Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

Molecular Pathways: Targeting PARP in Cancer Treatment

Khanh Do1, Alice P. Chen2

Affiliations: Medical Oncology Branch, Center for Cancer Research, NCI1, Cancer Therapy Evaluation Program, Division of Cancer Treatment and Diagnosis, NCI2

Corresponding Author Contact: Alice P. Chen, M.D., 6130 Executive Blvd, Suite 7130, Bethesda MD 20852, (301)496-1196, [email protected]

Authors have no conflict of interest to report.

1

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

Abstract

Poly (ADP-ribose) polymerases (PARPs) are a family of nuclear protein enzymes involved in the DNA damage response. The role of PARP-1 in base excisional repair has been extensively characterized. More recent in vitro studies additionally implicate a role for PARP-1 in facilitating homologous recombination and non-homologous end-joining. The more faithful process of homologous recombination repair of double-stranded DNA breaks, involves localization of BRCA-1 and BRCA-2 to sites of DNA damage, resection of the double-stranded break, and gap-filling DNA synthesis using the homologous sister chromatid as a template. Simultaneous dysfunction of both DNA repair pathways decreases the ability of cells to compensate, and forms the basis for the concept of synthetic lethality. Treatment strategies thus far have focused on two main principals: 1) exploitation of the concept of synthetic lethality in homologous recombination deficient tumors, primarily in breast and ovarian cancer patients with BRCA mutation, and 2) as radio-sensitizers and chemo-sensitizers. BRCA deficiency accounts for only a fraction of dysfunction in homologous recombination repair. Epigenetic alterations of BRCA function and defects within the Fanconi anemia pathway also result in defective DNA repair. Rational therapeutic combinations exploiting alternate mechanisms of defective DNA repair, abrogation of cell cycle checkpoints, and additional functions of PARP-1, present novel opportunities for further clinical development of PARP inhibitors. Based on the results of clinical studies of PARP inhibitors thus far, it is imperative that future development of PARP inhibitors take a more refined approach, identifying the unique subset of patients that would most benefit from these agents, determining the optimal time for use, and identifying the optimal combination partner in any particular setting.

I. Background

The integrity of the human genome is constantly under stress from both endogenous genotoxic insults such as reactive oxygen species generated during normal metabolism, or from exogenous insults such as chemotherapeutic agents. Cellular response depends upon the magnitude of the insult, resulting in induction of cell cycle checkpoint pathways and DNA repair mechanisms. If the damage is extensive and irreparable, induction of cell death occurs. Poly (ADP-ribose) polymerases (PARPs) are a family of nuclear protein enzymes involved in the post-translational modification of proteins and synthesis of poly (ADP-ribose). PARP-1 and PARP-2 are the best characterized members of the PARP family, and play a key role in the DNA damage response and repair of single-stranded breaks (SSBs) through base excisional repair (BER).

The role of PARP-1 in BER has been the most extensively. After binding to SSBs, PARP-1 transfers the ADP-ribosyl moiety from NAD+ to acceptor proteins, generating long chains of poly (ADP-ribosyl)ated polymers. This allows for the recruitment of DNA repair proteins such as DNA polymerase β, DNA ligase III, and scaffolding proteins such as x-ray cross complementing protein 1 (XRCC1) to sites of SSBs (1,2). PARP may also facilitate homologous 2

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

recombination (HR) via recruitment of factors like ataxia telangiectasia-mutated (ATM), mitotic recombination 11 (Mre11), and Nijmegen breakage syndrome 1 (Nbs1) to sites of double- stranded DNA damage (3), and has been shown to interact with the DNA protein kinase complex involved with non-homologous end-joining (NHEJ) (4) [Figure 1]. More recent studies by Helleday et al. propose that PARP inhibition results in trapping of PARP-1 on DNA repair intermediates at SSBs and stalling of replication forks that require BRCA-dependent HR for resolution (5). The more faithful process of HR repair, involves localization of BRCA-1 and BRCA-2 to sites of DNA damage, resection of double-stranded breaks (DSBs), and gap-filling DNA synthesis using the homologous sister chromatid as a template. Whereas, in NHEJ, DNA break ends are directly ligated without the use of a homologous template, resulting in the introduction of errors at ligated sites. DSBs, which can occur during repair of SSBs or secondarily from chemotherapeutic agents or ionizing radiation, are the most critical form of DNA damage and can result in problems for transcription and replication, eventually leading to apoptosis (6).

Loss of PARP activity in and of itself is not lethal to cells with intact HR pathway, as unrepaired SSBs are converted to DSBs which can be effectively repaired via the HR pathway. The concept that a double-hit in DNA repair would result in synthetic lethality is the rationale for investigating PARP inhibition in BRCA 1/2 mutated cell lines deficient in HR. BRCA-1 plays a role in the surveillance of DNA damage and transduction of DNA repair responses whereas BRCA-2 plays a direct role in DNA repair, via modulation of Rad51, which is involved in double-stranded repair by HR (7). Cells that are deficient in BRCA-1 or BRCA-2 are rendered defective in the ability to repair through HR and depend upon error-prone NHEJ for DNA repair. The resultant amplification of DNA instability and chromosomal aberrations results in cell death. This concept of synthetic lethality was first reported in two landmark studies in 2005 (8, 9). Investigators showed that inhibition of PARP-1 both by chemical inhibition and siRNA depletion in embryonic stem cells lacking BRCA 1 or BRCA 2 resulted in generation of chromatid breaks, G2 cell cycle arrest, and enhancement of apoptosis.

