Skip to main content

Application of PARP inhibitors combined with immune checkpoint inhibitors in ovarian cancer

Abstract

The advent of polyadenosine diphosphate ribose polymerase inhibitors (PARPi) has brought about significant changes in the field of ovarian cancer treatment. However, in 2022, Rucaparib, Olaparib, and Niraparib, had their marketing approval revoked for third-line and subsequent therapies due to an increased potential for adverse events. Consequently, the exploration of new treatment modalities remains imperative. Recently, the integration of PARPi with immune checkpoint inhibitors (ICIs) has emerged as a potential remedy option within the context of ovarian cancer. This article offers a comprehensive examination of the mechanisms and applications of PARPi and ICIs in the treatment of ovarian cancer. It synthesizes the existing evidence supporting their combined use and discusses key considerations that merit attention in ongoing development efforts.

Background

Ovarian cancer is the third most common cancer among female reproductive system tumors, with the highest mortality rate among gynecological malignancies [1]. Although tumor cytoreductive surgery and platinum-based chemotherapy have witnessed substantial progress in the management of ovarian cancer [2, 3], the recurrence rate in patients with ovarian cancer continues to be notably high [4]. Up to 70% of patients will relapse within 3 years, and the interval between subsequent relapses will become shorter and shorter until platinum resistance [5]. In order to delay and solve platinum resistance, it has become a breakthrough to improve the survival rate of ovarian cancer by finding effective maintenance therapy drugs [6]. Polyadenosine diphosphate ribose polymerase inhibitors (PARPi) are targeted drugs primarily designed for BRCA1/2 gene mutations or homologous recombination deficiency (HRD) in ovarian cancer. Their main mechanism is to kill tumor cells through the “synthetic lethality” effect [7]. The emergence of PARPi therapy, particularly in the context of BRCA1/2 mutations, has established a cornerstone for precision treatment in ovarian cancer [8]. Numerous clinical studies have consistently demonstrated that PARPi substantially extend progression-free survival (PFS) among ovarian cancer patients [9,10,11,12,13]. Currently, there are six PARPis approved by the Food and Drug Administration (FDA) for anticancer therapy. They are Olaparib, Rucaparib, Niraparib, Talazoparib, Fuzuloparib, and Pamiparib [14, 15]. Among them, Olaparib, Rucaparib, and Niraparib have obtained approvals from both the FDA and the European Medicines Agency (EMA) for using in epithelial ovarian cancer [16]. In 2022, the FDA withdrew the accelerated approval for certain indications of three PARPi (Rucaparib, Olaparib, and Niraparib) in the advanced treatment of ovarian cancer due to insufficient evidence of clinical benefit in confirmatory trials.

As the follow-up duration in corresponding clinical trials have been extended, the data reveals an elevated risk of mortality among patients who received second- or third-line or third-line maintenance treatment with PARPi compared to those who underwent chemotherapy [17]. This suggests that PARPi treatment did not result in overall survival (OS) benefits for those ovarian cancer patients who received it as a monotherapy in the third line or subsequent lines of treatment [18, 19]. Consequently, finding ways to enhance the survival outcomes for this specific group of patients have become a focal point of attention and research.

Immune checkpoint inhibitors (ICIs) have significantly altered the treatment paradigm for various malignant tumors, leading to substantial improvements in patient survival outcomes [20,21,22,23,24]. Preclinical data indicates that combining ICIs with PARPi could potentially generate synergistic effects, particularly in ovarian cancer patients who may not be suitable candidates for platinum-based retreatment [25,26,27,28]. In this article, we will provide an in-depth review of the development and clinical applications of PARPi and ICIs. We will also delve into their combined use in ovarian cancer, with a particular focus on their roles in second-line and subsequent-line treatments. To explore whether this combination can bring hope to ovarian cancer patients.

Treatment of PARP inhibitors in ovarian cancer

The synthetic lethal mechanism of PARP inhibitors

The human genome changes dynamically. According to statistics, each cell will experience more than 20,000 DNA damage events [29]. Healthy cells can resist the harmful effects of DNA damage through a series of interrelated molecular pathways, namely DNA damage response (DDR). These molecular pathways recognize DNA damage, delay the cell cycle and mediate DNA repair, thus maintaining genome integrity [30]. DNA damage repair includes: (1) DNA mismatch repair, (2) Base excision repair (BER), (3) Nucleotide excision repair, (4) DNA double-strand breaks (DSBs) repair, in which DNA DSBs repair is mediated by non-homologous end joining (NHEJ) and homologous recombination repair (HRR) pathways [31]. In the realm of DNA damage, the most severe forms of damage typically involve single-strand or double-strand breaks [31]. Single-strand DNA damage repair primarily depends on PARP enzymes, of which there are 17 in the human body. Among these enzymes, PARP-1 assumes a predominant role, contributing to approximately 90% of DNA repair processes [32]. DNA double-strand repair encompasses two primary pathways: NHEJ repair and HRR [31]. It is broadly recognized that BRCA proteins play a key role in the HRR pathway, being responsible for responsible for double-strand DNA repair. For normal cell DNA damage - single strand breaks, the cell can rely on PARP proteins for repair through the BER pathway. PARPi refers to small molecule inhibitors capable of triggering cell death by blocking the activity of PARP during DNA damage repair processes [33]. When PARPi acts on normal cells, PARP protein cannot play a role, and the inhibition of BER leads to the shortening of replication forks and the formation of double-strand breaks. At this time, BRCA1/2 can initiate the HRR pathway for cell repair. If the cell has homologous HRD, the double strand break caused by PARPi will not be repaired, and the “synthetic lethality” effect of the two will eventually lead to cell death [34, 35]. Of course, HRR repair is a complex process, and many genes and protein components are involved, including MRE11, RAD50, NBS1, ATM, ATR, etc., and BRCA1/2 are only some of the important components [36]. Mutations or silencing of expression in any gene in the HRR repair pathway will cause defects in the HRR repair pathway, and PARPi may exert anti-tumor activity through synthetic lethal effects. Its primary molecular mechanism involves competitive binding to the catalytic domain’s active site of the PARP enzyme using NAD + nicotinamide, which effectively inhibits the activity of the PARP enzyme. It cannot function by forming PAR polymers and recruiting DNA damage repair related proteins [37]. In this scenario, cells with impaired HRR mechanisms cannot effectively mend DNA damage via HRR. This gives rise to a synthetic lethal phenotype, ultimately causing the death of cancer cells. This phenomenon exemplifies the classic synthetic lethal mechanism [38] (Fig. 1).

Fig. 1
figure 1

Mechanism of PARP inhibitors. In cells with normal BRCA1/2 function, the homologous recombination repair (HRR) mechanism can repair DNA double-strand breaks (DSBs) and single-strand breaks (SSBs), preventing cell death. However, in patients with BRCA1/2 mutations, the HRR mechanism fails and cannot repair DNA DSBs. PARP inhibitors (PARPi) block PARP activity, further preventing the repair of SSBs and leading to the accumulation of DSBs. This induces synthetic lethality in patients with BRCA1/2 mutations and HRR deficiency, resulting in tumor cell death

Safety and tolerability of PARP inhibitors

The clinical application of PARPi is becoming increasingly widespread, not only in the treatment of ovarian cancer but also in other BRCA1/2-related cancers. While their therapeutic effects are significant, their safety and tolerability are also a major concern. The most common adverse events associated with PARPi include fatigue, nausea, and hematologic toxicity. Some patients also experience gastrointestinal, neurological, and respiratory toxicities [39], with serious side effects such as bone marrow suppression and secondary malignancies. Nearly all patients exposed to PARPi experience adverse events of any grade. Grade 1–2 toxicities are common in approximately two-thirds of patients, with similar incidence rates across all PARPi [40]. Grade 1–2 toxicities generally require monitoring while continuing treatment or pausing treatment for up to 28 days, but Grade 3 or higher toxicities should be carefully considered, with dose reduction or discontinuation of treatment if necessary [41].

Application of PARP inhibitors in ovarian cancer

Building upon the understanding of the PARP mechanism, a multitude of PARPi have been developed, primarily with a focus on cancer patients who possess BRCA1/2 mutations [42]. The prevalence of certain ovarian cancers is linked to mutations in the BRCA gene, which serves a critical function in DNA double-strand repair. Consequently, PARPi have gained extensive utilization in the management of ovarian cancer [43].

