H.; Ding, L.-W.; Sun, H.; Gery, S.; de Gramont, A.; Koeffler, H.P. Co-targeting poly(ADP-ribose) polymerase (PARP) and histone deacetylase (HDAC) in triple-negative breast cancer: Higher synergism in BRCA mutated cells. Biomed. Pharmacother. 2018, 99, 543–551. [Google Scholar] [CrossRef] [PubMed]
  119. Yin, L.; Liu, Y.; Peng, Y.; Peng, Y.; Yu, X.; Gao, Y.; Yuan, B.; Zhu, Q.; Cao, T.; He, L.; et al. PARP inhibitor veliparib and HDAC inhibitor SAHA synergistically co-target the UHRF1/BRCA1 DNA damage repair complex in prostate cancer cells. J. Exp. Clin. Cancer Res. 2018, 37, 153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Muvarak, N.E.; Chowdhury, K.; Xia, L.; Robert, C.; Choi, E.Y.; Cai, Y.; Bellani, M.; Zou, Y.; Singh, Z.N.; Duong, V.H.; et al. Enhancing the Cytotoxic Effects of PARP Inhibitors with DNA Demethylating Agents—A Potential Therapy for Cancer. Cancer Cell 2016, 30, 637–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Thurn, K.T.; Thomas, S.; Raha, P.; Qureshi, I.; Munster, P.N. Histone Deacetylase Regulation of ATM-Mediated DNA Damage Signaling. Mol. Cancer Ther. 2013, 12, 2078–2087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Zhang, Y.-W.; Zheng, Y.; Wang, J.-Z.; Lu, X.-X.; Wang, Z.; Chen, L.-B.; Guan, X.-X.; Tong, J.-D. Integrated analysis of DNA methylation and mRNA expression profiling reveals candidate genes associated with cisplatin resistance in non-small cell lung cancer. Epigenetics 2014, 9, 896–909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Camphausen, K.; Tofilon, P.J. Inhibition of Histone Deacetylation: A Strategy for Tumor Radiosensitization. J. Clin. Oncol. 2007, 25, 4051–4056. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, Y.; Carr, T.; Dimtchev, A.; Zaer, N.; Dritschilo, A.; Jung, M. Attenuated DNA Damage Repair by Trichostatin A through BRCA1 Suppression. Radiat. Res. 2007, 168, 115–124. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, J.; Wang, Y.; Mei, H.; Yin, Z.; Geng, Y.; Zhang, T.; Wu, G.; Lin, Z. The BET bromodomain inhibitor JQ1 radiosensitizes non-small cell lung cancer cells by upregulating p21. Cancer Lett. 2017, 391, 141–151. [Google Scholar] [CrossRef] [PubMed]
  126. Wu, C.; Jin, X.; Yang, J.; Yang, Y.; He, Y.; Ding, L.; Pan, Y.; Chen, S.; Jiang, J.; Huang, H. Inhibition of EZH2 by chemo- and radiotherapy agents and small molecule inhibitors induces cell death in castration-resistant prostate cancer. Oncotarget 2015, 7, 3440–3452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Zhu, X.; Wang, Y.; Tan, L.; Fu, X. The pivotal role of DNA methylation in the radio-sensitivity of tumor radiotherapy. Cancer Med. 2018, 7, 3812–3819. [Google Scholar] [CrossRef] [PubMed]
  128. Ahn, D.H.; Bekaii-Saab, T. Biliary tract cancer and genomic alterations in homologous recombinant deficiency: Exploiting synthetic lethality with PARP inhibitors. Chin. Clin. Oncol. 2020, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  129. Kim, H.; George, E.; Ragland, R.L.; Rafail, S.; Zhang, R.; Krepler, C.; Morgan, M.A.; Herlyn, M.; Brown, E.J.; Simpkins, F. Targeting the ATR/CHK1 Axis with PARP Inhibition Results in Tumor Regression in BRCA-Mutant Ovarian Cancer Models. Clin. Cancer Res. 2017, 23, 3097–3108. [Google Scholar] [CrossRef] [Green Version]
  130. Carrassa, L.; Chilà, R.; Lupi, M.; Ricci, F.; Celenza, C.; Mazzoletti, M.; Broggini, M.; Damia, G. Combined inhibition of Chk1 and Wee1: In vitro synergistic effect translates to tumor growth inhibition in vivo. Cell Cycle 2012, 11, 2507–2517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Restelli, V.; Vagni, M.; Arribas, A.J.; Bertoni, F.; Damia, G.; Carrassa, L. Inhibition of CHK 1 and WEE 1 as a new therapeutic approach in diffuse large B cell lymphomas with MYC deregulation. Br. J. Haematol. 2016, 181, 129–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Chilà, R.; Basana, A.; Lupi, M.; Guffanti, F.; Gaudio, E.; Rinaldi, A.; Cascione, L.; Restelli, V.; Tarantelli, C.; Bertoni, F.; et al. Combined inhibition of Chk1 and Wee1 as a new therapeutic strategy for mantle cell lymphoma. Oncotarget 2014, 6, 3394–3408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Davies, K.D.; Cable, P.L.; Garrus, J.E.; Sullivan, F.X.; Von Carlowitz, I.; Le Huérou, Y.; Wallace, E.; Woessner, R.D.; Gross, S. Chk1 inhibition and Wee1 inhibition combine synergistically to impede cellular proliferation. Cancer Biol. Ther. 2011, 12, 788–796. [Google Scholar] [CrossRef] [PubMed]
  134. Restelli, V.; Lupi, M.; Chilà, R.; Vagni, M.; Tarantelli, C.; Spriano, F.; Gaudio, E.; Bertoni, F.; Damia, G.; Carrassa, L. DNA Damage Response Inhibitor Combinations Exert Synergistic Antitumor Activity in Aggressive B-Cell Lymphomas. Mol. Cancer Ther. 2019, 18, 1255–1264. [Google Scholar] [CrossRef] [Green Version]
  135. Sanjiv, K.; Hagenkort, A.; Calderón-Montaño, J.M.; Koolmeister, T.; Reaper, P.M.; Mortusewicz, O.; Jacques, S.A.; Kuiper, R.V.; Schultz, N.; Scobie, M.; et al. Cancer-Specific Synthetic Lethality between ATR and CHK1 Kinase Activities. Cell Rep. 2016, 14, 298–309. [Google Scholar] [CrossRef] [PubMed]
  136. Ha, D.-H.; Min, A.; Kim, S.; Jang, H.; Kim, S.H.; Kim, H.-J.; Ryu, H.S.; Ku, J.-L.; Lee, K.-H.; Im, S.-A. Antitumor effect of a WEE1 inhibitor and potentiation of olaparib sensitivity by DNA damage response modulation in triple-negative breast cancer. Sci. Rep. 2020, 10, 9930. [Google Scholar] [CrossRef]
  137. Jin, J.; Fang, H.; Yang, F.; Ji, W.; Guan, N.; Sun, Z.; Shi, Y.; Zhou, G.; Guan, X. Combined Inhibition of ATR and WEE1 as a Novel Therapeutic Strategy in Triple-Negative Breast Cancer. Neoplasia 2018, 20, 478–488. [Google Scholar] [CrossRef] [PubMed]
  138. Bukhari, A.; Lewis, C.W.; Pearce, J.J.; Luong, D.; Chan, G.K.; Gamper, A.M. Inhibiting Wee1 and ATR kinases produces tumor-selective synthetic lethality and suppresses metastasis. J. Clin. Investig. 2019, 129, 1329–1344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Kim, H.; Xu, H.; George, E.; Hallberg, D.; Kumar, S.; Jagannathan, V.; Medvedev, S.; Kinose, Y.; Devins, K.; Verma, P.; et al. Combining PARP with ATR inhibition overcomes PARP inhibitor and platinum resistance in ovarian cancer models. Nat. Commun. 2020, 11, 3726. [Google Scholar] [CrossRef] [PubMed]
  140. Saxton, R.A.; Sabatini, D.M. mTOR Signaling in Growth, Metabolism, and Disease. Cell 2017, 169, 361–371. [Google Scholar] [CrossRef] [PubMed]
  141. Liu, G.Y.; Sabatini, D.M. mTOR at the nexus of nutrition, growth, ageing and disease. Nat. Rev. Mol. Cell Biol. 2020, 21, 183–203. [Google Scholar] [CrossRef]
  142. Ilagan, E.; Manning, B.D. Emerging Role of mTOR in the Response to Cancer Therapeutics. Trends Cancer 2016, 2, 241–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Ma, Y.; Vassetzky, Y.; Dokudovskaya, S. mTORC1 pathway in DNA damage response. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1293–1311. [Google Scholar] [CrossRef]
  144. Bandhakavi, S.; Kim, Y.-M.; Ro, S.-H.; Xie, H.; Onsongo, G.; Jun, C.-B.; Kim, D.-H.; Griffin, T.J. Quantitative Nuclear Proteomics Identifies mTOR Regulation of DNA Damage Response. Mol. Cell. Proteom. 2010, 9, 403–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chen, H.; Ma, Z.; Vanderwaal, R.P.; Feng, Z.; Gonzalez-Suarez, I.; Wang, S.; Zhang, J.; Roti, J.L.R.; Gonzalo, S.; Zhang, J. The mTOR Inhibitor Rapamycin Suppresses DNA Double-Strand Break Repair. Radiat. Res. 2010, 175, 214–224. [Google Scholar] [CrossRef] [Green Version]
