Next Article in Journal
Serum Myostatin Predicts the Risk of Hepatocellular Carcinoma in Patients with Alcoholic Cirrhosis: A Multicenter Study
Previous Article in Journal
Assessment of HER2 Protein Overexpression and Gene Amplification in Renal Collecting Duct Carcinoma: Therapeutic Implication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

BRCA in Gastrointestinal Cancers: Current Treatments and Future Perspectives

by
Eleonora Molinaro
,
Kalliopi Andrikou
,
Andrea Casadei-Gardini
* and
Giulia Rovesti
Department of Oncology and Hematology, Division of Oncology, University of Modena and Reggio Emilia, 41121 Modena, Italy
*
Author to whom correspondence should be addressed.
Cancers 2020, 12(11), 3346; https://doi.org/10.3390/cancers12113346
Submission received: 10 October 2020 / Revised: 26 October 2020 / Accepted: 11 November 2020 / Published: 12 November 2020

Abstract

:

Simple Summary

BRCA gene mutations are progressively gaining more attention in the context of gastrointestinal malignancies, especially in pancreatic cancer where their identification can have both therapeutic and surveillance relevance.

Abstract

A strong association between pancreatic cancer and BRCA1 and BRCA2 mutations is documented. Based on promising results of breast and ovarian cancers, several clinical trials with poly (ADP-ribose) polymerase inhibitors (PARPi) are ongoing for gastrointestinal (GI) malignancies, especially for pancreatic cancer. Indeed, the POLO trial results provide promising and awaited changes for the pancreatic cancer therapeutic landscape. Contrariwise, for other gastrointestinal tumors, the rationale is currently only alleged. The role of BRCA mutation in gastrointestinal cancers is the subject of this review. In particular, we aim to provide the latest updates about novel therapeutic strategies that, exploiting DNA repair defects, promise to shape the future therapeutic scenario of GI cancers.

Graphical Abstract

1. Introduction

BRCA1 and BRCA2 are famous tumor susceptibility genes. They encode for proteins playing a crucial role in the correct repair of damaged DNA. Indeed, these genes are key components of the homologous recombination (HR) pathway [1]. Particularly, during a normal cell cycle, the double-strand DNA can be damaged by internal and exogenous agents, producing a double-strand break (DSB). The most known mechanisms the cell uses to repair the DSB are the following: HR, as said, and non-homologous end joining (NHEJ). The first one allows a greater genomic stability compared to the second one. In fact, HR employs an undamaged homologous sequence as a template; instead, NHEJ links the DNA broken ends directly, without a template (Figure 1).
A pathogenic mutation in BRCA1 or BRCA2 genes causes an impaired HR. Therefore, the cell, being deficient in HR, utilizes NHEJ preferentially to repair the DSB. NHEJ, unlike HR, enhances the cellular genomic instability until carcinogenesis. For this reason, BRCA mutation carriers have a higher risk of developing cancers during their life [2].
In addition, several other proteins participate in the HR process such as PALB2, ATM, BRIP1, RAD51 and CHEK2 for the correct DSB repair. Particularly, ATM is a protein kinase able to find the DSB and monitor its reparation, PALB2 also has a modulatory role, stabilizing the BRCA2 protein, BRIP1 encodes a protein-terminal helicase 1 involved in the DSB repair machine, RAD51 forms a nucleoprotein filament catalyzing homologous pairing and CHEK2 blocks the cell cycle when a DSB occurs [3]. Consequently, also a mutation in these proteins determines a HR deficiency (HRD) and can produce the same effect as BRCA1 and BRCA2 pathogenic mutations (Figure 2).
All these proteins form the so-called BRCAness phenotype [4].
Overall, the diagnosis of a pathogenic mutation either in BRCA proteins or in BRCAness proteins is particularly important to establish a commensurate management of surveillance programs and treatment strategies in hereditary conditions [5,6].
In recent years, due to the impairment of the HR pathway, BRCA-related cancers showed a major sensitivity to old and new drugs such as platinum-based chemotherapies and inhibitors of poly (ADP-ribose) polymerase (PARP), respectively [7]. As a matter of fact, platinum-based chemotherapies work as alkylating agents and produce a DSB in a cell unable to repair it [8]. Instead, PARP inhibitors (PARPi) are responsible for the so-called “synthetic lethality”. This process consists of different events that only when taken together can bring the cell to death. In detail, since also a single-strand DNA can be damaged, its reparation is mediated by PARP enzymes through the mechanism of base excision repair (BER). When this pathway is interrupted by the action of PARPi, the single-strand DNA break (SSB) cannot be repaired and it becomes a DSB. At last, in patients with a HR deficiency, such as BRCA mutation carriers, also a DSB cannot be repaired. Therefore, the cell accumulates gene alterations that lead it to death [9] (Figure 3).
Moreover, PARPi also act with an intrinsic cytotoxic effect, known as the “PARP trapping” effect, forming an inseparable complex with DNA strands [10].
As mentioned above, individuals carrying a germline mutation in BRCA1 and BRCA2 genes present a higher susceptibility to develop solid tumors. Classically, cancers most frequently associated with mutations in these genes are breast and ovarian ones. Indeed, the lifetime risk to develop breast cancer is approximately 52–72% among BRCA1 mutation carriers and 45–84% among BRCA2 mutation carriers. Furthermore, the lifetime risk for ovarian cancers is about 39–63% in BRCA1 mutation carriers and 11–27% in BRCA2 mutation carriers [11].
The cancer spectrum in BRCA1 and BRCA2 germline mutation carriers has been more extensively described in females than in males. In an effort to at least in part fill this gap, Silvestri et al. recently analyzed a large dataset of males harboring a germline mutation in the BRCA1 or BRCA2 gene, showing that being affected by any tumor and developing multiple cancers, particularly those of the breast, prostate and pancreas, is linked to a higher probability of being a BRCA2, rather than a BRCA1, carrier [12].
Regarding the topic of BRCA mutation, the knowledge about its role in gastrointestinal cancers is still limited, although it is known that the main underlining molecular pathways are those previously described; for this reason, the aim of this review is to describe the relationship existing between BRCA pathogenic variants and gastrointestinal cancers and its potential therapeutic role.

2. Pancreatic Cancer

With a five-year relative survival rate of 9%, the lowest among all cancer types, pancreatic cancer (PC) is the tumor with the most dismal prognosis. Moreover, deaths from PC are projected to increase dramatically in the next 20 years and by 2030, PC is expected to become the second leading cause of cancer-related death in the United States [13].
Cigarette smoking, increased body mass index, dietary factors, heavy alcohol consumption and a recent diagnosis of diabetes mellitus have been associated with increased pancreatic cancer risk [14,15,16,17], but inherited genetic factors also play an important role in pancreatic cancer risk [18]. It is estimated that 3% of PC cases derive from hereditary cancer syndromes (Peutz–Jeghers syndrome, PJS, ORPHA:2869, gene LKB1/STK11; hereditary pancreatitis, HP, ORPHA:676, gene PRSS1; familial atypical multiple mole melanoma, FAMMM, ORPHA:404,560, gene CDKN2A; hereditary breast and ovarian cancer syndrome, HBOCS, ORPHA:145, genes BRCA1 and BRCA2; Lynch syndrome, LS, ORPHA:144, genes MLH1, MSH2, MSH6, PMS2; familial adenomatous polyposis, FAP, ORPHA:733, gene APC), and that another 4–10% of the cases are classified as familial PC (FPC), which is defined as an individual who has two or more first-degree relatives (FDRs) with PC and without association, with known hereditary genetic syndromes [19,20,21]. BRCA mutations are the most common germline genetic alterations known to occur in PC, inherited in an autosomal dominant pattern with incomplete penetrance [22]. BRCA1 and BRCA2 pathogenic mutations are found in 1% or less and in up to 2% of unselected PC cases, respectively [23,24,25,26]. Among Ashkenazi Jewish individuals with pancreatic cancer, these kinds of mutations are found in up to 13.7% of unselected cases. In FPC, BRCA2 mutations are found in about 5% to 10% of cases and BRCA1 mutations in approximately 1%. The lifetime risk of developing PC is 2.1–3.5 times higher in BRCA mutation carriers [27,28]. In particular, it is estimated to be 3% for carriers of mutations in BRCA1 and 5% to 10% for carriers of mutations in BRCA2 [23], certainly lower than the risk of developing breast or ovarian cancer [29].
As for breast and ovarian cancer, it is likely that mutations in a specific gene region may influence the risk and the characteristics of pancreatic cancer that is developed by BRCA mutation carriers; in their retrospective study of 5143 Italian families with history of BRCA-related malignancies, Toss A. et al. indicated two possible pancreatic cancer cluster regions (PCCR) that should be further verified in a larger cohort of BRCA-associated pancreatic cancer patients [11] and that are different from those previously identified for BRCA1 in breast and ovarian cancers (BCCR and OCCR, respectively) and only marginally overlapping for BRCA2.
The prognosis of PC in BRCA mutation carriers remains unclear. In their retrospective analysis, Reiss et al. suggested a better prognosis in BRCA1, BRCA2 or PALB2 mutations carriers compared with non-carriers (21.8 vs. 8.1 months OS, HR 0.35, 95% CI 0.2–0.62; p < 0.001) [30]. However, some other studies showed no difference in overall survival (OS) [31,32] or even suggested a worse prognosis in BRCA mutation carriers [33]. Currently, systemic therapies for PC determine only a small increase in OS, therefore research advances are compelling, possibly moving in the direction of personalized, biomarker-driven options. Recent large-scale cancer genomic studies demonstrated a heterogeneous mutational profile, with activating mutations of KRAS present in over 90% of mutations of TP53, CDKN2A and SMAD4 in over 50% of cases. Other mutations have been found with a prevalence of less than 5%, with frequent heterogeneity from case to case, thus involving a significant intertumoral heterogeneity. A whole-genome sequencing analysis of 100 PC patients showed that chromosomal structural variation is a relevant mechanism of DNA damage in pancreatic carcinogenesis, allowing the identification and classification of PC into four specific subtypes of pancreatic adenocarcinoma: stable, locally rearranged, scattered and unstable. The unstable subtype, exhibiting a large number (>200, maximum of 558) of structural variations, resulted in being associated with inactivation of homologous recombination DNA damage repair (HR-DDR) genes (BRCA1, BRCA2 or PALB2) exhibiting a unique mutational signature reflecting defects in DNA maintenance and displaying sensitivity to DNA-damaging agents [34]. KRAS, TP53, CDKN2A and SMAD4 are not currently actionable therapeutic targets, as the most commonly mutated genes. Notably, however, mutations in the DDR system, including BRCA1 and BRCA2, but also ATM and PALB2, are emerging biologic targets for therapy in advanced pancreatic cancer [35].

