Next Article in Journal
Quality of Life Assessment in Students from Polish Universities during the COVID-19 Pandemic According to WHO Quality of Life Questionnaire
Previous Article in Journal
Classic Psychedelics and Human–Animal Relations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epithelial Ovarian Cancer: Providing Evidence of Predisposition Genes

1
Department of Palliative Care, Guy’s and St Thomas’ Hospital, London SE1 9RT, UK
2
Department of Hematology/Medical Oncology, Medway NHS Foundation Trust, Windmill Road, Kent, Gillingham ME7 5NY, UK
3
Department of Urology, Medway NHS Foundation Trust, Windmill Road, Kent, Gillingham ME7 5NY, UK
4
King’s College London, Faculty of Life Sciences & Medicine, School of Cancer & Pharmaceutical Sciences, London SE1 9RT, UK
5
AELIA Organization, 9th Km Thessaloniki-Thermi, 57001 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(13), 8113; https://doi.org/10.3390/ijerph19138113
Submission received: 23 April 2022 / Revised: 28 June 2022 / Accepted: 30 June 2022 / Published: 1 July 2022

Abstract

:
Epithelial ovarian cancer (EOC) is one of the cancers most influenced by hereditary factors. A fourth to a fifth of unselected EOC patients carry pathogenic variants (PVs) in a number of genes, the majority of which encode for proteins involved in DNA mismatch repair (MMR) pathways. PVs in BRCA1 and BRCA2 genes are responsible for a substantial fraction of hereditary EOC. In addition, PV genes involved in the MMR pathway account for 10–15% of hereditary EOC. The identification of women with homologous recombination (HR)-deficient EOCs has significant clinical implications, concerning chemotherapy regimen planning and development as well as the use of targeted therapies such as poly(ADP-ribose) polymerase (PARP) inhibitors. With several genes involved, the complexity of genetic testing increases. In this context, next-generation sequencing (NGS) allows testing for multiple genes simultaneously with a rapid turnaround time. In this review, we discuss the EOC risk assessment in the era of NGS.

1. Introduction

Ovarian cancer is the fifth leading cause of cancer-related mortality in females [1]. Over two-thirds of women present with advanced stages of ovarian cancer, which leads to an estimated 5-year survival rate of between 20–40% [2]. In contrast, those diagnosed at an earlier stage (e.g., stage 1 disease) have a 5-year survival rate of over 90% [3]. The GLOBOCAN study predicted a worldwide increase of 55% in ovarian cancer cases and a 67% increase in deaths from the year 2012 to 2035 [4]. The median age of women at diagnosis is 63 years [5]. Epithelial ovarian cancer (EOC) is a heterogenous disease further classified into benign, borderline, and malignant [6,7]. Malignant EOC includes five main histological subtypes: high-grade serous ovarian cancer (70–80%), endometrioid (10%), clear cell (10%), mucinous (3%), and low-grade serous (<5%) [6,7]. Each subtype behaves as a distinct disease with differences in clinical presentations, mutations, and responses to treatment such as chemotherapy [8]. It is well-established that EOC develops according to two different carcinogenic pathways. The vast majority of these tumours are high-grade serous tumours that develop according to the type II pathway and present p53 and BRCA mutations. In contrast, low-grade serous tumours are characterised by KRAS, BRAF, PTEN, PIK3CA, CTNNB1, ARID1A, and PPP2R1A mutations and progress according to the type I pathway [9]. Beyond EOC, there are also non-epithelial counterparts that are further divided into germ cell (5%) and sex-cord stromal cell (5%) ovarian cancers [10,11,12]. Ovarian carcinosarcoma is even rarer; it is a biphasic, but challenging histologic subtype, accounting for only 1–4% of all ovarian cancers [13].
Newly diagnosed high-grade serous EOC patients are treated with radical surgery followed by adjuvant platinum and taxane combination chemotherapy. In EOC patients where upfront surgery is contraindicated for medical reasons, or where complete cytoreduction cannot be achieved, neoadjuvant chemotherapy prior to interval debulking surgery followed by adjuvant chemotherapy is an alternative therapeutic option [14]. The treatment of gestational ovarian malignancies depends on the histology, stage, and gestational weeks. When possible, surgical excision is indicated; fertility-sparing surgery can be offered to stage I EOC, germ cell ovarian, or sex-cord stromal ovarian tumours [15].
Approximately 20–30% of EOC occurs in females with an inherited predisposition; most of these hereditary ovarian cancers are due to germline mutations in BRCA1 and BRCA2 genes [16,17,18,19]. Lynch/hereditary nonpolyposis colorectal cancer (HNPCC) syndrome is also associated with an increased risk of ovarian cancer and is caused by germline mutations in the mismatch repair (MMR) genes [20,21]. The BRCA1 and BRCA2 genes are involved in the repair of double-strand DNA breaks (DSBs) through the homologous recombination (HR) system [22]. Pathogenic variants (PVs) in these genes cause a dysfunction in the BRCA proteins and the HR system, and increase the risk of ovarian cancer [22,23]. Carriers of germline mutations in BRCA1 and BRCA2 genes have a lifetime risk of ovarian cancer of 35–60% and 12–25%, respectively, along with an increased risk of peritoneal and fallopian tube malignancies [24,25,26]. Therefore, the identification of these germline BRCA1/2 mutations has implications for the therapeutic management of EOC [16]. BRCA1/2 carriers respond well to platinum-based chemotherapy and poly(ADP-ribose) polymerase (PARP) inhibitors [27,28,29]. New genes (RAD51C, RAD51D, and PALB2) have been identified as increasing susceptibility to ovarian cancer, but further research is required to investigate this [30,31,32,33]. In this review, we discuss susceptibility to EOC, testing, and EOC risk assessments in the context of next-generation sequencing (NGS).

2. Molecular Landscape

DNA damage occurs daily in cells and there are complex repair pathways to prevent the genomic instability that can be predisposed to cancer [34,35,36]. HR is a repair pathway for complex DNA damage, including DSBs and DNA gaps [37]. BRCA1 and BRCA2 genes are integral components of the HR repair pathway, acting as tumour suppressors and preserving the chromosome structure [23,38]. As part of the HR pathway, the BRCA2 protein recruits the recombinase RAD51 to the DSBs by forming a BRCA1–PALB2–BRCA2 complex [39]. Germline mutations in both BRCA1 and BRCA2 can, therefore, lead to a deficiency in the HR pathway and increase susceptibility to ovarian cancer [38]. Identifying these mutations is significant as those with BRCA1/2 mutations are sensitive to a therapy with PARP inhibitors [40,41,42,43]. Combinations of PARP inhibitors with drugs that inhibit HR may sensitise EOC with a primary or secondary HR proficiency to PARP inhibitors and potentially expand their use beyond HR-deficient ovarian cancers. Regarding this, PARP inhibitors may be combined separately with anti-angiogenics and immune checkpoint inhibitors as well as with phosphoinositide 3-kinase (PI3K), protein kinase B (AKT), mammalian target of rapamycin (mTOR), WEE1, mitogen-activated protein kinase (MEK), and cyclin dependent kinase (CDK) 4/6 inhibitors, or even with standard chemotherapy [44].
If the HR repair pathway is impaired, synthetic lethality may be induced by a PARP inhibition exclusively for the target tumour tissue and spare normal cells [2,29,45]. The PARPs are a group of 18 multifunctional enzymes, the most active of which are PARP1 and PARP2; these repair DNA single-strand breaks through a base excision repair [29,46,47]. PARP inhibition leads to accumulating single-strand DNA breaks (SSBs) and subsequently leads to DSBs at DNA replication forks [29]. These breaks are usually repaired by a functioning HR pathway [48]. Cells carrying the heterozygous BRCA1/2 mutation result in tumours that carry a DNA mismatch repair (MMR) deficiency unlike normal cells [29]. Therefore, PARP inhibition can selectively target the tumour tissue and lead to unrepaired SSBs, which then leads to an accumulation of DSBs and collapsed replication forks as well as chromosomal instability and consequent tumour cell apoptosis [29,49,50,51]. This synthetic lethality—a phenomenon where multiple, concurrent genetic events lead to cell apoptosis rather than a single event—allows the increased sensitivity of BRCA-mutated tumours to PARP inhibitors and is a novel approach that has changed the treatment of ovarian cancer [52,53]. Figure 1 provides a schematic of the roles of BRCA1 and BRCA2 in the DNA repair mechanism.
Apart from breast and ovarian cancer, several PARP inhibitors have been investigated in metastatic castration-resistant prostate cancer patients [54]. They are particularly effective in BRCA1/2 mutations with increased survival outcomes. Olaparib is used in this subset of patients after a progression on novel hormonal agents, e.g., enzalutamide or abiraterone, whilst rucaparib is considered in combination with androgen receptor-guided therapy and paclitaxel-based chemotherapy [55]. Moreover, PARP inhibitor monotherapy induces an objective anti-tumour activity in patients with PALB2, BRIP1, or FANCA aberrations. In contrast, those with ATM and CDK12 alterations do not seem to benefit [56].
MMR genes are another class of genes that are linked to a susceptibility to ovarian cancer [3]. The MMR system involves seven genes: MLH1, MSH2, MSH6, PMS2, PMS1, MSH3 and MLH3; these act sequentially to identify MMR mutations and then form a protein complex to correct them [57]. MSH2 and either MSH3 or MSH6 form a complex to identify the mismatch; this then binds to a complex of MLH1 with either PMS2, PMS3, or MLH3 to repair the defect [57]. An impaired MMR function through MMR mutations leads to microsatellite instability (MSI); this is characteristic of HNPCC, the third most common hereditary ovarian cancer [57]. This autosomal dominant condition is most often associated with germline mutations in the MSH6, MSH2, and MLH1 genes and increases the risk of the endometrioid and clear cell subtypes of ovarian cancer [6,57,58,59]. The cumulative lifetime risk of ovarian cancer in women with HNPCC is estimated at 4 to 12% and the prognosis is linked to the MMR variants [6,57]. Furthermore, the mean age of diagnosis with these germline MMR mutations is 9 to 13 years earlier than the general population with sporadic tumours [60]. Increasing research to improve the knowledge of the genetic mechanisms associated with hereditary ovarian cancer is, therefore, paramount for the future of its management.
There are five major histologies of EOC: high-grade serous, low-grade serous, mucinous, endometrioid, and clear cell [61]. Mutations in the DNA repair genes differ across different histological subtypes. For example, mutations in BRCA1/2 genes are often found in high-grade serous, but rarely in mucinous, endometrioid, and clear cell EOC [62,63,64]. The Cancer Genome Atlas project identified that 16% of high-grade serous EOC patients had germline BRCA1/2 mutations [65]. Sugino et al., investigated these differences further in high-grade serous, endometrioid, and clear cell EOC; a few of the common mutations are summarised in Table 1 [62]. Furthermore, MMR mutations were also investigated in these histologies. MMR mutations were found in 10% of endometrioid, 3% in clear cell, and 2% of high-grade serous EOC [62]. Lheureux et al., also found that MMR mutations accounted for 10–15% of hereditary ovarian cancer, especially in endometrioid and clear cell histological subtypes [66]. Norquist et al., found that mutations were similar in high-grade serous, endometrioid, and clear cell EOC, but there was an overall lower mutation frequency in low-grade compared with high-grade serous EOC (5.7% vs. 19.6%, respectively; p = 0.003) and there were no ovarian cancer-associated mutations in the mucinous histology [18].

