Next Article in Journal
Clinical and Molecular Features of Skin Malignancies in Muir-Torre Syndrome
Next Article in Special Issue
Identification of a Splice Variant (c.5074+3A>C) of BRCA1 by RNA Sequencing and TOPO Cloning
Previous Article in Journal
Amplified Fragments of an Autosome-Borne Gene Constitute a Significant Component of the W Sex Chromosome of Eremias velox (Reptilia, Lacertidae)
Previous Article in Special Issue
Making Sense of a Health Threat: Illness Representations, Coping, and Psychological Distress among BRCA1/2 Mutation Carriers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Imprecise Medicine: BRCA2 Variants of Uncertain Significance (VUS), the Challenges and Benefits to Integrate a Functional Assay Workflow with Clinical Decision Rules

by
Judit Jimenez-Sainz
1,* and
Ryan B. Jensen
1,2,*
1
Department of Therapeutic Radiology, Yale University School of Medicine, New Haven, CT 06520, USA
2
Department of Pathology, Yale University School of Medicine, New Haven, CT 06520, USA
*
Authors to whom correspondence should be addressed.
Submission received: 28 April 2021 / Revised: 14 May 2021 / Accepted: 18 May 2021 / Published: 20 May 2021
(This article belongs to the Special Issue BRCA1 and BRCA2: Genome Instability and Tumorigenesis)

Abstract

:
Pathological mutations in homology-directed repair (HDR) genes impact both future cancer risk and therapeutic options for patients. HDR is a high-fidelity DNA repair pathway for resolving DNA double-strand breaks throughout the genome. BRCA2 is an essential protein that mediates the loading of RAD51 onto resected DNA breaks, a key step in HDR. Germline mutations in BRCA2 are associated with an increased risk for breast, ovarian, prostate, and pancreatic cancer. Clinical findings of germline or somatic BRCA2 mutations in tumors suggest treatment with platinum agents or PARP inhibitors. However, when genetic analysis reveals a variant of uncertain significance (VUS) in the BRCA2 gene, precision medicine-based decisions become complex. VUS are genetic changes with unknown pathological impact. Current statistics indicate that between 10–20% of BRCA sequencing results are VUS, and of these, more than 50% are missense mutations. Functional assays to determine the pathological outcome of VUS are urgently needed to provide clinical guidance regarding cancer risk and treatment options. In this review, we provide a brief overview of BRCA2 functions in HDR, describe how BRCA2 VUS are currently assessed in the clinic, and how genetic and biochemical functional assays could be integrated into the clinical decision process. We suggest a multi-step workflow composed of robust and accurate functional assays to correctly evaluate the potential pathogenic or benign nature of BRCA2 VUS. Success in this precision medicine endeavor will offer actionable information to patients and their physicians.

1. Introduction

Breast Cancer Susceptibility Gene 2 (BRCA2) was identified in 1994 at the Institute of Cancer Research by the Stratton Laboratory [1,2], only one year after the gene mapping of BRCA1 [3]. Germline mutations in the BRCA genes account for 20–60% of breast cancer cases in families with multiple cancer incidence [4]. The lifetime risk for breast and ovarian cancer rises dramatically for BRCA mutation germline carriers compared to the general population (Figure 1). The overall risk varies by family history, type and location of the mutation, age at diagnosis, parity (number of term pregnancies), environmental factors, and genetic modifiers [5,6,7,8,9,10]. The discovery of the hereditary breast and ovarian cancer (HBOC) BRCA genes in the early 1990s laid the framework for the future of precision medicine as patients became empowered with the ability to make life-changing decisions to manage and mitigate cancer risks [11,12].
Rapid advances in DNA sequencing technology coupled with public awareness and quick adoption of genetic testing for HBOC genes in genetic counseling clinics have resulted in vast database repositories of BRCA mutations and variants. While the majority of BRCA mutations are associated with breast, ovarian, pancreatic, and prostate cancers, mutations have been documented in other tumor sites such as lung and malignant melanoma.
Soon after the BRCA gene sequences were determined, Myriad Genetics initiated a vigorous intellectual property protection strategy providing them with the exclusive rights to provide all diagnostic testing. A supreme court decision in 2013 overturned Myriads’ patents and opened up the genetic testing landscape to several commercial and academic laboratories. As a result, thousands of BRCA variants have been deposited into various databases. While the majority of BRCA sequencing information has been released, Myriad still maintains a large database of BRCA variants not accessible by the public research community (https://www.wsj.com/articles/geneticists-call-on-myriad-to-share-proprietary-data-to-aid-gene-tests-11578851248, accessed on 5 May 2021).
Genetic testing of patient germline DNA and primary tumor DNA (or circulating tumor DNA (ctDNA)) has resulted in more than 30,000 genetic changes in the BRCA genes that lack functional interpretation. This knowledge gap has led to uncertainty in the ability of genetic counselors to predict cancer risk and impedes the successful clinical management of patients who may benefit from targeted therapies. International collaborations and consortiums have joined forces to design integrative approaches to combine familial and clinical data, in silico predictions, and functional analysis of BRCA2 variants of uncertain significance (VUS). However, the challenge to evaluate extremely rare VUS looms large and functional assays may be the only avenue for some patients to rely on for fundamental life-changing preventative clinical options.
In this review, we briefly cover the role of BRCA2 in HDR repair of DNA double-strand breaks (DSBs), pathogenic BRCA2 mutations and their impact on cancer risk and treatment options, database and classification methods for VUS, current clinical management strategies for BRCA2 patients, and the challenges of adopting functional analyses into clinical practice.