2.1. BRCA Testing for Therapeutic Purpose

Identification of BRCA1 and BRCA2 mutations offers potential therapeutic advantages as they confer increased sensitivity to PARPi, reflecting a unique biology of BRCA-mutated pancreatic cancer cells [36,37].
The clinical evolution of PARPi in the context of PC has evolved from being used as monotherapies in refractory disease to maintenance therapies and in combination with other classes of therapeutics [36].
The only phase III trial has been conducted in the maintenance setting and is the international, randomized, placebo-controlled POLO (Pancreas cancer OLaparib Ongoing) trial, in which patients with metastatic pancreatic cancer and a germline BRCA1 and/or BRCA2 mutation whose disease had not progressed on first-line platinum-based chemotherapy derived a statistically significant and clinically meaningful improvement in progression-free survival (PFS; primary endpoint of the study) from maintenance treatment with PARP inhibitor olaparib vs. placebo. Median PFS was significantly longer in the olaparib arm (7.4 vs. 3.8 months, p = 0.004) and objective response rate (ORR) (23.1% vs. 11.5%) and median duration of response (24.9 vs. 3.7 months) were also improved. No difference in terms of median OS was observed between the two groups (18.9 vs. 18.1 months in the olaparib arm and placebo arm, respectively, p = 0.68), but data are still immature for this outcome since they derive from a planned interim analysis at data maturity of 46% [38]. Health-related quality of life (HRQoL) was preserved with maintenance olaparib treatment with no clinically meaningful difference compared with placebo, an important result for patients particularly when considering the cumulative toxicities of standard-of-care chemotherapies [39]. Of note, 21.7% of patients in the POLO trial progressed on first-line treatment and were ineligible for randomization, consistent with findings reported by Wattenberg et al. [40] in their retrospective real-world cohort study where over 40% of BRCA-defective PC patients did not respond to platinum-based chemotherapy and up to 20% had disease progression as best response even in the first-line setting. Clearly, a subset of PC patients harboring a BRCA germline mutation do not display the typical and possibly targetable HRD phenotype and the question on how to identify this subgroup of patients is still open [41]. Biomarker enrichment for the POLO trial was based on the presence of germline BRCA1 and BRCA2 mutations identified using the BRACAnalysis companion diagnostic assay; however, germline BRCA1 and BRCA2 mutations, which are typically found, respectively, in 1% and 2% of unselected PC cases, might reflect only the tip of the iceberg with regard to the potential target population [42]. Beyond germline BRCA1 and BRCA2 mutations, some BRCA-proficient tumors have defects in HR-DDR genes including ATM, ATR, CHK1, CHK2, PALB2 and RAD51 and also these cases, sharing the molecular features of BRCA-mutated tumors (BRCAness), are considered good targets for PARP inhibition treatment [21]. Nonetheless, on the basis of the randomized, placebo-controlled POLO trial, which showed that a biomarker-driven approach to PC treatment is achievable in practice, on December 27, 2019, the Food and Drug Administration (FDA) approved olaparib for the maintenance treatment of adult patients with deleterious or suspected deleterious germline BRCA-mutated metastatic pancreatic cancer, as detected by an FDA-approved test, whose disease has not progressed on at least 16 weeks of a first-line platinum-based chemotherapy regimen.
Outside the maintenance setting, the use of PARPi as single agents has generally underperformed in advanced-stage PC, suggesting that rational combination therapies are necessary in this disease and also relevant to the setting of PARP inhibitor resistance, which has both genomic mechanisms, such as BRCA1 and BRCA2 reversion mutations, and non-genomic mechanisms, including ATR pathway activation in order to bypass the impaired HR-DDR [35,43].
The combination of PARPi and chemotherapy has a strong rationale. Platinum-based drugs (cisplatin, carboplatin, oxaliplatin) are DNA cross-linking agents that kill tumor cells by interfering with DNA repair and inducing DSB, whereas topoisomerase inhibitors stall the replication fork by stabilizing the DNA complex in unrepaired state, enhancing SSBs. For this reason, the association between PARPi and these agents has been evaluated in several clinical trials. An open-label phase I/II clinical trial (NCT01489865) tested the combination of veliparib with mFOLFOX6 (modified Folinic acid + 5-FU + Oxaliplatin) in advanced PC patients. Patients in phase II of the trial were both pretreated (18 patients) and untreated (15 patients) and they were pre-selected for germline or somatic DDR mutations (69%) or had a family history suggestive of hereditary breast or ovarian cancer syndrome (HBOCS, 27%). The primary endpoint of the study was ORR, equal to 26% for the whole cohort. Interestingly, when considering different subgroups of patients separately, ORR was higher in platinum-naïve in respect to platinum-pretreated patients (33% vs. 7%, respectively), in patients with HBOCS in respect to no-HBOCS patients (30% vs. 14%, respectively) and in DDR mutation-positive in respect to negative ones (50% vs. 17%, respectively). The ORR for the platinum-naïve, HBOCS and DDR mutation-positive cohort was 58%, strongly highlighting the relevance of patient selection. Median PFS and OS for this selected cohort of patients was 8.7 and 11.8 months, respectively (vs. 3.7 and 8.5 months, respectively, for the unselected cohort of patients) [44]. The SWOG S1513 phase II trial (NCT02890355) randomized a biomarker unselected metastatic PC population to receive veliparib plus modified FOLFIRI (Folinic acid + 5-FU + Irinotecan) or FOLFIRI alone as second-line treatment. A total of 9% of patients had HRD genes mutations and 20% had other DDR genes, not classified as HRD, mutations. A planned interim futility analysis showed that the experimental arm did not have an OS benefit (5.1 vs. 5.9 months; HR 1.3, 95% CI 0.9–2.0, p = 0.21) for biomarker unselected patients, whereas it was likely to be superior to the control arm for patients with HRD in respect to patients without HRD. The incidence of grade 3/4 treatment-related adverse events (AEs), mainly fatigue, neutropenia and nausea, was higher in the veliparib plus chemotherapy arm [45]. The combination of mFOLFOX6 plus veliparib is promising, especially in highly selected patients who are platinum-naïve and have DDR mutation and HBOCS history, but the lack of direct comparison to chemotherapy alone limits the upfront use of this strategy. SWOG S1513 suggests that the inclusion of metastatic PC with any defect in the DNA maintenance system in clinical trials with irinotecan chemotherapy should be pursued. Both trials showed an increased toxicity profile with the combination arms, indicating that the benefit of the combination would potentially come at the cost of an increased toxicity. Further insights will probably be provided from the direct comparison of gemcitabine/cisplatin with and without veliparib in the front-line setting in BRCA1/BRCA2/PALB2-mutated PC [36].
Moving backwards to the neoadjuvant setting, Golan et al. recently showed that borderline resectable PC patients harboring a germline BRCA mutation have an increased chance of achieving a pathological complete response (44.4%, significantly higher than that reported for sporadic PC) and an improved survival after neoadjuvant treatment with FOLFIRINOX [46].

2.2. BRCA Testing for Surveillance Purposes

Based on the above-mentioned therapeutic implications of BRCA1/BRCA2 mutation, the most recent update of the National Comprehensive Cancer Network (NCCN) Pancreatic Cancer guideline (v1. 2020) now recommends germline testing (on peripheral blood) for any patient in clinical practice with confirmed pancreatic cancer, using comprehensive gene panels for hereditary cancer syndromes and performing widely validated methodologies (next-generation sequencing—NGS). The response to PARPi and DNA-damaging agents in PC patients with somatic (tumor) mutation in one or more DNA damage response genes has been evaluated by Lowery et al., who concluded that the presence of those genes failed to improve patient’s response to platinum-based chemotherapies [47]. The mosaicism and heterogeneity of tumor HRD that might be present in the setting of somatic mutations might represent one of the reasons for that result. Clearly, further studies are needed to understand the degree of HRD that somatic mutations might confer since, reasonably, somatic mutations within the HRD pathway and also the BRCAness phenotype are likely to further expand the proportion of PC patients that might benefit from HRD-directed therapies [48].
A genetic testing proposal should occur providing a deep and comprehensive knowledge on all aspects related to possible test results and respecting the decisional time of the patient. Genetic counseling is then recommended for patients who test positive for a pathogenic mutation or for patients with a positive family history of cancer, especially pancreatic cancer, regardless of mutation status [49]. According to the American Society of Clinical Oncology (ASCO), genetic testing is recommended for both affected and unaffected individuals from familial PC families and families with at least three cases of PC diagnoses, but also for individuals for whom testing criteria for hereditary cancer syndromes with high risk of PC are met. Of note, ASCO also suggests that genetic testing should be discussed with any individual diagnosed with pancreatic cancer, even in the presence of an unremarkable family history [50]. Thus, there is now movement in clinical practice toward genetic testing for all patients with pancreatic adenocarcinoma.
Clinical management for PC probands necessarily raises the problem of surveillance, including counseling, for unaffected relatives. The main purpose of surveillance for high-risk individuals (HRIs) is the detection of precursor lesions or early PC, which is the only point at which a surgical (curative) approach may be feasible at present. However, standard screening procedures have not been settled and it is not clear whether screening offers clinical benefit. Numerous studies in the past have failed to show a substantial benefit of screening for pancreatic cancer [21]. More recent studies offered suggestions of benefit in at least some high-risk patients and it is reasonable that stronger evidence supporting surveillance in HRIs derive from long-term follow-up studies compared to single-round ones [51,52,53,54,55]. Currently, no protocols are established, and disagreement remains as to the best screening modality, time of screening initiation or follow-up duration. Nonetheless, there is general agreement that screening is appropriate for individuals at highest risk of developing pancreatic cancer [23]. Some of the promising and conceivable criteria for screening are based upon the International Cancer of the Pancreas Screening (CAPS) Consortium consensus (Table 1) [56,57].
High-risk patients should perform endoscopic ultrasound (EUS) and magnetic resonance imaging (MRI)/magnetic resonance cholangiopancreatography (MRCP), but the scientific community has not reached a consensus on the optimal ages to begin screening or the appropriate screening interval. Generally, it is recommended to begin surveillance at age 50 (or 10 years earlier than the age of the youngest affected relative), with a level of evidence of IV (based on a retrospective cohort or case–control studies) [57]. According to a systematic review including five prospective controlled studies for familial high-risk individuals, subjects in a screening program, mainly by EUS, had a significantly higher curative resection rate (60% vs. 25%) and longer median OS (14.5 months vs. 4.0 months) compared with the control group, although economic and emotional impacts were adverse in the screening group [58]. Regarding the psychological burden of surveillance in particular, in their multicenter prospective trial with follow-up data up to three years, Konings et al. found instead that high-risk individuals feared their next investigation less with the progression of surveillance, with decreasing worries about possible cancer diagnosis and normal or stable levels of depression and anxiety [59]. Accordingly, Paiella et al. stated that PC annual screening with MRCP seemed not to negatively influence HRIs’ psychological wellbeing, with the exception of younger subjects showing higher level of stress [60].
Even in the case a suspicious lesion is detected, no consensus has been reached with respect to the extension of pancreatic resection (partial or total pancreatectomy). In this setting, a multidisciplinary team is needed and surgical intervention must be individualized. In gene mutation carriers without any precursor lesion, prophylactic pancreatectomy is not indicated [57].
Beyond directing pancreatic screening and treatment decision, there are other recognized benefits for genetic testing. Foremost is that the identification of a mutation in the patient will allow for cascade testing of at-risk family members for the same mutation with limited cost and high accuracy. Family members without the mutation will not need pancreatic cancer screening, whereas those with the mutation may. Secondly, identification of a responsible mutation also provides information on other possible cancer risks associated with the mutation/syndrome. For each of the clinically actionable genes on these testing panels, guidelines to direct screening of at-risk individuals (e.g., for breast, ovarian and prostate cancers in BRCA mutation carriers or colon, endometrial and other cancers in Lynch syndrome mutation carriers) are available. In some cases, prophylactic surgery or chemoprevention may also be offered [23].