  146. Pópulo, H.; Lopes, J.M.; Soares, P. The mTOR Signalling Pathway in Human Cancer. Int. J. Mol. Sci. 2012, 13, 1886–1918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Walsh, S.; Flanagan, L.; Quinn, C.; Evoy, D.; McDermott, E.; Pierce, A.; Duffy, M. mTOR in breast cancer: Differential expression in triple-negative and non-triple-negative tumors. Breast 2012, 21, 178–182. [Google Scholar] [CrossRef] [PubMed]
  148. Montero, J.C.; Esparís-Ogando, A.; Re-Louhau, M.F.; Seoane, S.; Abad, M.; Calero, R.; Ocaña, A.; Pandiella, A. Active kinase profiling, genetic and pharmacological data define mTOR as an important common target in triple-negative breast cancer. Oncogene 2012, 33, 148–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Guri, Y.; Hall, M.N. mTOR Signaling Confers Resistance to Targeted Cancer Drugs. Trends Cancer 2016, 2, 688–697. [Google Scholar] [CrossRef]
  150. Germano, G.; Lamba, S.; Rospo, G.; Barault, L.; Magrì, A.; Maione, F.; Russo, M.; Crisafulli, G.; Bartolini, A.; Lerda, G.; et al. Inactivation of DNA repair triggers neoantigen generation and impairs tumour growth. Nature 2017, 552, 116–120. [Google Scholar] [CrossRef]
  151. Zhao, P.; Li, L.; Jiang, X.; Li, Q. Mismatch repair deficiency/microsatellite instability-high as a predictor for anti-PD-1/PD-L1 immunotherapy efficacy. J. Hematol. Oncol. 2019, 12, 54. [Google Scholar] [CrossRef]
  152. Wang, B.; Qiao, L.; Wang, Y.; Zeng, J.; Chen, D.; Guo, H.; Zhang, Y. Mitochondrial DNA D-loop lesions with the enhancement of DNA repair contribute to gastrointestinal cancer progression. Oncol. Rep. 2018, 40, 3694–3704. [Google Scholar] [CrossRef] [Green Version]
  153. Ueta, E.; Sasabe, E.; Yang, Z.; Osaki, T.; Yamamoto, T. Enhancement of apoptotic damage of squamous cell carcinoma cells by inhibition of the mitochondrial DNA repairing system. Cancer Sci. 2008, 99, 2230–2237. [Google Scholar] [CrossRef] [PubMed]
  154. Shokolenko, I.N.; Alexeyev, M.F.; Robertson, F.M.; LeDoux, S.P.; Wilson, G.L. The expression of Exonuclease III from E. coli in mitochondria of breast cancer cells diminishes mitochondrial DNA repair capacity and cell survival after oxidative stress. DNA Repair 2003, 2, 471–482. [Google Scholar] [CrossRef] [PubMed]
  155. Chandran, E.A.; Kennedy, I. Significant Tumor Response to the Poly (ADP-ribose) Polymerase Inhibitor Olaparib in Heavily Pretreated Patient with Ovarian Carcinosarcoma Harboring a Germline RAD51D Mutation. JCO Precis. Oncol. 2018, 2, 1–4. [Google Scholar] [CrossRef] [PubMed]
  156. Gou, R.; Dong, H.; Lin, B. Application and reflection of genomic scar assays in evaluating the efficacy of platinum salts and PARP inhibitors in cancer therapy. Life Sci. 2020, 261, 118434. [Google Scholar] [CrossRef] [PubMed]
  157. Teyssonneau, D.; Margot, H.; Cabart, M.; Anonnay, M.; Sargos, P.; Vuong, N.-S.; Soubeyran, I.; Sevenet, N.; Roubaud, G. Prostate cancer and PARP inhibitors: Progress and challenges. J. Hematol. Oncol. 2021, 14, 51. [Google Scholar] [CrossRef]
  158. Cleary, J.M.; Aguirre, A.J.; Shapiro, G.I.; D’Andrea, A.D. Biomarker-Guided Development of DNA Repair Inhibitors. Mol. Cell 2020, 78, 1070–1085. [Google Scholar] [CrossRef]
  159. Lord, C.J.; Ashworth, A. BRCAness revisited. Nat. Rev. Cancer 2016, 16, 110–120. [Google Scholar] [CrossRef]
  160. Kondrashova, O.; Topp, M.; Nesic, K.; Lieschke, E.; Ho, G.Y.; Harrell, M.I.; Zapparoli, G.V.; Hadley, A.; Holian, R.; Boehm, E.; et al. Methylation of all BRCA1 copies predicts response to the PARP inhibitor rucaparib in ovarian carcinoma. Nat. Commun. 2018, 9, 3970. [Google Scholar] [CrossRef] [Green Version]
  161. Swisher, E.M.; Lin, K.K.; Oza, A.M.; Scott, C.L.; Giordano, H.; Sun, J.; Konecny, G.E.; Coleman, R.L.; Tinker, A.V.; O’Malley, D.M.; et al. Rucaparib in relapsed, platinum-sensitive high-grade ovarian carcinoma (ARIEL2 Part 1): An international, multicentre, open-label, phase 2 trial. Lancet Oncol. 2017, 18, 75–87. [Google Scholar] [CrossRef] [Green Version]
  162. Pennington, K.P.; Walsh, T.; Harrell, M.I.; Lee, M.K.; Pennil, C.C.; Rendi, M.H.; Thornton, A.; Norquist, B.M.; Casadei, S.; Nord, A.S.; et al. Germline and Somatic Mutations in Homologous Recombination Genes Predict Platinum Response and Survival in Ovarian, Fallopian Tube, and Peritoneal Carcinomas. Clin. Cancer Res. 2014, 20, 764–775. [Google Scholar] [CrossRef] [Green Version]
  163. Arora, S.; Balasubramaniam, S.; Zhang, H.; Berman, T.; Narayan, P.; Suzman, D.; Bloomquist, E.; Tang, S.; Gong, Y.; Sridhara, R.; et al. FDA Approval Summary: Olaparib Monotherapy or in Combination with Bevacizumab for the Maintenance Treatment of Patients with Advanced Ovarian Cancer. Oncologist 2020, 26, e164–e172. [Google Scholar] [CrossRef] [PubMed]
  164. Minchom, A.; Aversa, C.; Lopez, J. Dancing with the DNA damage response: Next-generation anti-cancer therapeutic strategies. Ther. Adv. Med Oncol. 2018, 10, 1758835918786658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Smith, M.A.; Hampton, O.A.; Reynolds, C.P.; Kang, M.H.; Maris, J.M.; Gorlick, R.; Kolb, E.A.; Lock, R.; Carol, H.; Keir, S.T.; et al. Initial testing (stage 1) of the PARP inhibitor BMN 673 by the pediatric preclinical testing program: PALB2 mutation predicts exceptional in vivo response to BMN 673. Pediatr. Blood Cancer 2014, 62, 91–98. [Google Scholar] [CrossRef] [Green Version]
  166. Tentori, L.; Ricci-Vitiani, L.; Muzi, A.; Ciccarone, F.; Pelacchi, F.; Calabrese, R.; Runci, D.; Pallini, R.; Caiafa, P.; Graziani, G. Pharmacological inhibition of poly(ADP-ribose) polymerase-1 modulates resistance of human glioblastoma stem cells to temozolomide. BMC Cancer 2014, 14, 151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Patch, A.-M.; Christie, E.L.; Etemadmoghadam, D.; Garsed, D.W.; George, J.; Fereday, S.; Nones, K.; Cowin, P.; Alsop, K.; Bailey, P.J.; et al. Whole–genome characterization of chemoresistant ovarian cancer. Nature 2015, 521, 489–494. [Google Scholar] [CrossRef] [PubMed]
  168. Robinson, D.; Van Allen, E.M.; Wu, Y.-M.; Schultz, N.; Lonigro, R.J.; Mosquera, J.-M.; Montgomery, B.; Taplin, M.-E.; Pritchard, C.C.; Attard, G.; et al. Integrative Clinical Genomics of Advanced Prostate Cancer. Cell 2015, 162, 454. [Google Scholar] [CrossRef] [Green Version]
  169. Miller, R.; Leary, A.; Scott, C.; Serra, V.; Lord, C.; Bowtell, D.; Chang, D.; Garsed, D.; Jonkers, J.; Ledermann, J.; et al. ESMO recommendations on predictive biomarker testing for homologous recombination deficiency and PARP inhibitor benefit in ovarian cancer. Ann. Oncol. 2020, 31, 1606–1622. [Google Scholar] [CrossRef] [PubMed]
  170. Abkevich, V.; Timms, K.M.; Hennessy, B.T.; Potter, J.; Carey, M.S.; Meyer, L.A.; Smith-McCune, K.; Broaddus, R.; Lu, K.H.; Chen, J.; et al. Patterns of genomic loss of heterozygosity predict homologous recombination repair defects in epithelial ovarian cancer. Br. J. Cancer 2012, 107, 1776–1782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Watkins, J.A.; Irshad, S.; Grigoriadis, A.; Tutt, A.N.J. Genomic scars as biomarkers of homologous recombination deficiency and drug response in breast and ovarian cancers. Breast Cancer Res. 2014, 16, 211. [Google Scholar] [CrossRef] [Green Version]
  172. Helleday, T.; Eshtad, S.; Nik-Zainal, S. Mechanisms underlying mutational signatures in human cancers. Nat. Rev. Genet. 2014, 15, 585–598. [Google Scholar] [CrossRef]
