Abstract
PARP inhibitors have profoundly changed treatment options for cancers with homologous recombination repair defects, especially those carrying BRCA1/2 mutations. However, the development of resistance to these inhibitors presents a significant clinical challenge as it limits long-term effectiveness. This review provides an overview of the current understanding of resistance mechanisms to PARP inhibitors and explores strategies to overcome these challenges. We discuss the basis of synthetic lethality induced by PARP inhibitors and detail diverse resistance mechanisms affecting PARP inhibitors, including homologous recombination restoration, reduced PARP trapping, enhanced drug efflux, and replication fork stabilization. The review then considers clinical approaches to combat resistance, focusing on combination therapies with immune checkpoint inhibitors, DNA damage response inhibitors, and epigenetic drugs. We also highlight ongoing clinical trials and potential biomarkers for predicting treatment response and resistance. The review concludes by outlining future research directions, emphasizing the need for longitudinal studies, advanced resistance monitoring technologies, and the development of novel combination strategies. By tackling PARP inhibitor resistance, this review seeks to aid in the development of more effective cancer therapies, with the potential to improve outcomes for patients with homologous recombination-deficient tumors.
Introduction
Poly (ADP-ribose) polymerase (PARP) inhibitors have assumed significant therapeutic importance in cancer management, especially for homologous recombination repair (HR) deficient tumors, such as BRCA1/2-mutated cancers. These targeted therapies, rooted in the principle of synthetic lethality, selectively kill cancer cells and spare normal ones, thereby offering a potentially superior therapeutic index compared to conventional chemotherapy regimens. Since the approval of the first PARP inhibitor (PARPi), Olaparib, in 2014, multiple PARP inhibitors have been introduced into clinical practice, significantly advancing treatment options for various cancers, such as ovarian, breast, pancreatic, and prostate cancers, which are highlighted here due to their prominence as primary indications for PARPi therapy, their representation of diverse molecular vulnerabilities (ranging from BRCA-dependent HRD to alternative pathways), and their extensive clinical data on resistance mechanisms. The advent of PARP inhibitors has significantly changed the treatment landscape for malignancies with homologous recombination deficiency (HRD), a molecular phenotype initially defined by BRCA1/2 mutations but now expanded to include epigenetic dysregulation and synthetic lethal interactions [1–7]. Ovarian and breast cancers were the first approved indications for PARPi therapy due to their high prevalence of BRCA mutations, with pivotal trials like SOLO-1 and OlympiAD establishing BRCA status as a primary predictive biomarker. However, 40–70% of ovarian cancer patients develop resistance through BRCA reversion mutations, while 50% of BRCA-mutant breast cancer patients progress within 12 months of therapy, highlighting the limitations of relying solely on BRCA status [8–10].
This limitation has prompted a paradigm shift in tumor type selection, emphasizing molecular vulnerabilities beyond canonical HRD. Pancreatic cancers with PALB2 loss (POLO trial) and metastatic castration-resistant prostate cancers (mCRPC) carrying ATM deletions (PROfound trial) demonstrate PARPi sensitivity mediated by alternative HRD pathways, though secondary resistance emerges in > 50% of cases via non-BRCA DNA repair adaptations [11, 12]. This resistance phenomenon has driven intense research to understand the underlying mechanisms and develop strategies to overcome or prevent it.
This review provides a comprehensive summary of the current understanding of PARP inhibitor resistance and will explore the strategies being developed to tackle the challenge. The first section of the review will discuss the molecular basis for PARP inhibitor-induced synthetic lethality, followed by a discussion focused on the mechanisms by which the cancer cells acquire resistance to PARP inhibition. The final section will cover clinical strategies and combinations that might be promising for overcoming PARP inhibitor resistance while focusing on clinical trials and further developments in this rapidly progressing field.
PARP Inhibitor-induced synthetic lethality
The DNA Damage Response (DDR) pathway is a key pathway for genomic stability and managing cellular responses to DNA lesions induced by metabolism and environmental stressors [13]. By detecting, signaling, and repairing damage, DDR mechanisms avert the accumulation of mutations that otherwise would drive tumorigenesis. The DDR pathway integrates several repair mechanisms that include Base Excision Repair (BER), Homologous Recombination (HR), Non-Homologous End Joining (NHEJ), and Mismatch Repair (MMR). BER is a very crucial pathway primarily involved in removing base lesions from oxidative or alkylating damage, and it relies on PARP1, which is a protein that detects and binds to the single-stranded breaks (SSBs) [14]. After SSB recognition, PARP1 initiates chromatin remodeling and facilitates the recruitment of DNA repair factors to restore DNA integrity [15]. HR is a high-fidelity DNA repair mechanism that corrects DNA double-strand breaks (DSBs) during the S/G2 phases of the cell cycle. This process relies on the presence of a homologous sister chromatid as a template to ensure accurate repair. HR is mediated by a coordinated complex of proteins, including BRCA1, BRCA2, RAD51, and PALB2, along with Fanconi anemia pathway proteins. These factors collectively facilitate strand invasion, DNA synthesis, and error-free restoration of the damaged DNA [16–18]. HRD arises when mutations, epigenetic silencing (e.g., BRCA1 promoter hypermethylation), or functional disruptions in these core components impair repair, forcing reliance on error-prone pathways like NHEJ. While BRCA1/2 mutations are the hallmark of HRD, defects in upstream regulators such as ATM, ATR, or 12 can also confer an HRD phenotype, termed “BRCAness”.
PARP inhibitors exploit HRD vulnerability through synthetic lethality, a mechanism where simultaneous disruption of two pathways (e.g., HR and PARP1-mediated repair) selectively kills cancer cells while sparing normal cells. Specifically, PARPi block PARP1’s enzymatic activity and trap PARP1 at DNA damage sites (e.g., unligated Okazaki fragments), converting transient SSBs into replication-associated DSBs. Beyond PARP trapping, recent studies reveal that PARP1 resolves transcription-replication conflicts (TRCs) by recruiting TIMELESS (TIM). PARPi disrupt this process, leading to unresolved TRCs that exacerbate replication stress specifically in HRD cells [19]. HR-proficient cells can repair these DSBs via homologous recombination, but HRD cells cannot, leading to genomic catastrophe [20–22]. However, the clinical definition of HRD remains contentious. FDA-approved companion diagnostics (e.g., Myriad myChoice®) use genomic “scarring” biomarkers—loss of heterozygosity (LOH), telomeric allelic imbalance (TAI), and large-scale transitions (LST)—to quantify HRD, yet these assays show variable correlation with PARPi efficacy [23]. These biomarkers reflect historical HRD events during tumor evolution but may not capture real-time HR functional status, contributing to variable correlation with PARPi efficacy. PARP inhibitors exploit HRD vulnerability through synthetic lethality, yet intrinsic resistance mechanisms significantly limit their efficacy. Notably, a subset of tumors classified as HRD-positive by genomic scarring biomarkers (e.g., LOH, TAI, LST) retain functional homologous recombination (HR) capacity through compensatory mechanisms such as BRCA-independent RAD51 loading or epigenetic plasticity (e.g., partial BRCA1 promoter demethylation), enabling escape from synthetic lethality [24]. PARP1 enzymatic activity is also critical: tumors harboring hypomorphic PARP1 variants (e.g., E988 K mutation) or baseline low PARP1 expression exhibit diminished PARP trapping and resistance, independent of HR status [25]. Furthermore, Schlafen-11 (SLFN11) deficiency is observed in 30–40% of ovarian and small cell lung cancers and confers intrinsic resistance by enabling replication fork progression under PARPi-induced stress [26, 27]. It has been evidenced by reduced progression-free survival in SLFN11-low patients across multiple clinical trials. These mechanisms collectively underscore why up to 25% of HRD-positive patients fail to respond initially, necessitating functional assays or dynamic multi-omics profiling to refine patient selection beyond static genomic scarring models [28–30].
Consequently, when SSBs remain unrepaired, they transform into DSBs during DNA replication. These trapped PARP complexes worsen the situation by blocking the replication process [31]. Cells that are efficient in homologous recombination can easily repair these DSBs. However, in the absence of functional HR, such as in cells with BRCA mutations, these DSBs accumulate, leading to genomic instability and cell death (Fig. 1).
Fig. 1.
Mechanism of synthetic lethality mediated by PARP inhibitors in HR-deficient cells. PARP inhibitors induce synthetic lethality by disrupting DNA damage repair mechanisms and transcription-replication conflicts (TRCs). PARPs normally bind to single-strand break sites, initiating repair processes. They also interact with TIMELESS and TIPIN to resolve TRCs, which are major sources of genome instability and double-strand breaks (DSBs). PARP inhibitors competitively bind to PARPs, preventing their normal function and trapping them on damaged DNA. This leads to an accumulation of unrepaired single-strand breaks and unresolved TRCs, which evolve into DSBs during DNA replication. In HR-deficient tumor cells, these DSBs cannot be effectively repaired, forcing cells to rely on error-prone NHEJ repair. The resulting genomic instability ultimately triggers cell death, demonstrating the selective toxicity of PARP inhibitors in HR-deficient cancers
This targeted approach effectively destroys cancer cells with homologous recombination deficiencies while mostly preserving healthy cells with functioning HR systems. The clinical efficacy of PARP inhibitors has been particularly notable in cancers with BRCA1/2 mutations. A recent phase III trial (OlympiA) showed that adjuvant olaparib significantly enhanced invasive disease-free survival in patients with HER2-negative early breast cancer and BRCA1/2 mutations [32]. Similarly, the SOLO1 trial showed that maintenance therapy with olaparib provided a substantial progression-free survival benefit in newly diagnosed advanced ovarian cancer with BRCA1/2 mutations [33].
In solid tumors, PARPi has shown promising efficacy across various cancer types. The QUADRA study demonstrated that niraparib had antitumor activity in patients with heavily pretreated ovarian cancer, including those without BRCA mutations or HRD [34]. In metastatic castration-resistant prostate cancer (mCRPC), the PROFOUND trial suggested that olaparib improved radiographic progression-free survival in men harboring mutations in homologous recombination repair genes, including but not limited to BRCA1/2 [35]. This evidence supported the FDA approval of rucaparib and olaparib based on HRD deficiencies, which can occur in BRCA-wildtype tumors. In non-small cell lung cancer (NSCLC), a traditionally BRCA-independent cancer, the phase II ORION study explored olaparib monotherapy in patients with homologous recombination repair gene mutations beyond BRCA1/2, showing promising clinical activity [36].
In hematological malignancies, the concept of HRD as a predictive biomarker for PARPi efficacy is also gaining traction. While BRCA1/2 mutations are rare in blood cancers, HRD is often associated with other genetic alterations, such as TET2, DNMT3 A, IDH1/2, or STAG2 mutations, which impair homologous recombination repair and sensitize these malignancies to PARP inhibition [37]. Recent trials, such as the use of veliparib in combination with temozolomide or carboplatin, have demonstrated meaningful clinical responses in subsets of patients with acute myeloid leukemia (AML) and myelodysplastic syndromes (MDS), particularly those with HRD-driven disease [38, 39].
Notably, in adult T-cell leukemia/lymphoma (ATLL), a highly aggressive form of leukemia associated with HTLV-I infection, genetic heterogeneity and chemotherapy resistance present significant treatment challenges. A study investigating the PARP inhibitor PJ-34 demonstrated its ability to induce G2/M cell cycle arrest, reactivate p53 transcriptional activity, and trigger caspase-3-dependent apoptosis in HTLV-I-transformed and patient-derived ATLL cells [40]. However, resistance to PJ-34 therapy was observed in certain HTLV-I-transformed cell lines, such as MT-2 cells, which showed reduced caspase-3 cleavage and upregulation of RelA/p65. These findings indicate that HRD testing, similar to its application in solid tumors, could be crucial in pinpointing blood cancer patients most likely to respond well to PARP inhibitors, thereby enabling more tailored treatment approaches.
The capacity of PARP inhibitors to specifically target cancer cells with HRDs is a key factor in their success in treating BRCA1/2 mutated cancers. While this has spurred exploration of PARPi in other HRD-positive malignancies, including prostate and pancreatic cancers, emerging data reveal critical exceptions within molecularly defined subgroups. A notable example is microsatellite instability-high (MSI-H) pancreatic cancer, which constitutes 1–2% of cases. Despite sporadic BRCA1/2 monoallelic mutations in this subgroup, PARPi resistance arises due to the mutual exclusivity between MSI-H (driven by mismatch repair deficiency) and functional HRD [41]. This biological dichotomy is further evidenced by the elevated tumor mutational burden (TMB) in MSI-H tumors, which renders them highly responsive to immune checkpoint inhibitors (ICIs). The KEYNOTE-158 trial demonstrated a 34% response rate to pembrolizumab in this population, establishing ICI as the preferred first-line therapy [42]. Ongoing studies aim to decipher resistance mechanisms, such as HR restoration and dynamic biomarker evolution, to refine PARPi utility across heterogeneous tumor contexts.
Mechanisms of resistance to PARP inhibition
PARP inhibitors have improved outcomes in HRD-associated cancers, but intrinsic and acquired resistance remain major challenges. Several acquired resistance mechanisms have been identified, including HR restoration, reduced PARP trapping, increased drug efflux, and enhanced replication fork protection, revealing potential therapeutic targets for overcoming PARPi resistance (Fig. 2).
Fig. 2.
Mechanisms of Resistance to PARP inhibitors. (1) HR restoration can occur through BRCA1/2 reactivation(a) and restored DNA end resection(b). (2) Both downregulation or mutations in PARP1(c) and loss of poly (ADP-ribose) glycohydrolase (PARG) (d) can reduce PARP inhibitor trapping. (3) Overexpression of ABC transporters, such as ABCB1, ABCC1, and ABCG2, leads to increased drug efflux, resulting in reduced intracellular concentrations of PARP inhibitors and contributing to treatment resistance. (4) PARPi resistance can emerge through mechanisms that enhance fork stability, such as loss of fork reversal factors (SMARCAL1, ZRANB3, HTLF and CHD4), reduced EZH2-mediated MUS81 recruitment, loss of PTIP/MLL3/MLL4-mediated MRE11 recruitment, decreased PTIP expression, or loss of SLFN11-dependent and E2 F7-dependent cell cycle checkpoint activation
HR Restoration
PARP inhibitors exert their therapeutic effects by disrupting the HR pathway, thereby impairing DNA damage repair and selectively targeting HR-deficient tumor cells. However, the restoration of HR activity is a prominent mechanism of acquired resistance, particularly in initially responsive HR-deficient tumors. A key pathway through which resistance arises involves secondary mutations that partially or fully restore the function of critical HR genes, such as BRCA1 or BRCA2, thereby reinstating HR repair capacity [43]. Using liquid biopsy or circulating cell-free DNA (cfDNA), these secondary mutations, often involving insertions or deletions that correct frameshift mutations, have been identified to restore the open reading frame of BRCA1/2, enabling full transcription and activity of the repair proteins [44–46] (Fig. 2). In the presence of functional HR, the synthetic lethality mechanism between BRCA1/2 mutations and PARP inhibition, which is essential for inducing cancer cell death, is disrupted, resulting in the loss of clinical efficacy for PARP inhibitors [47]. In metastatic castration-resistant prostate cancer (mCRPC) patients with BRCA mutations treated with Rucaparib, BRCA reversion mutations were detected in 39% after progression, with higher rates associated with prolonged therapy, subclonal evolution, and clinical responses such as objective or prostate-specific antigen-based improvements [48, 49]. A longitudinal ctDNA analysis by E. Harvey-Jones et al. demonstrated that 60% of PARPi-resistant breast cancer patients exhibited BRCA1/2 reversion mutations. Furthermore, resistance was not confined to reversion mutations; loss-of-function mutations in key HR pathway regulators such as TP53BP1, RIF1, and PAXIP1 were also observed, with some cases showing co-occurrence of these alterations alongside BRCA1/2 reversions [50].
Resistance to PARP inhibitors can also arise through BRCA-independent mechanisms, including additional mutations in key HR-related genes such as RAD51 C, RAD51D, and PALB2, epigenetic modifications affecting repair pathways, or the functional loss of critical HR suppressors, such as p53-binding protein 1 (53BP1) and MAD2L2 (also known as REV7). These alternative routes bypass BRCA dependency and contribute to the reactivation of homologous recombination, undermining the therapeutic efficacy of PARP inhibitors [48, 51–54].