3. Susceptibility to EOC

The parts of the gene responsible for an increased susceptibility or predisposition to cancer in individuals are called PVs [67]. In the case of ovarian cancer, at least one of these variations in the actionable genes was found in 49.2% of patients [67]. Moreover, the prevalence of PVs is said to increase with the number of primary cancers (PCs): 13.1% with 2 PCs, 15.9% with 3 PCs, and 18.0% with ≥4 PCs [68]. This shows that EOC has a higher incidence compared with the general subset of malignant malformations. There are plenty of well-established PVs in ovarian cancer that have been shown to have a significant correlation with EOC such as BRCA1 and BRCA2 [69]. There are studies that show promising results for other PVs that are connected to EOC and have shown a potential for diagnostic and treatment use. Several of them include MMR genes; e.g., MSH6, MSH2, and MLH1 [70,71].

3.1. Germline Predisposing Variants

3.1.1. BRCA1 and BRCA2 Genes

The highly penetrant genes BRCA1 and BRCA2 are associated with most of the identified germline mutations in EOC patients. These genes produce proteins involved in fundamental cellular processes such as cell cycle checkpoint control, chromatin remodelling, transcriptional regulation, and mitosis [72]. Therefore, an HR deficiency caused by these mutations is often utilised in the treatment of EOC with platinum-based chemotherapy or PARP inhibitors [73]. Apart from EOC, these germline variants have been shown to be associated with 22 first PCs [69]. However, these germline mutations are the strongest-known genetic risk factors for EOC and are found in 6–15% of women with EOC [74]. The BRCA1/2 status can be used by healthcare professionals for patient counselling regarding expected survival as BRCA1 and BRCA2 carriers with EOC respond better than non-carriers to platinum-based chemotherapies. This yields greater survival, even though the disease is generally diagnosed at a later stage and higher grade [75]. Figure 2 provides the distribution of germline PVs identified in unselected EOC patients. A fourth to a fifth of these patients carried PVs in a number of genes, the majority of which encode for proteins involved in the DNA repair pathways.

3.1.2. RAD51C, RAD51D, and BRIP1 Genes

Collectively, germline mutations in BRIP1, RAD51C, and RAD51D account for around 2% of ovarian cancer cases [76]. Despite significant evidence of a strong correlation with EOC, the risk attributed to particular genes varies substantially among studies (odds ratio (OR) values estimated for BRIP1 ~ 5–19, RAD51C ~ 5–15, and RAD51D ~ 6–12) [77,78,79]. BRCA1-interacting protein C-terminal helicase 1 (BRIP1), RAD51 homolog C (RAD51C), and RAD51 paralog D (RAD51D) are all coding for proteins that interact with BRCA1/2 and support the MMR process. Hence, patients with germline mutations in BRIP1, RAD51C, and RAD51D would also benefit from a therapy with PARP inhibitors.
RAD51C and RAD51D are essential for HR repair [30,80]. Mutations in these genes have a higher likelihood of high-grade serous EOC in women aged between 40 and 49 years old [60]. These genes have a few similarities with the BRCA1/2 genes, but there is little evidence to suggest they also increase the risk of breast cancer as with BRCA1/2 [60]. Furthermore, the location of their mutation is not variable as with BRCA1/2, which determines the risk of ovarian cancer [81]. Mutations in RAD51C occur between amino acid 143 and 319, which affect the RAD51B–RAD51C–RAD51D–XRCC2 and RAD51C–XRCC3 complexes; mutations in RAD51D occur in the C-terminal region, which affect binding to RAD51C and DSB repair [82].
Mutations in the BRIP1 gene have been found in the first two-thirds of the gene between nucleotides 68 and 2508; these cut the protein before the BRCA1 binding domain [77]. This domain is found in proteins acting as checkpoints for DNA damage; losing this interaction impairs the DNA damage repair in the epithelial cells in the ovaries or fallopian tubes [83,84,85]. Furthermore, an increased risk of BRIP1 missense variants has been associated with high-grade serous EOC [86]. Several countries have implemented recommendations such as those for BRCA1/2 carriers where BRIP1, RAD51C, and RAD51D mutation carriers are given an option of a salpingo-oophorectomy beginning at age 45–50 as a prophylactic measure [77].

3.1.3. PALB2 Gene

Although there are studies that identify PLAB2 as one of the PVs in EOC, a meta-analysis showed a non-significant increased risk of ovarian cancer (OR = 4.55; 95% confidence interval (CI), 0.76–27.24; p = 0.10) [71]. Moreover, it showed that germline PALB2 mutations were rare; PALB2 mutation carriers were found in only 0.21% of the ovarian cancer patients analysed, which is not significantly different from the frequency in the general population (0.05%) [84]. More research needs to be undertaken to further investigate the connection.

3.1.4. MLH1, MSH2, MSH6, and PMS2 Genes

DNA MMR consists of three stages: initiation, excision, and resynthesis. MutS homologs (MSH2, MSH3, and MSH6) are a few of the proteins involved in the initiation stage of MMR [87]. Lynch syndrome, also called HNPCC, is caused by mutations in these variants (MLH1, MSH2, MSH6, and PMS2). EOC patients with Lynch syndrome are often younger with a median age of diagnosis of 43 years; however, the prevalence of the disease in this syndrome is small at 0.9–2.7% [88,89]. Pal et al., reported that clear pathogenic germline MLH1, MSH2, and MSH6 mutations occurred in only 9/1893 (0.5%) of unselected EOC patients [90]. A small number of studies have investigated the survival of women with EOC due to MMR defects and the results are inconclusive [91]. This contrasts with the research undertaken in colorectal cancer, where MMR mutations are wildly established.

4. Tumour Testing in EOC

4.1. NGS for Cancer Risk Assessments

NGS technology has rapidly evolved over the last few decades. This revolutionary sequencing technique tests large genomic regions with a single test in a short period of time without needing a prior knowledge of the gene sequence [92]. NGS has almost replaced Sanger’s sequencing, which can only sequence a particular genomic region of interest due to the large workload and high costs [67]. Nowadays, multigene panels that analyse selected genes of interest are the most frequently used sequencing method in clinical applications [93]. Commonly, the included genes for testing are highly penetrant; however, they can also include moderate- and low-penetrance PV genes, depending on the actionability and significance of the pathological variant [94,95]. It provides valuable clinical information such as susceptibility to cancers and hereditary cancer syndromes in a cost-effective manner.
It also limits the risk of gathering large amounts of variants of unknown significance (VUS) that do not have enough data to change the clinical management. Gene panels with NGS also have a high accuracy rate in detecting all classes of mutations and genetic alterations in patients with suspected cancers [33,96]. This has significant clinical implications for patients with known hereditary gynaecological cancers who have a negative single-gene test as gene panels can often identify pathogenic mutations outside the high-penetrant genes [97]. Multigene panel testing has a much faster turnaround time and is much more cost-effective compared with single-gene testing [98,99]. It also reduces the chance of missing a mosaicism of pathogenic gene mutations in patients with hereditary cancer as NGS can identify even low levels of mutations [16].

4.2. Upfront Tumour Testing

Currently, tumour sample testing for ovarian cancer involves testing for BRCA1/2 mutations and HR deficiency status, which predicts the response to platinum agents and PARP inhibitors [38]. If a mutation is detected, sequencing is performed on normal cells to determine if the mutation is somatic or germline. Considering the above, performing tumour sequencing upfront with NGS can possibly avoid additional testing as it provides information on both the HR deficiency status and the gene mutation profile (germline or somatic). It also rules out mutations in susceptibility genes [100]. Recent studies have demonstrated that tumour sequencing has a high accuracy [101,102,103]. A tumour-first testing approach can facilitate early treatment decision-making for the use of PARP inhibitors and early cancer prevention, given that more than 20% of EOC patients have a hereditary link [33].

