Open Access

Identification of the most common BRCA alterations through analysis of germline mutation databases: Is droplet digital PCR an additional strategy for the assessment of such alterations in breast and ovarian cancer families?

  • Authors:
    • Alessandro Lavoro
    • Aurora Scalisi
    • Saverio Candido
    • Guido Nicola Zanghì
    • Roberta Rizzo
    • Giuseppe Gattuso
    • Giuseppe Caruso
    • Massimo Libra
    • Luca Falzone
  • View Affiliations

  • Published online on: April 6, 2022     https://doi.org/10.3892/ijo.2022.5349
  • Article Number: 58
  • Copyright: © Lavoro et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

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


Abstract

Breast and ovarian cancer represent two of the most common tumor types in females worldwide. Over the years, several non‑modifiable and modifiable risk factors have been associated with the onset and progression of these tumors, including age, reproductive factors, ethnicity, socioeconomic status and lifestyle factors, as well as family history and genetic factors. Of note, BRCA1 and BRCA2 are two tumor suppressor genes with a key role in DNA repair processes, whose mutations may induce genomic instability and increase the risk of cancer development. Specifically, females with a family history of breast or ovarian cancer harboring BRCA1/2 germline mutations have a 60‑70% increased risk of developing breast cancer and a 15‑40% increased risk for ovarian cancer. Different databases have collected the most frequent germline mutations affecting BRCA1/2. Through the analysis of such databases, it is possible to identify frequent hotspot mutations that may be analyzed with next‑generation sequencing (NGS) and novel innovative strategies. In this context, NGS remains the gold standard method for the assessment of BRCA1/2 mutations, while novel techniques, including droplet digital PCR (ddPCR), may improve the sensitivity to identify such mutations in the hereditary forms of breast and ovarian cancer. On these bases, the present study aimed to provide an update of the current knowledge on the frequency of BRCA1/2 mutations and cancer susceptibility, focusing on the diagnostic potential of the most recent methods, such as ddPCR.

Introduction

Breast and ovarian cancer are two of the most common malignancies affecting the female population both in industrialized and developing countries. According to GLOBOCAN 2020, breast cancer is the most frequently diagnosed tumor among females with 2,261,419 new cases (24.5%) and the first leading cause of cancer-related death with 684,996 deaths (15.5%). On the other hand, ovarian cancer represents the eighth most common cancer for incidence and mortality in females, accounting for 313,959 newly diagnosed cases (3,4%) and 207,252 deaths (4.7%) (1).

Over the years, several risk factors have been associated with the onset and progression of both breast and ovarian cancer (Fig. 1). As widely reported in the literature, more than half of all cases diagnosed are females aged >50 years, indicating that age is one of the major non-modifiable risk factors (2,3). Similarly, reproductive factors, such as age at menarche and menopause (before 11 and after 55 years of age, respectively), nulliparity and age at first full-term pregnancy (>35 years), are well-established risk factors for both breast and ovarian cancer (4,5). Furthermore, post-menopausal hormone therapies, based on the administration of estrogens plus progestin, significantly increase the risk of cancer development (6,7). Of note, the use of oral contraceptives for birth control has only been described as a risk factor for breast cancer, while oral contraceptive pills represent a protective factor for ovarian cancer, reducing the risk by at least 50% when used for 10 years or more (8-10). Other risk factors for both of these female cancer types include ethnicity, tobacco smoking, alcohol consumption, low physical activity, high-fat diet, obesity and socioeconomic status (11-16). Several studies have demonstrated that environmental, lifestyle and epigenetic factors are associated with the development of both breast and ovarian cancer (17-20).

Besides the aforementioned modifiable and non-modifiable risk factors, a family history of breast and/or ovarian cancer, as well as genetic mutations, may have a key role in increasing cancer susceptibility. Of note, a growing body of evidence suggests that the risk of developing these female cancers is significantly increased in females having a first-degree relative affected by breast cancer or ovarian cancer or when the affected relative was under 50 years of age (21-24). Of note, over the years, several genetic mutations have been reported to be highly associated with an increased risk of both breast and ovarian cancer. Among these, BReast CAncer 1 (BRCA1) and BRCA2 are two tumor suppressor genes with high penetrance, which are involved in the activation of DNA repair processes and cell-cycle checkpoints in response to DNA damage (25,26). Functional deficiencies due to BRCA1/2 mutations induce genome instability, cell-cycle dysregulation and accumulation of other mutations (27,28). In this field, it has been widely demonstrated that BRCA1/2 mutations increase the lifetime risk to develop breast or ovarian cancer. Specifically, 5-10% of patients with breast cancer and 25% of ovarian cancer cases are due to an inherited genetic mutation affecting BRCA1 or BRCA2 (29,30). Furthermore, healthy individuals harboring germline mutations of BRCA1/2 had a 60-70% increased risk to develop breast cancer and a 15-40% increased risk for ovarian cancer (31,32). Therefore, the assessment of cancer-related risk factors, particularly the identification of BRCA1 and BRCA2 gene mutations, is crucial for the clinical management of females at a high risk of developing breast or ovarian cancer, which includes annual screening, chemoprevention and preventive surgery (33,34).

Although the impairment of DNA repair mechanisms due to BRCA mutations is associated with an increased risk of breast and ovarian cancer, patients developing these tumors benefit from therapies further affecting the DNA repair machinery aimed at killing cancer cells through the accumulation of several DNA alterations resulting in tumor cell death (35,36). Among these therapies, the use of poly [ADP-ribose] polymerase (PARP) inhibitors directed at PARP proteins involved in DNA repair mechanisms proved to be highly efficient in both ovarian and breast cancer (37,38).

The precise molecular characterization of both breast and ovarian tumors is essential to correctly classify cancer lesions and predict the prognosis of patients. With regard to breast cancer, the presence of hormone receptors (estrogen receptor and progesterone receptor), membrane receptors (epidermal growth factor receptor) and other molecular markers, including BRCA1/2 mutations, are used to classify breast cancer into different molecular subtypes (39). Besides its molecular classification, breast cancer may be divided according to histological features, e.g. ductal carcinoma, lobular carcinoma, mucinous carcinoma or spindle cell carcinoma (40,41).

As mentioned above, the precise identification of molecular markers such as BRCA1/2 mutations is essential to predict the prognosis of patients. In this context, both BRCA1 and BRCA2 mutations are associated with a higher aggressiveness of both breast and ovarian cancer (42). However, due to the development of novel targeted therapies using high-effective PARP inhibitors (olaparib, niraparib, recuparib and veliparib) for the treatment of tumors with BRCA mutations, the presence of these mutations in ovarian cancer is associated with a favorable prognosis (43). Despite the prognostic importance of BRCA1/2 mutations, the identification of other proteins or genetic and epigenetic factors is essential to predict the efficacy of treatments and the survival of patients (44).

The BRCA1 gene is located on the long arm of chromosome 17 (17q21) and it is composed of 24 exons (45). BRCA1 encodes for a multi-functional protein of 1,863 amino acids, which consists of an amino (N)-terminal RING domain, a carboxyl (C)-terminus, also known as the BRCT domain, and coding regions of exons 11-13 (Fig. 2) (46,47). These domains have a crucial role in the interaction between BRCA1 and several partner proteins. The RING domain (amino acids 1-109) is a highly conserved domain encoded by exons 2-7, which is characterized by a RING finger motif involved in the ubiquitination pathway. Of note, it heterodimerizes with BRCA1 associated RING domain 1 (BARD1) to form a dimeric RING ubiquitin-ligase (E3) (48). The BRCT domain is encoded by exons 16-24 and spans from amino acids 1,650-1,863, including two tandem repeats (~100 amino acids) linked by 22 amino acids. This domain binds to the phosphorylated serine-proline-x-phenylalanine motifs of different partner proteins, such as BTB domain and CNC homolog 1 (BACH1), BRCA1 interacting helicase 1, BRCA1 A complex subunit and C-terminal binding protein 1, to form functional macromolecular complexes that allow selecting the substrate for BRCA1-BARD1 activity (49-52). Compared to other domains, exons 11-13 cover a large part of the BRCA1 protein. Of note, exon 11 comprises two nuclear localization sequences (NLS) (amino acids 501-507 and 607-614), which facilitate the nuclear import process of BRCA1 interacting with importin α (53).

As widely described in the literature, BRCA1 may be considered a tumor suppressor gene whose derived protein is involved in several molecular pathways both in the nucleus and cytoplasm, including DNA double-strand break (DSB) repair, cell-cycle checkpoints, genome stability, transcription regulation, apoptosis, chromosomal segregation, mitochondrial genome repair, cytoskeletal rearrangements and centrosome regulation (Fig. 2) (54-57).

It has been reported that DNA DSBs activate several kinases, such as ATM, ATM-related kinase, checkpoint kinase 1 (Chk1) and Chk2, which phosphorylate BRCA1 (58,59). The hyperphosphorylated BRCA1 then interacts with several protein complexes that repair DSBs via homologous recombination repair (HRR) and the activation of cell-cycle checkpoints. Of note, BRCA1 is involved in HRR through the interaction with the RAD50-MRE11-NBS1 complex, as well as partner and localizer of BRCA2 (PALB2) and RAD51 DNA repair proteins (60-62). Regarding cell-cycle checkpoints, other BRCA1 complexes have been described. Specifically, G2/M checkpoint signaling is activated by BRCA1-receptor-associated protein 80, while the BRCA1-BACH1 complex is required during the S-phase (63,64). BRCA1 also regulates gene expression at the transcriptional level, interacting with RNA polymerase II and several transcription factors, including c-Myc, p53, histone deacetylase 1 and 2, signal transducer and activator of transcription 1 and zinc finger and BRCA1-interacting protein with KRAB domain-1, as well as the SWItch/sucrose non-fermentable complex (65-70). Of note, the ubiquitin-ligase activity of the BRCA1-BARD1 complex has a critical role in centrosome regulation. In particular, BRCA1 and BARD1 bind to Obg-Like ATPase 1, favoring the maintenance of centrosome numbers at S- and G2/M-phases (71). Furthermore, BRCA1 has apoptotic properties due to its nuclear export and the activation of the p53-independent growth arrest and DNA damage-inducible 45 regulatory sequences (72).