2. BRCA2 Protein, Functions, and Patient Mutations

BRCA2 plays a key role in DNA damage repair via HDR, a high-fidelity repair pathway for DNA DSB repair [13,14]. In mammalian cells, DNA DSBs can be repaired by two major pathways, non-homologous end joining or HDR. HDR requires a template, usually the sister chromatid, and is thus restricted to the S or G2 phase of the cell cycle. If a cell encounters a DNA DSB and commits to the HDR pathway, several nucleases will resect the DNA end resulting in single-stranded DNA (ssDNA) 3’ terminated tails. The ssDNA is immediately coated with Replication Protein A (RPA) to remove the secondary structure and prevent self-annealing. BRCA2 then delivers RAD51 to the ssDNA displacing the RPA protein while simultaneously preventing RAD51 from binding to dsDNA and downregulating the ATPase activity of RAD51 [15]. These mediator activities of BRCA2 result in nucleation and stabilization of RAD51 nucleoprotein filaments on ssDNA, a necessary step preceding RAD51-dependent DNA strand invasion into a homologous donor [15,16,17,18,19,20,21]. Another interesting feature of BRCA2 is its role in replication fork protection. Utilizing the DNA fiber combing technique, several groups have found that under conditions of replicative stress, BRCA2 deficient cells undergo degradation of nascent DNA [22,23,24,25,26]. The hypothesis is that RAD51 is unable to protect the annealed nascent DNA strands at a reversed fork in the absence of BRCA2, and unchecked nucleolytic degradation by MRE11 and other resection nucleases leads to fork collapse and DNA DSBs [25]. A separation-of-function mutation located in the C-terminal domain of BRCA2 (Figure 2), S3291A, was shown to have a defect in fork protection while remaining functionally competent for HDR [25]. Described in detail elsewhere, an ever-growing list of functions has been ascribed to BRCA2, such as R loop processing, gap suppression following DNA replication, and cytokinesis [27,28,29,30,31].
The BRCA2 genomic locus is 84 kb, contains 27 exons, and is located on chromosome 13 [12]. The BRCA2 gene encodes a 3418-amino acid nuclear protein with four distinct domains: An N-terminal domain containing residues important for interactions with PALB2 and EMSY, a central BRC repeats region, a DNA binding domain (DBD) containing three tandem oligonucleotide/oligosaccharide-binding folds (OB-folds), and a C-terminal domain (CTD) (Figure 2). PALB2 (Partner and Localizer of BRCA2) mediates the interaction between BRCA1 and BRCA2, plays a role in the nuclear localization of BRCA2, aids in RAD51 loading at resected DNA breaks, and pathogenic mutations in PALB2 have been found to predispose to breast, ovarian, and prostate cancer. EMSY is a BRCA2-interacting transcriptional repressor that intriguingly is linked to sporadic breast and ovarian cancer [32,33,34,35,36]. The eight BRC repeats play critical roles in binding and regulating RAD51 loading and filament stability [37,38,39]. The CTD also functions in binding and stabilizing the RAD51 nucleoprotein filament. BRCA2 plays an important role in meiosis, and hence, binds the meiotic counterpart to RAD51, DMC1 [40]. DSS1 is a small acidic protein visible in the crystal structure of the DBD, makes several intimate contacts with BRCA2, and has been found to facilitate RPA displacement [41,42,43]. Together, all of the protein and DNA interaction domains of BRCA2 promote successful HDR and fork protection [32,35,37,38,43,44,45,46,47,48,49].
It is essential that BRCA2 localizes to the nucleus to carry out its genome integrity and tumor suppressor functions, and thus, contains several putative nuclear localization signals (NLS) [51,52,53,54]. Surprisingly, only a handful of studies have characterized both the NLS and nuclear export signals (NES) of BRCA2, and it remains unclear which specific sequences are necessary and sufficient for the proper regulation of BRCA2 nuclear/cytoplasmic trafficking [51,52,53,54,55].
In cellular models, disruption of BRCA2 leads to the accumulation of DNA DSBs, a severe decrease in HDR activity, and consequently, sensitivity to DNA damaging agents that induce replication fork stalling and/or DNA DSBs such as crosslinking agents or PARP inhibitors (PARPi) [44,56,57,58]. In addition, BRCA2 deficiency leads to micronuclei formation and centrosome amplification, ultimately contributing to genomic instability [59]. BRCA2 knockout mice are embryonic lethal, and only a handful of cell lines (VC8 (hamster); DLD1, CAPAN1, PEO1 (human)) are viable without the expression of full-length protein supporting the essential role that BRCA2 plays in early development and viability [60,61,62,63,64].
Cancer-associated mutations in BRCA2 are commonly found in the BRC repeat and OB-fold regions [9,65,66]. Population-based studies identified more than 10 BRCA2 founder mutations including: 6174delT (BRC repeat region) in Ashkenazi Jews and 7252C>T (DBD region) in the Korean and Finnish populations [67,68,69,70]. BRCA2 mutations found to increase cancer risk consist of indels, frameshifts, and nonsense mutations, ultimately leading to loss of expression or a truncated protein product. As the putative NLS signals of BRCA2 are located at the C-terminus of the protein, most truncated protein products will be mislocalized to the cytoplasm rendering the cells functionally null. Importantly, mutations that lead to loss of nuclear localization of BRCA2 are associated with HDR dysfunction and cancer predisposition [53,62].
The molecular route to tumorigenesis in BRCA2 mutant germline carriers remains an unsolved mystery. The majority of BRCA2 deficient tumors have lost the wild-type allele in a classic case of loss-of-heterozygosity (LOH), however, exceptions to the Knudson’s two-hit hypothesis rule have been described [71]. A generally accepted belief is that BRCA2 haploinsufficiency provides a chronic state of low-level genomic instability conducive to the accumulation of mutations that eventually drive the tumorigenic process. As BRCA2 functions seem necessary for cellular viability, the specific genetic background that is permissive for tumor growth is unknown. Furthermore, the genetic and pathological basis for why BRCA2 (and BRCA1) mutations lead preferentially to cancer of the breast and ovaries remains an ever-present enigma.
The ClinVar database contains over 13,000 reports of germline and somatic BRCA2 alterations with more than 5000 VUS (Figure 3 left panel). Approximately 6000 of the 9500 BRCA2 molecular mutations reported are single amino acid changes in the 3,418-amino acid BRCA2 protein (Figure 3 right panel). Many of the mutations evaluated and reviewed as pathogenic are indeed frameshift and nonsense mutations resulting in either a truncated protein or complete lack of BRCA2 expression.
The tumor spectrum of cancers associated with pathogenic mutations in BRCA2 include: breast, ovarian, fallopian tube, melanoma, prostate, pancreatic, and lung cancers (Figure 4A). The display of all BRCA2 germline and somatic mutations described in the curated set of non-redundant studies of cBioPortal for Cancer Genomics shows that BRCA2 is altered in 10–15% of breast, ovarian, and prostate cancer cases, whereas BRCA2 is altered in 18% of melanoma cases (Figure 4A). The location of BRCA2 mutations in the dataset does not suggest specific “hot-spots” or “cold-spots” but widespread alterations across the whole gene (Figure 4B). As of 2020, a new database, HotSpotsAnnotations, applied a statistical model to detect putative hotspots on The Cancer Genome Atlas (TCGA) cancer datasets which contain 10,182 patients and 33 cancer types. In the cancer HotSpots Database, six BRCA2 mutations were identified at positions 1393, 1689, 1782, 2842, 3308, and 3342 (http://bioinformatica.mty.itesm.mx:8080/HotSpotsAnnotations/PiquinSpot.jsp?gene=BRCA2, accessed on 25 April 2021) [72]. The first three were located in the BRC domain, while the latter were mapped to the DNA binding domain. Controversy remains as some studies demonstrated no correlation, and others, a positive correlation between BRCA2 mutations in the BRC domain and ovarian cancer patient survival [73,74]. Discrepancies arise in part due to the selection of the BRCA2 mutations (missense, indels, etc.), the number of patients included, or the treatment provided. A single study in 2020 classified BRCA2 exon 10 and 11 as “cold-spots”, or more tolerant to variation [75]. However, the analysis was based on limited data, and thus overall, additional studies are needed regarding the frequency of mutation locations across the BRCA2 genomic locus.
The data obtained from cBioPortal indicate a total of 1934 BRCA2 germline and somatic alterations in the curated set of non-redundant studies. Of the 518 driver mutations, 83.6% are truncations (Figure 5 left panel), whereas missense mutations make up 96.7% (1369) of VUS (1416) (Figure 5 right panel). While truncating mutations in BRCA2 are clearly pathogenic in the majority of cases, the curated data highlight the need to further understand missense mutations, as they represent the most abundant VUS.

3. Clinical Management of Patients

The American College of Obstetrician and Gynecologists (ACOG) and the American Cancer Society have created guidelines for clinicians to manage patients with BRCA2 gene mutations (Figure 6) [76,77].
Sequencing-based genetic tests to screen for BRCA mutations are offered to individuals with a family history of breast or ovarian cancer. As a risk-reduction strategy, pathogenic BRCA mutation carriers, who are cancer-free, will be informed of surgical interventions such as a double mastectomy and/or unilateral salpingo-oophorectomy. In many cases, female BRCA carriers during their early reproductive years may opt for careful surveillance in order to delay prophylactic surgeries that would prevent them from having children. In 2013, the so-called “Angelina effect” sparked an acute public awareness of genetic testing for hereditary cancer risk genes, which may have resulted in many people undergoing unnecessary testing without significant associated risk factors such as a family history of unusually high cancer incidence [78,79]. In a worst-case scenario, a recent article in the Wall Street Journal described a family where multiple women underwent prophylactic surgeries based on a “likely” pathogenic BRCA VUS finding that was later re-classified as a benign variant (https://www.wsj.com/articles/seven-women-in-a-family-chose-surgery-after-a-genetic-test-then-the-results-changed-11576860210, accessed on 24 April 2021). This failure of precision medicine underscores the need to cautiously interpret VUS findings and to bolster VUS evaluations with additional evidence for pathogenicity. We suggest that rigorous functional assays performed by laboratories with expertise in BRCA2 biology could provide meaningful criteria for evaluating rare VUS when familial, co-segregation, co-occurrence, and other classical genetic linkage data are lacking.
For patients who learn they are BRCA2 VUS carriers, the situation is complex and frustrating. Not only does the patient now know their BRCA2 gene is not normal (deviates from the wild-type sequence), the clinical significance of the VUS findings are rarely known, leaving the patient with unknown future cancer risk. Universally accepted standards for how to deal with BRCA VUS have yet to be established, revealing a shortcoming in our ability to make informed clinical decisions [78,79,80]. Thus, it is critical that genetic counselors have the ability to intercede and minimally inform patients of their options thus that rational choices can be made.
Patients diagnosed with cancer and for whom a germline or somatic pathogenic BRCA2 mutation is found often respond well to chemotherapy agents such as PARPi and platinum drugs. LOH of the wild-type BRCA2 allele is usually associated with a robust clinical response to these drugs but not in every case [71]. A tumor that is classified as HDR deficient is likely the best biomarker correlating with clinical response to platinum drugs and PARPi. Some patients still fail to respond to these drugs, and others may initially respond, but ultimately relapse as resistant tumor cells emerge and recolonize the tumor. Clinical trials with inhibitors such as AZD2281, 6-Mercaptopurine, and methotrexate are ongoing for BRCA2 defective ovarian tumors focused on resistance. Somatic findings of BRCA2 VUS in a tumor present a similar challenge to germline findings. In a tumor scenario, LOH of the wild-type BRCA2 allele may determine whether a patient is a good candidate for PARPi therapy. Again, if the tumor can be classified as HDR deficient, this result would provide the motivation for the use of synthetic lethal PARPi therapy. In summary, BRCA2 VUS, whether germline or somatic, presents an ongoing problem for clinical decision-making.

