Next Article in Journal
Actinic Keratoses: A Prospective Pilot Study on a Novel Formulation of 4% 5-Fluorouracil Cream and a Review of Other Current Topical Treatment Options
Next Article in Special Issue
Immunomodulatory Effects of Endoscopic Ultrasound-Guided Thermal Ablation in Patients with Pancreatic Ductal Adenocarcinoma
Previous Article in Journal
Metaplastic Breast Carcinoma in U.S. Population: Racial Disparities, Survival Benefit of Adjuvant Chemoradiation and Future Personalized Treatment with Genomic Landscape
Previous Article in Special Issue
CDK4/6 Inhibitors in Pancreatobiliary Cancers: Opportunities and Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pancreatic Cancer: BRCA Targeted Therapy and Beyond

by
Fergus Keane
1,2,
Catherine A. O’Connor
1,2,
Wungki Park
1,2,3,
Thomas Seufferlein
4 and
Eileen M. O’Reilly
1,2,3,*
1
Department of Medicine, Memorial Sloan Kettering Cancer Center, New York, NY 10065, USA
2
David M. Rubenstein Center for Pancreatic Cancer Research, New York, NY 10065, USA
3
Department of Medicine, Weill Cornell Medical College, New York, NY 10065, USA
4
Department of Internal Medicine, Ulm University Hospital, 89081 Ulm, Germany
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(11), 2955; https://doi.org/10.3390/cancers15112955
Submission received: 20 April 2023 / Revised: 18 May 2023 / Accepted: 26 May 2023 / Published: 28 May 2023

Abstract

:

Simple Summary

Pancreatic cancer is associated with poor outcomes for several reasons, including diagnosis at an advanced stage, the absence of effective screening for the diagnosis, and resistance to treatments. Pancreatic cancers associated with BRCA1/2 mutations have emerged as a distinct subgroup with sensitivity to other treatments and, in some cases, durable responses. Furthermore, beyond BRCA1/2 mutations, there is increasing recognition that other gene mutations may behave in a similar manner. The focus of this review is to discuss recent developments in the management of BRCA-associated pancreatic cancer, emerging therapeutic strategies, and future directions for this subgroup of patients.

Abstract

Pancreatic ductal adenocarcinoma (PDAC) is projected to become the second leading cause of cancer-related death in the US by 2030, despite accounting for only 5% of all cancer diagnoses. Germline gBRCA1/2-mutated PDAC represents a key subgroup with a favorable prognosis, due at least in part to additional approved and guideline-endorsed therapeutic options compared with an unselected PDAC cohort. The relatively recent incorporation of PARP inhibition into the treatment paradigm for such patients has resulted in renewed optimism for a biomarker-based approach to the management of this disease. However, gBRCA1/2 represents a small subgroup of patients with PDAC, and efforts to extend the indication for PARPi beyond BRCA1/2 mutations to patients with PDAC and other genomic alterations associated with deficient DNA damage repair (DDR) are ongoing, with several clinical trials underway. In addition, despite an array of approved therapeutic options for patients with BRCA1/2-associated PDAC, both primary and acquired resistance to platinum-based chemotherapies and PARPi presents a significant challenge in improving long-term outcomes. Herein, we review the current treatment landscape of PDAC for patients with BRCA1/2 and other DDR gene mutations, experimental approaches under investigation or in development, and future directions.

1. Introduction

By 2030, pancreatic ductal adenocarcinoma (PDAC) is projected to become the second-leading cause of cancer-related mortality in the US [1]. The 5-year survival rate for PDAC is increasing and is now 12%, up from 5–6% in the previous decade [2]. Approximately 80% of new diagnoses are in the setting of advanced or metastatic disease [3], and in those patients who do undergo curative-intent surgery, disease recurrence is observed in up to 80% [4]. Until recently, precision medicine approaches to the management of PDAC have demonstrated modest incremental benefits; however, recent developments have given cause for renewed hope for specific subsets of patients.
The Pancreas Olaparib Ongoing (POLO) trial was the first phase III trial to establish a biomarker-based approach to the management of PDAC for patients with germline gBRCA1/2 mutations [5,6], highlighting BRCA-associated PDAC as a distinct biological group with the potential for a personalized approach to treatment. Stadler et al. reported a pan-cancer analysis of almost 12,000 patients at Memorial Sloan Kettering Cancer Center (MSK) across over 50 tumor types who underwent comprehensive germline testing [7]. The OncoKB Precision Oncology Knowledge Base [8] was used to define the actionability of identified gene alterations, and 19.6% of PDAC cases were identified to harbor a pathogenic or likely pathogenic germline variant. Earlier, Lowery and colleagues carried out comprehensive germline testing on 615 patients with exocrine pancreatic neoplasms and identified pathogenic germline variants in 19.8% [9]. In both studies, pathogenic or likely pathogenic variants in BRCA1/2 and other genes associated with homologous recombination accounted for a substantial proportion of the variants identified. With the National Comprehensive Cancer Network (NCCN) [10] and American Society of Clinical Oncology (ASCO) guidelines [11] recommending both germline and somatic testing for PDAC, there is potential for the incorporation of precision medicine approaches in a greater number of patients with PDAC than before. Herein, we review recent advances and future directions in the approach to the management of patients with PDAC, with a focus on BRCA1/2 and other key HR gene mutations.

2. BRCA and DNA Damage Repair Pathways

Based on the type and frequency of structural variations identified by whole exome sequencing of a large cohort of PDAC tumor samples, Waddell et al. proposed four distinct PDAC subtypes: stable, locally rearranged, scattered, and unstable [12]. The unstable subtype is characterized by recurrent structural variation events owing to germline and somatic mutations of genes known to be involved in DNA damage repair (DDR), including BRCA1, BRCA2, PALB2, RAD51C, RAD51D, and ATM. These genes are relevant for the normal function of homologous recombination (HR) repair. Germline and somatic alterations in these key genes are now recognized to potentially lead to a HR deficient (HRD) phenotype within a tumor [13], which results in deficient double-stranded DNA break repair [14,15]. The roles of the BRCA1 and BRCA2 genes in encoding proteins required for HR are well described [16]. More recently, the function of PALB2 as a “binding protein” for BRCA2, localizing BRCA2 to sites of DNA breaks [17], has also been well supported. Together, BRCA1/2 and PALB2 are recognized as “core” genes that, when mutated, confer HRD in most settings of PDAC. Other “non-core” DDR variants and their role in inducing HR are less well defined but of growing interest.

2.1. Prevalence of BRCA and HRD in Pancreatic Cancer

The initial analysis of the Know Your Tumor registry trial identified mutations in the HR-DDR (DNA damage repair) pathway in 25% of 820 patients with PDAC, subdivided into three groups: Group 1: BRCA1/2 and PALB2; Group 2: ATM/ATR/ATRX; and Group 3: BAP1, BARD1, BRIP1, CHEK1/2, RAD50/51/51B, or FANCA/C/D2/E/F/G/L [18]. While gBRCA1/2 and gPALB2 mutations are identified in 5–6% of an unselected PDAC population [14], rates of gBRCA mutations are known to be enriched in certain populations, including those of Ashkenazi Jewish heritage (5–16%), patients with a family history of pancreatic cancer (5–19%), and those with a family history of ovarian or breast cancer (5–10%) [19,20,21,22]. A retrospective analysis of the first 2206 patients screened for enrollment in the POLO trial demonstrated an overall prevalence of BRCA1/2 mutations of 7.2% (5.8% after exclusion of populations known to be enriched with people of Ashkenazi Jewish heritage) and identified substantial geographic as well as some racial variability in the prevalence of gBRCA mutations amongst those screened, acknowledging a degree of selection bias given that almost 20% of patients enrolled in this trial had a prior documented gBRCA mutation [23].
While most patients with PDAC and an associated HRD phenotype have a pathogenic germline variant, a minority (2–4%) harbor a somatic-only HRD alteration. The widespread use of next-generation sequencing (NGS), as well as improvements in the sequencing assays utilized, has permitted more comprehensive identification of alterations in DDR genes. Waddell et al.’s whole genome sequencing and copy number variation analysis on 100 patients with PDAC identified two with a somatic BRCA1 mutation and three with a somatic BRCA2 alteration [12]. Work from our group demonstrated somatic HRD mutations in 4% of our cohort with PDAC and observed similarly favorable outcomes relative to those with germline HRD who received first-line platinum-based chemotherapy [14].