3. Other Gastrointestinal Cancers

The association between pathogenic mutations in BRCA1 and BRCA2 and other gastrointestinal tumors such as colorectal, gastric cancers, cholangiocarcinoma and hepatocellular carcinoma is unclear. Several population studies conducted over several years reported contradictory results. Therefore, these tumors are not accounted as criteria to select patients for BRCA genetic testing.

3.1. Colorectal Cancer

Colorectal cancer (CRC) is the second cause of cancer death in both men and women in the world. In most cases, CRCs are sporadic. However, different hereditary CRC syndromes exist; the best known are LS (or hereditary nonpolyposis CRC) and FAP. The first one is associated with a mutation in mismatch repair (MMR) genes, or rather MLH1, MSH2, MSH6 or PMS2 [61]. Lynch syndrome-related CRCs represent about 1%–3% of all CRC cases. FAP syndrome is caused by a germline mutation in the APC (adenomatous polyposis coli) gene and it is responsible for nearly 1% of all CRCs [62]. The link between CRC and a mutation in BRCA1/BRCA2 genes is less coded; in fact, individuals affected by CRC are not normally tested either for BRCA1 or for BRCA2.
In 1994, the Breast Cancer Linkage Consortium (BCLC) highlighted a statistically significant increased risk of CRC in a population of BRCA1 mutation carriers (RR = 4.11, 95% CI 2.36–7.15) [63]. The same result was not observed in BRCA2 mutation carriers. In subsequent years, different groups of investigators confirmed or disproved these findings. Thompson and Easton showed a 2-fold increased risk of colon cancer (RR = 2.03, 95% CI 1.45–2.85) and a decreased risk of rectal cancer (RR = 0.23, 95% CI 0.09–0.59) in BRCA1 mutation carriers [64]. Moreover, Brose et al. underlined a 2-fold increased risk (11%, 95% CI 8.2%–13.2%) to develop CRC in BRCA1 carriers compared to the risk reported by Surveillance, Epidemiology, and End Results (SEER) [65]. Phelan and his collaborators screened 7015 women carrying a BRCA mutation and found twenty-one CRC cases, an incidence not higher than that of the general population. Nevertheless, they observed an increased risk to develop CRC in women younger than 50 years carrying a BRCA1 pathogenic mutation. Instead, no differences with global population rates were reported in older women and in BRCA2 mutation carriers [66]. These findings are consistent with the results of other studies. Indeed, in a Polish population of 2398 unselected patients affected by CRC, Suchy et al. reported a mutation detection rate of about 0.42% (not higher than the rate of the control group, that was 0.48%). However, also in this trial, a major incidence of CRC in patients younger than 60 years (OR = 1.7) was underlined, though not statistically significant (p = 0.3) [67]. Thus, women with a BRCA1 mutation should undergo a CRC screening test, such as a high-sensitivity fecal occult blood test or colonoscopy at a younger age. Notably, they may also be good candidates for chemoprevention programs with low-dose aspirin. Indeed, several trials highlighted a decrease in CRC incidence in subjects taking daily aspirin at the dosage of ≥75 mg/day [68]. Particularly, Burn et al. tested 600 mg/day of aspirin in patients with Lynch syndrome and observed a reduction in CRC incidence [69]. However, since the regular use of acetylsalicylic acid might cause severe AEs such as cerebral and gastrointestinal bleedings, the US Preventive Services Task Force recommends chemoprevention only for high-risk individuals, e.g., Lynch syndrome or FAP individuals [70].
Mersch et al. investigated the risk of developing CRC in a group of 613 BRCA1 and 459 BRCA2 mutation carriers and found no statistically significant difference between carriers and non-carriers [71]. Moreover, in a cohort study, Lin and colleagues observed no significantly different risk in 164 BRCA1 and 88 BRCA2 mutation carriers compared to the general population [72]. Other studies investigated if a family history of breast cancer was associated with a higher CRC incidence. Niell and colleagues did not identify a correlation between a family history of breast cancer in a first-degree female relative and the risk of developing CRC [73]. Conversely, Slattery and Kerber reported a low, but statistically significant, increased risk for CRC in patients with a positive family history for breast cancer [74]. Of note, several hereditary syndromes could be involved and could explain the association. Peutz–Jeghers syndrome, Cowden syndrome and Muir–Torre syndrome are just some examples of inherited conditions with a spectrum of diseases in which both cancers (breast cancer and CRC) are accounted. Furthermore, APC polymorphism I1307K, as reported by Woodage et al. and Redston and colleagues, might be associated with low penetrance to breast cancer susceptibility [75,76].
In summary, some family-based studies and prospective cohort studies suggested a possible greater risk among early-onset CRCs in BRCA1 mutation carriers. These results seem to be confirmed by a systematic review and meta-analysis underlining a 1.49-fold higher risk of CRC in BRCA1 mutation carriers [77]. Nevertheless, more studies are needed to investigate the real linkage.
In recent years, also in the context of CRC, new and old drugs demonstrated efficacy in HRD conditions. In their case report, Lin et al. described a complete pathological response in a young man affected by rectal cancer carrying a BRCA2 pathogenic mutation with a platinum-based neoadjuvant chemotherapy [78]. The authors also reported a high tumor mutational burden (TMB), investigated by next-generation sequencing (NGS), without microsatellite instability (MSI), in their patient. Therefore, based on previous evidence for other cancers, they speculated also a possible rationale for the use of checkpoint inhibitors [78,79,80]. In their study involving 6396 CRC tumor samples, Naseem et al. recently detected BRCA1 and BRCA2 mutations in 1.1% and 2.8% of tumors, respectively. Interestingly, they found a higher frequency of BRCA1 and BRCA2 mutations in MSI-high (MSI-H) patients and found that those mutations were independently associated with higher TMB. Therefore, Naseem et al. also came to the conclusion that BRCA1 and BRCA2 mutations might potentially be predictive biomarkers for checkpoint inhibitors in CRC [81]. Notably, Harpaz and collaborators found a statistically significantly higher incidence of BRCA mutations and a higher TMB in CRCs with mucinous histology, compared to adenocarcinomas, suggesting that this association might lead to the use of histopathologic characterization, besides other tests, to identify patients who may be good candidates for immunotherapy [82].
As previously mentioned, PARPi are novel therapeutic agents. They currently play a very important role mostly in the treatment of ovarian cancer. Earlier studies investigated the role of ABT-888 (veliparib) in a CRC cell line pretreated with DNA-damaging chemotherapy agents, such as irinotecan and oxaliplatin [83], reporting a synergistic effect. ABT-888 showed a synergistic effect also in combination with radiation [84]. Based on these results, a phase II open-label study evaluating the action of veliparib in combination with temozolomide in metastatic CRC patients successfully met its primary endpoint with a disease control rate (DCR) of 24% and two confirmed partial responses [85]. However, PARPi demonstrated their efficacy also when a mutation occurred in the so-called BRCAness genes, such as in ATM. Wang et al. highlighted that CRC cell lines with an ATM-inactivating mutation had an increased sensitivity to olaparib [86].
Based on previous studies on myeloid malignancy [87], Leichman et al. tested a PARPi in microsatellite-stable (MSS) and -unstable (MSI) CRC patients in a phase II clinical trial. No differences between the two groups were observed. Therefore, the authors reported that microsatellite status is not a predictive marker of response to PARPi [88]. Certainly, more studies are needed, and for the time being, PARPi are not approved for the treatment of CRC.

3.2. Gastric Cancer

Gastric cancer (GC) is a heterogeneous disease, mostly sporadic, but hereditary in a small percentage of cases (1%–3%). Familial intestinal GC (FIGC) and hereditary diffuse GC (HDGC, ORPHA: 26106) are the principal hereditary GC conditions. HDGC syndrome is caused by a mutation in CDH1, the gene encoding for the E-cadherin protein, and it is characterized by an association between signet ring cell/diffuse GC and lobular breast cancer [89]. GC is also a key component of other hereditary cancer syndromes such as Lynch Syndrome, Li–Fraumeni syndrome (ORPHA:524, gene TP53) and Peutz–Jeghers Syndrome. Furthermore, GC is accounted in hereditary breast/ovarian cancer syndrome (HBOCS).
The BCLC reported a 6-fold increased risk of GC among first-degree relatives of both BRCA genes mutation carriers [28,63]. Brose et al. estimated a 4-fold higher lifetime risk to develop GC in BRCA1 mutation carriers [65]. Tulinius and colleagues investigated the risk of developing GC in 995 women and found a 2-fold greater risk in the BRCA2 mutation-positive cohort [90]. Conversely, van Asperen and collaborators highlighted no statistically significant higher risk of developing GC in BRCA2 families in the Dutch population [91]. Some authors explained the contradictory results, arguing that breast and ovarian cancers have an earlier onset in BRCA1/BRCA2 mutation carriers, therefore patients might not have time to develop GC afterwards. Notably, in their population-based study, Bermejo and colleagues found a major incidence of GC in males. Particularly, in 23 families with ovarian, breast and gastric cancers, they reported 23 GC cases in males and only 1 case in females [92]. Previously, also BCLC suggested a sex-related increased incidence of GC in males [28].
As mentioned above, an impairment in the proteins involved in HR causes a higher susceptibility to PARPi. A phase II study reported a significant improvement in OS with the combination of olaparib plus paclitaxel in Asian patients with advanced GC, especially in ATM mutation carriers [93]. A subsequent phase III trial (GOLD trial) did not confirm these results, showing an OS of 8.8 months (95% CI 7.4–9.6) in the olaparib group vs. 6.9 months (95% CI 6.3–7.9) in the placebo group. Moreover, among the ATM mutation carriers, the OS was 12 months (95% CI 7.8–18.1) in the experimental arm vs. 10 months (95% CI 6.4–13.3) in the standard arm [94].
Several other combination therapies were investigated. A phase II basket study demonstrated the tolerability and the reasonable efficacy (ORR 10%) of the association between olaparib and durvalumab, an anti PD-L1 monoclonal antibody, in patients with relapsed GC [95]. Currently, a study evaluating the combination of olaparib and ramucirumab (an angiogenesis inhibitor) is ongoing [96].