II. Clinical-Translational Advances

PARP inhibitors currently in clinical development compete with the nicotinamide moiety of NAD+ for the catalytic domain of the PARP enzyme. This prevents its release from sites of DNA damage and therefore, prevents DNA repair enzymes from accessing the site. (AZD-2281/KU-0059436) and veliparib (ABT-888) have been the most extensively studied in clinical trials. Additional PARP inhibitors currently under active clinical investigation also include (MK-4827), CEP-9722, E7016/GPI-21016, BMN-673, and (AG- 014699/PF-01367338). Iniparib (BSI-201) was initially studied in triple negative breast cancer (TNBC) (10); however, recent evidence suggests iniparib does not inhibit PARP-1 and PARP-2 at clinically relevant doses, and is no longer considered a PARP inhibitor (11). Given the variable potency of current PARP inhibitors in clinical development, it is not known to what degree of PARP inhibition is needed for maximal efficacy, and to what degree the off-target effects beyond PARP inhibition contribute to antitumor effect. A sandwich immunoassay to measure levels of poly(ADP-ribose) macromolecules (PAR) in both peripheral blood mononuclear cells (PBMCs) and tumor has been developed to determine on target effect of the 3

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

PARP inhibitors (12, 13). γH2AX levels and caspase 3 levels are also currently being evaluated as surrogate markers of efficacy.

Treatment strategies thus far have focused on two main principals: 1) exploitation of the concept of synthetic lethality in HR deficient tumors, primarily in BRCA mutation -related breast and ovarian cancers, and 2) as radio- and chemotherapy-sensitizers. Early Phase I studies evaluating single-agent olaparib in a cohort of patients enriched for BRCA 1/2 mutation carriers demonstrated clinical benefit in 12 of 17 (63%) with objective responses in 9 of 19 (47%) patients (14). In the expansion cohort of 50 BRCA mutation carriers with ovarian, peritoneal, and fallopian tube cancer, 46% received clinical benefit, with the highest benefit rate noted in patients with platinum-sensitive disease (15). Pharmacodynamic assays demonstrated more than 90% reduction of PAR levels in PBMCs at a dose of 60mg administered on a twice daily dosing schedule implicating on target effect.

The concept of PARP inhibition in sporadic cancers with epigenetic loss of BRCA function has also been evaluated (16, 17). In a phase 2 study evaluating olaparib in ovarian and TNBC patients stratified for the presence or absence of BRCA-1 or BRCA-2 mutation, of the 63 evaluable patients, objective responses were seen in 11 of 46 (24%) patients with no BRCA mutation, all in high-grade serous ovarian cancer, supporting the notion that alternative mechanisms for failure of HR may also predispose sensitivity to PARP inhibition (18). Similar findings have also been demonstrated with veliparib and niraparib. In the Phase I study evaluating continuous dosing of veliparib in patients with BRCA 1/2 mutated cancer, platinum- refractory ovarian cancer, and basal-like breast cancer, 1 partial response (PR) and 7 stable diseases (SD) were seen in 25 patients with no BRCA mutation (19). In the Phase I study evaluating niraparib in advanced solid tumors enriched for sporadic cancers associated with non- BRCA HR repair defects, there were 3 PR and 4 SD of the 20 evaluable patients (20).

The majority of clinical development of PARP inhibitors thus far, has focused on PARP inhibitors as chemo-sensitizers in combination with alkylating agents, topoisomerase inhibitors, platinum agents, and/or (21-31; Table 1). The use of rucaparib, olaparib and veliparib in combination with chemotherapy has been burdened with increased myelosuppression. This raises the question of whether the addition of PARP inhibitor improves efficacy in the presence of decreased doses of chemotherapy. The randomized study of and veliparib did not show improvement in overall survival over temozolomide alone (32). Another issue concerns the correlation between PAR levels and activity. A multicenter randomized trial comparing metronomic with or without veliparib showed no improvement in response rates of the combination, despite demonstration of 50% reduction of PAR levels in 9 of 10 patient PBMCs (33). One reason could be the lack of tight correlation between the PAR levels in PBMCs and tumor. Less myelosuppressive chemotherapy combinations and off target effects of PARP inhibitors are currently being investigated.

Extensive preclinical data also exists for PARP inhibition and augmentation of cellular radio-sensitivity (34-39). Radio-sensitization of a non-Hodgkin’s lymphoma cell line has been shown with olaparib or veliparib in combination with both external beam radiation and 131I- tositumomab (35). Radio-sensitization with veliparib has been demonstrated in head and neck 4

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

carcinoma cell lines (36) and lung cancer xenograft models (37), with niraparib in neuroblastoma cell lines (38), and with E7016 in combination with temozolomide and radiotherapy in a glioma mouse model (39). A clinical trial is ongoing evaluating veliparib in combination with whole brain radiation therapy in patients with brain metastases. Of 48 evaluable patients, best tumor response was seen in 37.5% in non-small cell lung carcinoma and 52.9% in breast cancer patients (40). No enhanced toxicity was noted unlike the myelosuppression seen with chemotherapy making this approach more attractive for further investigation.

