Open Access

PARP inhibitors and epithelial ovarian cancer: Molecular mechanisms, clinical development and future prospective (Review)

  • Authors:
    • Vera Loizzi
    • Girolamo Ranieri
    • Mariarita Laforgia
    • Cosmo   Damiano Gadaleta
    • Giulio Gargano
    • Anila Kardhashi
    • Maria De Liso
    • Emanuele Naglieri
    • Vittoria Del Vecchio
    • Ettore Cicinelli
    • Gennaro Cormio
  • View Affiliations

  • Published online on: August 6, 2020     https://doi.org/10.3892/ol.2020.11951
  • Article Number: 90
  • Copyright: © Loizzi et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Epithelial ovarian cancer (EOC) has a poor prognosis. Since the introduction of paclitaxel as antineoplastic agent >20 years ago, only a few phase III randomized trials have shown challenging data regarding different therapeutic options for facing its aggressive clinical course and granting active therapies to patients. Different studies have shown the utility of poly(ADP‑ribose) polymerase (PARP) inhibitors in women with EOC with or without BRCA mutations, both germline and somatic. Three PARP inhibitors, olaparib, rucaparib and niraparib, have been recently approved by the Food and Drug Administration for clinical use in EOC patients, though with different clinical indications and profiles of toxicity, while two other molecules, veliparib and talazoparib, are still under clinical investigation. The aim of the present paper is to evaluate the current status of PARP inhibitors in terms of molecular activity, pharmacodynamic properties and clinical applications.

Introduction

Since the introduction of paclitaxel >20 years ago for the treatment of epithelial ovarian cancer (EOC), only a few phase III trials testing other therapeutic agents have demonstrated notable data in terms of clinical outcome (1). Two studies demonstrated an increase in progression-free survival (PFS) time and, in a selected subgroup of patients, overall survival (OS) time (2) or only in PFS time (3), with the introduction of the anti-vascular endothelial growth factor (VEGF) monoclonal antibody, bevacizumab, using a therapeutic schedule based on paclitaxel (25).

More recently, poly(ADP-ribose) polymerase (PARP) inhibitors have begun to be used as a new therapeutic approach in the management of EOC (6), particularly for patients with assessed defects in the homologous recombination (HR) DNA repair process, which is strictly linked to BRCA1/2 gene mutations (7). Considering the few available therapeutic options for EOC treatment, this important discovery has addressed scientific research into novel strategies exploiting DNA repair deficiencies and PARP inhibitors are the first drugs with this peculiar mechanism of action and are active in patients with recurrent EOC with HR deficiencies (6,7). In spite of this potentially revolutionary evidence, these molecules have been demonstrated to be also active in patients without HR deficiencies (8). Three PARP inhibitors, olaparib, rucaparib and niraparib, are commercially available and approved by the Food and Drug Administration (FDA) for the treatment of patients with recurrent EOC, with different clinical indications and toxicity profiles. In addition, two other molecules, veliparib and talazoparib, are still under clinical investigation. In the literature, to the best of our knowledge, no comparisons among the three commercial drugs have been made so far; however, ongoing trials are now focusing their attention on new clinical indications and on additional therapeutic strategies in combination with conventional antineoplastic drugs.

The aim of the present review was to discuss the current status of PARP inhibitors in terms of the mechanisms of action, molecular activity and clinical applications, as well as to evaluate their future prospective in oncological therapy.

BRCA mutations and cancer risk

Previous studies have demonstrated the association between germline mutations of BRCA1 and BRCA2 genes and the early development of both breast and ovarian cancer (9), as well as other neoplasms caused by either germline or somatic mutations (10).

The techniques used for the detection of BRCA gene mutations depend on DNA sequencing procedures, which are, however, susceptible to yielding false-positive results. In fact, these genes can also be affected by certain benign non-pathogenic variations, termed variants of unknown significance, which represent ~13% of BRCA1 and BRCA2 mutations, suggesting clinical uncertainty and ambiguity in the risk assessment of patients undergoing the analysis (11,12). As a consequence, different polymorphisms of these genes complicate the identification of BRCA mutations.

Breast cancer-related to BRCA1 mutation is more likely estrogen receptor (ER)-negative when compared with BRCA2 and non-BRCA1 tumors (13). This evidence is substantial as estrogens influence certain genes controlling growth regulation; therefore, both breast and ovarian cancer are assessed for ER status to predict prognosis, future treatment or preventive and curative measures in both BRCA and non-BRCA tumors. The failure of BRCA function and estrogen signaling, together with other subcellular mechanisms, causes tumor growth due to the lack of appropriate DNA surveillance. Silencing the BRCA1 gene leads to increased expression of the gene codifying the aromatase enzyme that is responsible for the conversion of steroids into active estrogens, promoting their synthesis and biological activity (14).

Molecular mechanisms of PARP enzymes

Tumor genomic instability results in DNA aberrations consisting of point mutations, tandem duplications and translocations, which induce carcinogenesis and tumor progression (15,16). The integrity of chromosomal structure, transcription, replication, recombination and DNA repair are under the control of a pool of 17 enzymes, constituting the PARP family of proteins (17,18) (Fig. 1).

Human cells have at least five primary mechanisms of DNA repair (19), such as mismatch repair (MMR), base excision repair (BER), nucleotide excision repair (NER) and double-strand break (DSB) recombination repair, including both non-homologous end-joining (NHEJ) and homologous recombination repair. The dysfunction, reduction or absence of proteins involved in these pathways may lead to dangerous cellular implications, determining mutagenesis and toxicity (20).

Different insults can affect DNA, although single alterations are the most recurrent and are repaired by a combination of BER, NER and MMR pathways using the undamaged DNA strand as a template. The predominant mechanism of single-strand break (SSB) repair is BER with the activity of PARP enzymes (21).

PARP-1 and PARP-2 are activated by DNA damage. In particular, PARP-1 functions as a molecular sensor binding the N-terminal zinc finger domains to DNA SSBs and subsequently, by increasing its activity, catalyzes the transfer of ADP-ribose (poly ADP-ribosylation) to target proteins through their C-terminal catalytic domain. Following the activation of the nicotinamide-adenine-dinucleotide (NAD+), PARPs form PAR polymer chains that play an essential role for recruiting intermediates for the DNA repair pathway (20) (Fig. 2). PARP-1 covalently attaches PAR chains to several different proteins, in the process known as PARylation. Due to its role in DNA repair, PARP inhibition results in genomic instability and the accumulation of damaged cells in cell cycle arrest (2226).

If PARP activity is lacking, more deleterious DSBs can multiply, beginning from damaged SSBs, which require other different pathways for repair (19).