  173. Popova, T.; Manié, E.; Rieunier, G.; Caux-Moncoutier, V.; Tirapo, C.; Dubois, T.; Delattre, O.; Sigal-Zafrani, B.; Bollet, M.; Longy, M.; et al. Ploidy and Large-Scale Genomic Instability Consistently Identify Basal-like Breast Carcinomas with BRCA1/2 Inactivation. Cancer Res. 2012, 72, 5454–5462. [Google Scholar] [CrossRef] [PubMed]
  174. Birkbak, N.J.; Wang, Z.C.; Kim, J.-Y.; Eklund, A.C.; Li, Q.; Tian, R.; Bowman-Colin, C.; Li, Y.; Greene-Colozzi, A.; Iglehart, J.D.; et al. Telomeric Allelic Imbalance Indicates Defective DNA Repair and Sensitivity to DNA-Damaging Agents. Cancer Discov. 2012, 2, 366–375. [Google Scholar] [CrossRef] [Green Version]
  175. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Aparicio, S.A.J.R.; Behjati, S.; Biankin, A.V.; Bignell, G.R.; Bolli, N.; Borg, A.; Børresen-Dale, A.-L.; et al. Signatures of mutational processes in human cancer. Nature 2013, 500, 415–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Choi, W.; Lee, E.S. Therapeutic Targeting of DNA Damage Response in Cancer. Int. J. Mol. Sci. 2022, 23, 1701. [Google Scholar] [CrossRef] [PubMed]
  177. Polak, P.; Kim, J.; Braunstein, L.Z.; Karlic, R.; Haradhavala, N.J.; Tiao, G.; Rosebrock, D.; Livitz, D.; Kübler, K.; Mouw, K.W.; et al. A mutational signature reveals alterations underlying deficient homologous recombination repair in breast cancer. Nat. Genet. 2017, 49, 1476–1486. [Google Scholar] [CrossRef]
  178. Toledo, I.L.I.; Murga, M.; Zur, R.; Soria, R.; Rodriguez, A.; Martinez, S.; Oyarzabal, J.; Pastor, J.; Bischoff, J.R.; Fernandez-Capetillo, O. A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nat. Struct. Mol. Biol. 2011, 18, 721–727. [Google Scholar] [CrossRef] [Green Version]
  179. Murga, M.; Campaner, S.; Lopez-Contreras, A.J.; Toledo, L.I.; Soria, R.; Montaña, M.F.; Artista, L.D.; Schleker, T.; Guerra, C.; Garcia, E.O.; et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nat. Struct. Mol. Biol. 2011, 18, 1331–1335. [Google Scholar] [CrossRef] [PubMed]
  180. Grabocka, E.; Commisso, C.; Bar-Sagi, D. Molecular Pathways: Targeting the Dependence of Mutant RAS Cancers on the DNA Damage Response. Clin. Cancer Res. 2015, 21, 1243–1247. [Google Scholar] [CrossRef] [Green Version]
  181. Pfister, S.X.; Markkanen, E.; Jiang, Y.; Sarkar, S.; Woodcock, M.; Orlando, G.; Mavrommati, I.; Pai, C.-C.; Zalmas, L.-P.; Drobnitzky, N.; et al. Inhibiting WEE1 Selectively Kills Histone H3K36me3-Deficient Cancers by dNTP Starvation. Cancer Cell 2015, 28, 557–568. [Google Scholar] [CrossRef] [Green Version]
  182. Pilié, P.; George, A.; Yap, T. Patient selection biomarker strategies for PARP inhibitor therapy. Ann. Oncol. 2020, 31, 1603–1605. [Google Scholar] [CrossRef] [PubMed]
  183. Paulet, L.; Trecourt, A.; Leary, A.; Peron, J.; Descotes, F.; Devouassoux-Shisheboran, M.; Leroy, K.; You, B.; Lopez, J. Cracking the homologous recombination deficiency code: How to identify responders to PARP inhibitors. Eur. J. Cancer 2022, 166, 87–99. [Google Scholar] [CrossRef]
  184. Chen, Z.; Sun, T.; Yang, Z.; Zheng, Y.; Yu, R.; Wu, X.; Yan, J.; Shao, Y.W.; Shao, X.; Cao, W.; et al. Monitoring treatment efficacy and resistance in breast cancer patients via circulating tumor DNA genomic profiling. Mol. Genet. Genom. Med. 2019, 8, e1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. DiBardino, D.M.; Rawson, D.W.; Saqi, A.; Heymann, J.J.; Pagan, C.A.; Bulman, W.A. Next-generation sequencing of non-small cell lung cancer using a customized, targeted sequencing panel: Emphasis on small biopsy and cytology. CytoJournal 2017, 14, 7. [Google Scholar] [CrossRef]
  186. Coyne, G.O.; Karlovich, C.; Wilsker, D.; Voth, A.R.; Parchment, E.R.; Chen, A.P.; Doroshow, J.H. PARP Inhibitor Applicability: Detailed Assays for Homologous Recombination Repair Pathway Components. OncoTargets Ther. 2022, 15, 165–180. [Google Scholar] [CrossRef] [PubMed]
  187. Lin, K.K.; Harrell, M.I.; Oza, A.M.; Oaknin, A.; Ray-Coquard, I.; Tinker, A.V.; Helman, E.; Radke, M.R.; Say, C.; Vo, L.-T.; et al. BRCA Reversion Mutations in Circulating Tumor DNA Predict Primary and Acquired Resistance to the PARP Inhibitor Rucaparib in High-Grade Ovarian Carcinoma. Cancer Discov. 2019, 9, 210–219. [Google Scholar] [CrossRef] [Green Version]
  188. Bakhoum, S.F.; Landau, D.A. Chromosomal Instability as a Driver of Tumor Heterogeneity and Evolution. Cold Spring Harb. Perspect. Med. 2017, 7, a029611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Jallepalli, P.V.; Lengauer, C. Chromosome segregation and cancer: Cutting through the mystery. Nat. Rev. Cancer 2001, 1, 109–117. [Google Scholar] [CrossRef]
  190. Lytle, N.K.; Barber, A.G.; Reya, T. Stem cell fate in cancer growth, progression and therapy resistance. Nat. Rev. Cancer 2018, 18, 669–680. [Google Scholar] [CrossRef]
  191. Galluzzi, L.; Senovilla, L.; Vitale, I.; Michels, J.; Martins, I.; Kepp, O.; Castedo, M.; Kroemer, G. Molecular mechanisms of cisplatin resistance. Oncogene 2011, 31, 1869–1883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Wang, L.-E.; Yin, M.; Dong, Q.; Stewart, D.J.; Merriman, K.W.; Amos, C.I.; Spitz, M.R.; Wei, Q. DNA Repair Capacity in Peripheral Lymphocytes Predicts Survival of Patients with Non–Small-Cell Lung Cancer Treated with First-Line Platinum-Based Chemotherapy. J. Clin. Oncol. 2011, 29, 4121–4128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Oliver, T.G.; Mercer, K.L.; Sayles, L.C.; Burke, J.R.; Mendus, D.; Lovejoy, K.S.; Cheng, M.-H.; Subramanian, A.; Mu, D.; Powers, S.; et al. Chronic cisplatin treatment promotes enhanced damage repair and tumor progression in a mouse model of lung cancer. Genes Dev. 2010, 24, 837–852. [Google Scholar] [CrossRef] [PubMed]
  194. Banerjee, S.; Kaye, S.B.; Ashworth, A. Making the best of PARP inhibitors in ovarian cancer. Nat. Rev. Clin. Oncol. 2010, 7, 508–519. [Google Scholar] [CrossRef]
  195. Friboulet, L.; Barrios-Gonzales, D.; Commo, F.; Olaussen, K.A.; Vagner, S.; Adam, J.; Goubar, A.; Dorvault, N.; Lazar, V.; Job, B.; et al. Molecular Characteristics of ERCC1-Negative versus ERCC1-Positive Tumors in Resected NSCLC. Clin. Cancer Res. 2011, 17, 5562–5572. [Google Scholar] [CrossRef] [Green Version]
  196. Zhang, J.; Stevens, M.F.; Bradshaw, T.D. Temozolomide: Mechanisms of Action, Repair and Resistance. Curr. Mol. Pharmacol. 2012, 5, 102–114. [Google Scholar] [CrossRef] [PubMed]
  197. Hegi, M.E.; Diserens, A.-C.; Gorlia, T.; Hamou, M.-F.; De Tribolet, N.; Weller, M.; Kros, J.M.; Hainfellner, J.A.; Mason, W.; Mariani, L.; et al. MGMT Gene Silencing and Benefit from Temozolomide in Glioblastoma. N. Engl. J. Med. 2005, 352, 997–1003. [Google Scholar] [CrossRef] [Green Version]
  198. Stupp, R.; Mason, W.P.; van den Bent, M.J.; Weller, M.; Fisher, B.; Taphoorn, M.J.B.; Belanger, K.; Brandes, A.A.; Marosi, C.; Bogdahn, U.; et al. Radiotherapy plus Concomitant and Adjuvant Temozolomide for Glioblastoma. N. Engl. J. Med. 2005, 352, 987–996. [Google Scholar] [CrossRef] [Green Version]
  199. Li, L.-Y.; Guan, Y.-D.; Chen, X.-S.; Yang, J.-M.; Cheng, Y. DNA Repair Pathways in Cancer Therapy and Resistance. Front. Pharmacol. 2021, 11, 629266. [Google Scholar] [CrossRef]
  200. Oldrini, B.; Vaquero-Siguero, N.; Mu, Q.; Kroon, P.; Zhang, Y.; Galán-Ganga, M.; Bao, Z.; Wang, Z.; Liu, H.; Sa, J.K.; et al. MGMT genomic rearrangements contribute to chemotherapy resistance in gliomas. Nat. Commun. 2020, 11, 3883. [Google Scholar] [CrossRef]