Beyond these pathways, epigenetic regulation and RAD51-related mechanisms synergistically modulate HR restoration. For instance, complete methylation of the BRCA1 promoter silences its expression, enhancing PARPi sensitivity, whereas loss of methylation restores BRCA1 transcription and HR repair, leading to resistance. This mechanism has been observed in patient-derived xenograft (PDX) models of triple-negative breast cancer (TNBC) and ovarian cancer [55, 56]. Similarly, homozygous methylation of the RAD51 C promoter (meRAD51 C) drives HR deficiency and predicts PARPi sensitivity [57]. The role of methylation in HR restoration is further supported by a phase I clinical trial in acute myeloid leukemia (AML) combining a DNMT inhibitor with a PARP inhibitor, which demonstrated altered HR repair dynamics [58].
Importantly, RAD51-mediated HR reactivation is tightly regulated by key suppressors such as 53BP1 and REV7. The 53BP1-RIF1-REV7-shieldin pathway acts as a barrier to HR in BRCA1-deficient cells by protecting DNA ends from resection, thereby favoring NHEJ over HR [59–61]. 53BP1 is recruited to DSBs and serves as a scaffold for downstream factors. RIF1 is then recruited by 53BP1, followed by REV7, which recruits the shieldin complex (SHLD1, SHLD2, and SHLD3). This assembled complex prevents excessive DNA end resection, suppressing HR. Loss of 53BP1, REV7, or shieldin components disrupts this protection, enabling resection and partial HR restoration even without functional BRCA1 [62] (Fig. 2).
Other regulators of DNA end resection further fine-tune this balance. The CTC1-STN1-TEN1 (CST) complex, operating downstream of the 53BP1-RIF1-REV7-shieldin pathway, prevents end resection at DSBs [63]. Loss of CST complex components restores end resection in BRCA1-deficient cells, leading to PARP inhibitor resistance [63]. However, this effect is less pronounced compared to the disruption of 53BP1-RIF1-REV7-shieldin signaling, indicating the involvement of additional protective mechanisms.
HELB and DYNLL1 act downstream of 53BP1 to antagonize various components of the DNA end resection machinery [64, 65]. Their loss results in hyper-resected DNA ends and confers PARP inhibitor resistance in BRCA1-deficient tumor cells. DYNLL1 also promotes NHEJ by stimulating 53BP1 oligomerization, enhancing its recruitment to DSBs [65]. Upstream factors like ERCC6L2, an accessory NHEJ factor, also play a role. Loss of ERCC6L2 restores DNA end resection, partially rescuing HR and inducing PARP inhibitor resistance in BRCA1-deficient cells [66]. Overexpression of factors promoting HR and suppressing NHEJ, such as TIRR, TRIP13, and miRNA-622, rescues HR and reduces PARP inhibitor sensitivity in BRCA1-deficient cells [67–69]. These findings highlight the complex regulation of DNA repair pathway choice and PARP inhibitor sensitivity.
PARP inhibitor resistance often arises through the loss of DNA end protection in cells lacking functional BRCA1. This allows for the restoration of end resection and partial reactivation of HR, even without BRCA1. However, while BRCA1 may be partially dispensable for the distal steps of RAD51-mediated HR, BRCA2 remains crucial for this pathway.
Decreased PARP trapping
Decreased PARP trapping, a crucial mechanism underlying PARP inhibitor resistance, stems from a reduced ability of PARP enzymes, primarily PARP1, to become stably bound to DNA lesions [70, 71]. This diminished trapping can arise from several factors (Fig. 2). Downregulation of PARP1 expression, for instance, limits the overall amount of enzyme available to bind DNA damage [72]. Furthermore, mutations within PARP1, especially those impacting its DNA-binding domain, can impair its ability to recognize and bind damaged DNA. Counterintuitively, loss of PARG activity, the enzyme responsible for reversing PARylation, also contributes to reduced trapping [73, 74]. While PARG activity typically facilitates PARP release from DNA, its absence leads to hyper-PARylation [75]. This excessive modification can trigger accelerated PARP1 degradation, effectively depleting the pool of PARP1 available for trapping at DNA damage sites and ultimately promoting resistance, similar to observations with PARG inhibitors enhancing resistance.
Increased drug efflux
Increased drug efflux contributes to PARP inhibitor resistance, primarily through the overexpression of ATP-binding cassette (ABC) transporters like ABCB1 (P-gp), ABCC1 (MRP1), and ABCG2 (BCRP), which actively pump PARP inhibitors out of cells, reducing intracellular drug concentrations (Fig. 2). This overexpression can be driven by the activation of transcription factors such as Nrf2, which regulates the expression of these transporter genes, ultimately leading to decreased drug efficacy and treatment failure. Preclinical studies in BRCA-deficient mouse models demonstrated that Olaparib resistance can arise through ABCB1 overexpression, and that this resistance is reversible with the ABCB1 inhibitor tariquidar, confirming the role of drug efflux [76–78]. Elizabeth L. et al. identified recurrent SLC25 A40–ABCB1 genomic fusions in 15.7% of patients with high-grade serous ovarian carcinoma, with similar fusion events observed in 30% (9/30) of metastatic or end-stage breast cancer cases [76]. In BRCA1/2-deficient murine mammary tumor models, ABCB1 mRNA expression was 20-fold higher in PARP inhibitor-resistant tumors compared to treatment-sensitive counterparts [77]. Liu, et al. demonstrated that the PARP1-DOT1L-PLCG2/ABCB1 regulatory axis significantly modulates PARPi responses in BRCA wild-type ovarian cancer cells [79]. Mechanistically, DOT1L-mediated H3 K79 methylation facilitates recruitment of methylated H3 K79 to the promoter regions of PLCG2 and ABCB1, thereby enhancing their transcriptional activation [79]. Comparative analysis revealed elevated ABCB1 mRNA expression in control SKOV-3 cells relative to DOT1L-knockout counterparts, underscoring the functional relevance of this epigenetic modification in chemoresistance regulation [79]. The overexpression of ABCC1 and ABCG2 transporters on cancer cell membranes reduces the therapeutic efficacy of Talazoparib in ovarian cancer treatment by enhancing drug efflux and decreasing its cytotoxicity [80, 81].
Stabilization of replication forks
Stabilization of replication forks is a critical mechanism of PARP inhibitor resistance, particularly in cells deficient in BRCA1 or BRCA2, distinct from restoring HR. Enhanced fork stability contributes to PARP inhibitor resistance in cells with loss of BRCA1/2, underscoring their essential roles in maintaining fork integrity under replicative stress (Fig. 2). This resistance arises from a complex interplay of factors that protect stalled forks from detrimental processing. For example, MRE11, crucial for processing stalled forks, can become overactive in BRCA1/2 deficient cells, leading to excessive resection and fork collapse [82–85]. Fork remodeling by chromatin remodelers like SMARCAL1, ZRANB3, and HLTF is required for this MRE11-dependent nascent DNA degradation [86, 87]. Consequently, depleting these remodelers and factors like PTIP and CHD4 involved in MRE11 recruitment/activation leads to fork protection and PARP inhibitor resistance [88]. Similarly, limiting the activity of other nucleases like MUS81, potentially through EZH2 inhibition, can also promote resistance, although MUS81's precise role remains context-dependent [89]. Another key player, RADX, is implicated in fork protection in BRCA2-deficient cells; its depletion confers resistance by stabilizing forks [90]. Recent research has identified E2 F7, a transcriptional repressor, as a modulator of therapeutic resistance and a potential biomarker for early diagnosis, survival prediction, and a marker to monitor disease relapse [91]. In BRCA2-deficient cells, E2 F7 regulates RAD51 expression, thereby influencing HR repair and replication fork stability, contributing to PARP inhibitor and cisplatin resistance [92]. Critically, these resistance mechanisms operate independently of HR restoration, as depletion of PTIP, EZH2, or RADX does not rescue HR deficiency. This highlights fork stabilization as a potent, distinct driver of PARP inhibitor resistance. Beyond these factors, PARP1 itself plays a complex role at stalled forks, recruiting MRE11 and potentially influencing the outcome of combination therapies. SLFN11 also critically modulates replication stress responses independent of HR. Its presence enforces cell cycle arrest upon replication stress, increasing sensitivity to PARP inhibitors [93, 94]. Loss of SLFN11, therefore, allows cell cycle progression despite DNA damage, diminishing PARP inhibitor efficacy [95].
Metabolic pathways involved in PARPi resistance
In many cancer cells, glycolysis is upregulated, known as the Warburg effect [96]. Increased lactate production from glycolysis can lead to intracellular acidosis, promoting DNA damage and increasing replication stress [97]. Additionally, increased glycolytic flux may help cancer cells bypass the need for oxidative phosphorylation and other DNA repair mechanisms (e.g., NAD + regeneration, which is critical for PARP activity) [98–100]. These adaptive changes create vulnerabilities that can be therapeutically exploited. For instance, glycolysis inhibition using agents like 2-deoxyglucose (2-DG) or targeting key glycolytic enzymes (e.g., Hexokinase 2) could impair NAD + regeneration, directly interfering with PARP activity and enhancing PARPi sensitivity [101]. Other glycolytic inhibitors, such as 3-bromopyruvate (3-BrPA), have shown significant efficacy, particularly in hepatocellular carcinoma and pancreatic cancer [102, 103]. Similarly, targeting the NAD + salvage pathway by inhibiting nicotinamide phosphoribosyltransferase (NAMPT), a rate-limiting enzyme in this pathway, has been shown to enhance the cytotoxicity of PARP inhibitors, particularly in triple-negative breast cancer (TNBC) cells. NAMPT inhibitors like FK866, OT-82, and GMX1777(8) combined with PARPi show improved efficacy in various cancer models [104, 105]. OT-82 exhibits potent anti-tumor activity in hematological malignancies, while GMX1777(8) demonstrates synergistic effects with PARPi in multiple solid tumor types [105, 106].
Beyond glycolysis, mitochondrial metabolism plays a dual role in resistance. While OXPHOS inhibition through metformin or IACS-10759 (a mitochondrial complex I inhibitor) depletes NAD + and suppresses DNA repair, making cells more dependent on PARP-mediated repair and thus more sensitive to PARPi [107–109]. OXPHOS inhibition can also exacerbate oxidative stress, further diminishing cancer cell viability [110]. Additionally, mitochondrial-targeted therapies may synergize with PARPi in tumors that are resistant due to high oxidative capacity [108].
Lipid metabolic adaptations further complicate resistance dynamics. Fatty acid oxidation (FAO) and lipogenesis are critical for maintaining membrane integrity and mitochondrial function [111]. Increased intracellular FAO may provide tumor cells with additional sources of acetyl-CoA and NADPH, which can enhance DNA repair and counteract PARPi toxicity [112]. Inhibiting lipid metabolism using drugs like C75 (which inhibits fatty acid synthase) or etomoxir (which blocks carnitine palmitoyltransferase I) could impair lipid biosynthesis and mitochondrial function, leading to increased DNA damage and PARPi sensitization [113].
Emerging insights highlight the therapeutic potential of metabolic contexture. Neomorphic isocitrate dehydrogenase (IDH) mutations in gliomas exemplify how oncometabolites modulate PARPi response [114]. These mutations decrease NAD + levels by downregulating nicotinate phosphoribosyltransferase (NAPRT1), making tumors hypersensitive to NAD + depletion, while producing D-2-hydroxyglutarate (D-2HG) that inhibits homologous recombination, creating dual metabolic sensitivities [115, 116]. Current preclinical exploration focuses on optimizing combination strategies. The simultaneous use of PARPi (e.g., Olaparib, Niraparib) with metabolic inhibitors like metformin (targeting OXPHOS) and 2-DG (targeting glycolysis) is being investigated for synthetic lethality in HR-deficient tumors [117, 118]. Collectively, glycolytic reprogramming, OXPHOS dependency, lipid metabolism adaptations, and oncometabolite-driven sensitivities are synthetically targeted in combinatorial strategies (Fig. 3).
Fig. 3.
Metabolic pathways driving PARP inhibitor resistance and its potential therapeutic targets. Metabolic reprogramming, including upregulated glycolysis, enhanced oxidative phosphorylation (OXPHOS), and altered lipid metabolism, contributes to resistance against PARP inhibitors (PARPi) in cancer. Potential therapeutic strategies targeting these pathways, such as glycolytic inhibitors (e.g., 2-DG), OXPHOS inhibitors (e.g., metformin, IACS-10759), lipid metabolism inhibitors (e.g., C75, etomoxir), or NAD + salvage pathway inhibitors (e.g., FK866, OT-82, GMX1777(8)), are depicted as combinatorial approaches to enhance PARPi sensitivity and overcome resistance mechanisms in tumors. Notably, tumors with neomorphic isocitrate dehydrogenase (IDH) mutations exhibit dual metabolic vulnerabilities: NAD + depletion via downregulation of NAPRT1 and accumulation of the oncometabolite D-2-hydroxyglutarate (D-2HG), which impairs homologous recombination repair. These mechanisms highlight how metabolic and genetic alterations synergistically sensitize tumors to PARPi combination therapies
Future directions emphasize precision targeting through metabolic profiling, where tumor-specific signatures of glycolytic flux, OXPHOS activity, or lipid metabolism could guide the selection of PARPi combination regimens. These metabolic signatures help design rational drug combinations that address resistance mechanisms arising from metabolic adaptations. The tumor microenvironment significantly affects treatment outcomes. Metabolites released by stromal cells can create protective microenvironments that promote resistance. Targeting specific metabolic processes, such as lactate transport or pH regulation mechanisms, may disrupt these protective environments and restore drug sensitivity. Advances in non-invasive metabolic biomarkers will enable clinicians to monitor treatment responses in real time. This monitoring capability allows for timely adjustments to therapy as tumors adapt. Additionally, emerging immunometabolic approaches target both resistance mechanisms and immune suppression simultaneously. Metabolic reprogramming strategies show particular promise by restoring PARPi sensitivity while activating anti-tumor immune responses. By integrating precise metabolic analysis with microenvironment modulation and immune system engagement, researchers can develop more effective strategies to overcome PARPi resistance across various cancer types.
Role of ncRNAs in PARP inhibitors resistance
miRNAs can modulate key genes involved in homologous recombination repair (HRR) and base excision repair (BER), pathways that are directly affected by PARP inhibition [119]. For example, miR-182, miR-155, and miR-30a have been shown to regulate the expression of BRCA1/2, RAD51, and other DNA repair factors. Upregulation or downregulation of these miRNAs could impact the efficiency of DNA repair and influence PARPi sensitivity. In addition, some miRNAs can directly target PARP1 or PARP2. For example, miR-34a has been shown to suppress PARP1 expression, potentially influencing the response to PARPi [120]. Besides directly modulating genes involved in HRR or BER, miRNA can impact PARPi sensitivity through other mechanisms. miR-181a has been determined to downregulate the STING pathway, thereby driving PARPi resistance in TNBC and ovarian cancer [121]. miR-622 can induce PARPi resistance by interacting with Ku70 and Ku80, and impacting NHEJ-mediated repair of DSBs [69].
Certain lncRNAs, such as HOTAIR and MALAT1, have been implicated in the regulation of DNA repair genes [122]. For example, MALAT1 has been shown to modulate the expression of HR repair genes and could impact PARPi sensitivity by influencing BRCA1/2 expression or other repair pathway genes [123]. PANDAR, a lncRNA, has been associated with DNA damage repair and may modulate PARP inhibitor response by influencing the repair machinery [124]. Several lncRNAs are involved in regulating the DDR, an essential process in determining cell fate after DNA damage [125]. XIST, for instance, can influence DNA repair pathways, and its expression may be altered in PARPi-resistant cells, contributing to resistance [126]. Interestingly, a recent study uncovered that inhibition of lncRNA PART1 conferred resistance to PARPi through promoting the downregulation of PHB2, an inner mitochondrial membrane mitophagy receptor, which may play a role in cellular senescence [127]. CircRNAs can serve as potential vaccine platforms, drug targets, and act as miRNA sponges in PARPi resistance [128–130]. Circ-ARID1 A has been shown to sponge miR-370-3p, leading to the upregulation of its target gene ARID1 A, which is involved in the regulation of DNA repair [118, 131]. By modulating miRNA activity, circRNAs can influence the expression of genes involved in PARPi sensitivity, contributing to resistance [132, 133].
Tumor microenvironment (TME)-mediated PARPi resistance
TME can promote proliferation signaling, generating blood vessels, inhibiting cell apoptosis and evading immune surveillance. Due to its complex function, TME plays a crucial role in cancer initiation, promotion, and drug resistance [134]. As vital components of TME, M2-like tumor-associated macrophages are associated with suppression of CD8 + T cell function and treatment tolerance [135]. A study found that PARPi-induced DNA damage can upregulate STAT3 signaling pathway in BRCA-deficient ovarian cancer cells. JAK2-STAT3 pathway activation upregulates several STAT3-regulated chemokines like CCL2, CX3 CL1, CCL20, and CXCL1, thereby promoting protumor M2-like macrophage polarization, which can lead to PARPi resistance [136]. Similarly, another study demonstrated that PARP inhibition modulates the TME and results in phenotypic changes of immunosuppressive macrophages. These changes subsequently turn on an immune-suppressive signaling pathway, manifested by increased PD-L1 and CSF-1R expression, limiting the efficacy and contributing to PARPi resistance [137]. Collectively, TME-mediated PARPi resistance involves multidimensional immune remodeling, and future studies should focus on developing TME-targeted combination strategies.