Biological link between PARP and angiogenesis inhibition

Ample experimental data have indicated a link between PARP enzymes and angiogenesis. Angiogenesis is an important driver of EOC development and progression and it is a main target of antitumor therapy (27). In this regard, since 2011, anti-VEGF therapy with bevacizumab combined with paclitaxel and carboplatin has been the backbone of treatment with monoclonal antibodies in patients with locally advanced and metastatic EOC (FIGO classification stage III and IV) (28).

The PARP-1 pathway is able to regulate gene expression, controlling angiogenesis through hypoxia-inducible factor-1α (HIF-1α) (29). Experimentally, PARP-deficient mice have been shown to exhibit a decreased level of HIF-1α. This transcription factor plays a major role in stimulating tumor angiogenesis and is a subunit of the heterodimer HIF-1 together with HIF-1β. HIF-1β is a nuclear constitutively expressed protein that does not undergo regulation by the oxygen level (30); by contrast, HIF-1α is a cytoplasmic protein whose activation is dependent on the oxygen concentration. In particular, in oxygenated microenvironmental conditions, HIF-1α undergoes hydroxylation by prolyl hydroxylases on its prolyl residues in the oxygen-dependent degradation domain and this event leads to its binding to von Hippel-Lindau protein, before being degraded in the ubiquitin-proteasome pathway. Meanwhile, at low oxygen tension, prolyl hydroxylase is inactive, resulting in HIF-1α stabilization, which allows its migration to the nucleus, where it binds HIF-1β, finally forming the HIF-1 complex (31). The HIF-1 complex targets a consensus hypoxia response element in the promoter of several pro-angiogenic genes, in particular VEGF, activating their transcription (32).

From a biological point of view, in vivo and in vitro data have suggested that angiogenesis and tumorigenesis in EOC is promoted by PARP-1 overexpression and due to the increasing level of VEGF-A, PARP-1 can be considered a potential therapeutic target (33).

Experimental data employing reverse transcription- quantitative polymerase chain reaction demonstrated that SKOV3 human ovarian cancer cells transfected with PARP-1 small interfering RNA express lower levels of VEGF-A mRNA compared with SKOV3 cell cultures transfected with negative control-small interfering-RNA (26). Moreover, the knockdown of PARP-1 was shown to decrease VEGF-A levels in SKOV3 cells, as demonstrated by western blot analysis. These results were confirmed by ELISA, revealing the presence of VEGF-A in the supernatant of SKOV3 cell cultures transfected with negative control-small interfering RNA.

Notably, in addition to these data, the PARP inhibitor, N- (6-Oxo-5,6-dihydro-phenanthridin- 2-yl)- N,N-dimethylacetamide (PJ-34), has been demonstrated to be endowed with anti-angiogenic activity by the in vitro inhibition of growth and migration of human umbilical vein endothelial cells (34), probably due to the reduction of nitric oxide, guanylylcyclase and the cGMP pathway, which represent the drivers of the VEGF effect on endothelial cells. Evidently, the effects of VEGF are also mediated by the binding to VEGF receptors that, in turn, activates the intracellular pathways of Akt, ERK1/2 and p38 MAP kinase. In vitro studies have demonstrated that PJ-34 inhibits the phosphorylation of these kinases, suggesting that in the VEGF response in endothelial cells, PARP exerts a key role. Similarly, the PARP inhibitor, GPI 15427, has been found to exert anti-angiogenic effects in PARP-1-knockout mice (35).

Clinical application of PARP inhibitors in BRCA mutations, mechanisms of activity and resistance

PARP inhibitors prevent the repair of persistent SSBs and the reconstitution of DSBs, through the replication fork. PARP inhibitors have been developed for targeting cancer related to BRCA1 or BRCA2 gene mutations, these genes being responsible for the synthesis of proteins involved in the HR repair pathway. Individuals with the wild-type phenotype have both functioning copies of the BRCA genes; by contrast, patients carrying a BRCA mutation have only one functioning copy, which allows the correct process of DNA repair and viability. When a mutation occurs in the only functioning gene copy, cells lose their mitotic control, become susceptible to tumor growing and are unable to undertake HR (36).

When tumor cells carry both abnormal copies of the gene, being homozygous for BRCA1 or BRCA2 mutations, cells undergo inadequate DNA repair and become sensitive to PARP inhibitors. In the presence of an oncogene mutation, targeted drugs or gene therapy should theoretically induce synthetic lethality in neoplastic cells; in other words, synthetic lethality occurs when two cellular events occur independently, but permit cell survival, whereas in combination, they result in cell death (3739).

All PARP inhibitors can inhibit both PARP-1 and PARP-2 by suppressing PARP catalytic activity, avoiding the formation of PAR polymers and, consequently, blocking the binding of NAD+ at the site of DNA damage. These effects compromise the cellular ability to overcome DNA damage (20). Recent studies have investigated possible biomarkers of the response to PARP inhibition by measuring both the biosynthesis of RAD51 foci and PAR poly(ADP-ribose) and 53BP1 expression levels in cancer cells (37). However, efforts to identify an efficient and more specific biomarker are imperative in order to optimize clinical outcomes to PARP inhibitor treatment.

PARP inhibitors share similar toxic effects with other chemotherapeutic agents, such as nausea, fatigue, vomiting, anemia and abdominal pain, which are the most frequently reported adverse effects (40).

Recent evidence has shed light on an acquired resistance to PARP inhibitors developed by neoplastic cells. A number of mechanisms of pharmacological resistance have been proposed; these include, in particular, the ability of tumor cells to reverse the mutation in the BRCA gene, which restores HR function (41). Other possible mechanisms of resistance to PARP inhibitors can be either a decrease in NHEJ, the reduction of PARP-1 enzymatic activity or the increased activity of RAD51, an essential protein involved in HR function (37). In the light of all considerations and hypotheses, the resistance mechanisms to PARP inhibitors are important aspects to ascertain knowledge for, in order to forecast the efficacy of treatments.

Clinical trials with PARP inhibitors in EOC

DNA repair processes induced by both BRCA and PARP pathways are considered key elements for tumor aggressiveness. If PARP is inhibited by a molecule that modifies its function, cells genetically deficient in BRCA die, and this occurs in EOC with BRCA mutations (42).

Olaparib

Olaparib was the first PARP inhibitor to be used in clinical practice for the treatment of EOC; it acts as a single therapeutic agent, with a 30–50% response rate when used in second-line or subsequent line treatments in patients carrying BRCA1-2 mutations (40,4347). The clinical activity of the drug is greater in the presence of platinum-sensitive tumors, although platinum-resistant cancer also responds to therapy. In addition, olaparib has been tested in patients with the wild-type phenotype affected by high-grade serous EOC, although the response rates have not been satisfactory in this group of patients.