4.3. T Cell Repertoire Sequencing

There is an immune response to tumours that is facilitated by tumour-infiltrating T lymphocytes (TILs) [104]. These TILs are regulatory T cells that respond to shared tumour antigens [105]. Deep sequencing of T cell receptor genes has been used to identify TILs in ovarian cancer to compare the T cells in the blood. TIL repertoires within ovarian tumours are discrete from the T cell repertoire circulating in the peripheral blood. This makes TILs a prognostic marker and their presence (e.g., CD3+ and CD8+) has been shown to improve overall survival [104]. CD8+ TILs have particularly shown an increased survival benefit than CD3+ TILs. Hwang et al., suggest that CD8+ staining should be used to measure TIL proportions in ovarian cancer [106]. Zhang et al., found that ovarian tumours containing these T cells had a progression-free survival of 22.4 months and an overall survival of 50.3 months compared with 5.8 and 18 months, respectively, for tumours with no TILs [107]. Better outcomes with TILs indicate that they have an anti-tumour effect and reduce growth through the release of specific cytokines, e.g., interferon-gamma, which promotes inflammation and tumour elimination [107,108]. The Elispot (enzyme-linked immunospot) assay can detect antigen-specific T cells and this can detect TILs in the peripheral blood of patients [107]. This opens possibilities for testing using T cell repertoires from TILs. It also has implications for treatments with immunotherapy. There have been mixed results with the use of TILs in ovarian cancer. In advanced or recurrent ovarian cancer, Aoki et al., found clinical responses in five out of seven patients receiving TILs alone and in nine out of ten patients receiving chemotherapy only [109]. In other studies, there were no significant clinical responses and using TILs did not enhance the immune response [110,111]. Further studies need to be completed to conclude whether TIL therapy can be implemented alongside chemotherapy with the previously observed clinical significance practically translated into better survival outcomes.

4.4. Pre- and Post-Test Counselling

Genetic counselling is known to help patients decrease their anxiety and depression about the cancer diagnosis and the frequency of testing [112]. In the era of NGS, genetic counselling models must evolve alongside the rapidly changing genetic testing technologies and ovarian cancer management strategies. Despite current guidelines recommending that all women with a diagnosis of EOC should undergo genetic testing at the time of the diagnosis, only approximately 10–30% of these women are referred for germline genetic testing [113,114,115,116]. To ensure genetic counselling is delivered in a timely manner to avoid missed therapeutic and prevention opportunities, the authors of a Canadian review article on genetic assessments for BRCA-associated malignancies recommended a tumour-first testing model to detect both somatic and germline BRCA1/2 cases and to ensure that this is available to all patients without depending on a referral system [117].
With a multigene panel, the rates of VUS increase along with the detection rate [38]. A large retrospective study assessed individuals who underwent genetic testing over a decade and found that 91% of VUS were reclassified as benign or likely benign and their penetrance and cancer risks remained unknown [118]. The lack of actionability of VUS is often very distressing to patients [69]. Hence, in pre-test counselling, patients should be fully informed of the genes that will be tested, the level of penetrance, and the implications in cancer predisposition and clinical management. Fecteau et al., recommended that tested PV genes should be categorised into high, moderate, and unknown penetrance to help patients to understand the cancer risk profile of the genetic mutations, the understanding of them of scientists, and their clinical significance [119].
Post-test counselling should be guided by the findings of the test. The implications of the mutations should be explained to patients as well as their family members. Cascade testing for relatives of mutation carriers should also be recommended, if applicable [16]. It is essential that genetic counsellors fully explain what VUS results mean and their limitations. Patients should understand that although scientists currently have a limited knowledge of VUS, our understanding of them will improve with more data available in time and with international collaboration; hence the importance of regular follow-ups in the future.

4.5. Treatment and Recurrent Ovarian Cancer

As mentioned previously, different histologies of EOC may have different mutations; however, they have traditionally been grouped into one unit [120,121]. Most of the hereditary ovarian cancers are linked to BRCA1/2 mutations, especially in high-grade serous EOC. An analysis allows for genetic testing in relatives; if identified in asymptomatic carriers, preventative measures can be offered to reduce their risk of ovarian cancer [122]. Otherwise, if negative, they can be reassured their risk is similar to the general population risk. Risk-reducing options in BRCA1/2 carriers include a bilateral salpingo-oophorectomy (BSO), which is invasive compared with regular surveillance; that is, it is not always reliable in identifying the early stages [122]. A BSO is performed between the ages of 35 and 40 in BRCA1 carriers and 40 and 45 in BRCA2 carriers due to the late onset and can reduce the risk of ovarian cancer by 96% [123,124,125].
Careful counselling is required with planned pregnancies; menopausal symptoms can be induced, which have their own implications on quality of life and physical health (e.g., cardiovascular), and the use of hormone replacement therapy [123]. An analysis of mutations also allows tailored treatment planning, including PARP inhibitors, which have reformed ovarian cancer management [122]. Studies have found improved progression-free survival with olaparib and an efficacy of niraparib and rucaparib as maintenance treatments [122,126]. These studies examined recurrent EOC in particular and are summarised in Table 2.
Table 2. Studies of PARP inhibitors in recurrent EOC.
Table 2. Studies of PARP inhibitors in recurrent EOC.
Study/ReferencePopulationTreatment PlanMedian PFS
SOLO 1/[127]Stage III or IV high-grade serous or endometroid cancer with BRCA1/2 mutation and complete/partial response to CTH300 mg olaparib b.d. or placebo49.9 vs. 13.8 m (p < 0.0001)
SOLO 2/[126]Relapsed, platinum-sensitive EOC with a BRCA1/2 mutation300 mg olaparib b.d. or placebo19.9 vs. 5.5 m (p < 0.0001)
SOLO 3/[128]Germline BRCA-mutated ovarian cancer with relapse ≤ 12 m300 mg olaparib b.d. or non-platinum CTH of the choice of the physician13.4 vs. 9.2 m (p = 0.013)
NOVA/[129]Platinum-sensitive recurrent EOCNiraparib 300 mg o.d. or placeboGermline BRCA cohort: 21.0 vs. 5.5 m;
non-germline BRCA cohort: 9.3 vs. 3.9 m (p < 0.001)
EOC, epithelial ovarian cancer; PFS, progression-free survival; b.d., bis die (twice daily); CTH, chemotherapy; m, months.
Despite various treatment options, recurrent EOC is common, occurring in 20–25% of stage I or II patients and in 70% of advanced-disease patients. The first relapse can occur from a few months to five years after treatment and the median recurrence is 18 to 24 months [130]. Histology types do not appear to have a correlation with the recurrence rates [131]. The treatment of recurrent EOC is decided through platinum sensitivity based on the progression-free interval. If there is a response to platinum-based chemotherapy and the progression-free interval is over 6 months, the patients are considered to be platinum-sensitive [130]. Conversely, patients are considered to be platinum-resistant if the progression-free interval is less than 6 months, and refractory if less than 3 months [130,132]. For these patients, non-platinum chemotherapy may be used to increase the platinum-free interval, e.g., single-agent topotecan. This allows a better response to the platinum challenge at a later point [132]. A carboplatin-based combination is generally used for a platinum-sensitive disease; the options include gemcitabine, paclitaxel, and doxorubicin [130]. Surgery may be performed as a secondary cytoreductive surgery and more than 30% of these could involve a bowel resection; each patient should have specific treatment plans depending on their progression-free interval and prognostic factors [133].

5. Conclusions

Ovarian cancer is heavily influenced by hereditary factors and up to 25% of patients carry PVs in several genes. These PVs translate the proteins in various DNA repair pathways. The most common mutations are BRCA1/2 mutations with rarer mutations in other genes such as BRIP1, RAD51C, RAD51D, and MMR proteins. Most of the proteins are implicated in HR repair; treatments to target this pathway include PARP inhibitors, which allow highly specific chemotherapy planning. NGS has a high accuracy for detecting all mutation types as it allows the rapid testing of large genomic regions and identifies lower penetrance genes. Testing can also detect whether the mutations are somatic or germline. NGS is cost-effective as the cost is similar to single-gene sequencing. Testing for mutations allows asymptomatic carriers to be counselled and proactively managed as opposed to regular surveillance, which may not always be successful. More research is needed on the rarer mutations present in ovarian cancer as understanding and detecting the different mutations benefits from genetic counselling and improves the treatment options for ovarian cancer in both the immediate and longer future.