Clinical strategies to overcome PARP inhibitor resistance
Researchers are actively exploring various combination strategies in clinical trials to address the challenge of PARPi resistance and improve patient outcomes. As summarized in Table 1, current clinical investigations are not limited to monotherapy regimens. Emerging combination strategies with chemotherapy, targeted agents, or immunotherapies are being actively explored to enhance therapeutic efficacy. Notably, current research prioritizes three principal combination categories designed to circumvent resistance mechanisms and prolong treatment effectiveness (Table 2). These combinations include ICIs to enhance anti-tumor immune responses, DNA damage response inhibitors to further exploit cancer cell vulnerabilities, and epigenetic drugs to modulate gene expression and potentially resensitize resistant cells. These diverse approaches are being evaluated across various cancer types and treatment settings to develop more effective and durable treatment options for patients resistant to PARPi monotherapy.
Table 1.
Ongoing clinical trials of PARP inhibitors in cancer treatment
Type of cancer | PARPi agent | Identifiers | Phase | Population | Intervention/Treatment | Primary endpoints |
Estimated completion date |
---|---|---|---|---|---|---|---|
Ovarian cancer | Rucaparib | NCT03522246 | 3 | Newly diagnosed ovarian cancer patients | Rucaparib; Nivolumab; Placebo Oral Tablet; Placebo IV Infusion | PFS | 2030/12/30 |
Olaparib | NCT01844986 | 3 | Patients with BRCA mutated advanced ovarian cancer | Olaparib | PFS | 2028/08/29 | |
Olaparib | NCT03737643 | 3 | Patients with newly diagnosed advanced ovarian cancer | Bevacizumab; Durvalumab; Olaparib; Placebo olaparib; Durvalumab placebo; Carboplatin + Paclitaxel | PFS | 2028/03/30 | |
Pamiparib | NCT0351923 | 3 | Chinese participants with recurrent ovarian cancer who achieved a complete response or partial response after platinum-based chemotherapy | Pamiparib; Placebo | PFS | 2026/6/30 | |
Olaparib | NCT04729387 | 3 | Patients with platinum resistant or refractory high-grade serous ovarian cancer, with no germline BRCA mutation detected | Alpelisib; Olaparib; Paclitaxel; Pegylated liposomal doxorubicin | PFS | 2026/01/31 | |
Niraparib | NCT02655016 | 3 | Patients With Advanced Ovarian Cancer Following Response on Front-Line Platinum-Based Chemotherapy | Niraparib; Placebo | PFS | 2025/09/30 | |
Rucaparib | NCT04227522 | 3 | Patients with histologically confirmed, advanced high grade serous or high grade endometrioid ovarian cancer, fallopian tube cancer, primary peritoneal cancer and clear cell carcinoma of the ovary | Rucaparib; Placebo | PFS | 2025/01 | |
Prostate cancer | Saruparib | NCT06120491 | 3 | mCSPC | Saruparib; Placebo; Abiraterone Acetate; Darolutamide; Enzalutamide | rPFS | 2031/04/30 |
Talazoparib | NCT04821622 | 3 | mCSPC | Talazoparib + Enzalutamide; Placebo + Enzalutamide | rPFS | 2027/08/07 | |
Niraparib | NCT04497844 | 3 | Patients with deleterious germline or somatic HRR gene-mutated mCSPC | Niraparib; Abiraterone acetate; Prednisone; Placebo for Niraparib | rPFS | 2027/05/27 | |
Niraparib | NCT03748641 | 3 | mCRPC | Niraparib; Abiraterone Acetate; Prednisone; Placebo; New Formulation of Niraparib and Abiraterone Acetate | rPFS | 2027/02/19 | |
Rucaparib | NCT04455750 | 3 | mCRPC | Enzalutamide; Rucaparib camsylate; Placebo; Leuprolide Acetate; Goserelin Acetate; Degarelix | rPFS, OS | 2026/09/01 | |
Talazoparib | NCT03395197 | 3 | mCRPC | Talazoparib + enzalutamide; Placebo + enzalutamide | rPFS | 2025/12/31 | |
Olaparib | NCT03150576 | 3 | Patients with TNBC and/or gBRCA breast cancer | Olaparib; Paclitaxel + Carboplati | Number of participants with treatment-related adverse events as assessed by NCI CTCAE v4.03; pCR rate | 2034/06 | |
Breast cancer | Saruparib | NCT06380751 | 3 | Patients with BRCA1, BRCA2, or PALB2m, HR-positive, HER2-negative advanced breast cancer | Saruparib; Camizestrant; Abemaciclib; Ribociclib; Palbociclib; Fulvestrant; Letrozole; Anastrozole; Exemestane | PFS | 2030/10/18 |
Olaparib | NCT02032823 | 3 | Patients with gBRCA1/2 mutations and high risk HER2 negative primary breast cancer | Olaparib; Placebo | iDFS | 2029/05/28 | |
Niraparib | NCT04915755 | 3 | Participants with either HER2-negative BRCA-mutated or triple-negative breast cancer with molecular disease | Niraparib; Placebo | Number of participants with treatment emergent adverse event (TEAEs), serious adverse events (SAEs), and adverse events of special interest (AESIs) | 2025/12/31 | |
Niraparib | NCT04475939 | 3 | Patients with advanced or metastatic non-small cell lung cancer | Niraparib; Pembrolizumab; Placebo | PFS, OS | 2025/02/19 | |
Lung cancer | Niraparib | NCT04475939 | 3 | Patients with advanced or metastatic non-small cell lung cancer | Niraparib; Pembrolizumab; Placebo | PFS, OS | 2025/02/19 |
Abbreviations: iDFS Invasive disease-free Survival, PFS Progression-free Survival, OS: Overall survival, rPFS Radiological progression free survival, mCRPC Metastatic castration-resistant prostate cancer, NHA: New hormonal agent, mCSPC Metastatic castration-sensitive prostate cancer, HRR Homologous Recombination Repair
Table 2.
Comprehensive List of Clinical Trials Involving Combination Therapies with PARP Inhibitors
Cancer type | Year | Phase | PARPi agent | Combination agent | Combination therapy | No. of Patients | Design | Primary endpoints | Results | Recruitment status | Identifiers | |
---|---|---|---|---|---|---|---|---|---|---|---|---|
First submitted | Last verified | |||||||||||
Ovarian, Fallopian Tube, Peritoneal cancer | 2015 | 2023 | 3 | Veliparib | Carboplatin and Paclitaxel | PARPi + Platinum chemotherapeutic + Anti-microtubule agent | 1140 | Previously untreated stages III or IV high-grade serous epithelial OC, fallopian tube, or primary peritoneal cancer | PFS | Veliparib + Carboplatin + Paclitaxel → Veliparib: 34.7 (31.8-NA), Mo; Veliparib + Carboplatin + Paclitaxel → Placebo: 21.1 (17.0–25.5), Mo | Terminated | NCT02470585 |
2015 | 2024 | 2/3 | Olaparib | Cediranib | PARPi + VEGFRi | 587 | Recurrent platinum-resistant or refractory ovarian, fallopian tube, or primary peritoneal cancer | PFS, OS | NA | Active, not recruiting | NCT02502266 | |
2015 | 2024 | 3 | Olaparib | Cediranib | PARPi + VEGFRi | 579 | Recurrent platinum-sensitive ovarian, fallopian tube, or primary peritoneal cancer | PFS | 10.4 (8.5–12.5), Mo | Active, not recruiting | NCT02446600 | |
2017 | 2022 | 3 | Olaparib | Cediranib | PARPi + VEGFRi | 330 | Relapsed platinum-sensitive ovarian, fallopian tube or peritoneal cancer | PFS | NA | Unknown status | NCT03278717 | |
2018 | 2023 | 3 | Talazoparib | Avelumab + Carboplatin + Paclitaxel | PARPi + immunotherapy + Platinum chemotherapeutic + Anti-microtubule agent | 79 | Stage III/IV epithelial OC, fallopian tube, or primary peritoneal cancer with high-grade serous component | PFS | NA | Terminated | NCT03642132 | |
2021 | 2024 | 3 | Olaparib | Alpelisib | PARPi + PI3 Ki | 358 | Platinum resistant or refractory high-grade serous or endometrioid OC, fallopian tube cancer, or primary peritoneal cancer with no gBRCAm detected | PFS | NA | Active, not recruiting | NCT04729387 | |
Breast Cancer | 2014 | 2023 | 3 | Veliparib (ABT-888) | Carboplatin + Paclitaxel | PARPi + Platinum chemotherapeutic + Anti-microtubule agent | 509 | HER2-negative metastatic or locally advanced unresectable BRCA-associated BC | PFS | 14.5 (12.5–17.7), Mo | Completed | NCT02163694 |
2024 | 2024 | 3 | Fluzoparib | Dalpiciclib + Fulvestrant/AI | PARPi + CDK4/6i + Endocrine therapy | 307 | HR +/HER2- advanced BC | PFS | NA | Active, not recruiting | NCT06612814 | |
2024 | 2024 | 3 | Fluzoparib | Camrelizumab | PARPi + immunotherapy | 310 | High-risk non-pCR TNBC | iDFS | NA | Recruiting | NCT06533384 | |
Prostate Cancer | 2019 | 2024 | 3 | Niraparib | Abiraterone Acetate | PARPi + ARSI | 750 | Metastatic hormone-sensitive and castration-resistant prostate cancer | PFS | NA | Recruiting | NCT03903835 |
2021 | 2024 | 3 | Talazoparib | Enzalutamide | PARPi + ARSI | 599 | DDR gene mutated mCSPC | Radiological PFS | NA | Active, not recruiting | NCT04821622 | |
2024 | 2024 | 3 | Niraparib | Abiraterone Acetate | PARPi + ARSI | 8000 | Metastatic hormone sensitive prostate cancer | Radiographic PFS, OS | NA | Recruiting | NCT06320067 | |
Non-small cell lung cancer | 2020 | 2024 | 3 | Olaparib | Pembrolizumab + Carboplatin + Cisplatin | PARPi + immunotherapy + Platinum chemotherapeutic + Radiotherapy | 870 | Unresectable, locally advanced, stage III NSCLC | PFS, OS | NA | Active, not recruiting | NCT04380636 |
Uterine leiomyosarcoma | 2022 | 2024 | 2/3 | Olaparib | Temozolomide | PARPi + Alkylating agents | 190 | Advanced uterine leiomyosarcoma after progression on prior chemotherapy | PFS, OS | NA | Active, not recruiting | NCT05432791 |
Endometrial, ovarian carcinosarcoma | 2018 | 2024 | 2/3 | Niraparib | Dostarlimab (TSR-042) | PARPi + immunotherapy | 196 | Metastatic or recurrent endometrial or ovarian carcinosarcoma | ORR, OS | NA | Active, not recruiting | NCT03651206 |
Abbreviations: PFS progression-free survival; OS, overall survival; ORR, objective response rate, iDFS invasive disease-free survival; BC, breast cancer; OC, ovarian cancer, mCSPC metastatic castration-sensitive prostate cancer; Mo, Month; NA, not available, NSCLC non-small cell lung cancer, ARSI androgen receptor signaling inhibitors, gBRCAm germline BRCA mutated, DDR DNA damage repair
Combining PARP inhibitors with ICIs
The synergistic potential of PARPi and ICI is underpinned by several interconnected mechanisms (Fig. 4). These integrated mechanisms, including genomic instability-driven immunogenicity, cGAS-STING-IFN-I activation, PD-L1-mediated checkpoint targeting, and TME immune reprogramming, collectively define the synergistic framework of PARPi-ICI combinations, as comprehensively illustrated in Fig. 4. Primarily, PARPi-induced DNA damage escalates genomic instability, leading to an increased TMB [138]. This heightened TMB is a proxy for neoantigen load, a critical factor in ICI efficacy [139]. The augmented neoantigen presentation enhances tumor immunogenicity, potentially sensitizing malignancies to immunotherapeutic interventions. PARPi in LKB1-mutant NSCLC can simultaneously exploit synthetic lethality due to DNA repair deficiencies and restore IFN-γ signaling, creating a more immunogenic tumor microenvironment that synergizes with PD-1 blockade [140].
Fig. 4.
Mechanisms underlying the synergistic effects of PARP inhibitors with immunotherapies. PARP inhibitors enhance anti-tumor immunity by (1) increasing tumor mutational burden (TMB) and neoantigen production through DNA damage-induced genomic instability; (2) activating the cGAS-STING pathway via cytoplasmic double-stranded DNA accumulation, thereby stimulating type I interferon (IFN-I) production to promote dendritic cell maturation, cytotoxic T lymphocyte activation, and chemokine-mediated T cell infiltration; (3) upregulating PD-L1 expression via both IFN-dependent and DNA damage-triggered pathways, creating targetable immune checkpoints for PD-1/PD-L1 blockade; and (4) remodeling the tumor microenvironment by shifting chronic inflammation to acute Th1-polarized immunity, enhancing CD8 + T cell infiltration, suppressing myeloid-derived suppressor cells (MDSCs)
PARPi-mediated DNA lesions trigger the cytosolic DNA sensing pathway. Accumulation of cytoplasmic double-stranded DNA activates the cGAS-STING axis, culminating in robust type I interferon (IFN-I) production [141–145]. This interferon surge orchestrates a cascade of immunostimulatory events: enhancing dendritic cell maturation and antigen presentation, augmenting T cell trafficking via chemokine modulation, and bolstering cytotoxic T lymphocyte function through elevated perforin and granzyme B expression [146, 147]. Moreover, IFN-I attenuates regulatory T cell suppressive capacity, further tilting the balance towards an anti-tumor immune response. A preclinical trial demonstrated the therapeutic efficacy of the combination of PARPi and STING agonist, which can overcome TME-dependent adaptive resistance to Olaparib by remodeling TME and reprogramming myeloid cells [136]. Currently, several clinical trials have been conducted on STING agonists either as monotherapy or in combination therapies. For example, a phase I trial (NCT04144140) assessed the safety and clinical activity of a STING agonist E7766 in patients with advanced solid tumors or lymphomas. Another phase I/II trial (NCT04020185) evaluated the efficacy of a STING agonist IMSA101 monotherapy and combination therapy with ICI. However, no clinical trials have been initiated to evaluate the combination of STING agonists with PARPi, and the efficacy of such combination therapy remains unknown.
PARPi treatment has been observed to upregulate PD-L1 expression on tumor cells through both IFN-dependent and independent mechanisms [148]. While this might initially appear counterproductive, it provides additional targets for anti-PD-1/PD-L1 therapies, potentially enhancing their efficacy. This adaptive response to PARPi-induced DNA damage creates a therapeutic window for ICI intervention.
Furthermore, PARPi administration instigates a profound remodeling of the tumor immune microenvironment. In contrast to the chronic, low-grade inflammation associated with baseline DNA damage, PARPi elicits an acute inflammatory response. This shift from smoldering inflammation to an acute surge recalibrates the immune landscape, fostering a Th1-polarized milieu conducive to anti-tumor immunity [149]. The resultant microenvironment is characterized by enhanced CD8 + T cell infiltration, diminished myeloid-derived suppressor cell presence, and altered macrophage phenotypes [150]. The intricate interplay between DNA damage response pathways and immune system modulation forms the cornerstone of PARPi-ICI combinatorial strategies [137, 151]. By leveraging these multifaceted mechanisms, this therapeutic approach aims to overcome immune evasion strategies employed by malignant cells and reinvigorate anti-tumor immune responses.
The ongoing phase III trials ATHENA and FIRST are evaluating PARPi combined with ICIs in newly diagnosed advanced ovarian cancer. FIRST (NCT03602859) is a randomized, adaptive phase III trial investigating Rucaparib plus Nivolumab, while ATHENA (NCT03522246) is a randomized, placebo-controlled, four-arm phase III trial studying Niraparib plus Dostarlimab. FIRST included 1402 ovarian cancer patients, while ATHENA included 1097 patients. Both trials focus on patients with advanced (FIGO stage III or IV) high-grade serous or endometrioid ovarian cancer who have responded to first-line platinum-based chemotherapy. These studies aim to determine if the combination therapy improves progression-free survival compared to PARP inhibitor monotherapy or placebo. Results are expected around 2024 and could significantly impact first-line treatment strategies in ovarian cancer, potentially improving outcomes for patients with advanced ovarian cancer.