A double-blind randomized controlled phase II trial enrolled patients with high-grade OC who were pre-treated with a platinum-based second-line chemotherapeutic regimen, resulting in a complete response (CR) or partial response (PR), to receive either olaparib or a placebo as the maintenance therapy (48). The study demonstrated a better outcome with regard to PFS time (median, 8.4 months vs. 4.8 months; hazard ratio, 0.35; P<0.001) in patients treated with olaparib compared with placebo. Furthermore, patients with a documented germline BRCA mutation experienced a tripling of the time to disease progression (median, 11.2 months vs. 4.3 months; hazard ratio, 0.18; P<0.0001) (49).

The long-term follow-up of patients enrolled in the aforementioned study (48) demonstrated a favorable impact on OS time due to the maintenance strategy. A similar outcome was recorded in a subsequent randomized phase III trial testing olaparib vs. placebo in patients with platinum-sensitive recurrent EOC, with a CR or PR after ≥2 lines of platinum-based chemotherapy (50). The overall median PFS time was 19.1 months for active maintenance therapy vs. 5.5 months for placebo, respectively (hazard ratio, 0.30; P<0.0001).

Apart from its use in the maintenance approach, olaparib has recently been approved by the FDA in monotherapy for women who carry germline BRCA mutations with a diagnosis of EOC and who have received a minimum of three prior lines of cytotoxic chemotherapy (51). Promising data from more recent studies might extend the clinical indications of the drug within the near future.

Rucaparib

Rucaparib is a PARP inhibitor approved by the FDA as a single-agent treatment in patients with EOC who carry either a germline or a somatic BRCA mutation and who have received pre-treatment with a minimum of two prior lines of chemotherapy. A phase II trial demonstrated an objective response rate (ORR) of ~80% in patients with a BRCA mutation (52).

By contrast, in patients with the wild-type phenotype with a high loss of heterozygosity, the ORR was reduced to 44%, while in treated patients with a low loss of heterozygosity, only 20% experienced a response (52). These data suggest that defects in DNA repair mechanisms can be considered appropriate targets for therapy with a PARP inhibitor.

A recent randomized phase III trial evaluating the effects of maintenance therapy with rucaparib following second-line chemotherapy demonstrated a significantly improved PFS time (53), highlighting the candidacy of this clinical approach to be once more a new standard of care for patients with platinum-sensitive EOC.

Niraparib

Niraparib is the third PARP inhibitor available for clinical use in the USA; it has been approved on the basis of the results of a randomized phase III trial testing the drug compared with a placebo, for maintenance therapy in patients who obtained a CR or PR after second-line platinum-based chemotherapy (54). In women affected by EOC who carry a BRCA mutation, the median PFS time of 21 months was higher than the 5.5 months found for patients treated with placebo (hazard ratio, 0.27). Favorable results have been observed even in the overall non-germline BRCA patient population, with a median PFS time of 9.3 months compared with 3.9 months (hazard ratio, 0.45) for placebo (54).

Moreover, within the same niraparib trial (54), an attempt was made to identify a biomarker to recognize patients whose tumors could be particularly susceptible to treatment, despite the absence of a germline BRCA mutation. In patients with the wild-type phenotype with tumors characterized by homologous recombinant deficiency (HRD), maintenance therapy with niraparib exhibited a statistically significant 12.9-month median PFS time compared with the 3.8 months (hazard ratio, 0.38) found for placebo (54). From this objective evidence, the FDA approved niraparib for clinical use as a second-line maintenance strategy, following a response to platinum-based treatment without requiring the use of a molecular biomarker, either BRCA mutation or HRD-positive (Table I).

Table I.

Studies of PARP-inhibitors in epithelial ovarian cancer.

Table I.

Studies of PARP-inhibitors in epithelial ovarian cancer.

First author, yearPatient PopulationTherapies and dosesOutcomes(Refs.)
Audeh et al, 2010Two cohorts of women (aged ≥18 years) with confirmed genetic BRCA1 or BRCA2 mutations and recurrent, measurable disease.First cohort (n=33): Continuous oral olaparib at the maximum tolerated dose of 400 mg twice daily. Second cohort (n=24): Continuous oral olaparib at 100 mg twice daily.ORR was 11 (33%) out of 33 patients (95% CI, 20–51) in the cohort assigned 400 mg olaparib twice daily and 3 (13%) out of 24 (431) in the cohort assigned 100 mg twice daily.(45)
Gelmon et al, 2011Women with advanced high-grade serous and/or undifferentiated ovarian carcinoma or triple-negative breastancer were cstratified according to whether they had a BRCA1 or BRCA2 mutation or not. A total of 91 patients were enrolled (65 with ovarian cancer and 26 breast cancer).Olaparib at 400 mg twice daily.In the ovarian cancer cohorts, confirmed objective responses were seen in 7 (41%; 95% CI, 22–64) out of 17 patients with BRCA1 or BRCA2 mutations and 11 (24%; 95% CI, 14–38) out of 46 patients without mutations. No confirmed objective responses were reported in patients with breast cancer.(46)
Swisher et al, 2017A total of 204 patients with recurrent, platinum-sensitive, high-grade ovarian carcinoma were classified into one of three predefined homologous recombination deficiency subgroups on the basis of tumor mutational analysis: BRCA mutant (deleterious germline or somatic), BRCA wild-type and LOH high (LOH high group), or BRCA wild-type and LOH low (LOH low group).Rucaparib at 600 mg twice daily for continuous 28 day cycles.Progression-free survival was significantly longer in the BRCA mutant (hazard ratio, 0.27; 95% CI,0.16-0.44; P<0.0001) and LOH high (hazard ratio, 0.62; 95% CI, 0.42-0.90; P=0·011) subgroups compared with that in the LOH low subgroup.(52)
Mirza et al, 2016A total of 553 patients were enrolled and categorized according to the presence or absence of a germline BRCA mutation (gBRCA cohort andnon-gBRCA cohort). Of these, 203 were in the gBRCA cohort (with 138assigned niraparib and 65 placebo), and 350 patients were in the non-gBRCA cohort (with 234 assigned niraparib and 116 placebo).Niraparib at 300 mg or placebo once daily.Patients in the niraparib group had significantly longer median PFS times than those in the placebo group, namely 21.0 months vs. 5.5 in the gBRCA cohort (hazard ratio, 0.27; 95% confidence interval [CI], 0.17 to 0.41), as compared with 12.9 months vs. 3.8 months in the non-gBRCA cohort for patients who had tumors with homologous recombination deficiency (hazard ratio, 0.38; 95% CI, 0.24 to 0.59) and 9.3 months vs. 3.9 months in the overall non-gBRCA cohort (hazard ratio, 0.45; 95% CI, 0.34 to 0.61; P<0.001 for all three comparisons).(54)

[i] ORR, objective response rate; LOH, loss of heterozygosity.