Author Contributions

Conceptualization, S.S. and S.B.; methodology, M.S. and S.B.; software, A.C. and M.K.; validation, S.S., M.S. and S.B.; formal analysis, A.C.; investigation, M.K.; resources, S.B.; data curation, A.C.; writing—original draft preparation, S.S.; writing—review and editing, S.S., A.C., M.K., and S.B.; visualization, M.S.; supervision, S.B.; project administration, S.B.; funding acquisition, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chaudhry, S.R.; Lopes, J.; Levin, N.K.; Kalpage, H.; Tainsky, M.A. Germline mutations in apoptosis pathway genes in ovarian cancer; the functional role of a TP53I3 (PIG3) variant in ROS production and DNA repair. Cell Death Discov. 2021, 7, 62. [Google Scholar] [CrossRef] [PubMed]
  2. Boussios, S.; Moschetta, M.; Karihtala, P.; Samartzis, E.P.; Sheriff, M.; Pappas-Gogos, G.; Ozturk, M.A.; Uccello, M.; Karathanasi, A.; Tringos, M.; et al. Development of new poly(ADP-ribose) polymerase (PARP) inhibitors in ovarian cancer: Quo Vadis? Ann. Transl. Med. 2020, 8, 1706. [Google Scholar] [CrossRef] [PubMed]
  3. Jones, M.R.; Kamara, D.; Karlan, B.Y.; Pharoah, P.D.P.; Gayther, S.A. Genetic epidemiology of ovarian cancer and prospects for polygenic risk prediction. Gynecol. Oncol. 2017, 147, 705–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. The World Ovarian Cancer Coalition Atlas. Global Trends in Incidence, Mortality and Survival. In World Ovarian Cancer Coalition; The World Ovarian Cancer Coalition: Toronto, ON, Canada, 2018; pp. 1–39. Available online: https://worldovariancancercoalition.org/wp-content/uploads/2018/10/THE-WORLD-OVARIAN-CANCER-COALITION-ATLAS-2018.pdf (accessed on 23 April 2022).
  5. Cancer of the Ovary—Cancer Stat Facts. In Surveillance, Epidemiology, and End Results (SEER) Program; The National Cancer Institute: Bethesda, MD, USA, 2018. Available online: https://seer.cancer.gov/statfacts/html/ovary.html (accessed on 23 April 2022).
  6. Flaum, N.; Crosbie, E.J.; Edmondson, R.J.; Smith, M.J.; Evans, D.G. Epithelial ovarian cancer risk: A review of the current genetic landscape. Clin. Genet. 2020, 97, 54–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Reid, B.M.; Permuth, J.B.; Sellers, T.A. Epidemiology of ovarian cancer: A review. Cancer Biol. Med. 2017, 14, 9–32. [Google Scholar]
  8. Vaughan, S.; Coward, J.I.; Bast, R.C., Jr.; Berchuck, A.; Berek, J.S.; Brenton, J.D.; Coukos, G.; Crum, C.C.; Drapkin, R.; Etemadmoghadam, D.; et al. Rethinking ovarian cancer: Recommendations for improving outcomes. Nat. Rev. Cancer 2011, 11, 719–725. [Google Scholar] [CrossRef] [Green Version]
  9. Pavlidis, N.; Rassy, E.; Vermorken, J.B.; Assi, T.; Kattan, J.; Boussios, S.; Smith-Gagen, J. The outcome of patients with serous papillary peritoneal cancer, fallopian tube cancer, and epithelial ovarian cancer by treatment eras: 27 years data from the SEER registry. Cancer Epidemiol. 2021, 75, 102045. [Google Scholar] [CrossRef]
  10. Boussios, S.; Zarkavelis, G.; Seraj, E.; Zerdes, I.; Tatsi, K.; Pentheroudakis, G. Non-epithelial Ovarian Cancer: Elucidating Uncommon Gynaecological Malignancies. Anticancer Res. 2016, 36, 5031–5042. [Google Scholar] [CrossRef] [Green Version]
  11. Cheung, A.; Shah, S.; Parker, J.; Soor, P.; Limbu, A.; Sheriff, M.; Boussios, S. Non-Epithelial Ovarian Cancers: How Much Do We Really Know? Int. J. Environ. Res. Public Health 2022, 19, 1106. [Google Scholar] [CrossRef]
  12. Boussios, S.; Moschetta, M.; Zarkavelis, G.; Papadaki, A.; Kefas, A.; Tatsi, K. Ovarian sex-cord stromal tumours and small cell tumours: Pathological, genetic and management aspects. Crit. Rev. Oncol. Hematol. 2017, 120, 43–51. [Google Scholar] [CrossRef]
  13. Boussios, S.; Karathanasi, A.; Zakynthinakis-Kyriakou, N.; Tsiouris, A.K.; Chatziantoniou, A.A.; Kanellos, F.S.; Tatsi, K. Ovarian carcinosarcoma: Current developments and future perspectives. Crit. Rev. Oncol. Hematol. 2019, 134, 46–55. [Google Scholar] [CrossRef] [PubMed]
  14. Moschetta, M.; Boussios, S.; Rassy, E.; Samartzis, E.P.; Funingana, G.; Uccello, M. Neoadjuvant treatment for newly diagnosed advanced ovarian cancer: Where do we stand and where are we going? Ann. Transl. Med. 2020, 8, 1710. [Google Scholar] [CrossRef] [PubMed]
  15. Boussios, S.; Moschetta, M.; Tatsi, K.; Tsiouris, A.K.; Pavlidis, N. A review on pregnancy complicated by ovarian epithelial and non-epithelial malignant tumors: Diagnostic and therapeutic perspectives. J. Adv. Res. 2018, 12, 1–9. [Google Scholar] [CrossRef] [PubMed]
  16. McCuaig, J.M.; Stockley, T.L.; Shaw, P.; Fung-Kee-Fung, M.; Altman, A.D.; Bentley, J.; Bernardini, M.Q.; Cormier, B.; Hirte, H.; Kieser, K.; et al. Evolution of genetic assessment for BRCA-associated gynaecologic malignancies: A Canadian multisociety roadmap. J. Med. Genet. 2018, 55, 571–577. [Google Scholar] [CrossRef] [PubMed]
  17. Eoh, K.J.; Kim, J.E.; Park, H.S.; Lee, S.T.; Park, J.S.; Han, J.W.; Lee, J.Y.; Kim, S.; Kim, S.W.; Kim, J.H.; et al. Detection of Germline Mutations in Patients with Epithelial Ovarian Cancer Using Multi-gene Panels: Beyond BRCA1/2. Cancer Res. Treat. 2018, 50, 917–925. [Google Scholar] [CrossRef] [Green Version]
  18. Norquist, B.M.; Harrell, M.I.; Brady, M.F.; Walsh, T.; Lee, M.K.; Gulsuner, S.; Bernards, S.S.; Casadei, S.; Yi, Q.; Burger, R.A.; et al. Inherited Mutations in Women with Ovarian Carcinoma. JAMA Oncol. 2016, 2, 482–490. [Google Scholar] [CrossRef]
  19. Kuchenbaecker, K.B.; Hopper, J.L.; Barnes, D.R.; Phillips, K.A.; Mooij, T.M.; Roos-Blom, M.J.; Jervis, S.; van Leeuwen, F.E.; Milne, R.L.; Andrieu, N.; et al. Risks of Breast, Ovarian, and Contralateral Breast Cancer for BRCA1 and BRCA2 Mutation Carriers. JAMA 2017, 317, 2402–2416. [Google Scholar] [CrossRef] [Green Version]
  20. Chirasophon, S.; Manchana, T.; Teerapakpinyo, C. High-risk epithelial ovarian cancer patients for hereditary ovarian cancer. J. Obstet. Gynaecol. Res. 2017, 43, 929–934. [Google Scholar] [CrossRef]
  21. Boussios, S.; Mikropoulos, C.; Samartzis, E.; Karihtala, P.; Moschetta, M.; Sheriff, M.; Karathanasi, A.; Sadauskaite, A.; Rassy, E.; Pavlidis, N. Wise Management of Ovarian Cancer: On the Cutting Edge. J. Pers. Med. 2020, 10, 41. [Google Scholar] [CrossRef]
  22. Santonocito, C.; Rizza, R.; Paris, I.; De Marchis, L.; Paolillo, C.; Tiberi, G.; Scambia, G.; Capoluongo, E. Spectrum of Germline BRCA1 and BRCA2 Variants Identified in 2351 Ovarian and Breast Cancer Patients Referring to a Reference Cancer Hospital of Rome. Cancers 2020, 12, 1286. [Google Scholar] [CrossRef]
  23. Venkitaraman, A.R. Cancer susceptibility and the functions of BRCA1 and BRCA2. Cell 2002, 108, 171–182. [Google Scholar] [CrossRef] [Green Version]
  24. Ford, D.; Easton, D.F.; Stratton, M.; Narod, S.; Goldgar, D.; Devilee, P.; Bishop, D.T.; Weber, B.; Lenoir, G.; Chang-Claude, J.; et al. Genetic heterogeneity and penetrance analysis of the BRCA1 and BRCA2 genes in breast cancer families. The Breast Cancer Linkage Consortium. Am. J. Hum. Genet. 1998, 62, 676–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. King, M.C.; Marks, J.H.; Mandell, J.B. Breast and ovarian cancer risks due to inherited mutations in BRCA1 and BRCA2. Science 2003, 302, 643–646. [Google Scholar] [CrossRef] [PubMed]
  26. Antoniou, A.; Pharoah, P.D.; Narod, S.; Risch, H.A.; Eyfjord, J.E.; Hopper, J.L.; Loman, N.; Olsson, H.; Johannsson, O.; Borg, A.; et al. Average risks of breast and ovarian cancer associated with BRCA1 or BRCA2 mutations detected in case Series unselected for family history: A combined analysis of 22 studies. Am. J. Hum. Genet. 2003, 72, 1117–1130. [Google Scholar] [CrossRef] [Green Version]