PARP inhibitors with DNA damage response inhibitors
The strategic combination of DNA damage response inhibitors, such as ATR and CHK1, with PARPi emerges as a promising approach to overcome PARPi resistance and increase therapeutic efficacy.
ATR regulates HR repair by phosphorylating key HR-related proteins. One of its downstream targets, CHK1, enhances the transcription of RAD51, thereby facilitating the HR repair process [152]. Up to now, combining PARPi with ATR, CHK1, AXL inhibitors, and other targets has shown promising results in cancer therapy. (Fig. 5) [153]. In line with these findings, several studies have demonstrated that ATR and CHK1 inhibitors induce synthetic lethality with PARP inhibitors by causing HRD, highlighting their therapeutic potential to overcome PARP inhibitor resistance [154, 155]. SLFN11 is actively recruited to DNA damage sites to inhibit HR. Loss of SLFN11, which is common in cancer cells, could confer resistance to PARPi [156]. However, the addition of ATRi VE-821 can reverse such resistance, providing an option in treating SLFN11-negative cancer [157]. In ovarian cancer, the CHK1 inhibitor SRA737 has shown notable single-agent activity in PARPi-resistant cells. Although the synergistic effects of PARPi and CHK1 inhibitors (CHK1i) are well known, it has been shown that SRA737 can sensitize PARP inhibitor-resistant ovarian cancer by increasing replication stress and the accumulation of DNA double-strand breaks [158].
Fig. 5.
Mechanisms underlying the synergistic effects of PARP Inhibitors in combination with other drugs. Mechanisms involve targeting various cellular processes to overcome resistance and enhance therapeutic efficacy. These strategies include suppression of HR pathway reactivation through tyrosine kinase inhibitors and epigenetic regulators, enhancement of PARPi cytotoxicity via NAD + synthesis inhibition, and disruption of restored replication fork protection in resistant cells by targeting factors such as ATR and RAD52. The inhibition of alternative DNA repair pathways, particularly microhomology-mediated end-joining (MMEJ), presents a novel approach, with polymerase θ (POLQ) inhibitors emerging as promising agents. Importantly, combination therapies with PARPi can also modulate the tumor immune microenvironment. Tyrosine kinase inhibitors used in combination with PARPi may alter the infiltration and function of CD4 + and CD8 + T cells. Epigenetic drugs, such as HDAC inhibitors, when combined with PARPi, can enhance the expression of immunogenic markers on tumor cells, potentially increasing their recognition by immune cells. Some DNA damage response inhibitors used in combination strategies may influence the activation state of tumor-infiltrating lymphocytes and alter the balance between effector and regulatory T cells
WEE1 can inhibit CDK and activate the cell cycle checkpoint to allow DNA damage repair. Therefore, inhibiting WEE1 or CDK, which can induce cell cycle arrest, apoptosis, and suppression of transcription [159], presents another potential strategy for combination therapy [160]. Several studies have shown that combining PARPi with WEE1 inhibitors induces synthetic lethality, supporting the therapeutic potential of this combination for cancer treatment [153, 154]. The underlying mechanisms of overcoming PARPi resistance have been explored. Targeting WEE1 can downmodulate the key enzyme in dNTP synthesis RRM2 thereby inducing replication fork stalling and replication stress response, delaying the acquired resistance to PARPi in BRCA1-altered PDX models [161]. WEE1 inhibition sensitize BRCA1/2 wild-type TNBC to PARPi through increasing DNA damage, apoptosis, replication stress, and STING pathway activation [162]. A growing body of research is highlighting the potential of combining PARPi and CDK inhibitors (CDKi), providing new hope and perspectives for cancer treatment. Talazoparib-induced apoptosis was increased by Palbociclib, a CDK4/6 inhibitor. Talazoparib/Palbociclib combination therapy can induce both G0/G1 and G2 arrest, thereby leading to synergism. The combination of Palbociclib and Talazoparib effectively enhances bladder cancer therapy, and RB is a molecular biomarker of response to this treatment regimen [163]. Besides, in Olaparib-resistant BRCA-deficient TNBC models, CDKi can reinstate sensitivity to Olaparib by reverting the functional HR restoration, which is frequently associated with PARPi resistance [164]. Dinaciclib, a CDK inhibitor that can downregulate MYC expression and impair the HR pathway by inhibiting RAD51 and BRCA1 activity, induces TNBC sensitivity to PARPi. Treatment with Dinaciclib and Niraparib can overcome acquired resistance to PARPi [165].
Other DNA damage response inhibitors are also emerging as promising candidates in combination therapy, opening up exciting new possibilities for treatment. For example, by preventing RPA2/CHK1-mediated HR, AXL inhibitor Bemcentinib can augment PARPi Olaparib sensitivity, uncovering that AXL acts as a prognostic biomarker. A combination therapy with Bemcentinib can expand the application of PARPi in HCC patients, especially those with AXL expression [166]. Numerous studies have shown that RAD51 plays a key role in identifying PARP inhibitor sensitivity, correlating with the functional status of BRCA1/2 proteins [167–169]. Silencing of RAD51 C enhanced the sensitivity to Olaparib, and RAD51 C deficiency may be considered a biomarker for predicting the efficacy of treating with Olaparib [170]. In agreement with these observations, Ksenija et al. found that homozygous RAD51 C methylation is a positive predictive biomarker for sensitivity to PARPi [171]. Besides the dominant DNA repair pathway HR, DSBs can be repaired by other pathways, including NHEJ and microhomology-mediated end joining (MMEJ). These pathways promote DSB repair, especially in HR-deficient cells. Activation or upregulation of the main factor of these pathways can promote DNA repair and compensate HR pathway [172]. Upregulation of DNA polθ, a crucial factor in MMEJ, results in enhancement of DNA repair and Olaparib resistance in HR-deficient ovarian cancer cells [173]. Thereby, targeting DNA polθ can suppress DSB repair and improve PARPi therapeutic efficacy. Furthermore, a study has demonstrated that ART558, a Polθ inhibitor, enhances the effects of a PARPi, presenting a promising strategy to overcome PARPi resistance in HRR-deficient cancer [174].
ATM is a nuclear protein that participates in the initiation of DNA repair signaling and cell-cycle checkpoints during DNA damage. ATM deficiency was found to be associated with enhanced sensitivity to Veliparib and TOP1 inhibitor irinotecan. Combining PARP and TOP1 inhibitors could improve drug efficacy and increase the sensitivity of gastric cancer cells to chemotherapy in ATM-driven cancers, offering a potential strategy to overcome drug resistance [175]. Chu et al. discovered that ALK promotes HRR and PARP inhibitor resistance by phosphorylating CDK9 at tyrosine-19. Moreover, inhibiting ALK was shown to overcome PARP inhibitor resistance. These results suggest that ALK is a predictive biomarker for PARP inhibitor resistance, and that combining ALK and PARP inhibitors may represent a promising approach to enhance the effectiveness of PARPi [176].
Several studies have indicated that small molecules engaged in DNA damage response can serve as predictive biomarkers, potentially predicting the efficacy of PARPi treatment. BRCA1 expression continues to be one of the most widely studied biomarkers, despite the exploration of many others. Loss of BRCA1 methylation and biallelic loss of BRCA1/2 may serve as potential markers of PARP inhibitor resistance [177]. Additionally, the expression level of other molecules, like GLI1 [178], LIG3 [179], PPP2R2 A [180], ERCC1 [181], MED12 [182], and others, may also represent prognostic and predictive biomarkers for PARPi treatment.
PARPi and writers/erasers epigenetic drugs
Epigenetic dysregulation, recognized as a pivotal factor in tumor genesis and maintenance, regulates PARPi sensitivity. This presents opportunities for combining epigenetic drugs with PARPi to overcome acquired resistance to therapy, as demonstrated in studies of DNMT and HDAC inhibitors [18, 183, 184]. Up to now, a variety of epigenetic drugs have come into focus, showing promising therapeutic effects when combined with PARPi. Some epigenetic drugs, like DNMT, HDAC, and EZH2 inhibitors, have already been approved by the FDA in cancer therapy [185, 186].
BET inhibitors (BETi) disrupt the binding of BET proteins to acetylated lysine residues on chromatin, thereby suppressing the transcription of various genes. The synergy between BETi and PARPi has triggered synthetic lethality and improved therapeutic sensitivity [187]. Whereas utilizing BETi in conventional chemotherapy has shown limited effectiveness, there is a growing need to explore additional combination strategies and identify sensitive biomarkers [188]. Guadecitabine, a DNMT inhibitor, in combination with Talazoparib, has been shown to enhance the efficacy of PARPi and further sensitize breast and ovarian cancer cells to PARPi regardless of the BRCA status. Possible mechanisms are PARP trapping, ROS accumulation, and subsequent PKA activation [189]. However, a recent study demonstrated that malignant DNMT3 A-deficient leukemia cells were resistant to PARPi treatment both in vitro and in vivo [190]. HDAC inhibitors (HDACi) can inhibit histone deacetylases, thereby regulating gene expression [191]. In a preclinical trial, HDACi SAHA can enhance the anti-tumor effects of Olaparib in TNBC cells. The expression level of functional PTEN may serve as a biomarker to predict response to this combination therapy [192]. Similar results were also observed in the MDA-MB-231 xenograft model in vivo [193]. Finally, studies have shown that the downregulation of EZH2, which can methylate H3 K27 and exert a transcriptional repressor activity, prevents MUS81 recruitment and causes replication fork stabilization, promoting PARPi resistance in BRCA2-deficient cells [89].
Targeting other repair pathways
In addition to the aforementioned strategies, targeting other repair pathways could also hold significant promise for combination therapy. Vascular endothelial growth factor (VEGF) is a secreted molecule that contributes to angiogenesis and vascular permeability, which participates in oncogenesis. A combination of VEGF and VEGFR can enhance PI3 K/AKT signaling [194]. Additionally, VEGFR and PI3 K/AKT pathways have been shown to modulate HR function, providing rationale for combination therapy with PARPi [195]. VEGFR3 inhibition can downregulate BRCA1 and BRCA2 mRNA, resulting in cell cycle arrest and chemosensitization [196]. Another study demonstrated that VEGF/VEGFR/PI3 K signaling enhances HR activity by upregulating CRY1 expression [197], with parallel research showing spliceosome components such as SF3B4 regulate HR-related transcripts via alternative splicing [198]. Interestingly, several evidence have uncovered the interaction of androgen receptor (AR) signaling and HRR pathway [199, 200]. AR signaling activation can upregulate HRR genes, including BRCA1 and RAD51, so inhibiting this signaling may downregulate HRR genes. AR signaling inhibition may also upregulate PARP and induce sensitivity to PARPi, indicating the potential role of AR signaling inhibitors (ARSI) in reversing PARPi resistance [201]. These complementary mechanisms jointly promote DNA repair and cell survival.
A phase IIb clinical trial combining anti-VEGFR Cediranib and Olaparib in patients with ovarian cancer showed clinical activity despite failing to meet the target ORR of 20%. Further exploration for combination therapy and biomarkers are required [202]. Several clinical studies have shown that combining PARP inhibitors with PI3 K/AKT pathway inhibitors results in synergistic anti-tumor activity. In a phase I clinical trial, antitumor activity was observed when combining PARPi and AKT inhibitor Capivasertib in patients with BRCA1/2- and non-BRCA1/2-Mutant Cancers [203]. Furthermore, preliminary clinical evidence suggests synergistic activity between Olaparib and the PI3 K inhibitor Alpelisib [204]. Additionally, a phase III clinical trial has been initiated investigating the combination of PARP and PI3 K inhibition in ovarian cancer (NCT04729387). This study included patients with platinum-resistant or refractory high-grade serous ovarian cancer with no germline BRCA mutation detected, aiming to assess the efficacy and safety of the combination of Alpelisib and Olaparib. The primary endpoint is progression-free survival, and the estimated completion date of this study is June 2025. Two phase III trials are assessing the therapeutic effect of PARPi combined with ARIS in metastatic castration-sensitive prostate cancer. Both of these trials investigated whether the combination of Talazoparib with ARSI Enzalutamide improves patients’ survival compared to Enzalutamide monotherapy. Results of one trial demonstrated a hazard ratio (HR) of 0.447 in favor of the Talazoparib + Enzalutamide group compared to the placebo + Enzalutamide, indicating a formidable therapeutic effect of combination therapy (NCT03395197). Another trial is still ongoing, with an estimated completion date of August 2027 (NCT04821622). These trials may provide novel therapeutic options for combination therapy.
PARPi and ionizing radiation (IR)
IR can induce DNA damage and trigger cell death. The combination of PARPi and IR could enhance the cytotoxicity mediated by reactive oxygen species, preferentially targeting cancer cells with mutant TP53 [205].
A study in pancreatic ductal adenocarcinoma suggests that combining radiotherapy with the PARP inhibitor Olaparib could enhance treatment efficacy, with DDB2 expression serving as a potential predictive biomarker for response. Through patient stratification by DDB2 expression level, it is possible to expand the therapeutic options for Olaparib as a radiosensitizer in a wide range of patients [206]. Niraparib BMN673, a PARPi with unique radiosensitizing ability via unique mechanisms causing profound shifts in the balance among DNA double-strand break (DSB) repair pathways, sheds light on combining PARPi and IR to overcome PARPi resistance [207]. Combining Niraparib with IR shows significant potential, as indicated by several preclinical and clinical studies. However, further research is needed to identify biomarkers that could predict the response to this combination.
Future direction
Recent advances in understanding PARPi mechanisms and resistance patterns have significantly improved their clinical application. However, as PARP inhibitors expand beyond breast and ovarian cancers, acquired resistance remains a critical challenge, necessitating further investigation [208]. While laboratory studies have identified various resistance mechanisms, clinical scenarios often differ due to prior treatment-induced cross-resistance. This discrepancy underscores the need for more comprehensive clinical studies to elucidate real-world resistance patterns.
Future research directions should focus on several key areas to address these challenges. Longitudinal studies will be crucial to track the evolution of resistance mechanisms throughout treatment courses, particularly comparing first-line versus later-line PARPi use. For molecularly defined subgroups such as MSI-H pancreatic cancer, dedicated trials are needed to validate ICI as the preferred first-line therapy over PARPi. Additionally, developing combinatorial strategies (e.g., PARPi + ICI in MSS/HRD tumors vs. ICI monotherapy in MSI-H tumors) may maximize therapeutic efficacy while avoiding futile treatments. Utilizing liquid biopsy analysis of circulating tumor DNA can provide non-invasive, real-time monitoring of resistance development. Advanced single-cell omics technologies will enable the dissection of intratumoral heterogeneity and the identification of resistant subclones.
In-depth analysis of exceptional responders to PARPi may uncover novel sensitivity determinants, potentially leading to new therapeutic strategies. Further exploration of synergistic drug combinations, particularly concerning POLQ inhibitors and other emerging targets, is warranted. Identifying and validating predictive biomarkers for PARPi response and resistance will be essential for guiding personalized treatment approaches.
Investigating novel drug delivery systems, such as nanoparticle-based or targeted delivery approaches, as well as leveraging tumor-associated microbiota as a delivery pathway, may enhance PARPi efficacy and overcome resistance mechanisms [209–212]. A recent study developed a genomic-guided biomimetic co-delivery system using B7H3-targeted metal–organic frameworks (MOFs) to simultaneously deliver CDK4/6 inhibitors and PARP inhibitors, which not only achieved tumor microenvironment-responsive drug release but also amplified DNA damage through photodynamic therapy while reprogramming the immunosuppressive microenvironment via enhanced CD8 + T cell infiltration, demonstrating a 4.2-fold improvement in tumor targeting efficiency compared to non-targeted nanoparticles [213]. Nanoparticles can be designed to deliver RNA-based therapeutics (such as siRNA or miRNA) that target resistance mechanisms, such as restored HRR (via BRCA1/2 or RAD51 overexpression). In addition, nanoparticles could deliver siRNA targeting RAD51 to inhibit the HR pathway, thereby sensitizing tumors to PARPi [214, 215]. Preclinical models demonstrate that PARPi-loaded nanoparticles prolong drug retention in the cerebrospinal fluid with eightfold tumor accumulation, significantly enhancing antitumor activity over free drugs in BRCA-deficient cancers [216]. Furthermore, nanoparticles co-delivering PARPi and immune checkpoint inhibitors boost tumor penetration by 3.7-fold through hyaluronidase-mediated matrix remodeling while blocking PD-L1 upregulation, leading to sustained cytotoxic T-cell infiltration and improved survival [217, 218]. These advances underscore the potential of nanocarriers to overcome drug solubility limitations, minimize off-target effects, and enable combination therapies through controlled spatiotemporal release.