III. Future Directions

The majority of clinical development of PARP inhibitors thus far, has been limited to the setting of relapsed disease. More recently, olaparib has been studied in the maintenance setting. In a phase 2 study, patients with platinum-sensitive relapsed, high-grade serous ovarian cancer were randomized to receive either maintenance olaparib at 400mg twice daily dose or placebo. Progression-free survival was longer with olaparib (median of 8.4 months) versus placebo (median 4.8 months) [HR of 0.35, CI 0.25-0.49, P< 0.001] (41). Given the tolerability of single agent treatment and with evidence indicating activity with treating bulk and small volume disease in high-risk patients with known BRCA 1/2 mutation, the question of PARP inhibitor as a preventative agent has been raised. A major obstacle in evaluating PARP inhibitors in prevention is the concerns for risk of secondary malignancy. This will be answered as more data becomes available on the long term use of PARP inhibitors. In addition, BRCA deficiency accounts for only a fraction of causes of defective HR repair. Additional causes include defects in other HR pathway proteins, alterations of proteins involved in the Fanconi anemia (FA) pathway, defects of proteins in the sensing of DNA damage, as well as in the cell cycle checkpoints (42). Early in vitro studies of phosphatase and tensin homolog (PTEN) null cells implicated a role for Rad51-mediated double-stranded DNA repair, and sensitization to PARP inhibition similar to BRCA-deficient cells (43). PTEN mutations are common in multiple cancers in addition to endometrial cancers and glioblastomas (44). The mechanism by which PTEN loss results in genomic instability may involve perturbations of multiple DNA repair pathways, presenting a unique opportunity for further development of PARP inhibitors in this subset of patients. A study of abiraterone and veliparib in metastatic castration-resistant prostate cancer is currently underway to further explore this concept (ClinicalTrials.gov identifier: NCT01576172). The complex interaction between various components of the Fanconi pathway and other pathways involved in HR presents an opportunity to expand upon the subset of patients that may benefit from PARP inhibition. In response to DNA damage, a nuclear complex of FA proteins form, culminating in the downstream ubiquination of FANCD2 that is required for repair of DNA crosslinks and interaction with BRCA2 and DNA repair (45). Inactivation of the FA pathway through methylation of the FANCF promoter (46-48) has been detected in a number of sporadic cancers. Additional epigenetic alterations affecting FANCC expression in 53% of early-onset and 59% of late-onset sporadic breast cancers have also been identified (49). Studies are ongoing to evaluate this concept by testing patients with somatic deficiency in FA pathway with veliparib with or without (50).

5

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

Evidence also exists for sensitivity to PARP inhibition in the setting of defective detection of DNA damage and abrogation of cell cycle checkpoints. Cyclin-dependent kinase 1 (Cdk1), a key component of cell cycle regulation, phosphorylates BRCA1 forming BRCA1 foci at sites of DNA damage to facilitate checkpoint activation. Recent preclinical studies show inhibition of Cdk1 activity sensitizes BRCA-proficient cancers to PARP inhibition (51). The addition of Cdk1 inhibition to olaparib resulted in synergistic inhibition of growth of TNBC cells in vitro (52). A phase I trial of veliparib and SCH727965 is ongoing to evaluate this concept (ClinicalTrials.gov identifier: NCT01434316). Additional in vitro evidence also exists for sensitivity to PARP inhibition in the setting of deficiency of ATM involved in the initial detection of DNA damage, as well as checkpoint kinase 2 (Chk2), the downstream effector of cell cycle arrest (42). With the initial promise of PARP inhibitor in BRCA deficient patients, other populations are being investigated as well as combinations with other targeted agents. More recent evidence suggests pleiotropism of PARP inhibition, and additional off-target effects of PARP inhibitors. In addition to playing a central role in DNA repair, preclinical evidence additionally implicates a role for PARP in angiogenesis. In vitro studies showed inhibition of PARP resulted in enhancement of VEGF protein expression on endothelial cells (53) and inhibition of growth-factor stimulated migration and proliferation of human umbilical vein endothelial cells (54). PARP inhibition also resulted in down-regulation of genes involved in angiogenesis during skin carcinogenesis in mice (55). A phase I study combining olaparib with cediranib, a multi-targeted kinase inhibitor of VEGFR-1/2/3 and c-kit, in recurrent ovarian and TNBC, showed a 56% unconfirmed response rate in the cohort of ovarian cancer patients, but with hematologic dose-limiting toxicities (56). Recent comparative analyses of rucaparib, olaparib, and veliparib in TNBC cell lines suggest target pleiotropism of the various PARP inhibitors, and implicate additional PARP-independent signaling mechanisms thought to account for differing levels of anti-tumor activity (57). In this study, rucaparib demonstrated the highest antitumor potency in all cell lines evaluated, and was distinct in its ability to suppress Stat3 phosphorylation at concentrations of 2.5uM. Additional evidence suggests rucaparib and veliparib may also target the NF-κB pathway (58, 59). In response to DNA damage, NF-κB is activated via signaling cascades, culminating in the transcriptional modulation of genes involved in cellular proliferation. Aberrant NF-κB activation is thought to contribute to chemoresistance in multiple cancers (60). Recent cell viability assays evaluating veliparib in HR proficient HER2+ breast cancer cell lines demonstrate cytotoxicity with evidence of inhibition of NF-κB at the transcriptional level at concentrations of 10uM (59). Further investigation into PARP- independent off-target effects of PARP inhibition is ongoing.