2.2. BRCAness Phenotype

While pathogenic variants in the DDR pathway have the potential to result in HRD and impaired double-stranded DNA break repair, as exemplified by BRCA1/2 mutations, this does not appear to be the case for all DDR pathogenic variants. The concept of BRCAness or HRD phenotype, defined as double-strand break repair deficiency in the absence of a BRCA1/2 variant [24,25], is of growing interest across tumor types, including PDAC. Tumors exhibiting BRCAness appear to respond favorably to DNA-damaging therapies such as platinum-based chemotherapy and poly-ADP ribose polymerase inhibitors (PARPi). There is an increasing recognition that, beyond focused germline and somatic profiling of HR-associated genes, evaluating genomic signatures through validated HRD scores may represent a more inclusive method of determining tumors that may respond to DDR-targeted treatments. These methods are more established in the context of breast and ovarian cancers [26,27]. Candidate variants of particular interest in PDAC and whether they confer a BRCAness phenotype include PALB2, RAD51, RAD50, CHEK2, ATM, BRIP, and BLM variants.
Several methods for determining HRD status are in various stages of development. DNA “scars”, referring to gross chromosomal abnormalities and mutational signatures, are specific genomic features that are characteristic of HRD [28]. The three key types of chromosomal aberrations associated with HRD are telomeric allelic imbalance (TAI), large-scale state transitions (LSTs), and loss of heterozygosity (LOH) [29,30,31]. In triple-negative breast cancer, a composite score of TAI, LST, and LOH is a robust biomarker for identification of HRD status, demonstrating improved pathologic complete response rates to platinum-based chemotherapy in those with a high combined score, regardless of BRCA1/2 status [26]. This strategy of HRD detection is the basis of Myriad’s MyChoice HRD assay, which, in addition to testing for pathogenic BRCA1/2 mutations, also evaluates Genomic Instability Scores (GIS) based on TAI, LST, and LOH and is now an FDA-approved companion test to determine eligibility for olaparib and niraparib [32,33].
Beyond genomic scars, stereotyped mutational patterns based on the accumulation of single-base substitutions (SBS), insertion deletions (indels), and rearrangements are observed in tumors with HRD [28,34]. Signature 3, a SBS pattern associated with large deletions and microhomology, is observed regularly in BRCA1/2 and PALB2-mutated PDAC [12,35] and has been demonstrated to predict response to platinum-based chemotherapy in patients with PDAC [15]. Other tests based on NGS have been developed, most notably HRDetect, which employs a weighted model of mutational signatures based on whole genome sequencing to detect HRD in breast, ovarian, and pancreatic cancer [36]. A recent meta-analysis of over 60 studies and >21,000 patients with PDAC identified the prevalence of HRD to be 14.5–16.5% based on targeted NGS but 24–44% by whole genome or whole exome sequencing [37], highlighting the need to clarify the definition and optimal method of detection of HRD in PDAC.
In one large study of 391 patients with PDAC, including 49 with gBRCA1/2 or gPALB2 pathogenic variants, HRD classifiers, including (i) GIS using Myriad’s MyChoice assay, (ii) Signature 3, (iii) HRDetect, and (iv) structural variant burden, were applied [15]. In this study, GIS scores had a sensitivity of 91% and a specificity of 83% for the identification of HRD, and improved clinical outcomes with platinum chemotherapy were associated with higher GIS scores. Most notably, HRDetect appeared to outperform both Sig3 and GIS scores in its ability to identify HRD PDAC, with high HRDetect scores present in up to 10% of patients who did not have a germline HR gene mutation. A HRDetect score >0.7 predicted gBRCA1/PALB2 deficiency with a sensitivity of 98% and specificity of 100%. In addition, high HRDetect scores predicted improved survival in patients receiving platinum-based chemotherapy. A notable feature of this study was that pathogenic variants in CHEK2 and ATM (N = 2 and N = 6, respectively) did not result in HRD by any of the above criteria. Identifying which, if any, of the DDR gene mutations result in a HRD phenotype and would thus present similar therapeutic actionability as BRCA1/2 mutations is of considerable interest. As an example, our group at MSK recently evaluated 46 patients with PDAC and germline or somatic ATM variants [38] and determined that while ATM-mutated PDAC is associated with a favorable overall survival (OS) relative to genomically unselected PDAC, pathogenic variants in ATM did not appear to confer a HRD signature [38]. Several studies are underway evaluating other candidate HRD genes and whether they confer a HRD phenotype, as well as additional HRD/BRCAness detection methods, including integrating DNA- and RNA-based HRD detection methods together [39]. The selected tests for evaluating HRD status are summarized in Table 1.

3. Treatment of BRCA and HRD-Mutated Pancreatic Cancer

3.1. Platinum-Based Chemotherapy

Platinum chemotherapies can crosslink purine bases in DNA, interrupting DNA transcription and replication, which leads to the accumulation of DNA damage due to a limited DNA damage repair capacity. The body of evidence supporting the use of platinum-based therapy in patients harboring BRCA1/2 mutations is ever-growing, including in PDAC, with several studies demonstrating superior outcomes with platinum-based chemotherapy in patients with BRCA1/2 mutations compared with unselected PDAC cohorts. Data suggests that the pleiotropic effects of mutant BRCA1/2 are tumor-lineage dependent and that the therapeutic relevance of gBRCA1/2 mutations is of most relevance in “BRCA-associated cancer types”, namely pancreatic, breast, ovarian, and prostate cancer [40].
In the Know Your Tumor Program, for patients with advanced PDAC who received platinum-based chemotherapy, the mOS was 1.27 years for HR-DDR mutated patients (N = 53) versus 1.45 years for patients with proficient HR-DDR status (N = 258), representing a statistically significant and clinically meaningful difference (p = 0.000072; HR 0.44; 95% CR 0.29 to 0.66). Data from our group at MSK demonstrated that in 50 patients with advanced PDAC and a germline or somatic HR-associated mutation treated with systemic therapy, a median progression-free survival (mPFS) advantage was observed for those in receipt of a platinum regimen versus a non-platinum regimen (12.6 vs. 4.4 months, HR 0.44, 96% CI 0.29–0.67, p < 0.01) [14]. This analysis also demonstrated that patients with biallelic HR mutations had higher genomic instability and derived further benefit from front-line platinum therapy, with a significantly improved mPFS of 13.3 months (95% CI, 0.26–0.70) in platinum-treated patients versus 3.8 months (95% CI, 2.79—not reached (NR); p < 0.0001) in non-platinum-treated patients. Momtaz et al. reported on a large cohort of patients with PDAC and germline or somatic BRCA1/2 mutations [41]. Of 81 patients with stage IV disease, the mOS for patients who received upfront platinum-based therapy (N = 65) was 23 months (95% CI, 19–26), versus 29 months (95% CI, 19 months to NR) for those who did not receive frontline platinum-based therapy (N = 14). Of the 14 patients who did not receive platinum-based therapy in the first line, 10 did receive it in the second line, and favorable responses were noted (a PR was noted in 7 patients and SD in 1 patient as the best response), with a median duration of therapy of 11 months (range, 1–35). Notably, amongst those with metastatic disease who did receive front-line platinum-based therapy, the mOS for patients with biallelic status (N = 39) was 26 months (95% CI, 20–52 months) and 8.66 months (95% CI, 6.2—NR) for those with monoallelic status (N = 4). Beyond the advanced PDAC setting, several studies have also demonstrated an advantage associated with the receipt of platinum-based therapy in the neoadjuvant setting for patients with BRCA1/2-mutated PDAC [42,43].
To enhance response and potentially delay resistance, O’Reilly et al. evaluated the role of veliparib in combination with cisplatin and gemcitabine in untreated metastatic and locally advanced PDAC with BRCA1/2 or PALB2 mutations [44]. Patients were randomized to receive cisplatin and gemcitabine, with or without veliparib. Delayed emergence of resistance and improved survival were not observed with the addition of veliparib; however, high response rates in both groups were observed (74.1% in the experimental arm and 65.2% in the control arm, p = 0.55), as well as a favorable OS signal from first-line cisplatin-based therapy in both arms (15.5 months and 16.4 months, respectively, p = 0.73), establishing cisplatin and gemcitabine as a standard approach in gBRCA1/2- and PALB2- mutated PDAC and an alternative to FOLFIRINOX.

3.2. Poly-ADP Ribose Polymerase Inhibitors

The PARP enzymes have a key role in base excision repair of single-stranded DNA breaks [45], and inhibition of these enzymes results in the accumulation of single-stranded breaks (SSBs), leading to replication fork collapse, and the generation of double-stranded breaks (DSBs). As discussed earlier, DSBs rely on HRR for repair and to avoid cell cycle arrest and apoptosis [46]. Thus, in patients with mutations in HRR genes, PARPi can be employed to create a state of synthetic lethality in tumors [46]. Select recently-completed and ongoing trials evaluating the role of PARPi for BRCA-mutated and HRD PDAC are summarized in Table 2 and Table 3. In a multicenter phase 2 study of patients with breast, prostate, ovarian, and PDAC tumors with gBRCA1/2 mutations, the use of oral PARPi olaparib demonstrated an overall response rate of 26.2% (95% CA 21.3–31.6) across all patients [47]. Specifically, in the PDAC cohort, which included 23 patients, all with advanced disease, an overall response rate of 21.7% was observed, including one complete response and four partial responses.
Olaparib was prospectively evaluated in the phase III Pancreas Olaparib Ongoing (POLO) trial, which enrolled 154 patients with gBRCA1/2 mutations and advanced or metastatic PDAC [5,6]. Patients were required to have demonstrated disease response or stability following a minimum of 16 weeks of platinum-based chemotherapy. The primary end point of PFS was met, with a mPFS of 7.4 months in the olaparib arm compared with 3.8 months in the placebo arm (HR 0.53, p = 0.004). Following this data, FDA approval for olaparib was granted in December 2019. The secondary endpoint of OS, however, was not met (median 19.0 months versus 19.2 months; HR 0.83, 95% CI, 0.56–1.22, p = 0.3487). OS at 36 months was 33.9% for olaparib and 17.8% for placebo, and mPFS2, defined as median time from randomization to second disease progression, was 16.9 months for olaparib versus 9.3 months for placebo (hazard ratio, 0.66; p = 0.0613), indicating a trend toward benefit for olaparib, though not reaching statistical significance. Notably, quality of life scores between the arms were equivalent. POLO was the first trial to establish a biomarker-based approach to the management of PDAC in patients with gBRCA mutations; however, the lack of OS advantage and the placebo control arm have led to questions regarding the magnitude and importance of the observed PFS benefit.
Reiss et al. conducted a phase 2 trial evaluating PARPi rucaparib in the post-platinum maintenance setting in patients with PDAC harboring germline or somatic BRCA1/2 or PALB2 alterations [48]. They reported a mPFS of 13.1 months (95% CI, 4.4–21.8) and a mOS of 23.5 months (95% CI, 20–27). An ORR of 41.7% was observed, which included 3 complete responses and 12 partial responses. Rucaparib was also evaluated in the phase 2 RUCAPANC study, which included patients with germline or somatic BRCA1/2 mutations [49]. While two-thirds of patients had received prior platinum, platinum sensitivity and prior platinum were not mandated. The study was terminated early based on a modest ORR of 15.8%, and notably, none of the patients who had platinum-resistant disease at the time of enrollment had an objective response. Indeed, sensitivity to platinum has emerged as a key predictive biomarker for sensitivity to PARPi.
In a phase 2 study of veliparib in 16 patients with metastatic or locally advanced PDAC and gBRCA1/2 or PALB2 mutations, Lowery et al. reported no objective response, although stable disease for >4 months was noted in four patients [50]. Most patients enrolled in this study had platinum-resistant PDAC, which most likely explains the limited signal observed. Beyond gBRCA mutations, Javle and colleagues described the clinical outcomes of olaparib monotherapy in patients with PDAC and alterations in other DDR genes, including ATM, PALB2, ARID1A, sBRCA, PTEN, RAD51, CCNE, and FANCB. They identified that patients with platinum-sensitive PDAC had an improved mPFS (4.1 months, 95% CI, 3.6–7.8) over those with platinum-resistant PDAC (2.2 months, 95% CI, from 1.8 to not reached, p = 0.01) [51]. This benefit was also observed in the analysis of OS (10.5 vs. 5.4 months for platinum-sensitive versus platinum-resistant disease, respectively, p = 0.03). Based on this study, in addition to others, it is apparent that in PDAC, the role of PARPi is maximized in the platinum-sensitive maintenance setting.
Beyond their role in advanced disease, the utility of PARPi in the early-stage setting is under investigation in clinical trials. The APOLLO study (NCT04858334) is a randomized phase II study of adjuvant olaparib versus placebo in patients with pancreatic cancer and germline or somatic BRCA1/2 or PALB2 mutations in whom all curative intent standard treatment has been completed [52].