BRCA2 was described for the first time in 1995 by Wooster et al (73) analyzing breast cancer families. The BRCA2 gene is located on the long arm of chromosome 13 (13q12.3) and it is composed of 27 exons that encode for a protein of 3,418 amino acids. BRCA2 consists of an N-terminal domain, a middle region and a C-terminal domain, which mediate its interaction with different partner proteins (Fig. 3). Although the structures of BRCA1 and BRCA2 exhibit certain similarities, there is no sequence homology (74,75). Of note, the N-terminal domain, encoded by exons 1-10, spans from amino acids 1-636 and contains a transcription activation domain (TAD) essential for PALB2 binding (21-39 amino acids) (76). The central segment is encoded by exon 11 (637-2280 amino acids) and covers the major portion of the BRCA2 protein. This evolutionarily conserved region is characterized by eight repeats of ~35-40 amino acids, also known as BRC repeats, which represent the primary interaction sites for the recombination enzyme RAD51 (77). Regarding the C-terminal domain (exons 12-27), it spans from amino acids 2,281-3,418 and includes the DNA-binding domain composed of a helical domain (amino acids 2,482-2,668) and three oligonucleotide-binding folds (amino acids 2,670-3,184). Furthermore, the C-terminus of BRCA2 also includes two NLS, as well as an additional RAD51-binding site (TR2) (amino acids 3,270-3,395) (78).

Similar to BRCA1, the BRCA2 gene is considered a caretaker of genome stability that has a key role in several biological pathways. Of note, it has been reported that BRCA2 interacts with different partner proteins for the formation of macromolecular complexes performing distinct cellular functions, such as DNA DSBs repair by HRR, DNA replication fork stabilization, transcription regulation and cell cycle checkpoint regulation (Fig. 3) (79-82).

For instance, the interaction between BRCA2 and RAD51 is implicated in the repair of DNA damage by the HRR pathway. Of note, BRC repeats, as well as the C-terminal domain of BRCA2, regulate the assembly of RAD51 into a nucleoprotein filament, promoting strand invasion and the search for homologous DNA (83,84). Other partner proteins interact with BRCA2 forming macromolecular complexes that have a crucial role in DNA DSB repair and maintenance of genome stability, including BARD1, PDS5 cohesin-associated factor B and SEM1 26S proteasome subunit (85-87). BRCA2 also promotes the stabilization of stalled DNA replication forks. Specifically, BRCA2 directly interacts with PALB2 through TAD of the N-terminal domain to sustain the recruitment of polymerase η at blocked replication forks (88). Furthermore, Fanconi anemia complementation group D2 and biorientation of chromosomes in cell division 1-like proteins have been described to interact with BRCA2 and promote stalled fork protection (89,90). Of note, BRCA2 may act as a transcriptional co-regulator forming a functional complex with mothers against decapentaplegic homolog 3 (SMAD3). BRCA2 induces SMAD3-dependent transcriptional activation of plasminogen activator inhibitor 1, while SMAD3 increases the transcriptional activity of BRCA2, indicating a synergistic activity of these two proteins (91). Furthermore, it has been reported that the binding between BRCA2 exon 3 and the nuclear protein EMSY (BRCA2-interacting transcriptional repressor) is involved in chromatin remodeling and transcription regulation (92). Finally, although the direct role of BRCA2 in cell-cycle checkpoints remains to be fully clarified, the interaction between BRCA2 and BRCA2-associated factor 35 may be responsible for G2/M checkpoint modulation (93).

On these bases, the present study aimed to provide an update of the current knowledge on BRCA1 and BRCA2 mutations and cancer susceptibility, focusing on their physiological functions, mutation frequency and clinical impact, as well as the available genetic tests and new potential technologies for the detection of BRCA1/2 mutations.

Materials and methods

Identification of the most common BRCA1/2 genetic variants and their clinical significance

The analysis of BRCA1 and BRCA2 mutations was performed using the Breast Cancer Information Core (BIC) (https://research.nhgri.nih.gov/bic/, accessed on 3 February 2022), BRCA Exchange (https://brcaexchange.org/, accessed on 15 February 2022) and ClinVar (https://www.ncbi.nlm.nih.gov/clinvar/, accessed on 7 February 2022) public databases. Specifically, a total of 15,311 BRCA1 and 14,914 BRCA2 mutations were registered in the BIC database. Among these, the top 10 most common BRCA1 and BRCA2 genetic variants were selected for evaluation of their clinical significance. In this regard, the BIC designation of BRCA1/2 mutations was matched with data reported in the BRCA Exchange and ClinVar databases in order to classify them into variants that are benign, likely benign, pathogenetic or of uncertain significance.

Identification of BRCA1 and BRCA2 founder mutations

A literature search of studies published from 1995 until November 2021 was conducted using the PubMed public database (https://pubmed.ncbi.nlm.nih.gov/) in order to investigate interethnic mutation frequencies. The key words 'breast cancer', 'ovarian cancer', 'BRCA1 mutations', 'BRCA2 mutations', 'BRCA founder mutations', 'BRCA germline mutations' or 'BRCA somatic mutations' were used to identify potentially relevant studies. In addition, the references contained in the most relevant studies were manually retrieved to find relevant articles not retrieved by PubMed exploration. Of note, the mutation frequency analysis was performed focusing on specific regions and ethnic groups, including Ashkenazi Jews, as well as populations from China, Denmark, Finland, France, Germany, Italy, Japan, Korea, Norway, Philippines, Poland, Russia and Sweden. In cases of overlapping data with the other published articles, the latest published and/or the larger sample size study was selected. Articles with ambiguous annotation of data, published in non-English or in Chinese language and duplicate publications were not included.

Results

BRCA1/2 mutation frequency and cancer susceptibility

Both endogenous and external DNA-damaging agents, including mutagens and radiation, as well as spontaneously occurring mutations, constantly threaten the integrity of the genome. Of note, DSBs represent the most damaging lesions of DNA, which may lead to chromosomal aberrations and mutations, increasing the risk of developing genetic disorders strictly related to cancer susceptibility (94). In this field, it has been widely reported that BRCA1 and BRCA2 tumor suppressor genes are involved in the repair of DNA DSBs by HRR. Specifically, BRCA1 promotes end resection and recruits PALB2, inducing chromatin localization of BRCA2. On the other hand, BRCA2 facilitates the recruitment of RAD51 recombinase, which inhibits the annealing of complementary ssDNA into the deleterious single strand (95). However, BRCA1 and BRCA2 mutations may increase the susceptibility to several tumor types, particularly breast and ovarian cancer (96,97). Indeed, BRCA1/2 loss of functions leads to genomic instability, which may result in the oncogenic transformation of normal cells into tumor-initiating cells (42).

BRCA1/2 mutations may be classified into germline and somatic mutations. Germline mutations are inherited in an autosomal dominant manner, while somatic mutations may arise de novo in tumor tissues due to a combination of genetic and environmental factors (98). Regarding BRCA1/2 germline mutations, loss of heterozygosity results in a non-functional protein that leads to Hereditary Breast and Ovarian Cancer (HBOC) syndrome, which is associated with increased susceptibility for these female tumors (99). Furthermore, inherited bi-allelic mutations of both BRCA1 and BRCA2 may cause congenital syndromes that are strictly associated with developmental anomalies, chromosomal fragility and increased cancer risk (100). In particular, the risk of breast cancer for females with a pathogenic BRCA1 or BRCA2 germline variant is 55-72% and 45-69%, respectively. Similarly, the risk for ovarian cancer is 39-44% for females with a BRCA1 germline variant and 11-17% for those with a BRCA2 germline variant (101). On the other hand, it has been reported that BRCA somatic mutations account for 15-30% of all BRCA1 and BRCA2 mutations. In addition, these non-inherited mutations are only present in 3% of all breast cancer cases (102).

Over the years, a large number of BRCA1/2 mutations have been described, and several of them are reliably known to increase cancer susceptibility (103). As reported in ClinVar, thousands of pathogenic or likely pathogenic BRCA1 and BRCA2 variants have been identified (>2,900 and >3,500, respectively) (https://www.ncbi.nlm.nih.gov/clinvar/, accessed 7 February 2022). The pathogenic or likely pathogenic mutations account for 80% of all mutations and result in a premature termination codon and truncated protein. Furthermore, missense mutations encoding a stable mutant protein account for 10% of all missense variants. Of note, frameshift mutations are more common in BRCA1, whereas missense mutations are more frequent in BRCA2 (104).

In the BIC database, a total of 15,311 BRCA1 mutations were registered, of which 6,133 are frameshift mutations, 4,577 are missense mutations and 1,421 are nonsense mutations. Regarding BRCA2, 14,914 mutations were registered (3,567 frameshift, 7,156 missense and 1,040 nonsense mutations). Of note, the mutation with the highest number of entries for BRCA1 was 185delAG, followed by 5382insC, whereas the most frequent mutation for BRCA2 was 6174delT (https://research.nhgri.nih.gov/bic/, accessed on 3 February 2022). Table I summarizes the most common BRCA1 and BRCA2 mutations registered in the BIC database.

Table I

Top 10 most common BRCA1/2 mutations according to the BIC database.

Table I

Top 10 most common BRCA1/2 mutations according to the BIC database.

BIC designationEntries, nGeneHGVS nucleotideHGVS proteinMutation
185delAG2038BRCA1c.68_69delp.Glu23fsFrameshift
5382insC1093BRCA1c.5266dupp.Gln1756fsFrameshift
4427T>C251BRCA1c.4308T>Cp.Ser1436=Synonymous
S1613G248BRCA1c.4837A>Gp.Ser1613GlyMissense
C61G239BRCA1c.181T>Gp.Cys61GlyMissense
2430T>C229BRCA1c.2311T>Cp.Leu771=Synonymous
2201C>T227BRCA1c.2082C>Tp.Ser694=Synonymous
IVS18+66G>A222BRCA1 c.5152+66G>A/Intervening sequence
IVS16°68A>G216BRCA1 c.4987-68A>G/Intervening sequence
IVS16°92A>G216BRCA1 c.4987-92A>G/Intervening sequence
6174delT1093BRCA2c.943T>Ap.Cys315SeFrameshift
H372N396BRCA2c.1114=p.Asn372=Missense
10°90A>C346BRCA2c.*105A>C/3′UTR
F599S345BRCA2c.1796=p.Ser599=Missense
IVS16-14T>C332BRCA2 c.7806-14T>C/Intervening sequence
IVS21-66T>C319BRCA2 c.8755-66T>C/Intervening sequence
K3326X301BRCA2c.9976A>Tp.Lys3326TerNonsense
I2490T240BRCA2c.7469T>Cp.Ile2490ThrMissense
3°24A>G234BRCA2c.3396A>Gp.Lys1132=Synonymous
IVS11+80delTTAA221BRCA2 c.6841+80_6841+83del/Intervening sequence

[i] HGVS, Human Genome Variation Society; BRCA1, BReast CAncer 1; BIC, Breast Cancer Information Core.