  27. Yang, D.; Khan, S.; Sun, Y.; Hess, K.; Shmulevich, I.; Sood, A.K.; Zhang, W. Association of BRCA1 and BRCA2 mutations with survival, chemotherapy sensitivity, and gene mutator phenotype in patients with ovarian cancer. JAMA 2011, 306, 1557–1565. [Google Scholar] [CrossRef] [Green Version]
  28. Vencken, P.M.L.H.; Kriege, M.; Hoogwerf, D.; Beugelink, S.; van der Burg, M.E.L.; Hooning, M.J.; Berns, E.M.; Jager, A.; Collée, M.; Burger, C.W.; et al. Chemosensitivity and outcome of BRCA1- and BRCA2-associated ovarian cancer patients after first-line chemotherapy compared with sporadic ovarian cancer patients. Ann. Oncol. 2011, 22, 1346–1352. [Google Scholar] [CrossRef]
  29. Fong, P.C.; Boss, D.S.; Yap, T.A.; Tutt, A.; Wu, P.; Mergui-Roelvink, M.; Mortimer, P.; Swaisland, H.; Lau, A.; O’Connor, M.J.; et al. Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N. Engl. J. Med. 2009, 361, 123–134. [Google Scholar] [CrossRef] [Green Version]
  30. Osher, D.J.; De Leeneer, K.; Michils, G.; Hamel, N.; Tomiak, E.; Poppe, B.; Leunen, K.; Legius, E.; Shuen, A.; Smith, E.; et al. Mutation analysis of RAD51D in non-BRCA1/2 ovarian and breast cancer families. Br. J. Cancer 2012, 106, 1460–1463. [Google Scholar] [CrossRef] [Green Version]
  31. Loveday, C.; Turnbull, C.; Ramsay, E.; Hughes, D.; Ruark, E.; Frankum, J.R.; Bowden, G.; Kalmyrzaev, B.; Warren-Perry, M.; Snape, K.; et al. Germline mutations in RAD51D confer susceptibility to ovarian cancer. Nat. Genet. 2011, 43, 879–882. [Google Scholar] [CrossRef]
  32. Meindl, A.; Hellebrand, H.; Wiek, C.; Erven, V.; Wappenschmidt, B.; Niederacher, D.; Freund, M.; Lichtner, P.; Hartmann, L.; Schaal, H.; et al. Germline mutations in breast and ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility gene. Nat. Genet. 2010, 42, 410–414. [Google Scholar] [CrossRef]
  33. Walsh, T.; Casadei, S.; Lee, M.K.; Pennil, C.C.; Nord, A.S.; Thornton, A.M.; Roeb, W.; Agnew, K.J.; Stray, S.M.; Wickramanayake, A.; et al. Mutations in 12 genes for inherited ovarian, fallopian tube, and peritoneal carcinoma identified by massively parallel sequencing. Proc. Natl. Acad. Sci. USA 2011, 108, 18032–18037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. O’Connor, M.J. Targeting the DNA Damage Response in Cancer. Mol. Cell 2015, 60, 547–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ciccia, A.; Elledge, S.J. The DNA damage response: Making it safe to play with knives. Mol. Cell 2010, 40, 179–204. [Google Scholar] [CrossRef] [Green Version]
  36. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Li, X.; Heyer, W.D. Homologous recombination in DNA repair and DNA damage tolerance. Cell Res. 2008, 18, 99–113. [Google Scholar] [CrossRef] [Green Version]
  38. Da Cunha Colombo Bonadio, R.R.; Fogace, R.N.; Miranda, V.C.; Diz, M.D.P.E. Homologous recombination deficiency in ovarian cancer: A review of its epidemiology and management. Clinics 2018, 73, e450s. [Google Scholar] [CrossRef] [PubMed]
  39. Angeli, D.; Salvi, S.; Tedaldi, G. Genetic Predisposition to Breast and Ovarian Cancers: How Many and Which Genes to Test? Int. J. Mol. Sci. 2020, 21, 1128. [Google Scholar] [CrossRef] [Green Version]
  40. de Bono, J.S.; Ashworth, A. Translating cancer research into targeted therapeutics. Nature 2010, 467, 543–549. [Google Scholar] [CrossRef]
  41. Byrski, T.; Gronwald, J.; Huzarski, T.; Grzybowska, E.; Budryk, M.; Stawicka, M.; Mierzwa, T.; Szwiec, M.; Wisniowski, R.; Siolek, M.; et al. Pathologic complete response rates in young women with BRCA1-positive breast cancers after neoadjuvant chemotherapy. J. Clin. Oncol. 2010, 28, 375–379. [Google Scholar] [CrossRef]
  42. Boussios, S.; Karathanasi, A.; Cooke, D.; Neille, C.; Sadauskaite, A.; Moschetta, M.; Zakynthinakis-Kyriakou, N.; Pavlidis, N. PARP Inhibitors in Ovarian Cancer: The Route to “Ithaca”. Diagnostics 2019, 9, 55. [Google Scholar] [CrossRef] [Green Version]
  43. Boussios, S.; Abson, C.; Moschetta, M.; Rassy, E.; Karathanasi, A.; Bhat, T.; Ghumman, F.; Sheriff, M.; Pavlidis, N. Poly (ADP-Ribose) Polymerase Inhibitors: Talazoparib in Ovarian Cancer and Beyond. Drugs R D 2020, 20, 55–73. [Google Scholar] [CrossRef] [Green Version]
  44. Boussios, S.; Karihtala, P.; Moschetta, M.; Karathanasi, A.; Sadauskaite, A.; Rassy, E.; Pavlidis, N. Combined Strategies with Poly (ADP-Ribose) Polymerase (PARP) Inhibitors for the Treatment of Ovarian Cancer: A Literature Review. Diagnostics 2019, 9, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Boussios, S.; Karihtala, P.; Moschetta, M.; Abson, C.; Karathanasi, A.; Zakynthinakis-Kyriakou, N.; Ryan, J.E.; Sheriff, M.; Rassy, E.; Pavlidis, N. Veliparib in ovarian cancer: A new synthetically lethal therapeutic approach. Investig. N. Drugs 2020, 38, 181–193. [Google Scholar] [CrossRef] [PubMed]
  46. Amé, J.C.; Spenlehauer, C.; de Murcia, G. The PARP superfamily. Bioessays 2004, 26, 882–893. [Google Scholar] [CrossRef] [PubMed]
  47. Dantzer, F.; de La Rubia, G.; Ménissier-De Murcia, J.; Hostomsky, Z.; de Murcia, G.; Schreiber, V. Base excision repair is impaired in mammalian cells lacking Poly(ADP-ribose) polymerase-1. Biochemistry 2000, 39, 7559–7569. [Google Scholar] [CrossRef]
  48. McCabe, N.; Turner, N.C.; Lord, C.J.; Kluzek, K.; Bialkowska, A.; Swift, S.; Giavara, S.; O’Connor, M.J.; Tutt, A.N.; Zdzienicka, M.Z.; et al. Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition. Cancer Res. 2006, 66, 8109–8115. [Google Scholar] [CrossRef] [Green Version]
  49. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef]
  50. Bryant, H.E.; Schultz, N.; Thomas, H.D.; Parker, K.M.; Flower, D.; Lopez, E.; Kyle, S.; Meuth, M.; Curtin, N.J.; Helleday, T. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 2005, 434, 913–917. [Google Scholar] [CrossRef]
  51. 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]
  52. Topatana, W.; Juengpanich, S.; Li, S.; Cao, J.; Hu, J.; Lee, J.; Suliyanto, K.; Ma, D.; Zhang, B.; Chen, M.; et al. Advances in synthetic lethality for cancer therapy: Cellular mechanism and clinical translation. J. Hematol. Oncol. 2020, 13, 118. [Google Scholar] [CrossRef]
  53. Eskander, R.N.; Tewari, K.S. PARP inhibition and synthetic lethality in ovarian cancer. Expert Rev. Clin. Pharmacol. 2014, 7, 613–622. [Google Scholar] [CrossRef] [PubMed]
  54. Ghose, A.; Moschetta, M.; Pappas-Gogos, G.; Sheriff, M.; Boussios, S. Genetic Aberrations of DNA Repair Pathways in Prostate Cancer: Translation to the Clinic. Int. J. Mol. Sci. 2021, 22, 9783. [Google Scholar] [CrossRef] [PubMed]
  55. Shah, S.; Rachmat, R.; Enyioma, S.; Ghose, A.; Revythis, A.; Boussios, S. BRCA Mutations in Prostate Cancer: Assessment, Implications and Treatment Considerations. Int. J. Mol. Sci. 2021, 22, 12628. [Google Scholar] [CrossRef]
  56. Boussios, S.; Rassy, E.; Shah, S.; Ioannidou, E.; Sheriff, M.; Pavlidis, N. Aberrations of DNA repair pathways in prostate cancer: A cornerstone of precision oncology. Expert Opin. Ther. Targets 2021, 25, 329–333. [Google Scholar] [CrossRef] [PubMed]
  57. Pal, T.; Permuth-Wey, J.; Sellers, T.A. A review of the clinical relevance of mismatch-repair deficiency in ovarian cancer. Cancer 2008, 113, 733–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Wajed, S.A.; Laird, P.W.; DeMeester, T.R. DNA methylation: An alternative pathway to cancer. Ann. Surg. 2001, 234, 10–20. [Google Scholar] [CrossRef]
  59. Cederquist, K.; Emanuelsson, M.; Wiklund, F.; Golovleva, I.; Palmqvist, R.; Grönberg, H. Two Swedish founder MSH6 mutations, one nonsense and one missense, conferring high cumulative risk of Lynch syndrome. Clin. Genet. 2005, 68, 533–541. [Google Scholar] [CrossRef]