3.3. Cholangiocarcinoma and Hepatocellular Carcinoma

Cholangiocarcinoma (CCA) is the second most common hepatic neoplasm after hepatocellular carcinoma. In recent years, with the purpose of improving CCA treatment, Nakamura et al. found in a series of CCA several somatic alterations in potentially targetable genes, such as kinases FGFR1, FGFR2, FGFR3, AKT3, BRAF, PIK3CA, EGFR and ALK and oncogenes MDM2, CCND3, CCND1, IDH1 and IDH2, but also in the tumor suppressor proteins BRCA1 and BRCA2 [97]. Moreover, Churi and his collaborators also highlighted targetable somatic mutations in MSH2, MLH1, ATM, BAP1, MSH6, BRCA1 and BRCA2 in 74 CCA cases [98]. The role of BRCA1 and BRCA2 in the pathogenesis of CCA was primarily suggested by BCLC. In 1999, they described a relative risk (RR) of developing CCA of about 4.97 (95% CI 1.50–16.52) among BRCA2 mutation carriers [28]. Encouraged by these results, Golan and other authors identified 18 cases of CCA with genetic alterations in BRCA1 and BRCA2 genes: five of those were germline, thirteen were somatic mutations. Thirteen CCA patients were treated with a platinum-based chemotherapy and four patients received PARPi. Notably, one of the patients treated with PARPi experienced a progression-free survival of 42.6 months [99]. Cheng et al. also described a good response with olaparib monotherapy in a patient affected by intrahepatic CCA [100]. Recently, Spizzo et al. analyzed 1288 CCA samples and detected BRCA1 and BRCA2 mutations in 46 cases, at 3.6% (0.6% BRCA1 and 3% BRCA2). They also underscored that these mutations were associated with a high mutational burden and suggested a potential rationale for the combination of PARPi with immunotherapies [101]. More evidence for the use of PARPi in CCA comes from pre-clinical experiences. Indeed, Fehling et al. highlighted a synergistic action of BET inhibitors (JQ1) with PARPi in CCA cell lines [102]. Another pre-clinical study suggested a potential role of olaparib in sensitizing CCA cells to radiation [103], as reported above for CRC. Moving from the benchside to the bedside, clinical trials are currently ongoing. For instance, a phase II trial is evaluating olaparib in patients with metastatic CCA and aberrant DNA repair genes (BRCA1, BRCA2, ATM, RAD51 and others) [104]. The results of this trial, which are expected in 2021, and those from other trials are needed to clarify, firstly, the real role of BRCA1 and BRCA2 mutations in CCA, and secondly, the potential role for PARPi use either in monotherapy or in combination with other drugs.
Regarding hepatocellular carcinoma (HCC), evidence about its link with BRCA1 and BRCA2 mutations is extremely limited. In a recent study, Lin J. and colleagues analyzed a population of 357 patients with primary liver cancers: 214 HCC, 122 CCA and 21 mixed HCC and CCA. They found a mutation in BRCA1 or BRCA2 genes only in five HCC patients. However, they reported an ATM mutation rate of 6.07%, higher than that of CCA. Clearly, because of the scarce data available, the real role of BRCA proteins in the pathogenesis of HCC cannot be postulated [105].

3.4. Gastrointestinal Cancer Minorities

Lastly, but not least, coming to gastrointestinal cancer minorities, evidence is still little or absent. Recently, Hännimen et al. [106] and Quaas et al. [107] highlighted a possible role of BRCA1 and BRCA2 mutations in small bowel cancer pathogenesis, but data are very preliminary. The linkage between gastrointestinal tumors and BRCA mutation is currently drawing more and more attention and since PARPi might open a new therapeutic scenario in a wide range of cancers, potentially also in rare and orphan ones, evidence is expected to increasingly grow in the near future.

4. Future Perspectives

Remarkable progress on the genomic profiling of gastrointestinal cancers has been achieved in recent years. Based on encouraging results in breast and ovarian cancers, understanding the role of BRCA1/BRCA2 mutations in the pathogenesis, prognosis and therapeutic decision of gastrointestinal malignancies, especially in PC, has gained particular interest. Moreover, the presence of BRCA1/BRCA2 mutations is associated with increased risk for pancreatic cancer [108,109], but its role in the other gastrointestinal malignancies is still to be defined. In addition, the effect of BRAC1/2 mutations on the survival of these patients remains controversial in the literature [109,110]. Therefore, large-scale genetic testing seems to be necessary in order to clarify doubts and broaden our knowledge in this field.
As mentioned above, increased evidence underlines the potential role of BRCA1/BRCA2 mutations as predictive markers for the efficacy of platinum-based agents and other treatments in various types of cancer because of DNA repair defects [34,37,111,112]. In particular, pathological complete responses to platinum agents have been described in patients affected by BRCA-defective gastrointestinal tumors, whereas their efficacy in gastrointestinal tumors with somatic BRCA mutations still remains controversial [113,114]. Actually, several clinical trials are ongoing to evaluate the association between BRCA1/BRCA2 gene mutations and platinum sensitivity in several tumors including gastrointestinal malignancies (Table 2).
Based on promising results of breast and ovarian cancers [115,116], several clinical trials of PARPi are undergoing for gastrointestinal malignancies, especially for pancreatic cancers. Recently, a phase III trial demonstrated that PARPi can significantly improve outcomes as maintenance treatment in patients affected by platinum-sensitive advanced pancreatic cancer harboring germline BRCA mutations [38]. However, the effect of these novel agents in tumors with somatic BRCA mutations is still not well defined. Therefore, randomized trials of PARPi alone or in association with chemotherapy in gastrointestinal tumors are ongoing; their results are expected to clarify the position of PARPi in the therapeutic armamentarium of these tumors.
Considering the relationship between high mutational burden and BRCA mutations reported for breast and ovarian cancers, these characteristics provide a rationale for the evaluation of immune checkpoint inhibitors in BRCA-mutated tumors [117,118]. However, further studies are required to better define the potential clinical benefit of immunotherapy alone or in combination with PARPi or platinum agents on this subset of gastrointestinal tumors. Hopefully, several ongoing trials will answer some of these unresolved issues, trying to improve the therapeutic landscape of GI malignancies in the coming years (Table 2).
Challenges for the future are definitely many. Firstly, the genetic background of patients affected by gastrointestinal tumors with BRCA mutations seems to be heterogeneous and underexamined and many questions remain to be explored; use of sophisticated software risk assessment tools may facilitate their better identification. Secondly, additional studies are necessary to better clarify the prevalence and penetrance of BRCA1/BRCA2 mutations, and whether they might not be driver but rather passenger mutations only, and consequently to define the impact of available genomic information with clinical outcomes and eventually of personalized therapeutic approaches. Thirdly, with the advent of emerging agents (PARPi and immunotherapy) and their possible association with cytotoxic agents, further studies are necessary to define the most suitable use of such agents depending on disease status. Fourthly, since only a subset of patients with BRCA mutations seems to benefit from available treatments, there is a compelling need to identify predictive biomarkers able to direct the right treatment to the right patient. Finally, since clinical and genomic data about BRCA-mutated gastrointestinal cancers are still limited, though progressively expanding, international spreading of hereditary gastrointestinal tumors registries seems to be crucial for future studies.