Furthermore, according to the BRCA Exchange and ClinVar database, numerous BRCA1/2 mutations among the 10 most frequent in the BIC database are classified as benign or likely benign variants. Specifically, 4427T>C, S1613G, 2430T>C, 2201C>T, IVS18+66G>A, IVS16°68A>G and IVS16°92A>G are benign or likely benign BRCA1 variants, while 185delAG, 5382insC and C61G are certainly pathogenetic. Regarding BRCA2, 10°90A>C, IVS16-14T>C, IVS21-66T>C, K3326X, I2490T, 3°24A>G and IVS11+80delTTAA are registered as benign variants, H372N and F599S are variants with uncertain significance, and 6174delT is a pathogenic variant (https://brcaexchange.org/, accessed on 15 February 2022; https://www.ncbi.nlm.nih.gov/clinvar/, accessed on 7 February 2022).

BRCA1/2 founder mutations among the worldwide population

In the last decades, an increasing number of studies have reported that certain BRCA1/2 mutations, also known as founder mutations, were more frequent in specific regions and ethnic groups. For instance, 185delAG, 5382insC and 6174delT mutations have been detected in Ashkenazi Jews, which accounted for 99% of the pathogenic variants identified in this population. Of note, the 5382insC founder mutation has also been identified in other countries, such as Poland and Russia, accounting for 94 and 60% of BRCA1 mutations, respectively. Other BRCA1 founder mutations have been observed in these countries, including C61G and 4154delA among the Polish population and G1706A in Russians (105-108).

Regarding the Northern European countries, the most common BRCA1 founder mutations detected among the Norwegian population were represented by 1675delA, 816delGT, 3347delAG and 1135insA (109-111). Furthermore, several BRCA1/2 mutations have been exclusively detected in individuals born in Finland, including IVS1°+3A>G and R1443X for BRCA1 and IVS23+1G>A, 7708C>T and T8555G for BRCA2 (112). Of note, 2594delC, E1107X and G1706A represented the most common BRCA1 mutations in the Danish population, while 3171ins5 and 6601delA were detected among the Swedish (113,114).

As for Central European regions, it has been reported that 3600del11 and G1710X were the most frequently detected BRCA1 mutations in the French population (115). On the other hand, numerous founder mutations were detected among Germans, such as G1706A, C61G, 2804delAA and IVS12-1643del3835 for BRCA1 and 5579insA and 6503delTT for BRCA2 (116). Of note, the 5083del19 mutation exhibited a high rate in Italian families from Calabria, while 8765delAG was observed in certain Sardinian families (117-119).

Finally, different BRCA1/2 founder mutations have been also discovered in Asian populations. Of note, L63X, Q934X and 5802delAATT have been identified in Japanese, while the most common mutations among Koreans were 1041del3insT for BRCA1 and 7708C>T for BRCA2 (120-123). In addition, 1100delAT, 3337C>T and 9325insA were more frequently detected among Chinese, whereas 5454delC and 4859delA were the most common BRCA1/2 founder mutations in the Philippine population (124,125). The distribution of BRCA1/2 founder mutations is summarized in Tables II and III.

Table II

BRCA1 founder mutations compared among different countries.

Table II

BRCA1 founder mutations compared among different countries.

BIC DesignationHGVS nucleotideMutationPopulation(Refs.)
185delAGc.68_69delFrameshiftAshkenazi Jewish(106,107)
5382insCc.5266dupFrameshiftAshkenazi Jewish(106,107)
1100delATc.981_982delFrameshiftChina(125)
2594delC c.9613_9614delinsCTFrameshiftDenmark(114)
E1107Xc.3319G>TNonsenseDenmark(114)
G1706Ac.5117G>CMissmatchDenmark(114)
IVS1°+3A>Gc.409°+3A>GSplicing variantFinland(113)
R1443Xc.4327C>TNonsenseFinland(113)
3600del11c.3481_3491delFrameshiftFrance(116)
G1710Xc.5128G>TNonsenseFrance(116)
G1706Ac.5117G>CMissenseGermany(117)
C61Gc.181T>GMissenseGermany(117)
2804delAAc.2685_2686delFrameshiftGermany(117)
IVS12-1643del3835 c.4186-1643_4357+2020delIntervening sequenceGermany(117)
5083del19c.4964_4982delFrameshiftItaly(118-120)
L63Xc.188T>ANonsenseJapan(121,122)
Q934Xc.2800C>TNonsenseJapan(121,122)
1041del3insT c.922_924delinsTNonsenseKorea(123,124)
1675delAc.1556delFrameshiftNorway(110-112)
816delGTc.697_698delFrameshiftNorway(110-112)
3347delAGc.3228_3229delFrameshiftNorway(110-112)
1135insAc.1016dupFrameshiftNorway(110-112)
5454delCc.5335delFrameshiftPhilippines(126)
C61Gc.181T>GMissensePoland(108,109)
4154delAc.4035delFrameshiftPoland(108,109)
5382insCc.5266dupFrameshiftPoland(108,109)
G1706Ac.5117G>CMissenseRussia(106,107)
5382insCc.5266dupFrameshiftRussia(106,107)
3171ins5c.3048_3052dupFrameshiftSweden(115)

[i] HGVS, Human Genome Variation Society; BRCA1, BReast CAncer 1; BIC, Breast Cancer Information Core.

Table III

BRCA2 founder mutations compared among different countries.

Table III

BRCA2 founder mutations compared among different countries.

BIC designationHGVS nucleotideMutationPopulation(Refs.)
6174delTc.5946delc.5946delAshkenazi Jewish(106,107)
3337C>Tc.3109C>Tc.3109C>TChina(125)
9325insAc.9097dupc.9097dupChina(125)
IVS23+1G>Ac.9117+1G>AIntervening sequenceFinland(113)
7708C>Tc.7480C>TNonsenseFinland(113)
T8555Gc.8327T>GNonsenseFinland(113)
5579insAc.5351dupFrameshiftGermany(117)
6503delTTc.6275_6276delFrameshiftGermany(117)
8765delAGc.8537_8538delFrameshiftItaly(118-120)
5802delAATT/FrameshiftJapan(121,122)
7708C>Tc.7480C>TNonsenseKorea(123,124)
4859delAc.4631delFrameshiftPhilippines(126)
6601delAc.6373delFrameshiftSweden(115)

[i] HGVS, Human Genome Variation Society; BRCA2, BReast CAncer 2; BIC, Breast Cancer Information Core.

Discussion

To date, thousands of mutations have been identified in the BRCA1 and BRCA2 genes. Most of the described mutations are caused by small insertions or deletions, large genomic rearrangements (LGRs), as well as nonsense mutations and splice variants (126). As previously described, certain mutations have been more frequently detected in specific geographical areas and ethnicities (127). Furthermore, it has been reported that each family group may carry a specific mutation that may be considered unique (128). Functional deficiencies due to these pathogenic mutations increase the lifetime risk to develop both breast and ovarian cancer. Of note, the most common cause of these female tumors is HBOC syndrome. Females harboring BRCA1/2 germline mutations have a higher risk of cancer development compared to other subjects (129,130). Therefore, the usage of reliable genetic tests for the assessment of BRCA1/2 mutations in high-risk females is crucial for the clinical management of patients.

Over the years, different platforms have been developed for the detection of BRCA mutations (Fig. 4). Among these, the Sanger method has been widely employed for genomic DNA sequencing. In brief, this method allows division of large genomic DNA into small fragments that are sequenced separately (131,132). Although the Sanger sequencing method is considered a reliable technology with a relatively simple workflow, it has several limits, such as sequencing throughput (a single DNA fragment at a time), time-consuming analysis and low cost-effectiveness (133).

Currently, the gold standard genetic test for the identification of BRCA1/2 mutations is next-generation sequencing (NGS). NGS is a high-throughput technology based on synthesis by sequencing millions of DNA fragments at once. Compared to Sanger sequencing, NGS has numerous other advantages, including automated analysis, faster turnaround time and higher sensitivity to detect LGRs and low-frequency variants with deep sequencing (134,135). However, NGS technology may also have certain disadvantages, such as complex workflow, high rates of variants with uncertain significance, higher cost and the reduction of sensitivity for large insertions/deletions (>20 bp) (136,137). The currently used standard protocol for the identification of BRCA1 and BRCA2 mutations includes comprehensive sequencing and the assessment of LGRs. Furthermore, a single-site target may also be analyzed for patients with a first-degree relative affected by BRCA1/2 mutation (138,139).

The high cost of these technologies has prompted researchers and clinicians to develop novel high-sensitivity and low-cost strategies for the detection of BRCA1/2 mutations. Among these, reliable results were obtained by using the droplet digital PCR (ddPCR) platform.

ddPCR has recently emerged as a reliable tool with high sensitivity and specificity. In brief, the ddPCR system is based on a water-oil emulsion of the reaction mixture, which consists of a DNA sample, ddPCR Mastermix, primers and probe in a final volume of 20 µl (140). This procedure allows division of the sample into ~20,000 droplets that are transferred into a 96-well PCR plate for amplification. Since the sample is fractioned into thousands of droplets, PCR amplification takes place in each droplet. A droplet reader then detects the positive/negative signal of droplets depending on their amplified target and fluorescence amplitude (141).

Of note, ddPCR represents a valuable alternative to standard methods for the analysis of different clinical samples, such as fresh tumor biopsies and formalin-fixed paraffin-embedded (FFPE) tissues, as well as liquid biopsy samples (peripheral blood, sputum, urine, cerebrospinal fluid, stool, pleural effusions and ascites fluid), overcoming the limits due to poor DNA quality (142). Over the years, the potential clinical application of ddPCR has been demonstrated for absolute allele quantification, viral load quantification, DNA methylation, DNA copy number variation, germline/somatic mutation detection, circulating mutation detection, analysis of microRNAs, long non-coding RNAs and gene rearrangements (143-147). In this field, a growing number of recent studies have focused on the ddPCR system as a promising tool for the identification of specific BRCA1/2 mutations.