  201. Neboori, H.J.; Haffty, B.G.; Wu, H.; Yang, Q.; Aly, A.; Goyal, S.; Schiff, D.; Moran, M.S.; Golhar, R.; Chen, C.; et al. Low p53 Binding Protein 1 (53BP1) Expression Is Associated with Increased Local Recurrence in Breast Cancer Patients Treated with Breast-Conserving Surgery and Radiotherapy. Int. J. Radiat. Oncol. 2012, 83, e677–e683. [Google Scholar] [CrossRef]
  202. Ma, J.; Benitez, J.A.; Li, J.; Miki, S.; de Albuquerque, C.P.; Galatro, T.; Orellana, L.; Zanca, C.; Reed, R.; Boyer, A.; et al. Inhibition of Nuclear PTEN Tyrosine Phosphorylation Enhances Glioma Radiation Sensitivity through Attenuated DNA Repair. Cancer Cell 2019, 35, 504–518.e7. [Google Scholar] [CrossRef] [Green Version]
  203. Naidu, M.D.; Agarwal, R.; Peña, L.A.; Cunha, L.; Mezei, M.; Shen, M.; Wilson, D.M.; Liu, Y.; Sanchez, Z.; Chaudhary, P.; et al. Lucanthone and Its Derivative Hycanthone Inhibit Apurinic Endonuclease-1 (APE1) by Direct Protein Binding. PLoS ONE 2011, 6, e23679. [Google Scholar] [CrossRef]
  204. Li, Q.; Wei, X.; Zhou, Z.-W.; Wang, S.-N.; Jin, H.; Chen, K.-J.; Luo, J.; Westover, K.; Wang, J.-M.; Wang, D.; et al. GADD45α sensitizes cervical cancer cells to radiotherapy via increasing cytoplasmic APE1 level. Cell Death Dis. 2018, 9, 524. [Google Scholar] [CrossRef]
  205. Harima, Y.; Sawada, S.; Miyazaki, Y.; Kin, K.; Ishihara, H.; Imamura, M.; Sougawa, M.; Shikata, N.; Ohnishi, T. Expression of Ku80 in Cervical Cancer Correlates with Response to Radiotherapy and Survival. Am. J. Clin. Oncol. 2003, 26, e80–e85. [Google Scholar] [CrossRef]
  206. Hayashi, J.; Sakata, K.-I.; Someya, M.; Matsumoto, Y.; Satoh, M.; Nakata, K.; Hori, M.; Takagi, M.; Kondoh, A.; Himi, T.; et al. Analysis and results of Ku and XRCC4 expression in hypopharyngeal cancer tissues treated with chemoradiotherapy. Oncol. Lett. 2012, 4, 151–155. [Google Scholar] [CrossRef] [Green Version]
  207. Beskow, C.; Skikuniene, J.; Holgersson, Å.; Nilsson, B.; Lewensohn, R.; Kanter, L.; Viktorsson, K. Radioresistant cervical cancer shows upregulation of the NHEJ proteins DNA-PKcs, Ku70 and Ku86. Br. J. Cancer 2009, 101, 816–821. [Google Scholar] [CrossRef] [Green Version]
  208. Liu, C.; Liao, K.; Gross, N.; Wang, Z.; Li, G.; Zuo, W.; Zhong, S.; Zhang, Z.; Zhang, H.; Yang, J.; et al. Homologous recombination enhances radioresistance in hypopharyngeal cancer cell line by targeting DNA damage response. Oral Oncol. 2019, 100, 104469. [Google Scholar] [CrossRef]
  209. Wang, Z.; Zuo, W.; Zeng, Q.; Li, Y.; Lu, T.; Bu, Y.; Hu, G. The Homologous Recombination Repair Pathway is Associated with Resistance to Radiotherapy in Nasopharyngeal Carcinoma. Int. J. Biol. Sci. 2020, 16, 408–419. [Google Scholar] [CrossRef]
  210. Golan, T.; Hammel, P.; Reni, M.; Van Cutsem, E.; Macarulla, T.; Hall, M.J.; Park, J.-O.; Hochhauser, D.; Arnold, D.; Oh, D.-Y.; et al. Maintenance Olaparib for Germline BRCA-Mutated Metastatic Pancreatic Cancer. N. Engl. J. Med. 2019, 381, 317–327. [Google Scholar] [CrossRef]
  211. Kettner, N.M.; Vijayaraghavan, S.; Durak, M.G.; Bui, T.; Kohansal, M.; Ha, M.J.; Liu, B.; Rao, X.; Wang, J.; Yi, M.; et al. Combined Inhibition of STAT3 and DNA Repair in Palbociclib-Resistant ER-Positive Breast Cancer. Clin. Cancer Res. 2019, 25, 3996–4013. [Google Scholar] [CrossRef] [Green Version]
  212. Pulliam, N.; Fang, F.; Ozes, A.R.; Tang, J.; Adewuyi, A.; Keer, H.; Lyons, J.; Baylin, S.B.; Matei, D.; Nakshatri, H.; et al. An Effective Epigenetic-PARP Inhibitor Combination Therapy for Breast and Ovarian Cancers Independent of BRCA Mutations. Clin. Cancer Res. 2018, 24, 3163–3175. [Google Scholar] [CrossRef] [Green Version]
  213. Lheureux, S.; Oaknin, A.; Garg, S.; Bruce, J.P.; Madariaga, A.; Dhani, N.C.; Bowering, V.; White, J.; Accardi, S.; Tan, Q.; et al. EVOLVE: A Multicenter Open-Label Single-Arm Clinical and Translational Phase II Trial of Cediranib Plus Olaparib for Ovarian Cancer after PARP Inhibition Progression. Clin. Cancer Res. 2020, 26, 4206–4215. [Google Scholar] [CrossRef]
  214. Parmar, K.; Kochupurakkal, B.S.; Lazaro, J.-B.; Wang, Z.C.; Palakurthi, S.; Kirschmeier, P.T.; Yang, C.; Sambel, L.A.; Färkkilä, A.; Reznichenko, E.; et al. The CHK1 Inhibitor Prexasertib Exhibits Monotherapy Activity in High-Grade Serous Ovarian Cancer Models and Sensitizes to PARP Inhibition. Clin. Cancer Res. 2019, 25, 6127–6140. [Google Scholar] [CrossRef]
  215. Yang, X.-D.; Kong, F.-E.; Qi, L.; Lin, J.-X.; Yan, Q.; Loong, J.H.C.; Xi, S.-Y.; Zhao, Y.; Zhang, Y.; Yuan, Y.-F.; et al. PARP inhibitor Olaparib overcomes Sorafenib resistance through reshaping the pluripotent transcriptome in hepatocellular carcinoma. Mol. Cancer 2021, 20, 20. [Google Scholar] [CrossRef]
  216. Patel, P.R.; Senyuk, V.; Rodriguez, N.S.; Oh, A.L.; Bonetti, E.; Mahmud, D.; Barosi, G.; Mahmud, N.; Rondelli, D. Synergistic Cytotoxic Effect of Busulfan and the PARP Inhibitor Veliparib in Myeloproliferative Neoplasms. Biol. Blood Marrow Transplant. 2019, 25, 855–860. [Google Scholar] [CrossRef]
  217. Sun, C.; Fang, Y.; Labrie, M.; Li, X.; Mills, G.B. Systems approach to rational combination therapy: PARP inhibitors. Biochem. Soc. Trans. 2020, 48, 1101–1108. [Google Scholar] [CrossRef]
  218. Higuchi, F.; Nagashima, H.; Ning, J.; Koerner, M.V.A.; Wakimoto, H.; Cahill, D.P. Restoration of Temozolomide Sensitivity by PARP Inhibitors in Mismatch Repair Deficient Glioblastoma is Independent of Base Excision Repair. Clin. Cancer Res. 2020, 26, 1690–1699. [Google Scholar] [CrossRef] [PubMed]
  219. Gupta, S.K.; Smith, E.J.; Mladek, A.C.; Tian, S.; Decker, P.A.; Kizilbash, S.H.; Kitange, G.J.; Sarkaria, J.N. PARP Inhibitors for Sensitization of Alkylation Chemotherapy in Glioblastoma: Impact of Blood-Brain Barrier and Molecular Heterogeneity. Front. Oncol. 2019, 8, 670. [Google Scholar] [CrossRef] [Green Version]
  220. Matthews, H.K.; Bertoli, C.; de Bruin, R.A.M. Cell cycle control in cancer. Nat. Rev. Mol. Cell Biol. 2021, 23, 74–88. [Google Scholar] [CrossRef]
  221. Zhang, T.; Shen, Y.; Chen, Y.; Hsieh, J.-T.; Kong, Z. The ATM inhibitor KU55933 sensitizes radioresistant bladder cancer cells with DAB2IP gene defect. Int. J. Radiat. Biol. 2015, 91, 368–378. [Google Scholar] [CrossRef]
  222. Golding, S.E.; Rosenberg, E.; Valerie, N.; Hussaini, I.; Frigerio, M.; Cockcroft, X.F.; Wei, Y.C.; Hummersone, M.; Rigoreau, L.; Menear, K.A.; et al. Improved ATM kinase inhibitor KU-60019 radiosensitizes glioma cells, compromises insulin, AKT and ERK prosurvival signaling, and inhibits migration and invasion. Mol. Cancer Ther. 2009, 8, 2894–2902. [Google Scholar] [CrossRef] [Green Version]
  223. Karlin, J.; Allen, J.; Ahmad, S.F.; Hughes, G.; Sheridan, V.; Odedra, R.; Farrington, P.; Cadogan, E.B.; Riches, L.C.; Garcia-Trinidad, A.; et al. Orally Bioavailable and Blood–Brain Barrier-Penetrating ATM Inhibitor (AZ32) Radiosensitizes Intracranial Gliomas in Mice. Mol. Cancer Ther. 2018, 17, 1637–1647. [Google Scholar] [CrossRef]
  224. Fujisawa, H.; Nakajima, N.I.; Sunada, S.; Lee, Y.; Hirakawa, H.; Yajima, H.; Fujimori, A.; Uesaka, M.; Okayasu, R. VE-821, an ATR inhibitor, causes radiosensitization in human tumor cells irradiated with high LET radiation. Radiat. Oncol. 2015, 10, 175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Huang, R.-X.; Zhou, P.-K. DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal Transduct. Target. Ther. 2020, 5, 60. [Google Scholar] [CrossRef] [PubMed]
  226. Gorecki, L.; Andrs, M.; Rezacova, M.; Korabecny, J. Discovery of ATR kinase inhibitor berzosertib (VX-970, M6620): Clinical candidate for cancer therapy. Pharmacol. Ther. 2020, 210, 107518. [Google Scholar] [CrossRef] [PubMed]