4. Curation of BRCA2 VUS

Several international groups and consortiums have joined efforts to create databases reporting BRCA2 genetic variants generated from sequencing analysis (Table 1 and Table 2). While publicly available databases are useful resources, curation of the information and classification of the variants is an ongoing challenge.
There are several commonly used online database resources that provide some interpretation of BRCA2 sequence variants (Table 2). For example, BRCA Exchange was created by the Global Alliance for Genomics and Health (GA4GH) and is the largest open-source database on BRCA variant information [83]. BRCA Exchange contains information from ClinVar, the Breast Cancer Information Core database, Leiden Open Variant Database (LOVD), several population databases, and is conveniently accessible through a smartphone app. Many of these databases, such as ClinVar and LOVD, are freely accessible and contain submissions reporting BRCA variants found in patient samples [84]. These databases display BRCA variants as expert-reviewed, interpreted, and classified by supporting evidence such as familial cases and co-segregation studies. Some of the databases display the BRCA information by molecular alteration and/or clinical significance (Table 2). The comparison of five databases with BRCA2 information shows that missense mutations are around 44–96% of the data deposited in the databases, and BRCA2 VUS make up 41–83% of ClinVar and BRCA Exchange databases. Thus, although collaborative efforts have unified the information of BRCA2 mutations, most remain unknown for cancer risk.

5. Methods of Classification

In the past decade, enormous effort has been dedicated to classifying BRCA2 VUS using a variety of clinical genetic methods and probability models. A standardized method of classifying BRCA2 variants was created by the International Agency for Research on Cancer (IARC) and the American College of Medical Genetics and Genomic (ACMG) based on a 5-tier system with a multifactorial likelihood model [85,86,87] (Figure 7). Multifactorial likelihood models describe the odds in favor of causality or neutrality based on analysis of personal and family history of cancer and co-segregation of the variant with disease in pedigrees [88]. Several genetic risk prediction models have been created based on familial and personal history. These genetic risk models include: BRCAPRO [89], IBIS [90], and BOADICEA [91]. Several multifactorial models include pathology or immunohistochemistry analysis of the tumors together with the cancer family history [92,93,94], and other models with co-segregation studies consider gender, genotype, and onset age of cancer [95].
Computational methods have been developed to help predict amino acid changes in protein function [96,97,98]. Polyphen2, SIFT, Align-GVGD, HCI database, GeneSplicer, and Human Splicing Finder are in silico online tools that can be utilized to evaluate sequence conservation and biochemical properties of changed amino acid residues in a clinical variant (see [99]) [100]. The posterior probability model combines sequence conservation and biochemical properties of the amino acid residue with the likelihood ratios [101].
Models hold the most utility in BRCA2 variants where the frequency and/or linkage analysis of the variant is high. However, rare variants (allele frequency < 0.01) are impossible to classify due to a lack of statistical power. Another layer of complexity is that the guidelines governing BRCA2 variant classification change very rapidly necessitating reinterpretation and leading to debate. Finally, if the clinical, genetic, and in silico data for BRCA2 VUS are not deposited in a freely accessible database, it becomes an almost impossible barrier for researchers and clinicians to overcome. In summary, functional studies accompanied by clinical annotation, and the aforementioned tools, may improve classification of BRCA2 variants.

6. Functional Studies

BRCA2 is a large and complex protein, and the literature to date has ascribed multiple functions to the protein. Loss of one particular activity does not necessarily directly link to cancer risk, and thus, evaluation of germline variants by a single assay should be interpreted carefully and in combination with clinical and familial data. In contrast, somatically mutated BRCA2 variants identified by sequencing of tumor DNA may correlate well with functional assays designed to measure chemotherapeutic response. For example, confirmation of tumor HDR deficiency is clinically actionable by stratifying patients for treatment with PARPi or platinum agents. Multiple studies have validated that in vitro cell culture models of BRCA2 deficiency respond in a similar manner to patients who receive PARPi therapy [102,103,104]. Of course, the genetic background of the patient, immune system makeup, prior treatments, tumor site, and other factors will impact the ultimate therapeutic benefit to patients.
The low frequency of rare VUS coupled with limited or no family history of cancer can make the correct evaluation of BRCA2 pathogenic classification very challenging. We and others have developed numerous biological and biochemical assays to assess the consequences of BRCA2 variants on protein function. These assays require high specificity and sensitivity to complement genetic linkage and epidemiological classification methods mentioned above successfully.
Functional complementation assays in BRCA2 deficient hamster, mouse, and human cells have been used extensively by several groups [65,67,105,106,107,108,109,110,111,112,113,114,115,116,117]. These assays have been documented in-depth in a previous review [65] but include the following: The DR-GFP reporter assay designed to measure gene conversion, yeast recombination assays, centrosome amplification, drug sensitivity assays, mouse ES cell complementation, the SyVal human DLD1 cell model (introduction of variants by gene targeting of BRCA2), nuclear localization, protein-protein, and protein-DNA interactions, and RNA splicing and mRNA maturation assays utilizing transcript analysis and minigenes [30,32,53,54,57,58,61,105,106,107,110,111,112,113,115,116,117,118,119,120,121,122,123,124,125,126]. Recently, a high-throughput functional evaluation termed the MANO-B method was used to measure the sensitivity of BRCA2 variants to PARPi in a mixed population of cells containing several BRCA2 variants. This large-scale method demonstrated a high correlation with IARC, ClinVar, and Align-GVGD classifications of BRCA2 pathogenic variants [127]. A highly useful variant analysis system was developed by the Sharan group in 2003 using a mouse embryonic stem (ES) cell-based complementation assay (functional BRCA2 VUS will rescue the lethality of a conditional BRCA2 allele). In addition to utilizing the endogenous BRCA2 promoter, this system has the advantage of maintaining the exon/intron structure of BRCA2 (the entire genomic region on a bacterial artificial chromosome (BAC) is utilized for complementation), and thus, has the ability to analyze splice variants in addition to variants within the open reading frame. BRCA2 VUS that provide viable complementation can be further tested for DNA damage sensitivity, HDR by the DR-GFP assay, or subtle changes in genomic integrity to identify hypomorphic alleles [47,105,106,107,111]. Finally, in 2018, the Shendure lab published a groundbreaking approach to analyze VUS by mutating each individual base in the RING and BRCT regions of BRCA1 utilizing a CRISPR-based technique termed saturation genome editing [128]. The approach relies on the lethality of human HAP1 cells (a unique haploid human cell line) upon the expression of a non-functional BRCA1 allele. Saturation genome editing analysis of BRCA1 variants appears to correlate well with clinically annotated variants, and this scalable approach should be applicable to other HDR genes critical for cell viability such as BRCA2, PALB2, and the RAD51 paralogs.
In total, while functional assays have been instrumental in the classification of BRCA2 VUS where familial and genetic information is scarce, clinical geneticists have been slow to incorporate lab-based results into counseling decisions for patients. Lack of standardization and uniformity, assays that can only measure specific regions of the protein, low throughput, uncertainty as to which functions of BRCA2 are necessary and sufficient for tumor suppression, and discordance between functional assay results with known variants of clinical significance have all likely contributed to the slow adoption. Furthermore, as with any clinical diagnostic assay, the risk of false negatives or false positives must be carefully addressed. However, in situations where rare VUS lack any clear genetic linkage data, we suggest that functional assays should be considered if used judiciously and with full disclosure.
In our laboratory, we have designed a simple multi-step workflow to cover many BRCA2 functions described to date. We have taken advantage of biochemical, genetic, cellular, and molecular complementation analysis to investigate the functionality and cellular localization of the BRCA2 protein (Figure 8). To do so, we incorporate each missense VUS into a mammalian expression vector placing the BRCA2 cDNA in frame with an N-terminal tandem maltose-binding protein (2XMBP) tag, which provides robust expression and stability. We have further demonstrated that the 2XMBP tag does not interfere with any of the cellular functions of BRCA2 [15]. The constructs are stably integrated into the human DLD1 BRCA2−/− cell line for cell-based and genetic evaluation [61]. Constructs containing BRCA2 variants are also introduced into facile human cell models such as 293T cells to assess protein-protein interactions and cellular localization. Finally, we perform complementation studies in immortalized human MCF10A breast epithelial and FTSEC (fallopian tube secretory epithelial) cells, relevant to breast and ovarian cancer, respectively, in which we have introduced an inducible shRNA against BRCA2. In contrast to CRISPR, or gene-targeted knockouts of BRCA2, besides being lethal in most cells, a conditional shRNA system for BRCA2 has the advantage of acute depletion in an otherwise genetically stable background. In the near future, we plan to take advantage of CRISPR knock-in technology to create conditional BRCA2 alleles at the endogenous loci.
The advantage of using a human cell model (e.g., DLD1 BRCA2−/−) that tolerates a genetically BRCA2 null background is that we can evaluate pathogenic BRCA2 variants directly. Cellular complementation systems that rely on viability are limited to benign, or in some cases, hypomorphic BRCA2 alleles for further analysis as pathogenic variants do not give rise to clones. While we acknowledge that our CMV-driven cDNA complementation system forgoes expression by the endogenous promoter, evaluation of several hundred stable cell clones has revealed little difference in complementation ability due to variability in expression levels. Using our approach, we have identified multiple missense mutations predicted to be pathogenic that are mislocalized to the cytoplasm (manuscript in preparation). This unexpected observation was only possible through the expression of full-length BRCA2 proteins in the DLD1 BRCA2−/− cell line. The phenotypic consequences of pathogenic BRCA2 allele cellular localization significantly impact our workflow decision process of whether further biochemical workup is warranted. A further advantage of our ability to interrogate deleterious BRCA2 alleles is that considerable uncertainty still exists as to which functions of BRCA2, HDR versus fork protection, are critical for chemotherapeutic response in patients. In summary, the potential for biological insight derived from analyzing separation-of-function alleles, and the capability to directly measure drug sensitivity for somatically mutated BRCA2 alleles in patient tumors, highlight the importance of our approach.
Our workflow for BRCA2 VUS evaluation (Figure 8) begins at BRCA2 localization by immunofluorescence. At this decision point, if BRCA2 is cytosolic, we conclude pathogenicity as cytoplasmic localization precludes any ability of BRCA2 to carry out its necessary genomic integrity functions in the nucleus. It remains unclear how and why certain BRCA2 missense VUS are retained in the cytosol, and this is an active research area in our laboratory. If BRCA2 is nuclear, we analyze biochemically and genetically the functionality of BRCA2 VUS. To do so, we leverage our ability to purify the full-length human BRCA2 protein [15,38] and interrogate altered biochemical functions such as DNA substrate and RAD51 binding, RAD51 filament stabilization, and stimulation of RAD51-dependent DNA strand exchange activity. In parallel, our cell-based models allow us to evaluate drug sensitivity (e.g., PARPi and crosslinking agents), RAD51 foci, fork protection, HDR functionality by DR-GFP assay, and gene targeting ability (Cas9/mClover LMNA assay).
Our approach complements and extends prior functional characterization of BRCA2 VUS by providing an exhaustive genetic and biochemical workup of pathogenic alleles. Pathogenic BRCA2 missense variants, in particular, present a rich source of information into the underlying biological functions of BRCA2 and can be leveraged in a “reverse translational” approach to decipher which specific attributes of BRCA2 are necessary for HDR, chemotherapeutic response, and tumor suppressor functions.
While many functional assays exist to examine HDR or fork protection activities, no specific test for the actual tumor-suppressive functions of BRCA2 has been developed. The lack of a definitive in vitro or in vivo model for the cancer-promoting activity of BRCA2 poses an issue when one considers the evaluation of germline VUS to predict future cancer risk. How does one decide which functional assay is critical for the long-term effects of BRCA2 haploinsufficiency leading to the eventual outgrowth of a clinically diagnosable cancer? The oncogene field has developed multiple approaches to interrogate the tumorigenic potential of specific gain-of-function mutations in genes such as kinases (e.g., SRC) or GTPases (e.g., RAS) including: foci assays, anchorage-independent growth in soft agar, genetically engineered mouse models (GEMM), or tumor xenograft studies. To date, no equivalent experimental approach exists for either BRCA gene. Hopefully, in the future, we will be able to parse out the exact functions of BRCA2 that are required to spark tumorigenesis, and this knowledge can be adapted into an assay that will accurately quantify the tumorigenic capacity of BRCA2 variants.