In 2009, PARPi’s first human clinical trial conducted a clinical evaluation of Olaparib, confirming for the first time the synthetic lethal interaction between PARPi and BRCA1/BRCA2 mutations [8, 44]. In the following ten years, relevant clinical research has been carried out in breast cancer, ovarian cancer, prostate cancer, gastric cancer, pancreatic cancer and other cancers [45, 46]. Olaparib, Rucaparib, and Niraparib have garnered approval from both the FDA and the EMA for their application in treating epithelial ovarian cancer [16]. Olaparib holds the distinction of being the inaugural PARPi to undergo extensive research and remains the most thoroughly evaluated to date. In April 2014, the European Commission granted the marketing license for Olaparib as the first treatment drug for platinum sensitive recurrent BRCA mutations in advanced serous epithelial ovarian cancer, fallopian tube cancer, or primary peritoneal cancer in adult patients with complete or partial response to platinum chemotherapy (AstraZeneca press release 18 December 2014). Subsequently, Rucaparib was approved in 2016, followed by Niraparib in 2017 [47].

In 2018, the SOLO-1 trial, which investigated Olaparib as a first-line maintenance treatment for primary ovarian cancer, reached the significant milestone of completing its five-year follow-up. Building upon this development, the FDA has now granted approval for Olaparib as a first-line maintenance treatment specifically for high-grade serous ovarian cancer (HGSC) with BRCA1/2 mutations in either embryonic or somatic cell lines [48]. In 2022, the SOLO-1 study’s seven-year follow-up data was unveiled, marking the most extended follow-up period to date for first-line maintenance therapy employing PARPi. This outcome reaffirms the long-term survival benefits associated with Olaparib maintenance therapy for patients with BRCA mutations [49]. The PRIMA trial, Study 19 trial, and PRIME trial (Table 1)have once more underscored the substantial effectiveness of single-agent PARPi in the maintenance treatment of ovarian cancer [9, 50]. In light of the aforementioned research findings, the use of specific PARPi is strongly recommended for first-line maintenance therapy in appropriate cases. Study 19 trial provided initial evidence suggesting that Olaparib could offer benefits to patients experiencing platinum-sensitive relapses [13]. The SOLO-2 study marked the inaugural confirmation that Olaparib maintenance treatment can significantly enhance OS in patients with gBRCA mutation [51]. The OPINION study further substantiated the effectiveness of Olaparib in the maintenance treatment of platinum-sensitive relapses in patients without gBRCA mutations [52]. The L-MOCA study demonstrated that Olaparib maintenance treatment could bring significant benefits to Asian platinum-sensitive relapse patients, irrespective of their BRCA mutation status [12]. The OReO study and NORA study further validated the substantial effectiveness of PARPi monotherapy in second-line maintenance therapy for ovarian cancer [11, 53]. Building on the research mentioned, PARPi have been established as the standard treatment for maintenance therapy in platinum-sensitive relapse cases. In real world studies, ovarian cancer patients newly diagnosed with the condition received PARPi maintenance therapy in natural clinical settings. These studies demonstrated significant PFS benefits when compared to patients who did not receive PARPi maintenance therapy [54, 55] (Table 1).

Table 1 Clinical trials of PARP inhibitors in ovarian cancer

However, in 2022, due to the potential increased risk of death in the quality of Olaparib, Niraparib, and Rucaparib [17, 18, 56], as a result of the FDA’s review, the indications for advanced treatment of ovarian cancer using these medications were revoked [57]. This has led to a rising clinical demand for the treatment of late-stage ovarian cancer.

The combination of PARPi and ICIs in ovarian cancer

ICIs treatment in ovarian cancer

The immune system plays a pivotal role in both restraining and facilitating cancer development by actively engaging in various phases of the body’s response to cancer [58]. Current immunotherapy for cancer is mainly divided into oncolytic virus therapy, cancer vaccines, cytokine therapy, adoptive cell, and ICIs [59]. Recent studies have confirmed that targeting immune checkpoint pathways has significant clinical efficiency [60]. Over the past few decades, significant breakthroughs have been made in the exploration of the treatment of tumor ICIs. Interventions aimed at programmed cell death 1 (PD-1), programmed cell death ligand 1 (PD-L1), and cytotoxic T lymphocyte-associated protein 4 (CTLA-4) have achieved broad acceptance for treating solid tumors [61, 62] (Fig. 2). PD-1 is a co-inhibitory receptor that exhibits widespread expression on T cells, natural killer (NK) cells, and B cells [63,64,65]. Indeed, PD-1 has two ligands: PD-L1 and PD-L2 [66]. Interaction between PD-1 and PD-L1 can suppress T cell proliferation, dampen T cell activation, and contribute to the prevalence of a tumor microenvironment characterized by helper T cell 2 (Th2) cytokines, which tends to favor tumor growth and development. Hence, inhibiting the interaction between PD-L1 and PD-1 can rejuvenate cytotoxic T lymphocyte (CTL) function, which had been compromised, and reestablish the capacity to eliminate cancer cells [67]. CTLA-4 and PD-1 have distinct roles in modulating T cell immunity. CTLA-4, a member of the CD28 immunoglobulin subfamily, serves as an inhibitory receptor primarily expressed by T cells. Its ligands, CD80 and CD86, are typically present on the surface of antigen-presenting cells. These ligands can interact with CD28, resulting in co-stimulation, or with CTLA-4, triggering co-inhibition reactions [68].

Fig. 2
figure 2

Mechanism of immune checkpoint inhibitors. PD-1/PD-L1 Pathway: Ovarian cancer cells express PD-L1 on their surface, which binds to PD-1 receptors on T cells, thereby inhibiting T cell activity. Anti-PD-1 and anti-PD-L1 antibodies can block this interaction, restoring the cytotoxic activity of T cells. CTLA-4/B7 Pathway: Dendritic cells express B7 molecules on their surface, which bind to CTLA-4 on T cells, inhibiting T cell activity. Anti-CTLA-4 antibodies can block this interaction, enhancing T cell activity. MHC/TCR Pathway: Dendritic cells present antigens via MHC class I molecules, which activate T cell receptors (TCR), further activating T cells

PD-L1 is present in approximately one-third of advanced ovarian cancer tumors, while the majority of tumor-infiltrating lymphocytes exhibit PD-1 expression [69]. Early studies have shown that PD-L1 expression in tumors is positively correlated with ovarian cancer survival [70]. This may also indicate that PD-1/PD-L1 plays a key role in the tumor immune response to ovarian cancer. Due to the lack of cytotoxic T lymphocytes and the immunosuppressive tumor microenvironment, ovarian cancer is included in the cold tumor range [71]. Cold tumors are tumors that have low activity to suppress immune cells and respond to treatment with tumors that are lower than those in the heat-immune category. Other names with the same category include invasive exclusion, non-inflammatory, or non-immunoreactive tumors [72].

The clinical use of ICIs in ovarian cancer has demonstrated limited efficacy, with clinical trial objective response rates typically ranging from approximately 6% to 15%. This outcome falls short of the substantial impact that immunotherapy has achieved in the treatment of metastatic and recurrent cervical cancer [73]. This may be related to low expression levels of PD-L1, low mutation load, and weak immunogenicity in ovarian cancer [71, 74, 75]. In September 2021, the National Comprehensive Cancer Network (NCCN) in the United States endorsed the utilization of Pembrolizumab for recurrent ovarian cancer patients who exhibit high microsatellite instability (MSI-H) or have deficient DNA mismatch repair (dMMR) and are either platinum-sensitive or platinum-resistant [76,77,78]. Currently, immunotherapy is often considered as a post-treatment option for ovarian cancer. This means that ICIs have emerged as a potential treatment choice for ovarian cancer. However, the outcomes have not been entirely satisfactory [79,80,81,82]. Finding new combinations to improve the efficacy of immunotherapy for ovarian cancer is a feasible approach.

Effects of PARPi on Immune Regulation

The effectiveness of ICIs relies on several factors, such as the level of PD-L1 expression, the quantity of tumor-infiltrating lymphocytes, the neoantigen load, and the tumor mutation burden [83, 84]. In terms of their mechanism of action, PARPi work by impeding DNA repair processes, amplifying DNA damage, generating new antigens and cytoplasmic DNA, activating the interferon pathway and initiating anti-tumor immune responses. Next, let’s outline the rationale behind combining PARPi with ICIs.

PARPi upregulates PD-L1

Research has revealed that Niraparib has the capacity to increase the expression of PD-L1 on the outer surface of ovarian cancer cells. This leads to an augmentation in the quantity and effectiveness of CD8 + T cells while simultaneously exerting immunosuppressive effects within the tumor microenvironment [85]. Several studies have demonstrated that PARPi can elevate the expression of PD-L1 in breast cancer cell lines and animal models. PARPi weakens the anti-cancer immune response by upregulating PD-L1, while concurrently blocking PD-L1 to enhance the sensitivity of cancer cells treated with PARPi to T cell-mediated killing. When compared to each drug administered individually, the combined treatment of PARPi and anti-PD-L1 significantly enhances the therapeutic efficacy in vivo [86]. Furthermore, research has identified that in an advanced serous ovarian cancer mouse model, the administration of Olaparib leads to an increase in PD-L1 expression. The combination of Olaparib and PD-L1 exhibits a notable effect, whereas the efficacy of anti-PD-L1 treatment alone is minimal or unresponsive [87]. Shen et al. conducted a study using a different model than the previous one, which involved combining talazoparib with anti-PD-L1 under normal homologous recombination (HR) conditions. In this study, they once again observed an upregulation of PD-L1 expression and a significant tumor response to the combined treatment of PARPi and ICIs [25]. In a recent study conducted by Jiao et al., it was suggested that PARPi can induce the expression of PD-L1. Their research involved treating MDA-MB-231 and BT549 breast cancer cells with olaparib or talazoparib, and the results demonstrated an increase in PD-L1 expression both in vitro and in vivo [86].