IV. Conclusions Clinical development of PARP inhibitors in combination with chemotherapy thus far have been limited by the fact that when used in combination with chemotherapeutic agents, they could not be used in full doses due to enhanced myelosuppression, often requiring dose reduction of the partnered cytotoxic agent. The question of whether PARP inhibitor combinations in the setting of dose reduction is superior to full dose therapy without PARP inhibitor needs to be addressed in future studies. In addition, given the variable potency of current PARP inhibitors in 6

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

clinical development, the degree of PARP inhibition needed for maximal efficacy, and to what degree the off-target effects are contributing to antitumor effect is not known. Based on known data to date, it is imperative that future development of PARP inhibitors take a more refined approach, identifying the unique subset of patients that would most benefit from these agents: determining the optimal time for use, identifying the optimal combination partner in any particular setting, and the optimal population. Further investigation is needed with regard to how various additional DNA repair pathways interact and whether synthetic lethality may be utilized in this setting as well. Further delineation of off-target effects of current PARP inhibitors may present additional opportunities for drug development. Finally, identification and screening for mechanisms of resistance, reactivation of HR by secondary mechanisms, and alternate DNA repair pathways which may compensate for loss of PARP activity would allow for optimal use of the PARP inhibitors in an oncologic setting.

7

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

References:

1. Houtgraaf JH, Versmissen J, van der Giessen WJ. A concise review of DNA damage checkpoints and repair in mammalian cells. Cardiovasc Revasc Med 2006;7:165-172. 2. El-Khamisy SF, Masutani M, Suzuki H, Caldecott KW. A requirement for PARP-1 for the assembly or stability of XRCC1 nuclear foci at sites of oxidative DNA damage. Nucleic Acids Res 2003;31:5526-5533. 3. Haince JF, McDonald D, Rodrigue A, Déry U, Masson JY, Hendzel MJ, Poirier GG. PARP-1 dependent kinetics of recruitment of MRE11 and NBS1 proteins to multiple DNA damage sites. J Biol Chem 2008;283:1197-1208. 4. Wang M, Wu W, Wu W, Rosidi B, Zhang L, Wang H et al. PARP-1 and Ku compete for repair of DNA double strand breaks by distinct NHEJ pathways. Nucleic Acids Res 2006;34:6170-6182. 5. Helleday T. The underlying mechanism for the PARP and BRCA synthetic lethality: clearing up the misunderstandings. Mol Oncol 2011;5:387-393. 6. Hoeijmakers JH. Genome maintenance mechanisms for preventing cancer. Nature 2001;411:366-374. 7. Tutt A, Ashworth A. The relationship between the roles of BRCA genes in DNA repair and cancer predisposition. Trends Mol Med 2002;8:571-576. 8. Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D, Lopez E et al. Specific killing of BRCA-2 deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 2005;434:913-917. 9. Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005;434:917-921. 10. O’Shaughnessy J, Osborne C, Pippen JE, Yoffe M, Patt D, Rocha C et al. Iniparib plus chemotherapy in metastatic triple-negative breast cancer. N Engl J Med 2011;364:205- 214. 11. Patel AG, De Lorenzo SB, Flatten KS, Poirier GG, Kaufmann SH. Failure of iniparib to inhibit poly(ADP-Ribose) polymerase in vitro. Clin Cancer Res 2012;18:1655-1662. 12. Kinders RJ, Hollingshead M, Khin S, Rubinstein L, Tomaszewski JE, Doroshow JH et al. Preclinical modeling of a phase 0 clinical trial: qualification of a pharmacodynamic assay of poly (ADP-ribose) polymerase in tumor biopsies of mouse xenografts. Clin Cancer Res 2008;14:6877-6885. 13. Ji J, Kinders RJ, Zhang Y, Rubinstein L, Kummar S, Parchment RE et al. Modeling pharmacodynamic response to the poly(ADP-Ribose) polymerase inhibitor ABT-888 in human peripheral blood mononuclear cells. PLoS One 2011;6:e26152. 14. 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;361:123-134. 15. Fong PC, Yap TA, Boss DS, Carden CP, Mergui-Roelvink M, Gourley C et al. Poly(ADP)-ribose polymerase inhibition: frequent durable responses in BRCA carrier ovarian cancer correlating with platinum-free interval. J Clin Oncol 2010;28:2512-2519.