7. Conclusions and Future Directions

In the future, the volume of BRCA2 VUS findings will undoubtedly increase as sequencing technology becomes more economical, higher throughput, and adopted into clinical practice as a routine diagnostic. Patients, clinicians, and researchers will hopefully demand more open access to this information, with appropriate privacy controls in place, thus that VUS have a better chance of being correctly classified. The end goal of functional analysis is to derive a highly sensitive, specific, and predictive score that can be integrated with prior methods to correctly classify the large number of BRCA2 VUS and unmask both cancer risk and response to chemotherapeutics.
The current diversity and non-standard resources and databases for deciphering BRCA2 VUS information is a constant challenge for physicians and genetic counselors. Without proper curation and standardized guidelines, clinical decisions surrounding VUS are complex at best and irresolvable at worst. To improve classification efforts, a universal workflow could be adopted with agreed-upon expert guidelines to evaluate BRCA2 variants in a systematic fashion. These universal guidelines will need to escalate quickly as VUS are rapidly clogging the variant pipeline in step with clinical genetic testing panels and sequencing technology. Commercial companies such as Myriad, Ambry, deCODE, and Labcorp are adding to this VUS pipeline, and public awareness has further increased findings of VUS as more and more people seek out their own genetic information.
The collective sequences of BRCA variants are being centralized in databases such as BRCA Exchange and are growing exponentially. Just a few days before the submission of this review, BRCA Exchange announced on social media that the number of BRCA variants in their database would increase from 40,000 to over 60,000. The pace of VUS findings concurrent with inconclusive assessments of cancer risk or potential response to targeted therapy will undoubtedly accelerate in the future. Hence, the need for functional clarification intensifies.
We conclude that rationale, rigorous, and statistically sound functional analyses of BRCA2 VUS will unleash the current lack of clinical significance for thousands of rare VUS. Although laborious, biochemical and cellular functional assays have the potential to correctly assign a pathogenic or benign classification to a BRCA2 VUS. Functional information combined with familial, co-segregation, co-occurrence, and other classical genetic epidemiological data will reveal the full catalog of BRCA2 pathogenic mutations. Additionally, we will continue to leverage patient-derived BRCA2 pathogenic variants to enhance our understanding of the molecular events that lead to tumorigenesis. Our long-term goal is to remove the frustration associated with clinical BRCA2 VUS findings such that patients feel empowered with clinically actionable information fulfilling the modern promise of delivering precision medicine.

Author Contributions

Conceptualization, writing—original draft preparation, review, and editing, J.J.-S. and R.B.J.; supervision, R.B.J.; funding acquisition, R.B.J. All authors have read and agreed to the published version of the manuscript.