  60. Song, H.; Cicek, M.S.; Dicks, E.; Harrington, P.; Ramus, S.J.; Cunningham, J.M.; Fridley, B.L.; Tyrer, J.P.; Alsop, J.; Jimenez-Linan, M.; et al. The contribution of deleterious germline mutations in BRCA1, BRCA2 and the mismatch repair genes to ovarian cancer in the population. Hum. Mol. Genet. 2014, 23, 4703–4709. [Google Scholar] [CrossRef] [Green Version]
  61. Jones, M.R.; Peng, P.C.; Coetzee, S.G.; Tyrer, J.; Reyes, A.L.P.; Corona, R.I.; Davis, B.; Chen, S.; Dezem, F.; Seo, J.H.; et al. Ovarian Cancer Risk Variants Are Enriched in Histotype-Specific Enhancers and Disrupt Transcription Factor Binding Sites. Am. J. Hum. Genet. 2020, 107, 622–635. [Google Scholar] [CrossRef]
  62. Sugino, K.; Tamura, R.; Nakaoka, H.; Yachida, N.; Yamaguchi, M.; Mori, Y.; Yamawaki, K.; Suda, K.; Ishiguro, T.; Adachi, S.; et al. Germline and somatic mutations of homologous recombination-associated genes in Japanese ovarian cancer patients. Sci. Rep. 2019, 9, 17808. [Google Scholar] [CrossRef] [Green Version]
  63. Alsop, K.; Fereday, S.; Meldrum, C.; deFazio, A.; Emmanuel, C.; George, J.; Dobrovic, A.; Birrer, M.J.; Webb, P.M.; Stewart, C.; et al. BRCA mutation frequency and patterns of treatment response in BRCA mutation-positive women with ovarian cancer: A report from the Australian Ovarian Cancer Study Group. J. Clin. Oncol. 2012, 30, 2654–2663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Klotz, D.M.; Wimberger, P. Cells of origin of ovarian cancer: Ovarian surface epithelium or fallopian tube? Arch. Gynecol. Obstet. 2017, 296, 1055–1062. [Google Scholar] [CrossRef] [PubMed]
  65. Cancer Genome Atlas Research Network. Integrated genomic analyses of ovarian carcinoma. Nature 2011, 474, 609–615. [Google Scholar] [CrossRef] [PubMed]
  66. Lheureux, S.; Gourley, C.; Vergote, I.; Oza, A.M. Epithelial ovarian cancer. Lancet 2019, 393, 1240–1253. [Google Scholar] [CrossRef] [Green Version]
  67. Kang, S.; Yu, Y.L.; Cho, S.Y.; Park, S.Y. Prevalence of pathogenic variants in actionable genes in advanced ovarian cancer: A next-generation sequencing analysis of a nationwide registry study. Eur. J. Cancer 2020, 141, 185–192. [Google Scholar] [CrossRef] [PubMed]
  68. Bychkovsky, B.L.; Lo, M.T.; Yussuf, A.; Horton, C.; Richardson, M.; LaDuca, H.; Garber, J.E.; Rana, H.Q. Prevalence and spectrum of pathogenic variants among patients with multiple primary cancers evaluated by clinical characteristics. Cancer 2022, 128, 1275–1283. [Google Scholar] [CrossRef] [PubMed]
  69. Li, S.; Silvestri, V.; Leslie, G.; Rebbeck, T.R.; Neuhausen, S.L.; Hopper, J.L.; Nielsen, H.R.; Lee, A.; Yang, X.; McGuffog, L.; et al. Cancer Risks Associated with BRCA1 and BRCA2 Pathogenic Variants. J. Clin. Oncol. 2022, 40, 1529. [Google Scholar] [CrossRef]
  70. Boni, J.; Idani, A.; Roca, C.; Feliubadaló, L.; Tomiak, E.; Weber, E.; Foulkes, W.D.; Orthwein, A.; El Haffaf, Z.; Lazaro, C.; et al. A decade of RAD51C and RAD51D germline variants in cancer. Hum. Mutat. 2022, 43, 285–298. [Google Scholar] [CrossRef]
  71. Kotsopoulos, J.; Sopik, V.; Rosen, B.; Fan, I.; McLaughlin, J.R.; Risch, H.; Sun, P.; Narod, S.A.; Akbari, M.R. Frequency of germline PALB2 mutations among women with epithelial ovarian cancer. Fam. Cancer 2017, 16, 29–34. [Google Scholar] [CrossRef]
  72. Manchana, T.; Phoolcharoen, N.; Tantbirojn, P. BRCA mutation in high grade epithelial ovarian cancers. Gynecol. Oncol. Rep. 2019, 29, 102–105. [Google Scholar] [CrossRef]
  73. Matulonis, U.A.; Sood, A.K.; Fallowfield, L.; Howitt, B.E.; Sehouli, J.; Karlan, B.Y. Ovarian cancer. Nat. Rev. Dis. Primers 2016, 2, 16061. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, S.; Royer, R.; Li, S.; McLaughlin, J.R.; Rosen, B.; Risch, H.A.; Fan, I.; Bradley, L.; Shaw, P.A.; Narod, S.A. Frequencies of BRCA1 and BRCA2 mutations among 1342 unselected patients with invasive ovarian cancer. Gynecol. Oncol. 2011, 121, 353–357. [Google Scholar] [CrossRef] [PubMed]
  75. Bolton, K.L.; Chenevix-Trench, G.; Goh, C.; Sadetzki, S.; Ramus, S.J.; Karlan, B.Y.; 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] [PubMed] [Green Version]
  76. Suszynska, M.; Ratajska, M.; Kozlowski, P. BRIP1, RAD51C, and RAD51D mutations are associated with high susceptibility to ovarian cancer: Mutation prevalence and precise risk estimates based on a pooled analysis of ~30,000 cases. J. Ovarian Res. 2020, 13, 50. [Google Scholar] [CrossRef]
  77. Ramus, S.J.; Song, H.; Dicks, E.; Tyrer, J.P.; Rosenthal, A.N.; Intermaggio, M.P.; Fraser, L.; Gentry-Maharaj, A.; Hayward, J.; Philpott, S.; et al. Germline Mutations in the BRIP1, BARD1, PALB2, and NBN Genes in Women with Ovarian Cancer. J. Natl. Cancer Inst. 2015, 107, djv214. [Google Scholar] [CrossRef]
  78. Lilyquist, J.; LaDuca, H.; Polley, E.; Davis, B.T.; Shimelis, H.; Hu, C.; Hart, S.N.; Dolinsky, J.S.; Couch, F.J.; Goldgar, D.E. Frequency of mutations in a large series of clinically ascertained ovarian cancer cases tested on multi-gene panels compared to reference controls. Gynecol. Oncol. 2017, 147, 375–380. [Google Scholar] [CrossRef]
  79. Weber-Lassalle, N.; Hauke, J.; Ramser, J.; Richters, L.; Groß, E.; Blümcke, B.; Gehrig, A.; Kahlert, A.K.; Müller, C.R.; Hackmann, K.; et al. BRIP1 loss-of-function mutations confer high risk for familial ovarian cancer, but not familial breast cancer. Breast Cancer Res. 2018, 20, 7. [Google Scholar] [CrossRef]
  80. Clague, J.; Wilhoite, G.; Adamson, A.; Bailis, A.; Weitzel, J.N.; Neuhausen, S.L. RAD51C germline mutations in breast and ovarian cancer cases from high-risk families. PLoS ONE 2011, 6, e25632. [Google Scholar] [CrossRef]
  81. Thompson, D.; Easton, D. Variation in cancer risks, by mutation position, in BRCA2 mutation carriers. Am. J. Hum. Genet. 2001, 68, 410–419. [Google Scholar] [CrossRef] [Green Version]
  82. Miller, K.A.; Sawicka, D.; Barsky, D.; Albala, J.S. Domain mapping of the Rad51 paralog protein complexes. Nucleic Acids Res. 2004, 32, 169–178. [Google Scholar] [CrossRef]
  83. Yu, X.; Chini, C.C.; He, M.; Mer, G.; Chen, J. The BRCT domain is a phospho-protein binding domain. Science. 2003, 302, 639–642. [Google Scholar] [CrossRef] [PubMed]
  84. Bork, P.; Hofmann, K.; Bucher, P.; Neuwald, A.F.; Altschul, S.F.; Koonin, E.V. A superfamily of conserved domains in DNA damage-responsive cell cycle checkpoint proteins. FASEB J. 1997, 11, 68–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Seal, S.; Thompson, D.; Renwick, A.; Elliott, A.; Kelly, P.; Barfoot, R.; Chagtai, T.; Jayatilake, H.; Ahmed, M.; Spanova, K.; et al. Truncating mutations in the Fanconi anemia J gene BRIP1 are low-penetrance breast cancer susceptibility alleles. Nat. Genet. 2006, 38, 1239–1241. [Google Scholar] [CrossRef] [PubMed]
  86. Tyrer, J.P.; Guo, Q.; Easton, D.F.; Pharoah, P.D. The admixture maximum likelihood test to test for association between rare variants and disease phenotypes. BMC Bioinform. 2013, 14, 177. [Google Scholar] [CrossRef] [Green Version]
  87. Helleman, J.; van Staveren, I.L.; Dinjens, W.N.; van Kuijk, P.F.; Ritstier, K.; Ewing, P.C.; van der Burg, M.E.; Stoter, G.; Berns, E.M. Mismatch repair and treatment resistance in ovarian cancer. BMC Cancer 2006, 6, 201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Nakamura, K.; Banno, K.; Yanokura, M.; Iida, M.; Adachi, M.; Masuda, K.; Ueki, A.; Kobayashi, Y.; Nomura, H.; Hirasawa, A.; et al. Features of ovarian cancer in Lynch syndrome (Review). Mol. Clin. Oncol. 2014, 2, 909–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Bernards, S.S.; Norquist, B.M.; Harrell, M.I.; Agnew, K.J.; Lee, M.K.; Walsh, T.; Swisher, E.M. Genetic characterization of early onset ovarian carcinoma. Gynecol. Oncol. 2016, 140, 221–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Pal, T.; Akbari, M.R.; Sun, P.; Lee, J.H.; Fulp, J.; Thompson, Z.; Coppola, D.; Nicosia, S.; Sellers, T.A.; McLaughlin, J.; et al. Frequency of mutations in mismatch repair genes in a population-based study of women with ovarian cancer. Br. J. Cancer 2012, 107, 1783–1790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Crijnen, T.E.; Janssen-Heijnen, M.L.; Gelderblom, H.; Morreau, J.; Nooij, M.A.; Kenter, G.G.; Vasen, H.F. Survival of patients with ovarian cancer due to a mismatch repair defect. Fam. Cancer 2005, 4, 301–305. [Google Scholar] [CrossRef]