Preobrazhenskaya et al (148) developed a ddPCR assay to investigate BRCA1 LGRs in blood-derived DNA samples of patients with breast cancer (n=141). Using ribonuclease P RNA component H1 as a reference gene and PCR primers covering the entire coding region of BRCA1, the researchers identified three cases with exon 8 deletion and one case with exons 5-7 deletion (148). Furthermore, they identified a total of four cases with exon 8 deletion in an additional cohort of patients with breast and ovarian cancer (720 and 184, respectively). Collectively, the ddPCR data exhibited high concordance with multiplex ligation-dependent probe amplification (MLPA), a standard method for the detection of LGRs, suggesting that ddPCR may be a valuable tool for the assessment of BRCA1 LGRs (149). Similarly, another study evaluated the sensitivity and specificity of ddPCR to detect BRCA1 rearrangements in patients with serous ovarian cancer (149). In brief, the authors performed a multiplex ddPCR assay dividing primers/probe for BRCA1 exons into 8 groups along with albumin or ribonuclease P/MRP subunit P30 as a reference gene and identified nine cases with different BRCA1 deletions (100% concordance with MLPA) (149). Of note, Khalique et al (150) provided a case report of high-grade serous ovarian cancer (HGSOC) with secondary somatic mutation of BRCA2 (c.5446_5449delCTAG, p.Ser1816Leu fs*23). Specifically, the study highlighted that the ddPCR assay detected a higher mutation frequency (9.4%) in FFPE tissue samples compared to the whole-exome sequencing genetic test (2.3%) (150). Recently, De Paolis et al (151) further investigated the specificity of ddPCR in a cohort of patients with HGSOC positive for BRCA1 LGRs (exons 2, 20 and 21). By using ribosomal protein lateral stalk subunit P0 as a reference gene and specific primers/probe, the ddPCR assay confirmed the detection of BRCA1 LGRs in blood, FFPE and fresh frozen tissue samples with 100% specificity (151).

Although no studies have been performed on the use of ddPCR for the detection of specific BRCA mutations in relatives of BRCA-positive cancer patients, the use of ddPCR for the detection of already known mutations in familial clusters would reduce the higher costs related to the whole sequencing of both BRCA1 and BRCA2 genes. Indeed, at present, the cost of a single test for the detection of BRCA mutations is 172$, while Garcia and colleagues described that the analysis of gene mutations performed by ddPCR cost half as much as NGS or other methods (152,153).

Besides the cost-effectiveness, ddPCR may also be used for the detection of circulating BRCA1/2 mutations in liquid biopsy samples. Indeed, ddPCR was effectively used for the detection of circulating mutations in patients with ovarian cancer, demonstrating how the detection of both germline and somatic ctDNA mutations is a valuable complementary tool for the diagnosis of BRCA-positive tumors and to establish the effectiveness of PARP inhibitors and the prognosis of patients with ovarian cancer (154).

All of these data suggest that ddPCR may be used for the detection of already known BRCA1/2 mutations in familial clusters of patients with a diagnosis of hereditary breast or ovarian cancer.

In conclusion, BRCA1 and BRCA2 germline mutations are well-established risk factors for the onset of breast or ovarian cancer. Overall, the computational investigations performed in the present study suggested that the BRCA1/2 mutations with the highest number of entries are classified as certainly pathogenic variants, such as 185delAG, 5382insC and C61G for BRCA1 and 6174delT for BRCA2. Furthermore, several of these mutations, also known as founder mutations, appear to be more frequently detected in specific geographical regions and ethnic groups. In this context, the development of low-cost and reliable genetic tests is fundamental to improve the screening program for the identification of females with germline BRCA1/2 mutations and the management of relatives of patients with a known diagnosis of hereditary breast or ovarian cancer. Currently, NGS is the gold standard for the assessment of BRCA1/2 mutations. In this field, ddPCR has recently emerged as a highly sensitive and specific technique. According to the studies described in the present article, ddPCR may be considered a promising alternative strategy for the detection of BRCA1/2 pathogenetic mutations. However, further studies should be performed to validate the application of ddPCR in routine clinical practice and surveillance strategies for hereditary female tumors.

Availability of data and materials

The data reported in the manuscript are available from the corresponding author on request. The original data examined in the study are publicly available and may be retrieved from the following sources: www.pubmed.com; https://www.ncbi.nlm.nih.gov/clinvar/; https://research.nhgri.nih.gov/bic/; and https://brcaexchange.org/.

Authors' contributions

ML and AS conceptualized the study. AL, RR, GG and SC wrote the original draft of the manuscript. ML, AS, GNZ, GC and LF provided critical revisions. AL and RR prepared the tables and figures, conducted the formal analysis and critically analyzed the literature. All authors contributed to manuscript revision and read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Acknowledgments

Not applicable.

Funding

This work was supported by the Italian League Against Cancer (LILT) - Grant Ricerca Sanitaria 2018 LILT.

References

1 

Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A and Bray F: Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 Cancers in 185 countries. CA Cancer J Clin. 71:209–249. 2021. View Article : Google Scholar : PubMed/NCBI

2 

Łukasiewicz S, Czeczelewski M, Forma A, Baj J, Sitarz R and Stanisławek A: Breast cancer-epidemiology, risk factors, classification, prognostic markers, and current treatment strategies-an updated review. Cancers (Basel). 13:42872021. View Article : Google Scholar

3 

Falzone L, Scandurra G, Lombardo V, Gattuso G, Lavoro A, Distefano AB, Scibilia G and Scollo P: A multidisciplinary approach remains the best strategy to improve and strengthen the management of ovarian cancer (Review). Int J Oncol. 59:532021. View Article : Google Scholar : PubMed/NCBI

4 

Winters S, Martin C, Murphy D and Shokar NK: Breast cancer epidemiology, prevention, and screening. Prog Mol Biol Transl Sci. 151:1–32. 2017. View Article : Google Scholar : PubMed/NCBI

5 

La Vecchia C: Ovarian cancer: Epidemiology and risk factors. Eur J Cancer Prev. 26:55–62. 2017. View Article : Google Scholar

6 

D'Alonzo M, Bounous VE, Villa M and Biglia N: Current evidence of the oncological benefit-risk profile of hormone replacement therapy. Medicina (Kaunas). 55:5732019. View Article : Google Scholar

7 

Prentice RL, Aragaki AK, Chlebowski RT, Rossouw JE, Anderson GL, Stefanick ML, Wactawski-Wende J, Kuller LH, Wallace R, Johnson KC, et al: Randomized trial evaluation of the benefits and risks of menopausal hormone therapy among women 50-59 years of age. Am J Epidemiol. 190:365–375. 2021. View Article : Google Scholar

8 

Beaber EF, Malone KE, Tang MT, Barlow WE, Porter PL, Daling JR and Li CI: Oral contraceptives and breast cancer risk overall and by molecular subtype among young women. Cancer Epidemiol Biomarkers Prev. 23:755–764. 2014. View Article : Google Scholar : PubMed/NCBI

9 

Havrilesky LJ, Moorman PG, Lowery WJ, Gierisch JM, Coeytaux RR, Urrutia RP, Dinan M, McBroom AJ, Hasselblad V, Sanders GD and Myers ER: Oral contraceptive pills as primary prevention for ovarian cancer: A systematic review and meta-analysis. Obstet Gynecol. 112:139–147. 2013. View Article : Google Scholar

10 

Benfatto G, Zanghì G, Catalano F, Di Stefano G, Fancello R, Mugavero F and Giovanetto A: Day surgery for breast cancer in the elderly. G Chir. 27:49–52. 2006.In Italian. PubMed/NCBI

11 

Shoemaker ML, White MC, Wu M, Weir HK and Romieu I: Differences in breast cancer incidence among young women aged 20-49 years by stage and tumor characteristics, age, race, and ethnicity, 2004-2013. Breast Cancer Res Treat. 169:595–606. 2018. View Article : Google Scholar : PubMed/NCBI

12 

Sarink D, Wu AH, Le Marchand L, White KK, Park SY, Setiawan VW, Hernandez BY, Wilkens LR and Merritt MA: Racial/ethnic differences in ovarian cancer risk: Results from the multiethnic cohort study. Cancer Epidemiol Biomarkers Prev. 29:2019–2025. 2020. View Article : Google Scholar : PubMed/NCBI

13 

Zeinomar N, Knight JA, Genkinger JM, Phillips KA, Daly MB, Milne RL, Dite GS, Kehm RD, Liao Y, Southey MC, et al: Alcohol consumption, cigarette smoking, and familial breast cancer risk: Findings from the prospective family study cohort (ProF-SC). Breast Cancer Res. 21:1282019. View Article : Google Scholar : PubMed/NCBI

14 

Friedenreich CM, Ryder-Burbidge C and McNeil J: Physical activity, obesity and sedentary behavior in cancer etiology: Epidemiologic evidence and biologic mechanisms. Mol Oncol. 15:790–800. 2021. View Article : Google Scholar :

15 

Dunneram Y, Greenwood DC and Cade JE: Diet, menopause and the risk of ovarian, endometrial and breast cancer. Proc Nutr Soc. 78:438–448. 2019. View Article : Google Scholar : PubMed/NCBI

16 

Bryere J, Dejardin O, Launay L, Colonna M, Grosclaude P and Launoy G; French Network of Cancer Registries (FRANCIM): Socioeconomic status and site-specific cancer incidence, a Bayesian approach in a French cancer registries network study. Eur J Cancer Prev. 27:391–398. 2018. View Article : Google Scholar

17 

Falzone L, Grimaldi M, Celentano E, Augustin LSA and Libra M: Identification of modulated MicroRNAs associated with breast cancer, diet, and physical activity. Cancers (Basel). 12:25552020. View Article : Google Scholar

18 

Park HL: Epigenetic biomarkers for environmental exposures and personalized breast cancer prevention. Int J Environ Res Public Health. 17:11812020. View Article : Google Scholar :