Veliparib

Veliparib is a PARP inhibitor that is still under investigation. A phase II study evaluated the effects of the use of oral veliparib at 400 mg twice daily in 50 patients who underwent a maximum of three prior chemotherapy regimens, with measurable disease and who had never benefitted from another previous treatment with a PARP inhibitor. The response rate was 26%, with a median PFS time of 8.18 months (55). Another phase I/II study revealed a 65% overall response rate in platinum-resistant or partially platinum-sensitive patients with a relapse of EOC carrying a germline BRCA mutation treated with maintenance oral veliparib 300 mg twice daily (56). These preliminary data gave raise to other ongoing phase III clinical trials testing veliparib not only in ovarian cancer, but also in non-small cell lung and triple-negative breast cancer.

Talazoparib

Talazoparib is a new PARP inhibitor in clinical development for patients with advanced or recurrent solid tumors. In vitro studies have demonstrated that talazoparib exhibits potent activity in tumor cells with BRCA or PTEN mutations compared to other PARP inhibitors (57). In a multicenter phase I study, among patients with BRCA-mutated ovarian cancer, talazoparib exhibited a response in 5 out of 12 patients (42%), with a median PFS time of 36.4 weeks (58).

Selection of PARP inhibitors in EOC

To date, to the best of our knowledge, no direct trial comparing the three commercially available PARP inhibitors has been performed; therefore, a summary report about the relative efficacy or toxicity of each drug is not immediately being attempted. Thanks to the results of the discussed clinical trials and despite the different adverse events (AEs) they induce, all agents have obtained regulatory approval for use in different clinical settings. The large majority of patients enrolled in non-randomized and randomized trials for all drugs have continued treatment (if permitted by the protocol), despite recognized toxicity, often with appropriate dose modifications and treatment interruptions to permit recovery from the AEs (40,4354).

All currently available PARP inhibitors have some common side effects, in particular low-grade nausea, fatigue and myelosuppression, which can compromise the quality of life of patients, despite no evidenced cancer-related symptoms (59).

The selection of the right PARP inhibitor remains a challenge. At the current time, olaparib is the only PARP inhibitor approved for first-line therapy owing to the results of the SOLO1 clinical trial. By contrast, for second-line treatment the matter is still under debate and the selection could be based on certain specific differences in toxicity. Niraparib exerts a more potent effect on platelet counts (54), while rucaparib induces an increase in creatinine and transaminases (52,53), both of which are essentially false-positives, as they are not associated with real kidney toxicity or liver toxicity. The cause of this peculiar effect remains under investigation, as it seems to be associated with the interaction of the PARP inhibitor with certain transport proteins, which is responsible for the difficulty in monitoring the clinical effect of the drug in patients with kidney and liver comorbidities (6062).

More frequent toxicities of PARP inhibitors can be grouped as follows: i) Hematological: Anemia, thrombocytopenia and neutropenia (55,63,64). These outcomes provoke different degrees of bone marrow suppression, depending upon the dose, which must be prescribed only after blood counts have been taken. It is considered best practice to follow-up new patients with lower blood counts at a weekly frequency, particularly in the case of bone marrow suppression due to previous chemotherapeutic regimens. ii) Gastrointestinal: Nausea, vomiting, diarrhea, constipation, difficulty in eating and anorexia (65), which tend to reduce in severity with time. iii) Other: Fatigue.

In the SOLO-1/2 trial, 3 patients in the olaparib group developed acute myeloid leukemia (AML) or myelodysplastic syndrome (MDS). Therefore, patients who begin therapy with a PARP inhibitor must be advised about this 3% risk of developing AML or MDS (50).

Niraparib can cause hypertension, tachycardia and headaches due to its interaction with the norepinephrine dopamine carriers; therefore, patients affected by hypertension have to monitor their blood pressure regularly (66).

Rucaparib can cause a benign elevation both in liver function tests and in serum cholesterol during the first few weeks of treatment (60).

Future development of PARP inhibitors in EOC

Ongoing trials are questioning the new possible clinical applications of PARP inhibitors in EOC, as first-line maintenance therapy or in combination with chemotherapy, but also the potential cross-resistance among all PARP inhibitors. In fact, the pharmacological strategy based on PARP-after-PARP could be considered in women who have failed to respond to one PARP agent or who have progressed after the initial response.

Maintenance therapy plays a central role in the clinical use of PARP agents, both as a first-line and a second-line response to platinum-based chemotherapy and as third-line maintenance.

Other clinically relevant questions to elaborate on in the treatment of EOC with PARP inhibitors are the possible combination therapies with standard platinum-based cytotoxic therapy, bevacizumab, a checkpoint inhibitor and a topoisomerase I inhibitor, such as topotecan.

The last cited strategy is based on the activity of topoisomerase I to bind DNA during its replication or repair, tempering SSBs and diminishing the associated distortional tension (67). Topotecan is still approved for EOC therapy due to its ability to induce de-stabilization of the replication forks promoting DNA lesions (68,69). PARP-1 is activated by topotecan and induces DNA lesion-reducing DNA breakage (69). In the light of these considerations, topoisomerase I inhibition with topotecan in combination with PARP inhibition could lead to a magnification and strengthening of the anti-EOC response. This therapeutic strategy has been explored in pre-clinical studies and is now under clinical investigation, as in vitro anti-tumor effects have been shown to be highly potentiated (7073).

Another clinical approach involving PARP inhibitors is represented by their potential ability to radiosensitize EOC (74). In a recently published study, BRCA1-deficient high-grade ovarian cancer cells were shown to be more sensitive to radiotherapy alone after olaparib-mediated radiosensitization, compared with BRCA1-proficient cells. Furthermore, when used in association with radiotherapy, olaparib inhibited DNA damage repair and PARP-1 activity, increased apoptosis and increased OS.

All these notable and potentially revolutionary clinical applications of PARP inhibitors must be considered alongside the potential associated toxicities, first of which is the onset of MDS and AML; this toxicity seems to be due to prior DNA damage caused by cytotoxic chemotherapy. The overall risk of MDS and AML after PARP inhibitors is <3%, with a number of patients having received >5 years of continuous PARP inhibitor therapy without onset. In the future, longer treatments in a larger population of patients will be required; therefore, the incidence of these serious events must be carefully evaluated.

Finally, PARP inhibitors represent a new important weapon against EOC, which is known to be associated with a poor prognosis, with a few therapeutic options. The potential clinical efficacy of PARP inhibitors lies not only in their peculiar mechanisms of action, but also in the number of clinical approaches they are involved with. The near future may provide the answers to all questions related to PARP inhibitors in this context.