  92. Kamps, R.; Brandão, R.D.; Bosch, B.J.; Paulussen, A.D.; Xanthoulea, S.; Blok, M.J.; Romano, A. Next-Generation Sequencing in Oncology: Genetic Diagnosis, Risk Prediction and Cancer Classification. Int. J. Mol. Sci. 2017, 18, 308. [Google Scholar] [CrossRef]
  93. Cobain, E.F.; Milliron, K.J.; Merajver, S.D. Updates on breast cancer genetics: Clinical implications of detecting syndromes of inherited increased susceptibility to breast cancer. Semin. Oncol. 2016, 43, 528–535. [Google Scholar] [CrossRef] [PubMed]
  94. Lynce, F.; Isaacs, C. How Far Do We Go with Genetic Evaluation? Gene, Panel, and Tumor Testing. Am. Soc. Clin. Oncol. Educ. Book. 2016, 35, e72–e78. [Google Scholar] [CrossRef] [PubMed]
  95. Andoni, T.; Wiggins, J.; Robinson, R.; Charlton, R.; Sandberg, M.; Eeles, R. Half of germline pathogenic and likely pathogenic variants found on panel tests do not fulfil NHS testing criteria. Sci. Rep. 2022, 12, 2507. [Google Scholar] [CrossRef] [PubMed]
  96. Pritchard, C.C.; Smith, C.; Salipante, S.J.; Lee, M.K.; Thornton, A.M.; Nord, A.S.; Gulden, C.; Kupfer, S.S.; Swisher, E.M.; Bennett, R.L.; et al. ColoSeq provides comprehensive lynch and polyposis syndrome mutational analysis using massively parallel sequencing. J. Mol. Diagn. 2012, 14, 357–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Frey, M.K.; Kim, S.H.; Bassett, R.Y.; Martineau, J.; Dalton, E.; Chern, J.Y.; Blank, S.V. Rescreening for genetic mutations using multi-gene panel testing in patients who previously underwent non-informative genetic screening. Gynecol. Oncol. 2015, 139, 211–215. [Google Scholar] [CrossRef]
  98. Sun, L.; Brentnall, A.; Patel, S.; Buist, D.; Bowles, E.; Evans, D.G.; Eccles, D.; Hopper, J.; Li, S.; Duffy, S.; et al. Should we offer multi-gene testing to all patients with breast cancer: A cost-effectiveness analysis. Int. J. Gynecol. Cancer 2019, 29, A31–A32. [Google Scholar]
  99. Asphaug, L.; Melberg, H.O. The Cost-Effectiveness of Multigene Panel Testing for Hereditary Breast and Ovarian Cancer in Norway. MDM Policy Pr. 2019, 4, 2381468318821103. [Google Scholar] [CrossRef] [Green Version]
  100. Bonadio, R.C.; Crespo, J.R.; Estevez-Diz, M.D.P. Ovarian cancer risk assessment in the era of next-generation sequencing. Ann. Transl. Med. 2020, 8, 1704. [Google Scholar] [CrossRef]
  101. Bekos, C.; Grimm, C.; Kranawetter, M.; Polterauer, S.; Oberndorfer, F.; Tan, Y.; Müllauer, L.; Singer, C.F. Reliability of Tumor Testing Compared to Germline Testing for Detecting BRCA1 and BRCA2 Mutations in Patients with Epithelial Ovarian Cancer. J. Pers. Med. 2021, 11, 593. [Google Scholar] [CrossRef]
  102. Dougherty, B.A.; Lai, Z.; Hodgson, D.R.; Orr, M.C.M.; Hawryluk, M.; Sun, J.; Yelensky, R.; Spencer, S.K.; Robertson, J.D.; Ho, T.W.; et al. Biological and clinical evidence for somatic mutations in BRCA1 and BRCA2 as predictive markers for olaparib response in high-grade serous ovarian cancers in the maintenance setting. Oncotarget 2017, 8, 43653–43661. [Google Scholar] [CrossRef] [Green Version]
  103. Fumagalli, C.; Tomao, F.; Betella, I.; Rappa, A.; Calvello, M.; Bonanni, B.; Bernard, L.; Peccatori, F.; Colombo, N.; Viale, G.; et al. Tumor BRCA Test for Patients with Epithelial Ovarian Cancer: The Role of Molecular Pathology in the Era of PARP Inhibitor Therapy. Cancers 2019, 11, 1641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Emerson, R.O.; Sherwood, A.M.; Rieder, M.J.; Guenthoer, J.; Williamson, D.W.; Carlson, C.S.; Drescher, C.W.; Tewari, M.; Bielas, J.H.; Robins, H.S. High-throughput sequencing of T-cell receptors reveals a homogeneous repertoire of tumour-infiltrating lymphocytes in ovarian cancer. J. Pathol. 2013, 231, 433–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ostmeyer, J.; Lucas, E.; Christley, S.; Lea, J.; Monson, N.; Tiro, J.; Cowell, L.G. Biophysicochemical motifs in T cell receptor sequences as a potential biomarker for high-grade serous ovarian carcinoma. PLoS ONE 2020, 15, e0229569. [Google Scholar] [CrossRef] [PubMed]
  106. Hwang, W.T.; Adams, S.F.; Tahirovic, E.; Hagemann, I.S.; Coukos, G. Prognostic significance of tumor-infiltrating T cells in ovarian cancer: A meta-analysis. Gynecol. Oncol. 2012, 124, 192–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Zhang, L.; Conejo-Garcia, J.R.; Katsaros, D.; Gimotty, P.A.; Massobrio, M.; Regnani, G.; Makrigiannakis, A.; Gray, H.; Schlienger, K.; Liebman, M.N.; et al. Intratumoral T cells, recurrence, and survival in epithelial ovarian cancer. N. Engl. J. Med. 2003, 348, 203–213. [Google Scholar] [CrossRef] [Green Version]
  108. Santoiemma, P.P.; Powell, D.J., Jr. Tumor infiltrating lymphocytes in ovarian cancer. Cancer Biol. Ther. 2015, 16, 807–820. [Google Scholar] [CrossRef]
  109. Aoki, Y.; Takakuwa, K.; Kodama, S.; Tanaka, K.; Takahashi, M.; Tokunaga, A.; Takahashi, T. Use of adoptive transfer of tumor-infiltrating lymphocytes alone or in combination with cisplatin-containing chemotherapy in patients with epithelial ovarian cancer. Cancer Res. 1991, 51, 1934–1939. [Google Scholar]
  110. Freedman, R.S.; Edwards, C.L.; Kavanagh, J.J.; Kudelka, A.P.; Katz, R.L.; Carrasco, C.H.; Atkinson, E.N.; Scott, W.; Tomasovic, B.; Templin, S.; et al. Intraperitoneal adoptive immunotherapy of ovarian carcinoma with tumor-infiltrating lymphocytes and low-dose recombinant interleukin-2: A pilot trial. J. Immunother. Emphas. Tumor Immunol. 1994, 16, 198–210. [Google Scholar] [CrossRef]
  111. Wright, S.E.; Rewers-Felkins, K.A.; Quinlin, I.S.; Phillips, C.A.; Townsend, M.; Philip, R.; Dobrzanski, M.J.; Lockwood-Cooke, P.R.; Robinson, W. Cytotoxic T-lymphocyte immunotherapy for ovarian cancer: A pilot study. J. Immunother. 2012, 35, 196–204. [Google Scholar] [CrossRef] [Green Version]
  112. Owens, D.K.; Davidson, K.W.; Krist, A.H.; Barry, M.J.; Cabana, M.; Caughey, A.B.; Doubeni, C.A.; Epling, J.W., Jr.; Kubik, M.; Landefeld, C.S.; et al. Risk Assessment, Genetic Counseling, and Genetic Testing for BRCA-Related Cancer: US Preventive Services Task Force Recommendation Statement. JAMA 2019, 322, 652–665. [Google Scholar]
  113. Konstantinopoulos, P.A.; Norquist, B.; Lacchetti, C.; Armstrong, D.; Grisham, R.N.; Goodfellow, P.J.; Kohn, E.C.; Levine, D.A.; Liu, J.F.; Lu, K.H.; et al. Germline and Somatic Tumor Testing in Epithelial Ovarian Cancer: ASCO Guideline. J. Clin. Oncol. 2020, 38, 1222–1245. [Google Scholar] [CrossRef] [PubMed]
  114. Childers, K.K.; Maggard-Gibbons, M.; Macinko, J.; Childers, C.P. National Distribution of Cancer Genetic Testing in the United States: Evidence for a Gender Disparity in Hereditary Breast and Ovarian Cancer. JAMA Oncol. 2018, 4, 876–879. [Google Scholar] [CrossRef] [PubMed]
  115. McGee, J.; Panabaker, K.; Leonard, S.; Ainsworth, P.; Elit, L.; Shariff, S.Z. Genetics Consultation Rates Following a Diagnosis of High-Grade Serous Ovarian Carcinoma in the Canadian Province of Ontario. Int. J. Gynecol. Cancer 2017, 27, 437–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Metcalfe, K.A.; Fan, I.; McLaughlin, J.; Risch, H.A.; Rosen, B.; Murphy, J.; Bradley, L.; Armel, S.; Sun, P.; Narod, S.A. Uptake of clinical genetic testing for ovarian cancer in Ontario: A population-based study. Gynecol. Oncol. 2009, 112, 68–72. [Google Scholar] [CrossRef] [Green Version]