Research on the mechanism behind the upregulation of PD-L1 can generally be categorized into three main aspects. Firstly, studies have revealed that Niraparib triggers the activation of the cyclic GMP-AMP synthase (cGAS) –stimulator of interferon genes (STING) pathway, consequently leading to the upregulation of IFN-β. This represents a direct mechanism responsible for the increased expression of PD-L1 [85]. Secondly, in a separate study, it was demonstrated that PARPi primarily upregulates PD-L1 expression by deactivating glycogen synthase kinase-3 (GSK3β) [86, 88]. At last, Another study suggests that an alternative pathway through which PARPi upregulates PD-L1 involves the ATM-ATR-Checkpoint Kinase 1 (CHEK1) pathway, which functions as a kinase sensor for DSB. Upon activation of ATM, the signal kinase transitions from ATM to ATR. This switch from ATM to ATR activation subsequently triggers CHEK1 activation, further initiating the activation of the Janus kinase/signal transducer and transcriptional protein activator (JAK/STAT) signaling pathway, ultimately resulting in the upregulation of PD-L1 expression [89, 90].

Furthermore, in experiments utilizing a BRCA1 deficient ovarian cancer mouse model, it has been demonstrated that PARPi enhances the therapeutic effectiveness of CTLA-4 blockade [91]. The above findings illustrate PARPi’s capability to augment the therapeutic impact of ICIs.

PARPi activate the cGAS-STING signaling pathway

The immune pathway primarily comprises several signaling pathways, including the Toll-like receptors (TLRs) signaling pathway, C-type lectin receptors (CLR) signaling pathway, RIG-I-like receptor signaling pathway, and the cyclic cGAS-STING signaling pathway [92,93,94,95]. PARPi has the capability to activate the cGAS-STING innate immune pathway, which ultimately results in the production of type I interferons (IFN). This activation leads to a range of immunogenic effects and associated immune responses [96,97,98]. As a crucial innate immune sensor, the cGAS-STING pathway plays a pivotal role in regulating tumor growth and progression by facilitating the recruitment, initiation, and activation of anti-tumor immune cells [99, 100]. As mentioned previously, Niraparib triggers the activation of the cGAS-STING pathway, resulting in the upregulation of IFN-β and the elevation of PD-L1 expression. Conversely, when DNA binds to cGAS, it initiates the recruitment and activation of STING, which in turn facilitates the significant expression of CCL5 and CXCL10 via the involvement of TRAF family member-associated NF-kappa-B activator (TANK) binding kinase 1 and interferon regulatory factor 3. This process results in the recruitment of T cells and enhances the function of lymphocytes within ovarian cancer [85, 101, 102]. Olaparib’s inhibition of PARP elicits robust local and systemic anti-tumor immunity, encompassing both adaptive and innate immune responses, primarily via STING-dependent mechanisms in mice with BRCA1-deficient ovarian cancer. Furthermore, when combined with a PD-1 inhibitor, this effect is further intensified [87]. Additionally, certain studies have discovered that the treatment of ovarian cancer cells with talazoparib leads to a notable increase in the phosphorylation of two pivotal components within the STING pathway, namely IRF3 and TBK1 [25]. The sustained release of IFN-I from PARPi within the tumor microenvironment (TME) plays a vital role in promoting various immune functions. This includes the activation of dendritic cells, the preservation of cross-presentation of tumor-derived antigens to T cells, the support of NK cell-mediated anti-tumor immunity, and synergistic activation with Toll-like receptor 4 (TLR4) ligands like HMGB1 to stimulate M1 anti-tumor macrophages [103, 104].

PARPi increases genomic instability

Because of HRD and BRCA mutations, cancer cells exhibit heightened genomic instability, rendering them more immunogenic. Following PARPi administration, the induction of severe DNA damage leads to the accumulation of DNA fragments within the cytoplasm. As a consequence, this process generates a greater number of novel antigens and exposes them on the cell surface, resulting in heightened activation of the immune response. This, in turn, leads to an increase in tumor mutational burden (TMB) and elevated immunogenicity [105, 106]. Especially in cells with HRD, it becomes feasible to reestablish a productive Th1 immune response and reset the tumor microenvironment. Moreover, PARPi induces sustained DNA damage, resulting in epigenetic alterations within tumor cells. These changes render tumor cells more receptive to the influence of T cells and NK cells, ultimately culminating in an enhanced intrinsic immunogenicity of the tumor cells [107]. Conversely, ICIs can rescue the tumor microenvironment from the consequences of inadequate immune cell infiltration and facilitate the recognition of newly formed antigens resulting from chronic DNA damage induced by PARPi [27].

In summary, PARPi exert their effects through several key mechanisms. Firstly, they can activate the cGAS-STING pathway, leading to the activation of GSK3β. Concurrently, through the ATM-ATR-CHEK1 pathway, they upregulate PD-L1 expression. Secondly, PARPi activate the immune pathway, resulting in the release of IFN-I and the promotion of the expression of chemokines CCL5 and CXCL10. Finally, PARPi can increase genetic instability, resulting in a higher TMB and heightened immune responsiveness (Fig. 3).

Fig. 3
figure 3

The interaction between PARP inhibitors and immune checkpoint inhibitors. PARP inhibitors enhance tumor immunogenicity and the presentation of neoantigens in the tumor microenvironment (TME) by increasing antigen presentation and upregulating PD-L1 expression. Dendritic cells, through the cGAS/STING pathway, activate and present these neoantigens, promoting the recruitment and activation of T cells. Immune checkpoint inhibitors restore the cytotoxic activity of T cells by blocking the PD-1/PD-L1 and CTLA-4/B7 pathways, further enhancing the anti-tumor immune response. The combined use of PARP inhibitors and immune checkpoint inhibitors not only increases antigen presentation and T cell activity but also upregulates PD-L1 expression. Additionally, this combination promotes the secretion of IFN-β through the JAK-STAT and GSK3β pathways, further enhancing the immune system’s attack on tumor cells

Clinical study of PARPi combined with ICIs in ovarian cancer

In recent years, ICIs including anti-PD-1 antibodies and anti-PD-L1 antibodies, have demonstrated notable efficacy in various types of tumors [108]. The combination of ICIs with PARPi appears to be a viable approach. Preclinical data suggests that this combination therapy of ICIs and PARPi may yield synergistic effects, particularly benefiting ovarian cancer patients who may not be suitable candidates for platinum-based retreatment [25, 50]. This could potentially offer novel treatment alternatives for a subset of advanced ovarian cancer patients who currently lack effective therapeutic options. In summary, it is a feasible approach to improve the efficacy of ICIs by using PARPi to influence the immune system and tumor microenvironment. And this combination has been carried out in many clinical trials.

The MEDIOLA trial is a multicenter, open-label, phase 1/2 clinical trial conducted to assess the treatment of solid tumors using Duvalizumab and Olaparib. In this trial, the objective response rate (ORR) for the combination of Duvalizumab and Olaparib reached an impressive 71.9% in patients with platinum-sensitive recurrent ovarian cancer who had gBRCA mutations [109]. The early results of the Phase II study of MEDIOLA (NCT02734004) showed good efficacy and safety in the combination of olapanib and Duvalizumab in platinum sensitive recurrent ovarian cancer with germline BRCA1 and/or BRCA2 mutations (gBRCAm). In gBRCAm patients, the ORR of Olapanib combined with Duvalizumab was 92.2%, and over 40% of patients had CR [110]. The findings from this study suggest that the combination of Olaparib and Duvalizumab demonstrates encouraging anti-tumor activity and safety in the context of recurrent ovarian cancer.

Based on the preliminary research results of the MEDIOLA trial, a Phase III DUO-O trial was conducted, The DUO-O study is a randomized, double-blind, placebo-controlled multicenter phase III study aimed at exploring the efficacy of Bevacizumab + Olaparib + Duvalizumab in first-line maintenance therapy for BRCA wild-type newly diagnosed ovarian cancer. The results showed that patients receiving triple maintenance therapy had significantly higher PFS than those receiving bevacizumab monotherapy, at 24.2 months and 19.3 months, respectively [111]. TOPACIO/KEYNOTE-162 is a Single-Arm, Phases 1/2 trial. The study is designed to evaluate the efficacy of Niraparib in combination with pembrolizumab in patients with relapsing platinum resistance. In the population with ovarian cancer, the ORR was 18% and the disease control rate was 65%. Among them, 3 (5%) confirmed complete remission, 8 (13%) confirmed partial remission, 28 (47%) were stable, and 20 (33%) were progressing [112]. This study demonstrates that the combination of Niraparib and Pembrolizumab therapy exhibits promising anti-tumor activity in ovarian cancer patients and is worthy of further research.