19 

Singh A, Gupta S and Sachan M: Epigenetic biomarkers in the management of ovarian cancer: Current prospectives. Front Cell Dev Biol. 7:1822019. View Article : Google Scholar : PubMed/NCBI

20 

Huang M, Xiao J, Nasca PC, Liu C, Lu Y, Lawrence WR, Wang L, Chen Q and Lin S: Do multiple environmental factors impact four cancers in women in the contiguous United States? Environ Res. 179(PtA): 1087822019. View Article : Google Scholar : PubMed/NCBI

21 

Brewer HR, Jones ME, Schoemaker MJ, Ashworth A and Swerdlow AJ: Family history and risk of breast cancer: An analysis accounting for family structure. Breast Cancer Res Treat. 165:193–200. 2017. View Article : Google Scholar : PubMed/NCBI

22 

Flaum N, Crosbie EJ, Edmondson RJ, Smith MJ and Evans DG: Epithelial ovarian cancer risk: A review of the current genetic landscape. Clin Genet. 97:54–63. 2020. View Article : Google Scholar :

23 

Bethea TN, Ochs-Balcom HM, Bandera EV, Beeghly-Fadiel A, Camacho F, Chyn D, Cloyd EK, Harris HR, Joslin CE, Myers E, et al: First- and second-degree family history of ovarian and breast cancer in relation to risk of invasive ovarian cancer in African American and white women. Int J Cancer. 148:2964–2973. 2021. View Article : Google Scholar : PubMed/NCBI

24 

Liu L, Hao X, Song Z, Zhi X, Zhang S and Zhang J: Correlation between family history and characteristics of breast cancer. Sci Rep. 11:63602021. View Article : Google Scholar : PubMed/NCBI

25 

Welcsh PL and King MC: BRCA1 and BRCA2 and the genetics of breast and ovarian cancer. Hum Mol Genet. 10:705–713. 2001. View Article : Google Scholar : PubMed/NCBI

26 

Yoshida K and Miki Y: Role of BRCA1 and BRCA2 as regulators of DNA repair, transcription, and cell cycle in response to DNA damage. Cancer Sci. 95:866–871. 2004. View Article : Google Scholar : PubMed/NCBI

27 

Paul A and Paul S: The breast cancer susceptibility genes (BRCA) in breast and ovarian cancers. Front Biosci (Landmark Ed). 19:605–618. 2014. View Article : Google Scholar

28 

Venkitaraman AR: How do mutations affecting the breast cancer genes BRCA1 and BRCA2 cause cancer susceptibility? DNA Repair (Amst). 81:1026682019. View Article : Google Scholar

29 

Paalosalo-Harris K and Skirton H: Mixed method systematic review: The relationship between breast cancer risk perception and health-protective behaviour in women with family history of breast cancer. J Adv Nurs. 73:760–764. 2017. View Article : Google Scholar

30 

Hanley GE, McAlpine JN, Miller D, Huntsman D, Schrader KA, Gilks CB and Mitchell G: A population-based analysis of germline BRCA1 and BRCA2 testing among ovarian cancer patients in an era of histotype-specific approaches to ovarian cancer prevention. BMC Cancer. 18:2542018. View Article : Google Scholar : PubMed/NCBI

31 

Moschetta M, George A, Kaye SB and Banerjee S: BRCA somatic mutations and epigenetic BRCA modifications in serous ovarian cancer. Ann Oncol. 27:1449–1455. 2016. View Article : Google Scholar : PubMed/NCBI

32 

Kuchenbaecker KB, Hopper JL, Barnes DR, Phillips KA, Mooij TM, Roos-Blom MJ, Jervis S, van Leeuwen FE, Milne RL, Andrieu N, et al: Risks of breast, ovarian, and contralateral breast cancer for BRCA1 and BRCA2 mutation carriers. JAMA. 317:2402–2416. 2017. View Article : Google Scholar : PubMed/NCBI

33 

Neff RT, Senter L and Salani R: BRCA mutation in ovarian cancer: Testing, implications and treatment considerations. Ther Adv Med Oncol. 9:519–531. 2017. View Article : Google Scholar : PubMed/NCBI

34 

Kotsopoulos J: BRCA mutations and breast cancer prevention. Cancers (Basel). 10:5242018. View Article : Google Scholar

35 

Cortesi L, Piombino C and Toss A: Germline mutations in other homologous recombination repair-related genes than BRCA1/2: Predictive or prognostic factors? J Pers Med. 11:2452021. View Article : Google Scholar :

36 

Zhao W, Hu H, Mo Q, Guan Y, Li Y, Du Y and Li L: Function and mechanism of combined PARP-1 and BRCA genes in regulating the radiosensitivity of breast cancer cells. Int J Clin Exp Pathol. 12:3915–3920. 2019.

37 

Liu X, Wu K, Zheng D, Luo C, Fan Y, Zhong X and Zheng H: Efficacy and safety of PARP inhibitors in advanced or metastatic triple-negative breast cancer: A systematic review and meta-analysis. Front Oncol. 11:7421392021. View Article : Google Scholar : PubMed/NCBI

38 

Dickson KA, Xie T, Evenhuis C, Ma Y and Marsh DJ: PARP inhibitors display differential efficacy in models of BRCA mutant high-grade serous ovarian cancer. Int J Mol Sci. 22:85062021. View Article : Google Scholar :

39 

Al-Thoubaity FK: Molecular classification of breast cancer: A retrospective cohort study. Ann Med Surg (Lond). 49:44–48. 2019. View Article : Google Scholar

40 

Makki J: Diversity of breast carcinoma: Histological subtypes and clinical relevance. Clin Med Insights Pathol. 8:23–31. 2015. View Article : Google Scholar

41 

Magro G, Salvatorelli L, Puzzo L, Piombino E, Bartoloni G, Broggi G and Vecchio GM: Practical approach to diagnosis of bland-looking spindle cell lesions of the breast. Pathologica. 111:344–360. 2019. View Article : Google Scholar

42 

Gorodetska I, Kozeretska I and Dubrovska A: BRCA genes: The role in genome stability, cancer stemness and therapy resistance. J Cancer. 10:2109–2127. 2019. View Article : Google Scholar :

43 

Tung NM and Garber JE: BRCA1/2 testing: Therapeutic implications for breast cancer management. Br J Cancer. 119:141–152. 2018. View Article : Google Scholar : PubMed/NCBI

44 

Broggi G, Filetti V, Ieni A, Rapisarda V, Ledda C, Vitale E, Varricchio S, Russo D, Lombardo C, Tuccari G, et al: MacroH2A1 immunoexpression in breast cancer. Front Oncol. 10:15192020. View Article : Google Scholar : PubMed/NCBI

45 

Miki Y, Swensen J, Shattuck-Eidens D, Futreal PA, Harshman K, Tavtigian S, Liu Q, Cochran C, Bennett LM and Ding W: A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science. 266:66–71. 1994. View Article : Google Scholar : PubMed/NCBI

46 

Nelson AC and Holt JT: Impact of RING and BRCT domain mutations on BRCA1 protein stability, localization and recruitment to DNA damage. Radiat Res. 174:1–13. 2010. View Article : Google Scholar : PubMed/NCBI

47 

Christou CM and Kyriacou K: BRCA1 and its network of interacting partners. Biology (Basel). 2:40–63. 2013.

48 

Xia Y, Pao GM, Chen HW, Verma IM and Hunter T: Enhancement of BRCA1 E3 ubiquitin ligase activity through direct interaction with the BARD1 protein. J Biol Chem. 278:5255–5263. 2003. View Article : Google Scholar

49 

Manke IA, Lowery DM, Nguyen A and Yaffe MB: BRCT repeats as phosphopeptide-binding modules involved in protein targeting. Science. 302:636–639. 2003. View Article : Google Scholar : PubMed/NCBI

50 

Clapperton JA, Manke IA, Lowery DM, Ho T, Haire LF, Yaffe MB and Smerdon SJ: Structure and mechanism of BRCA1 BRCT domain recognition of phosphorylated BACH1 with implications for cancer. Nat Struct Mol Biol. 11:512–518. 2004. View Article : Google Scholar : PubMed/NCBI

51 

Thanassoulas A, Nomikos M, Theodoridou M, Yannoukakos D, Mastellos D and Nounesis G: Thermodynamic study of the BRCT domain of BARD1 and its interaction with the -pSER-X-X-Phemotif-containing BRIP1 peptide. Biochim Biophys Acta. 1804:1908–1916. 2010. View Article : Google Scholar : PubMed/NCBI

52 

Wang B, Matsuoka S, Ballif BA, Zhang D, Smogorzewska A, Gygi SP and Elledge SJ: Abraxas and RAP80 form a BRCA1 protein complex required for the DNA damage response. Science. 316:1194–1198. 2007. View Article : Google Scholar : PubMed/NCBI

53 

Chen CF, Li S, Chen Y, Chen PL, Sharp ZD and Lee WH: The nuclear localization sequences of the BRCA1 protein interact with the importin-alpha subunit of the nuclear transport signal receptor. J Biol Chem. 271:32863–32868. 1996. View Article : Google Scholar : PubMed/NCBI

54 

Caestecker KW and Van de Walle GR: The role of BRCA1 in DNA double-strand repair: past and present. Exp Cell Res. 319:575–587. 2013. View Article : Google Scholar

55 

Savage KI and Harkin DP: BRCA1, a 'complex' protein involved in the maintenance of genomic stability. FEBS J. 282:630–646. 2015. View Article : Google Scholar

56 

Sharma B, Kaur RP, Raut S and Munshi A: BRCA1 mutation spectrum, functions, and therapeutic strategies: The story so far. Curr Probl Cancer. 42:189–207. 2018. View Article : Google Scholar : PubMed/NCBI

57 

Takaoka M and Miki Y: BRCA1 gene: Function and deficiency. Int J Clin Oncol. 23:36–44. 2018. View Article : Google Scholar

58 

Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER III, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y, et al: ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science. 316:1160–1166. 2007. View Article : Google Scholar : PubMed/NCBI

59 

Smith J, Tho LM, Xu N and Gillespie DA: The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer. Adv Cancer Res. 108:73–112. 2010. View Article : Google Scholar : PubMed/NCBI

60 

Syed A and Tainer JA: The MRE11-RAD50-NBS1 complex conducts the orchestration of damage signaling and outcomes to stress in DNA replication and repair. Annu Rev Biochem. 87:263–294. 2018. View Article : Google Scholar : PubMed/NCBI