3.3. Resistance to PARP Inhibition

Despite the rationale and enthusiasm for the use of PARPi in patients with BRCA1/2 mutations (and other DDR gene alterations), primary resistance is observed in a notable proportion of patients, including those with platinum-sensitive PDAC. In the POLO trial, despite all patients enrolled having confirmed platinum sensitivity in the context of gBRCA1/2 mutations, approximately one fifth of patients had evidence of progression of disease on first assessment [5,6]. Similarly, in the phase II study by Reiss et al. evaluating maintenance rucaparib in the post-platinum maintenance setting for patients with germline or somatic BRCA1/2 or PALB2 pathogenic variants, 16% of patients experienced progression of disease within the first 8 weeks of treatment [48]. This builds upon the data from Golan and colleagues, in which patients with PDAC and gBRCA1/2 mutations had HRDetect performed on their pancreatic cancer tumor samples, and 12% did not have a HRD signature [15], which may account for primary resistance to PARPi in a proportion of patients.
Secondary acquired resistance to PARPi is common, and several mechanisms are postulated [53]. In several studies, hyperactivation of the ATR/CHK1 pathway, resulting in maintained genomic stability, has been suggested as a mechanism of PARPi resistance in BRCA-driven tumors [54,55], and forms the basis of several efforts to overcome resistance. This pathway is discussed in greater detail below. In addition, the emergence of reversion mutations, resulting in the restoration of HRR following treatment with PARPi, has been described. In a study by Pettitt et al., data from the literature pertaining to 308 reversion mutations in the setting of PARPi or platinum resistance identified in 91 patients were reported and suggest that for BRCA2 mutations, the position of the mutation may affect the risk of reversion [56]. In addition, they concluded that many reversions are predicted to encode highly immunogenic neopeptides, which may provide a potential opportunity for avoiding resistance, for example, through the integration of immunotherapy. Specific to PDAC, preclinical models of ATM-deficient pancreatic cancer cells identified an association between the upregulation of alternative-end joining via upregulation of drug efflux transporters and detoxication enzymes and resistance to PARP inhibition [57]. However, the mechanisms of PARPi resistance in pancreatic cancer appear complex and, as of yet, are less well defined [58]. Ongoing efforts to deepen the responses to PARPi as well as to delay resistance are underway, with several combination strategies under investigation in PDAC (Table 2).

4. Combination Treatment Strategies

4.1. PARP Inhibition and Chemotherapy

As discussed earlier, our group evaluated veliparib in combination with cisplatin and gemcitabine in untreated metastatic and locally advanced PDAC with BRCA1/2 or PALB2 mutations in a randomized phase 2 study [43] and demonstrated that while a favorable overall survival signal from first-line cisplatin and gemcitabine chemotherapy (+/− veliparib) was observed in both arms (15.5 months and 16.4 months, respectively, p = 0.73), improved survival and delayed resistance were not observed with the addition of veliparib in this trial. The SWOG S1513 trial was a randomized phase 2 study that evaluated veliparib with or without FOLFIRI in the second line for patients with stage IV PDAC [59]. Accrual was halted early, with 123 patients enrolled, due to the lack of benefit from the addition of veliparib. For the entire group, mOS was not improved with the addition of veliparib (5.4 vs. 6.5 months, HR 1.23, p = 0.28). Notably, for patients with HR-DDR defects (N = 22, 19%) compared to those without HR-DDR gene defects, mPFS and mOS were 7.3 vs. 2.5 months (p = 0.05) and 10.1 vs. 5.9 months (p = 0.17), respectively, with FOLFIRI alone, and 2.0 vs. 2.1 months (p = 0.62) and 7.4 vs. 5.1 months (p = 0.10), respectively, with veliparib plus mFOLFIRI. Both arms performed favorably compared with the overall group, indicating a potential role for irinotecan-based therapy in patients with such defects. In addition, an ongoing non-randomized phase 2 study is evaluating 5-FU and liposomal irinotecan in combination with rucaparib in patients with metastatic PDAC with BRCA1/2, PALB2, or other DDR-HRD genomic alterations who have not received systemic therapy in the metastatic setting (NCT03337087).

4.2. PARPi and Immune Checkpoint Inhibitors

In pre-clinical studies, PARPi have been demonstrated to enhance cancer-associated immunosuppression and upregulate PD-L1 expression [60]. In addition, the combination of PARPi and immune checkpoint blockade has demonstrated an accumulation of tumor neoantigens and activation of interferon pathways, resulting in a synergistic antitumoral effect [61,62] in pre-clinical studies. The phase 1b/2 PARPVAX study of niraparib plus nivolumab and niraparib plus ipilimumab in patients with platinum-sensitive PDAC was evaluated and compared to a historical reference [63]. Subgroup analysis of patients with BRCA1/2 or PALB2 variants demonstrated a mPFS of 3.7 months and a mOS of 12.2 months in the niraparib and nivolumab arm, versus a prolonged mPFS of 10.4 months and a mOS of 38 months in the niraparib plus ipilimumab arm, suggesting that CTLA-4 inhibitors may potentially have an enhanced benefit in combination with PARPi compared with PD-1 inhibitors, and this signal warrants further investigation.
The non-randomized phase 2 POLAR study (NCT04666740) is evaluating the combination of maintenance pembrolizumab and olaparib in patients with metastatic pancreatic cancer and HR deficiency (or exceptional response to platinum) post-platinum-based chemotherapy. The SWOG S2001 trial (NCT04548752) is a randomized study evaluating olaparib with/without pembrolizumab, also in the post-platinum maintenance setting, in patients with metastatic pancreatic cancer and gBRCA1/2 mutations. Several studies are evaluating PARPi and immunotherapy +/− other agents, for example, PD-1 inhibitor dostarlimab and PARPi niraparib in patients with advanced BRCA-mutated tumors, including PDAC (NCT04673448, NCT04493060), dostarlimab, niraparib, and radiotherapy (NCI04409002), PD-L1 inhibitor avelumab, PARPi talazoparib, and MEK inhibitor binimetinib in patients with locally advanced or metastatic RAS-mutated solid tumors (NCT03637491, closed), Beyond BRCA1/2 variants, the MAZEPPA trial is evaluating PARPi of olaparib or kinase inhibitor selumetinib plus PD-1 inhibitor durvalumab in patients with metastatic PDAC and BRCAness (NCT04348045).

4.3. PARPi and Anti-Angiogenic Agents

Agents inhibiting vascular epidermal growth factor (VEGF) families result in a hypoxic state and subsequent downregulation of HRR gene expression [64]. Series suggest that this hypoxic tumor microenvironment results in the functional inactivation of RAD51 and BRCA without a genetic alteration in these genes, with a resulting BRCAness phenotype [65,66]. Some promise has been observed in combining PARPi and VEGF inhibitors in both ovarian cancer and prostate cancer. A randomized phase 2 study of olaparib with or without the pan-VEGF inhibitor cediranib demonstrated a significant radiographic PFS advantage over olaparib alone in men with metastatic castrate-resistant prostate cancer [67]. In ovarian cancer, the phase 3 PAOLA-1/ENGOT-ov25 trial of olaparib and bevacizumab in the post-platinum maintenance setting demonstrated a significant PFS advantage in those with BRCA and non-BRCA HRD mutations [32]. Olaparib combined with the anti-VEGF agent cediranib is under investigation in advanced solid tumors, including PDAC, in an ongoing phase 2 study (NCT02498613).

4.4. PARPi and Other Agents

ATR kinase-mediated DDR pathways are thought to promote tumor survival during PARP inhibition, and in pre-clinical studies, the ATR-inhibitor AZD6738 (ceralasertib) in combination with olaparib has demonstrated synergistic tumor inhibition [68]. A phase 2 study is currently underway investigating the role of AZD6738 with or without olaparib in patients with advanced solid tumors, including PDAC, who harbor ARID1A expression or ATM loss/mutations (NCT03682289). In addition, pre-clinical data from one study demonstrated that cells resistant to PARP inhibition showed evidence of elevated RAS/MAP kinase signaling, and a signal for partial reversal with the addition of MEK inhibition was observed [69]. On this basis, PARPi and MEK inhibition are being investigated both in the neoadjuvant (NCT04005690) and advanced settings (NCT03637491).

4.5. Immune Checkpoint Blockade in BRCA and HRD Pancreatic Cancer

PDAC is considered a prototypical immunogenically “cold” tumor, owing to an inherently immunosuppressive, hypoxic tumor microenvironment as well as a dense surrounding stroma. Outside of the small proportion of patients with microsatellite-instable PDAC (approximately 1%), a role for immune checkpoint blockade has been challenging to support, with poor responses observed in a phase 2 study from our group [70]. However, greater genomic instability in patients with PDAC with biallelic loss of BRCA1 and BRCA2 has been demonstrated across several series [14,71], suggesting a possible role for the incorporation of immune checkpoint blockade in this subset of patients. Terraro et al. reported the results of a series of 12 patients with chemotherapy-refractory pancreaticobiliary cancers and pathogenic germline variants in HRD genes (specifically BRCA1, RAD51C, ATM, BRCA2, and RAD51D) who received dual checkpoint blockade with ipilimumab and nivolumab. Of the ten patients with PDAC (out of a total of 12; 10 PDAC, 1 cholangiocarcinoma, and 1 ampullary carcinoma), two had a complete response to therapy, one had a partial response, and two had stable disease [72], indicating a potential role for immune checkpoint blockade as a therapeutic strategy in BRCA-mutated and other HRD PDAC.