The MOONSTONE study is an open-label, single-arm Phase 2 trial. It aims to assess the effectiveness and safety of combining Niraparib and Dostarlimab. This study includes participants with advanced, relapsed high-grade ovarian, fallopian tube, endometrioid, clear cell ovarian, or primary peritoneal cancer. These participants do not possess a known breast cancer susceptibility gene (BRCA) mutation. They also have platinum-resistant disease and have previously undergone treatment with Bevacizumab. However, the results of this study did not meet the desired level of satisfaction [113] (Table 2).

Table 2 Clinical trials combining PARP inhibitors and immune checkpoint inhibitors in ovarian cancer

Both preclinical and clinical data suggest that monotherapy with PARPi and ICIs have limitations in the management of ovarian cancer [114]. The clinical studies mentioned above have been completed, but it’s worth noting that there are numerous ongoing or actively recruiting clinical trials that are exploring the combination of PARPi and ICIs. These trials hold significant promise and merit our attention.

Discussion

Recent clinical studies have demonstrated that the treatment combination of PARPi and ICIs holds practical clinical significance, particularly for advanced ovarian cancer patients with limited treatment options. Notably, both the MEDIOLA study and the TOPACIO study have highlighted that the synergy between PARPi and ICIs can be harnessed effectively in the context of ovarian cancer treatment.

Immunotherapy has transformed cancer treatment, but it’s important to note that the effectiveness of ICIs, whether used as standalone treatments or in combination with chemotherapy is not yet satisfactory. Hence, there is a need to investigate alternative combination approaches and novel immunotherapy techniques for targeted medications. The triple combination of ICIs, PARPi, and anti-angiogenic drugs appears to yield promising outcomes in the treatment of recurrent ovarian cancer. The results from the MEDIOLA trial demonstrate that the response rate to the triple therapy is significantly higher compared to the dual therapy of PARPi and ICIs. This presents a novel clinical treatment approach for consideration. Furthermore, the quest for novel biomarkers can aid in the identification of patients with a higher likelihood of responding favorably to combination therapy protocols. Moreover, it is imperative to uncover distinct resistance mechanisms to PARPi and ICIs, thereby establishing a fresh theoretical framework for the integration of these strategies. There is an urgent need to explore novel biomarkers for the precise screening of individuals suitable for this combination approach. The toxic attributes of PARPi, primarily associated with bone marrow suppression, result in adverse events for some patients undergoing monotherapy. These toxicities also impose limitations on the application of PARPi within certain otherwise viable combination strategies, such as in conjunction with chemotherapy [115]. However, it’s worth noting that immunosuppressants typically do not exhibit a significant bone marrow suppression effect. Consequently, the combination regimen can mitigate the overlap of drug side effects, thereby enhancing the safety profile of this combination approach. During the execution of clinical trials, it is imperative to take into account the potential adverse reactions associated with combination therapy. This entails a thorough investigation of the optimal dosage and timing of administration, with the ultimate goal of ensuring patient tolerance and safety. To steer clinical trials toward a sensible combination of treatments.

PARPi currently face several issues in clinical use, such as acquired resistance in a significant portion of patients after initial treatment [116]. The restoration of HRR is a primary cause of PARPi resistance. Other factors, including reversion mutations, replication fork protection, epigenetic modifications, restoration of ADP-ribosylation (PARylation), and pharmacological changes, also contribute to PARPi resistance [117, 118]. HRR restoration includes secondary mutations in BRCA1/2 genes and regulation of other proteins in the HRR pathway. Recent studies have revealed that PARPi can activate STING-dependent intrinsic immunity in tumor cells [97, 119]. This immune activation has inspired the combination of ICIs with PARPi, as suggested by clinical trial results, which seem to indicate that ICIs can reverse PARPi resistance. Additionally, PARPi primarily target BRCA1/2 gene mutations or HRD. However, in high-grade serous ovarian cancer, only 25% of patients with BRCA mutations are more sensitive to platinum-based chemotherapy and PARPi, while the remaining 75% are BRCAwt patients [120]. Clinical exploration for the benefits of PARPi in BRCAwt patients is ongoing. The NORA study showed that in the BRCAwt subgroup, the niraparib group had a significantly extended PFS compared to the placebo group (11.1 vs. 3.9 months) [53]. Results from the 2022 SGO Annual Meeting showed that niraparib monotherapy maintenance extended the median PFS for the “BRCA and HRD double-negative” population to 14 months, compared to 5.5 months for the placebo group. This demonstrates the efficacy of PARPi in BRCAwt populations, but more clinical trials are needed to verify this. Moreover, long-term use of PARPi can lead to severe bone marrow suppression and other adverse effects. To address these issues, combining PARPi with other drugs such as ICIs can overcome resistance. Additionally, developing PARPi with higher selectivity and potency but fewer side effects is essential.

It is noteworthy that in recent years, RAS has played a crucial role in the development and progression of ovarian cancer. RAS is an important signaling protein belonging to the small GTPase family, including three main proteins: KRAS, NRAS, and HRAS [121]. In ovarian cancer, the mutation status of RAS, especially KRAS at codons 12, 13, and 61, accounts for 6% to 65%. KRAS mutations are also considered biomarkers of poor outcomes and resistance to various drugs in ovarian cancer [122]. RAS is also a determinant for several small molecule therapies, such as MEK inhibitors [123]. Interestingly, studies have found that KRAS mutant tumor models exhibit resistance to PARPi, anti-PD-L1, and the combination of PARPi and PD-L1 inhibitors. MEK inhibitors can trigger and amplify PARPi-induced DNA damage, cytoplasmic dsDNA accumulation, STING pathway activation, and CD8 + T cell recruitment. Additionally, MEKi reduces myeloid-derived suppressor cell (MDSC) infiltration, at least partially by decreasing IL-6 and GM-CSF. The results of this study demonstrated significant efficacy of the triple therapy of PARPi, MEKi, and anti-PD-L1 blockade in KRAS mutant tumor models [124]. This suggests that RAS and possibly other proteins play significant roles in resistance to ovarian cancer treatments. Utilizing drugs targeting these proteins in combination with PARPi and ICIs might yield better results. This area warrants further research and clinical trials, and we look forward to better therapeutic combinations in the future.

It is well-known that glutamine metabolism plays a central role in altered metabolism in cancer cells [125]. As mentioned earlier, KRas itself is considered undruggable. Some studies have utilized the late G1 glutamine (Gln)-dependent cell cycle checkpoint bypass in cancer cells with KRAS mutations [126]. Upon Gln deprivation, KRas-driven cancer cells enter the S phase and stall due to insufficient nucleotide biosynthesis [127]. Cells stalled in the S phase are more susceptible to cytotoxic drugs, resulting in better cancer cell killing effects. For example, in ovarian cancer, platinum-resistant tumor cells show increased glutamine metabolism, and glutaminase (GLS) inhibitors BPTES and 968 can sensitize chemotherapy-resistant ovarian cancer cells to platinum-based chemotherapeutic agents [128]. Recent studies have found that treatment with the GLS inhibitor CB-839 makes cells susceptible to Olaparib and extends survival in tumor-bearing mice, suggesting that combined treatment with GLS inhibitors and PARPi can effectively treat chemotherapy-resistant ovarian cancer [129]. Therefore, could the combination of GLS inhibitors, PARPi, and ICIs achieve better efficacy in ovarian cancer? This could be a future direction for exploration.

Conclusion

PARPi and ICIs are both effective treatments for ovarian cancer. PARPi have become a standard element in the treatment of ovarian cancer, while immunotherapy with ICIs is well-suited for addressing non-resectable or metastatic microsatellite instability or mismatch repair deficiency solid tumors. PARPi can regulate the immune microenvironment by upregulating PD-L1, activating immune pathways, increasing genomic instability, and promoting tumor response to ICIs. Current conversion and preclinical data provide strong evidence for the synergistic potential of combining PARPi with ICIs. However, additional investigation is necessary to delve deeper into the clinical trial results. In summary, the synergistic combination of PARPi and ICIs holds promise for improving outcomes in patients with advanced ovarian cancer.

Data availability

No data was used for the research described in the article.