61 

Simhadri S, Vincelli G, Huo Y, Misenko S, Foo TK, Ahlskog J, Sørensen CS, Oakley GG, Ganesan S, Bunting SF and Xia B: PALB2 connects BRCA1 and BRCA2 in the G2/M checkpoint response. Oncogene. 38:1585–1596. 2019. View Article : Google Scholar :

62 

Zhao W, Steinfeld JB, Liang F, Chen X, Maranon DG, Ma CJ, Kwon Y, Rao T, Wang W, Sheng C, et al: BRCA1-BARD1 promotes RAD51-mediated homologous DNA pairing. Nature. 550:360–365. 2017. View Article : Google Scholar : PubMed/NCBI

63 

Coleman KA and Greenberg RA: The BRCA1-RAP80 complex regulates DNA repair mechanism utilization by restricting end resection. J Biol Chem. 286:13669–13680. 2011. View Article : Google Scholar : PubMed/NCBI

64 

Kumaraswamy E and Shiekhattar R: Activation of BRCA1/BRCA2-associated helicase BACH1 is required for timely progression through S phase. Mol Cell Biol. 27:6733–6741. 2007. View Article : Google Scholar : PubMed/NCBI

65 

Wang Q, Zhang H, Kajino K and Greene MI: BRCA1 binds c-Myc and inhibits its transcriptional and transforming activity in cells. Oncogene. 17:1939–1948. 1998. View Article : Google Scholar : PubMed/NCBI

66 

Chai YL, Cui J, Shao N, Shyam E, Reddy P and Rao VN: The second BRCT domain of BRCA1 proteins interacts with p53 and stimulates transcription from the p21WAF1/CIP1 promoter. Oncogene. 18:263–268. 1999. View Article : Google Scholar : PubMed/NCBI

67 

Thurn KT, Thomas S, Raha P, Qureshi I and Munster PN: Histone deacetylase regulation of ATM-mediated DNA damage signaling. Mol Cancer Ther. 12:2078–2087. 2013. View Article : Google Scholar : PubMed/NCBI

68 

Buckley NE, Hosey AM, Gorski JJ, Purcell JW, Mulligan JM, Harkin DP and Mullan PB: BRCA1 regulates IFN-gamma signaling through a mechanism involving the type I IFNs. Mol Cancer Res. 5:261–270. 2007. View Article : Google Scholar : PubMed/NCBI

69 

Tan W, Zheng L, Lee WH and Boyer TG: Functional dissection of transcription factor ZBRK1 reveals zinc fingers with dual roles in DNA-binding and BRCA1-dependent transcriptional repression. J Biol Chem. 279:6576–6587. 2004. View Article : Google Scholar

70 

Harte MT, O'Brien GJ, Ryan NM, Gorski JJ, Savage KI, Crawford NT, Mullan PB and Harkin DP: BRD7, a subunit of SWI/SNF complexes, binds directly to BRCA1 and regulates BRCA1-dependent transcription. Cancer Res. 70:2538–2547. 2010. View Article : Google Scholar : PubMed/NCBI

71 

Yoshino Y, Qi H, Fujita H, Shirota M, Abe S, Komiyama Y, Shindo K, Nakayama M, Matsuzawa A, Kobayashi A, et al: BRCA1-interacting protein OLA1 requires interaction with BARD1 to regulate centrosome number. Mol Cancer Res. 16:1499–1511. 2018. View Article : Google Scholar : PubMed/NCBI

72 

Harkin DP, Bean JM, Miklos D, Song YH, Truong VB, Englert C, Christians FC, Ellisen LW, Maheswaran S, Oliner JD and Haber DA: Induction of GADD45 and JNK/SAPK-dependent apoptosis following inducible expression of BRCA1. Cell. 97:575–586. 1999. View Article : Google Scholar : PubMed/NCBI

73 

Wooster R, Bignell G, Lancaster J, Swift S, Seal S, Mangion J, Collins N, Gregory S, Gumbs C and Micklem G: Identification of the breast cancer susceptibility gene BRCA2. Nature. 378:789–792. 1995. View Article : Google Scholar : PubMed/NCBI

74 

Chen J, Silver DP, Walpita D, Cantor SB, Gazdar AF, Tomlinson G, Couch FJ, Weber BL, Ashley T, Livingston DM and Scully R: Stable interaction between the products of the BRCA1 and BRCA2 tumor suppressor genes in mitotic and meiotic cells. Mol Cell. 2:317–328. 1998. View Article : Google Scholar : PubMed/NCBI

75 

Sharan SK and Bradley A: Functional characterization of BRCA1 and BRCA2: Clues from their interacting proteins. J Mammary Gland Biol Neoplasia. 3:413–421. 1998. View Article : Google Scholar

76 

Oliver AW, Swift S, Lord CJ, Ashworth A and Pearl LH: Structural basis for recruitment of BRCA2 by PALB2. EMBO Rep. 10:990–996. 2009. View Article : Google Scholar : PubMed/NCBI

77 

Carreira A, Hilario J, Amitani I, Baskin RJ, Shivji MK, Venkitaraman AR and Kowalczykowski SC: The BRC repeats of BRCA2 modulate the DNA-binding selectivity of RAD51. Cell. 136:1032–1043. 2009. View Article : Google Scholar : PubMed/NCBI

78 

Davies OR and Pellegrini L: Interaction with the BRCA2 C terminus protects RAD51-DNA filaments from disassembly by BRC repeats. Nat Struct Mol Biol. 14:475–483. 2007. View Article : Google Scholar : PubMed/NCBI

79 

Moynahan ME, Pierce AJ and Jasin M: BRCA2 is required for homology-directed repair of chromosomal breaks. Mol Cell. 7:263–272. 2001. View Article : Google Scholar : PubMed/NCBI

80 

Roy R, Chun J and Powell SN: BRCA1 and BRCA2: Different roles in a common pathway of genome protection. Nat Rev Cancer. 12:68–78. 2011. View Article : Google Scholar : PubMed/NCBI

81 

Yuan SS, Lee SY, Chen G, Song M, Tomlinson GE and Lee EY: BRCA2 is required for ionizing radiation-induced assembly of Rad51 complex in vivo. Cancer Res. 59:3547–3551. 1999.PubMed/NCBI

82 

Milner J, Ponder B, Hughes-Davies L, Seltmann M and Kouzarides T: Transcriptional activation functions in BRCA2. Nature. 386:772–773. 1997. View Article : Google Scholar : PubMed/NCBI

83 

Davies AA, Masson JY, McIlwraith MJ, Stasiak AZ, Stasiak A, Venkitaraman AR and West SC: Role of BRCA2 in control of the RAD51 recombination and DNA repair protein. Mol Cell. 7:273–282. 2001. View Article : Google Scholar : PubMed/NCBI

84 

Esashi F, Galkin VE, Yu X, Egelman EH and West SC: Stabilization of RAD51 nucleoprotein filaments by the C-terminal region of BRCA2. Nat Struct Mol Biol. 14:468–474. 2007. View Article : Google Scholar : PubMed/NCBI

85 

Henderson BR: Regulation of BRCA1, BRCA2 and BARD1 intracellular trafficking. Bioessays. 27:884–893. 2005. View Article : Google Scholar : PubMed/NCBI

86 

Couturier AM, Fleury H, Patenaude AM, Bentley VL, Rodrigue A, Coulombe Y, Niraj J, Pauty J, Berman JN, Dellaire G, et al: Roles for APRIN (PDS5B) in homologous recombination and in ovarian cancer prediction. Nucleic Acids Res. 44:10879–10897. 2016. View Article : Google Scholar : PubMed/NCBI

87 

Yang H, Jeffrey PD, Miller J, Kinnucan E, Sun Y, Thoma NH, Zheng N, Chen PL, Lee WH and Pavletich NP: BRCA2 function in DNA binding and recombination from a BRCA2-DSS1-ssDNA structure. Science. 297:1837–1848. 2002. View Article : Google Scholar : PubMed/NCBI

88 

Buisson R, Niraj J, Pauty J, Maity R, Zhao W, Coulombe Y, Sung P and Masson JY: Breast cancer proteins PALB2 and BRCA2 stimulate polymerase η in recombination-associated DNA synthesis at blocked replication forks. Cell Rep. 6:553–564. 2014. View Article : Google Scholar : PubMed/NCBI

89 

Hussain S, Wilson JB, Medhurst AL, Hejna J, Witt E, Ananth S, Davies A, Masson JY, Moses R, West SC, et al: Direct interaction of FANCD2 with BRCA2 in DNA damage response pathways. Hum Mol Genet. 13:1241–1248. 2004. View Article : Google Scholar : PubMed/NCBI

90 

Higgs MR and Stewart GS: Protection or resection: BOD1L as a novel replication fork protection factor. Nucleus. 7:34–40. 2016. View Article : Google Scholar : PubMed/NCBI

91 

Preobrazhenska O, Yakymovych M, Kanamoto T, Yakymovych I, Stoika R, Heldin CH and Souchelnytskyi S: BRCA2 and Smad3 synergize in regulation of gene transcription. Oncogene. 21:5660–5664. 2002. View Article : Google Scholar : PubMed/NCBI

92 

Hughes-Davies L, Huntsman D, Ruas M, Fuks F, Bye J, Chin SF, Milner J, Brown LA, Hsu F, Gilks B, et al: EMSY links the BRCA2 pathway to sporadic breast and ovarian cancer. Cell. 115:523–535. 2003. View Article : Google Scholar : PubMed/NCBI

93 

Marmorstein LY, Kinev AV, Chan GK, Bochar DA, Beniya H, Epstein JA, Yen TJ and Shiekhattar R: A human BRCA2 complex containing a structural DNA binding component influences cell cycle progression. Cell. 104:247–257. 2001. View Article : Google Scholar : PubMed/NCBI

94 

Singh JK, Smith R, Rother MB, de Groot AJL, Wiegant WW, Vreeken K, D'Augustin O, Kim RQ, Qian H, Krawczyk PM, et al: Zinc finger protein ZNF384 is an adaptor of Ku to DNA during classical non-homologous end-joining. Nat Commun. 12:65602021. View Article : Google Scholar : PubMed/NCBI