4.6. Targeting DNA Replication Stress

While much of the focus of novel therapeutic strategies for PDAC has been on targeting HR, defective DDR may result in “replication stress”, defined as the perturbation of error-free DNA replication and the slowing of DNA synthesis, resulting in genomic instability and oncogenic transformation [73]. The ATR kinase has an important role in the cellular response to replication stress [74]. Replication protein A (RPA) binds single-stranded DNA (ssDNA) and recruits ATR, which results in the downstream activation of ATR-interacting proteins and subsequently ATR and CHEK1 through phosphorylation [74]. ATR-CHEK1 activation results in s-G2 cell cycle arrest and activation of HR repair, as well as replication fork stabilization [73,74]. CHEK1 is negatively regulated by WEE1 and MYT1 via phosphorylation and results in the activation of WEE1 and degradation of CDC25A [75,76]. Attempts at targeting replication stress have led to the development of several compounds, many of which are in preclinical and clinical development, including ATR inhibitors (ATRi), CHEK1 inhibitors, and WEE1 inhibitors.
ATRi elimusertib was evaluated as a monotherapy in a phase 1 trial of 21 patients with advanced solid tumors (NCT03188965). An ORR of 19% (4 of 21) was observed, all 4 occurring in patients harboring alterations in or loss of ATM [77]. Intravenous ATRi berzosertib was evaluated as a monotherapy and in combination with carboplatin in 40 patients with advanced solid tumors (NCT02157792). One patient with metastatic colorectal cancer harboring ATM loss and an ARID1A mutation had a complete response; another patient with platinum-refractory and PARPi-resistant ovarian cancer harboring a gBRCA1 mutation had a partial response [78].
CHEK1/2i prexasertib has been evaluated as a monotherapy in two studies. In a phase 1 study (NCT01115790) of 45 patients with advanced solid tumors, 2 patients had a partial response, and 15 patients had stable disease as the best overall response, with a more favorable signal in patients with squamous histology. Thus, an expansion cohort, including patients with squamous histology only, was carried out and demonstrated partial responses in 15% of patients with anal cancer and 5% of patients with head and neck tumors [79]. In the second study (NCT02203513), a phase 2 study, 42 patients with ovarian cancer were enrolled. Of 24 patients without BRCA1/2 mutations, an ORR of 33% was observed, and among 18 patients harboring BRCA1/2 mutations, an ORR of 11% was observed [80,81]. WEE1i adavosertib was evaluated in combination with chemotherapy (gemcitabine, paclitaxel, carboplatin, or pegylated liposomal doxorubicin) in a phase II study in 94 patients with primary platinum-resistant gynecologic malignancies (NCT02272790) [82]. A promising ORR of 32% was observed. ZN-C3 has been evaluated in a phase 1 study (NCT04158336) of patients with high-grade serous endometrial carcinoma and resulted in a complete response in one patient and two partial responses in eleven patients in total [83]. Several studies combining chemotherapy with inhibitors of the pathway are underway, including gemcitabine, carboplatin, and berzosertib in advanced ovarian cancer (NCT02627443), and gemcitabine, nab-paclitaxel, and adavosertib in PDAC (not specific to patients with HRD) (NCT02194829—active, not recruiting).

4.7. Other Promising Strategies in BRCA and HRD

Inhibitors of polymerase theta (POLθ, which is encoded by POLQ) are gaining interest as an additional therapeutic strategy in HRD malignancies. POLθ is the key enzyme in microhomology-mediated end-joining (MMEJ), and several studies have demonstrated the potential of POLθ inhibitors in evasion of MMEJ and inducing synthetic lethality [84,85]. In a key recent study in BRCA-mutated breast cancer and PDAC cell lines, either novobiocin, an inhibitor of the POLθ ATPase domain, or ART558, an inhibitor of the POLθ polymerase domain, demonstrated activation of the cGAS/STING pathway. Activation of this pathway drives the expression of type I interferon response elements, including PD-L1, and increases CD8+ T-cell tumor infiltration and activation, as well as the activation of antigen-presenting dendritic cells [86]. In addition, the antitumor activity of novobiocin was augmented with the addition of PD-1 blockade in a BRCA2-deficient mouse model in this study. The first human study of novobiocin will begin accruing patients with tumors with alterations in DNA repair genes in 2023 (NCT05687110). Current and emergent targeted strategies are summarized in Figure 1.

5. Conclusions and Future Directions

With modest progress in survival outcomes for patients with PDAC over the last two decades, the emergence of a personalized, biomarker-based approach to the management of PDAC is promising. Patients with BRCA1/2 mutations represent a biologically distinct subgroup with a favorable predictive profile compared with an unselected PDAC population. In addition, patients with HRD status are emerging as a group that may also benefit from the incorporation of a similar personalized approach. However, despite developments in the management of patients with BRCA1/2- and HRD-associated PDAC, resistance is a critical limitation. Future efforts should focus on the continued development of accurate HRD detection methods. In addition, well-designed, adequately powered, randomized clinical trials of novel combination approaches to deepen responses and overcome resistance are imperative. Lastly, the incorporation of translational, interdisciplinary science and, most importantly, universal patient access to somatic and germline testing as recommended by international guidelines should be prioritized.

Author Contributions

Conceptualization, E.M.O. and F.K.; methodology, F.K. and E.M.O.; formal analysis, F.K., C.A.O., W.P., T.S. and E.M.O.; data curation, F.K. and C.A.O.; writing—original draft preparation, F.K. and C.A.O.; writing—review and editing, F.K., C.A.O., W.P., T.S. and E.M.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

F.K. and C.O. have no conflicts of interest. W.P.: consultancy: Astellas; research funding to MSK: Merck, Astellas, and Miracogen; funding relevant to this MS: NIH Pancreas SPORE (1P50CA257881-01A1), NIH K12 CA184746 Paul Calabresi Career Development Award for Clinical Oncology, and MSK Parker Institute for Cancer Immunotherapy (PICI) Pilot Grant Award; funding is not relevant to this MS: Breakthrough Cancer—Conquering KRAS for Pancreatic Cancer, Society of MSKCC Research Grant Award—iBTC, TimIOs: Society of Immunotherapy for Cancer (SITC), Merck Investigator Studies Program: Investigator-Initiated Trial. T.S.: honoraria for scientific presentations and advisory boards from Astra Zeneca and Servier. E.O.: research funding to MSK from Genentech/Roche, BioNTech, AstraZeneca, Arcus, Elicio, Parker Institute, NIH/NCI, and Pertzye; consulting/DSMB: Cytomx Therapeutics (DSMB), Rafael Therapeutics (DSMB), Seagen, Boehringer Ingelheim, BioNTech, Ipsen, Merck, IDEAYA, Fibrogen, Merus, Silenseed, Novartis, AstraZeneca, Novartis, BMS, Noxxon, BioSapien, Cend Therapeutics, Thetis, Autem, ZielBio, Tempus, Agios (spouse), Genentech-Roche (spouse), and Eisai (spouse).