Acknowledgements

The authors would like to thank Mrs Daniela Simone (IRCCS Istituto di Ricovero e Cura a Carattere Scientifico ‘Giovanni Paolo II’, Bari, Italy) for her assistance in the purchase of the manuscripts useful for the elaboration of this review.

Funding

No funding was received.

Availability of data and materials

Not applicable.

Authors' contributions

GR and GC conceived and designed the study of comparison among PARP inhibitors. VL, AK, MDL, VDV and EN performed the literature review, selecting information and clinical trials. VL wrote the original draft of the manuscript and ML revised English language and syntaxes. EN, ML, CDG, GG, EC and GR critically revised the manuscript for important intellectual content in terms of clinical trial results, adverse reactions and future perspectives in therapy. GC, GG, EN, CDG and GR supervised the study. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Ozols RF, Bundy BN, Greer BE, Fowler JM, Clarke-Pearson D, Burger RA, Mannel RS, DeGeest K, Hartenbach EM and Baergen R; Gynecologic Oncology Group, : Phase III trial of carboplatin and paclitaxel compared with cisplatin and paclitaxel in patients with optimally resected stage III ovarian cancer: A gynecologic oncology group study. J Clin Oncol. 21:3194–3200. 2003. View Article : Google Scholar : PubMed/NCBI

2 

Perren TJ, Swart AM, Pfisterer J, Ledermann JA, Pujade- Lauraine E, Kristensen G, Carey MS, Beale P, Cervantes A, Kurzeder C, et al: A phase 3 trial of bevacizumab in ovarian cancer. N Engl J Med. 365:2484–2496. 2011. View Article : Google Scholar : PubMed/NCBI

3 

Burger RA, Brady MF, Bookman MA, Fleming GF, Monk BJ, Huang H, Mannel RS, Homesley HD, Fowler J, Greer BE, et al: Incorporation of bevacizumab in the primary treatment of ovarian cancer. N Engl J Med. 365:2473–2483. 2011. View Article : Google Scholar : PubMed/NCBI

4 

Ranieri G, Ferrari C, Di Palo A, Marech I, Porcelli M, Falagario G, Ritrovato F, Ramunni L, Fanelli M, Rubini G and Gadaleta CD: Bevacizumab-based chemotherapy combined with regional deep capacitive hyperthermia in metastatic cancer patients: A pilot study. Int J Mol Sci. 18:14582017. View Article : Google Scholar

5 

Ranieri G, Patruno R, Ruggieri E, Montemurro S, Valerio P and Ribatti D: Vascular endothelial growth factor (VEGF) as a target of bevacizumab in cancer: From the biology to the clinic. Curr Med Chem. 13:1845–1857. 2006. View Article : Google Scholar : PubMed/NCBI

6 

Gadducci A and Guerrieri ME: PARP inhibitors alone and in combination with other biological agents in homologous recombination deficient epithelial ovarian cancer: From the basic research to the clinic. Crit Rev Oncol Hematol. 114:153–165. 2017. View Article : Google Scholar : PubMed/NCBI

7 

Sunada S, Nakanishi A and Miki Y: Crosstalk of DNA double-strand break repair pathways in poly(ADP-ribose) polymerase inhibitor treatment of breast cancer susceptibility gene 1/2-mutated cancer. Cancer Sci. 109:893–899. 2018. View Article : Google Scholar : PubMed/NCBI

8 

Bitler BG, Watson ZL, Wheeler LJ and Behbakht K: Behbakht K: PARP inhibitors: Clinical utility and possibilities of overcoming resistance. Gynecol Oncol. 147:695–704. 2017. View Article : Google Scholar : PubMed/NCBI

9 

Alsop K, Fereday S, Meldrum C, deFazio A, Emmanuel C, George J, Dobrovic A, Birrer MJ, Webb PM, Stewart C, et al: BRCA mutation frequency and patterns of treatment response in BRCA mutation-positive women with ovarian cancer: A report from the Australian ovarian cancer study group. J Clin Oncol. 30:2654–2663. 2012. View Article : Google Scholar : PubMed/NCBI

10 

Hennessy BT, Timms KM, Carey MS, Gutin A, Meyer LA, Flake DD II, Abkevich V, Potter J, Pruss D, Glenn P, et al: Somatic mutations in BRCA1 and BRCA2 could expand the number of patients that benefit from poly (ADP ribose) polymerase inhibitors in ovarian cancer. J Clin Oncol. 28:3570–3576. 2010. View Article : Google Scholar : PubMed/NCBI

11 

Richter S, Haroun I, Graham TC, Eisen A, Kiss A and Warner E: Variants of unknown significance in BRCA testing: Impact on risk perception, worry, prevention and counseling. Ann Oncol. 24 (Suppl 8):viii69–viii74. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Frank TS, Deffenbaugh AM, Reid JE, Hulick M, Ward BE, Lingenfelter B, Gumpper KL, Scholl T, Tavtigian SV, Pruss DR and Critchfield GC: Clinical characteristics of individuals with germline mutations in BRCA1 and BRCA2: Analysis of 10,000 individuals. J Clin Oncol. 20:1480–1490. 2002. View Article : Google Scholar : PubMed/NCBI

13 

Foulkes WD, Metcalfe K, Sun P, Hanna WM, Lynch HT, Ghadirian P, Tung N, Olopade OI, Weber BL, McLennan J, et al: Estrogen receptor status in BRCA1- and BRCA2-related breast cancer: The influence of age, grade, and histological type. Clin Cancer Res. 10:2029–2034. 2004. View Article : Google Scholar : PubMed/NCBI

14 

Bhattacharjee S and Nandi S: Rare genetic diseases with defects in DNA repair: Opportunities and challenges in orphan drug development for targeted cancer therapy. Cancers (Basel). 10:2982018. View Article : Google Scholar

15 

Andor N, Maley CC and Ji HP: Genomic instability in cancer: Teetering on the limit of tolerance. Cancer Res. 77:2179–2185. 2017. View Article : Google Scholar : PubMed/NCBI

16 

O'Connor MJ: Targeting the DNA damage response in cancer. Mol Cell. 60:547–560. 2015. View Article : Google Scholar : PubMed/NCBI

17 

Langelier MF, Eisemann T, Riccio AA and Pascal JM: PARP family enzymes: Regulation and catalysis of the poly(ADP-ribose) posttranslational modification. Curr Opin Struct Biol. 53:187–198. 2018. View Article : Google Scholar : PubMed/NCBI