Funding

Research in the author’s laboratory was funded by the National Institutes of Health grant R01 CA215990 to R.B.J.; further support was provided by the American Cancer Society RSG-17-038-01, Women’s Health Research at Yale, and the Gray Foundation to R.B.J.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We thank members of the Jensen Lab for helpful discussions and comments on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wooster, R.; Bignell, G.; Lancaster, J.; Swift, S.; Seal, S.; Mangion, J.; Collins, N.; Gregory, S.; Gumbs, C.; Micklem, G.; et al. Identification of the breast cancer susceptibility gene brca2. Nature 1995, 378, 789–792. [Google Scholar] [CrossRef] [PubMed]
  2. Wooster, R.; Neuhausen, S.L.; Mangion, J.; Quirk, Y.; Ford, D.; Collins, N.; Nguyen, K.; Seal, S.; Tran, T.; Averill, D.; et al. Localization of a breast cancer susceptibility gene, BRCA2, to chromosome 13q12-13. Science 1994, 265, 2088–2090. [Google Scholar] [CrossRef] [PubMed]
  3. Bowcock, A.; Anderson, L.A.; Friedman, L.S.; Black, D.M.; Osborne-Lawrence, S.; Rowell, S.E.; Hall, J.M.; Solomon, E.; King, M.C. THRA1 and D17S183 flank an interval of <4 cm for the breast-ovarian cancer gene (BRCA1) on chromosome 17q21. Am. J. Hum. Genet. 1993, 52, 718–722. [Google Scholar]
  4. Nathanson, K.L.; Wooster, R.; Weber, B.L. Breast cancer genetics: What we know and what we need. Nat. Med. 2001, 7, 552–556. [Google Scholar] [CrossRef] [PubMed]
  5. Andrieu, N.; Goldgar, D.E.; Easton, D.F.; Rookus, M.; Brohet, R.; Antoniou, A.C.; Chang-Claude, J. Pregnancies, breast-feeding, and breast cancer risk in the international brca1/2 carrier cohort study (ibccs). J. Natl. Cancer Inst. 2006, 98, 535–544. [Google Scholar] [CrossRef] [PubMed]
  6. Cullinane, C.A.; Lubinski, J.; Neuhausen, S.L.; Ghadirian, P.; Lynch, H.T.; Isaacs, C.; Weber, B.; Moller, P.; Offit, K.; Kim-Sing, C.; et al. Effect of pregnancy as a risk factor for breast cancer inBRCA1/BRCA2 mutation carriers. Int. J. Cancer 2005, 117, 988–991. [Google Scholar] [CrossRef] [PubMed]
  7. Metcalfe, K.; Lubinski, J.; Lynch, H.T.; Ghadirian, P.; Foulkes, W.D.; Kim-Sing, C.; Neuhausen, S.; Tung, N.; Rosen, B.; Gronwald, J.; et al. Family History of Cancer and Cancer Risks in Women with BRCA1 or BRCA2 Mutations. J. Natl. Cancer Inst. 2010, 102, 1874–1878. [Google Scholar] [CrossRef] [Green Version]
  8. Milne, R.L.; Osorio, A.; Cajal, T.R.Y.; Baiget, M.; Lasa, A.; Diaz-Rubio, E.; De La Hoya, M.; Caldés, T.; Teulé, A.; Lázaro, C.; et al. Parity and the risk of breast and ovarian cancer in BRCA1 and BRCA2 mutation carriers. Breast Cancer Res. Treat. 2009, 119, 221–232. [Google Scholar] [CrossRef] [Green Version]
  9. 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]
  10. Mavaddat, N.; Peock, S.; Frost, D.; Ellis, S.; Platte, R.; Fineberg, E.; Evans, D.G.; Izatt, L.; Eeles, R.A.; Adlard, J.; et al. Cancer Risks for BRCA1 and BRCA2 Mutation Carriers: Results From Prospective Analysis of EMBRACE. J. Natl. Cancer Inst. 2013, 105, 812–822. [Google Scholar] [CrossRef] [Green Version]
  11. Kuchenbaecker, K.B.; Hopper, J.L.; Barnes, D.R.; Phillips, K.; Mooij, T.M.; Jervis, S.; Van Leeuwen, F.E.; Milne, R.L.; Andrieu, N.; Goldgar, D.E.; et al. Risks of Breast, Ovarian, and Contralateral Breast Cancer for BRCA1 and BRCA2 Mutation Carriers. JAMA 2017, 317, 2402–2416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Easton, D.F.; Steele, L.; Fields, P.; Ormiston, W.; Averill, D.; Daly, P.A.; McManus, R.; Neuhausen, S.L.; Ford, D.; Wooster, R.; et al. Cancer Risks in Two Large Breast Cancer Families Linked to BRCA2 on Chromosome 13q12-13. Am. J. Hum. Genet. 1997, 61, 120–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Scully, R.; Livingston, D.M. In search of the tumour-suppressor functions of BRCA1 and BRCA2. Nat. Cell Biol. 2000, 408, 429–432. [Google Scholar] [CrossRef]
  14. Venkitaraman, A.R. Cancer Susceptibility and the Functions of BRCA1 and BRCA2. Cell 2002, 108, 171–182. [Google Scholar] [CrossRef] [Green Version]
  15. Jensen, R.B.; Carreira, A.; Kowalczykowski, S.C. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 2010, 467, 678–683. [Google Scholar] [CrossRef] [Green Version]
  16. Davies, A.A.; Masson, J.-Y.; McIlwraith, M.J.; Stasiak, A.Z.; Stasiak, A.; Venkitaraman, A.R.; West, S.C. Role of BRCA2 in Control of the RAD51 Recombination and DNA Repair Protein. Mol. Cell 2001, 7, 273–282. [Google Scholar] [CrossRef]
  17. Liu, J.; Doty, T.; Gibson, B.; Heyer, W.-D. Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA. Nat. Struct. Mol. Biol. 2010, 17, 1260–1262. [Google Scholar] [CrossRef]
  18. Powell, S.N.; Willers, H.; Xia, F. Brca2 keeps rad51 in line. High-fidelity homologous recombination prevents breast and ovarian cancer? Mol. Cell 2002, 10, 1262–1263. [Google Scholar] [CrossRef]
  19. Thorslund, T.; McIlwraith, M.J.; Compton, S.A.; Lekomtsev, S.; Petronczki, M.; Griffith, J.D.; West, S.C. The breast cancer tumor suppressor BRCA2 promotes the specific targeting of RAD51 to single-stranded DNA. Nat. Struct. Mol. Biol. 2010, 17, 1263–1265. [Google Scholar] [CrossRef]
  20. Yang, H.; Li, Q.; Fan, J.; Holloman, W.K.; Pavletich, N.P. The brca2 homologue brh2 nucleates rad51 filament formation at a dsdna-ssdna junction. Nature 2005, 433, 653–657. [Google Scholar] [CrossRef]
  21. Yuan, S.S.; Lee, S.Y.; Chen, G.; Song, M.; Tomlinson, G.E.; Lee, E.Y. BRCA2 is required for ionizing radiation-induced assembly of Rad51 complex in vivo. Cancer Res. 1999, 59, 3547–3551. [Google Scholar]
  22. Eckelmann, B.J.; Bacolla, A.; Wang, H.; Ye, Z.; Guerrero, E.N.; Jiang, W.; El-Zein, R.; Hegde, M.L.; Tomkinson, A.E.; Tainer, J.A.; et al. XRCC1 promotes replication restart, nascent fork degradation and mutagenic DNA repair in BRCA2-deficient cells. NAR Cancer 2020, 2, zcaa013. [Google Scholar] [CrossRef] [PubMed]
  23. Mijic, S.; Zellweger, R.; Chappidi, N.; Berti, M.; Jacobs, K.; Mutreja, K.; Lopes, M. Replication fork reversal triggers fork degradation in BRCA2-defective cells. Nat. Commun. 2017, 8, 859. [Google Scholar] [CrossRef] [PubMed]
  24. Rickman, K.A.; Noonan, R.J.; Lach, F.; Sridhar, S.; Wang, A.; Abhyankar, A.; Huang, A.; Kelly, M.; Auerbach, A.D.; Smogorzewska, A. Distinct roles of BRCA2 in replication fork protection in response to hydroxyurea and DNA interstrand cross-links. Genes Dev. 2020, 34, 832–846. [Google Scholar] [CrossRef] [PubMed]
  25. Schlacher, K.; Christ, N.; Siaud, N.; Egashira, A.; Wu, H.; Jasin, M. Double-Strand Break Repair-Independent Role for BRCA2 in Blocking Stalled Replication Fork Degradation by MRE11. Cell 2011, 145, 529–542. [Google Scholar] [CrossRef] [Green Version]
  26. Taglialatela, A.; Alvarez, S.; Leuzzi, G.; Sannino, V.; Ranjha, L.; Huang, J.-W.; Madubata, C.; Anand, R.; Levy, B.; Rabadan, R.; et al. Restoration of Replication Fork Stability in BRCA1- and BRCA2-Deficient Cells by Inactivation of SNF2-Family Fork Remodelers. Mol. Cell 2017, 68, 414–430. [Google Scholar] [CrossRef] [Green Version]
  27. Bhatia, V.; Barroso, S.I.; García-Rubio, M.L.; Tumini, E.; Herrera-Moyano, E.; Aguilera, A. BRCA2 prevents R-loop accumulation and associates with TREX-2 mRNA export factor PCID2. Nat. Cell Biol. 2014, 511, 362–365. [Google Scholar] [CrossRef]