5. Conclusions

In summary, a strong association between pancreatic cancer and BRCA1 and BRCA2 mutations is documented. Indeed, the POLO trial results can change the approach to newly diagnosed pancreatic cancers. Contrariwise, for other gastrointestinal tumors, the association is currently only alleged. Some colorectal epidemiological studies speculate a greater incidence of colorectal cancer in women younger than 50 years carrying a BRCA1 mutation. A lower link emerged for gastric cancer and cholangiocarcinoma. Notably, for gastric cancer, a major incidence in males is presumed. Several pre-clinical studies and clinical trials, also in the absence of a genetic predisposition, investigated the effects of PARPi in combination with DNA-harmful agents (e.g., radiation, chemotherapies), which appear to be amplified in BRCA-defective patients.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Toss, A.; Cortesi, L. Molecular Mechanisms of PARP Inhibitors in BRCA-related Ovarian Cancer. J. Cancer Sci. Ther. 2013, 5, 409–416. [Google Scholar] [CrossRef]
  2. Symington, L.S.; Gautier, J. Double-Strand Break End Resection and Repair Pathway Choice. Annu. Rev. Genet. 2011, 45, 247–271. [Google Scholar] [CrossRef]
  3. Stoppa-Lyonnet, D. The biological effects and clinical implications of BRCA mutations: Where do we go from here? Eur. J. Hum. Genet. 2016, 24, S3–S9. [Google Scholar] [CrossRef]
  4. Toss, A.; Molinaro, E.; Sammarini, M.; Del Savio, M.C.; Cortesi, L.; Facchinetti, F.; Grandi, G. Hereditary ovarian cancers: State of the art. Minerva Med. 2019, 110, 301–319. [Google Scholar] [CrossRef] [PubMed]
  5. Cortesi, L.; De Matteis, E.; Toss, A.; Marchi, I.; Medici, V.; Contu, G.; Xholli, A.; Grandi, G.; Cagnacci, A.; Federico, M. Evaluation of Transvaginal Ultrasound plus CA-125 Measurement and Prophylactic Salpingo-Oophorectomy in Women at Different Risk Levels of Ovarian Cancer: The Modena Study Group Cohort Study. Oncology 2017, 93, 377–386. [Google Scholar] [CrossRef] [PubMed]
  6. Tyrer, J.; Duffy, S.W.; Cuzick, J. A breast cancer prediction model incorporating familial and personal risk factors. Stat. Med. 2004, 23, 1111–1130. [Google Scholar] [CrossRef] [Green Version]
  7. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.J.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.D.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nat. Cell Biol. 2005, 434, 917–921. [Google Scholar] [CrossRef]
  8. Bolton, K.L.; Chenevix-Trench, G.; Goh, C.; Sadetzki, S.; Ramus, S.J.; Karlan, B.; Lambrechts, D.; Despierre, E.; Barrowdale, D.; McGuffog, L.; et al. Association Between BRCA1 and BRCA2 Mutations and Survival in Women with Invasive Epithelial Ovarian Cancer. JAMA 2012, 307, 382–390. [Google Scholar] [CrossRef] [Green Version]
  9. Kurnit, K.; Coleman, R.L.; Westin, S.N. Using PARP Inhibitors in the Treatment of Patients with Ovarian Cancer. Curr. Treat. Opt. Oncol. 2018, 19, 1. [Google Scholar] [CrossRef]
  10. Murai, J.; Huang, S.-Y.N.; Renaud, A.; Zhang, Y.; Ji, J.; Takeda, S.; Morris, J.; Teicher, B.; Doroshow, J.H.; Pommier, Y. Stereospecific PARP Trapping by BMN 673 and Comparison with Olaparib and Rucaparib. Mol. Cancer Ther. 2014, 13, 433–443. [Google Scholar] [CrossRef] [Green Version]
  11. Toss, A.; Venturelli, M.; Molinaro, E.; Pipitone, S.; Barbieri, E.; Marchi, I.; Tenedini, E.; Artuso, L.; Castellano, S.; Marino, M.; et al. Hereditary Pancreatic Cancer: A Retrospective Single-Center Study of 5143 Italian Families with History of BRCA-Related Malignancies. Cancers 2019, 11, 193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Silvestri, V.; Leslie, G.; Barnes, D.R.; Agnarsson, B.A.; Aittomäki, K.; Alducci, E.; Andrulis, I.L.; Barkardottir, R.B.; Barroso, A.; CIMBA Group; et al. Characterization of the Cancer Spectrum in Men with Germline BRCA1 and BRCA2 Pathogenic Variants: Results from the Consortium of Investigators of Modifiers of BRCA1/2 (CIMBA). JAMA Oncol. 2020, 6, 1218–1230. [Google Scholar] [CrossRef] [PubMed]
  13. 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] [Green Version]
  14. Lynch, S.M.; Vrieling, A.; Lubin, J.H.; Kraft, P.; Mendelsohn, J.B.; Hartge, P.; Canzian, F.; Steplowski, E.; Arslan, A.A.; Gross, M.; et al. Cigarette Smoking and Pancreatic Cancer: A Pooled Analysis from the Pancreatic Cancer Cohort Consortium. Am. J. Epidemiol. 2009, 170, 403–413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Iodice, S.; Gandini, S.; Maisonneuve, P.; Lowenfels, A.B. Tobacco and the risk of pancreatic cancer: A review and meta-analysis. Langenbeck Arch. Surg. 2008, 393, 535–545. [Google Scholar] [CrossRef]
  16. Everhart, J.; Wright, D. Diabetes mellitus as a risk factor for pancreatic cancer. A meta-analysis. JAMA 1995, 273, 1605–1609. [Google Scholar] [CrossRef]
  17. Jiao, L.; Berrington de Gonzalez, A.; Hartge, P.; Pfeiffer, R.M.; Park, Y.; Freedman, D.M.; Gail, M.H.; Alavanja, M.C.; Albanes, D.; Beane Freeman, L.E.; et al. Body mass index, effect modifiers, and risk of pancreatic cancer: A pooled study of seven prospective cohorts. Cancer Causes Control. 2010, 21, 1305–1314. [Google Scholar] [CrossRef]
  18. Klein, A.P. Genetic susceptibility to pancreatic cancer. Mol. Carcinog. 2012, 51, 14–24. [Google Scholar] [CrossRef] [Green Version]
  19. Welinsky, S.; Lucas, A.L. Familial Pancreatic Cancer and the Future of Directed Screening. Gut Liver 2017, 11, 761–770. [Google Scholar] [CrossRef] [Green Version]
  20. Matsubayashi, H.; Takaori, K.; Morizane, C.; Maguchi, H.; Mizuma, M.; Takahashi, H.; Wada, K.; Hosoi, H.; Yachida, S.; Suzuki, M.; et al. Familial pancreatic cancer: Concept, management and issues. World J. Gastroenterol. 2017, 23, 935–948. [Google Scholar] [CrossRef]
  21. Ohmoto, A.; Yachida, S.; Morizane, C. Genomic Features and Clinical Management of Patients with Hereditary Pancreatic Cancer Syndromes and Familial Pancreatic Cancer. Int. J. Mol. Sci. 2019, 20, 561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Rebelatto, T.F.; Falavigna, M.; Pozzari, M.; Spada, F.; Cella, C.A.; Laffi, A.; Pellicori, S.; Fazio, N. Should platinum-based chemotherapy be preferred for germline BReast CAncer genes (BRCA) 1 and 2-mutated pancreatic ductal adenocarcinoma (PDAC) patients? A systematic review and meta-analysis. Cancer Treat. Rev. 2019, 80, 101895. [Google Scholar] [CrossRef] [PubMed]
  23. Pilarski, R. The Role of BRCA Testing in Hereditary Pancreatic and Prostate Cancer Families. Am. Soc. Clin. Oncol. Educ. Book 2019, 39, 79–86. [Google Scholar] [CrossRef] [PubMed]
  24. Shindo, K.; Yu, J.; Suenaga, M.; Fesharakizadeh, S.; Cho, C.; Macgregor-Das, A.; Siddiqui, A.; Witmer, P.D.; Tamura, K.; Song, T.J.; et al. Deleterious Germline Mutations in Patients with Apparently Sporadic Pancreatic Adenocarcinoma. J. Clin. Oncol. 2017, 35, 3382–3390. [Google Scholar] [CrossRef] [PubMed]
  25. Peters, M.L.; Tseng, J.F.; Miksad, R.A. Genetic Testing in Pancreatic Ductal Adenocarcinoma: Implications for Prevention and Treatment. Clin Ther. 2016, 38, 1622–1635. [Google Scholar] [CrossRef] [Green Version]
  26. Young, E.L.; Thompson, B.A.; Neklason, D.W.; Firpo, M.A.; Werner, T.; Bell, R.; Berger, J.; Fraser, A.; Gammon, A.; Koptiuch, C.; et al. Pancreatic cancer as a sentinel for hereditary cancer predisposition. BMC Cancer 2018, 18, 697. [Google Scholar] [CrossRef] [Green Version]
  27. Iqbal, J.; Ragone, A.V.; Lubinski, J.; Lynch, H.T.; Moller, P.; Ghadirian, P.; Foulkes, W.D.; Armel, S.; Eisen, A.Z.; Neuhausen, S.L.; et al. The incidence of pancreatic cancer in BRCA1 and BRCA2 mutation carriers. Br. J. Cancer 2012, 107, 2005–2009. [Google Scholar] [CrossRef]
  28. The Breast Cancer Linkage Consortium. Cancer Risks in BRCA2 Mutation Carriers. J. Natl. Cancer Inst. 1999, 91, 1310–1316. [Google Scholar] [CrossRef]
  29. Roberts, N.J.; Norris, A.L.; Petersen, G.M.; Bondy, M.L.; Brand, R.E.; Gallinger, S.; Kurtz, R.C.; Olson, S.H.; Rustgi, A.K.; Schwartz, A.G.; et al. Whole Genome Sequencing Defines the Genetic Heterogeneity of Familial Pancreatic Cancer. Cancer Discov. 2015, 6, 166–175. [Google Scholar] [CrossRef] [Green Version]
  30. Reiss, K.A.; Yu, S.; Judy, R.; Symecko, H.; Nathanson, K.L.; Domchek, S.M. Retrospective Survival Analysis of Patients with Advanced Pancreatic Ductal Adenocarcinoma and Germline BRCA or PALB2 Mutations. JCO Precis. Oncol. 2018, 2, 1–9. [Google Scholar] [CrossRef]
  31. 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. J. Natl. Cancer Inst. 2018, 110, 1067–1074. [Google Scholar] [CrossRef] [Green Version]
  32. Golan, T.; Sella, T.; O’Reilly, E.; 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] [Green Version]
  33. Blair, A.B.; Groot, V.P.; Gemenetzis, G.; Wei, J.; Cameron, J.L.; Weiss, M.J.; Goggins, M.; Wolfgang, C.L.; Yu, J.; He, J. BRCA1/BRCA2 Germline Mutation Carriers and Sporadic Pancreatic Ductal Adenocarcinoma. J. Am. Coll. Surg. 2018, 226, 630–637.e1. [Google Scholar] [CrossRef] [PubMed]
  34. Waddell, N.; Initiative, A.P.C.G.; Pajic, M.; Patch, A.-M.; Chang, D.K.; Kassahn, K.S.; Bailey, P.; Johns, A.L.; Miller, D.; Nones, K.; et al. Whole genomes redefine the mutational landscape of pancreatic cancer. Nat. Cell Biol. 2015, 518, 495–501. [Google Scholar] [CrossRef] [Green Version]
  35. Pant, S.; Maitra, A.; Yap, T.A. PARP inhibition—Opportunities in pancreatic cancer. Nat. Rev. Clin. Oncol. 2019, 16, 595–596. [Google Scholar] [CrossRef]
  36. Gupta, M.; Iyer, R.; Fountzilas, C. Poly(ADP-Ribose) Polymerase Inhibitors in Pancreatic Cancer: A New Treatment Paradigms and Future Implications. Cancers 2019, 11, 1980. [Google Scholar] [CrossRef] [Green Version]