Examining the role of the tumor immune microenvironment in PARPi resistance and exploring immunomodulatory combination strategies could yield valuable insights. Tumor-associated fibroblasts (TAFs) and the extracellular matrix (ECM) contribute to the tumor microenvironment's mechanical properties, affecting drug delivery, immune infiltration, and PARPi resistance. TAFs can produce collagen, fibronectin, and other ECM components, creating physical barriers to immune cell infiltration and drug diffusion [219]. Developing new strategies to target fibroblasts or alter ECM composition would enhance the penetration of PARPi into the tumor and promote more efficient immune cell infiltration. In addition, cytokines such as IL-10, TGF-β, and IL-6 are abundant in the tumor microenvironment and play critical roles in immune suppression and PARPi resistance. Cytokine-targeted therapies can be developed to block immunosuppressive cytokines and boost antitumor immunity. Furthermore, TGF-β is a key mediator of immune suppression in tumors, and blocking its signaling can reverse immune evasion, sensitize tumors to immune checkpoint inhibitors, and enhance the efficacy of PARPi.
Targeting key signaling pathways involved in DNA repair, cell cycle regulation, and tumor survival can sensitize tumors to PARPi by inhibiting alternative repair pathways or overcoming resistance mechanisms. Targeting specific kinases may provide a powerful tool to enhance PARP therapy efficacy. For example, Olaparib has been tested in the presence of Lapatinib (HER2 Inhibitor). Clinical studies in HER2-positive breast cancer are evaluating the synergistic effects of Olaparib and Lapatinib. This combination targets HER2-mediated signaling and DNA repair, aiming to overcome resistance and increase treatment efficacy. In some leukemia models, the combination of Imatinib (BCR-ABL inhibitor) and Olaparib is being tested in clinical trials, particularly in CML and Ph + ALL, where BCR-ABL signaling contributes to DNA repair and survival, and PARPi can exacerbate DNA damage.
Leveraging artificial intelligence and machine learning algorithms to predict resistance patterns and optimize treatment strategies represents an exciting frontier in PARPi research [220]. Cutting-edge AI tools are revolutionizing PARPi drug design. AlphaFold2 now accurately predicts PARP1–ligand binding interfaces, while Rosetta’s COMBS algorithm optimizes molecular interactions [221, 222]. Together, these technologies enabled de novo creation of PARPi-binding proteins (PiB) with sub-nanomolar affinity. By analyzing data from BRCA-mutant cancers (including single-cell profiles and patient treatment histories), AI systems can forecast resistance development with 85% accuracy. These models guide the engineering of “PARPi-capturing” proteins, whose drug-binding strength is enhanced by simulating key molecular interactions like the Q54-D58-N131 network [221]. In lab tests, such AI-designed proteins restored PARPi sensitivity in resistant cells, cutting the required drug dose by 78% [221]. Meanwhile, AI accelerates personalized treatment by predicting optimal drug concentrations for individual patients. Advanced models calculate drug-protein binding efficiency with near-atomic precision, enabling tailored dosing strategies. By combining computational predictions with experimental validation, AI is proving indispensable to overcoming resistance and improving PARPi therapies.
By pursuing these multifaceted research directions and integrating insights from basic science, translational research, and clinical observations, we can develop more effective strategies to overcome PARPi resistance and improve patient outcomes. Furthermore, understanding the mechanisms of PARPi resistance may provide valuable insights into cancer biology and DNA repair pathways, potentially leading to novel therapeutic approaches beyond PARPi. As we unravel the complexities of PARPi resistance, a collaborative, interdisciplinary approach will be essential to translate these findings into meaningful clinical advancements.
Acknowledgements
None.
Abbreviations
- 2-DG
2-Deoxyglucose
- 3-BrPA
3-Bromopyruvate
- 53BP1
P53-binding protein 1
- ABC
ATP-binding cassette
- AML
Acute myeloid leukemia
- AR
Androgen receptor
- ARSI
Androgen receptor signaling inhibitors
- ATLL
Adult T-cell leukemia/lymphoma
- BER
Base Excision Repair
- BETi
BET inhibitors
- CDKi
CDK inhibitors
- cfDNA
Cell-free DNA
- CHK1i
CHK1 inhibitors
- CST
CTC1-STN1-TEN1
- D-2HG
D-2-hydroxyglutarate
- DDR
DNA Damage Response
- DSBs
Double-strand breaks
- ECM
Extracellular matrix
- FAO
Fatty acid oxidation
- HDACi
HDAC inhibitors
- HR
Homologous Recombination
- HRD
Homologous recombination deficiency
- HRR
Homologous recombination repair
- ICIs
Immune checkpoint inhibitors
- IDH
Isocitrate dehydrogenase
- IFN-I
Type I interferon
- IR
Ionizing radiation
- LOH
Loss of heterozygosity
- LST
Large-scale transitions
- mCRPC
Metastatic castration-resistant prostate cancer
- MDS
Myelodysplastic syndromes
- meRAD51 C
RAD51C promoter methylation
- MMR
Mismatch Repair
- MMEJ
Microhomology-mediated end joining
- MSI-H
Microsatellite instability-high
- NAMPT
Nicotinamide phosphoribosyltransferase
- NAPRT1
Nicotinate phosphoribosyltransferase
- NHEJ
Non-Homologous End Joining
- NSCLC
Non-small cell lung cancer
- PARP
Poly (ADP-ribose) polymerase
- PARPi
PARP inhibitors
- PDX
Patient-derived xenograft
- PiB
PARPi-binding proteins
- SLFN11
Schlafen-11
- SSBs
Single-strand breaks
- TAFs
Tumor-associated fibroblasts
- TAI
Telomeric allelic imbalance
- TIM
TIMELESS
- TMB
Tumor mutational burden
- TME
Tumor microenvironment
- TNBC
Triple-negative breast cancer
- TRCs
Transcription-replication conflicts
- VEGF
Vascular endothelial growth factor
- VEGFR
Vascular endothelial growth factor receptor
Authors’ contributions
Conceptualization and design: H.T., C.N., X.L., and Z.G.; Literature search: H.Z. and P.C.; Data interpretation: Y.Z. and H.Z.; Writing: Y.Z., H.Z., P.C., J.T. and S.Y.; Reviewing and editing: C.N., H.T., X.L., and Z.G. All authors read and approve the manuscript.
Funding
This research was funded by the Guangdong Basic and Applied Basic Research Foundation (2022B1515120087, Hailin Tang; 2021B1515120053, Ziyun Guan; 2025A1515012712, Yutian Zou), the National Natural Science Foundation of China (No. 82403976, Yutian Zou) and the China National Postdoctoral Program for Innovative Talents (No. BX20230447, Yutian Zou).
Data availability
No datasets were generated or analysed during the current study.
Declarations
Ethics approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Yutian Zou, Hanqi Zhang and Pangzhou Chen contributed equally to this work.
Contributor Information
Ziyun Guan, Email: lyguanzy@scut.edu.cn.
Xing Li, Email: lixing@sysucc.org.cn.
Hailin Tang, Email: tanghl@sysucc.org.cn.
References
- 1.Monk BJ, Parkinson C, Lim MC, O’Malley DM, Oaknin A, Wilson MK, et al. A Randomized, Phase III Trial to Evaluate Rucaparib Monotherapy as Maintenance Treatment in Patients With Newly Diagnosed Ovarian Cancer (ATHENA-MONO/GOG-3020/ENGOT-ov45). J Clin Oncol. 2022;40(34):3952–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Fizazi K, Piulats JM, Reaume MN, Ostler P, McDermott R, Gingerich JR, et al. Rucaparib or Physician’s Choice in Metastatic Prostate Cancer. N Engl J Med. 2023;388(8):719–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Agarwal N, Azad AA, Carles J, Fay AP, Matsubara N, Heinrich D, et al. Talazoparib plus enzalutamide in men with first-line metastatic castration-resistant prostate cancer (TALAPRO-2): a randomised, placebo-controlled, phase 3 trial. Lancet. 2023;402(10398):291–303. [DOI] [PubMed] [Google Scholar]
- 4.Monk BJ, Barretina-Ginesta MP, Pothuri B, Vergote I, Graybill W, Mirza MR, et al. Niraparib first-line maintenance therapy` in patients with newly diagnosed advanced ovarian cancer: final overall survival results from the PRIMA/ENGOT-OV26/GOG-3012 trial. Ann Oncol. 2024;35(11):981–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Pujade-Lauraine E, Selle F, Scambia G, Asselain B, Marmé F, Lindemann K, et al. Maintenance olaparib rechallenge in patients with platinum-sensitive relapsed ovarian cancer previously treated with a PARP inhibitor (OReO/ENGOT-ov38): a phase IIIb trial. Ann Oncol. 2023;34(12):1152–64. [DOI] [PubMed] [Google Scholar]
- 6.Besse B, Pons-Tostivint E, Park K, Hartl S, Forde PM, Hochmair MJ, et al. Biomarker-directed targeted therapy plus durvalumab in advanced non-small-cell lung cancer: a phase 2 umbrella trial. Nat Med. 2024;30(3):716–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Geyer CE Jr, Garber JE, Gelber RD, Yothers G, Taboada M, Ross L, et al. Overall survival in the OlympiA phase III trial of adjuvant olaparib in patients with germline pathogenic variants in BRCA1/2 and high-risk, early breast cancer. Ann Oncol. 2022;33(12):1250–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Banerjee S, Gonzalez-Martin A, Harter P, Lorusso D, Moore KN, Oaknin A, et al. First-line PARP inhibitors in ovarian cancer: summary of an ESMO Open - Cancer Horizons round-table discussion. ESMO Open. 2020;5(6): e001110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Litton JK, Rugo HS, Ettl J, Hurvitz SA, Gonçalves A, Lee KH, et al. Talazoparib in Patients with Advanced Breast Cancer and a Germline BRCA Mutation. N Engl J Med. 2018;379(8):753–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Robson M, Im SA, Senkus E, Xu B, Domchek SM, Masuda N, et al. Olaparib for Metastatic Breast Cancer in Patients with a Germline BRCA Mutation. N Engl J Med. 2017;377(6):523–33. [DOI] [PubMed] [Google Scholar]
- 11.de Bono J, Mateo J, Fizazi K, Saad F, Shore N, Sandhu S, et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N Engl J Med. 2020;382(22):2091–102. [DOI] [PubMed] [Google Scholar]
- 12.Kindler HL, Hammel P, Reni M, Van Cutsem E, Macarulla T, Hall MJ, et al. Overall Survival Results From the POLO Trial: A Phase III Study of Active Maintenance Olaparib Versus Placebo for Germline BRCA-Mutated Metastatic Pancreatic Cancer. J Clin Oncol. 2022;40(34):3929–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Jeggo PA, Pearl LH, Carr AM. DNA repair, genome stability and cancer: a historical perspective. Nat Rev Cancer. 2016;16(1):35–42. [DOI] [PubMed] [Google Scholar]
- 14.van Beek L, McClay É, Patel S, Schimpl M, Spagnolo L, Maia de Oliveira T. PARP Power: A Structural Perspective on PARP1, PARP2, and PARP3 in DNA Damage Repair and Nucleosome Remodelling. Int J Mol Sci. 2021;22(10). [DOI] [PMC free article] [PubMed]
- 15.Pascal JM, Ellenberger T. The rise and fall of poly(ADP-ribose): An enzymatic perspective. DNA Repair (Amst). 2015;32:10–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Rodgers K, McVey M. Error-Prone Repair of DNA Double-Strand Breaks. J Cell Physiol. 2016;231(1):15–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Chen CC, Feng W, Lim PX, Kass EM, Jasin M. Homology-Directed Repair and the Role of BRCA1, BRCA2, and Related Proteins in Genome Integrity and Cancer. Annu Rev Cancer Biol. 2018;2:313–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Bayley R, Sweatman E, Higgs MR. New perspectives on epigenetic modifications and PARP inhibitor resistance in HR-deficient cancers. Cancer Drug Resist. 2023;6(1):35–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Petropoulos M, Karamichali A, Rossetti GG, Freudenmann A, Iacovino LG, Dionellis VS, et al. Transcription-replication conflicts underlie sensitivity to PARP inhibitors. Nature. 2024;628(8007):433–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D, Lopez E, et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature. 2005;434(7035):913–7. [DOI] [PubMed] [Google Scholar]
- 21.Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature. 2005;434(7035):917–21. [DOI] [PubMed] [Google Scholar]
- 22.Hanzlikova H, Kalasova I, Demin AA, Pennicott LE, Cihlarova Z, Caldecott KW. The Importance of Poly(ADP-Ribose) Polymerase as a Sensor of Unligated Okazaki Fragments during DNA Replication. Mol Cell. 2018;71(2):319–31.e3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Miller RE, Leary A, Scott CL, Serra V, Lord CJ, Bowtell D, et al. ESMO recommendations on predictive biomarker testing for homologous recombination deficiency and PARP inhibitor benefit in ovarian cancer. Ann Oncol. 2020;31(12):1606–22. [DOI] [PubMed] [Google Scholar]
- 24.Noordermeer SM, Adam S, Setiaputra D, Barazas M, Pettitt SJ, Ling AK, et al. The shieldin complex mediates 53BP1-dependent DNA repair. Nature. 2018;560(7716):117–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Pettitt SJ, Krastev DB, Brandsma I, Dréan A, Song F, Aleksandrov R, et al. Genome-wide and high-density CRISPR-Cas9 screens identify point mutations in PARP1 causing PARP inhibitor resistance. Nat Commun. 2018;9(1):1849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Onji H, Tate S, Sakaue T, Fujiwara K, Nakano S, Kawaida M, et al. Schlafen 11 further sensitizes BRCA-deficient cells to PARP inhibitors through single-strand DNA gap accumulation behind replication forks. Oncogene. 2024;43(32):2475–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Scattolin D, Maso AD, Ferro A, Frega S, Bonanno L, Guarneri V, et al. The emerging role of Schlafen-11 (SLFN11) in predicting response to anticancer treatments: Focus on small cell lung cancer. Cancer Treat Rev. 2024;128: 102768. [DOI] [PubMed] [Google Scholar]
- 28.Leman R, Muller E, Legros A, Goardon N, Chentli I, Atkinson A, et al. Validation of the Clinical Use of GIScar, an Academic-developed Genomic Instability Score Predicting Sensitivity to Maintenance Olaparib for Ovarian Cancer. Clin Cancer Res. 2023;29(21):4419–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Petrelli A, Rizzolio S, Pietrantonio F, Bellomo SE, Benelli M, De Cecco L, et al. BRCA2 Germline Mutations Identify Gastric Cancers Responsive to PARP Inhibitors. Cancer Res. 2023;83(10):1699–710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Llop-Guevara A, Loibl S, Villacampa G, Vladimirova V, Schneeweiss A, Karn T, et al. Association of RAD51 with homologous recombination deficiency (HRD) and clinical outcomes in untreated triple-negative breast cancer (TNBC): analysis of the GeparSixto randomized clinical trial. Ann Oncol. 2021;32(12):1590–6. [DOI] [PubMed] [Google Scholar]
- 31.Cong K, Peng M, Kousholt AN, Lee WTC, Lee S, Nayak S, et al. Replication gaps are a key determinant of PARP inhibitor synthetic lethality with BRCA deficiency. Mol Cell. 2021;81(15):3128–44.e7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Tutt ANJ, Garber JE, Kaufman B, Viale G, Fumagalli D, Rastogi P, et al. Adjuvant Olaparib for Patients with BRCA1- or BRCA2-Mutated Breast Cancer. N Engl J Med. 2021;384(25):2394–405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.DiSilvestro P, Banerjee S, Colombo N, Scambia G, Kim BG, Oaknin A, et al. Overall Survival With Maintenance Olaparib at a 7-Year Follow-Up in Patients With Newly Diagnosed Advanced Ovarian Cancer and a BRCA Mutation: The SOLO1/GOG 3004 Trial. J Clin Oncol. 2023;41(3):609–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Moore KN, Secord AA, Geller MA, Miller DS, Cloven N, Fleming GF, et al. Niraparib monotherapy for late-line treatment of ovarian cancer (QUADRA): a multicentre, open-label, single-arm, phase 2 trial. Lancet Oncol. 2019;20(5):636–48. [DOI] [PubMed] [Google Scholar]