  28. Shivji, M.K.; Renaudin, X.; Williams, Ç.H.; Venkitaraman, A.R. BRCA2 Regulates Transcription Elongation by RNA Polymerase II to Prevent R-Loop Accumulation. Cell Rep. 2018, 22, 1031–1039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Daniels, M.J.; Wang, Y.; Lee, M.; Venkitaraman, A.R. Abnormal Cytokinesis in Cells Deficient in the Breast Cancer Susceptibility Protein BRCA2. Science 2004, 306, 876–879. [Google Scholar] [CrossRef]
  30. Mondal, G.; Rowley, M.; Guidugli, L.; Wu, J.; Pankratz, V.S.; Couch, F.J. BRCA2 Localization to the Midbody by Filamin A Regulates CEP55 Signaling and Completion of Cytokinesis. Dev. Cell 2012, 23, 137–152. [Google Scholar] [CrossRef] [Green Version]
  31. Panzarino, N.J.; Krais, J.J.; Cong, K.; Peng, M.; Mosqueda, M.; Nayak, S.U.; Bond, S.M.; Calvo, J.A.; Doshi, M.B.; Bere, M.; et al. Replication Gaps Underlie BRCA Deficiency and Therapy Response. Cancer Res. 2021, 81, 1388–1397. [Google Scholar] [CrossRef] [PubMed]
  32. Xia, B.; Sheng, Q.; Nakanishi, K.; Ohashi, A.; Wu, J.; Christ, N.; Liu, X.; Jasin, M.; Couch, F.J.; Livingston, D.M. Control of BRCA2 Cellular and Clinical Functions by a Nuclear Partner, PALB2. Mol. Cell 2006, 22, 719–729. [Google Scholar] [CrossRef]
  33. Dray, E.; Etchin, J.; Wiese, C.; Saro, D.; Williams, G.J.; Hammel, M.; Yu, X.; Galkin, V.E.; Liu, D.; Tsai, M.-S.; et al. Enhancement of RAD51 recombinase activity by the tumor suppressor PALB2. Nat. Struct. Mol. Biol. 2010, 17, 1255–1259. [Google Scholar] [CrossRef] [PubMed]
  34. Foo, T.K.; Tischkowitz, M.; Simhadri, S.; Boshari, T.; Zayed, N.; Burke, K.A.; Berman, S.H.; Blecua, P.; Riaz, N.; Huo, Y.; et al. Compromised BRCA1–PALB2 interaction is associated with breast cancer risk. Oncogene 2017, 36, 4161–4170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hughes-Davies, L.; Huntsman, D.; Ruas, M.; Fuks, F.; Bye, J.; Chin, S.-F.; Milner, J.; Brown, L.; Hsu, F.; Gilks, B.; et al. EMSY Links the BRCA2 Pathway to Sporadic Breast and Ovarian Cancer. Cell 2003, 115, 523–535. [Google Scholar] [CrossRef] [Green Version]
  36. Tischkowitz, M.; Capanu, M.; Sabbaghian, N.; Li, L.; Liang, X.; Vallée, M.P.; Tavtigian, S.V.; Concannon, P.; Foulkes, W.D.; Bernstein, L.; et al. Rare germline mutations inPALB2and breast cancer risk: A population-based study. Hum. Mutat. 2012, 33, 674–680. [Google Scholar] [CrossRef] [Green Version]
  37. Carreira, A.; Hilario, J.; Amitani, I.; Baskin, R.J.; Shivji, M.K.; Venkitaraman, A.R.; Kowalczykowski, S.C. The BRC Repeats of BRCA2 Modulate the DNA-Binding Selectivity of rad51. Cell 2009, 136, 1032–1043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Chatterjee, G.; Jimenez-Sainz, J.; Presti, T.; Nguyen, T.; Jensen, R.B. Distinct binding of BRCA2 BRC repeats to RAD51 generates differential DNA damage sensitivity. Nucleic Acids Res. 2016, 44, 5256–5270. [Google Scholar] [CrossRef] [Green Version]
  39. Carreira, A.; Kowalczykowski, S.C. Two classes of BRC repeats in BRCA2 promote RAD51 nucleoprotein filament function by distinct mechanisms. Proc. Natl. Acad. Sci. USA 2011, 108, 10448–10453. [Google Scholar] [CrossRef] [Green Version]
  40. Thorslund, T.; Esashi, F.; West, S.C. Interactions between human BRCA2 protein and the meiosis-specific recombinase DMC1. EMBO J. 2007, 26, 2915–2922. [Google Scholar] [CrossRef] [Green Version]
  41. Marston, N.J.; Richards, W.J.; Hughes, D.; Bertwistle, D.; Marshall, C.J.; Ashworth, A. Interaction between the Product of the Breast Cancer Susceptibility Gene BRCA2 and DSS1, a Protein Functionally Conserved from Yeast to Mammals. Mol. Cell. Biol. 1999, 19, 4633–4642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Zhao, W.; Vaithiyalingam, S.; Filippo, J.S.; Maranon, D.G.; Jimenez-Sainz, J.; Fontenay, G.V.; Kwon, Y.; Leung, S.G.; Lu, L.; Jensen, R.B.; et al. Promotion of BRCA2-Dependent Homologous Recombination by DSS1 via RPA Targeting and DNA Mimicry. Mol. Cell 2015, 59, 176–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Yang, H.; Jeffrey, P.D.; Miller, J.; Kinnucan, E.; Sun, Y.; Thomä, N.H.; Zheng, N.; Chen, P.-L.; Lee, W.-H.; Pavletich, N.P. BRCA2 Function in DNA Binding and Recombination from a BRCA2-DSS1-ssDNA Structure. Science 2002, 297, 1837–1848. [Google Scholar] [CrossRef] [Green Version]
  44. Chen, P.-L.; Chen, C.-F.; Chen, Y.; Xiao, J.; Sharp, Z.D.; Lee, W.-H. The BRC repeats in BRCA2 are critical for RAD51 binding and resistance to methyl methanesulfonate treatment. Proc. Natl. Acad. Sci. USA 1998, 95, 5287–5292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Mizuta, R.; LaSalle, J.; Cheng, H.-L.; Shinohara, A.; Ogawa, H.; Copeland, N.; Jenkins, N.A.; Lalande, M.; Alt, F.W. RAB22 and RAB163/mouse BRCA2: Proteins that specifically interact with the RAD51 protein. Proc. Natl. Acad. Sci. USA 1997, 94, 6927–6932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Pellegrini, L.; Yu, D.S.; Lo, T.; Anand, S.; Lee, M.; Blundell, T.L.; Venkitaraman, A.R. Insights into DNA recombination from the structure of a RAD51–BRCA2 complex. Nat. Cell Biol. 2002, 420, 287–293. [Google Scholar] [CrossRef] [PubMed]
  47. Sharan, S.K.; Morimatsu, M.; Albrecht, U.; Lim, D.-S.; Regel, E.; Dinh, C.; Sands, A.; Eichele, G.; Hasty, P.; Bradley, A. Embryonic lethality and radiation hypersensitivity mediated by Rad51 in mice lacking Brca2. Nat. Cell Biol. 1997, 386, 804–810. [Google Scholar] [CrossRef]
  48. Wong, A.K.C.; Pero, R.; Ormonde, P.A.; Tavtigian, S.V.; Bartel, P.L. RAD51 Interacts with the Evolutionarily Conserved BRC Motifs in the Human Breast Cancer Susceptibility Gene brca2. J. Biol. Chem. 1997, 272, 31941–31944. [Google Scholar] [CrossRef] [Green Version]
  49. Oliver, A.W.; Swift, S.; Lord, C.J.; Ashworth, A.; Pearl, L.H. Structural basis for recruitment of BRCA2 by PALB2. EMBO Rep. 2009, 10, 990–996. [Google Scholar] [CrossRef]
  50. Ren, J.; Wen, L.; Gao, X.; Jin, C.; Xue, Y.; Yao, X. DOG 1.0: Illustrator of protein domain structures. Cell Res. 2009, 19, 271–273. [Google Scholar] [CrossRef]
  51. Bertwistle, D.; Swift, S.; Marston, N.J.; Jackson, L.E.; Crossland, S.; Crompton, M.R.; Marshall, C.J.; Ashworth, A. Nuclear location and cell cycle regulation of the BRCA2 protein. Cancer Res. 1997, 57, 5485–5488. [Google Scholar] [PubMed]
  52. Han, X.; Saito, H.; Miki, Y.; Nakanishi, A. A CRM1-mediated nuclear export signal governs cytoplasmic localization of BRCA2 and is essential for centrosomal localization of BRCA2. Oncogene 2007, 27, 2969–2977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Spain, B.H.; Larson, C.J.; Shihabuddin, L.S.; Gage, F.H.; Verma, I.M. Truncated BRCA2 is cytoplasmic: Implications for cancer-linked mutations. Proc. Natl. Acad. Sci. USA 1999, 96, 13920–13925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yano, K.-I.; Morotomi, K.; Saito, H.; Kato, M.; Matsuo, F.; Miki, Y. Nuclear Localization Signals of the BRCA2 Protein. Biochem. Biophys. Res. Commun. 2000, 270, 171–175. [Google Scholar] [CrossRef]
  55. Jeyasekharan, A.; Liu, Y.; Hattori, H.; Pisupati, V.; Jonsdottir, A.B.; Rajendra, E.; Lee, M.; Sundaramoorthy, E.; Schlachter, S.; Kaminski, C.F.; et al. A cancer-associated BRCA2 mutation reveals masked nuclear export signals controlling localization. Nat. Struct. Mol. Biol. 2013, 20, 1191–1198. [Google Scholar] [CrossRef] [Green Version]