  37. Dedes, K.J.; Wilkerson, P.M.; Wetterskog, D.; Weigelt, B.; Ashworth, A.; Reis-Filho, J.S. Synthetic lethality of PARP inhibition in cancers lackingBRCA1andBRCA2mutations. Cell Cycle 2011, 10, 1192–1199. [Google Scholar] [CrossRef] [Green Version]
  38. 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]
  39. Hammel, P.; Kindler, H.L.; Reni, M.; Van Cutsem, E.; Macarulla, T.; Hall, M.J.; Park, J.O.; Hochhauser, D.; Arnold, D.; Oh, D.-Y.; et al. Health-related quality of life in patients with a germline BRCA mutation and metastatic pancreatic cancer receiving maintenance olaparib. Ann. Oncol. 2019, 30, 1959–1968. [Google Scholar] [CrossRef] [Green Version]
  40. Wattenberg, M.M.; Asch, D.; Yu, S.; O’Dwyer, P.J.; Domchek, S.M.; Nathanson, K.L.; Rosen, M.A.; Beatty, G.L.; Siegelman, E.S.; Reiss, K.A. Platinum response characteristics of patients with pancreatic ductal adenocarcinoma and a germline BRCA1, BRCA2 or PALB2 mutation. Br. J. Cancer 2019, 122, 333–339. [Google Scholar] [CrossRef]
  41. Lowery, M.A. Genotype–phenotype correlation in BRCA1/2 mutation-associated pancreatic cancer. Br. J. Cancer 2019, 122, 293–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Hu, C.; Hart, S.N.; Polley, E.C.; Gnanaolivu, R.; Shimelis, H.; Lee, K.Y.; Lilyquist, J.; Na, J.; Moore, R.M.; Antwi, S.O.; et al. Association Between Inherited Germline Mutations in Cancer Predisposition Genes and Risk of Pancreatic Cancer. JAMA 2018, 319, 2401–2409. [Google Scholar] [CrossRef] [PubMed]
  43. Pilié, P.G.; Tang, C.; Mills, G.B.; Yap, T.A. State-of-the-art strategies for targeting the DNA damage response in cancer. Nat. Rev. Clin. Oncol. 2019, 16, 81–104. [Google Scholar] [CrossRef] [PubMed]
  44. Pishvaian, M.J.; Hongkun Wang, H.; Parenti, S.; He, A.R.; Hwang, J.J.; Ley, L.; Difebo, H.; Smaglo, B.G.; Kim, S.S.; Weinberg, B.A.; et al. Final report of a phase I/II study of veliparib (Vel) in combination with 5-FU and oxaliplatin (FOLFOX) in patients (pts) with metastatic pancreatic cancer (mPDAC). J. Clin. Oncol. 2019, 37, 4015. [Google Scholar] [CrossRef]
  45. 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 second-line modified FOLFIRI with PARP inhibitor ABT-888 (Veliparib) (NSC-737664) versus FOLFIRI in metastatic pancreatic cancer (mPC): SWOG S1513. J. Clin. Oncol. 2019, 37, 4014. [Google Scholar] [CrossRef]
  46. Golan, T.; Barenboim, A.; Lahat, G.; Nachmany, I.; Goykhman, Y.; Shacham-Shmueli, E.; Halpern, N.; Brazowski, E.; Geva, R.; Wolf, I.; et al. Increased Rate of Complete Pathologic Response After Neoadjuvant FOLFIRINOX for BRCA Mutation Carriers with Borderline Resectable Pancreatic Cancer. Ann. Surg. Oncol. 2020, 27, 3963–3970. [Google Scholar] [CrossRef]
  47. Lowery, M.A.; Jordan, E.J.; Basturk, O.; Ptashkin, R.N.; Zehir, A.; Berger, M.; Leach, T.; Herbst, B.; Askan, G.; Maynard, H.; et al. Real-Time Genomic Profiling of Pancreatic Ductal Adenocarcinoma: Potential Actionability and Correlation with Clinical Phenotype. Clin. Cancer Res. 2017, 23, 6094–6100. [Google Scholar] [CrossRef] [Green Version]
  48. Wong, W.; Raufi, A.G.; Safyan, R.A.; Bates, S.E.; Manji, G.A. BRCA Mutations in Pancreas Cancer: Spectrum, Current Management, Challenges and Future Prospects. Cancer Manag. Res. 2020, 12, 2731–2742. [Google Scholar] [CrossRef] [Green Version]
  49. Tempero, M.A.; Malafa, M.P.; Chiorean, E.G.; Czito, B.; Scaife, C.; Narang, A.K.; Fountzilas, C.; Wolpin, B.M.; Al-Hawary, M.; Asbun, H.; et al. Guidelines Insights: Pancreatic Adenocarcinoma, Version 1.2019. J. Natl. Compr. Cancer Netw. 2019, 17, 202–210. [Google Scholar] [CrossRef] [Green Version]
  50. Stoffel, E.M.; McKernin, S.E.; Brand, R.; Canto, M.; Goggins, M.; Moravek, C.; Nagarajan, A.; Petersen, G.M.; Simeone, D.M.; Yurgelun, M.; et al. Evaluating Susceptibility to Pancreatic Cancer: ASCO Provisional Clinical Opinion. J. Clin. Oncol. 2019, 37, 153–164. [Google Scholar] [CrossRef]
  51. Vasen, H.F.; Ibrahim, I.; Ponce, C.G.; Slater, E.P.; Matthäi, E.; Carrato, A.; Earl, J.; Robbers, K.; Van Mil, A.M.; Potjer, T.; et al. Benefit of Surveillance for Pancreatic Cancer in High-Risk Individuals: Outcome of Long-Term Prospective Follow-Up Studies from Three European Expert Centers. J. Clin. Oncol. 2016, 34, 2010–2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Paiella, S.; Capurso, G.; Cavestro, G.M.; Butturini, G.; Pezzilli, R.; Salvia, R.; Signoretti, M.; Crippa, S.; Carrara, S.; Frigerio, I.; et al. Results of First-Round of Surveillance in Individuals at High-Risk of Pancreatic Cancer from the AISP (Italian Association for the Study of the Pancreas) Registry. Am. J. Gastroenterol. 2019, 114, 665–670. [Google Scholar] [CrossRef] [PubMed]
  53. Paiella, S.; Salvia, R.; De Pastena, M.; Pollini, T.; Casetti, L.; Landoni, L.; Esposito, A.; Marchegiani, G.; Malleo, G.; De Marchi, G.; et al. Screening/surveillance programs for pancreatic cancer in familial high-risk individuals: A systematic review and proportion meta-analysis of screening results. Pancreatology 2018, 18, 420–428. [Google Scholar] [CrossRef] [PubMed]
  54. Signoretti, M.; Bruno, M.J.; Zerboni, G.; Poley, J.-W.; Fave, G.D.; Capurso, G. Results of surveillance in individuals at high-risk of pancreatic cancer: A systematic review and meta-analysis. United Eur. Gastroenterol. J. 2018, 6, 489–499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Canto, M.I.; Almario, J.A.; Schulick, R.D.; Yeo, C.J.; Klein, A.; Blackford, A.; Shin, E.J.; Sanyal, A.; Yenokyan, G.; Lennon, A.M.; et al. Risk of Neoplastic Progression in Individuals at High Risk for Pancreatic Cancer Undergoing Long-term Surveillance. Gastroenterology 2018, 155, 740–751.e2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Goggins, M.; Overbeek, K.A.; Brand, R.; Syngal, S.; Del Chiaro, M.; Bartsch, D.K.; Bassi, C.; Carrato, A.; Farrell, J.; Fishman, E.K.; et al. Management of patients with increased risk for familial pancreatic cancer: Updated recommendations from the International Cancer of the Pancreas Screening (CAPS) Consortium. Gut 2019, 69, 7–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Stjepanovic, N.; Moreira, L.; Carneiro, F.; Balaguer, F.; Cervantes, A.; Balmaña, J.; Martinelli, E. Hereditary gastrointestinal cancers: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2019, 30, 1558–1571. [Google Scholar] [CrossRef] [Green Version]
  58. Lu, C.; Xu, C.F.; Wan, X.Y.; Zhu, H.T.; Yu, C.H.; Li, Y.M. Screening for pancreatic cancer in familial high-risk individuals: A systematic review. World J. Gastroenterol. 2015, 21, 8678–8686. [Google Scholar] [CrossRef]
  59. Konings, I.C.A.W.; Sidharta, G.N.; Harinck, F.; Aalfs, C.M.; Poley, J.W.; Kieffer, J.M.; Kuenen, M.A.; Smets, E.M.A.; Wagner, A.; Van Hooft, J.E.; et al. Repeated participation in pancreatic cancer surveillance by high-risk individuals imposes low psychological burden. Psycho Oncol. 2015, 25, 971–978. [Google Scholar] [CrossRef]
  60. Paiella, S.; Marinelli, V.; Secchettin, E.; Mazzi, M.A.; Ferretto, F.; Casolino, R.; Bassi, C.; Salvia, R. The emotional impact of surveillance programs for pancreatic cancer on high-risk individuals: A prospective analysis. Psycho Oncol. 2020, 29, 1004–1011. [Google Scholar] [CrossRef]
  61. Lynch, H.T.; Shaw, M.W.; Magnuson, C.W.; Larsen, A.L.; Krush, A.J. Hereditary factors in cancer. Study of two large midwestern kindreds. Arch. Intern. Med. 1966, 117, 206–212. [Google Scholar] [CrossRef] [PubMed]
  62. Karstensen, J.G.; Burisch, J.; Pommergaard, H.-C.; Aalling, L.; Højen, H.; Jespersen, N.; Schmidt, P.N.; Bülow, S. Colorectal Cancer in Individuals with Familial Adenomatous Polyposis, Based on Analysis of the Danish Polyposis Registry. Clin. Gastroenterol. Hepatol. 2019, 17, 2294–2300.e1. [Google Scholar] [CrossRef] [PubMed]
  63. Ford, D.; Easton, D.F.; Bishop, D.T.; Narod, S.A.; Goldgar, D.E. Risks of cancer in BRCA1-mutation carriers. Breast Cancer Linkage Consortium. Lancet 1994, 343, 692–695. [Google Scholar] [CrossRef]
  64. Thompson, D.; Easton, D.F. Breast Cancer Linkage Consortium. Cancer Incidence in BRCA1 mutation carriers. J. Natl. Cancer Inst. 2002, 94, 1358–1365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Brose, M.S.; Rebbeck, T.R.; Calzone, K.A.; Stopfer, J.E.; Nathanson, K.L.; Weber, B.L. Cancer Risk Estimates for BRCA1 Mutation Carriers Identified in a Risk Evaluation Program. J. Natl. Cancer Inst. 2002, 94, 1365–1372. [Google Scholar] [CrossRef]
  66. Phelan, C.M.; Iqbal, J.; Lynch, H.T.; Lubinski, J.; Gronwald, J.; Moller, P.; Ghadirian, P.; Foulkes, W.D.; Armel, S.; Eisen, A.; et al. Incidence of colorectal cancer in BRCA1 and BRCA2 mutation carriers: Results from a follow-up study. Br. J. Cancer 2013, 110, 530–534. [Google Scholar] [CrossRef] [Green Version]
  67. Suchy, J.; Cybulski, C.; Górski, B.; Huzarski, T.; Byrski, T.; Dębniak, T.; Gronwald, J.; Jakubowska, A.; Wokołorczyk, M.; Kurzawski, G.; et al. BRCA1 mutations and colorectal cancer in Poland. Fam. Cancer 2010, 9, 541–544. [Google Scholar] [CrossRef]
  68. Rothwell, P.M.; Wilson, M.; Elwin, C.-E.; Norrving, B.; Algra, A.; Warlow, C.P.; Meade, T.W. Long-term effect of aspirin on colorectal cancer incidence and mortality: 20-year follow-up of five randomised trials. Lancet 2010, 376, 1741–1750. [Google Scholar] [CrossRef]
  69. Burn, J.; Gerdes, A.-M.; Macrae, F.A.; Mecklin, J.-P.; Moeslein, G.; Olschwang, S.; Eccles, D.; Evans, D.G.; Maher, E.R.; Bertario, L.; et al. Long-term effect of aspirin on cancer risk in carriers of hereditary colorectal cancer: An analysis from the CAPP2 randomised controlled trial. Lancet 2011, 378, 2081–2087. [Google Scholar] [CrossRef] [Green Version]