References

  1. Rahib, L.; Smith, B.D.; Aizenberg, R.; Rosenzweig, A.B.; Fleshman, J.M.; Matrisian, L.M. Projecting cancer incidence and deaths to 2030: The unexpected burden of thyroid, liver, and pancreas cancers in the United States. Cancer Res. 2014, 74, 2913–2921. [Google Scholar] [CrossRef] [PubMed]
  2. Siegel, R.L.; Miller, K.D.; Wagle, N.S.; Jemal, A. Cancer statistics, 2023. CA A Cancer J. Clin. 2023, 73, 17–48. [Google Scholar] [CrossRef] [PubMed]
  3. Mizrahi, J.D.; Surana, R.; Valle, J.W.; Shroff, R.T. Pancreatic cancer. Lancet 2020, 395, 2008–2020. [Google Scholar] [CrossRef] [PubMed]
  4. Dunne, R.F.; Hezel, A.F. Genetics and Biology of Pancreatic Ductal Adenocarcinoma. Hematol. Clin. N. Am. 2015, 29, 595–608. [Google Scholar] [CrossRef]
  5. Golan, T.; Hammel, P.; Reni, M.; Van Cutsem, E.; Macarulla, T.; Hall, M.J.; Park, J.-O.; Hochhauser, D.; Arnold, D.; Oh, D.-Y.; et al. Maintenance Olaparib for Germline BRCA-Mutated Metastatic Pancreatic Cancer. N. Engl. J. Med. 2019, 381, 317–327. [Google Scholar] [CrossRef]
  6. Kindler, H.L.; Hammel, P.; Reni, M.; Van Cutsem, E.; Macarulla, T.; Hall, M.J.; Park, J.O.; Hochhauser, D.; Arnold, D.; Oh, D.-Y.; et al. Overall Survival Results from the POLO Trial: A Phase III Study of Active Maintenance Olaparib Versus Placebo for Germline BRCA-Mutated Metastatic Pancreatic Cancer. J. Clin. Oncol. 2022, 40, 3929–3939. [Google Scholar] [CrossRef]
  7. Stadler, Z.K.; Maio, A.; Chakravarty, D.; Kemel, Y.; Sheehan, M.; Salo-Mullen, E.; Tkachuk, K.; Fong, C.J.; Nguyen, B.; Erakky, A.; et al. Therapeutic Implications of Germline Testing in Patients with Advanced Cancers. J. Clin. Oncol. 2021, 39, 2698–2709. [Google Scholar] [CrossRef]
  8. Chakravarty, D.; Gao, J.; Phillips, S.; Kundra, R.; Zhang, H.; Wang, J.; Rudolph, J.E.; Yaeger, R.; Soumerai, T.; Nissan, M.H.; et al. OncoKB: A Precision Oncology Knowledge Base. JCO Precis. Oncol. 2017, 2017, 1–16. [Google Scholar] [CrossRef]
  9. Lowery, M.A.; Wong, W.; Jordan, E.J.; Lee, J.W.; Kemel, Y.; Vijai, J.; Mandelker, D.; Zehir, A.; Capanu, M.; Salo-Mullen, E.; et al. Prospective Evaluation of Germline Alterations in Patients with Exocrine Pancreatic Neoplasms. Gynecol. Oncol. 2018, 110, 1067–1074. [Google Scholar] [CrossRef]
  10. National Comprehensive Cancer Network. Pancreatic Cancer, Version 2. 2021. Available online: https://www.nccn.org/professionals/physician_gls/pdf/pancreatic.pdf (accessed on 17 April 2023).
  11. Sohal, D.P.S.; Kennedy, E.B.; Cinar, P.; Conroy, T.; Copur, M.S.; Crane, C.H.; Garrido-Laguna, I.; Lau, M.W.; Johnson, T.; Krishnamurthi, S.; et al. Metastatic Pancreatic Cancer: ASCO Guideline Update. J. Clin. Oncol. 2020, 38, 3217–3230. [Google Scholar] [CrossRef]
  12. Waddell, N.; Pajic, M.; Patch, A.-M.; Chang, D.K.; Kassahn, K.S.; Bailey, P.; Johns, A.L.; Miller, D.; Nones, K.; Quek, K.; et al. Whole genomes redefine the mutational landscape of pancreatic cancer. Nature 2015, 518, 495–501. [Google Scholar] [CrossRef]
  13. Jasin, M.; Rothstein, R. Repair of Strand Breaks by Homologous Recombination. Cold Spring Harb. Perspect. Biol. 2013, 5, a012740. [Google Scholar] [CrossRef]
  14. Park, W.; Chen, J.; Chou, J.F.; Varghese, A.M.; Yu, K.H.; Wong, W.; Capanu, M.; Balachandran, V.; McIntyre, C.A.; El Dika, I.; et al. Genomic Methods Identify Homologous Recombination Deficiency in Pancreas Adenocarcinoma and Optimize Treatment Selection. Clin. Cancer Res. 2020, 26, 3239–3247. [Google Scholar] [CrossRef]
  15. Golan, T.; O’kane, G.M.; Denroche, R.E.; Raitses-Gurevich, M.; Grant, R.C.; Holter, S.; Wang, Y.; Zhang, A.; Jang, G.H.; Stossel, C.; et al. Genomic Features and Classification of Homologous Recombination Deficient Pancreatic Ductal Adenocarcinoma. Gastroenterology 2021, 160, 2119–2132.e9. [Google Scholar] [CrossRef]
  16. Holter, S.; Borgida, A.; Dodd, A.; Grant, R.; Semotiuk, K.; Hedley, D.; Dhani, N.; Narod, S.; Akbari, M.; Moore, M.; et al. Germline BRCA Mutations in a Large Clinic-Based Cohort of Patients with Pancreatic Adenocarcinoma. J. Clin. Oncol. 2015, 33, 3124–3129. [Google Scholar] [CrossRef]
  17. Xia, B.; Sheng, Q.; Nakanishi, K.; Ohashi, A.; Wu, J.; Christ, N.; Liu, X.; Jasin, M.; Couch, F.J.; Livingston, D.M. Control of BRCA2 Cellular and Clinical Functions by a Nuclear Partner, PALB2. Mol. Cell 2006, 22, 719–729. [Google Scholar] [CrossRef]
  18. Pishvaian, M.J.; Blais, E.M.; Brody, J.R.; Rahib, L.; Lyons, E.; De Arbeloa, P.; Hendifar, A.; Mikhail, S.; Chung, V.; Sohal, D.P.; et al. Outcomes in Patients with Pancreatic Adenocarcinoma with Genetic Mutations in DNA Damage Response Pathways: Results From the Know Your Tumor Program. JCO Precis. Oncol. 2019, 3, 1–10. [Google Scholar] [CrossRef]
  19. Salo-Mullen, E.E.; O’Reilly, E.M.; Kelsen, D.P.; Ashraf, A.M.; Lowery, M.A.; Yu, K.h.; Reidy, D.L.; Epstein, A.S.; Lincoln, A.; Saldia, A.; et al. Identification of germline genetic mutations in patients with pancreatic cancer. Cancer 2015, 121, 4382–4388. [Google Scholar] [CrossRef]
  20. Özcelik, H.; Schmocker, B.; Di Nicola, N.; Shi, X.-H.; Langer, B.; Moore, M.; Taylor, B.R.; Narod, S.A.; Darlington, G.; Andrulis, I.L.; et al. Germline BRCA2 6174delT mutations in Ashkenazi Jewish pancreatic cancer patients. Nat. Genet. 1997, 16, 17–18. [Google Scholar] [CrossRef]
  21. Chaffee, K.G.; Oberg, A.L.; McWilliams, R.R.; Majithia, N.; Allen, B.A.; Kidd, J.; Singh, N.; Hartman, A.-R.; Wenstrup, R.J.; Petersen, G.M. Prevalence of germ-line mutations in cancer genes among pancreatic cancer patients with a positive family history. Anesthesia Analg. 2017, 20, 119–127. [Google Scholar] [CrossRef]
  22. Ferrone, C.R.; Levine, D.A.; Tang, L.H.; Allen, P.J.; Jarnagin, W.; Brennan, M.F.; Offit, K.; Robson, M.E. BRCA Germline Mutations in Jewish Patients with Pancreatic Adenocarcinoma. J. Clin. Oncol. 2009, 27, 433–438. [Google Scholar] [CrossRef] [PubMed]
  23. Golan, T.; Kindler, H.L.; Park, J.O.; Reni, M.; Macarulla, T.; Hammel, P.; Van Cutsem, E.; Arnold, D.; Hochhauser, D.; McGuinness, D.; et al. Geographic and Ethnic Heterogeneity of Germline BRCA1 or BRCA2 Mutation Prevalence Among Patients With Metastatic Pancreatic Cancer Screened for Entry Into the POLO Trial. J. Clin. Oncol. 2020, 38, 1442–1454. [Google Scholar] [CrossRef] [PubMed]
  24. Turner, N.; Tutt, A.; Ashworth, A. Hallmarks of ‘BRCAness’ in sporadic cancers. Nat. Rev. Cancer 2004, 4, 814–819. [Google Scholar] [CrossRef] [PubMed]
  25. Lord, C.J.; Ashworth, A. BRCAness revisited. Nat. Rev. Cancer 2016, 16, 110–120. [Google Scholar] [CrossRef] [PubMed]
  26. Telli, M.L.; Timms, K.M.; Reid, J.; Hennessy, B.; Mills, G.B.; Jensen, K.C.; Szallasi, Z.; Barry, W.T.; Winer, E.P.; Tung, N.M.; et al. Homologous Recombination Deficiency (HRD) Score Predicts Response to Platinum-Containing Neoadjuvant Chemotherapy in Patients with Triple-Negative Breast Cancer. Clin. Cancer Res. 2016, 22, 3764–3773. [Google Scholar] [CrossRef]
  27. Takaya, H.; Nakai, H.; Takamatsu, S.; Mandai, M.; Matsumura, N. Homologous recombination deficiency status-based classification of high-grade serous ovarian carcinoma. Sci. Rep. 2020, 10, 2757. [Google Scholar] [CrossRef]
  28. Wattenberg, M.M.; Reiss, K.A. Determinants of Homologous Recombination Deficiency in Pancreatic Cancer. Cancers 2021, 13, 4716. [Google Scholar] [CrossRef]
  29. Birkbak, N.J.; Wang, Z.C.; Kim, J.-Y.; Eklund, A.C.; Li, Q.; Tian, R.; Bowman-Colin, C.; Li, Y.; Greene-Colozzi, A.; Iglehart, J.D.; et al. Telomeric Allelic Imbalance Indicates Defective DNA Repair and Sensitivity to DNA-Damaging Agents. Cancer Discov. 2012, 2, 366–375. [Google Scholar] [CrossRef]
  30. Popova, T.; Manié, E.; Rieunier, G.; Caux-Moncoutier, V.; Tirapo, C.; Dubois, T.; Delattre, O.; Sigal-Zafrani, B.; Bollet, M.; Longy, M.; et al. Ploidy and Large-Scale Genomic Instability Consistently Identify Basal-like Breast Carcinomas with BRCA1/2 Inactivation. Cancer Res. 2012, 72, 5454–5462. [Google Scholar] [CrossRef]
  31. Abkevich, V.; Timms, K.M.; Hennessy, B.T.; Potter, J.; Carey, M.S.; Meyer, L.A.; Smith-McCune, K.; Broaddus, R.; Lu, K.H.; Chen, J.; et al. Patterns of genomic loss of heterozygosity predict homologous recombination repair defects in epithelial ovarian cancer. Br. J. Cancer 2012, 107, 1776–1782. [Google Scholar] [CrossRef]