  227. Tu, X.; Kahila, M.M.; Zhou, Q.; Yu, J.; Kalari, K.R.; Wang, L.; Harmsen, W.S.; Yuan, J.; Boughey, J.C.; Matthew, P.; et al. ATR Inhibition Is a Promising Radiosensitizing Strategy for Triple-Negative Breast Cancer. Mol. Cancer Ther. 2018, 17, 2462–2472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Restelli, V.; Lupi, M.; Vagni, M.; Chilà, R.; Bertoni, F.; Damia, G.; Carrassa, L. Combining Ibrutinib with Chk1 Inhibitors Synergistically Targets Mantle Cell Lymphoma Cell Lines. Target. Oncol. 2018, 13, 235–245. [Google Scholar] [CrossRef]
  229. Peasland, A.; Wang, L.-Z.; Rowling, E.; Kyle, S.; Chen, T.; Hopkins, A.; Cliby, A.W.; Sarkaria, J.; Beale, G.; Edmondson, R.J.; et al. Identification and evaluation of a potent novel ATR inhibitor, NU6027, in breast and ovarian cancer cell lines. Br. J. Cancer 2011, 105, 372–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Stefanski, C.D.; Keffler, K.; McClintock, S.; Milac, L.; Prosperi, J.R. APC loss affects DNA damage repair causing doxorubicin resistance in breast cancer cells. Neoplasia 2019, 21, 1143–1150. [Google Scholar] [CrossRef] [PubMed]
  231. Fokas, E.; Prevo, R.; Pollard, J.R.; Reaper, P.M.; Charlton, P.A.; Cornelissen, B.; Vallis, K.A.; Hammond, E.M.; Olcina, M.M.; Gillies McKenna, W.; et al. Targeting ATR in vivo using the novel inhibitor VE-822 results in selective sensitization of pancreatic tumors to radiation. Cell Death Dis. 2012, 3, e441. [Google Scholar] [CrossRef] [Green Version]
  232. Kwok, M.; Davies, N.; Agathanggelou, A.; Smith, E.; Oldreive, C.; Petermann, E.; Stewart, G.; Brown, J.; Lau, A.; Pratt, G.; et al. ATR inhibition induces synthetic lethality and overcomes chemoresistance in TP53- or ATM-defective chronic lymphocytic leukemia cells. Blood 2016, 127, 582–595. [Google Scholar] [CrossRef] [Green Version]
  233. Kantidze, O.L.; Velichko, A.K.; Luzhin, A.V.; Petrova, N.V.; Razin, S.V. Synthetically Lethal Interactions of ATM, ATR, and DNA-PKcs. Trends Cancer 2018, 4, 755–768. [Google Scholar] [CrossRef] [PubMed]
  234. Esposito, F.; Giuffrida, R.; Raciti, G.; Puglisi, C.; Forte, S. Wee1 Kinase: A Potential Target to Overcome Tumor Resistance to Therapy. Int. J. Mol. Sci. 2021, 22, 10689. [Google Scholar] [CrossRef] [PubMed]
  235. Schmidt, M.; Rohe, A.; Platzer, C.; Najjar, A.; Erdmann, F.; Sippl, W. Regulation of G2/M Transition by Inhibition of WEE1 and PKMYT1 Kinases. Molecules 2017, 22, 2045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Nakanishi, M.; Ando, H.; Watanabe, N.; Kitamura, K.; Ito, K.; Okayama, H.; Miyamoto, T.; Agui, T.; Sasaki, M. Identification and characterization of human Wee1B, a new member of the Wee1 family of Cdk-inhibitory kinases. Genes Cells 2000, 5, 839–847. [Google Scholar] [CrossRef] [PubMed]
  237. Heald, R.; McLoughlin, M.; McKeon, F. Human wee1 maintains mitotic timing by protecting the nucleus from cytoplasmically activated cdc2 kinase. Cell 1993, 74, 463–474. [Google Scholar] [CrossRef] [PubMed]
  238. Watanabe, N.; Broome, M.; Hunter, T. Regulation of the human WEE1Hu CDK tyrosine 15-kinase during the cell cycle. EMBO J. 1995, 14, 1878–1891. [Google Scholar] [CrossRef]
  239. Masaki, T.; Shiratori, Y.; Rengifo, W.; Igarashi, K.; Yamagata, M.; Kurokohchi, K.; Uchida, N.; Miyauchi, Y.; Yoshiji, H.; Watanabe, S.; et al. Cyclins and cyclin-dependent kinases: Comparative study of hepatocellular carcinoma versus cirrhosis. Hepatology 2003, 37, 534–543. [Google Scholar] [CrossRef]
  240. Harris, P.S.; Venkataraman, S.; Alimova, I.; Birks, D.K.; Balakrishnan, I.; Cristiano, B.; Donson, A.M.; Dubuc, A.M.; Taylor, M.D.; Foreman, N.K.; et al. Integrated genomic analysis identifies the mitotic checkpoint kinase WEE1 as a novel therapeutic target in medulloblastoma. Mol. Cancer 2014, 13, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Mueller, S.; Hashizume, R.; Yang, X.; Kolkowitz, I.; Olow, A.; Phillips, J.; Smirnov, I.; Tom, M.W.; Prados, M.D.; James, C.D.; et al. Targeting Wee1 for the treatment of pediatric high-grade gliomas. Neuro-Oncology 2014, 16, 352–360. [Google Scholar] [CrossRef] [PubMed]
  242. Slipicevic, A.; Holth, A.; Hellesylt, E.; Tropé, C.G.; Davidson, B.; Flørenes, V.A. Wee1 is a novel independent prognostic marker of poor survival in post-chemotherapy ovarian carcinoma effusions. Gynecol. Oncol. 2014, 135, 118–124. [Google Scholar] [CrossRef]
  243. Magnussen, G.I.; Holm, R.; Emilsen, E.; Rosnes, A.K.R.; Slipicevic, A.; Flørenes, V.A. High Expression of Wee1 Is Associated with Poor Disease-Free Survival in Malignant Melanoma: Potential for Targeted Therapy. PLoS ONE 2012, 7, e38254. [Google Scholar] [CrossRef]
  244. Ronco, C.; Martin, A.R.; Demange, L.; Benhida, R. ATM, ATR, CHK1, CHK2 and WEE1 inhibitors in cancer and cancer stem cells. MedChemComm 2016, 8, 295–319. [Google Scholar] [CrossRef]
  245. Du, X.; Li, J.; Luo, X.; Li, R.; Li, F.; Zhang, Y.; Shi, J.; He, J. Structure-activity relationships of Wee1 inhibitors: A review. Eur. J. Med. Chem. 2020, 203, 112524. [Google Scholar] [CrossRef]
  246. Caretti, V.; Hiddingh, L.; Lagerweij, T.; Schellen, P.; Koken, P.W.; Hulleman, E.; van Vuurden, D.G.; Vandertop, W.P.; Kaspers, G.J.; Noske, D.P.; et al. WEE1 Kinase Inhibition Enhances the Radiation Response of Diffuse Intrinsic Pontine Gliomas. Mol. Cancer Ther. 2013, 12, 141–150. [Google Scholar] [CrossRef] [Green Version]
  247. Sarcar, B.; Kahali, S.; Prabhu, A.H.; Shumway, S.D.; Xu, Y.; Demuth, T.; Chinnaiyan, P. Targeting Radiation-Induced G(2) Checkpoint Activation with the Wee-1 Inhibitor MK-1775 in Glioblastoma Cell Lines. Mol. Cancer Ther. 2011, 10, 2405–2414. [Google Scholar] [CrossRef] [Green Version]
  248. Karnak, D.; Engelke, C.G.; Parsels, L.A.; Kausar, T.; Wei, D.; Robertson, J.R.; Marsh, K.B.; Davis, M.A.; Zhao, L.; Maybaum, J.; et al. Combined Inhibition of Wee1 and PARP1/2 for Radiosensitization in Pancreatic Cancer. Clin. Cancer Res. 2014, 20, 5085–5096. [Google Scholar] [CrossRef] [Green Version]
  249. Hirai, H.; Iwasawa, Y.; Okada, M.; Arai, T.; Nishibata, T.; Kobayashi, M.; Kimura, T.; Kaneko, N.; Ohtani, J.; Yamanaka, K.; et al. Small-molecule inhibition of Wee1 kinase by MK-1775 selectively sensitizes p53-deficient tumor cells to DNA-damaging agents. Mol. Cancer Ther. 2009, 8, 2992–3000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Bridges, K.A.; Hirai, H.; Buser, C.A.; Brooks, C.; Liu, H.; Buchholz, T.A.; Molkentine, J.M.; Mason, K.A.; Meyn, R.E. MK-1775, a Novel Wee1 Kinase Inhibitor, Radiosensitizes p53-Defective Human Tumor Cells. Clin. Cancer Res. 2011, 17, 5638–5648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Leijen, S.; van Geel, R.M.; Sonke, G.S.; de Jong, D.; Rosenberg, E.H.; Marchetti, S.; Pluim, D.; van Werkhoven, E.; Rose, S.; Lee, M.A.; et al. Phase II Study of WEE1 Inhibitor AZD1775 Plus Carboplatin in Patients with TP53-Mutated Ovarian Cancer Refractory or Resistant to First-Line Therapy Within 3 Months. J. Clin. Oncol. 2016, 34, 4354–4361. [Google Scholar] [CrossRef] [Green Version]
  252. Osman, A.A.; Monroe, M.M.; Ortega Alves, M.V.; Patel, A.A.; Katsonis, P.; Fitzgerald, A.L.; Neskey, D.M.; Frederick, M.J.; Woo, S.H.; Caulin, C.; et al. Wee-1 Kinase Inhibition Overcomes Cisplatin Resistance Associated with High-Risk TP53 Mutations in Head and Neck Cancer through Mitotic Arrest Followed by Senescence. Mol. Cancer Ther. 2015, 14, 608–619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Van Linden, A.A.; Baturin, D.; Ford, J.B.; Fosmire, S.P.; Gardner, L.; Korch, C.; Reigan, P.; Porter, C.C. Inhibition of Wee1 Sensitizes Cancer Cells to Antimetabolite Chemotherapeutics in vitro and in vivo, Independent of p53 Functionality. Mol. Cancer Ther. 2013, 12, 2675–2684. [Google Scholar] [CrossRef]