Abbreviations

PARPi:

Polyadenosine diphosphate ribose polymerase inhibitors

ICIs:

Immune checkpoint inhibitors

HRD:

Homologous recombination deficiency

BER:

Base excision repair

DSBs:

DNA double-strand breaks

NHEJ:

Non-homologous end joining

PFS:

Progression-free survival

FDA:

Food and Drug Administration

EMA:

European Medicines Agency

OS:

Overall survival

NHEJ:

Non-homologous end joining

HRR:

Homologous recombination repair

HRD:

Recombination repair defect

HGSC:

High-grade serous ovarian cancer

PD-1:

Programmed cell death 1

PD-L1:

Programmed cell death ligand 1

CTLA-4:

Cytotoxic T lymphocyte-associated protein 4

Th2:

T cell 2

CTL:

Cytotoxic T lymphocyte

NCCN:

Comprehensive Cancer Network

MSI-H:

High microsatellite instability

dMMR:

Deficient DNA mismatch repair

GSK3β:

Glycogen synthase kinase-3

CHEK1:

Checkpoint Kinase 1

TMB:

Tumor mutational burden

ORR:

Objective response rate

DDR:

DNA damage response

References

  1. Siegel RL, Miller KD, Wagle NS, Jemal A. Cancer statistics, 2023. CA Cancer J Clin. 2023;73(1):17–48.

    Article  PubMed  Google Scholar 

  2. Goh J, Mohan GR, Ladwa R, Ananda S, Cohen PA, Baron-Hay S. Frontline treatment of epithelial ovarian cancer. Asia Pac J Clin Oncol. 2015;11(Suppl 6):1–16.

    Article  PubMed  Google Scholar 

  3. Provencher DM, Gallagher CJ, Parulekar WR, Ledermann JA, Armstrong DK, Brundage M, et al. OV21/PETROC: a randomized Gynecologic Cancer Intergroup phase II study of intraperitoneal versus intravenous chemotherapy following neoadjuvant chemotherapy and optimal debulking surgery in epithelial ovarian cancer. Ann Oncol. 2018;29(2):431–8.

    Article  CAS  PubMed  Google Scholar 

  4. Zhang J, Li XB, Ji ZH, Ma R, Bai WP, Li Y. Cytoreductive Surgery plus Hyperthermic Intraperitoneal Chemotherapy Improves Survival with Acceptable Safety for Advanced Ovarian Cancer: A Clinical Study of 100 Patients. Biomed Res Int. 2021, 2021: 5533134.

  5. Lheureux S, Braunstein M, Oza AM. Epithelial ovarian cancer: evolution of management in the era of precision medicine. CA Cancer J Clin. 2019;69(4):280–304.

    Article  PubMed  Google Scholar 

  6. Pujade-Lauraine E, Banerjee S, Pignata S. Management of Platinum-Resistant, relapsed epithelial ovarian Cancer and New Drug perspectives. J Clin Oncol. 2019;37(27):2437–48.

    Article  CAS  PubMed  Google Scholar 

  7. Curtin NJ, Szabo C. Poly(ADP-ribose) polymerase inhibition: past, present and future. Nat Rev Drug Discov. 2020;19(10):711–36.

    Article  CAS  PubMed  Google Scholar 

  8. 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.

    Article  CAS  PubMed  Google Scholar 

  9. Gonzalez-Martin A, Pothuri B, Vergote I, DePont Christensen R, Graybill W, Mirza MR, et al. Niraparib in patients with newly diagnosed Advanced Ovarian Cancer. N Engl J Med. 2019;381(25):2391–402.

    Article  CAS  PubMed  Google Scholar 

  10. Hardesty MM, Krivak TC, Wright GS, Hamilton E, Fleming EL, Belotte J, et al. OVARIO phase II trial of combination niraparib plus bevacizumab maintenance therapy in advanced ovarian cancer following first-line platinum-based chemotherapy with bevacizumab. Gynecol Oncol. 2022;166(2):219–29.

    Article  CAS  PubMed  Google Scholar 

  11. Morgan RD, Clamp AR, White DJ, Price M, Burghel GJ, Ryder WDJ, et al. Multi-maintenance Olaparib Therapy in Relapsed, Germline BRCA1/2-Mutant high-Grade Serous Ovarian Cancer (MOLTO): a phase II trial. Clin Cancer Res. 2023;29(14):2602–11.

    Article  CAS  PubMed  Google Scholar 

  12. Gao Q, Zhu J, Zhao W, Huang Y, An R, Zheng H, et al. Olaparib Maintenance Monotherapy in Asian patients with platinum-sensitive relapsed ovarian Cancer: phase III trial (L-MOCA). Clin Cancer Res. 2022;28(11):2278–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Ledermann J, Harter P, Gourley C, Friedlander M, Vergote I, Rustin G, et al. Olaparib maintenance therapy in platinum-sensitive relapsed ovarian cancer. N Engl J Med. 2012;366(15):1382–92.

    Article  CAS  PubMed  Google Scholar 

  14. 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.

    Article  PubMed  Google Scholar 

  15. Markham A, Pamiparib. First Approval Drugs. 2021;81(11):1343–8.

    CAS  PubMed  Google Scholar 

  16. Boussios S, Karathanasi A, Cooke D, Neille C, Sadauskaite A, Moschetta M et al. PARP inhibitors in ovarian Cancer: the Route to Ithaca. Diagnostics (Basel) 2019, 9(2).

  17. Penson RT, Valencia RV, Cibula D, Colombo N, Leath CA 3rd, Bidzinski M, et al. Olaparib Versus Nonplatinum Chemotherapy in patients with platinum-sensitive relapsed ovarian Cancer and a germline BRCA1/2 mutation (SOLO3): a Randomized Phase III Trial. J Clin Oncol. 2020;38(11):1164–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Oza AM, Matulonis UA, Malander S, Hudgens S, Sehouli J, Del Campo JM, et al. Quality of life in patients with recurrent ovarian cancer treated with niraparib versus placebo (ENGOT-OV16/NOVA): results from a double-blind, phase 3, randomised controlled trial. Lancet Oncol. 2018;19(8):1117–25.

    Article  PubMed  Google Scholar 

  19. Coleman RL, Oza AM, Lorusso D, Aghajanian C, Oaknin A, Dean A, et al. Rucaparib maintenance treatment for recurrent ovarian carcinoma after response to platinum therapy (ARIEL3): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet. 2017;390(10106):1949–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Robert C, Ribas A, Schachter J, Arance A, Grob JJ, Mortier L, et al. Pembrolizumab versus Ipilimumab in advanced melanoma (KEYNOTE-006): post-hoc 5-year results from an open-label, multicentre, randomised, controlled, phase 3 study. Lancet Oncol. 2019;20(9):1239–51.

    Article  CAS  PubMed  Google Scholar 

  21. Middleton G, Brock K, Savage J, Mant R, Summers Y, Connibear J, et al. Pembrolizumab in patients with non-small-cell lung cancer of performance status 2 (PePS2): a single arm, phase 2 trial. Lancet Respir Med. 2020;8(9):895–904.

    Article  CAS  PubMed  Google Scholar 

  22. Necchi A, Anichini A, Raggi D, Briganti A, Massa S, Luciano R, et al. Pembrolizumab as Neoadjuvant Therapy before Radical Cystectomy in patients with muscle-invasive urothelial bladder carcinoma (PURE-01): an Open-Label, Single-Arm, phase II study. J Clin Oncol. 2018;36(34):3353–60.

    Article  CAS  PubMed  Google Scholar 

  23. Muro K, Chung HC, Shankaran V, Geva R, Catenacci D, Gupta S, et al. Pembrolizumab for patients with PD-L1-positive advanced gastric cancer (KEYNOTE-012): a multicentre, open-label, phase 1b trial. Lancet Oncol. 2016;17(6):717–26.

    Article  CAS  PubMed  Google Scholar 

  24. Makker V, Rasco D, Vogelzang NJ, Brose MS, Cohn AL, Mier J, et al. Lenvatinib plus Pembrolizumab in patients with advanced endometrial cancer: an interim analysis of a multicentre, open-label, single-arm, phase 2 trial. Lancet Oncol. 2019;20(5):711–8.

    Article  CAS  PubMed  Google Scholar 

  25. 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.

    Article  CAS  PubMed  Google Scholar 

  26. Vikas P, Borcherding N, Chennamadhavuni A, Garje R. Therapeutic potential of combining PARP inhibitor and immunotherapy in solid tumors. Front Oncol. 2020;10:570.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Li T, Wang X, Qin S, Chen B, Yi M, Zhou J. Targeting PARP for the optimal immunotherapy efficiency in gynecologic malignancies. Biomed Pharmacother. 2023;162:114712.

    Article  CAS  PubMed  Google Scholar 

  28. Bhamidipati D, Haro-Silerio JI, Yap TA, Ngoi N. PARP inhibitors: enhancing efficacy through rational combinations. Br J Cancer 2023.