  32. Ray-Coquard, I.; Pautier, P.; Pignata, S.; Pérol, D.; González-Martín, A.; Berger, R.; Fujiwara, K.; Vergote, I.; Colombo, N.; Mäenpää, J.; et al. Olaparib plus Bevacizumab as First-Line Maintenance in Ovarian Cancer. N. Engl. J. Med. 2019, 381, 2416–2428. [Google Scholar] [CrossRef]
  33. González-Martín, A.; Pothuri, B.; Vergote, I.; DePont Christensen, R.; Graybill, W.; Mirza, M.R.; McCormick, C.; Lorusso, D.; Hoskins, P.; Freyer, G.; et al. Niraparib in Patients with Newly Diagnosed Advanced Ovarian Cancer. N. Engl. J. Med. 2019, 381, 2391–2402. [Google Scholar] [CrossRef]
  34. Zámborszky, J.; Szikriszt, B.; Gervai, J.Z.; Pipek, O.; Póti, A.; Krzystanek, M.; Ribli, D.; Szalai-Gindl, J.M.; Csabai, I.; Szallasi, Z.; et al. Loss of BRCA1 or BRCA2 markedly increases the rate of base substitution mutagenesis and has distinct effects on genomic deletions. Oncogene 2016, 36, 746–755. [Google Scholar] [CrossRef]
  35. Aguirre, A.J.; Nowak, J.A.; Camarda, N.D.; Moffitt, R.A.; Ghazani, A.A.; Hazar-Rethinam, M.; Raghavan, S.; Kim, J.; Brais, L.K.; Ragon, D.; et al. Real-time Genomic Characterization of Advanced Pancreatic Cancer to Enable Precision Medicine. Cancer Discov. 2018, 8, 1096–1111. [Google Scholar] [CrossRef]
  36. Davies, H.; Glodzik, D.; Morganella, S.; Yates, L.R.; Staaf, J.; Zou, X.; Ramakrishna, M.; Martin, S.; Boyault, S.; Sieuwerts, A.M.; et al. HRDetect is a predictor of BRCA1 and BRCA2 deficiency based on mutational signatures. Nat. Med. 2017, 23, 517–525. [Google Scholar] [CrossRef]
  37. Casolino, R.; Paiella, S.; Azzolina, D.; Beer, P.A.; Corbo, V.; Lorenzoni, G.; Gregori, D.; Golan, T.; Braconi, C.; Froeling, F.E.M.; et al. Homologous Recombination Deficiency in Pancreatic Cancer: A Systematic Review and Prevalence Meta-Analysis. J. Clin. Oncol. 2021, 39, 2617–2631. [Google Scholar] [CrossRef]
  38. Park, W.; O’Connor, C.A.; Bandlamudi, C.; Forman, D.; Chou, J.F.; Umeda, S.; Reyngold, M.; Varghese, A.M.; Keane, F.; Balogun, F.; et al. Clinico-genomic Characterization of ATM and HRD in Pancreas Cancer: Application for Practice. Clin. Cancer Res. 2022, 28, 4782–4792. [Google Scholar] [CrossRef]
  39. Leibowitz, B.D.; Dougherty, B.V.; Bell, J.S.K.; Kapilivsky, J.; Michuda, J.; Sedgewick, A.J.; Munson, W.A.; Chandra, T.A.; Dry, J.R.; Beaubier, N.; et al. Validation of genomic and transcriptomic models of homologous recombination deficiency in a real-world pan-cancer cohort. BMC Cancer 2022, 22, 587. [Google Scholar] [CrossRef]
  40. Jonsson, P.; Bandlamudi, C.; Cheng, M.L.; Srinivasan, P.; Chavan, S.S.; Friedman, N.D.; Rosen, E.Y.; Richards, A.L.; Bouvier, N.; Selcuklu, S.D.; et al. Tumour lineage shapes BRCA-mediated phenotypes. Nature 2019, 571, 576–579. [Google Scholar] [CrossRef]
  41. Momtaz, P.; O’connor, C.A.; Chou, J.F.; Capanu, M.; Park, W.; Bandlamudi, C.; Berger, M.F.; Kelsen, D.P.; Suehnholz, S.P.; Chakravarty, D.; et al. Pancreas cancer and BRCA: A critical subset of patients with improving therapeutic outcomes. Cancer 2021, 127, 4393–4402. [Google Scholar] [CrossRef]
  42. Golan, T.; Sella, T.; O’Reilly, E.M.; Katz, M.H.G.; Epelbaum, R.; Kelsen, D.P.; Borgida, A.; Maynard, H.; Kindler, H.; Friedmen, E.; et al. Overall survival and clinical characteristics of BRCA mutation carriers with stage I/II pancreatic cancer. Br. J. Cancer 2017, 116, 697–702. [Google Scholar] [CrossRef] [PubMed]
  43. Golan, T.; Kanji, Z.S.; Epelbaum, R.; Devaud, N.; Dagan, E.; Holter, S.; Aderka, D.; Paluch-Shimon, S.; Kaufman, B.; Gershoni-Baruch, R.; et al. Overall survival and clinical characteristics of pancreatic cancer in BRCA mutation carriers. Br. J. Cancer 2014, 111, 1132–1138. [Google Scholar] [CrossRef] [PubMed]
  44. O’reilly, E.M.; Lee, J.W.; Zalupski, M.; Capanu, M.; Park, J.; Golan, T.; Tahover, E.; Lowery, M.A.; Chou, J.F.; Sahai, V.; et al. Randomized, Multicenter, Phase II Trial of Gemcitabine and Cisplatin with or Without Veliparib in Patients with Pancreas Adenocarcinoma and a Germline BRCA/PALB2 Mutation. J. Clin. Oncol. 2020, 38, 1378–1388. [Google Scholar] [CrossRef] [PubMed]
  45. Mateo, J.; Lord, C.J.; Serra, V.; Tutt, A.; Balmaña, J.; Castroviejo-Bermejo, M.; Cruz, C.; Oaknin, A.; Kaye, S.B.; de Bono, J.S. A decade of clinical development of PARP inhibitors in perspective. Ann. Oncol. 2019, 30, 1437–1447. [Google Scholar] [CrossRef] [PubMed]
  46. Ashworth, A. A Synthetic Lethal Therapeutic Approach: Poly(ADP) Ribose Polymerase Inhibitors for the Treatment of Cancers Deficient in DNA Double-Strand Break Repair. J. Clin. Oncol. 2008, 26, 3785–3790. [Google Scholar] [CrossRef]
  47. Kaufman, B.; Shapira-Frommer, R.; Schmutzler, R.K.; Audeh, M.W.; Friedlander, M.; Balmaña, J.; Mitchell, G.; Fried, G.; Stemmer, S.M.; Hubert, A.; et al. Olaparib Monotherapy in Patients with Advanced Cancer and a Germline BRCA1/2 Mutation. J. Clin. Oncol. 2015, 33, 244–250. [Google Scholar] [CrossRef]
  48. Reiss, K.A.; Mick, R.; O’Hara, M.H.; Teitelbaum, U.; Karasic, T.B.; Schneider, C.; Cowden, S.; Southwell, T.; Romeo, J.; Izgur, N.; et al. Phase II Study of Maintenance Rucaparib in Patients with Platinum-Sensitive Advanced Pancreatic Cancer and a Pathogenic Germline or Somatic Variant in BRCA1, BRCA2, or PALB2. J. Clin. Oncol. 2021, 39, 2497–2505. [Google Scholar] [CrossRef]
  49. Shroff, R.T.; Hendifar, A.; McWilliams, R.R.; Geva, R.; Epelbaum, R.; Rolfe, L.; Goble, S.; Lin, K.K.; Biankin, A.V.; Giordano, H.; et al. Rucaparib Monotherapy in Patients with Pancreatic Cancer and a Known Deleterious BRCA Mutation. JCO Precis. Oncol. 2018, 2, 1–15. [Google Scholar] [CrossRef]
  50. Lowery, M.A.; Kelsen, D.P.; Capanu, M.; Smith, S.C.; Lee, J.W.; Stadler, Z.K.; Moore, M.J.; Kindler, H.L.; Golan, T.; Segal, A.; et al. Phase II trial of veliparib in patients with previously treated BRCA-mutated pancreas ductal adenocarcinoma. Eur. J. Cancer 2017, 89, 19–26. [Google Scholar] [CrossRef]
  51. Javle, M.; Shacham-Shmueli, E.; Xiao, L.; Varadhachary, G.; Halpern, N.; Fogelman, D.; Boursi, B.; Uruba, S.; Margalit, O.; Wolff, R.A.; et al. Olaparib Monotherapy for Previously Treated Pancreatic Cancer with DNA Damage Repair Genetic Alterations Other Than Germline BRCA Variants. JAMA Oncol. 2021, 7, 693. [Google Scholar] [CrossRef]
  52. Teke, M.E.; Saif, A.; Ryan, C.E.; Lux, S.C.; Hernandez, J.M.; Reiss, K.A. A Randomized Study of Olaparib or Placebo in Patients with Surgically Removed Pancreatic Cancer who have a BRCA1, BRCA2 or PALB2 Mutation (The APOLLO Trial). Ann. Surg. Oncol. 2022, 29, 5375–5376. [Google Scholar] [CrossRef]
  53. Bouwman, P.; Jonkers, J. Molecular Pathways: How Can BRCA-Mutated Tumors Become Resistant to PARP Inhibitors? Clin. Cancer Res. 2014, 20, 540–547. [Google Scholar] [CrossRef]
  54. Kim, H.; George, E.; Ragland, R.L.; Rafail, S.; Zhang, R.; Krepler, C.; Morgan, M.A.; Herlyn, M.; Brown, E.J.; Simpkins, F. Targeting the ATR/CHK1 Axis with PARP Inhibition Results in Tumor Regression in BRCA-Mutant Ovarian Cancer Models. Clin. Cancer Res. 2017, 23, 3097–3108. [Google Scholar] [CrossRef]
  55. Yazinski, S.A.; Comaills, V.; Buisson, R.; Genois, M.-M.; Nguyen, H.D.; Ho, C.K.; Kwan, T.T.; Morris, R.; Lauffer, S.; Nussenzweig, A.; et al. ATR inhibition disrupts rewired homologous recombination and fork protection pathways in PARP inhibitor-resistant BRCA-deficient cancer cells. Genes Dev. 2017, 31, 318–332. [Google Scholar] [CrossRef]
  56. Pettitt, S.J.; Frankum, J.R.; Punta, M.; Lise, S.; Alexander, J.; Chen, Y.; Yap, T.A.; Haider, S.; Tutt, A.N.; Lord, C.J. Clinical BRCA1/2 Reversion Analysis Identifies Hotspot Mutations and Predicted Neoantigens Associated with Therapy Resistance. Cancer Discov. 2020, 10, 1475–1488. [Google Scholar] [CrossRef]
  57. Gout, J.; Perkhofer, L.; Morawe, M.; Arnold, F.; Ihle, M.; Biber, S.; Lange, S.; Roger, E.; Kraus, J.M.; Stifter, K.; et al. Synergistic targeting and resistance to PARP inhibition in DNA damage repair-deficient pancreatic cancer. Gut 2020, 70, 743–760. [Google Scholar] [CrossRef]
  58. Perkhofer, L.; Golan, T.; Cuyle, P.-J.; Matysiak-Budnik, T.; Van Laethem, J.-L.; Macarulla, T.; Cauchin, E.; Kleger, A.; Beutel, A.K.; Gout, J.; et al. Targeting DNA Damage Repair Mechanisms in Pancreas Cancer. Cancers 2021, 13, 4259. [Google Scholar] [CrossRef]