  56. Chen, J.J.; Silver, D.; Cantor, S.; Livingston, D.M.; Scully, R. BRCA1, BRCA2, and Rad51 operate in a common DNA damage response pathway. Cancer Res. 1999, 59, 1752–1756. [Google Scholar]
  57. Moynahan, M.E.; Pierce, A.J.; Jasin, M. BRCA2 Is Required for Homology-Directed Repair of Chromosomal Breaks. Mol. Cell 2001, 7, 263–272. [Google Scholar] [CrossRef] [Green Version]
  58. Patel, K.J.; Yu, V.P.; Lee, H.; Corcoran, A.; Thistlethwaite, F.C.; Evans, M.J.; Colledge, W.H.; Friedman, L.S.; Ponder, B.A.; Venkitaraman, A.R. Involvement of Brca2 in DNA Repair. Mol. Cell 1998, 1, 347–357. [Google Scholar] [CrossRef]
  59. Tutt, A.; Gabriel, A.; Bertwistle, D.; Connor, F.; Paterson, H.; Peacock, J.; Ross, G.; Ashworth, A. Absence of Brca2 causes genome instability by chromosome breakage and loss associated with centrosome amplification. Curr. Biol. 1999, 9, 1107–1110. [Google Scholar] [CrossRef] [Green Version]
  60. Chen, J.; Silver, D.P.; Walpita, D.; Cantor, S.B.; Gazdar, A.F.; Tomlinson, G.; Couch, F.J.; Weber, B.L.; Ashley, T.; Livingston, D.M.; et al. Stable Interaction between the Products of the BRCA1 and BRCA2 Tumor Suppressor Genes in Mitotic and Meiotic Cells. Mol. Cell 1998, 2, 317–328. [Google Scholar] [CrossRef]
  61. Hucl, T.; Rago, C.; Gallmeier, E.; Brody, J.R.; Gorospe, M.; Kern, S.E. A Syngeneic Variance Library for Functional Annotation of Human Variation: Application to BRCA2. Cancer Res. 2008, 68, 5023–5030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sakai, W.; Swisher, E.M.; Jacquemont, C.; Chandramohan, K.V.; Couch, F.J.; Langdon, S.P.; Wurz, K.; Higgins, J.; Villegas, E.; Taniguchi, T. Functional Restoration of BRCA2 Protein by Secondary BRCA2 Mutations in BRCA2-Mutated Ovarian Carcinoma. Cancer Res. 2009, 69, 6381–6386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Suzuki, A.; De La Pompa, J.L.; Hakem, R.; Elia, A.; Yoshida, R.; Mo, R.; Nishina, H.; Chuang, T.; Wakeham, A.; Itie, A.; et al. Brca2 is required for embryonic cellular proliferation in the mouse. Genes Dev. 1997, 11, 1242–1252. [Google Scholar] [CrossRef] [Green Version]
  64. Wiegant, W.W.; Overmeer, R.M.; Godthelp, B.C.; van Buul, P.P.; Zdzienicka, M.Z. Chinese hamster cell mutant, V-C8, a model for analysis of Brca2 function. Mutat. Res. Mol. Mech. Mutagen. 2006, 600, 79–88. [Google Scholar] [CrossRef]
  65. Guidugli, L.; Carreira, A.; Caputo, S.M.; Ehlen, A.; Galli, A.; Monteiro, A.N.; Neuhausen, S.L.; Hansen, T.V.; Couch, F.J.; Vreeswijk, M.P.; et al. Functional Assays for Analysis of Variants of Uncertain Significance inBRCA2. Hum. Mutat. 2014, 35, 151–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Karchin, R.; Agarwal, M.; Sali, A.; Couch, F.; Beattie, M.S. Classifying Variants of Undetermined Significance in BRCA2 with Protein Likelihood Ratios. Cancer Inform. 2008, 6, 203–216. [Google Scholar] [CrossRef] [Green Version]
  67. Abul-Husn, N.S.; Team, C.G.; Soper, E.; Odgis, J.A.; Cullina, S.; Bobo, D.; Moscati, A.; Rodriguez, J.E.; Loos, R.; Cho, J.H.; et al. Exome sequencing reveals a high prevalence of BRCA1 and BRCA2 founder variants in a diverse population-based biobank. Genome Med. 2019, 12, 2. [Google Scholar] [CrossRef]
  68. Neuhausen, S.L.; Gilewski, T.; Norton, L.; Tran, T.; McGuire, P.; Swensen, J.; Hampel, H.; Borgen, P.I.; Brown, K.L.; Skolnick, M.; et al. Recurrent BRCA2 6174delT mutations in Ashkenazi Jewish women affected by breast cancer. Nat. Genet. 1996, 13, 126–128. [Google Scholar] [CrossRef]
  69. Sarantaus, L.; Huusko, P.; Eerola, H.; Launonen, V.; Vehmanen, P.; Rapakko, K.; Gillanders, E.M.; 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. 2000, 8, 757–763. [Google Scholar] [CrossRef]
  70. Seong, M.-W.; Cho, S.; Noh, D.-Y.; Han, W.; Kim, S.-W.; Park, C.-M.; Park, H.-W.; Kim, J.Y.; Park, S.S. Comprehensive mutational analysis ofBRCA1/BRCA2for Korean breast cancer patients: Evidence of a founder mutation. Clin. Genet. 2009, 76, 152–160. [Google Scholar] [CrossRef]
  71. Maxwell, K.N.; Wubbenhorst, B.; Wenz, B.M.; De Sloover, D.; Pluta, J.; Emery, L.; Barrett, A.; Kraya, A.A.; Anastopoulos, I.N.; Yu, S.; et al. BRCA locus-specific loss of heterozygosity in germline BRCA1 and BRCA2 carriers. Nat. Commun. 2017, 8, 319. [Google Scholar] [CrossRef] [PubMed]
  72. Trevino, V. Hotspotannotations-a database for hotspot mutations and annotations in cancer. Database (Oxford) 2020, 2020, baaa025. [Google Scholar] [CrossRef] [PubMed]
  73. 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] [Green Version]
  74. Labidi-Galy, S.I.; Olivier, T.; Rodrigues, M.; Ferraioli, D.; Derbel, O.; Bodmer, A.; Petignat, P.; Rak, B.; Chopin, N.; Tredan, O.; et al. Location of Mutation in BRCA2 Gene and Survival in Patients with Ovarian Cancer. Clin. Cancer Res. 2018, 24, 326–333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Dines, J.N.; Shirts, B.H.; Slavin, T.P.; Walsh, T.; King, M.-C.; Fowler, D.M.; Pritchard, C.C.; Shirts, B.H.; Pritchard, C.C. Systematic misclassification of missense variants in BRCA1 and BRCA2 “coldspots”. Genet. Med. 2020, 22, 825–830. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pruthi, S.; Gostout, B.S.; Lindor, N.M. Identification and management of women with brca mutations or hereditary predisposition for breast and ovarian cancer. Mayo Clin. Proc. 2010, 85, 1111–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Salhab, M.; Bismohun, S.; Mokbel, K. Risk-reducing strategies for women carrying brca1/2 mutations with a focus on prophylactic surgery. BMC Womens Health 2010, 10, 28. [Google Scholar] [CrossRef] [Green Version]
  78. Evans, D.G.R.; Barwell, J.; Eccles, D.M.; Collins, A.; Izatt, L.; Jacobs, C.; Donaldson, A.; Brady, A.F.; Cuthbert, A.; Harrison, R.; et al. The Angelina Jolie effect: How high celebrity profile can have a major impact on provision of cancer related services. Breast Cancer Res. 2014, 16, 442. [Google Scholar] [CrossRef]
  79. James, P.A.; Mitchell, G.; Bogwitz, M.; Lindeman, G.J. The Angelina Jolie effect. Med. J. Aust. 2013, 199, 646. [Google Scholar] [CrossRef]
  80. Welsh, J.L.; Hoskin, T.L.; Day, C.N.; Thomas, A.S.; Cogswell, J.A.; Couch, F.J.; Boughey, J.C. Clinical Decision-Making in Patients with Variant of Uncertain Significance in BRCA1 or BRCA2 Genes. Ann. Surg. Oncol. 2017, 24, 3067–3072. [Google Scholar] [CrossRef]
  81. Domchek, S.M.; Friebel, T.M.; Singer, C.F.; Evans, D.G.; Lynch, H.T.; Isaacs, C.; Garber, J.E.; Neuhausen, S.L.; Matloff, E.; Eeles, R.; et al. Association of risk-reducing surgery in brca1 or brca2 mutation carriers with cancer risk and mortality. JAMA 2010, 304, 967–975. [Google Scholar] [CrossRef] [Green Version]
  82. Spurdle, A.B.; Healey, S.; Devereau, A.; Hogervorst, F.B.L.; Monteiro, A.N.A.; Nathanson, K.L.; Radice, P.; Stoppa-Lyonnet, D.; Tavtigian, S.; Wappenschmidt, B.; et al. ENIGMA-Evidence-based network for the interpretation of germline mutant alleles: An international initiative to evaluate risk and clinical significance associated with sequence variation in BRCA1 and BRCA2 genes. Hum. Mutat. 2012, 33, 2–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Cline, M.S.; Liao, R.G.; Parsons, M.T.; 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. 2018, 14, e1007752. [Google Scholar] [CrossRef] [Green Version]