95 

Prakash R, Zhang Y, Feng W and Jasin M: Homologous recombination and human health: The roles of BRCA1, BRCA2, and associated proteins. Cold Spring Harb Perspect Biol. 7:a0166002015. View Article : Google Scholar : PubMed/NCBI

96 

Huang KL, Mashl RJ, Wu Y, Ritter DI, Wang J, Oh C, Paczkowska M, Reynolds S, Wyczalkowski MA, Oak N, et al: Pathogenic Germline variants in 10,389 adult cancers. Cell. 173:355–370. 2018. View Article : Google Scholar : PubMed/NCBI

97 

Van Hout CV, Tachmazidou I, Backman JD, Hoffman JD, Liu D, Pandey AK, Gonzaga-Jauregui C, Khalid S, Ye B, Banerjee N, et al: Exome sequencing and characterization of 49,960 individuals in the UK Biobank. Nature. 586:749–756. 2020. View Article : Google Scholar : PubMed/NCBI

98 

Engel C and Fischer C: Breast cancer risks and risk prediction models. Breast Care (Basel). 10:7–12. 2015. View Article : Google Scholar

99 

Wu H, Wu X and Liang Z: Impact of germline and somatic BRCA1/2 mutations: Tumor spectrum and detection platforms. Gene Ther. 24:601–609. 2017. View Article : Google Scholar : PubMed/NCBI

100 

Seo A, Steinberg-Shemer O, Unal S, Casadei S, Walsh T, Gumruk F, Shalev S, Shimamura A, Akarsu NA, Tamary H and King MC: Mechanism for survival of homozygous nonsense mutations in the tumor suppressor gene BRCA1. Proc Natl Acad Sci USA. 115:5241–5246. 2018. View Article : Google Scholar :

101 

Petrucelli N, Daly MB and Pal T: BRCA1- and BRCA2-Associated Hereditary Breast and Ovarian Cancer. GeneReviews®. Adam MP, Ardinger HH, Pagon RA, Wallace SE, Bean LJH, Gripp KW, Mirzaa GM and Amemiya A: University of Washington; Seattle: pp. 1993–2022. 2022

102 

Winter C, Nilsson MP, Olsson E, George AM, Chen Y, Kvist A, Törngren T, Vallon-Christersson J, Hegardt C, Häkkinen J, et al: Targeted sequencing of BRCA1 and BRCA2 across a large unselected breast cancer cohort suggests that one-third of mutations are somatic. Ann Oncol. 27:1532–1538. 2016. View Article : Google Scholar : PubMed/NCBI

103 

Cline MS, Liao RG, Parsons MT, Paten B, Alquaddoomi F, Antoniou A, Baxter S, Brody L, Cook-Deegan R, Coffin A, et al: BRCA challenge: BRCA exchange as a global resource for variants in BRCA1 and BRCA2. PLoS Genet. 26:e10077522018. View Article : Google Scholar

104 

Anczuków O, Ware MD, Buisson M, Zetoune AB, Stoppa-Lyonnet D, Sinilnikova OM and Mazoyer S: Does the nonsense-mediated mRNA decay mechanism prevent the synthesis of truncated BRCA1, CHK2, and p53 proteins? Hum Mutat. 29:65–73. 2008. View Article : Google Scholar

105 

Roa BB, Boyd AA, Volcik K and Richards CS: Ashkenazi Jewish population frequencies for common mutations in BRCA1 and BRCA2. Nat Genet. 14:185–187. 1996. View Article : Google Scholar : PubMed/NCBI

106 

Struewing JP, Abeliovich D, Peretz T, Avishai N, Kaback MM, Collins FS and Brody LC: The carrier frequency of the BRCA1 185delAG mutation is approximately 1 percent in Ashkenazi Jewish individuals. Nat Genet. 11:198–200. 1995. View Article : Google Scholar : PubMed/NCBI

107 

Ozolina S, Sinicka O, Jankevics E, Inashkina I, Lubinski J, Gorski B, Gronwald J, Nasedkina T, Fedorova O, Lyubchenko L and Tihomirova L: The 4154delA mutation carriers in the BRCA1 gene share a common ancestry. Fam Cancer. 8:1–4. 2009. View Article : Google Scholar

108 

Kaufman B, Laitman Y, Gronwald J, Lubinski J and Friedman E: Haplotype of the C61G BRCA1 mutation in Polish and Jewish individuals. Genet Test Mol Biomarkers. 13:465–469. 2009. View Article : Google Scholar : PubMed/NCBI

109 

Borg A, Dørum A, Heimdal K, Maehle L, Hovig E and Møller P: BRCA1 1675delA and 1135insA account for one third of Norwegian familial breast-ovarian cancer and are associated with later disease onset than less frequent mutations. Dis Markers. 15:79–84. 1999. View Article : Google Scholar : PubMed/NCBI

110 

Møller P, Heimdal K, Apold J, Fredriksen A, Borg A, Hovig E, Hagen A, Hagen B, Pedersen JC, Maehle L, et al: Genetic epidemiology of BRCA1 mutations in Norway. Eur J Cancer. 37:2428–2434. 2001. View Article : Google Scholar : PubMed/NCBI

111 

Heimdal K, Maehle L, Apold J, Pedersen JC and Møller P: The Norwegian founder mutations in BRCA1: High penetrance confirmed in an incident cancer series and differences observed in the risk of ovarian cancer. Eur J Cancer. 39:2205–2213. 2003. View Article : Google Scholar : PubMed/NCBI

112 

Sarantaus L, Huusko P, Eerola H, Launonen V, Vehmanen P, Rapakko K, Gillanders E, Syrjäkoski K, Kainu T, Vahteristo P, et al: Multiple founder effects and geographical clustering of BRCA1 and BRCA2 families in Finland. Eur J Hum Genet. 8:757–763. 2000. View Article : Google Scholar : PubMed/NCBI

113 

Thomassen M, Hansen TV, Borg A, Lianee HT, Wikman F, Pedersen IS, Bisgaard ML, Nielsen FC, Kruse TA and Gerdes AM: BRCA1 and BRCA2 mutations in Danish families with hereditary breast and/or ovarian cancer. Acta Oncol. 47:772–777. 2008. View Article : Google Scholar : PubMed/NCBI

114 

Einbeigi Z, Bergman A, Kindblom LG, Martinsson T, Meis-Kindblom JM, Nordling M, Suurküla M, Wahlström J, Wallgren A and Karlsson P: A founder mutation of the BRCA1 gene in Western Sweden associated with a high incidence of breast and ovarian cancer. Eur J Cancer. 37:1904–1909. 2001. View Article : Google Scholar : PubMed/NCBI

115 

Muller D, Bonaiti-Pellié C, Abecassis J, Stoppa-Lyonnet D and Fricker JP: BRCA1 testing in breast and/or ovarian cancer families from northeastern France identifies two common mutations with a founder effect. Fam Cancer. 3:15–20. 2004. View Article : Google Scholar : PubMed/NCBI

116 

Hartmann C, John AL, Klaes R, Hofmann W, Bielen R, Koehler R, Janssen B, Bartram CR, Arnold N and Zschocke J: Large BRCA1 gene deletions are found in 3% of German high-risk breast cancer families. Hum Mutat. 24:5342004. View Article : Google Scholar : PubMed/NCBI

117 

Pisano M, Cossu A, Persico I, Palmieri G, Angius A, Casu G, Palomba G, Sarobba MG, Rocca PC, Dedola MF, et al: Identification of a founder BRCA2 mutation in Sardinia. Br J Cancer. 82:553–559. 2000. View Article : Google Scholar : PubMed/NCBI

118 

Baudi F, Quaresima B, Grandinetti C, Cuda G, Faniello C, Tassone P, Barbieri V, Bisegna R, Ricevuto E, Conforti S, et al: Evidence of a founder mutation of BRCA1 in a highly homogeneous population from southern Italy with breast/ovarian cancer. Hum Mutat. 18:163–164. 2001. View Article : Google Scholar : PubMed/NCBI

119 

Cipollini G, Tommasi S, Paradiso A, Aretini P, Bonatti F, Brunetti I, Bruno M, Lombardi G, Schittulli F, Sensi E, et al: Genetic alterations in hereditary breast cancer. Ann Oncol. 15(Supp 1): SI7–SI13. 2004. View Article : Google Scholar

120 

Ikeda N, Miyoshi Y, Yoneda K, Shiba E, Sekihara Y, Kinoshita M and Noguchi S: Frequency of BRCA1 and BRCA2 germline mutations in Japanese breast cancer families. Int J Cancer. 91:83–88. 2001. View Article : Google Scholar : PubMed/NCBI

121 

Sekine M, Nagata H, Tsuji S, Hirai Y, Fujimoto S, Hatae M, Kobayashi I, Fujii T, Nagata I, Ushijima K, et al: Japanese Familial Ovarian Cancer Study Group. Mutational analysis of BRCA1 and BRCA2 and clinicopathologic analysis of ovarian cancer in 82 ovarian cancer families: Two common founder mutations of BRCA1 in Japanese population. Clin Cancer Res. 7:3144–3150. 2001.PubMed/NCBI

122 

Kang E and Kim SW: The korean hereditary breast cancer study: Review and future perspectives. J Breast Cancer. 16:245–253. 2013. View Article : Google Scholar : PubMed/NCBI

123 

Kang E, Seong MW, Park SK, Lee JW, Lee J, Kim LS, Lee JE, Kim SY, Jeong J, Han SA, et al: Korean hereditary breast cancer study group. The prevalence and spectrum of BRCA1 and BRCA2 mutations in Korean population: Recent update of the Korean hereditary breast cancer (KOHBRA) study. Breast Cancer Res Treat. 151:157–168. 2015. View Article : Google Scholar : PubMed/NCBI

124 

Kwong A, Ng EK, Wong CL, Law FB, Au T, Wong HN, Kurian AW, West DW, Ford JM and Ma ES: Identification of BRCA1/2 founder mutations in Southern Chinese breast cancer patients using gene sequencing and high resolution DNA melting analysis. PLoS One. 7:e439942012. View Article : Google Scholar : PubMed/NCBI

125 

De Leon Matsuda ML, Liede A, Kwan E, Mapua CA, Cutiongco EM, Tan A, Borg A and Narod SA: BRCA1 and BRCA2 mutations among breast cancer patients from the Philippines. Int J Cancer. 98:596–603. 2002. View Article : Google Scholar : PubMed/NCBI