  59. Chiorean, E.G.; Guthrie, K.A.; Philip, P.A.; Swisher, E.M.; Jalikis, F.; Pishvaian, M.J.; Berlin, J.; Noel, M.S.; Suga, J.M.; Garrido-Laguna, I.; et al. Randomized Phase II Study of PARP Inhibitor ABT-888 (Veliparib) with Modified FOLFIRI versus FOLFIRI as Second-line Treatment of Metastatic Pancreatic Cancer: SWOG S1513. Clin. Cancer Res. 2021, 27, 6314–6322. [Google Scholar] [CrossRef]
  60. Jiao, S.; Xia, W.; Yamaguchi, H.; Wei, Y.; Chen, M.K.; Hsu, J.M.; Hsu, J.L.; Yu, W.H.; Du, Y.; Lee, H.H.; et al. PARP Inhibitor Upregulates PD-L1 Expression and Enhances Cancer-Associated Immunosuppression. Clin. Cancer Res. 2017, 23, 3711–3720. [Google Scholar] [CrossRef]
  61. Mouw, K.W.; Goldberg, M.S.; Konstantinopoulos, P.A.; D’Andrea, A.D. DNA Damage and Repair Biomarkers of Immunotherapy Response. Cancer Discov. 2017, 7, 675–693. [Google Scholar] [CrossRef] [PubMed]
  62. Nesic, K.; Wakefield, M.; Kondrashova, O.; Scott, C.L.; McNeish, I.A. Targeting DNA repair: The genome as a potential biomarker. J. Pathol. 2018, 244, 586–597. [Google Scholar] [CrossRef] [PubMed]
  63. Reiss, K.A.; Mick, R.; Teitelbaum, U.; O’Hara, M.; Schneider, C.; Massa, R.; Karasic, T.; Tondon, R.; Onyiah, C.; Gosselin, M.K.; et al. Niraparib plus nivolumab or niraparib plus ipilimumab in patients with platinum-sensitive advanced pancreatic cancer: A randomised, phase 1b/2 trial. Lancet Oncol. 2022, 23, 1009–1020. [Google Scholar] [CrossRef] [PubMed]
  64. Kaplan, A.R.; Gueble, S.E.; Liu, Y.; Oeck, S.; Kim, H.; Yun, Z.; Glazer, P.M. Cediranib suppresses homology-directed DNA repair through down-regulation of BRCA1/2 and RAD51. Sci. Transl. Med. 2019, 11, eaav4508. [Google Scholar] [CrossRef] [PubMed]
  65. Bindra, R.S.; Gibson, S.L.; Meng, A.; Westermark, U.; Jasin, M.; Pierce, A.J.; Bristow, R.G.; Classon, M.K.; Glazer, P.M. Hypoxia-Induced Down-regulation of BRCA1 Expression by E2Fs. Cancer Res. 2005, 65, 11597–11604. [Google Scholar] [CrossRef] [PubMed]
  66. Hegan, D.C.; Lu, Y.; Stachelek, G.C.; Crosby, M.E.; Bindra, R.S.; Glazer, P.M. Inhibition of poly(ADP-ribose) polymerase down-regulates BRCA1 and RAD51 in a pathway mediated by E2F4 and p130. Proc. Natl. Acad. Sci. USA 2010, 107, 2201–2206. [Google Scholar] [CrossRef]
  67. Kim, J.W.; McKay, R.R.; Radke, M.R.; Zhao, S.; Taplin, M.-E.; Davis, N.B.; Monk, P.; Appleman, L.J.; Lara, P.N.; Vaishampayan, U.N.; et al. Randomized Trial of Olaparib with or Without Cediranib for Metastatic Castration-Resistant Prostate Cancer: The Results From National Cancer Institute 9984. J. Clin. Oncol. 2023, 41, 871–880. [Google Scholar] [CrossRef]
  68. Lloyd, R.L.; Wijnhoven, P.W.G.; Ramos-Montoya, A.; Wilson, Z.; Illuzzi, G.; Falenta, K.; Jones, G.N.; James, N.; Chabbert, C.D.; Stott, J.; et al. Combined PARP and ATR inhibition potentiates genome instability and cell death in ATM-deficient cancer cells. Oncogene 2020, 39, 4869–4883. [Google Scholar] [CrossRef]
  69. Sun, C.; Fang, Y.; Yin, J.; Chen, J.; Ju, Z.; Zhang, D.; Chen, X.; Vellano, C.P.; Jeong, K.J.; Ng, P.K.-S.; et al. Rational combination therapy with PARP and MEK inhibitors capitalizes on therapeutic liabilities in RAS mutant cancers. Sci. Transl. Med. 2017, 9, eaal5148. [Google Scholar] [CrossRef]
  70. O’Reilly, E.M.; Oh, D.-Y.; Dhani, N.; Renouf, D.J.; Lee, M.A.; Sun, W.; Fisher, G.; Hezel, A.; Chang, S.-C.; Vlahovic, G.; et al. Durvalumab With or Without Tremelimumab for Patients with Metastatic Pancreatic Ductal Adenocarcinoma: A Phase 2 Randomized Clinical Trial. JAMA Oncol. 2019, 5, 1431–1438. [Google Scholar] [CrossRef]
  71. Sokol, E.S.; Pavlick, D.; Khiabanian, H.; Frampton, G.M.; Ross, J.S.; Gregg, J.P.; Lara, P.N.; Oesterreich, S.; Agarwal, N.; Necchi, A.; et al. Pan-Cancer Analysis of BRCA1 and BRCA2 Genomic Alterations and Their Association with Genomic Instability as Measured by Genome-Wide Loss of Heterozygosity. JCO Precis. Oncol. 2020, 4, 442–465. [Google Scholar] [CrossRef]
  72. Terrero, G.; Datta, J.; Dennison, J.; Sussman, D.A.; Lohse, I.; Merchant, N.B.; Hosein, P.J. Ipilimumab/Nivolumab Therapy in Patients with Metastatic Pancreatic or Biliary Cancer with Homologous Recombination Deficiency Pathogenic Germline Variants. JAMA Oncol. 2022, 8, 938–940. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, J.; Dai, Q.; Park, D.; Deng, X. Targeting DNA Replication Stress for Cancer Therapy. Genes 2016, 7, 51. [Google Scholar] [CrossRef] [PubMed]
  74. da Costa, A.A.B.A.; Chowdhury, D.; Shapiro, G.I.; D’andrea, A.D.; Konstantinopoulos, P.A. Targeting replication stress in cancer therapy. Nat. Rev. Drug Discov. 2022, 22, 38–58. [Google Scholar] [CrossRef] [PubMed]
  75. Lecona, E.; Fernandez-Capetillo, O. Targeting ATR in cancer. Nat. Rev. Cancer 2018, 18, 586–595. [Google Scholar] [CrossRef]
  76. Karnitz, L.M.; Zou, L. Molecular Pathways: Targeting ATR in Cancer Therapy. Clin. Cancer Res. 2015, 21, 4780–4785. [Google Scholar] [CrossRef]
  77. Yap, T.A.; Tan, D.S.P.; Terbuch, A.; Caldwell, R.; Guo, C.; Goh, B.C.; Heong, G.B.C.V.; Haris, N.R.M.; Bashir, S.; Drew, Y.; et al. First-in-Human Trial of the Oral Ataxia Telangiectasia and RAD3-Related (ATR) Inhibitor BAY 1895344 in Patients with Advanced Solid Tumors. Cancer Discov. 2021, 11, 80–91. [Google Scholar] [CrossRef]
  78. Yap, T.A.; O’carrigan, B.; Penney, M.S.; Lim, J.S.; Brown, J.S.; Luken, M.J.D.M.; Tunariu, N.; Perez-Lopez, R.; Rodrigues, D.N.; Riisnaes, R.; et al. Phase I Trial of First-in-Class ATR Inhibitor M6620 (VX-970) as Monotherapy or in Combination with Carboplatin in Patients With Advanced Solid Tumors. J. Clin. Oncol. 2020, 38, 3195–3204. [Google Scholar] [CrossRef]
  79. Hong, D.S.; Moore, K.; Patel, M.; Grant, S.C.; Burris, H.A.; William, W.N.; Jones, S.; Meric-Bernstam, F.; Infante, J.; Golden, L.; et al. Evaluation of Prexasertib, a Checkpoint Kinase 1 Inhibitor, in a Phase Ib Study of Patients with Squamous Cell Carcinoma. Clin. Cancer Res. 2018, 24, 3263–3272. [Google Scholar] [CrossRef]
  80. Lee, J.M.; Nair, J.; Zimmer, A.; Lipkowitz, S.; Annunziata, C.M.; Merino, M.J.; Swisher, E.M.; Harrell, M.I.; Trepel, J.B.; Lee, M.J.; et al. Prexasertib, a cell cycle checkpoint kinase 1 and 2 inhibitor, in BRCA wild-type recurrent high-grade serous ovarian cancer: A first-in-class proof-of-concept phase 2 study. Lancet Oncol. 2018, 19, 207–215. [Google Scholar] [CrossRef]
  81. Lampert, E.J.; An, D.; McCoy, A.; Kohn, E.C.; Annunziata, C.M.; Trewhitt, K.; Zimmer, A.D.S.; Lipkowitz, S.; Lee, J.-M. Prexasertib, a cell cycle checkpoint kinase 1 inhibitor, in BRCA mutant recurrent high-grade serous ovarian cancer (HGSOC): A proof-of-concept single arm phase II study. J. Clin. Oncol. 2020, 38, 6038. [Google Scholar] [CrossRef]
  82. Moore, K.N.; Chambers, S.K.; Hamilton, E.P.; Chen, L.-M.; Oza, A.M.; Ghamande, S.A.; Konecny, G.E.; Plaxe, S.C.; Spitz, D.L.; Geenen, J.J.; et al. Adavosertib with Chemotherapy in Patients with Primary Platinum-Resistant Ovarian, Fallopian Tube, or Peritoneal Cancer: An Open-Label, Four-Arm, Phase II Study. Clin. Cancer Res. 2022, 28, 36–44. [Google Scholar] [CrossRef]
  83. Meric-Bernstam, F.; Chalsani, P.; Mamdani, H.; Zheng, C.; Viana, M.; Lambersky, R.; Pultar, P.; Tolcher, A.W. Abstract CT029: Safety and clinical activity of single-agent ZN-c3, an oral WEE1 inhibitor, in a phase 1 trial in subjects with recurrent or advanced uterine serous carcinoma (USC). Cancer Res. 2022, 82, CT029. [Google Scholar] [CrossRef]
  84. Zhou, J.; Gelot, C.; Pantelidou, C.; Li, A.; Yücel, H.; Davis, R.E.; Färkkilä, A.; Kochupurakkal, B.; Syed, A.; Shapiro, G.I.; et al. A first-in-class polymerase theta inhibitor selectively targets homologous-recombination-deficient tumors. Nat. Cancer 2021, 2, 598–610. [Google Scholar] [CrossRef]
  85. Zatreanu, D.; Robinson, H.M.R.; Alkhatib, O.; Boursier, M.; Finch, H.; Geo, L.; Grande, D.; Grinkevich, V.; Heald, R.A.; Langdon, S.; et al. Polθ inhibitors elicit BRCA-gene synthetic lethality and target PARP inhibitor resistance. Nat. Commun. 2021, 12, 3636. [Google Scholar] [CrossRef]