  254. Jette, N.; Lees-Miller, S.P. The DNA-dependent protein kinase: A multifunctional protein kinase with roles in DNA double strand break repair and mitosis. Prog. Biophys. Mol. Biol. 2015, 117, 194–205. [Google Scholar] [CrossRef] [Green Version]
  255. Wang, Y.; Xu, H.; Liu, T.; Huang, M.; Butter, P.-P.; Li, C.; Zhang, L.; Kao, G.D.; Gong, Y.; Maity, A.; et al. Temporal DNA-PK activation drives genomic instability and therapy resistance in glioma stem cells. JCI Insight 2018, 3, e98096. [Google Scholar] [CrossRef] [Green Version]
  256. Alikarami, F.; Safa, M.; Faranoush, M.; Hayat, P.; Kazemi, A. Inhibition of DNA-PK enhances chemosensitivity of B-cell precursor acute lymphoblastic leukemia cells to doxorubicin. Biomed. Pharmacother. 2017, 94, 1077–1093. [Google Scholar] [CrossRef]
  257. Pospisilova, M.; Seifrtova, M.; Rezacova, M. Small molecule inhibitors of DNA-PK for tumor sensitization to anticancer therapy. J. Physiol. Pharmacol. Off. J. Pol. Physiol. Soc. 2017, 68, 337–344. [Google Scholar]
  258. Liang, X.-M.; Qin, Q.; Liu, B.-N.; Li, X.-Q.; Zeng, L.-L.; Wang, J.; Kong, L.-P.; Zhong, D.-S.; Sun, L.-L. Targeting DNA-PK overcomes acquired resistance to third-generation EGFR-TKI osimertinib in non-small-cell lung cancer. Acta Pharmacol. Sin. 2021, 42, 648–654. [Google Scholar] [CrossRef]
  259. Fang, X.; Huang, Z.; Zhai, K.; Huang, Q.; Tao, W.; Kim, L.; Wu, Q.; Almasan, A.; Yu, J.S.; Li, X.; et al. Inhibiting DNA-PK induces glioma stem cell differentiation and sensitizes glioblastoma to radiation in mice. Sci. Transl. Med. 2021, 13. [Google Scholar] [CrossRef]
  260. Prados-Carvajal, R.; Irving, E.; Lukashchuk, N.; Forment, J.V. Preventing and Overcoming Resistance to PARP Inhibitors: A Focus on the Clinical Landscape. Cancers 2021, 14, 44. [Google Scholar] [CrossRef]
  261. D’Andrea, A.D. Mechanisms of PARP inhibitor sensitivity and resistance. DNA Repair 2018, 71, 172–176. [Google Scholar] [CrossRef] [PubMed]
  262. Lee, E.K.; Matulonis, U.A. PARP Inhibitor Resistance Mechanisms and Implications for Post-Progression Combination Therapies. Cancers 2020, 12, 2054. [Google Scholar] [CrossRef]
  263. Rottenberg, S.; Jaspers, J.E.; Kersbergen, A.; Van Der Burg, E.; Nygren, A.O.H.; Zander, S.A.L.; Derksen, P.W.B.; De Bruin, M.; Zevenhoven, J.; Lau, A.; et al. High sensitivity of BRCA1-deficient mammary tumors to the PARP inhibitor AZD2281 alone and in combination with platinum drugs. Proc. Natl. Acad. Sci. USA 2008, 105, 17079–17084. [Google Scholar] [CrossRef] [Green Version]
  264. Pettitt, S.J.; Krastev, D.B.; Brandsma, I.; Dréan, A.; Song, F.; Aleksandrov, R.; Harrell, M.I.; Menon, M.; Brough, R.; Campbell, J.; et al. Genome-wide and high-density CRISPR-Cas9 screens identify point mutations in PARP1 causing PARP inhibitor resistance. Nat. Commun. 2018, 9, 1849. [Google Scholar] [CrossRef]
  265. Gogola, E.; Duarte, A.A.; de Ruiter, J.R.; Wiegant, W.W.; Schmid, J.; de Bruijn, R.; James, D.I.; Llobet, S.G.; Vis, D.J.; Annunziato, S.; et al. Selective Loss of PARG Restores PARylation and Counteracts PARP Inhibitor-Mediated Synthetic Lethality. Cancer Cell 2018, 33, 1078–1093.e12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Konstantinopoulos, P.A.; Ceccaldi, R.; Shapiro, G.I.; D’Andrea, A.D. Homologous Recombination Deficiency: Exploiting the Fundamental Vulnerability of Ovarian Cancer. Cancer Discov. 2015, 5, 1137–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Chaudhuri, A.R.; Callen, E.; Ding, X.; Gogola, E.; Duarte, A.A.; Lee, J.-E.; Wong, N.; Lafarga, V.; Calvo, J.A.; Panzarino, N.J.; et al. Replication fork stability confers chemoresistance in BRCA-deficient cells. Nature 2016, 535, 382–387. [Google Scholar] [CrossRef] [Green Version]
  268. Rondinelli, B.; Gogola, E.; Yücel, H.; Duarte, A.A.; Van De Ven, M.; Van Der Sluijs, R.; Konstantinopoulos, P.A.; Jonkers, J.; Ceccaldi, R.; Rottenberg, S.; et al. EZH2 promotes degradation of stalled replication forks by recruiting MUS81 through histone H3 trimethylation. Nat. Cell Biol. 2017, 19, 1371–1378. [Google Scholar] [CrossRef]
  269. Yazinski, S.A.; Comaills, V.; Buisson, R.; Genois, M.-M.; Nguyen, H.D.; Ho, C.K.; Kwan, T.T.; Morris, R.; Lauffer, S.; Nussenzweig, A.; et al. ATR inhibition disrupts rewired homologous recombination and fork protection pathways in PARP inhibitor-resistant BRCA-deficient cancer cells. Genes Dev. 2017, 31, 318–332. [Google Scholar] [CrossRef] [Green Version]
  270. Edwards, S.L.; Brough, R.; Lord, C.J.; Natrajan, R.; Vatcheva, R.; Levine, D.A.; Boyd, J.; Reis-Filho, J.S.; Ashworth, A. Resistance to therapy caused by intragenic deletion in BRCA2. Nature 2008, 451, 1111–1115. [Google Scholar] [CrossRef] [PubMed]
  271. Sakai, W.; Swisher, E.M.; Karlan, B.Y.; Agarwal, M.K.; Higgins, J.; Friedman, C.; Villegas, E.; Jacquemont, C.; Farrugia, D.J.; Couch, F.J.; et al. Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers. Nature 2008, 451, 1116–1120. [Google Scholar] [CrossRef] [Green Version]
  272. Drost, R.; Bouwman, P.; Rottenberg, S.; Boon, U.; Schut, E.; Klarenbeek, S.; Klijn, C.; van der Heijden, I.; van der Gulden, H.; Wientjens, E.; et al. BRCA1 RING Function Is Essential for Tumor Suppression but Dispensable for Therapy Resistance. Cancer Cell 2011, 20, 797–809. [Google Scholar] [CrossRef] [Green Version]
  273. Noordermeer, S.M.; Adam, S.; Setiaputra, D.; Barazas, M.; Pettitt, S.J.; Ling, A.K.; Olivieri, M.; Álvarez-Quilón, A.; Moatti, N.; Zimmermann, M.; et al. The shieldin complex mediates 53BP1-dependent DNA repair. Nature 2018, 560, 117–121. [Google Scholar] [CrossRef] [PubMed]
  274. Dev, H.; Chiang, T.-W.W.; Lescale, C.; De Krijger, I.; Martin, A.G.; Pilger, D.; Coates, J.; Sczaniecka-Clift, M.; Wei, W.; Ostermaier, M.; et al. Shieldin complex promotes DNA end-joining and counters homologous recombination in BRCA1-null cells. Nat. Cell Biol. 2018, 20, 954–965. [Google Scholar] [CrossRef]
  275. Clairmont, C.S.; Sarangi, P.; Ponnienselvan, K.; Galli, L.D.; Csete, I.; Moreau, L.; Adelmant, G.; Chowdhury, D.; Marto, J.A.; D’Andrea, A.D. TRIP13 regulates DNA repair pathway choice through REV7 conformational change. Nat. Cell Biol. 2020, 22, 87–96. [Google Scholar] [CrossRef] [PubMed]
  276. Lim, J.S.J.; Tan, D.S.P.; Lim, J.S.J.; Tan, D.S.P. Understanding Resistance Mechanisms and Expanding the Therapeutic Utility of PARP Inhibitors. Cancers 2017, 9, 109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Ceccaldi, R.; Liu, J.C.; Amunugama, R.; Hajdu, I.; Primack, B.; Petalcorin, M.I.R.; O’Connor, K.W.; Konstantinopoulos, P.A.; Elledge, S.J.; Boulton, S.J.; et al. Homologous-recombination-deficient tumours are dependent on Polθ-mediated repair. Nature 2015, 518, 258–262. [Google Scholar] [CrossRef] [Green Version]
  278. Henneman, L.; van Miltenburg, M.H.; Michalak, E.M.; Braumuller, T.M.; Jaspers, J.E.; Drenth, A.P.; de Korte-Grimmerink, R.; Gogola, E.; Szuhai, K.; Schlicker, A.; et al. Selective resistance to the PARP inhibitor olaparib in a mouse model for BRCA1-deficient metaplastic breast cancer. Proc. Natl. Acad. Sci. USA 2015, 112, 8409–8414. [Google Scholar] [CrossRef] [Green Version]