8

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

16. Esteller M, Silva JM, Dominquez G, Bonilla F, Matias-Guiu X, Lerma E et al. Promoter hypermethylation and BRCA1 inactivation in sporadic breast and ovarian tumors. J Natl Cancer Inst 2000;92:564-569. 17. Catteau A, Harris WH, Xu CF, and Solomon E. Methylation of the BRCA1 promoter region in sporadic breast and ovarian cancer: correlation with disease characteristics. Oncogene 1999;18:1957-1965. 18. Gelmon KA, Tischkowitz M, Mackay H, Swenerton K, Robidoux A, Tonkin K et al. Olaparib in patients with recurrent high-grade serous or poorly differentiated ovarian carcinoma or triple-negative breast cancer: a phase 2, multicentre, open-label, non- randomised study. Lancet Oncol 2011;12:852-861. 19. Huggins-Puhalla SL, Beumer JH, Appleman LJ, Tawbi HA-H, Stoller RG, Lin Y et al. A phase I study of chronically dosed, single-agent veliparib (ABT-888) in patients (pts) with either BRCA 1/2-mutated cancer (BRCA+), platinum-refractory ovarian cancer, or basal-like breast cancer (BRCA-wt). J Clin Oncol 2012;30(suppl.) [abstract 3054]. 20. Schelman WR, Sandhu SK, GarciaVM, Wilding G, Sun L, Toniatti C et al. First-in- human trial of poly(ADP)-ribose polymerase (PARP) inhibitor MK-4827 in advanced cancer patients with antitumor activity in BRCA-deficient tumors and sporadic ovarian cancers (soc). J Clin Oncol 2012;29(suppl.) [abstract 3102]. 21. Plummer R, Middleton M, Wilson R, Evans J, Olsen A, Curtin N et al. Phase I study of the poly(ADP-ribose) polymerase inhibitor, AG014699, in combination with temozolomide in patients with advanced solid tumors. Clin Cancer Res 2008;14:7917- 7923. 22. Plummer R, Lorigan P, Evans J,Steven N, Middleton M, Wilson R et al. First and final report of a phase II study of the poly(ADP-ribose) polymerase (PARP) inhibitor, AG014699, in combination with temozolomide (TMZ) in patients with metastatic malignant melanoma. JCO, 2006 ASCO Annual Meeting Proceedings Part I. Vol 24, No. 18S (June 20 Supplement), 2006: 8013. 23. Lee J, Annunziata CM, Minasian LM, Zujewski J, Prindiville SA, Kotz HL et al. Phase I study of the PARP inhibitor olaparib (O) in combination with (C) in BRCA 1/2 mutation carriers with breast (Br) or ovarian (Ov) cancer (Ca). J Clin Oncol 2011;29(suppl.)[abstract2520]. 24. Dent RA, Linderman GJ, Clemons M, Wildiers H, Chan A, McCarthy NJ et al. Safety and efficacy of the oral PARP inhibitor olaparib (AZD2281) in combination with for the first- or second-line treatment of patients with metastatic triple-negative breast cancer: Results from the safety cohort of a Phase I/II multicentre trial. J Clin Oncol 2010;28(suppl.)[abstract:1018]. 25. Khan OA, Gore M, Lorigan P, Stone J, Greystoke A, Burke W et al. A phase I study of the safety and tolerability of olaparib (AZD2281, KU0059436) and in patients with advanced solid tumours. Br J Cancer 2011;104:750-755. 26. Giaccone G, Rajan A, Kelly RJ, Gutierrez M, Kummar S, Yancey M et al. A phase I combination study of olaparib (AZD2281; KU-0059436) and (C) plus (G) in adults with solid tumors. J Clin Oncol 2010;28(suppl.) [abstract 3027].

9

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

27. Isakoff SJ, Overmoyer B, Tung NM, Gelman RS, Giranda VL, Bernhard KM et al. A phase II trial of the PARP inhibitor veliparib (ABT888) and temozolomide for metastatic breast cancer. J Clin Oncol 2010;28(suppl.) [abstract 1019]. 28. Kummar S, Chen A, Ji J, Zhang Y, Ames M, Jia L et al. Phase I study of PARP inhibitor ABT-888 in combination with in adults with refractory solid tumors and lymphomas. Cancer Res 2011;71:5626-5634. 29. LoRusso P, Ji J, Li J, Heilbrun LK, Shapiro G, Sausville EA et al. Phase I study of the safety, pharmacokinetics (PK), and pharmacodynamics (PD) of the poly(ADP-ribose) polymerase (PARP) inhibitor veliparib (ABT-888;V) in combination with (CPT-111;Ir) in patients (pts) with advanced solid tumors. J Clin Oncol 2011;29(suppl.) [abstract 3000]. 30. Tan AR, Toppmeyer D, Stein MN, Moss RA, Gounder M, Lindquist DC et al. Phase I trial of veliparib, (ABT-888), a poly(ADP-ribose) polymerase (PARP) inhibitor, in combination with and cyclophosphamide in breast cancer and other solid tumors. J Clin Oncol 2011;29(suppl.) [abstract 3041]. 31. Kummar S, Ji J, Morgan R, Lenz HJ, Puhalla SL, Belani CP et al. A phase I study of veliparib in combination with metronomic cyclophosphamide in adults with refractory solid tumors and lymphomas. Clin Cancer Res 2012;18:1726-1734. 32. Middleton M, Friedlander P, Hamid O, Daud A, Plummer R, Schuster R et al. Efficacy of veliparib (ABT-888) plus temozolomide versus temozolomide alone: a randomized, double-blind, placebo-controlled trial in patients with metastatic melanoma. Presentation at ECCO/ESMO 2011. 33. Kummar S, Oza AM, Fleming G, Sullivan D, Gandara D, Erlichman C et al. Randomized trial of oral cyclophosphamide (C) with or without veliparib (V), and oral poly (ADP- ribose) polymerase (PARP) inhibitor, in patients with recurrent BRCA-positive ovarian, or primary peritoneal, or high-grade serous ovarian carcinoma. J Clin Oncol 2012;30(suppl.) [abstract 5020]. 34. Lunec J, George AM, Hedges M, Cramp WA, Whish WJ, Hunt B. Post-irradiation sensitization with the ADP-ribosyltransferase inhibitor 3-acetamidobenzamide. Br J Cancer Suppl 1984;6:19-25. 35. Schaefer NG, James E, Wahl RL. Poly(ADP-ribose) polymerase inhibitors combined with external beam and radioimmunotherapy to treat aggressive lymphoma. Nucl Med Commun 2011;32:1046-1051. 36. Nowsheen S, Bonner JA, Yang ES. The poly(ADP-ribose) polymerase inhibitor ABT- 888 reduces radiation-induced nuclear EGFR and augments head and neck tumor response to radiotherapy. Radiother Oncol 2011;99:331-338. 37. Albert JM, Cao C, Kim KW, Willey CD, Geng L, Xiao D et al. Inhibition of poly(ADP- ribose) polymerase enhances cell death and improves tumor growth delay in irradiated lung cancer models. Clin Cancer Res 2007;13:3033-3042. 38. Bhargava S, Mueller S, Wehmeijer L, Yang X, Gragg A, Matthay K et al. PARP-1 inhibitor MK-4827 in combination with radiation as a treatment strategy for metastatic neuroblastoma. J Clin Oncol 2011;29(suppl.) [abstract 9559]. 39. Russo AL, Kwon HC, Burgan WE, Carter D, Beam K, Weizheng X et al. In vitro and in vivo radiosensitization of glioblastoma cells by the poly (ADP-ribose) polymerase inhibitor E7016. Clin Cancer Res 2009;15:607-612. 10