- 35.Mateo J, de Bono JS, Fizazi K, Saad F, Shore N, Sandhu S, et al. Olaparib for the Treatment of Patients With Metastatic Castration-Resistant Prostate Cancer and Alterations in BRCA1 and/or BRCA2 in the PROfound Trial. J Clin Oncol. 2024;42(5):571–83. [DOI] [PubMed] [Google Scholar]
- 36.Chabanon RM, Muirhead G, Krastev DB, Adam J, Morel D, Garrido M, et al. PARP inhibition enhances tumor cell-intrinsic immunity in ERCC1-deficient non-small cell lung cancer. J Clin Invest. 2019;129(3):1211–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Padella A, Ghelli Luserna Di Rorà A, 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(1):10. [DOI] [PMC free article] [PubMed]
- 38.Gojo I, Beumer JH, Pratz KW, McDevitt MA, Baer MR, Blackford AL, et al. A Phase 1 Study of the PARP Inhibitor Veliparib in Combination with Temozolomide in Acute Myeloid Leukemia. Clin Cancer Res. 2017;23(3):697–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Pratz KW, Rudek MA, Gojo I, Litzow MR, McDevitt MA, Ji J, et al. A Phase I Study of Topotecan, Carboplatin and the PARP Inhibitor Veliparib in Acute Leukemias, Aggressive Myeloproliferative Neoplasms, and Chronic Myelomonocytic Leukemia. Clin Cancer Res. 2017;23(4):899–907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Bai XT, Moles R, Chaib-Mezrag H, Nicot C. Small PARP inhibitor PJ-34 induces cell cycle arrest and apoptosis of adult T-cell leukemia cells. J Hematol Oncol. 2015;8:117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Sokol ES, Jin DX, Fine A, Trabucco SE, Maund S, Frampton G, et al. PARP Inhibitor Insensitivity to BRCA1/2 Monoallelic Mutations in Microsatellite Instability-High Cancers. JCO Precis Oncol. 2022;6: e2100531. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Maio M, Ascierto PA, Manzyuk L, Motola-Kuba D, Penel N, Cassier PA, et al. Pembrolizumab in microsatellite instability high or mismatch repair deficient cancers: updated analysis from the phase II KEYNOTE-158 study. Ann Oncol. 2022;33(9):929–38. [DOI] [PubMed] [Google Scholar]
- 43.Edwards SL, Brough R, Lord CJ, Natrajan R, Vatcheva R, Levine DA, et al. Resistance to therapy caused by intragenic deletion in BRCA2. Nature. 2008;451(7182):1111–5. [DOI] [PubMed] [Google Scholar]
- 44.Tobalina L, Armenia J, Irving E, O’Connor MJ, Forment JV. A meta-analysis of reversion mutations in BRCA genes identifies signatures of DNA end-joining repair mechanisms driving therapy resistance. Ann Oncol. 2021;32(1):103–12. [DOI] [PubMed] [Google Scholar]
- 45.Lheureux S, Prokopec SD, Oldfield LE, Gonzalez-Ochoa E, Bruce JP, Wong D, et al. Identifying Mechanisms of Resistance by Circulating Tumor DNA in EVOLVE, a Phase II Trial of Cediranib Plus Olaparib for Ovarian Cancer at Time of PARP Inhibitor Progression. Clin Cancer Res. 2023;29(18):3706–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Kim YN, Shim Y, Seo J, Choi Z, Lee YJ, Shin S, et al. Investigation of PARP Inhibitor Resistance Based on Serially Collected Circulating Tumor DNA in Patients With BRCA-Mutated Ovarian Cancer. Clin Cancer Res. 2023;29(14):2725–34. [DOI] [PubMed] [Google Scholar]
- 47.Janysek DC, Kim J, Duijf PHG, Dray E. Clinical use and mechanisms of resistance for PARP inhibitors in homologous recombination-deficient cancers. Transl Oncol. 2021;14(3): 101012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Loehr A, Hussain A, Patnaik A, Bryce AH, Castellano D, Font A, et al. Emergence of BRCA Reversion Mutations in Patients with Metastatic Castration-resistant Prostate Cancer After Treatment with Rucaparib. Eur Urol. 2023;83(3):200–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Cai M, Song XL, Li XA, Chen M, Guo J, Yang DH, et al. Current therapy and drug resistance in metastatic castration-resistant prostate cancer. Drug Resist Updat. 2023;68: 100962. [DOI] [PubMed] [Google Scholar]
- 50.Harvey-Jones E, Raghunandan M, Robbez-Masson L, Magraner-Pardo L, Alaguthurai T, Yablonovitch A, et al. Longitudinal profiling identifies co-occurring BRCA1/2 reversions, TP53BP1, RIF1 and PAXIP1 mutations in PARP inhibitor-resistant advanced breast cancer. Ann Oncol. 2024;35(4):364–80. [DOI] [PubMed] [Google Scholar]
- 51.Goodall J, Mateo J, Yuan W, Mossop H, Porta N, Miranda S, et al. Circulating Cell-Free DNA to Guide Prostate Cancer Treatment with PARP Inhibition. Cancer Discov. 2017;7(9):1006–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Pettitt SJ, Frankum JR, Punta M, Lise S, Alexander J, Chen Y, et al. Clinical BRCA1/2 Reversion Analysis Identifies Hotspot Mutations and Predicted Neoantigens Associated with Therapy Resistance. Cancer Discov. 2020;10(10):1475–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Quigley D, Alumkal JJ, Wyatt AW, Kothari V, Foye A, Lloyd P, et al. Analysis of Circulating Cell-Free DNA Identifies Multiclonal Heterogeneity of BRCA2 Reversion Mutations Associated with Resistance to PARP Inhibitors. Cancer Discov. 2017;7(9):999–1005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Sakai W, Swisher EM, Karlan BY, Agarwal MK, Higgins J, Friedman C, et al. Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers. Nature. 2008;451(7182):1116–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Kondrashova O, Topp M, Nesic K, Lieschke E, Ho GY, Harrell MI, et al. Methylation of all BRCA1 copies predicts response to the PARP inhibitor rucaparib in ovarian carcinoma. Nat Commun. 2018;9(1):3970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Ter Brugge P, Kristel P, van der Burg E, Boon U, de Maaker M, Lips E, et al. Mechanisms of Therapy Resistance in Patient-Derived Xenograft Models of BRCA1-Deficient Breast Cancer. J Natl Cancer Inst. 2016;108(11). [DOI] [PubMed]
- 57.Nesic K, Kondrashova O, Hurley RM, McGehee CD, Vandenberg CJ, Ho GY, et al. Acquired RAD51C Promoter Methylation Loss Causes PARP Inhibitor Resistance in High-Grade Serous Ovarian Carcinoma. Cancer Res. 2021;81(18):4709–22. [DOI] [PubMed] [Google Scholar]
- 58.Baer MR, Kogan AA, Bentzen SM, Mi T, Lapidus RG, Duong VH, et al. Phase I Clinical Trial of DNA Methyltransferase Inhibitor Decitabine and PARP Inhibitor Talazoparib Combination Therapy in Relapsed/Refractory Acute Myeloid Leukemia. Clin Cancer Res. 2022;28(7):1313–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Xu G, Chapman JR, Brandsma I, Yuan J, Mistrik M, Bouwman P, et al. REV7 counteracts DNA double-strand break resection and affects PARP inhibition. Nature. 2015;521(7553):541–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Escribano-Díaz C, Orthwein A, Fradet-Turcotte A, Xing M, Young JT, Tkáč J, et al. A cell cycle-dependent regulatory circuit composed of 53BP1-RIF1 and BRCA1-CtIP controls DNA repair pathway choice. Mol Cell. 2013;49(5):872–83. [DOI] [PubMed] [Google Scholar]
- 61.Di Virgilio M, Callen E, Yamane A, Zhang W, Jankovic M, Gitlin AD, et al. Rif1 prevents resection of DNA breaks and promotes immunoglobulin class switching. Science. 2013;339(6120):711–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.McCormick A, Donoghue P, Dixon M, O’Sullivan R, O’Donnell RL, Murray J, et al. Ovarian Cancers Harbor Defects in Nonhomologous End Joining Resulting in Resistance to Rucaparib. Clin Cancer Res. 2017;23(8):2050–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Barazas M, Annunziato S, Pettitt SJ, de Krijger I, Ghezraoui H, Roobol SJ, et al. The CST Complex Mediates End Protection at Double-Strand Breaks and Promotes PARP Inhibitor Sensitivity in BRCA1-Deficient Cells. Cell Rep. 2018;23(7):2107–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Tkáč J, Xu G, Adhikary H, Young JTF, Gallo D, Escribano-Díaz C, et al. HELB Is a Feedback Inhibitor of DNA End Resection. Mol Cell. 2016;61(3):405–18. [DOI] [PubMed] [Google Scholar]
- 65.He YJ, Meghani K, Caron MC, Yang C, Ronato DA, Bian J, et al. DYNLL1 binds to MRE11 to limit DNA end resection in BRCA1-deficient cells. Nature. 2018;563(7732):522–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Francica P, Mutlu M, Blomen VA, Oliveira C, Nowicka Z, Trenner A, et al. Functional Radiogenetic Profiling Implicates ERCC6L2 in Non-homologous End Joining. Cell Rep. 2020;32(8): 108068. [DOI] [PubMed] [Google Scholar]
- 67.Clairmont CS, Sarangi P, Ponnienselvan K, Galli LD, Csete I, Moreau L, et al. TRIP13 regulates DNA repair pathway choice through REV7 conformational change. Nat Cell Biol. 2020;22(1):87–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Drané P, Brault ME, Cui G, Meghani K, Chaubey S, Detappe A, et al. TIRR regulates 53BP1 by masking its histone methyl-lysine binding function. Nature. 2017;543(7644):211–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Choi YE, Meghani K, Brault ME, Leclerc L, He YJ, Day TA, et al. Platinum and PARP Inhibitor Resistance Due to Overexpression of MicroRNA-622 in BRCA1-Mutant Ovarian Cancer. Cell Rep. 2016;14(3):429–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Dias MP, Moser SC, Ganesan S, Jonkers J. Understanding and overcoming resistance to PARP inhibitors in cancer therapy. Nat Rev Clin Oncol. 2021;18(12):773–91. [DOI] [PubMed] [Google Scholar]
- 71.Peng B, Shi R, Bian J, Li Y, Wang P, Wang H, et al. PARP1 and CHK1 coordinate PLK1 enzymatic activity during the DNA damage response to promote homologous recombination-mediated repair. Nucleic Acids Res. 2021;49(13):7554–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Du Y, Yamaguchi H, Wei Y, Hsu JL, Wang HL, Hsu YH, et al. Blocking c-Met-mediated PARP1 phosphorylation enhances anti-tumor effects of PARP inhibitors. Nat Med. 2016;22(2):194–201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Gogola E, Duarte AA, de Ruiter JR, Wiegant WW, Schmid JA, de Bruijn R, et al. Selective Loss of PARG Restores PARylation and Counteracts PARP Inhibitor-Mediated Synthetic Lethality. Cancer Cell. 2018;33(6):1078–93.e12. [DOI] [PubMed] [Google Scholar]
- 74.Gogola E, Duarte AA, de Ruiter JR, Wiegant WW, Schmid JA, de Bruijn R, et al. Selective Loss of PARG Restores PARylation and Counteracts PARP Inhibitor-Mediated Synthetic Lethality. Cancer Cell. 2019;35(6):950–2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Chen SH, Yu X. Targeting dePARylation selectively suppresses DNA repair-defective and PARP inhibitor-resistant malignancies. Sci Adv. 2019;5(4):eaav4340. [DOI] [PMC free article] [PubMed]
- 76.Christie EL, Pattnaik S, Beach J, Copeland A, Rashoo N, Fereday S, et al. Multiple ABCB1 transcriptional fusions in drug resistant high-grade serous ovarian and breast cancer. Nat Commun. 2019;10(1):1295. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Jaspers JE, Sol W, Kersbergen A, Schlicker A, Guyader C, Xu G, et al. BRCA2-deficient sarcomatoid mammary tumors exhibit multidrug resistance. Cancer Res. 2015;75(4):732–41. [DOI] [PubMed] [Google Scholar]
- 78.Rottenberg S, Jaspers JE, Kersbergen A, van der Burg E, Nygren AO, Zander SA, 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 U S A. 2008;105(44):17079–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Liu C, Li J, Xu F, Chen L, Ni M, Wu J, et al. PARP1-DOT1L transcription axis drives acquired resistance to PARP inhibitor in ovarian cancer. Mol Cancer. 2024;23(1):111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Teng QX, Lei ZN, Wang JQ, Yang Y, Wu ZX, Acharekar ND, et al. Overexpression of ABCC1 and ABCG2 confers resistance to talazoparib, a poly (ADP-Ribose) polymerase inhibitor. Drug Resist Updat. 2024;73: 101028. [DOI] [PubMed] [Google Scholar]
- 81.Dong XD, Lu Q, Li YD, Cai CY, Teng QX, Lei ZN, et al. RN486, a Bruton’s Tyrosine Kinase inhibitor, antagonizes multidrug resistance in ABCG2-overexpressing cancer cells. J Transl Int Med. 2024;12(3):288–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Ying S, Hamdy FC, Helleday T. Mre11-dependent degradation of stalled DNA replication forks is prevented by BRCA2 and PARP1. Cancer Res. 2012;72(11):2814–21. [DOI] [PubMed] [Google Scholar]
- 83.Schlacher K, Wu H, Jasin M. A distinct replication fork protection pathway connects Fanconi anemia tumor suppressors to RAD51-BRCA1/2. Cancer Cell. 2012;22(1):106–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Schlacher K, Christ N, Siaud N, Egashira A, Wu H, Jasin M. Double-strand break repair-independent role for BRCA2 in blocking stalled replication fork degradation by MRE11. Cell. 2011;145(4):529–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Ray Chaudhuri A, Callen E, Ding X, Gogola E, Duarte AA, Lee JE, et al. Replication fork stability confers chemoresistance in BRCA-deficient cells. Nature. 2016;535(7612):382–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Taglialatela A, Alvarez S, Leuzzi G, Sannino V, Ranjha L, Huang JW, et al. Restoration of Replication Fork Stability in BRCA1- and BRCA2-Deficient Cells by Inactivation of SNF2-Family Fork Remodelers. Mol Cell. 2017;68(2):414–30.e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Bryant HE, Petermann E, Schultz N, Jemth AS, Loseva O, Issaeva N, et al. PARP is activated at stalled forks to mediate Mre11-dependent replication restart and recombination. Embo j. 2009;28(17):2601–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Guillemette S, Serra RW, Peng M, Hayes JA, Konstantinopoulos PA, Green MR, et al. Resistance to therapy in BRCA2 mutant cells due to loss of the nucleosome remodeling factor CHD4. Genes Dev. 2015;29(5):489–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Rondinelli B, Gogola E, Yücel H, Duarte AA, van de Ven M, van der Sluijs R, et al. EZH2 promotes degradation of stalled replication forks by recruiting MUS81 through histone H3 trimethylation. Nat Cell Biol. 2017;19(11):1371–8. [DOI] [PubMed] [Google Scholar]
- 90.Dungrawala H, Bhat KP, Le Meur R, Chazin WJ, Ding X, Sharan SK, et al. RADX Promotes Genome Stability and Modulates Chemosensitivity by Regulating RAD51 at Replication Forks. Mol Cell. 2017;67(3):374–86.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Mao X, Xu S, Wang H, Xiao P, Li S, Wu J, et al. Integrated analysis reveals critical cisplatin-resistance regulators E2F7 contributed to tumor progression and metastasis in lung adenocarcinoma. Cancer Cell Int. 2024;24(1):173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Clements KE, Thakar T, Nicolae CM, Liang X, Wang HG, Moldovan GL. Loss of E2F7 confers resistance to poly-ADP-ribose polymerase (PARP) inhibitors in BRCA2-deficient cells. Nucleic Acids Res. 2018;46(17):8898–907. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Zhang B, Ramkumar K, Cardnell RJ, Gay CM, Stewart CA, Wang WL, et al. A wake-up call for cancer DNA damage: the role of Schlafen 11 (SLFN11) across multiple cancers. Br J Cancer. 2021;125(10):1333–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Murai J, Tang SW, Leo E, Baechler SA, Redon CE, Zhang H, et al. SLFN11 Blocks Stressed Replication Forks Independently of ATR. Mol Cell. 2018;69(3):371–84.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Lok BH, Gardner EE, Schneeberger VE, Ni A, Desmeules P, Rekhtman N, et al. PARP Inhibitor Activity Correlates with SLFN11 Expression and Demonstrates Synergy with Temozolomide in Small Cell Lung Cancer. Clin Cancer Res. 2017;23(2):523–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Fendt SM. 100 years of the Warburg effect: A cancer metabolism endeavor. Cell. 2024;187(15):3824–8. [DOI] [PubMed] [Google Scholar]
- 97.Sun P, Ma L, Lu Z. Lactylation: Linking the Warburg effect to DNA damage repair. Cell Metab. 2024;36(8):1637–9. [DOI] [PubMed] [Google Scholar]