18 

Grimaldi G and Corda D: ADP-ribosylation and intracellular traffic: An emerging role for PARP enzymes. Biochem Soc Trans. 47:357–370. 2019. View Article : Google Scholar : PubMed/NCBI

19 

Ricks TK, Chiu HJ, Ison G, Kim G, McKee AE, Kluetz P and Pazdur R: Successes and challenges of PAPR inhibitors in cancer therapy. Front Oncol. 5:2222015. View Article : Google Scholar : PubMed/NCBI

20 

Dziadkowiec KN, Gasiorowska E, Nowak-Markwitz E and Jankowska A: PARP inhibitors: Review of mechanisms of action and BRCA1/2 mutation targeting. Prz Menopauzalny. 15:215–219. 2016.PubMed/NCBI

21 

Davar D, Beumer JH, Hamieh L and TawbiH: Role of PARP inhibitors in cancer biology and therapy. Curr Med Chem. 19:3907–3921. 2012. View Article : Google Scholar : PubMed/NCBI

22 

Bhattacharjee S and Nandi S: Choices have consequences: The nexus between DNA repair pathways and genomic instability in cancer. Clin Transl Med. 5:452016. View Article : Google Scholar : PubMed/NCBI

23 

Bhattacharjee S and Nandi S: DNA damage response and cancer therapeutics through the lens of the fanconi anemia DNA repair pathway. Cell Commun Signal. 15:412017. View Article : Google Scholar : PubMed/NCBI

24 

Ghosh S, Lu Y, Katz A, Hu Y and Li R: Tumor suppressor BRCA1 inhibits a breast cancer-associated promoter of the aromatase gene (CYP19) in human adipose stromal cells. Am J Physiol Endocrinol Metab. 292:E246–E252. 2007. View Article : Google Scholar : PubMed/NCBI

25 

Bhattacharjee S and Nandi S: Synthetic lethality in DNA repair network: A novel avenue in targeted cancer therapy and combination therapeutics. IUBMB Life. 69:929–937. 2017. View Article : Google Scholar : PubMed/NCBI

26 

Chen A: PARP inhibitors: Its role in treatment of cancer. Chin J Cancer. 30:463–471. 2011. View Article : Google Scholar : PubMed/NCBI

27 

Ranieri G: Biological basis of tumor angiogenesis and therapeutic intervention: Past, present, and future. Int J Mol Sci. 19:16552018. View Article : Google Scholar

28 

Prat J; FIGO Committee on Gynecologic Oncology, : FIGO's staging classification for cancer of the ovary, fallopian tube, and peritoneum: Abridged republication. J Gynecol Oncol. 26:87–89. 2015. View Article : Google Scholar : PubMed/NCBI

29 

Martin-Oliva D, Aguilar-Quesada R, O'valle F, Muñoz-Gámez JA, Martínez-Romero R, García Del Moral R, Ruiz de Almodóvar JM, Villuendas R, Piris MA and Oliver FJ: Inhibition of poly(ADP-ribose) polymerase modulates tumor-related gene expression, including hypoxia-inducible factor-1 activation, during skin carcinogenesis. Cancer Res. 66:5744–5756. 2006. View Article : Google Scholar : PubMed/NCBI

30 

Mizukami Y, Kohgo Y and Chung DC: Hypoxia inducible factor-1 independent pathways in tumor angiogenesis. Clin Cancer Res. 13:5670–5674. 2007. View Article : Google Scholar : PubMed/NCBI

31 

Zimna A and Kurpisz M: Hypoxia-inducible factor-1 in physiological and pathophysiological angiogenesis: Applications and therapies. Biomed Res Int. 2015:5494122015. View Article : Google Scholar : PubMed/NCBI

32 

Balamurugan K: HIF-1 at the crossroads of hypoxia, inflammation, and cancer. Int J Cancer. 138:1058–1066. 2016. View Article : Google Scholar : PubMed/NCBI

33 

Del Rivero J and Kohn EC: PARP inhibitors: The cornerstone of DNA repair-targeted therapies. Oncology (Williston Park). 31:265–273. 2017.PubMed/NCBI

34 

Wei W, Li Y, Lv S, Zhang C and Tian Y: PARP-1 may be involved in angiogenesis in epithelial ovarian cancer. Oncol Lett. 12:4561–4567. 2016. View Article : Google Scholar : PubMed/NCBI

35 

Tentori L, Lacal PM, Muzi A, Dorio AS, Leonetti C, Scarsella M, Ruffini F, Xu W, Min W, Stoppacciaro A, et al: Poly(ADP-ribose) polymerase (PARP) inhibition or PARP-1 gene deletion reduces angiogenesis. Eur J Cancer. 43:2124–2133. 2007. View Article : Google Scholar : PubMed/NCBI

36 

Benafif S and Hall M: An update on PARP inhibitors for the treatment of cancer. Onco Targets Ther. 8:519–528. 2015.PubMed/NCBI

37 

Oplustilova L, Wolanin K, Mistrik M, Korinkova G, Simkova D, Bouchal J, Lenobel R, Bartkova J, Lau A, O'Connor MJ, et al: Evaluation of candidate biomarkers to predict cancer cell sensitivity or resistance to PARP-1 inhibitor treatment. Cell Cycle. 11:3837–3850. 2012. View Article : Google Scholar : PubMed/NCBI

38 

Arnaudeau C, Lundin C and Helleday T: DNA double-strand breaks associated with replication forks are predominantly repaired by homologous recombination involving an exchange mechanism in mammalian cells. J Mol Biol. 307:1235–1245. 2001. View Article : Google Scholar : PubMed/NCBI

39 

Tutt AN, Lord CJ, McCabe N, Farmer H, Turner N, Martin NM, Jackson SP, Smith GC and Ashworth A: Exploiting the DNA repair defect in BRCA mutant cells in the design of new therapeutic strategies for cancer. Cold Spring Harb Symp Quant Biol. 70:139–148. 2005. View Article : Google Scholar : PubMed/NCBI

40 

Domchek SM, Aghajanian C, Shapira-Frommer R, Schmutzler RK, Audeh MW, Friedlander M, Balmaña J, Mitchell G, Fried G, Stemmer SM, et al: Efficacy and safety of olaparib monotherapy in germline BRCA1/2 mutation carriers with advanced ovarian cancer and three or more lines of prior therapy. Gynecol Oncol. 140:199–203. 2016. View Article : Google Scholar : PubMed/NCBI

41 

Sakai W, Swisher EM, Karlan BY, Agarwal MK, Higgins J, Friedman C, Villegas E, Jacquemont C, Farrugia DJ, Couch FJ, et al: Secondary mutations as a mechanism of cisplatin resistance in BRCA2-mutated cancers. Nature. 451:1116–1120. 2008. View Article : Google Scholar : PubMed/NCBI