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

40. Mehta MP, Curran WJ, Wang D, Wang F, Kleinberg L, Brade AM et al. Phase I safety and pharmacokinetic (PK) study of veliparib in combination with whole brain radiation therapy (WBRT) in patients (pts) with brain metastases. J Clin Oncol 2012;30(suppl.) [abstract 2013]. 41. 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;366:1382-1392. 42. McCabe N, Turner NC, Lord CJ, Kluzek K, Biakowska A, Swift S et al. Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP- ribose) polymerase inhibition. Cancer Res 2006;66:8109-8115. 43. Shen WH, BalajeeAS, Wang J, Wu H, Eng C, Pandolfi PP et al. Essential role for nuclear PTEN in maintaining chromosomal integrity. Cell 2007;128:157-170. 44. Keniry M, Parsons R. The role of PTEN signaling perturbations in cancer and in targeted therapy. Oncogene 2008;27:5477-5485. 45. Wang X, D’Andrea AD. The interplay of Fanconi anemia proteins in the DNA damage response. DNA Repair (Amst) 2004;3:1063-1069. 46. Wei M, Xu J, Dignam J, Nanda R, Sveen L, Fackenthal J et al. Estrogen receptor alpha, BRCA1, and FANCF promoter methylation occur in distinct subsets of sporadic breast cancers. Breast Cancer Res Treat 2008;111:113-120. 47. Wang Z, Li M, Lu S, Zhang Y, Wang H. Promoter hypermethylation of FANCF plays an important role in the occurrence of ovarian cancer through disrupting Fanconi anemia- BRCA pathway. Cancer Biol Ther 2006;5:256-260. 48. Narayan G, Arias-Pulido H, Nandula SV, Basso K, Sugirtharaj DD, Vargas H et al. Promoter hypermethylation of FANCF: disruption of Fanconi Anemia-BRCA pathway in cervical cancer. Cancer Res 2004;64:2994-2997. 49. Sinha S, Singh RK, Alam N, Roy A, Roychoudhury S, Panda CK. Alterations in candidate genes PHF2, FANCC, PTCH1, and XPA at chromosomal 9q22.3 region: pathologic significance in early- and late-onset breast carcinoma. Mol Cancer 2008;7:84. 50. Villalona-Calero MA, Duan W, Zhao W, Gao L, Thurmond J, Leon M et al. Phase I trial of veliparib or mitomycin C + veliparib in pts with Fanconi anemia pathway (FA) repair defects. J Clin Oncol 2011;29(suppl.) [abstract A99]. 51. Johnson N, Li YC, Walton ZE, Cheng KA, Li D, Rodig SJ et al. Compromised CDK1 activity sensitizes BRCA-proficient cancers to PARP inhibition. Nat Med 2011;17:875- 882. 52. Cotter MB, Pierce A, McGowan PM, Flanagan L, Quinn C, Evoy D et al. Preclinical evaluation of PARP inhibition in breast cancer: Comparative effectiveness of olaparib and iniparib. J Clin Oncol 2012;30(suppl.) [abstract 1042]. 53. Beckert S, Farrahi F, Perveen Ghani Q, Aslam R, Scheuenstuhl H, Coerper S et al. IGF-I- induced VEGF expression in HUVEC involves phosphorylation and inhibition of poly(ADP-ribose)polymerase. Biochem Biophys Res Commun 2006;341:67-72. 54. Rajesh M, Mukhopadhyay P, Bátkai S, Godlewski G, Haskó G, Liaudet L et al. Pharmacological inhibition of poly(ADP-ribose) polymerase inhibits angiogenesis. Biochem Biophys Res Commun 2006;350:352-357. 55. Martin-Oliva D, Aguilar-Quesada R, O'valle F, Muñoz-Gámez JA, Martínez-Romero R, García Del Moral R et al. Inhibition of poly(ADP-ribose) polymerase modulates tumor-