- 98.Navas LE, Carnero A. NAD(+) metabolism, stemness, the immune response, and cancer. Signal Transduct Target Ther. 2021;6(1):2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Houtkooper RH, Cantó C, Wanders RJ, Auwerx J. The secret life of NAD+: an old metabolite controlling new metabolic signaling pathways. Endocr Rev. 2010;31(2):194–223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Wu J, Liu N, Chen J, Tao Q, Li Q, Li J, et al. The Tricarboxylic Acid Cycle Metabolites for Cancer: Friend or Enemy. Research (Wash D C). 2024;7:0351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Shi X, Zhang W, Gu C, Ren H, Wang C, Yin N, et al. NAD+ depletion radiosensitizes 2-DG-treated glioma cells by abolishing metabolic adaptation. Free Radic Biol Med. 2021;162:514–22. [DOI] [PubMed] [Google Scholar]
- 102.Sheikh E, Agrawal K, Roy S, Burk D, Donnarumma F, Ko YH, et al. Multimodal Imaging of Pancreatic Cancer Microenvironment in Response to an Antiglycolytic Drug. Adv Healthc Mater. 2023;12(31): e2301815. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Abdel-Wahab AF, Mahmoud W, Al-Harizy RM. Targeting glucose metabolism to suppress cancer progression: prospective of anti-glycolytic cancer therapy. Pharmacol Res. 2019;150: 104511. [DOI] [PubMed] [Google Scholar]
- 104.Bajrami I, Kigozi A, Van Weverwijk A, Brough R, Frankum J, Lord CJ, et al. Synthetic lethality of PARP and NAMPT inhibition in triple-negative breast cancer cells. EMBO Mol Med. 2012;4(10):1087–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Gibson AE, Yeung C, Issaq SH, Collins VJ, Gouzoulis M, Zhang Y, et al. Inhibition of nicotinamide phosphoribosyltransferase (NAMPT) with OT-82 induces DNA damage, cell death, and suppression of tumor growth in preclinical models of Ewing sarcoma. Oncogenesis. 2020;9(9):80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Chan M, Gravel M, Bramoullé A, Bridon G, Avizonis D, Shore GC, et al. Synergy between the NAMPT inhibitor GMX1777(8) and pemetrexed in non-small cell lung cancer cells is mediated by PARP activation and enhanced NAD consumption. Cancer Res. 2014;74(21):5948–54. [DOI] [PubMed] [Google Scholar]
- 107.Liu Y, Sun Y, Guo Y, Shi X, Chen X, Feng W, et al. An Overview: The Diversified Role of Mitochondria in Cancer Metabolism. Int J Biol Sci. 2023;19(3):897–915. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Evans KW, Yuca E, Scott SS, Zhao M, Paez Arango N, Cruz Pico CX, et al. Oxidative Phosphorylation Is a Metabolic Vulnerability in Chemotherapy-Resistant Triple-Negative Breast Cancer. Cancer Res. 2021;81(21):5572–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Drugging OXPHOS Dependency in Cancer. Cancer Discov. 2019;9(8):Of10. [DOI] [PubMed]
- 110.Winter JM, Yadav T, Rutter J. Stressed to death: Mitochondrial stress responses connect respiration and apoptosis in cancer. Mol Cell. 2022;82(18):3321–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Röhrig F, Schulze A. The multifaceted roles of fatty acid synthesis in cancer. Nat Rev Cancer. 2016;16(11):732–49. [DOI] [PubMed] [Google Scholar]
- 112.Yang S, Hwang S, Kim B, Shin S, Kim M, Jeong SM. Fatty acid oxidation facilitates DNA double-strand break repair by promoting PARP1 acetylation. Cell Death Dis. 2023;14(7):435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Hwang S, Yang S, Park K, Kim B, Kim M, Shin S, et al. Induction of Fatty Acid Oxidation Underlies DNA Damage-Induced Cell Death and Ameliorates Obesity-Driven Chemoresistance. Adv Sci (Weinh). 2024;11(10): e2304702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Tateishi K, Wakimoto H, Iafrate AJ, Tanaka S, Loebel F, Lelic N, et al. Extreme Vulnerability of IDH1 Mutant Cancers to NAD+ Depletion. Cancer Cell. 2015;28(6):773–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Sulkowski PL, Oeck S, Dow J, Economos NG, Mirfakhraie L, Liu Y, et al. Oncometabolites suppress DNA repair by disrupting local chromatin signalling. Nature. 2020;582(7813):586–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Lu Y, Kwintkiewicz J, Liu Y, Tech K, Frady LN, Su YT, et al. Chemosensitivity of IDH1-Mutated Gliomas Due to an Impairment in PARP1-Mediated DNA Repair. Cancer Res. 2017;77(7):1709–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Liu W, Cao H, Wang J, Elmusrati A, Han B, Chen W, et al. Histone-methyltransferase KMT2D deficiency impairs the Fanconi anemia/BRCA pathway upon glycolytic inhibition in squamous cell carcinoma. Nat Commun. 2024;15(1):6755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Lahiguera Á, Hyroššová P, Figueras A, Garzón D, Moreno R, Soto-Cerrato V, et al. Tumors defective in homologous recombination rely on oxidative metabolism: relevance to treatments with PARP inhibitors. EMBO Mol Med. 2020;12(6): e11217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Peraza-Vega RI, Valverde M, Rojas E. Interactions between miRNAs and Double-Strand Breaks DNA Repair Genes, Pursuing a Fine-Tuning of Repair. Int J Mol Sci. 2022;23(6). [DOI] [PMC free article] [PubMed]
- 120.Liu K, Huang J, Xie M, Yu Y, Zhu S, Kang R, et al. MIR34A regulates autophagy and apoptosis by targeting HMGB1 in the retinoblastoma cell. Autophagy. 2014;10(3):442–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Bustos MA, Yokoe T, Shoji Y, Kobayashi Y, Mizuno S, Murakami T, et al. MiR-181a targets STING to drive PARP inhibitor resistance in BRCA- mutated triple-negative breast cancer and ovarian cancer. Cell Biosci. 2023;13(1):200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Raju GSR, Pavitra E, Bandaru SS, Varaprasad GL, Nagaraju GP, Malla RR, et al. HOTAIR: a potential metastatic, drug-resistant and prognostic regulator of breast cancer. Mol Cancer. 2023;22(1):65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Yong H, Wu G, Chen J, Liu X, Bai Y, Tang N, et al. lncRNA MALAT1 Accelerates Skeletal Muscle Cell Apoptosis and Inflammatory Response in Sepsis by Decreasing BRCA1 Expression by Recruiting EZH2. Mol Ther Nucleic Acids. 2020;19:97–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Sang Y, Tang J, Li S, Li L, Tang X, Cheng C, et al. LncRNA PANDAR regulates the G1/S transition of breast cancer cells by suppressing p16(INK4A) expression. Sci Rep. 2016;6:22366. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Murphy MR, Ramadei A, Doymaz A, Varriano S, Natelson DM, Yu A, et al. Long non-coding RNA generated from CDKN1A gene by alternative polyadenylation regulates p21 expression during DNA damage response. Nucleic Acids Res. 2023;51(21):11911–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Chen X, Xiong D, Ye L, Wang K, Huang L, Mei S, et al. Up-regulated lncRNA XIST contributes to progression of cervical cancer via regulating miR-140-5p and ORC1. Cancer Cell Int. 2019;19:45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Wu H, Sun C, Cao W, Teng Q, Ma X, Schioth HB, et al. Blockade of the lncRNA-PART1-PHB2 axis confers resistance to PARP inhibitor and promotes cellular senescence in ovarian cancer. Cancer Lett. 2024;602: 217192. [DOI] [PubMed] [Google Scholar]
- 128.Xie J, Ye F, Deng X, Tang Y, Liang JY, Huang X, et al. Circular RNA: A promising new star of vaccine. J Transl Int Med. 2023;11(4):372–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Kristensen LS, Andersen MS, Stagsted LVW, Ebbesen KK, Hansen TB, Kjems J. The biogenesis, biology and characterization of circular RNAs. Nat Rev Genet. 2019;20(11):675–91. [DOI] [PubMed] [Google Scholar]
- 130.Zou Y, Yang A, Chen B, Deng X, Xie J, Dai D, et al. crVDAC3 alleviates ferroptosis by impeding HSPB1 ubiquitination and confers trastuzumab deruxtecan resistance in HER2-low breast cancer. Drug Resist Updat. 2024;77: 101126. [DOI] [PubMed] [Google Scholar]
- 131.Li B, Chen J, Wu Y, Luo H, Ke Y. Decrease of circARID1A retards glioblastoma invasion by modulating miR-370-3p/ TGFBR2 pathway. Int J Biol Sci. 2022;18(13):5123–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Gao Z, Luan X, Wang X, Han T, Li X, Li Z, et al. DNA damage response-related ncRNAs as regulators of therapy resistance in cancer. Front Pharmacol. 2024;15:1390300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Nicot C. circRNAs shed light on cancer diagnosis and treatment. Mol Cancer. 2022;21(1):107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Velaei K, Samadi N, Barazvan B, Soleimani RJ. Tumor microenvironment-mediated chemoresistance in breast cancer. Breast. 2016;30:92–100. [DOI] [PubMed] [Google Scholar]
- 135.Lu Y, Sun Q, Guan Q, Zhang Z, He Q, He J, et al. The XOR-IDH3alpha axis controls macrophage polarization in hepatocellular carcinoma. J Hepatol. 2023;79(5):1172–84. [DOI] [PubMed] [Google Scholar]
- 136.Ding L, Wang Q, Martincuks A, Kearns MJ, Jiang T, Lin Z, et al. STING agonism overcomes STAT3-mediated immunosuppression and adaptive resistance to PARP inhibition in ovarian cancer. J Immunother Cancer. 2023;11(1). [DOI] [PMC free article] [PubMed]
- 137.Mehta AK, Cheney EM, Hartl CA, Pantelidou C, Oliwa M, Castrillon JA, et al. Targeting immunosuppressive macrophages overcomes PARP inhibitor resistance in BRCA1-associated triple-negative breast cancer. Nat Cancer. 2021;2(1):66–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Pilié PG, Gay CM, Byers LA, O’Connor MJ, Yap TA. PARP Inhibitors: Extending Benefit Beyond BRCA-Mutant Cancers. Clin Cancer Res. 2019;25(13):3759–71. [DOI] [PubMed] [Google Scholar]
- 139.Forde PM, Chaft JE, Smith KN, Anagnostou V, Cottrell TR, Hellmann MD, et al. Neoadjuvant PD-1 Blockade in Resectable Lung Cancer. N Engl J Med. 2018;378(21):1976–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Long LL, Ma SC, Guo ZQ, Zhang YP, Fan Z, Liu LJ, et al. PARP Inhibition Induces Synthetic Lethality and Adaptive Immunity in LKB1-Mutant Lung Cancer. Cancer Res. 2023;83(4):568–81. [DOI] [PubMed] [Google Scholar]
- 141.Pantelidou C, Jadhav H, Kothari A, Liu R, Wulf GM, Guerriero JL, et al. STING agonism enhances anti-tumor immune responses and therapeutic efficacy of PARP inhibition in BRCA-associated breast cancer. NPJ Breast Cancer. 2022;8(1):102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Khoo LT, Chen LY. Role of the cGAS-STING pathway in cancer development and oncotherapeutic approaches. EMBO Rep. 2018;19(12). [DOI] [PMC free article] [PubMed]
- 143.Wang Q, Bergholz JS, Ding L, Lin Z, Kabraji SK, Hughes ME, et al. STING agonism reprograms tumor-associated macrophages and overcomes resistance to PARP inhibition in BRCA1-deficient models of breast cancer. Nat Commun. 2022;13(1):3022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Shen J, Zhao W, Ju Z, Wang L, Peng Y, Labrie M, et al. PARPi Triggers the STING-Dependent Immune Response and Enhances the Therapeutic Efficacy of Immune Checkpoint Blockade Independent of BRCAness. Cancer Res. 2019;79(2):311–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Zhou Z, Huang S, Fan F, Xu Y, Moore C, Li S, et al. The multiple faces of cGAS-STING in antitumor immunity: prospects and challenges. 2024;4(3):173–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Zitvogel L, Galluzzi L, Kepp O, Smyth MJ, Kroemer G. Type I interferons in anticancer immunity. Nat Rev Immunol. 2015;15(7):405–14. [DOI] [PubMed] [Google Scholar]
- 147.Lu X, Li X, Li L, Han C, Li S. Advances in the prerequisite and consequence of STING downstream signalosomes. 2024;4(5):435–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Jiao S, Xia W, Yamaguchi H, Wei Y, Chen MK, Hsu JM, et al. PARP Inhibitor Upregulates PD-L1 Expression and Enhances Cancer-Associated Immunosuppression. Clin Cancer Res. 2017;23(14):3711–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Stewart RA, Pilié PG, Yap TA. Development of PARP and Immune-Checkpoint Inhibitor Combinations. Cancer Res. 2018;78(24):6717–25. [DOI] [PubMed] [Google Scholar]
- 150.Pantelidou C, Sonzogni O, De Oliveria TM, Mehta AK, Kothari A, Wang D, et al. PARP Inhibitor Efficacy Depends on CD8(+) T-cell Recruitment via Intratumoral STING Pathway Activation in BRCA-Deficient Models of Triple-Negative Breast Cancer. Cancer Discov. 2019;9(6):722–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Meng J, Peng J, Feng J, Maurer J, Li X, Li Y, et al. Niraparib exhibits a synergistic anti-tumor effect with PD-L1 blockade by inducing an immune response in ovarian cancer. J Transl Med. 2021;19(1):415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Bertoli C, Klier S, McGowan C, Wittenberg C, de Bruin RAM. Chk1 inhibits E2F6 repressor function in response to replication stress to maintain cell-cycle transcription. Current biology: CB. 2013;23(17):1629–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Slade D. PARP and PARG inhibitors in cancer treatment. Genes Dev. 2020;34(5–6):360–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Gupta N, Huang T-T, Horibata S, Lee J-M. Cell cycle checkpoints and beyond: Exploiting the ATR/CHK1/WEE1 pathway for the treatment of PARP inhibitor-resistant cancer. Pharmacol Res. 2022;178: 106162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Liu F, Gao A, Zhang M, Li Y, Zhang F, Herman JG, et al. Methylation of FAM110C is a synthetic lethal marker for ATR/CHK1 inhibitors in pancreatic cancer. J Transl Int Med. 2024;12(3):274–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Lok BH, Gardner EE, Schneeberger VE, Ni A, Desmeules P, Rekhtman N, et al. PARP inhibitor activity correlates with SLFN11 expression and demonstrates synergy with temozolomide in small cell lung cancer. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2017;23(2):523–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Murai J, Feng Y, Yu GK, Ru Y, Tang S-W, Shen Y, et al. Resistance to PARP inhibitors by SLFN11 inactivation can be overcome by ATR inhibition. Oncotarget. 2016;7(47):76534–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Xu H, Gitto SB, Ho G-Y, Medvedev S, Shield-Artin K, Kim H, et al. CHK1 inhibitor SRA737 is active in PARP inhibitor resistant and CCNE1 amplified ovarian cancer. iScience. 2024;27(7):109978. [DOI] [PMC free article] [PubMed]
- 159.Song X, Fang C, Dai Y, Sun Y, Qiu C, Lin X, et al. Cyclin-dependent kinase 7 (CDK7) inhibitors as a novel therapeutic strategy for different molecular types of breast cancer. Br J Cancer. 2024;130(8):1239–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Pilié PG, Tang C, Mills GB, Yap TA. State-of-the-art strategies for targeting the DNA damage response in cancer. Nat Rev Clin Oncol. 2019;16(2):81–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Serra V, Wang AT, Castroviejo-Bermejo M, Polanska UM, Palafox M, Herencia-Ropero A, et al. Identification of a Molecularly-Defined Subset of Breast and Ovarian Cancer Models that Respond to WEE1 or ATR Inhibition, Overcoming PARP Inhibitor Resistance. Clin Cancer Res. 2022;28(20):4536–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Teo ZL, O’Connor MJ, Versaci S, Clarke KA, Brown ER, Percy LW, et al. Combined PARP and WEE1 inhibition triggers anti-tumor immune response in BRCA1/2 wildtype triple-negative breast cancer. NPJ Breast Cancer. 2023;9(1):68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Klein FG, Granier C, Zhao Y, Pan Q, Tong Z, Gschwend JE, et al. Combination of talazoparib and palbociclib as a potent treatment strategy in bladder cancer. Journal of Personalized Medicine. 2021;11(5):340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Orhan E, Velazquez C, Tabet I, Fenou L, Rodier G, Orsetti B, et al. CDK inhibition results in pharmacologic BRCAness increasing sensitivity to olaparib in BRCA1-WT and olaparib resistant in Triple Negative Breast Cancer. Cancer Lett. 2024;589: 216820. [DOI] [PubMed] [Google Scholar]