42 

Lord CJ and Ashworth A: PARP inhibitors: Synthetic lethality in the clinic. Science. 355:1152–1158. 2017. View Article : Google Scholar : PubMed/NCBI

43 

Fong PC, Boss DS, Yap TA, Tutt A, Wu P, Mergui-Roelvink M, Mortimer P, Swaisland H, Lau A, O'Connor MJ, et al: Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N Engl J Med. 361:123–134. 2009. View Article : Google Scholar : PubMed/NCBI

44 

Fong PC, Yap TA, Boss DS, Carden CP, Mergui-Roelvink M, Gourley C, De Greve J, Lubinski J, Shanley S, Messiou C, et al: Poly(ADP)-ribose polymerase inhibition: Frequent durable responses in BRCA carrier ovarian cancer correlating with platinum-free interval. J Clin Oncol. 28:2512–2519. 2010. View Article : Google Scholar : PubMed/NCBI

45 

Audeh MW, Carmichael J, Penson RT, Friedlander M, Powell B, Bell-McGuinn KM, Scott C, Weitzel JN, Oaknin A, Loman N, et al: Oral poly(ADP-ribose) polymerase inhibitor olaparib in patients with BRCA1 or BRCA2 mutations and recurrent ovarian cancer: A proof-of-concept trial. Lancet. 376:245–251. 2010. View Article : Google Scholar : PubMed/NCBI

46 

Gelmon KA, Tischkowitz M, Mackay H, Swenerton K, Robidoux A, Tonkin K, Hirte H, Huntsman D, Clemons M, Gilks B, et al: Olaparib in patients with recurrent high-grade serous or poorly differentiated ovarian carcinoma or triple-negative breast cancer: A phase 2, multicentre, open-label, nonrandomised study. Lancet Oncol. 12:852–861. 2011. View Article : Google Scholar : PubMed/NCBI

47 

Kaufman B, Shapira-Frommer R, Schmutzler RK, Audeh MW, Friedlander M, Balmaña J, Mitchell G, Fried G, Stemmer SM, Hubert A, et al: Olapari bmonotherapy in patients with advanced cancer and a germline BRCA1/2 mutation. J Clin Oncol. 33:244–250. 2015. View Article : Google Scholar : PubMed/NCBI

48 

Ledermann J, Harter P, Gourley C, Friedlander M, Vergote I, Rustin G, Scott C, Meier W, Shapira-Frommer R, Safra T, et al: Olaparib maintenance therapy in platinum-sensitive relapsed ovarian cancer. N Engl J Med. 366:1382–1392. 2012. View Article : Google Scholar : PubMed/NCBI

49 

Ledermann J, Harter P, Gourley C, Friedlander M, Vergote I, Rustin G, Scott CL, Meier W, Shapira-Frommer R, Safra T, et al: Olaparib maintenance therapy in patients with platinum-sensitive relapsed serous ovarian cancer: a preplanned retrospective analysis of outcomes by BRCA status in a randomized phase 2 trial. Lancet Oncol. 15:852–861. 2014. View Article : Google Scholar : PubMed/NCBI

50 

Pujade-Lauraine E, Ledermann JA, Selle F, Gebski V, Penson RT, Oza AM, Korach J, Huzarski T, Poveda A, Pignata S, et al: Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): A double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 18:1274–1284. 2017. View Article : Google Scholar : PubMed/NCBI

51 

Kim G, Ison G, McKee AE, Zhang H, Tang S, Gwise T, Sridhara R, Lee E, Tzou A, Philip R, et al: FDA approval summary: Olaparib monotherapy in patients with deleterious germline BRCA-mutated advanced ovarian cancer treated with three or more lines of chemotherapy. Clin Cancer Res. 21:4257–4261. 2015. View Article : Google Scholar : PubMed/NCBI

52 

Swisher EM, Lin KK, Oza AM, Scott CL, Giordano H, Sun J, Konecny GE, Coleman RL, Tinker AV, O'Malley DM, et al: Rucaparib in relapsed, platinum-sensitive high-grade ovarian carcinoma (ARIEL2 Part 1): An international, multicentre, open-label, phase 2 trial. Lancet Oncol. 18:75–87. 2017. View Article : Google Scholar : PubMed/NCBI

53 

Coleman RL, Oza AM, Lorusso D, Aghajanian C, Oaknin A, Dean A, Colombo N, Weberpals JI, Clamp A, Scambia G, et al: Rucaparib maintenance treatment for recurrent ovarian carcinoma after response to platinum therapy (ARIEL3): A randomised, double-blind, placebo-controlled, phase 3 trial. Lancet. 390:1949–1961. 2017. View Article : Google Scholar : PubMed/NCBI

54 

Mirza MR, Monk BJ, Herrstedt J, Oza AM, Mahner S, Redondo A, Fabbro M, Ledermann JA, Lorusso D, Vergote I, et al: Niraparib maintenance therapy in platinum-sensitive, recurrent ovarian cancer. N Engl J Med. 375:2154–2164. 2016. View Article : Google Scholar : PubMed/NCBI

55 

Coleman RL, Sill MW, Bell-McGuinn K, Aghajanian C, Gray HJ, Tewari KS, Rubin SC, Rutherford TJ, Chan JK, Chen A and Swisher EM: A phase II evaluation of the potent, highly selective PARP inhibitor veliparib in the treatment of persistent or recurrent epithelial ovarian, fallopian tube, or primary peritoneal cancer in patients who carry a germline BRCA1 or BRCA2 mutation-an NRG oncology/gynecologic oncology group study. Gynecol Oncol. 137:386–391. 2015. View Article : Google Scholar : PubMed/NCBI

56 

Steffensen KD, Adimi P and Jakobsen A: Veliparib monotherapy to patients with BRCA germ line mutation and platinum-resistant or partially platinum-sensitive relapse of epithelial ovarian cancer: A phase I/II study. Int J Gynecol Cancer. 27:1842–1849. 2017. View Article : Google Scholar : PubMed/NCBI

57 

Shen Y, Rehman FL, Feng Y, Boshuizen J, Bajrami I, Elliott R, Wang B, Lord CJ, Post LE and Ashworth A: BMN 673, a novel and highly potent PARP1/2 inhibitor for the treatment of human cancers with DNA repair deficiency. Clin Cancer Res. 19:5003–5015. 2013. View Article : Google Scholar : PubMed/NCBI

58 

de Bono J, Ramanathan RK, Mina L, Chugh R, Glaspy J, Rafii S, Kaye S, Sachdev J, Heymach J, Smith DC, et al: Phase I, dose-escalation, two-part trial of the PARP inhibitor talazoparib in patients with advanced germline BRCA1/2 mutations and selected sporadic cancers. Cancer Discov. 7:620–629. 2017. View Article : Google Scholar : PubMed/NCBI

59 

LaFargue CJ, Dal Molin GZ, Sood AK and Coleman RL: Exploring and comparing adverse events between PARP inhibitors. Lancet Oncol. 20:e15–e28. 2019. View Article : Google Scholar : PubMed/NCBI

60 

Rucaparib, . LiverTox: Clinical and research information on drug-induced liver injury [Internet]. (Bethesda (MD)). National Institute of Diabetes and Digestive and Kidney Diseases. Jun 1–2012-2017.