  70. US Preventive Services Task Force. Routine aspirin or nonsteroidal anti-inflammatory drugs for the primary prevention of colorectal cancer: U.S. Preventive Services Task Force recommendation statement. Ann. Intern. Med. 2007, 146, 361–364. [Google Scholar] [CrossRef]
  71. Mersch, J.; Jackson, M.A.; Park, M.; Nebgen, D.; Peterson, S.K.; Singletary, C.; Arun, B.K.; Litton, J.K. Cancers associated withBRCA1andBRCA2mutations other than breast and ovarian. Cancer 2015, 121, 269–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Lin, K.M.; Ternent, C.A.; Adams, D.R.; Thorson, A.G.; Blatchford, G.J.; Christensen, M.A.; Watson, P.; Lynch, H.T. Colorectal cancer in hereditary breast cancer kindreds. Dis. Colon Rectum 1999, 42, 1041–1045. [Google Scholar] [CrossRef] [PubMed]
  73. Lin, K.M.; Ternent, C.A.; Adams, D.R.; Thorson, A.G.; Blatchford, G.J.; Christensen, M.A.; Watson, P.; Lynch, H.T. BRCA1 and BRCA2 founder mutations and the risk of colorectal cancer. J. Natl Cancer Inst. 2004, 96, 15–21. [Google Scholar]
  74. Slattery, M.L.; Kerber, R.A. Family History of Cancer and Colon Cancer Risk: The Utah Population Database. J. Natl. Cancer Inst. 1994, 86, 1618–1626. [Google Scholar] [CrossRef]
  75. Woodage, T.; King, S.M.; Wacholder, S.; Hartge, P.; Struewing, J.P.; McAdams, M.; Laken, S.J.; Tucker, M.A.; Brody, L.C. The APCI1307K allele and cancer risk in a community-based study of Ashkenazi Jews. Nat. Genet. 1998, 20, 62–65. [Google Scholar] [CrossRef]
  76. Redston, M.; Nathanson, K.L.; Yuan, Z.Q.; Neuhausen, S.L.; Satagopan, J.M.; Wong, N.; Yang, D.; Nafa, D.; Abrahamson, J.; Ozcelik, H.; et al. The APC I1307K allele and breast cancer risk. Nat. Genet. 1998, 20, 13–14. [Google Scholar] [CrossRef]
  77. Oh, M.; McBride, A.; Yun, S.; Bhattacharjee, S.; Slack, M.; Jeter, J.M.; Abraham, I. BRCA1 and BRCA2 gene mutations and colorectal cancer risk: Systematic review and meta-analysis. J. Clin. Oncol. 2018, 36, 605. [Google Scholar] [CrossRef]
  78. Lin, Z.; Liu, J.; Peng, L.; Zhang, D.; Jin, M.; Wang, J.; Xue, J.; Liu, H.; Zhang, T. Complete pathological response following neoadjuvant FOLFOX chemotherapy in BRCA2-mutant locally advanced rectal cancer: A case report. BMC Cancer 2018, 18, 1253. [Google Scholar] [CrossRef]
  79. Nolan, E.; Savas, P.; Policheni, A.N.; Darcy, P.K.; Vaillant, F.; Mintoff, C.P.; Dushyanthen, S.; Mansour, M.; Pang, J.-M.B.; Fox, S.; et al. Combined immune checkpoint blockade as a therapeutic strategy forBRCA1-mutated breast cancer. Sci. Transl. Med. 2017, 9, eaal4922. [Google Scholar] [CrossRef]
  80. Strickland, K.C.; Howitt, B.E.; Shukla, S.A.; Rodig, S.; Ritterhouse, L.L.; Liu, J.F.; Garber, J.E.; Chowdhury, D.; Wu, C.J.; D’Andrea, A.D.; et al. Association and prognostic significance of BRCA1/2-mutation status with neoantigen load, number of tumor-infiltrating lymphocytes and expression of PD-1/PD-L1 in high grade serous ovarian cancer. Oncotarget 2016, 7, 13587–13598. [Google Scholar] [CrossRef] [Green Version]
  81. Naseem, M.; Xiu, J.; Salem, M.E.; Goldberg, R.M.; VanderWalde, A.M.; Grothey, A.; Philip, P.A.; Seeber, A.; Puccini, A.; Tokunaga, R.; et al. Characteristics of colorectal cancer (CRC) patients with BRCA1 and BRCA2 mutations. J. Clin. Oncol. 2019, 37, 606. [Google Scholar] [CrossRef] [Green Version]
  82. Harpaz, N.; Gatt, Y.E.; Granit, R.Z.; Fruchtman, H.; Hubert, A.; Grinshpun, A. Mucinous Histology, BRCA1/2 Mutations, and Elevated Tumor Mutational Burden in Colorectal Cancer. J. Oncol. 2020, 2020, 1–10. [Google Scholar] [CrossRef] [Green Version]
  83. Davidson, D.; Wang, Y.; Aloyz, R.; Panasci, L. The PARP inhibitor ABT-888 synergizes irinotecan treatment of colon cancer cell lines. Investig. New Drugs 2012, 31, 461–468. [Google Scholar] [CrossRef] [PubMed]
  84. Shelton, J.W.; Waxweiler, T.V.; Landry, J.; Gao, H.; Xu, Y.; Wang, L.; El-Rayes, B.; Shu, H.-K.G. In Vitro and In Vivo Enhancement of Chemoradiation Using the Oral PARP Inhibitor ABT-888 in Colorectal Cancer Cells. Int. J. Radiat. Oncol. Biol. Phys. 2013, 86, 469–476. [Google Scholar] [CrossRef]
  85. Pishvaian, M.J.; Slack, R.S.; Jiang, W.; He, A.R.; Hwang, J.J.; Hankin, A.; Dorsch-Vogel, K.; Kukadiya, D.; Weiner, L.M.; Marshall, J.L.; et al. A phase 2 study of the PARP inhibitor veliparib plus temozolomide in patients with heavily pretreated metastatic colorectal cancer. Cancer 2018, 124, 2337–2346. [Google Scholar] [CrossRef]
  86. Wang, C.; Jette, N.; Moussienko, D.; Bebb, D.G.; Lees-Miller, S.P. ATM-Deficient Colorectal Cancer Cells Are Sensitive to the PARP Inhibitor Olaparib. Transl. Oncol. 2017, 10, 190–196. [Google Scholar] [CrossRef]
  87. Gaymes, T.J.; Mohamedali, A.M.; Patterson, M.; Matto, N.; Smith, A.; Kulasekararaj, A.; Chelliah, R.; Curtin, N.J.; Farzaneh, F.; Shall, S.; et al. Microsatellite instability induced mutations in DNA repair genes CtIP and MRE11 confer hypersensitivity to poly (ADP-ribose) polymerase inhibitors in myeloid malignancies. Haematologica 2013, 98, 1397–1406. [Google Scholar] [CrossRef]
  88. Leichman, L.; Groshen, S.; O’Neil, B.H.; Messersmith, W.; Berlin, J.; Chan, E.; Leichman, C.G.; Cohen, S.J.; Cohen, D.; Lenz, H.; et al. Phase II Study of Olaparib (AZD-2281) After Standard Systemic Therapies for Disseminated Colorectal Cancer. Oncology 2016, 21, 172–177. [Google Scholar] [CrossRef] [Green Version]
  89. Hansford, S.; Kaurah, P.; Li-Chang, H.; Woo, M.; Senz, J.; Pinheiro, H.; Schrader, K.A.; Schaeffer, D.F.; Shumansky, K.; Zogopoulos, G.; et al. Hereditary Diffuse Gastric Cancer Syndrome: CDH1 Mutations and Beyond. JAMA Oncol. 2015, 1, 23–32. [Google Scholar] [CrossRef] [Green Version]
  90. Tulinius, H.; Olafsdottir, G.H.; Sigvaldason, H.; Arason, A.; Barkardottir, R.B.; Egilsson, V.; Ogmundsdottir, H.M.; Tryggvadottir, L.; Gudlaugsdottir, S.; Eyfjord, J.E. The effect of a single BRCA2 mutation on cancer in Iceland. J. Med. Genet. 2002, 39, 457–462. [Google Scholar] [CrossRef] [Green Version]
  91. Van Asperen, C.J.; Brohet, R.M.; Meijers-Heijboer, E.J.; Hoogerbrugge, N.; Verhoef, S.; Vasen, H.F.A.; Ausems, M.G.E.M.; Menko, F.H.; Garcia, E.B.G.; Klijn, J.G.M.; et al. Cancer risks in BRCA2 families: Estimates for sites other than breast and ovary. J. Med. Genet. 2005, 42, 711–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Bermejo, J.L.; García-Pérez, A.; Hemminki, K. Contribution of the Defective BRCA1, BRCA2 and CHEK2 Genes to the Familial Aggregation of Breast Cancer: A Simulation Study Based on the Swedish Family-Cancer Database. Hered. Cancer Clin. Pr. 2004, 2, 185–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Bang, Y.-J.; Im, S.-A.; Lee, K.-W.; Cho, J.Y.; Song, E.-K.; Lee, K.H.; Kim, Y.H.; Park, J.O.; Chun, H.G.; Zang, D.Y.; et al. Randomized, Double-Blind Phase II Trial With Prospective Classification by ATM Protein Level to Evaluate the Efficacy and Tolerability of Olaparib Plus Paclitaxel in Patients With Recurrent or Metastatic Gastric Cancer. J. Clin. Oncol. 2015, 33, 3858–3865. [Google Scholar] [CrossRef] [PubMed]
  94. Bang, Y.-J.; Xu, R.-H.; Chin, K.; Lee, K.-W.; Park, S.H.; Rha, S.Y.; Shen, L.; Qin, S.; Xu, N.; Im, S.-A.; et al. Olaparib in combination with paclitaxel in patients with advanced gastric cancer who have progressed following first-line therapy (GOLD): A double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 2017, 18, 1637–1651. [Google Scholar] [CrossRef]
  95. ClinicalTrials.gov Identifier NCT02734004. A Phase I/II Study of MEDI4736 in Combination with Olaparib in Patients with Advanced Solid Tumors. (MEDIOLA). Available online: https://clinicaltrials.gov/ct2/show/NCT02734004 (accessed on 2 September 2020).
  96. ClinicalTrials.gov Identifier NCT03008278. Olaparib and Ramucirumab in Treating Patients with Metastatic or Locally Recurrent Gastric or Gastroesophageal Junction Cancer That Cannot be Removed by Surgery. Available online: https://clinicaltrials.gov/ct2/show/NCT03008278 (accessed on 2 September 2020).
  97. Nakamura, H.; Arai, Y.; Totoki, Y.; Shirota, T.; ElZawahry, A.; Kato, M.; Hama, N.; Hosoda, F.; Urushidate, T.; Ohashi, S.; et al. Genomic spectra of biliary tract cancer. Nat. Genet. 2015, 47, 1003–1010. [Google Scholar] [CrossRef]
  98. Churi, C.R.; Shroff, R.; Wang, Y.; Rashid, A.; Kang, H.C.; Weatherly, J.; Zuo, M.; Zinner, R.; Hong, D.; Meric-Bernstam, F.; et al. Mutation Profiling in Cholangiocarcinoma: Prognostic and Therapeutic Implications. PLoS ONE 2014, 9, e115383. [Google Scholar] [CrossRef] [Green Version]
  99. Golan, T.; Raitses-Gurevich, M.; Kelley, R.K.; Bocobo, A.G.; Borgida, A.; Shroff, R.T.; Holter, S.; Gallinger, S.; Ahn, D.H.; Aderka, D.; et al. Overall Survival and Clinical Characteristics of BRCA-Associated Cholangiocarcinoma: A Multicenter Retrospective Study. Oncology 2017, 22, 804–810. [Google Scholar] [CrossRef] [Green Version]
  100. Cheng, Y.; Zhang, J.; Qin, S.K.; Hua, H.Q. Treatment with olaparib monotherapy for BRCA2-mutated refractory intrahepatic cholangiocarcinoma: A case report. Onco Targets Ther. 2018, 11, 5957–5962. [Google Scholar] [CrossRef]
  101. Spizzo, G.; Puccini, A.; Xiu, J.; Goldberg, R.M.; Grothey, A.; Shields, A.F.; Arora, S.P.; Khushman, M.M.; Salem, M.E.; Battaglin, F.; et al. Frequency of BRCA mutation in biliary tract cancer and its correlation with tumor mutational burden (TMB) and microsatellite instability (MSI). J. Clin. Oncol. 2019, 37, 4085. [Google Scholar] [CrossRef]
  102. Fehling, S.C.; Miller, A.L.; Garcia, P.L.; Vance, R.B.; Yoon, K.J. The combination of BET and PARP inhibitors is synergistic in models of cholangiocarcinoma. Cancer Lett. 2020, 468, 48–58. [Google Scholar] [CrossRef]