  29. Loeb LA. Human cancers express mutator phenotypes: origin, consequences and targeting. Nat Rev Cancer. 2011;11(6):450–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. D’Amours D, Desnoyers S, D’Silva I, Poirier GG. Poly(ADP-ribosyl)ation reactions in the regulation of nuclear functions. Biochem J. 1999;342(Pt 2):249–68.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Jalal S, Earley JN, Turchi JJ. DNA repair: from genome maintenance to biomarker and therapeutic target. Clin Cancer Res. 2011;17(22):6973–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Luscher B, Ahel I, Altmeyer M, Ashworth A, Bai P, Chang P, et al. ADP-ribosyltransferases, an update on function and nomenclature. FEBS J. 2022;289(23):7399–410.

    Article  PubMed  Google Scholar 

  33. Zong C, Zhu T, He J, Huang R, Jia R, Shen J. PARP mediated DNA damage response, genomic stability and immune responses. Int J Cancer. 2022;150(11):1745–59.

    Article  CAS  PubMed  Google Scholar 

  34. Bhamidipati D, Haro-Silerio JI, Yap TA, Ngoi N. PARP inhibitors: enhancing efficacy through rational combinations. Br J Cancer. 2023;129(6):904–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Kyo S, Kanno K, Takakura M, Yamashita H, Ishikawa M, Ishibashi T et al. Clinical Landscape of PARP inhibitors in Ovarian Cancer: Molecular mechanisms and clues to overcome resistance. Cancers (Basel) 2022, 14(10).

  36. Rass E, Willaume S, Bertrand P. 53BP1: keeping it under control, even at a Distance from DNA damage. Genes (Basel) 2022, 13(12).

  37. Min A, Im SA. PARP inhibitors as therapeutics: beyond modulation of PARylation. Cancers (Basel) 2020, 12(2).

  38. Lord CJ, Ashworth A. PARP inhibitors: synthetic lethality in the clinic. Science. 2017;355(6330):1152–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Cecere SC, Casartelli C, Forte M, Pignata S, Pisano C. Safety of PARP inhibitors as maintenance therapy in ovarian cancer. Expert Opin Drug Saf. 2023;22(10):897–908.

    Article  CAS  PubMed  Google Scholar 

  40. Chelariu-Raicu A, Trillsch F, Burges A, Czogalla B, Hester A, Wuerstlein R, et al. PARP inhibitors: risk factors for toxicity and matching patients to the proper poly (ADP-ribose) polymerase inhibitor (PARPi) therapy. Int J Gynecol Cancer. 2023;33(5):812–22.

    Article  PubMed  Google Scholar 

  41. Tian X, Chen L, Gai D, He S, Jiang X, Zhang N. Adverse event profiles of PARP inhibitors: analysis of spontaneous reports submitted to FAERS. Front Pharmacol. 2022;13:851246.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhao D, Long X, Wang J. Dose Adjustment of Poly (ADP–Ribose) polymerase inhibitors in patients with hepatic or renal impairment. Drug Des Devel Ther. 2022;16:3947–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Varol U, Kucukzeybek Y, Alacacioglu A, Somali I, Altun Z, Aktas S, et al. BRCA genes: BRCA 1 and BRCA 2. J BUON. 2018;23(4):862–6.

    PubMed  Google Scholar 

  44. Fong PC, Boss DS, Yap TA, Tutt A, Wu P, Mergui-Roelvink M, et al. Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N Engl J Med. 2009;361(2):123–34.

    Article  CAS  PubMed  Google Scholar 

  45. Wang Y, Zheng K, Huang Y, Xiong H, Su J, Chen R, et al. PARP inhibitors in gastric cancer: beacon of hope. J Exp Clin Cancer Res. 2021;40(1):211.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Shao C, Wan J, Lam FC, Tang H, Marley AR, Song Y, et al. A comprehensive literature review and meta-analysis of the prevalence of pan-cancer BRCA mutations, homologous recombination repair gene mutations, and homologous recombination deficiencies. Environ Mol Mutagen. 2022;63(6):308–16.

    Article  CAS  PubMed  Google Scholar 

  47. O’Malley DM. New therapies for Ovarian Cancer. J Natl Compr Canc Netw. 2019;17(55):619–21.

    PubMed  Google Scholar 

  48. Banerjee S, Moore KN, Colombo N, Scambia G, Kim BG, Oaknin A, et al. Maintenance olaparib for patients with newly diagnosed advanced ovarian cancer and a BRCA mutation (SOLO1/GOG 3004): 5-year follow-up of a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 2021;22(12):1721–31.

    Article  CAS  PubMed  Google Scholar 

  49. 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.

    Article  CAS  PubMed  Google Scholar 

  50. Li N, Zhu J, Yin R, Wang J, Pan L, Kong B et al. Treatment with Niraparib Maintenance Therapy in patients with newly diagnosed Advanced Ovarian Cancer: a phase 3 Randomized Clinical Trial. JAMA Oncol 2023.

  51. Pujade-Lauraine E, Ledermann JA, Selle F, Gebski V, Penson RT, Oza AM, et al. Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): a double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 2017;18(9):1274–84.

    Article  CAS  PubMed  Google Scholar 

  52. Poveda A, Lheureux S, Colombo N, Cibula D, Lindemann K, Weberpals J, et al. Olaparib maintenance monotherapy in platinum-sensitive relapsed ovarian cancer patients without a germline BRCA1/BRCA2 mutation: OPINION primary analysis. Gynecol Oncol. 2022;164(3):498–504.

    Article  CAS  PubMed  Google Scholar 

  53. Wu XH, Zhu JQ, Yin RT, Yang JX, Liu JH, Wang J, et al. Niraparib maintenance therapy in patients with platinum-sensitive recurrent ovarian cancer using an individualized starting dose (NORA): a randomized, double-blind, placebo-controlled phase III trial(). Ann Oncol. 2021;32(4):512–21.

    Article  CAS  PubMed  Google Scholar 

  54. Chan JK, Liu J, Song J, Xiang C, Wu EQ, Kalilani L, et al. Real-world trends of PARPi maintenance treatment uptake and progression-free survival (PFS) in patients (pts) with newly diagnosed advanced ovarian cancer (AOC) in the United States. J Clin Oncol. 2022;40(16suppl):6580–6580.

    Article  Google Scholar 

  55. Chan JK, Liu J, Song J, Xiang C, Wu E, Kalilani L, et al. Real-world outcomes Associated with Poly(ADP-ribose) polymerase inhibitor monotherapy maintenance in patients with primary Advanced Ovarian Cancer. Am J Clin Oncol. 2023;46(7):314–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Kristeleit R, Lisyanskaya A, Fedenko A, Dvorkin M, de Melo AC, Shparyk Y, et al. Rucaparib versus standard-of-care chemotherapy in patients with relapsed ovarian cancer and a deleterious BRCA1 or BRCA2 mutation (ARIEL4): an international, open-label, randomised, phase 3 trial. Lancet Oncol. 2022;23(4):465–78.

    Article  CAS  PubMed  Google Scholar 

  57. Lee JY, Lee YY, Park JY, Shim SH, Kim SI, Kong TW, et al. Major clinical research advances in gynecologic cancer in 2022: highlight on late-line PARP inhibitor withdrawal in ovarian cancer, the impact of ARIEL-4, and SOLO-3. J Gynecol Oncol. 2023;34(2):e51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Schreiber RD, Old LJ, Smyth MJ. Cancer immunoediting: integrating immunity’s roles in cancer suppression and promotion. Science. 2011;331(6024):1565–70.

    Article  CAS  PubMed  Google Scholar 

  59. Zhang Y, Zhang Z. The history and advances in cancer immunotherapy: understanding the characteristics of tumor-infiltrating immune cells and their therapeutic implications. Cell Mol Immunol. 2020;17(8):807–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Bagchi S, Yuan R, Engleman EG. Immune Checkpoint inhibitors for the treatment of Cancer: clinical impact and mechanisms of response and resistance. Annu Rev Pathol. 2021;16:223–49.

    Article  CAS  PubMed  Google Scholar 

  61. Zhu S, Zhang T, Zheng L, Liu H, Song W, Liu D, et al. Combination strategies to maximize the benefits of cancer immunotherapy. J Hematol Oncol. 2021;14(1):156.

    Article  PubMed  PubMed Central  Google Scholar 

  62. Seidel JA, Otsuka A, Kabashima K. Anti-PD-1 and Anti-CTLA-4 therapies in Cancer: mechanisms of Action, Efficacy, and limitations. Front Oncol. 2018;8:86.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Keir ME, Butte MJ, Freeman GJ, Sharpe AH. PD-1 and its ligands in tolerance and immunity. Annu Rev Immunol. 2008;26:677–704.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Blank C, Mackensen A. Contribution of the PD-L1/PD-1 pathway to T-cell exhaustion: an update on implications for chronic infections and tumor evasion. Cancer Immunol Immunother. 2007;56(5):739–45.