61 

Olaparib, . LiverTox: Clinical and research information on drug-induced liver injury [Internet]. (Bethesda (MD)). National Institute of Diabetes and Digestive and Kidney Diseases. Jun 1–2012-2017.

62 

Niraparib, . LiverTox: Clinical and research information on drug-induced liver injury [Internet]. (Bethesda (MD)). National Institute of Diabetes and Digestive and Kidney Diseases. Jun 20–2012-2017.

63 

Zhou J, Feng L and Zhang X: Risk of severe hematologic toxicities in cancer patients treated with PARP inhibitors: A meta-analysis of randomized controlled trials. Drug Des Devel Ther. 11:3009–3017. 2017. View Article : Google Scholar : PubMed/NCBI

64 

Alecu I, Milenkova T and Turner SR: Risk of severe hematologic toxicities in cancer patients treated with PARP inhibitors: Results of monotherapy and combination therapy trials. Drug Des Devel Ther. 12:347–348. 2018. View Article : Google Scholar : PubMed/NCBI

65 

Liu Y, Meng J and Wang G: Risk of selected gastrointestinal toxicities associated with poly (ADP-ribose) polymerase (PARP) inhibitors in the treatment of ovarian cancer: A meta-analysis of published trials. Drug Des Devel Ther. 12:3013–3019. 2018. View Article : Google Scholar : PubMed/NCBI

66 

Moore KN, Mirza MR and Matulonis UA: The poly (ADP ribose) polymerase inhibitor niraparib: Management of toxicities. Gynecol Oncol. 149:214–220. 2018. View Article : Google Scholar : PubMed/NCBI

67 

Hsiang YH and Liu LF: Identification of mammalian DNA topoisomerase I as an intracellular target of the anticancer drug camptothecin. Cancer Res. 48:1722–1726. 1988.PubMed/NCBI

68 

Staker BL, Hjerrild K, Feese MD, Behnke CA, Burgin AB Jr and Stewart L: The mechanism of topoisomerase I poisoning by a camptothecin analog. Proc Natl Acad Sci USA. 99:15387–15392. 2002. View Article : Google Scholar : PubMed/NCBI

69 

ten Bokkel Huinink W, Gore M, Carmichael J, Gordon A, Malfetano J, Hudson I, Broom C, Scarabelli C, Davidson N, Spanczynski M, et al: Topotecan versus paclitaxel for the treatment of recurrent epithelial ovarian cancer. J Clin Oncol. 15:2183–2193. 1997. View Article : Google Scholar : PubMed/NCBI

70 

D'Onofrio G, Tramontano F, Dorio AS, Muzi A, Maselli V, Fulgione D, Graziani G, Malanga M and Quesada P: Poly(ADP-ribose) polymerase signaling of topoisomerase 1-dependent DNA damage in carcinoma cells. Biochem Pharmacol. 81:194–202. 2011. View Article : Google Scholar : PubMed/NCBI

71 

Patel AG, Flatten KS, Schneider PA, Dai NT, McDonald JS, Poirier GG and Kaufmann SH: Enhanced killing of cancer cells by poly(ADP-ribose) polymerase inhibitors and topoisomerase I inhibitors reflects poisoning of both enzymes. J Biol Chem. 287:4198–4210. 2012. View Article : Google Scholar : PubMed/NCBI

72 

Wahner Hendrickson AE, Menefee ME, Hartmann LC, Long HJ, Northfelt DW, Reid JM, Boakye-Agyeman F, Kayode O, Flatten KS, Harrell MI, et al: A phase I clinical trial of the Poly(ADP-ribose) polymerase inhibitor veliparib and weekly topotecan in patients with solid tumors. Clin Cancer Res. 24:744–752. 2018. View Article : Google Scholar : PubMed/NCBI

73 

Hjortkjær M, Kanstrup H, Jakobsen A and Steffensen KD: Veliparib and topotecan for patients with platinum-resistant or partially platinum-sensitive relapse of epithelial ovarian cancer with BRCA negative or unknown BRCA status. Cancer Treat Res Commun. 14:7–12. 2018. View Article : Google Scholar : PubMed/NCBI

74 

Bi Y, Verginadis II, Dey S, Lin L, Guo L, Zheng Y and Koumenis C: Radiosensitization by the PARP inhibitor olaparib in BRCA1-proficient and deficient high-grade serous ovarian carcinomas. Gynecol Oncol. 150:534–544. 2018. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

October-2020
Volume 20 Issue 4

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Loizzi V, Ranieri G, Laforgia M, Gadaleta CD, Gargano G, Kardhashi A, De Liso M, Naglieri E, Del Vecchio V, Cicinelli E, Cicinelli E, et al: PARP inhibitors and epithelial ovarian cancer: Molecular mechanisms, clinical development and future prospective (Review). Oncol Lett 20: 90, 2020
APA
Loizzi, V., Ranieri, G., Laforgia, M., Gadaleta, C.D., Gargano, G., Kardhashi, A. ... Cormio, G. (2020). PARP inhibitors and epithelial ovarian cancer: Molecular mechanisms, clinical development and future prospective (Review). Oncology Letters, 20, 90. https://doi.org/10.3892/ol.2020.11951
MLA
Loizzi, V., Ranieri, G., Laforgia, M., Gadaleta, C. D., Gargano, G., Kardhashi, A., De Liso, M., Naglieri, E., Del Vecchio, V., Cicinelli, E., Cormio, G."PARP inhibitors and epithelial ovarian cancer: Molecular mechanisms, clinical development and future prospective (Review)". Oncology Letters 20.4 (2020): 90.
Chicago
Loizzi, V., Ranieri, G., Laforgia, M., Gadaleta, C. D., Gargano, G., Kardhashi, A., De Liso, M., Naglieri, E., Del Vecchio, V., Cicinelli, E., Cormio, G."PARP inhibitors and epithelial ovarian cancer: Molecular mechanisms, clinical development and future prospective (Review)". Oncology Letters 20, no. 4 (2020): 90. https://doi.org/10.3892/ol.2020.11951