11

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

related gene expression, including hypoxia-inducible factor-1 activation, during skin carcinogenesis. Cancer Res 2006;66:5744-5756. 56. Liu J, Fleming GF, Tolaney SM, Birrer MJ, Penson RT, Berlin ST et al. A phase I trial of the PARP inhibitor olaparib (AZD2281) in combination with the antiangiogenic cediranib (AZD2171) in recurrent ovarian or triple-negative breast cancer. J Clin Oncol 2011;29(suppl.) [abstract 5028]. 57. Chuang HC, Kapuriya N, Kulp SK, Chen CS, Shapiro CL. Differential anti-proliferative activities of poly(ADP-ribose) polymerase (PARP) inhibitors in triple-negative breast cancer cells. Breast Cancer Res Treat 2012;134:649-659. 58. Hunter JE, Willmore E, Irving JA, Hostomsky Z, Veuger SJ, Durkacz BW. NF-kB mediates radio-sensitization by the PARP-1 inhibitor, AG-014699. Oncogene 2012;31:251-264. 59. Yang ES, Nowsheen S, Cooper T, LoBuglio A, Bonner J. Susceptibility of HER2+ breast cancer cells to poly (ADP-ribose) polymerase (PARP) inhibition independent of an inherent DNA repair defect. J Clin Oncol 2012;30(suppl.) [abstract 621]. 60. Chaturvedi MM, Sung B, Yadav VR, Kannappan R, Aggarwal BB. NF-kB addiction and its role in cancer: ‘one size does not fit all’. Oncogene 2011;30:1615-1630. 61. Rouleau M, Patel A, Hendzel MJ, Kaufmann SH, Poirier GG. PARP inhibition: PARP1 and beyond. Nat Rev Cancer 2010;10:293-301.

12

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Targeting PARP

Figure 1. After binding to SSBs, PARP-1 transfers the ADP-ribosyl moiety from NAD+ to acceptor proteins, generating long chains of poly (ADP-ribosyl)ated polymers. This allows for the recruitment of DNA repair proteins such as DNA polymerase β, DNA ligase III, and scaffolding proteins XRCC1 to sites of SSBs. PARP may also facilitate homologous recombination via recruitment of factors like ATM, Mre11, and Nbs1 to sites of double-stranded DNA damage, and has been shown to interact with the DNA protein kinase complex involved with non-homologous end-joining. ATM, ataxia telangiectasia-mutated; BER, base excision repair; DNA-PKcs, DNA-protein kinase catalytic subunit; DSB, double-strand break; HR, homologous recombination; Mre11, mitotic recombination 11; Nbs1, Nijmegen breakage syndrome 1; NHEJ, non-homologous end-joining; PPi, inorganic pyrophosphate; SSB, single- strand break; XRCC1, x-ray cross complementing protein 1. (Adapted by permission from Macmillan Publishers Ltd: Nature Reviews Cancer (61), copyright 2010.)

13

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Table 1 – Key Clinical trials of PARP Inhibitors in Combination with Various Chemotherapeutic Agents and Related Hematologic Toxicities. CR = complete response, MR = moderate response, PR = partial response, SD = stable disease,

Downloaded from Combination Tumor Type Response Hematologic Toxicity Trial Design Author ManuscriptPublishedOnlineFirstonDecember26,2012;DOI:10.1158/1078-0432.CCR-12-0163 Author manuscriptshavebeenpeerreviewedandacceptedforpublicationbutnotyetedited.

Phase I: Advanced solid tumors and CR: none reported DLT: neutropenia Starting dose of: ABT-888 (Veliparib) + lymphomas PR: none reported thrombocytopenia; Veliparib 10mg PO BID D1-7+ clincancerres.aacrjournals.org Topotecan SD: 4/24 3 of 6 pts enrolled on DL1 Topotecan 1.2mg/m2/day D1-5 Kummar S et al. (2011) experienced DLTs

Gr 3/4: neutropenia 20.8% febrile neutropenia 8.3% thrombocytopenia 12.5%

Phase I: Advanced solid tumors CR: none reported DLT – febrile neutropenia Escalating doses of Veliparib PO ABT-888 (Veliparib) + MR: 2/28 leucopenia BID D1-14 + irinotecan Irinotecan PR: 5/28 neutropenia 100mg/m2 D1,8 on September 25, 2021. © 2012American Association for Cancer LoRusso P et al. (2011) SD: 10/28 Research.

Phase I: Breast cancer and advanced CR: none reported DLT – febrile neutropenia Escalating doses of Veliparib 50- ABT-888 (Veliparib) + solid tumors PR: 3/18 (TNBC; BRCA 150mg PO BID D1-4 + doxorubicin + mutation) doxorubicin 60mg/m2 D3 + cyclophosphamide SD: 8/18 (breast cancer) cyclophosphamide 600mg/m2 D3 Tan AR et al. (2011)

Phase I: Advanced solid tumors and CR: none reported Gr 2-4 lymphopenia 34.3% Escalating doses of Veliparib 20- ABT-888 (Veliparib) + lymphomas PR: 7/35 (6 BRCA mutation) 80mg PO daily for 7,14,or 21 days Oral Cyclophosphamide SD: 6/35 (3 BRCA mutation) + escalating doses of Kummar S et al. (2012) cyclophosphamide 50-100mg for 21 days

Phase II: Metastatic breast cancer (unconfirmed) Gr 3/4: neutropenia 11.4% Veliparib 40mg PO BID D1-7 ABT-888 (Veliparib) + CR: 1/24 thrombocytopenia 20% TMZ 150mg/m2 PO daily D1-5 Temozolomide (TMZ) PR: 2/24 Isakoff SJ et al. (2010) SD: 7/24

Randomized, double-blind, Metastatic melanoma No statistically significant Gr 3/4 hematologic toxicities: 1:1:1 randomization to: placebo-controlled trial improvement in PFS or OS TMZ alone 38% Placebo BID + TMZ, or ABT-888 (Veliparib) + Veliparib 20mg + TMZ 54% Veliparib 20mg BID + TMZ, or Temozolomide (TMZ) vs. Veliparib 40mg + TMZ 57% Veliparib 40mg BID + TMZ TMZ alone Middleton M et al. (2011) Downloaded from Author ManuscriptPublishedOnlineFirstonDecember26,2012;DOI:10.1158/1078-0432.CCR-12-0163 Author manuscriptshavebeenpeerreviewedandacceptedforpublicationbutnotyetedited.