126 

Concolino P and Capoluongo E: Detection of BRCA1/2 large genomic rearrangements in breast and ovarian cancer patients: An overview of the current methods. Expert Rev Mol Diagn. 19:795–802. 2019. View Article : Google Scholar : PubMed/NCBI

127 

Bhaskaran SP, Chandratre K, Gupta H, Zhang L, Wang X, Cui J, Kim YC, Sinha S, Jiang L, Lu B, et al: Germline variation in BRCA1/2 is highly ethnic-specific: Evidence from over 30,000 Chinese hereditary breast and ovarian cancer patients. Int J Cancer. 145:962–973. 2019. View Article : Google Scholar : PubMed/NCBI

128 

Hoogerbrugge N and Jongmans MC: Finding all BRCA pathogenic mutation carriers: Best practice models. Eur J Hum Genet. 24(Suppl 1): S19–S26. 2016. View Article : Google Scholar : PubMed/NCBI

129 

Hawsawi YM, Al-Numair NS, Sobahy TM, Al-Ajmi AM, Al-Harbi RM, Baghdadi MA, Oyouni AA and Alamer OM: The role of BRCA1/2 in hereditary and familial breast and ovarian cancers. Mol Genet Genomic Med. 7:e8792019. View Article : Google Scholar : PubMed/NCBI

130 

Hodgson A and Turashvili G: Pathology of hereditary breast and ovarian cancer. Front Oncol. 10:5317902020. View Article : Google Scholar : PubMed/NCBI

131 

Venter JC, Adams MD, Sutton GG, Kerlavage AR, Smith HO and Hunkapiller M: Shotgun sequencing of the human genome. Science. 280:1540–1542. 1998. View Article : Google Scholar : PubMed/NCBI

132 

Sanger F and Coulson AR: A rapid method for determining sequences in DNA by primed synthesis with DNA polymerase. J Mol Biol. 94:441–448. 1975. View Article : Google Scholar : PubMed/NCBI

133 

Wallace AJ: New challenges for BRCA testing: A view from the diagnostic laboratory. Eur J Hum Genet. 24(Suppl 1): S10–S18. 2016. View Article : Google Scholar : PubMed/NCBI

134 

Serratì S, De Summa S, Pilato B, Petriella D, Lacalamita R, Tommasi S and Pinto R: Next-generation sequencing: Advances and applications in cancer diagnosis. Onco Targets Ther. 9:7355–7365. 2016. View Article : Google Scholar : PubMed/NCBI

135 

Kumar KR, Cowley MJ and Davis RL: Next-generation sequencing and emerging technologies. Semin Thromb Hemost. 45:661–673. 2019. View Article : Google Scholar : PubMed/NCBI

136 

Idris SF, Ahmad SS, Scott MA, Vassiliou GS and Hadfield J: The role of high-throughput technologies in clinical cancer genomics. Expert Rev Mol Diagn. 13:167–181. 2013. View Article : Google Scholar : PubMed/NCBI

137 

Cheon JY, Mozersky J and Cook-Deegan R: Variants of uncertain significance in BRCA: A harbinger of ethical and policy issues to come? Genome Med. 6:1212014. View Article : Google Scholar

138 

Wong RSJ and Lee SC: BRCA sequencing of tumors: Understanding its implications in the oncology community. Chin Clin Oncol. 9:662020. View Article : Google Scholar : PubMed/NCBI

139 

Pujol P, Barberis M, Beer P, Friedman E, Piulats JM, Capoluongo ED, Foncillas JG, Ray-Coquard I, Penault-Llorca F, Foulkes WD, et al: Clinical practice guidelines for BRCA1 and BRCA2 genetic testing. Eur J Cancer. 146:30–47. 2021. View Article : Google Scholar : PubMed/NCBI

140 

Hindson BJ, Ness KD, Masquelier DA, Belgrader P, Heredia NJ, Makarewicz AJ, Bright IJ, Lucero MY, Hiddessen AL, Legler TC, et al: High-throughput droplet digital PCR system for absolute quantitation of DNA copy number. Anal Chem. 83:8604–8610. 2011. View Article : Google Scholar : PubMed/NCBI

141 

Pinheiro LB, Coleman VA, Hindson CM, Herrmann J, Hindson BJ, Bhat S and Emslie KR: Evaluation of a droplet digital polymerase chain reaction format for DNA copy number quantification. Anal Chem. 84:1003–1011. 2012. View Article : Google Scholar :

142 

Olmedillas-López S, García-Arranz M and García-Olmo D: Current and emerging applications of droplet digital PCR in Oncology. Mol Diagn Ther. 21:493–510. 2017. View Article : Google Scholar : PubMed/NCBI

143 

Postel M, Roosen A, Laurent-Puig P, Taly V and Wang-Renault SF: Droplet-based digital PCR and next generation sequencing for monitoring circulating tumor DNA: A cancer diagnostic perspective. Expert Rev Mol Diagn. 18:7–17. 2018. View Article : Google Scholar

144 

Stella M, Falzone L, Caponnetto A, Gattuso G, Barbagallo C, Battaglia R, Mirabella F, Broggi G, Altieri R, Certo F, et al: Serum extracellular vesicle-derived circHIPK3 and circS-MARCA5 are two novel diagnostic biomarkers for glioblastoma multiforme. Pharmaceuticals (Basel). 14:6182021. View Article : Google Scholar

145 

Crimi S, Falzone L, Gattuso G, Grillo CM, Candido S, Bianchi A and Libra M: droplet digital PCR analysis of liquid biopsy samples unveils the diagnostic role of hsa-miR-133a-3p and hsa-miR-375-3p in oral cancer. Biology (Basel). 9:3792020.

146 

Falzone L, Gattuso G, Tsatsakis A, Spandidos D and Libra M: Current and innovative methods for the diagnosis of COVID-19 infection (Review). Int J Mol Med. 47:1002021. View Article : Google Scholar :

147 

Falzone L, Musso N, Gattuso G, Bongiorno D, Palermo CI, Scalia G, Libra M and Stefani S: Sensitivity assessment of droplet digital PCR for SARS-CoV-2 detection. Int J Mol Med. 46:957–964. 2020. View Article : Google Scholar : PubMed/NCBI

148 

Preobrazhenskaya EV, Bizin IV, Kuligina ES, Shleykina AY, Suspitsin EN, Zaytseva OA, Anisimova EI, Laptiev SA, Gorodnova TV, Belyaev AM, et al: Detection of BRCA1 gross rearrangements by droplet digital PCR. Breast Cancer Res Treat. 165:765–770. 2017. View Article : Google Scholar : PubMed/NCBI

149 

Oscorbin I, Kechin A, Boyarskikh U and Filipenko M: Multiplex ddPCR assay for screening copy number variations in BRCA1 gene. Breast Cancer Res Treat. 178:545–555. 2019. View Article : Google Scholar : PubMed/NCBI

150 

Khalique S, Pettitt SJ, Kelly G, Tunariu N, Natrajan R, Banerjee S and Lord CJ: Longitudinal analysis of a secondary BRCA2 mutation using digital droplet PCR. J Pathol Clin Res. 6:3–11. 2020. View Article : Google Scholar :

151 

De Paolis E, De Bonis M, Concolino P, Piermattei A, Fagotti A, Urbani A, Scambia G, Minucci A and Capoluongo E: Droplet digital PCR for large genomic rearrangements detection: A promising strategy in tissue BRCA1 testing. Clin Chim Acta. 513:17–24. 2021. View Article : Google Scholar

152 

Manchanda R, Sun L, Patel S, Evans O, Wilschut J, De Freitas Lopes AC, Gaba F, Brentnall A, Duffy S, Cui B, et al: Economic evaluation of population-based BRCA1/BRCA2 mutation testing across multiple countries and health systems. Cancers (Basel). 12:19292020. View Article : Google Scholar

153 

Garcia J, Forestier J, Dusserre E, Wozny AS, Geiguer F, Merle P, Tissot C, Ferraro-Peyret C, Jones FS, Edelstein DL, et al: Cross-platform comparison for the detection of RAS mutations in cfDNA (ddPCR Biorad detection assay, BEAMing assay, and NGS strategy). Oncotarget. 9:21122–21131. 2018. View Article : Google Scholar : PubMed/NCBI

154 

Ratajska M, Koczkowska M, Żuk M, Gorczyński A, Kuźniacka A, Stukan M, Biernat W, Limon J and Wasąg B: Detection of BRCA1/2 mutations in circulating tumor DNA from patients with ovarian cancer. Oncotarget. 8:101325–101332. 2017. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

May-2022
Volume 60 Issue 5

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Lavoro A, Scalisi A, Candido S, Zanghì GN, Rizzo R, Gattuso G, Caruso G, Libra M and Falzone L: Identification of the most common BRCA alterations through analysis of germline mutation databases: Is droplet digital PCR an additional strategy for the assessment of such alterations in breast and ovarian cancer families?. Int J Oncol 60: 58, 2022
APA
Lavoro, A., Scalisi, A., Candido, S., Zanghì, G.N., Rizzo, R., Gattuso, G. ... Falzone, L. (2022). Identification of the most common BRCA alterations through analysis of germline mutation databases: Is droplet digital PCR an additional strategy for the assessment of such alterations in breast and ovarian cancer families?. International Journal of Oncology, 60, 58. https://doi.org/10.3892/ijo.2022.5349
MLA
Lavoro, A., Scalisi, A., Candido, S., Zanghì, G. N., Rizzo, R., Gattuso, G., Caruso, G., Libra, M., Falzone, L."Identification of the most common BRCA alterations through analysis of germline mutation databases: Is droplet digital PCR an additional strategy for the assessment of such alterations in breast and ovarian cancer families?". International Journal of Oncology 60.5 (2022): 58.
Chicago
Lavoro, A., Scalisi, A., Candido, S., Zanghì, G. N., Rizzo, R., Gattuso, G., Caruso, G., Libra, M., Falzone, L."Identification of the most common BRCA alterations through analysis of germline mutation databases: Is droplet digital PCR an additional strategy for the assessment of such alterations in breast and ovarian cancer families?". International Journal of Oncology 60, no. 5 (2022): 58. https://doi.org/10.3892/ijo.2022.5349