  279. Vaidyanathan, A.; Sawers, L.; Gannon, A.-L.; Chakravarty, P.; Scott, A.L.; Bray, E.S.; Ferguson, M.J.; Smith, G. ABCB1 (MDR1) induction defines a common resistance mechanism in paclitaxel- and olaparib-resistant ovarian cancer cells. Br. J. Cancer 2016, 115, 431–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Murai, J.; Huang, S.-Y.N.; Das, B.B.; Renaud, A.; Zhang, Y.; Doroshow, J.H.; Ji, J.; Takeda, S.; Pommier, Y. Trapping of PARP1 and PARP2 by Clinical PARP Inhibitors. Cancer Res. 2012, 72, 5588–5599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Chen, J.; Saha, P.; Kornbluth, S.; Dynlacht, B.D.; Dutta, A. Cyclin-binding motifs are essential for the function of p21CIP1. Mol. Cell. Biol. 1996, 16, 4673–4682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Lapenna, S.; Giordano, A. Cell cycle kinases as therapeutic targets for cancer. Nat. Rev. Drug Discov. 2009, 8, 547–566. [Google Scholar] [CrossRef] [PubMed]
  283. Carrassa, L.; Damia, G. Unleashing Chk1 in cancer therapy. Cell Cycle 2011, 10, 2121–2128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Restelli, V.; Chilà, R.; Lupi, M.; Rinaldi, A.; Kwee, I.; Bertoni, F.; Damia, G.; Carrassa, L. Characterization of a mantle cell lymphoma cell line resistant to the Chk1 inhibitor PF-00477736. Oncotarget 2015, 6, 37229–37240. [Google Scholar] [CrossRef] [PubMed]
  285. Chandrasekaran, A.; Woo, S.; Sarodaya, N.; Rhie, B.; Tyagi, A.; Das, S.; Suresh, B.; Ko, N.; Oh, S.; Kim, K.-S.; et al. Ubiquitin-Specific Protease 29 Regulates Cdc25A-Mediated Tumorigenesis. Int. J. Mol. Sci. 2021, 22, 5766. [Google Scholar] [CrossRef]
  286. Ruiz, S.; Mayor-Ruiz, C.; Lafarga, V.; Murga, M.; Vega-Sendino, M.; Ortega, S.; Fernandez-Capetillo, O. A Genome-wide CRISPR Screen Identifies CDC25A as a Determinant of Sensitivity to ATR Inhibitors. Mol. Cell 2016, 62, 307–313. [Google Scholar] [CrossRef] [Green Version]
  287. Neophytou, C.M.; Trougakos, I.P.; Erin, N.; Papageorgis, P. Apoptosis Deregulation and the Development of Cancer Multi-Drug Resistance. Cancers 2021, 13, 4363. [Google Scholar] [CrossRef] [PubMed]
  288. Li, Y.; Saini, P.; Sriraman, A.; Dobbelstein, M. Mdm2 inhibition confers protection of p53-proficient cells from the cytotoxic effects of Wee1 inhibitors. Oncotarget 2015, 6, 32339–32352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Kumar, R.J.; Chao, H.X.; Simpson, A.D.; Feng, W.; Cho, M.-G.; Roberts, V.R.; Sullivan, A.R.; Shah, S.J.; Wozny, A.-S.; Fagan-Solis, K.; et al. Dual inhibition of DNA-PK and DNA polymerase theta overcomes radiation resistance induced by p53 deficiency. NAR Cancer 2020, 2, zcaa038. [Google Scholar] [CrossRef] [PubMed]
  290. Schnog, J.-J.B.; Samson, M.J.; Gans, R.O.B.; Duits, A.J. An urgent call to raise the bar in oncology. Br. J. Cancer 2021, 125, 1477–1485. [Google Scholar] [CrossRef]
  291. Martorana, F.; Da Silva, L.A.; Sessa, C.; Colombo, I. Everything Comes with a Price: The Toxicity Profile of DNA-Damage Response Targeting Agents. Cancers 2022, 14, 953. [Google Scholar] [CrossRef] [PubMed]
  292. LaFargue, C.J.; Molin, G.Z.D.; Sood, A.K.; Coleman, R.L. Exploring and comparing adverse events between PARP inhibitors. Lancet Oncol. 2019, 20, e15–e28. [Google Scholar] [CrossRef] [PubMed]
  293. Madariaga, A.; Bowering, V.; Ahrari, S.; Oza, A.M.; Lheureux, S. Manage wisely: Poly (ADP-ribose) polymerase inhibitor (PARPi) treatment and adverse events. Int. J. Gynecol. Cancer 2020, 30, 903–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Valabrega, G.; Scotto, G.; Tuninetti, V.; Pani, A.; Scaglione, F. Differences in PARP Inhibitors for the Treatment of Ovarian Cancer: Mechanisms of Action, Pharmacology, Safety, and Efficacy. Int. J. Mol. Sci. 2021, 22, 4203. [Google Scholar] [CrossRef] [PubMed]
  295. Csizmar, C.M.; Saliba, A.N.; Swisher, E.M.; Kaufmann, S.H. PARP Inhibitors and Myeloid Neoplasms: A Double-Edged Sword. Cancers 2021, 13, 6385. [Google Scholar] [CrossRef] [PubMed]
  296. Kwan, T.T.; Oza, A.M.; Tinker, A.V.; Ray-Coquard, I.; Oaknin, A.; Aghajanian, C.; Lorusso, D.; Colombo, N.; Dean, A.; Weberpals, J.; et al. Preexisting TP53-Variant Clonal Hematopoiesis and Risk of Secondary Myeloid Neoplasms in Patients with High-grade Ovarian Cancer Treated with Rucaparib. JAMA Oncol. 2021, 7, 1772–1781. [Google Scholar] [CrossRef]
  297. Kusne, Y.; Xie, Z.; Patnaik, M.M. Clonal hematopoiesis: Molecular and clinical implications. Leuk. Res. 2022, 113. [Google Scholar] [CrossRef]
  298. Bolton, K.L.; Ptashkin, R.N.; Gao, T.; Braunstein, L.; Devlin, S.M.; Kelly, D.; Patel, M.; Berthon, A.; Syed, A.; Yabe, M.; et al. Cancer therapy shapes the fitness landscape of clonal hematopoiesis. Nat. Genet. 2020, 52, 1219–1226. [Google Scholar] [CrossRef]
  299. Navitski, A.; Al-Rawi, D.H.; Liu, Y.; Rubinstein, M.M.; Friedman, C.F.; Rampal, R.K.; Mandelker, D.L.; Cadoo, K.; O’Cearbhaill, R.E. Baseline risk of hematologic malignancy at initiation of frontline PARP inhibitor maintenance for BRCA1/2-associated ovarian cancer. Gynecol. Oncol. Rep. 2021, 38, 100873. [Google Scholar] [CrossRef]
  300. Oliveira, J.L.; Greipp, P.T.; Rangan, A.; Jatoi, A.; Nguyen, P.L. Myeloid malignancies in cancer patients treated with poly(ADP-ribose) polymerase (PARP) inhibitors: A case series. Blood Cancer J. 2022, 12, 11. [Google Scholar] [CrossRef]
  301. Pujade-Lauraine, E.; Ledermann, J.A.; Selle, F.; Gebski, V.; Penson, R.T.; Oza, A.M.; Korach, J.; Huzarski, T.; Poveda, A.; Pignata, S.; et al. 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. 2017, 18, 1274–1284. [Google Scholar] [CrossRef] [Green Version]
  302. Morice, P.-M.; Leary, A.; Dolladille, C.; Chrétien, B.; Poulain, L.; González-Martín, A.; Moore, K.; O’Reilly, E.M.; Ray-Coquard, I.; Alexandre, J. Myelodysplastic syndrome and acute myeloid leukaemia in patients treated with PARP inhibitors: A safety meta-analysis of randomised controlled trials and a retrospective study of the WHO pharmacovigilance database. Lancet Haematol. 2020, 8, e122–e134. [Google Scholar] [CrossRef]
  303. Nitecki, R.; Melamed, A.; Gockley, A.A.; Floyd, J.; Krause, K.J.; Coleman, R.L.; Matulonis, U.A.; Giordano, S.H.; Lu, K.H.; Rauh-Hain, J.A. Incidence of myelodysplastic syndrome and acute myeloid leukemia in patients receiving poly-ADP ribose polymerase inhibitors for the treatment of solid tumors: A meta-analysis of randomized trials. Gynecol. Oncol. 2021, 161, 653–659. [Google Scholar] [CrossRef] [PubMed]
  304. Ma, Z.; Sun, X.; Lu, W.; Zhao, Z.; Xu, Z.; Lyu, J.; Zhao, P.; Liu, L. Poly(ADP-ribose) polymerase inhibitor-associated myelodysplastic syndrome/acute myeloid leukemia: A pharmacovigilance analysis of the FAERS database. ESMO Open 2021, 6, 100033. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of resistance to DDR inhibitors. Cancer cells develop resistance to DDR through several mechanisms. The molecular mechanisms of resistance to PARPi include HR capacity restoration, decreased trapping of PARP1, stabilization of replication forks, and P-gp-mediated drug efflux. The resistance to WEE1 inhibitor is induced by AXL overexpression, mTOR signaling, CHK1 activation, and through the overexpression of MYT1 that decrease CDK1 activity. The resistance to CHK1 inhibitor is associated with increased E2F/G2M/SAC expression and reduced replication stress. The resistance to ATR inhibitor is induced by the loss of PGBD5 and CDC25A deficiency. Finally, the DNA-PK inhibitor resistance is caused by the loss of MLH1/MSH3 and the overexpression of ABCG2.