    Article  PubMed  Google Scholar 

  65. Flies DB, Sandler BJ, Sznol M, Chen L. Blockade of the B7-H1/PD-1 pathway for cancer immunotherapy. Yale J Biol Med. 2011;84(4):409–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Mahoney KM, Freeman GJ, McDermott DF. The Next Immune-Checkpoint inhibitors: PD-1/PD-L1 blockade in Melanoma. Clin Ther. 2015;37(4):764–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Pardoll DM. The blockade of immune checkpoints in cancer immunotherapy. Nat Rev Cancer. 2012;12(4):252–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Van Coillie S, Wiernicki B, Xu J. Molecular and Cellular functions of CTLA-4. Adv Exp Med Biol. 2020;1248:7–32.

    Article  PubMed  Google Scholar 

  69. Drakes ML, Mehrotra S, Aldulescu M, Potkul RK, Liu Y, Grisoli A, et al. Stratification of ovarian tumor pathology by expression of programmed cell death-1 (PD-1) and PD-ligand- 1 (PD-L1) in ovarian cancer. J Ovarian Res. 2018;11(1):43.

    Article  PubMed  PubMed Central  Google Scholar 

  70. Darb-Esfahani S, Kunze CA, Kulbe H, Sehouli J, Wienert S, Lindner J, et al. Prognostic impact of programmed cell death-1 (PD-1) and PD-ligand 1 (PD-L1) expression in cancer cells and tumor-infiltrating lymphocytes in ovarian high grade serous carcinoma. Oncotarget. 2016;7(2):1486–99.

    Article  PubMed  Google Scholar 

  71. Mortezaee K. Enriched cancer stem cells, dense stroma, and cold immunity: interrelated events in pancreatic cancer. J Biochem Mol Toxicol. 2021;35(4):e22708.

    Article  CAS  PubMed  Google Scholar 

  72. Mortezaee K, Najafi M. Immune system in cancer radiotherapy: resistance mechanisms and therapy perspectives. Crit Rev Oncol Hematol. 2021;157:103180.

    Article  PubMed  Google Scholar 

  73. Yang C, Xia BR, Zhang ZC, Zhang YJ, Lou G, Jin WL. Immunotherapy for Ovarian Cancer: adjuvant, combination, and Neoadjuvant. Front Immunol. 2020;11:577869.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Cancer Genome Atlas Research N. Integrated genomic analyses of ovarian carcinoma. Nature. 2011;474(7353):609–15.

    Article  Google Scholar 

  75. Ciriello G, Miller ML, Aksoy BA, Senbabaoglu Y, Schultz N, Sander C. Emerging landscape of oncogenic signatures across human cancers. Nat Genet. 2013;45(10):1127–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Marabelle A, Le DT, Ascierto PA, Di Giacomo AM, De Jesus-Acosta A, Delord JP, et al. Efficacy of Pembrolizumab in patients with Noncolorectal high microsatellite Instability/Mismatch repair-deficient Cancer: results from the phase II KEYNOTE-158 study. J Clin Oncol. 2020;38(1):1–10.

    Article  CAS  PubMed  Google Scholar 

  77. O’Malley DM, Bariani GM, Cassier PA, Marabelle A, Hansen AR, De Jesus Acosta A, et al. Pembrolizumab in patients with microsatellite instability-high Advanced Endometrial Cancer: results from the KEYNOTE-158 study. J Clin Oncol. 2022;40(7):752–61.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Marcus L, Fashoyin-Aje LA, Donoghue M, Yuan M, Rodriguez L, Gallagher PS, et al. FDA approval Summary: Pembrolizumab for the treatment of Tumor Mutational Burden-High Solid tumors. Clin Cancer Res. 2021;27(17):4685–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Disis ML, Taylor MH, Kelly K, Beck JT, Gordon M, Moore KM, et al. Efficacy and safety of Avelumab for patients with recurrent or refractory ovarian Cancer: phase 1b results from the JAVELIN Solid Tumor Trial. JAMA Oncol. 2019;5(3):393–401.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Matulonis UA, Shapira-Frommer R, Santin AD, Lisyanskaya AS, Pignata S, Vergote I, et al. Antitumor activity and safety of pembrolizumab in patients with advanced recurrent ovarian cancer: results from the phase II KEYNOTE-100 study. Ann Oncol. 2019;30(7):1080–7.

    Article  CAS  PubMed  Google Scholar 

  81. Pujade-Lauraine E, Fujiwara K, Ledermann JA, Oza AM, Kristeleit R, Ray-Coquard IL, et al. Avelumab alone or in combination with chemotherapy versus chemotherapy alone in platinum-resistant or platinum-refractory ovarian cancer (JAVELIN ovarian 200): an open-label, three-arm, randomised, phase 3 study. Lancet Oncol. 2021;22(7):1034–46.

    Article  CAS  PubMed  Google Scholar 

  82. Moore KN, Bookman M, Sehouli J, Miller A, Anderson C, Scambia G, et al. Atezolizumab, Bevacizumab, and Chemotherapy for newly diagnosed stage III or IV ovarian Cancer: placebo-controlled Randomized Phase III Trial (IMagyn050/GOG 3015/ENGOT-OV39). J Clin Oncol. 2021;39(17):1842–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Wu Y, Chen W, Xu ZP, Gu W. PD-L1 distribution and perspective for Cancer Immunotherapy-Blockade, Knockdown, or inhibition. Front Immunol. 2019;10:2022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Spranger S, Gajewski TF. Impact of oncogenic pathways on evasion of antitumour immune responses. Nat Rev Cancer. 2018;18(3):139–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Ding L, Kim HJ, Wang Q, Kearns M, Jiang T, Ohlson CE, et al. PARP inhibition elicits STING-Dependent Antitumor immunity in Brca1-Deficient ovarian Cancer. Cell Rep. 2018;25(11):2972–e29802975.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Phukan S, Babu VS, Kannoji A, Hariharan R, Balaji VN. GSK3beta: role in therapeutic landscape and development of modulators. Br J Pharmacol. 2010;160(1):1–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Shiotani B, Zou L. Single-stranded DNA orchestrates an ATM-to-ATR switch at DNA breaks. Mol Cell. 2009;33(5):547–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Sato H, Niimi A, Yasuhara T, Permata TBM, Hagiwara Y, Isono M, et al. DNA double-strand break repair pathway regulates PD-L1 expression in cancer cells. Nat Commun. 2017;8(1):1751.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Higuchi T, Flies DB, Marjon NA, Mantia-Smaldone G, Ronner L, Gimotty PA, et al. CTLA-4 Blockade synergizes therapeutically with PARP inhibition in BRCA1-Deficient ovarian Cancer. Cancer Immunol Res. 2015;3(11):1257–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Owen AM, Fults JB, Patil NK, Hernandez A, Bohannon JK. TLR agonists as mediators of trained immunity: mechanistic insight and immunotherapeutic potential to combat infection. Front Immunol. 2020;11:622614.

    Article  CAS  PubMed  Google Scholar 

  93. Brubaker SW, Bonham KS, Zanoni I, Kagan JC. Innate immune pattern recognition: a cell biological perspective. Annu Rev Immunol. 2015;33:257–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Ren Z, Yu Y, Chen C, Yang D, Ding T, Zhu L, et al. The triangle relationship between long noncoding RNA, RIG-I-like receptor signaling pathway, and Glycolysis. Front Microbiol. 2021;12:807737.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Jiang M, Chen P, Wang L, Li W, Chen B, Liu Y, et al. cGAS-STING, an important pathway in cancer immunotherapy. J Hematol Oncol. 2020;13(1):81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Staniszewska AD, Armenia J, King M, Michaloglou C, Reddy A, Singh M, et al. PARP inhibition is a modulator of anti-tumor immune response in BRCA-deficient tumors. Oncoimmunology. 2022;11(1):2083755.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Pantelidou C, Sonzogni O, De Oliveria Taveira M, 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.

    Article  PubMed  PubMed Central  Google Scholar 

  98. Bakhoum SF, Ngo B, Laughney AM, Cavallo JA, Murphy CJ, Ly P, et al. Chromosomal instability drives metastasis through a cytosolic DNA response. Nature. 2018;553(7689):467–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Shakfa N, Li D, Nersesian S, Wilson-Sanchez J, Koti M. The STING pathway: therapeutic vulnerabilities in ovarian cancer. Br J Cancer. 2022;127(4):603–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Wang Y, Luo J, Alu A, Han X, Wei Y, Wei X. cGAS-STING pathway in cancer biotherapy. Mol Cancer. 2020;19(1):136.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Turinetto M, Scotto G, Tuninetti V, Giannone G, Valabrega G. The role of PARP inhibitors in the Ovarian Cancer Microenvironment: moving Forward from Synthetic lethality. Front Oncol. 2021;11:689829.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Chen Q, Sun L, Chen ZJ. Regulation and function of the cGAS-STING pathway of cytosolic DNA sensing. Nat Immunol. 2016;17(10):1142–9.