  117. Nielsen, S.M.; Eccles, D.M.; Romero, I.L.; Al-Mulla, F.; Balmaña, J.; Biancolella, M.; Bslok, R.; Caligo, M.A.; Calvello, M.; Capone, G.L.; et al. Genetic Testing and Clinical Management Practices for Variants in Non-BRCA1/2 Breast (and Breast/Ovarian) Cancer Susceptibility Genes: An International Survey by the Evidence-Based Network for the Interpretation of Germline Mutant Alleles (ENIGMA) Clinical Working Group. JCO Precis. Oncol. 2018, 2, PO.18.00091. [Google Scholar]
  118. Mersch, J.; Brown, N.; Pirzadeh-Miller, S.; Mundt, E.; Cox, H.C.; Brown, K.; Aston, M.; Esterling, L.; Manley, S.; Ross, T. Prevalence of Variant Reclassification Following Hereditary Cancer Genetic Testing. JAMA 2018, 320, 1266–1274. [Google Scholar] [CrossRef]
  119. Fecteau, H.; Vogel, K.J.; Hanson, K.; Morrill-Cornelius, S. The evolution of cancer risk assessment in the era of next generation sequencing. J. Genet. Couns. 2014, 23, 633–639. [Google Scholar] [CrossRef]
  120. Gilks, C.B.; Prat, J. Ovarian carcinoma pathology and genetics: Recent advances. Hum. Pathol. 2009, 40, 1213–1223. [Google Scholar] [CrossRef]
  121. Prat, J.; D’Angelo, E.; Espinosa, I. Ovarian carcinomas: At least five different diseases with distinct histological features and molecular genetics. Hum. Pathol. 2018, 80, 11–27. [Google Scholar] [CrossRef]
  122. Fostira, F.; Papadimitriou, M.; Papadimitriou, C. Current practices on genetic testing in ovarian cancer. Ann. Transl. Med. 2020, 8, 1703. [Google Scholar] [CrossRef]
  123. Gasparri, M.L.; Taghavi, K.; Fiacco, E.; Zuber, V.; Di Micco, R.; Gazzetta, G.; Valentini, A.; Mueller, M.D.; Papadia, A.; Gentilini, O.D. Risk-Reducing Bilateral Salpingo-Oophorectomy for BRCA Mutation Carriers and Hormonal Replacement Therapy: If It Should Rain, Better a Drizzle than a Storm. Medicina 2019, 55, 415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Finch, A.P.; Lubinski, J.; Møller, P.; Singer, C.F.; Karlan, B.; Senter, L.; Rosen, B.; Maehle, L.; Ghadirian, P.; Cybulski, C.; et al. Impact of oophorectomy on cancer incidence and mortality in women with a BRCA1 or BRCA2 mutation. J. Clin. Oncol. 2014, 32, 1547–1553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Domchek, S.M.; Rebbeck, T.R. Prophylactic oophorectomy in women at increased cancer risk. Curr. Opin. Obstet. Gynecol. 2007, 19, 27–30. [Google Scholar] [CrossRef] [PubMed]
  126. Pujade-Lauraine, E.; Ledermann, J.A.; Selle, F.; Gebski, V.; Penson, R.T.; Oza, A.M.; Korach, J.; Huzarski, T.; Poveda, A.; Pignata, S.; et al. Olaparib tablets as maintenance therapy in patients with platinum-sensitive, relapsed ovarian cancer and a BRCA1/2 mutation (SOLO2/ENGOT-Ov21): A double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 2017, 18, 1274–1284. [Google Scholar] [CrossRef] [Green Version]
  127. Moore, K.; Colombo, N.; Scambia, G.; Kim, B.G.; Oaknin, A.; Friedlander, M.; Lisyanskaya, A.; Floquet, A.; Leary, A.; Sonke, G.S.; et al. Maintenance Olaparib in Patients with Newly Diagnosed Advanced Ovarian Cancer. N. Engl. J. Med. 2018, 379, 2495–2505. [Google Scholar] [CrossRef]
  128. Penson, R.T.; Valencia, R.V.; Cibula, D.; Colombo, N.; Leath, C.A., 3rd; Bidziński, M.; Kim, J.W.; Nam, J.H.; Madry, R.; Hernández, C.; et al. Olaparib Versus Nonplatinum Chemotherapy in Patients with Platinum-Sensitive Relapsed Ovarian Cancer and a Germline BRCA1/2 Mutation (SOLO3): A Randomized Phase III Trial. J. Clin. Oncol. 2020, 38, 1164–1174. [Google Scholar] [CrossRef]
  129. Mirza, M.R.; Monk, B.J.; Herrstedt, J.; Oza, A.M.; Mahner, S.; Redondo, A.; Fabbro, M.; Ledermann, J.A.; Lorusso, D.; Vergote, I.; et al. Niraparib Maintenance Therapy in Platinum-Sensitive, Recurrent Ovarian Cancer. N. Engl. J. Med. 2016, 375, 2154–2164. [Google Scholar] [CrossRef]
  130. Ushijima, K. Treatment for recurrent ovarian cancer-at first relapse. J. Oncol. 2010, 2010, 497429. [Google Scholar] [CrossRef]
  131. Gadducci, A.; Fuso, L.; Cosio, S.; Landoni, F.; Maggino, T.; Perotto, S.; Sartori, E.; Testa, A.; Galletto, L.; Zola, P. Are surveillance procedures of clinical benefit for patients treated for ovarian cancer?: A retrospective Italian multicentric study. Int. J. Gynecol. Cancer 2009, 19, 367–374. [Google Scholar] [CrossRef]
  132. Spriggs, D. Optimal sequencing in the treatment of recurrent ovarian cancer. Gynecol. Oncol. 2003, 90, S39–S44. [Google Scholar] [CrossRef]
  133. Onda, T.; Yoshikawa, H.; Yasugi, T.; Yamada, M.; Matsumoto, K.; Taketani, Y. Secondary cytoreductive surgery for recurrent epithelial ovarian carcinoma: Proposal for patients selection. Br. J. Cancer 2005, 92, 1026–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Overview of the roles of BRCA1 and BRCA2 in the DNA repair mechanism. (A). BRCA1 and BRCA2 loci on chromosomes 17 and 13, respectively. (B). The initiation of the double-stranded DNA (DSBs) break correction starts with BRCA1 binding to the site of damage, thus initiating the precise repair via homologous repair (HR) and preventing non-homologous end joining (NHEJ).
Figure 1. Overview of the roles of BRCA1 and BRCA2 in the DNA repair mechanism. (A). BRCA1 and BRCA2 loci on chromosomes 17 and 13, respectively. (B). The initiation of the double-stranded DNA (DSBs) break correction starts with BRCA1 binding to the site of damage, thus initiating the precise repair via homologous repair (HR) and preventing non-homologous end joining (NHEJ).
Ijerph 19 08113 g001
Figure 2. Distribution of germline pathogenic variants (PVs) identified in unselected epithelial ovarian cancer (EOC) patients. Among 21.5% of those PVs, 15% represent alterations to BRCA1/2 genes and 3.5% of genetic aberrations in other genes compromise the homologous recombination (HR) pathway whilst 3% are PV genes involved in the DNA mismatch repair (MMR) pathway.
Figure 2. Distribution of germline pathogenic variants (PVs) identified in unselected epithelial ovarian cancer (EOC) patients. Among 21.5% of those PVs, 15% represent alterations to BRCA1/2 genes and 3.5% of genetic aberrations in other genes compromise the homologous recombination (HR) pathway whilst 3% are PV genes involved in the DNA mismatch repair (MMR) pathway.
Ijerph 19 08113 g002
Table 1. Common mutations according to ovarian cancer histologies.
Table 1. Common mutations according to ovarian cancer histologies.
Genes Affected (%)Histology
HGSOCEnOCCCOC
BRCA18.0--
BRCA24.05.15.1
ATM4.09.117.9
BRIP12.02.0-
PALB2-2.02.6
RAD50-1.02.6
HGSOC, high-grade serous ovarian cancer; EnOC, endometrioid ovarian cancer; CCOC, clear cell ovarian cancer.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shah, S.; Cheung, A.; Kutka, M.; Sheriff, M.; Boussios, S. Epithelial Ovarian Cancer: Providing Evidence of Predisposition Genes. Int. J. Environ. Res. Public Health 2022, 19, 8113. https://doi.org/10.3390/ijerph19138113

AMA Style

Shah S, Cheung A, Kutka M, Sheriff M, Boussios S. Epithelial Ovarian Cancer: Providing Evidence of Predisposition Genes. International Journal of Environmental Research and Public Health. 2022; 19(13):8113. https://doi.org/10.3390/ijerph19138113

Chicago/Turabian Style

Shah, Sidrah, Alison Cheung, Mikolaj Kutka, Matin Sheriff, and Stergios Boussios. 2022. "Epithelial Ovarian Cancer: Providing Evidence of Predisposition Genes" International Journal of Environmental Research and Public Health 19, no. 13: 8113. https://doi.org/10.3390/ijerph19138113

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