Phase I: Stage I: advanced solid CR: 1/32 No DLTs Stage I:TMZ 100mg/m2/day + AG-014699/PF-01367338 tumors PR: 2/32 escalating doses rucaparib to (Rucaparib) + SD: 7/32 Myelosuppresion seen in 13% PARP inhibitory dose (PID) clincancerres.aacrjournals.org Temozolomide (TMZ) Stage II: chemo-naïve pts with rucaparib 18mg/m2/day Plummer R et al. (2008) metastatic melanoma + TMZ 200mg/m2/day Stage II: Rucaparib fixed at PID + TMZ escalating doses up to 200mg/m2/day Phase I: Advanced solid tumors CR: 0/21 DLT: febrile neutropenia Starting dose of: AZD-2281 (Olaparib) + PR: 2/21 (1 BRCA mutation) thrombocytopenia; Olaparib 100mg PO BID D1-4 Gemcitabine + SD: none reported 2 of first 3 pts enrolled on DL1 Gemcitabine 500mg/m2 D3-10 Cisplatin experienced DLTs Cisplatin 60mg/m2 D3 Giaccone G et al. (2010) on September 25, 2021. © 2012American Association for Cancer Gr 3/4: neutropenia 61% Research. lymphopenia 61% thrombocytopenia 57% Phase I: BRCA-1/2 mutation carriers CR: none reported DLT: thrombocytopenia and Escalating doses of olaparib AZD-2281 (Olaparib) + breast and ovarian cancer PR: 8/23 (ovarian cancer) delayed neutropenic recovery continuously or D1-7 + escalating Carboplatin 3/4 (breast cancer) doses of carboplatin (AUC 3-5) on Lee J et al. (2011) SD: 11/23 (ovarian cancer) Gr 3/4: Anemia 10% D2 or D8 1/1 (breast cancer) Thrombocytopenia 20% Neutropenia 17% Phase I: Advanced solid tumors; CR: 0/40 DLT: neutropenia, Escalating doses of AZD-2281 (Olaparib) + Expansion phase in chemo- PR: 2/40 thrombocytopenia Olaparib 10-100mg PO BID D1-7 Dacarbazine naïve stage III/IV melanoma SD: 8/40 + Dacarbazine 600-800mg/m2 D1 Khan OA et al. (2011) Gr 3/4: Anemia 5% Neutropenia 22.5% Thrombocytopenia 7.5% Phase I/II: Triple negative breast CR: none reported DLT: neutropenia requiring dose Olaparib 200mg PO BID + AZD-2281 (Olaparib) + cancer PR: 7/19 delay and GCSF prophylaxis Paclitaxel 90mg/m2 IV weekly for Paclitaxel SD: none reported 3 of 4 weeks; Dent RA et al. (2010)

Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Figure 1:

PARP1 DNA strand break detection by PARP1

Binding of PARP1 to DNA strand breaks

+ ADP-ribose NAD AMP  PPi PARP PARP1 inhibitors Nicotinamide ATP  PRPP

Poly(ADP-ribosyl)ation of PARP1

Transient recruitment and noncovalent and covalent modifications of various proteins

Histone DNA-PKcs, XRCC1 ATMMRE11 Topoisomerase I H1 Ku70 and Ku80

Recruitment Altered DSB repair DSB repair HR and Genomic Other of XRCC1 and chromatin by NHEJ by HR and restarting maintenance unknown DNA ligase III binding checkpoint of collapsed targets to SSBs and during DNA activation replication repair by BER damage and forks transcription

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research. Author Manuscript Published OnlineFirst on December 26, 2012; DOI: 10.1158/1078-0432.CCR-12-0163 Author manuscripts have been peer reviewed and accepted for publication but have not yet been edited.

Molecular Pathways: Targeting PARP in Cancer Treatment

Khanh Do and Alice Chen

Clin Cancer Res Published OnlineFirst December 26, 2012.

Updated version Access the most recent version of this article at: doi:10.1158/1078-0432.CCR-12-0163

Author Author manuscripts have been peer reviewed and accepted for publication but have not yet been Manuscript edited.

E-mail alerts Sign up to receive free email-alerts related to this article or journal.

Reprints and To order reprints of this article or to subscribe to the journal, contact the AACR Publications Subscriptions Department at [email protected].

Permissions To request permission to re-use all or part of this article, use this link http://clincancerres.aacrjournals.org/content/early/2012/12/22/1078-0432.CCR-12-0163. Click on "Request Permissions" which will take you to the Copyright Clearance Center's (CCC) Rightslink site.

Downloaded from clincancerres.aacrjournals.org on September 25, 2021. © 2012 American Association for Cancer Research.