Figure 1. Mechanisms of resistance to DDR inhibitors. Cancer cells develop resistance to DDR through several mechanisms. The molecular mechanisms of resistance to PARPi include HR capacity restoration, decreased trapping of PARP1, stabilization of replication forks, and P-gp-mediated drug efflux. The resistance to WEE1 inhibitor is induced by AXL overexpression, mTOR signaling, CHK1 activation, and through the overexpression of MYT1 that decrease CDK1 activity. The resistance to CHK1 inhibitor is associated with increased E2F/G2M/SAC expression and reduced replication stress. The resistance to ATR inhibitor is induced by the loss of PGBD5 and CDC25A deficiency. Finally, the DNA-PK inhibitor resistance is caused by the loss of MLH1/MSH3 and the overexpression of ABCG2.
Ijms 23 14672 g001
Table 1. Selected DNA damage response inhibitors approved and under research.
Table 1. Selected DNA damage response inhibitors approved and under research.
InhibitorClass/Mechanisms of ActionStage and Indication
OlaparibPARP inhibitorApproved: HER2-negative BC; FC; OC; PaC; PeC; PC
Phase III: BC; CRC; EC; NSCLC; SCLC; SCC
Phase II/III: TNBC
Phase II: BlaC; CC; GC; CBM; HNC; OSC; RCC; UC; HER2-positive; Solid tumors
TalazoparibPARP inhibitorApproved: BC; HER2-negative BC
Phase III: OC; PC
Phase II: EC, FC; SCLC; SCC; TNBC; Solid tumors
Phase I/II: AML
NiraparibPARP1 inhibitor
PARP2 inhibitor
Approved: FC; OC; PeC
Phase III: HER2-negative BC; NSCLC; SCLC
Phase II: BC; CC; CCA; CNSC; EC; GC; GBM; Glioma; HNC; Mesothelioma; NT; EsC; PaC; RCC; UC; TNBC; Solid tumors; Uveal melanoma
RucaparibPARP1 inhibitor
PARP2 inhibitor
Approved: FC; OC; PeC; PC
Phase II: BC; CC; EC; GC; UC; Mesothelioma; PaC; TNBC; Solid tumors
Phase I/II: NSCLC
PamiparibPARP1 inhibitor
PARP2 inhibitor
Approved: FC; OC; PeC (China)
Phase III: Cancer
Phase II: HER2-negative BC; GC
Phase I/II: GBM; Glioma; Solid tumors
VeliparibPARP inhibitorApproved: advanced LSCC; OC; BC, LC
Phase III: BC; HER2-negative BC; TNBC; NSCLC; OC
Phase II/III: GBM
Phase II: Brain metastases; CRC; GCEN; Malignant melanoma; PaC; RC; Solid tumors
Phase I/II: Glioma; HNC; SCLC
StenoparibPARP1 inhibitor
PARP2 inhibitorTankyrase inhibitor
Phase II: BC; OC
Phase I/II: Solid tumor
MethoxyamineAPE1 inhibitorPhase II. GBM; Mesothelioma; NSCLC
Phase I/II: Solid tumors
PalbociclibCDK4 inhibitor
CDK6 inhibitor
Approved: BC
Phase II/III: NSCLC
Phase II: Bone metastases; Brain metastases; GIST; Mantle cell lymphoma; PaC; PC; SCC; UC
Phase I/II: CRC; Malignant melanoma
IniparibCell cycle inhibitor
H2AFX protein stimulantTumor protein modulator
Phase II: Glioma
CeralasertibATR protein inhibitorPhase III: NSCLC
Phase II: TNBC; CCA; GC; GyC; Malignant melanoma; OSC; OC; PaC; PC; SCLCN; Solid tumors
Phase I/II: CLL
ElimusertibATR protein inhibitorPhase I: HNC; Lymphoma; OC; Solid tumors
M 4344ATR protein inhibitorPhase I: Lymphoma; Solid tumors
BerzosertibATR protein inhibitorPhase II: FC; LS; OC; PeC; PC; SCLC; UC; Solid tumors
Phase I: HNC
AZD 1390ATM protein inhibitorPhase I: GBM; NSCLC; Soft tissue sarcoma
Preclinical: BC; Meningioma
PrexasertibCHEK1 inhibitor
CHEK2 inhibitor
Phase II: OC; Solid tumors
Phase I/II: EC; UCP
reclinical: BC
SRA 737CHEK1 inhibitorPhase II: Solid tumors
Phase I/II: OC; PC
PeposertibDNA-PK inhibitorPhase I/II: RC/SCLC
Phase I: GBM; Cancer; Solid tumors
AZD 7648DNA-PK inhibitorPhase I: Soft tissue sarcoma
AdavosertibWEE1 inhibitorPhase II: TNBC; EC; FC; NSCLC; PeC; PC; RCC; SCLC; UtC; Solid tumors
Phase I: Hematological disorders; HNC; UtC
Preclinical: DLBCL; GC
VolasertibPLK1 inhibitorPhase III: AML
Phase II: MDS; NSCLC; OC
Phase I: Rhabdomyosarcoma; Solid tumors
OnvansertibPLK1 inhibitorPhase II: AML; PaC; PC; SCLC
Phase I/II: CRC; TNBC
Preclinical: CMML; Medulloblastoma; OC
AML, acute myeloid leukemia; ATM, serine protein kinase ATM; ATR, serine/threonine protein kinase ATR; BC, breast cancer; BlaC, bladder cancer; CC, cervical cancer; CCA, cholangiocarcinoma; CDK, cyclin-dependent kinase; CHEK, checkpoint kinase; CLL, chronic lymphocytic leukemia; CMML, chronic myelomonocytic leukemia; CNSC, central nerve system cancer; CRC, colorectal cancer; DLBCL, diffuse large B-cell lymphoma; DNA-PK, DNA-activated protein kinase; EC, endometrial cancer; EsC, esophageal cancer; FC, fallopian tube cancer; GBM, glioblastoma; GC, gastric cancer; GCEN, germ cell and embryonal neoplasms; GIST, gastrointestinal stromal tumor; GyC, gynecological cancer; HNC, head and neck cancer; LC, lung cancer; LS, leiomyosarcoma; LSCC, lung squamous cell carcinoma; MDS, myelodysplastic syndromes; NSCLC, non-small cell lung cancer; NT, neuroendocrine tumors; OC, ovarian cancer; OSC, osteosarcoma; PaC, pancreatic cancer; PARP, Poly(ADP-ribose) polymerase; PC, prostate cancer; PeC, peritoneal cancer; PLK1; polo-like kinase; RC; rectal cancer; RCC, renal cell carcinoma; SCC, squamous cell cancer; SCLC, small cell lung cancer; TNBC, triple-negative breast cancer; UC, urogenital cancer; UtC, uterine cancer; WEE1, WEE1-like protein kinase. ClinicalTrials.gov and AdisInsight data accessed on 20 November 2022.
Table 2. DNA repair pathway inhibitors currently in clinical trials to sensitize cells to therapy.
Table 2. DNA repair pathway inhibitors currently in clinical trials to sensitize cells to therapy.
Targeting ProteinInhibitorClinical StatusDisease StateNCT Number
ATR
PARP
AZD6738
Olaparib
Phase 2,
Recruiting
Neoadjuvant
Chemotherapy-resistant residual triple-negative breast cancer
NCT03740893
PARP 1/2 and Tankyrase 1/22X-121Phase 2, Active not recruitingMetastatic breast cancer, PARPi-resistant cancerNCT03562832
PARPTalazoparibPhase 2,
Recruiting
Multiple cancers, patients with aberrations in DNA damage response genesNCT04550494
PD-1
PARP
Pembrolizumab
Olaparib
Phase 2,
Not recruiting
Nasopharyngeal carcinoma, resistant to platinum agentsNCT04825990
Testosterone
PARP
OlaparibPhase 2,
Active, not recruiting
Castration-resistant prostate cancerNCT03516812
ATR
PARP
BAY1895344 NiraparibPhase 1,
Recruiting
Advanced solid tumors, ovarian cancerNCT04267939
PARPTalazoparib TemozolomidePhase 1,2
Recruiting
Metastatic castration-resistant prostate cancer and no mutations in DNA damage repairNCT04019327
PARPRucaparibPhase 2
Recruiting
Prostate cancer metastatic, resistant to androgen deprivation therapy and who carry a DNA repair gene mutationNCT03413995
PARPOlaparib
Temozolomide
IMRT
Phase 1, 2
Recruiting
Unresectable high-grade gliomasNCT03212742
IMRT, intensity-modulated radiotherapy.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jurkovicova, D.; Neophytou, C.M.; Gašparović, A.Č.; Gonçalves, A.C. DNA Damage Response in Cancer Therapy and Resistance: Challenges and Opportunities. Int. J. Mol. Sci. 2022, 23, 14672. https://doi.org/10.3390/ijms232314672

AMA Style

Jurkovicova D, Neophytou CM, Gašparović AČ, Gonçalves AC. DNA Damage Response in Cancer Therapy and Resistance: Challenges and Opportunities. International Journal of Molecular Sciences. 2022; 23(23):14672. https://doi.org/10.3390/ijms232314672

Chicago/Turabian Style

Jurkovicova, Dana, Christiana M. Neophytou, Ana Čipak Gašparović, and Ana Cristina Gonçalves. 2022. "DNA Damage Response in Cancer Therapy and Resistance: Challenges and Opportunities" International Journal of Molecular Sciences 23, no. 23: 14672. https://doi.org/10.3390/ijms232314672

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

Article Metrics

Back to TopTop