  103. Mao, X.; Du, S.; Yang, Z.; Zhang, L.; Peng, X.; Jiang, N.; Zhou, H. Inhibitors of PARP-1 exert inhibitory effects on the biological characteristics of hepatocellular carcinoma cells in vitro. Mol. Med. Rep. 2017, 16, 208–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. ClinicalTrials.gov Identifier NCT04042831. Olaparib in Treating Patients with Metastatic Biliary Tract Cancer with Aberrant DNA Repair Gene Mutations. Available online: https://clinicaltrials.gov/ct2/show/NCT04042831 (accessed on 2 September 2020).
  105. Lin, J.; Shi, J.; Guo, H.; Yang, X.; Jiang, Y.; Long, J.; Bai, Y.; Wang, D.; Yang, X.; Wan, X.; et al. Alterations in DNA Damage Repair Genes in Primary Liver Cancer. Clin. Cancer Res. 2019, 25, 4701–4711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hänninen, U.A.; Katainen, R.; Tanskanen, T.; Plaketti, R.-M.; Laine, R.; Hamberg, J.; Ristimäki, A.; Pukkala, E.; Taipale, M.; Mecklin, J.-P.; et al. Exome-wide somatic mutation characterization of small bowel adenocarcinoma. PLoS Genet. 2018, 14, e1007200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Quaas, A.; Heydt, C.; Waldschmidt, D.; Alakus, H.; Zander, T.; Goeser, T.; Kasper, P.; Bruns, C.J.; Brunn, A.; Roth, W.; et al. Alterations in ERBB2 and BRCA and microsatellite instability as new personalized treatment options in small bowel carcinoma. BMC Gastroenterol. 2019, 19, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Beger, C.; Ramadani, M.; Meyer, S.; Leder, G.; Krüger, M.; Welte, K.; Gansauge, F.; Beger, H.G. Down-Regulation of BRCA1 in Chronic Pancreatitis and Sporadic Pancreatic Adenocarcinoma. Clin. Cancer Res. 2004, 10, 3780–3787. [Google Scholar] [CrossRef] [Green Version]
  109. Golan, T.D.; Kanji, Z.S.; Epelbaum, R.; Devaud, N.; Dagan, E.; Holter, S.; Aderka, D.; Paluchshimon, S.; Kaufman, B.A.; Gershonibaruch, 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]
  110. 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]
  111. Xu, R.-H.; Yu, X.; Hao, J.; Wang, L.; Pan, H.; Han, G.; Xu, J.; Zhang, Y.; Yang, S.; Chen, J.; et al. Efficacy and safety of weekly nab-paclitaxel plus gemcitabine in Chinese patients with metastatic adenocarcinoma of the pancreas: A phase II study. BMC Cancer 2017, 17, 885. [Google Scholar] [CrossRef]
  112. Kobayashi, N.; Shimamura, T.; Tokuhisa, M.; Goto, A.; Endo, I.; Ichikawa, Y. Effect of FOLFIRINOX as second-line chemotherapy for metastatic pancreatic cancer after gemcitabine-based chemotherapy failure. Medicine 2017, 96, e6769. [Google Scholar] [CrossRef]
  113. Soyano, A.E.; Baldeo, C.; Kasi, P.M. BRCA Mutation and Its Association with Colorectal Cancer. Clin. Color. Cancer 2018, 17, e647–e650. [Google Scholar] [CrossRef] [Green Version]
  114. Cercek, A.; Roxburgh, C.S.; Strombom, P.; Smith, J.J.; Temple, L.K.; Nash, G.M.; Guillem, J.G.; Paty, P.B.; Yaeger, R.; Stadler, Z.K.; et al. Adoption of Total Neoadjuvant Therapy for Locally Advanced Rectal Cancer. JAMA Oncol. 2018, 4, e180071. [Google Scholar] [CrossRef]
  115. Menear, K.A.; Adcock, C.; Boulter, R.; Cockcroft, X.-L.; Copsey, L.; Cranston, A.; Dillon, K.J.; Drzewiecki, J.; Garman, S.; Gomez, S.; et al. 4-[3-(4-Cyclopropanecarbonylpiperazine-1-carbonyl)-4-fluorobenzyl]-2H-phthalazin-1-one: A Novel Bioavailable Inhibitor of Poly(ADP-ribose) Polymerase-1. J. Med. Chem. 2008, 51, 6581–6591. [Google Scholar] [CrossRef] [PubMed]
  116. Yi, M.; Dong, B.; Qin, S.; Chu, Q.; Wu, K.; Luo, S. Advances and perspectives of PARP inhibitors. Exp. Hematol. Oncol. 2019, 8, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Vinayak, S.; Tolaney, S.M.; Schwartzberg, L.; Mita, M.; McCann, G.; Tan, A.R.; Wahner-Hendrickson, A.E.; Forero, A.; Anders, C.; Wulf, G.M.; et al. Open-label Clinical Trial of Niraparib Combined With Pembrolizumab for Treatment of Advanced or Metastatic Triple-Negative Breast Cancer. JAMA Oncol. 2019, 5, 1132–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Konstantinopoulos, P.A.; Waggoner, S.; Vidal, G.A.; Mita, M.; Moroney, J.W.; Holloway, R.; Van Le, L.; Sachdev, J.C.; Chapman-Davis, E.; Colon-Otero, G.; et al. Single-Arm Phases 1 and 2 Trial of Niraparib in Combination With Pembrolizumab in Patients With Recurrent Platinum-Resistant Ovarian Carcinoma. JAMA Oncol. 2019, 5, 1141–1149. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Types of DNA damage and repair mechanisms with related repair enzymes. A single-strand break (SSB) is accomplished by base excision repair (BER) through PARP enzymes. A double-strand break (DSB) is accomplished either with non-homologous end joining (NHEJ) through, mainly, DNA-dependent protein kinase (DNA-PKcs) or with homologous recombination (HR) through several enzymes (BRCA1, BRCA2, ATM, PALB2, BRIP1, RAD51, CHEK2).
Figure 1. Types of DNA damage and repair mechanisms with related repair enzymes. A single-strand break (SSB) is accomplished by base excision repair (BER) through PARP enzymes. A double-strand break (DSB) is accomplished either with non-homologous end joining (NHEJ) through, mainly, DNA-dependent protein kinase (DNA-PKcs) or with homologous recombination (HR) through several enzymes (BRCA1, BRCA2, ATM, PALB2, BRIP1, RAD51, CHEK2).
Cancers 12 03346 g001
Figure 2. Effect of mutation in genes encoding for homologous recombination (HR) enzymes. A pathogenic mutation in BRCA1, BRCA2, ATM, PALB2, BRIP1, RAD51 and CHEK2 causes an impaired HR. The cell is, therefore, deficient in HR and it uses NHEJ preferentially to repair the DSB. However, NHEJ, unlike HR, enhances genomic instability until carcinogenesis.
Figure 2. Effect of mutation in genes encoding for homologous recombination (HR) enzymes. A pathogenic mutation in BRCA1, BRCA2, ATM, PALB2, BRIP1, RAD51 and CHEK2 causes an impaired HR. The cell is, therefore, deficient in HR and it uses NHEJ preferentially to repair the DSB. However, NHEJ, unlike HR, enhances genomic instability until carcinogenesis.
Cancers 12 03346 g002
Figure 3. Mechanism of action of PARP inhibitors (PARPi). PARP inhibitors block PARP enzymes’ action, thus inducing an SSB to become a DSB. In the case of HR deficiency, a DSB cannot be repaired and therefore NHEJ is aberrantly activated, thus leading to accumulation of gene alterations until cancer cell death.
Figure 3. Mechanism of action of PARP inhibitors (PARPi). PARP inhibitors block PARP enzymes’ action, thus inducing an SSB to become a DSB. In the case of HR deficiency, a DSB cannot be repaired and therefore NHEJ is aberrantly activated, thus leading to accumulation of gene alterations until cancer cell death.
Cancers 12 03346 g003
Table 1. Surveillance criteria according to the International Cancer of the Pancreas Screening (CAPS) Consortium consensus [56].
Table 1. Surveillance criteria according to the International Cancer of the Pancreas Screening (CAPS) Consortium consensus [56].
CAPS Consortium Consensus for Surveillance of HRIs.
Individuals with at least three or more affected relatives, of whom at least one is an FDR to the individual considered for screening
Individuals with at least two affected relatives who are FDRs to each other, of whom at least one is an FDR to the individual considered for surveillance
Individuals with at least two affected relatives on the same side of the family, of whom at least one is an FDR to the individual considered for surveillance
LKB1/STK11 mutation carriers (PJS) regardless of family history
CDKN2A mutation carriers regardless of family history
BRCA1 mutation carriers with one affected FDR
BRCA2 mutation carriers with one affected FDR (or two affected family members, no FDR) with PC
PALB2 mutation carriers with one affected FDR
MMR gene mutation carriers (LS) with one affected FDR
ATM mutation carriers with one affected FDR
Table 2. Summary of ongoing clinical trials in BRCA-mutated gastrointestinal tumors.
Table 2. Summary of ongoing clinical trials in BRCA-mutated gastrointestinal tumors.
NCI Trial NumberInterventionCancerPrimary EndpointPhase
NCT03337087Nal-IRI+ Fluorouracil + RucaparibPancreatic, colorectal, gastroesophageal or biliary cancerToxicity, ORR, best response rateI, II
NCT03838406FOLFOX/CAPOXGastric cancerORRNot Applicable
NCT03565991Talazoparib + AvelumabAdvanced solid tumorsORII
NCT02286687TalazoparibRecurrent, refractory, advanced or metastatic cancersClinical benefitII
NCT01989546BMN 673 (Talazoparib)Advanced solid tumorsPharmacodynamic effect of talazoparib; response rateI, II
NCT03140670RucaparibPancreatic cancerSafetyII
NCT03428802PembrolizumabAdvanced solid tumorsResponse rateII
NCT02723864Veliparib + VX-970 + CisplatinRefractory solid tumorsSafety, MTD, tolerabilityI
NCT04182516NMS-03305293Advanced/metastatic solid tumorsToxicityI
NCT03875313CB-839 + TalazoparibSolid tumorsSafety, MTD, tolerabilityI, II
ORR: objective response rate; OR: objective response; MTD: maximum tolerated dose.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Molinaro, E.; Andrikou, K.; Casadei-Gardini, A.; Rovesti, G. BRCA in Gastrointestinal Cancers: Current Treatments and Future Perspectives. Cancers 2020, 12, 3346. https://doi.org/10.3390/cancers12113346

AMA Style

Molinaro E, Andrikou K, Casadei-Gardini A, Rovesti G. BRCA in Gastrointestinal Cancers: Current Treatments and Future Perspectives. Cancers. 2020; 12(11):3346. https://doi.org/10.3390/cancers12113346

Chicago/Turabian Style

Molinaro, Eleonora, Kalliopi Andrikou, Andrea Casadei-Gardini, and Giulia Rovesti. 2020. "BRCA in Gastrointestinal Cancers: Current Treatments and Future Perspectives" Cancers 12, no. 11: 3346. https://doi.org/10.3390/cancers12113346

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