    Article  CAS  PubMed  Google Scholar 

  103. Muller E, Speth M, Christopoulos PF, Lunde A, Avdagic A, Oynebraten I, et al. Both type I and type II interferons can activate Antitumor M1 macrophages when combined with TLR stimulation. Front Immunol. 2018;9:2520.

    Article  PubMed  PubMed Central  Google Scholar 

  104. Karimi K, Karimi Y, Chan J, Boudreau JE, Basset J, Chew MV, et al. Type I IFN signaling on dendritic cells is required for NK cell-mediated anti-tumor immunity. Innate Immun. 2015;21(6):626–34.

    Article  CAS  PubMed  Google Scholar 

  105. Strickland KC, Howitt BE, Shukla SA, Rodig S, Ritterhouse LL, Liu JF, et al. Association and prognostic significance of BRCA1/2-mutation status with neoantigen load, number of tumor-infiltrating lymphocytes and expression of PD-1/PD-L1 in high grade serous ovarian cancer. Oncotarget. 2016;7(12):13587–98.

    Article  PubMed  PubMed Central  Google Scholar 

  106. Mouw KW, Goldberg MS, Konstantinopoulos PA, D’Andrea AD. DNA damage and repair biomarkers of Immunotherapy Response. Cancer Discov. 2017;7(7):675–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Maio M, Covre A, Fratta E, Di Giacomo AM, Taverna P, Natali PG, et al. Molecular pathways: at the crossroads of Cancer epigenetics and Immunotherapy. Clin Cancer Res. 2015;21(18):4040–7.

    Article  CAS  PubMed  Google Scholar 

  108. Bopp T, Becker C, Klein M, Klein-Hessling S, Palmetshofer A, Serfling E, et al. Cyclic adenosine monophosphate is a key component of regulatory T cell-mediated suppression. J Exp Med. 2007;204(6):1303–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Domchek SM, Postel-Vinay S, Im SA, Park YH, Delord JP, Italiano A, et al. Olaparib and durvalumab in patients with germline BRCA-mutated metastatic breast cancer (MEDIOLA): an open-label, multicentre, phase 1/2, basket study. Lancet Oncol. 2020;21(9):1155–64.

    Article  CAS  PubMed  Google Scholar 

  110. Drew Y, Kim JW, Penson RT, O’Malley DM, Parkinson C, Roxburgh P, et al. Olaparib plus Durvalumab, with or without Bevacizumab, as treatment in PARP inhibitor-naive platinum-sensitive relapsed ovarian Cancer: a phase II Multi-cohort Study. Clin Cancer Res. 2024;30(1):50–62.

    Article  CAS  PubMed  Google Scholar 

  111. Harter P, Bidziński M, Colombo N, Floquet A, Pérez MJR, Kim JW et al. DUO-O: a randomized phase III trial of durvalumab (durva) in combination with chemotherapy and bevacizumab (bev), followed by maintenance durva, bev and olaparib (olap), in newly diagnosed advanced ovarian cancer patients. J Clin Oncol 2019.

  112. Konstantinopoulos PA, Waggoner S, Vidal GA, Mita M, Moroney JW, Holloway R, et al. Single-arm phases 1 and 2 trial of Niraparib in Combination with Pembrolizumab in patients with recurrent platinum-resistant ovarian carcinoma. JAMA Oncol. 2019;5(8):1141–9.

    Article  PubMed  PubMed Central  Google Scholar 

  113. Randall LM, O’Malley DM, Monk BJ, Coleman RL, Gaillard S, Adams S, et al. Niraparib and Dostarlimab for the treatment of recurrent platinum-resistant ovarian cancer: results of a phase II study (MOONSTONE/GOG-3032). Gynecol Oncol. 2023;178:161–9.

    Article  CAS  PubMed  Google Scholar 

  114. Jin N, Xia Y, Gao Q. Combined PARP inhibitors and small molecular inhibitors in solid tumor treatment (review). Int J Oncol 2023, 62(2).

  115. Ngoi NYL, Leo E, O’Connor MJ, Yap TA. Development of next-generation poly(ADP-Ribose) polymerase 1-Selective inhibitors. Cancer J. 2021;27(6):521–8.

    Article  CAS  PubMed  Google Scholar 

  116. Noordermeer SM, van Attikum HPARP, Inhibitor Resistance. A tug-of-war in BRCA-Mutated cells. Trends Cell Biol. 2019;29(10):820–34.

    Article  CAS  PubMed  Google Scholar 

  117. Chiappa M, Guffanti F, Bertoni F, Colombo I, Damia G. Overcoming PARPi resistance: preclinical and clinical evidence in ovarian cancer. Drug Resist Updat. 2021;55:100744.

    Article  CAS  PubMed  Google Scholar 

  118. Li H, Liu ZY, Wu N, Chen YC, Cheng Q, Wang J. PARP inhibitor resistance: the underlying mechanisms and clinical implications. Mol Cancer. 2020;19(1):107.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. 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.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Buttarelli M, Ciucci A, Palluzzi F, Raspaglio G, Marchetti C, Perrone E, et al. Identification of a novel gene signature predicting response to first-line chemotherapy in BRCA wild-type high-grade serous ovarian cancer patients. J Exp Clin Cancer Res. 2022;41(1):50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Mukhopadhyay S, Vander Heiden MG, McCormick F. The Metabolic Landscape of RAS-Driven cancers from biology to therapy. Nat Cancer. 2021;2(3):271–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Therachiyil L, Anand A, Azmi A, Bhat A, Korashy HM, Uddin S. Role of RAS signaling in ovarian cancer. F1000Res. 2022;11:1253.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Nakayama N, Nakayama K, Yeasmin S, Ishibashi M, Katagiri A, Iida K, et al. KRAS or BRAF mutation status is a useful predictor of sensitivity to MEK inhibition in ovarian cancer. Br J Cancer. 2008;99(12):2020–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Yang B, Li X, Fu Y, Guo E, Ye Y, Li F, et al. MEK Inhibition remodels the Immune Landscape of Mutant KRAS Tumors to overcome resistance to PARP and Immune Checkpoint inhibitors. Cancer Res. 2021;81(10):2714–29.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Mukhopadhyay S, Saqcena M, Foster DA. Synthetic lethality in KRas-driven cancer cells created by glutamine deprivation. Oncoscience. 2015;2(10):807–8.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Saqcena M, Menon D, Patel D, Mukhopadhyay S, Chow V, Foster DA. Amino acids and mTOR mediate distinct metabolic checkpoints in mammalian G1 cell cycle. PLoS ONE. 2013;8(8):e74157.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Gaglio D, Soldati C, Vanoni M, Alberghina L, Chiaradonna F. Glutamine deprivation induces abortive s-phase rescued by deoxyribonucleotides in k-ras transformed fibroblasts. PLoS ONE. 2009;4(3):e4715.

    Article  PubMed  PubMed Central  Google Scholar 

  128. Hudson CD, Savadelis A, Nagaraj AB, Joseph P, Avril S, DiFeo A, et al. Altered glutamine metabolism in platinum resistant ovarian cancer. Oncotarget. 2016;7(27):41637–49.

    Article  PubMed  PubMed Central  Google Scholar 

  129. Shen YA, Hong J, Asaka R, Asaka S, Hsu FC, Suryo Rahmanto Y, et al. Inhibition of the MYC-Regulated Glutaminase Metabolic Axis is an effective Synthetic Lethal Approach for Treating Chemoresistant ovarian cancers. Cancer Res. 2020;80(20):4514–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

This study was supported by the Research Team for Reproduction Health and Translational Medicine of Hunan Normal University (2023JC101), Key Project of Developmental Biology and Breeding from Hunan Province (2022XKQ0205), National Natural Science Foundation of China (82003050 and 81874193), Hunan Provincial Natural Science Foundation of China (2023JJ30375, 2023JJ40415 and 2024JJ5285), Scientific Research Project of Hunan Provincial Health Commission Changsha (B2023047708)and the Hunan Provincial Science and Technology Department(2023ZJ1120).

Author information

Authors and Affiliations

Authors

Contributions

Fen Xiao: Investigation, Visualization, Writing original draft. ZhiBin Wang: Writing & editing. Liu Qiao: Writing & editing. Xiu Zhang: Writing & editing. Nayiyuan Wu: Conceptualization, Writing & editing. Jing Wang: Supervision, Writing & editing. Xing Yu: Supervision, Writing & editing.

Corresponding authors

Correspondence to NaYiYuan Wu, Jing Wang or Xing Yu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xiao, F., Wang, Z., Qiao, L. et al. Application of PARP inhibitors combined with immune checkpoint inhibitors in ovarian cancer. J Transl Med 22, 778 (2024). https://doi.org/10.1186/s12967-024-05583-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-024-05583-z

Keywords