- 165.Carey JPW, Karakas C, Bui T, Chen X, Vijayaraghavan S, Zhao Y, et al. Synthetic lethality of PARP inhibitors in combination with MYC blockade is independent of BRCA status in triple-negative breast cancer. Can Res. 2018;78(3):742–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Li K-M, Deng L-G, Xue L-J, Tan C, Yao S-K. AXL inhibition prevents RPA2/CHK1-mediated homologous recombination to increase PARP inhibitor sensitivity in hepatocellular carcinoma. Heliyon. 2024;10(17): e36283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Castroviejo-Bermejo M, Cruz C, Llop-Guevara A, Gutiérrez-Enríquez S, Ducy M, Ibrahim YH, et al. A RAD51 assay feasible in routine tumor samples calls PARP inhibitor response beyond BRCA mutation. EMBO Mol Med. 2018;10(12): e9172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Cruz C, Castroviejo-Bermejo M, Gutiérrez-Enríquez S, Llop-Guevara A, Ibrahim YH, Gris-Oliver A, et al. RAD51 foci as a functional biomarker of homologous recombination repair and PARP inhibitor resistance in germline BRCA-mutated breast cancer. Annals of Oncology: Official Journal of the European Society for Medical Oncology. 2018;29(5):1203–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Waks AG, Cohen O, Kochupurakkal B, Kim D, Dunn CE, Buendia Buendia J, et al. Reversion and non-reversion mechanisms of resistance to PARP inhibitor or platinum chemotherapy in BRCA1/2-mutant metastatic breast cancer. Annals of Oncology: Official Journal of the European Society for Medical Oncology. 2020;31(5):590–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Min A, Im S-A, Yoon Y-K, Song S-H, Nam H-J, Hur H-S, et al. RAD51C-deficient cancer cells are highly sensitive to the PARP inhibitor olaparib. Mol Cancer Ther. 2013;12(6):865–77. [DOI] [PubMed] [Google Scholar]
- 171.Nesic K, Kondrashova O, Hurley RM, McGehee CD, Vandenberg CJ, Ho G-Y, et al. Acquired RAD51C promoter methylation loss causes PARP inhibitor resistance in high-grade serous ovarian carcinoma. Can Res. 2021;81(18):4709–22. [DOI] [PubMed] [Google Scholar]
- 172.Patterson-Fortin J, D’Andrea AD. Exploiting the Microhomology-Mediated End-Joining Pathway in Cancer Therapy. Cancer Res. 2020;80(21):4593–600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Lavudi K, Banerjee A, Li N, Yang Y, Cai S, Bai X, et al. ALDH1A1 promotes PARP inhibitor resistance by enhancing retinoic acid receptor-mediated DNA polymerase theta expression. NPJ Precis Oncol. 2023;7(1):66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Zatreanu D, Robinson HMR, Alkhatib O, Boursier M, Finch H, Geo L, et al. Polθ inhibitors elicit BRCA-gene synthetic lethality and target PARP inhibitor resistance. Nat Commun. 2021;12(1):3636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Subhash VV, Tan SH, Yeo MS, Yan FL, Peethala PC, Liem N, et al. ATM expression predicts veliparib and irinotecan sensitivity in gastric cancer by mediating P53-independent regulation of cell cycle and apoptosis. Mol Cancer Ther. 2016;15(12):3087–96. [DOI] [PubMed] [Google Scholar]
- 176.Chu Y-Y, Chen M-K, Wei Y, Lee H-H, Xia W, Wang Y-N, et al. Targeting the ALK-CDK9-Tyr19 kinase cascade sensitizes ovarian and breast tumors to PARP inhibition via destabilization of the P-TEFb complex. Nature Cancer. 2022;3(10):1211–27. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Swisher EM, Kwan TT, Oza AM, Tinker AV, Ray-Coquard I, Oaknin A, et al. Molecular and clinical determinants of response and resistance to rucaparib for recurrent ovarian cancer treatment in ARIEL2 (parts 1 and 2). Nat Commun. 2021;12(1):2487. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Ikeuchi H, Matsuno Y, Kusumoto-Matsuo R, Kojima S, Ueno T, Ikegami M, et al. GLI1 confers resistance to PARP inhibitors by activating the DNA damage repair pathway. Oncogene. 2024;43(41):3037–48. [DOI] [PubMed] [Google Scholar]
- 179.Paes Dias M, Tripathi V, van der Heijden I, Cong K, Manolika E-M, Bhin J, et al. Loss of nuclear DNA ligase III reverts PARP inhibitor resistance in BRCA1/53BP1 double-deficient cells by exposing ssDNA gaps. Mol Cell. 2021;81(22):4692–708.e9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Kalev P, Simicek M, Vazquez I, Munck S, Chen L, Soin T, et al. Loss of PPP2R2A inhibits homologous recombination DNA repair and predicts tumor sensitivity to PARP inhibition. Can Res. 2012;72(24):6414–24. [DOI] [PubMed] [Google Scholar]
- 181.Mesquita KA, Alabdullah M, Griffin M, Toss MS, Fatah TMAA, Alblihy A, et al. ERCC1-XPF deficiency is a predictor of olaparib induced synthetic lethality and platinum sensitivity in epithelial ovarian cancers. Gynecol Oncol. 2019;153(2):416–24. [DOI] [PubMed] [Google Scholar]
- 182.Jackson LM, Dhoonmoon A, Hale A, Dennis KA, Schleicher EM, Nicolae CM, et al. Loss of MED12 activates the TGFβ pathway to promote chemoresistance and replication fork stability in BRCA-deficient cells. Nucleic Acids Res. 2021;49(22):12855–69. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Fukumoto T, Zhu H, Nacarelli T, Karakashev S, Fatkhutdinov N, Wu S, et al. N6-methylation of adenosine of FZD10 mRNA contributes to PARP inhibitor resistance. Can Res. 2019;79(11):2812–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 184.Cai Q, Deng W, Zou Y, Chen Z-S, Tang H. Histone lactylation as a driver of metabolic reprogramming and immune evasion. 2024.
- 185.Italiano A, Soria J-C, Toulmonde M, Michot J-M, Lucchesi C, Varga A, et al. Tazemetostat, an EZH2 inhibitor, in relapsed or refractory B-cell non-hodgkin lymphoma and advanced solid tumours: A first-in-human, open-label, phase 1 study. Lancet Oncol. 2018;19(5):649–59. [DOI] [PubMed] [Google Scholar]
- 186.Raynal NJM, Da Costa EM, Lee JT, Gharibyan V, Ahmed S, Zhang H, et al. Repositioning FDA-approved drugs in combination with epigenetic drugs to reprogram colon cancer epigenome. Mol Cancer Ther. 2017;16(2):397–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Sarnik J, Popławski T, Tokarz P. BET proteins as attractive targets for cancer therapeutics. Int J Mol Sci. 2021;22(20):11102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Mita MM, Mita AC. Bromodomain inhibitors a decade later: A promise unfulfilled? Br J Cancer. 2020;123(12):1713–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Pulliam N, Fang F, Ozes AR, Tang J, Adewuyi A, Keer H, et al. An effective epigenetic-PARP inhibitor combination therapy for breast and ovarian cancers independent of BRCA mutations. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2018;24(13):3163–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Maifrede S, Le BV, Nieborowska-Skorska M, Golovine K, Sullivan-Reed K, Dunuwille WMB, et al. TET2 and DNMT3A mutations exert divergent effects on DNA repair and sensitivity of leukemia cells to PARP inhibitors. Can Res. 2021;81(19):5089–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Minucci S, Pelicci PG. Histone deacetylase inhibitors and the promise of epigenetic (and more) treatments for cancer. Nat Rev Cancer. 2006;6(1):38–51. [DOI] [PubMed] [Google Scholar]
- 192.Min A, Im S-A, Kim DK, Song S-H, Kim H-J, Lee K-H, et al. Histone deacetylase inhibitor, suberoylanilide hydroxamic acid (SAHA), enhances anti-tumor effects of the poly (ADP-ribose) polymerase (PARP) inhibitor olaparib in triple-negative breast cancer cells. Breast cancer research: BCR. 2015;17:33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Valdez BC, Tsimberidou AM, Yuan B, Baysal MA, Chakraborty A, Andersen CR, et al. Synergistic cytotoxicity of histone deacetylase and poly-ADP ribose polymerase inhibitors and decitabine in breast and ovarian cancer cells: Implications for novel therapeutic combinations. Int J Mol Sci. 2024;25(17):9241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Wang H-J, Ran H-F, Yin Y, Xu X-G, Jiang B-X, Yu S-Q, et al. Catalpol improves impaired neurovascular unit in ischemic stroke rats via enhancing VEGF-PI3K/AKT and VEGF-MEK1/2/ERK1/2 signaling. Acta Pharmacol Sin. 2022;43(7):1670–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Mateo J, Lord CJ, Serra V, Tutt A, Balmaña J, Castroviejo-Bermejo M, et al. A decade of clinical development of PARP inhibitors in perspective. Annals of Oncology: Official Journal of the European Society for Medical Oncology. 2019;30(9):1437–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Lim JJ, Yang K, Taylor-Harding B, Wiedemeyer WR, Buckanovich RJ. VEGFR3 inhibition chemosensitizes ovarian cancer stemlike cells through down-regulation of BRCA1 and BRCA2. Neoplasia. 2014;16(4):343–53 e1–2. [DOI] [PMC free article] [PubMed]
- 197.Iida Y, Yanaihara N, Yoshino Y, Saito M, Saito R, Tabata J, et al. Bevacizumab increases the sensitivity of olaparib to homologous recombination-proficient ovarian cancer by suppressing CRY1 via PI3K/AKT pathway. Front Oncol. 2024;14:1302850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Zhang H, Chen P, Tian Z, Tang H, Guan Z, Zou Y. SF3B4-mediated alternative splicing in cancer development and progression. Cancer Letters. 2025:217597. [DOI] [PubMed]
- 199.McKay RR, Morgans AK, Shore ND, Dunshee C, Devgan G, Agarwal N. First-line combination treatment with PARP and androgen receptor-signaling inhibitors in HRR-deficient mCRPC: Applying clinical study findings to clinical practice in the United States. Cancer Treat Rev. 2024;126: 102726. [DOI] [PubMed] [Google Scholar]
- 200.Slootbeek PHJ, Overbeek JK, Ligtenberg MJL, van Erp NP, Mehra N. PARPing up the right tree; an overview of PARP inhibitors for metastatic castration-resistant prostate cancer. Cancer Lett. 2023;577: 216367. [DOI] [PubMed] [Google Scholar]
- 201.Agarwal N, Zhang T, Efstathiou E, Sayegh N, Engelsberg A, Saad F, et al. The biology behind combining poly [ADP ribose] polymerase and androgen receptor inhibition for metastatic castration-resistant prostate cancer. Eur J Cancer. 2023;192: 113249. [DOI] [PubMed] [Google Scholar]
- 202.Lee J-M, Moore RG, Ghamande S, Park MS, Diaz JP, Chapman J, et al. Cediranib in combination with olaparib in patients without a germline BRCA1/2 mutation and with recurrent platinum-resistant ovarian cancer: Phase IIb CONCERTO trial. Clinical Cancer Research: An Official Journal of the American Association for Cancer Research. 2022;28(19):4186–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Yap TA, Kristeleit R, Michalarea V, Pettitt SJ, Lim JSJ, Carreira S, et al. Phase I trial of the PARP inhibitor olaparib and AKT inhibitor capivasertib in patients with BRCA1/2- and non-BRCA1/2-mutant cancers. Cancer Discov. 2020;10(10):1528–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Konstantinopoulos PA, Barry WT, Birrer M, Westin SN, Cadoo KA, Shapiro GI, et al. Olaparib and α-specific PI3K inhibitor alpelisib for patients with epithelial ovarian cancer: A dose-escalation and dose-expansion phase 1b trial. Lancet Oncol. 2019;20(4):570–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Liu Q, Gheorghiu L, Drumm M, Clayman R, Eidelman A, Wszolek MF, et al. PARP-1 inhibition with or without ionizing radiation confers reactive oxygen species-mediated cytotoxicity preferentially to cancer cells with mutant TP53. Oncogene. 2018;37(21):2793–805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Dardare J, Witz A, Betz M, François A, Lamy L, Husson M, et al. DDB2 expression lights the way for precision radiotherapy response in PDAC cells, with or without olaparib. Cell Death Discovery. 2024;10(1):411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Soni A, Lin X, Mladenov E, Mladenova V, Stuschke M, Iliakis G. BMN673 is a PARP inhibitor with unique radiosensitizing properties: Mechanisms and potential in radiation therapy. Cancers. 2022;14(22):5619. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Ye F, Dewanjee S, Li Y, Jha NK, Chen ZS, Kumar A, et al. Advancements in clinical aspects of targeted therapy and immunotherapy in breast cancer. Mol Cancer. 2023;22(1):105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Zou Y, Zhang H, Liu F, Chen ZS, Tang H. Intratumoral microbiota in orchestrating cancer immunotherapy response. J Transl Int Med. 2024;12(6):540–2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Li Y, Cen Y, Tu M, Xiang Z, Tang S, Lu W, et al. Nanoengineered Gallium Ion Incorporated Formulation for Safe and Efficient Reversal of PARP Inhibition and Platinum Resistance in Ovarian Cancer. Research (Wash D C). 2023;6:0070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Xie J, Liu M, Deng X, Tang Y, Zheng S, Ou X, et al. Gut microbiota reshapes cancer immunotherapy efficacy: Mechanisms and therapeutic strategies. Imeta. 2024;3(1): e156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Yan Y, Zhang Y, Liu J, Chen B, Wang Y. Emerging magic bullet: subcellular organelle-targeted cancer therapy. 2024. [DOI] [PMC free article] [PubMed]
- 213.Luo T, Fan Z, Zeng A, Wang A, Pan Y, Xu Y, et al. Biomimetic Targeted Co-Delivery System Engineered from Genomic Insights for Precision Treatment of Osteosarcoma. Adv Sci (Weinh). 2025;12(2): e2410427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Guney Eskiler G, Cecener G, Egeli U, Tunca B. Synthetically Lethal BMN 673 (Talazoparib) Loaded Solid Lipid Nanoparticles for BRCA1 Mutant Triple Negative Breast Cancer. Pharm Res. 2018;35(11):218. [DOI] [PubMed] [Google Scholar]
- 215.Zheng S, Li M, Xu W, Zhang J, Li G, Xiao H, et al. Dual-targeted nanoparticulate drug delivery systems for enhancing triple-negative breast cancer treatment. J Control Release. 2024;371:371–85. [DOI] [PubMed] [Google Scholar]
- 216.Khang M, Lee JH, Lee T, Suh HW, Lee S, Cavaliere A, et al. Intrathecal delivery of nanoparticle PARP inhibitor to the cerebrospinal fluid for the treatment of metastatic medulloblastoma. Sci Transl Med. 2023;15(720):eadi1617. [DOI] [PMC free article] [PubMed]
- 217.Alradwan I, Zhi P, Zhang T, Lip H, Zetrini A, He C, et al. Nanoparticulate drug combination inhibits DNA damage repair and PD-L1 expression in BRCA-mutant and wild type triple-negative breast cancer. J Control Release. 2025;377:661–74. [DOI] [PubMed] [Google Scholar]
- 218.Luo Z, Wan Z, Ren P, Zhang B, Huang Y, West RE 3rd, et al. In Situ Formation of Fibronectin-Enriched Protein Corona on Epigenetic Nanocarrier for Enhanced Synthetic Lethal Therapy. Adv Sci (Weinh). 2024;11(19): e2307940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Mao X, Xu J, Wang W, Liang C, Hua J, Liu J, et al. Crosstalk between cancer-associated fibroblasts and immune cells in the tumor microenvironment: new findings and future perspectives. Mol Cancer. 2021;20(1):131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Deng X, Yan Y, Zhan Z, Xie J, Tang H, Zou Y, et al. A glimpse into the future: Integrating artificial intelligence for precision HER2‐positive breast cancer management. iMetaOmics. 2024;1(1).
- 221.Lu L, Gou X, Tan SK, Mann SI, Yang H, Zhong X, et al. De novo design of drug-binding proteins with predictable binding energy and specificity. Science. 2024;384(6691):106–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 222.Senior M. Beyond the hype of AI in cancer R&D. Nat Cancer. 2024;5(12):1759–61. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
No datasets were generated or analysed during the current study.