  86. Patterson-Fortin, J.; Jadhav, H.; Pantelidou, C.; Phan, T.; Grochala, C.; Mehta, A.K.; Guerriero, J.L.; Wulf, G.M.; Wolpin, B.M.; Stanger, B.Z.; et al. Polymerase θ inhibition activates the cGAS-STING pathway and cooperates with immune checkpoint blockade in models of BRCA-deficient cancer. Nat. Commun. 2023, 14, 1390. [Google Scholar] [CrossRef]
Figure 1. Targeting the DNA damage repair pathway in pancreatic cancer (created using Biorender.com accessed on 2 April 2023).
Figure 1. Targeting the DNA damage repair pathway in pancreatic cancer (created using Biorender.com accessed on 2 April 2023).
Cancers 15 02955 g001
Table 1. Select Tests to Evaluate for HRD (not specific to pancreatic cancer).
Table 1. Select Tests to Evaluate for HRD (not specific to pancreatic cancer).
TestDescription
Telomeric Allelic Imbalance (TAI) [28] Allelic imbalance at the telomere of the chromosome is due to the propensity for inappropriate end joining in HRD, identified by single nucleotide polymorphism (SNP) genotyping.
Large Scale Transitions (LST) [29] Chromosomal breaks larger than 10 Mb, which arise in HRD cells secondary to erroneous recombination between segments of the chromosome, are identified by single nucleotide polymorphism (SNP) genotyping.
Loss of Heterozygosity (LOH) [30] Uniparental disomy is owing to inaccurate repair of sister chromatids during S/G2 phase, resulting in the loss of entire genes and the surrounding chromosomal region, as identified by single nucleotide polymorphism (SNP) genotyping.
Genomic Instability Score eg by Myriad Genetics MyChoice AssayTAI + LST + LOH
Signature 3 (Sig 3) [35] A single base substitution mutational pattern, associated with microhomology and large deletions, was identified by whole exome sequencing.
HRDetect [36]A weighted model incorporating a weighted score of base substitution/rearrangement signatures, microhomology-mediated deletions, and an HRD score based on genomic scars identified by whole exome sequencing.
Table 2. Select Recent Completed Trials in BRCA and HRD PDAC.
Table 2. Select Recent Completed Trials in BRCA and HRD PDAC.
NCT IdentifierPhaseTarget PopulationExperimental ArmControl ArmPrimary OutcomeOther Key Findings
NCT02184195(POLO)3Metastatic PDAC, gBRCA1/2 mutated, with platinum sensitivityOlaparibPlaceboPFS 7.4 mo (experimental arm) vs. 3.8 mo (control arm)PFS 2 16.9 mo (experimental arm) vs. 9.3 mo (placebo arm).
Quality of life scores equivalent.
OS 19.0 mo (olaparib) vs. 19.2 months (placebo)
NCT031406702Metastatic PDAC, s/g BRCA1/2 or PALB2 mutated, with platinum sensitivityRucaparibN/APFS6 59.5%mPFS 13.1 mo
mOS 23.5 mo
ORR 41.7%
NCT02042378 (RUCAPANC)2Metastatic or advanced PDAC, s/g BRCA mutation, 1–2 lines of previous treatmentRucaparibN/ARR 15.8%
(1 CR, 2 PR)
DCR 31.6%
Terminated early due to insufficient response rate
NCT021841952Metastatic or locally advanced PDAC, previously treated, with gBRCA1/2 or gPALB2 mutationsCisplatin, gemcitabine (Arm B) plus veliparib (Arm A)Cisplatin, gemcitabine (Arm B)RR 74.1% for Arm A, 65.2% for Arm B DCR 100% for Arm A, 78.3% for Arm B.
mPFS 10.1 mo for Arm A, 9.7mo for Arm B.
mOS 15.5 mo for Arm A, 16.4 mo for Arm B
NCT034049602Locally advanced or metastatic PDAC with platinum sensitivity, in the platinum-sensitive, maintenance settingNiraparib plus nivolumab and niraparib plus ipilimumabN/A6m PFS 20.6% in niraparib + nivo arm versus 59.6% in the niraparib + ipi armmOS 10.2 mo in niraparib + nivo, 38 mo for niraparib + ipi
Higher grade 3 toxicity with niraparib + ipilimumab
NCT021841952Metastatic or locally advanced PDAC, gBRCA1/2 or gPALB2, 1–2 lines of previous treatmentVeliparib 300mg twice daily or veliparib 400mg twice dailyN/AORR 0%mPFS 1.7 mo
OS 3.1 mo
NCT01296731Unresectable PDAC, not confined to those with BRCA or HRD mutationsOlaparib, cisplatin, irinotecan, and mitomycinN/A ORR 23%
Grade 3 AEs in 89%
11% pts developed MDS
1 pt with gBRCA2 had durable response for >4 years
Abbreviations: PDAC: pancreatic ductal adenocarcinoma; gBRCA: germline BRCA; DLT: dose-limiting toxicities; mPFS: median progression-free survival; AE: adverse events; mo: months; ORR: objective response rate; s/g: somatic or germline; N/A: not applicable; PFS6: progression-free survival at six months; mOS: median overall survival; ORR: overall response rate; CR: complete response; GnP: gemcitabine plus nab-paclitaxel; PR: partial response; mFFX: modified FOLFIRINOX; SD: stable disease; DCR: disease control rate; POD: progression of disease; MDS: myelodysplastic syndrome.
Table 3. Select ongoing trials for BRCA, BRCAness, and HRD pancreatic cancer.
Table 3. Select ongoing trials for BRCA, BRCAness, and HRD pancreatic cancer.
NCT IdentifierPhaseTarget PopulationExperimental ArmControl ArmPrimary OutcomeAnticipated Completion
NCT026770382Metastatic PDAC with s/gHRD gene mutations (except BRCA1/2) or a family history suggestive of a HRD mutation.OlaparibN/AORRNovember 2022
NCT04858334
(APOLLO)
2Surgically removed PDAC, post-adjuvant chemo(radiotherapy), and a g/sBRCA1/2 or PALB2 mutation.OlaparibN/ARFS relapse free survivalNovember 2024
NCT04666740
(POLAR)
2Metastatic PDAC with core or non-core HRD mutations or platinum sensitivity.Olaparib plus pembrolizumabN/APFSJanuary 2024
NCT045487522Metastatic PDAC with gBRCA1/2 mutations.Olaparib plus pembrolizumabOlaparibPFSMarch 2025
NCT04171700
(LODESTAR)
2Metastatic solid tumors, including PDA with mutations in BRCA1/2, PALB2, RAD51C, RAD51D, BARD1, FANCA, NBN, RAD51, or RAD51B.RucaparibN/AORRMarch 2023
NCT03553004
(NIRAPANC)
2Locally advanced or metastatic PDAC with g/sHRD mutations.NiraparibN/AORRFebruary 2025
NCT036019232Locally advanced or metastatic PDAC with g/sBRCA1/2, PALB2, CHEK2, and ATM mutations.NiraparibN/APFSFebruary 2025
NCT044930602Metastatic PDAC with a BRCA1/2 or PALB2 mutation.Niraparib plus dostarlimabN/ADCR12September 2022
NCT046734481Metastatic solid tumors, including PDAC with a g/sBRCA1/2 mutation.Niraparib plus TSR-042N/APFSMarch 2026
NCT036019232Previously treated PDAC with g/sHRD mutations.Niraparib plus small field palliative radiationN/APFSFebruary 2025
NCT040056901Locally advanced or metastatic PDAC.Arm 1: cobimetinib
Arm 2: olaparib
N/AFeasibility of obtaining tumor tissue pre and post treatmentFebruary 2025
NCT045504942Locally advanced or metastatic solid tumors, including PDAC with g/s DDR gene alterations.TalazoparibN/APDAugust 2022
NCT04348045
MAZEPPA)
2Metastatic PDAC and known somatic profile—stratified by BRCAness and KRAS mutation status.Pts with BRCAness (somatic profile): olaparib (Arm A)
Pts without BRCAness and with KRAS mutation are randomized to durvalumab + selumetinib (Arm B) or FOLFIRI (Arm C)
N/APFSDecember 2024
NCT033370871/2Metastatic PDAC (colorectal, gastroesophageal, and biliary cancer) and for the phase 2 component, BRCA1/2, PALB2, or HRD, untreated in the metastatic setting.Liposomal irinotecan, fluorouracil, calcium leucovorin, and rucaparibN/ADLT rate (Phase 1)
ORR (Phase 1b)
Best response (Phase 2)
Not recruiting
NCT021948291Metastatic or unresectable PDAC with alterations in DDR genes, previously untreated.NovobiocinN/AMTD
RP2D
Anticipated to open in 2023
Abbreviations: AE: adverse events; PDAC: pancreatic ductal adenocarcinoma; DCR12: disease control rate at 12 weeks; ORR: overall response rate; DLT: dose limiting toxicities; HRD: homologous recombination deficiency; N/A: not applicable; DDR: DNA damage repair; MTD: maximum tolerated dose; PFS: progression free survival; GnP: Gemcitabine/Nab-Paclitaxel; OS: overall survival; mFFX: modified FOLFIRINOX; PK: pharmacokinetics; ctDNA: circulating tumor DNA; RP2D: recommended phase 2 dose.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Keane, F.; O’Connor, C.A.; Park, W.; Seufferlein, T.; O’Reilly, E.M. Pancreatic Cancer: BRCA Targeted Therapy and Beyond. Cancers 2023, 15, 2955. https://doi.org/10.3390/cancers15112955

AMA Style

Keane F, O’Connor CA, Park W, Seufferlein T, O’Reilly EM. Pancreatic Cancer: BRCA Targeted Therapy and Beyond. Cancers. 2023; 15(11):2955. https://doi.org/10.3390/cancers15112955

Chicago/Turabian Style

Keane, Fergus, Catherine A. O’Connor, Wungki Park, Thomas Seufferlein, and Eileen M. O’Reilly. 2023. "Pancreatic Cancer: BRCA Targeted Therapy and Beyond" Cancers 15, no. 11: 2955. https://doi.org/10.3390/cancers15112955

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

Article Metrics

Back to TopTop