  84. Landrum, M.J.; Chitipiralla, S.; Brown, G.R.; Chen, C.; Gu, B.; Hart, J.; Hoffman, D.; Jang, W.; Kaur, K.; Liu, C.; et al. ClinVar: Improvements to accessing data. Nucleic Acids Res. 2020, 48, D835–D844. [Google Scholar] [CrossRef]
  85. Plon, S.E.; Eccles, D.M.; Easton, D.; Foulkes, W.D.; Genuardi, M.; Greenblatt, M.S.; Hogervorst, F.B.; Hoogerbrugge, N.; Spurdle, A.B.; Tavtigian, S.V.; et al. Sequence variant classification and reporting: Recommendations for improving the interpretation of cancer susceptibility genetic test results. Hum. Mutat. 2008, 29, 1282–1291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Tavtigian, S.V.; Greenblatt, M.S.; Goldgar, D.E.; Boffetta, P.; for the IARC Unclassified Genetic Variants Working Group. Assessing pathogenicity: Overview of results from the IARC Unclassified Genetic Variants Working Group. Hum. Mutat. 2008, 29, 1261–1264. [Google Scholar] [CrossRef] [PubMed]
  87. Lindor, N.M.; Guidugli, L.; Wang, X.; Vallée, M.P.; Monteiro, A.N.A.; Tavtigian, S.; Goldgar, D.E.; Couch, F.J. A review of a multifactorial probability-based model for classification of BRCA1 and BRCA2 variants of uncertain significance (VUS). Hum. Mutat. 2012, 33, 8–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Easton, D.F.; Deffenbaugh, A.M.; Pruss, D.; Frye, C.; Wenstrup, R.J.; Allen-Brady, K.; Tavtigian, S.V.; Monteiro, A.N.; Iversen, E.S.; Couch, F.J.; et al. A Systematic Genetic Assessment of 1,433 Sequence Variants of Unknown Clinical Significance in the BRCA1 and BRCA2 Breast Cancer–Predisposition Genes. Am. J. Hum. Genet. 2007, 81, 873–883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Parmigiani, G.; Berry, D.A.; Aguilar, O. Determining Carrier Probabilities for Breast Cancer–Susceptibility Genes BRCA1 and BRCA2. Am. J. Hum. Genet. 1998, 62, 145–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Tyrer, J.; Duffy, S.W.; Cuzick, J. A breast cancer prediction model incorporating familial and personal risk factors. Stat. Med. 2004, 23, 1111–1130. [Google Scholar] [CrossRef] [Green Version]
  91. Antoniou, A.C.; Cunningham, A.P.; Peto, J.; Evans, D.G.; Lalloo, F.; Narod, S.A.; Risch, H.A.; Eyfjord, J.E.; Hopper, J.L.; Southey, M.C.; et al. The BOADICEA model of genetic susceptibility to breast and ovarian cancers: Updates and extensions. Br. J. Cancer 2008, 98, 1457–1466. [Google Scholar] [CrossRef] [PubMed]
  92. García, E.B.G.; Oosterwijk, J.C.; Timmermans, M.; van Asperen, C.J.; Hogervorst, F.B.; Hoogerbrugge, N.; Vreeswijk, M.P. A method to assess the clinical significance of unclassified variants in the brca1 and brca2 genes based on cancer family history. Breast Cancer Res. 2009, 11, R8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Spurdle, A.B.; Lakhani, S.R.; Healey, S.; Parry, S.; Da Silva, L.M.; Brinkworth, R.; Hopper, J.L.; Brown, M.A.; Babikyan, D.; Chenevix-Trench, G.; et al. Clinical Classification of BRCA1 and BRCA2 DNA Sequence Variants: The Value of Cytokeratin Profiles and Evolutionary Analysis—A Report From the kConFab Investigators. J. Clin. Oncol. 2008, 26, 1657–1663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Spearman, A.D.; Sweet, K.; Zhou, X.-P.; McLennan, J.; Couch, F.J.; Toland, A.E. Clinically Applicable Models to Characterize BRCA1 and BRCA2 Variants of Uncertain Significance. J. Clin. Oncol. 2008, 26, 5393–5400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Mohammadi, L.; Vreeswijk, M.P.; Oldenburg, R.; Ouweland, A.V.D.; Oosterwijk, J.C.; Van Der Hout, A.H.; Hoogerbrugge, N.; Ligtenberg, M.; Ausems, M.G.; Van Der Luijt, R.B.; et al. A simple method for co-segregation analysis to evaluate the pathogenicity of unclassified variants; BRCA1 and BRCA2 as an example. BMC Cancer 2009, 9, 211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Adzhubei, I.A.; Schmidt, S.; Peshkin, L.; Ramensky, V.E.; Gerasimova, A.; Bork, P.; Kondrashov, A.S.; Sunyaev, S.R. A method and server for predicting damaging missense mutations. Nat. Methods 2010, 7, 248–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kumar, P.; Henikoff, S.; Ng, P.C. Predicting the effects of coding non-synonymous variants on protein function using the SIFT algorithm. Nat. Protoc. 2009, 4, 1073–1081. [Google Scholar] [CrossRef]
  98. Mathe, E.; Olivier, M.; Kato, S.; Ishioka, C.; Hainaut, P.; Tavtigian, S.V. Computational approaches for predicting the biological effect of p53 missense mutations: A comparison of three sequence analysis based methods. Nucleic Acids Res. 2006, 34, 1317–1325. [Google Scholar] [CrossRef]
  99. Eccles, D.M.; Mitchell, G.; Monteiro, A.N.A.; Schmutzler, R.; Couch, F.J.; Spurdle, A.B.; Gómez-García, E.B.; Driessen, R.; Lindor, N.; Blok, M.; et al. BRCA1 and BRCA2 genetic testing—pitfalls and recommendations for managing variants of uncertain clinical significance. Ann. Oncol. 2015, 26, 2057–2065. [Google Scholar] [CrossRef]
  100. Tavtigian, S.V.; Byrnes, G.B.; Goldgar, D.E.; Thomas, A. Classification of rare missense substitutions, using risk surfaces, with genetic- and molecular-epidemiology applications. Hum. Mutat. 2008, 29, 1342–1354. [Google Scholar] [CrossRef] [Green Version]
  101. Goldgar, D.E.; Easton, D.F.; Byrnes, G.B.; Spurdle, A.B.; Iversen, E.S.; Greenblatt, M.S.; for the IARC Unclassified Genetic Variants Working Group. Genetic evidence and integration of various data sources for classifying uncertain variants into a single model. Hum. Mutat. 2008, 29, 1265–1272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. 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] [PubMed]
  103. Lord, C.J.; Ashworth, A. Mechanisms of resistance to therapies targeting BRCA-mutant cancers. Nat. Med. 2013, 19, 1381–1388. [Google Scholar] [CrossRef] [PubMed]
  104. Lord, C.J.; Ashworth, A. BRCAness revisited. Nat. Rev. Cancer 2016, 16, 110–120. [Google Scholar] [CrossRef]
  105. Biswas, K.; Das, R.; Alter, B.P.; Kuznetsov, S.G.; Stauffer, S.; North, S.L.; Burkett, S.; Brody, L.C.; Meyer, S.; Byrd, R.A.; et al. A comprehensive functional characterization of BRCA2 variants associated with Fanconi anemia using mouse ES cell–based assay. Blood 2011, 118, 2430–2442. [Google Scholar] [CrossRef] [Green Version]
  106. Biswas, K.; Das, R.; Eggington, J.M.; Qiao, H.; North, S.L.; Stauffer, S.; Burkett, S.; Martin, B.K.; Southon, E.; Sizemore, S.C.; et al. Functional evaluation of BRCA2 variants mapping to the PALB2-binding and C-terminal DNA-binding domains using a mouse ES cell-based assay. Hum. Mol. Genet. 2012, 21, 3993–4006. [Google Scholar] [CrossRef] [Green Version]
  107. Biswas, K.; Lipton, G.B.; Stauffer, S.; Sullivan, T.; Cleveland, L.; Southon, E.; Reid, S.; Magidson, V.; Edwin, S.I., Jr.; Sharan, S.K. A computational model for classification of BRCA2 variants using mouse embryonic stem cell-based functional assays. NPJ Genom. Med. 2020, 5, 52. [Google Scholar] [CrossRef]
  108. Carvalho, M.; Couch, F.J.; Monteiro, A.N. Functional assays for BRCA1 and BRCA2. Int. J. Biochem. Cell Biol. 2007, 39, 298–310. [Google Scholar] [CrossRef] [Green Version]
  109. Couch, F.J.; Rasmussen, L.J.; Hofstra, R.; Monteiro, A.N.; Greenblatt, M.S.; de Wind, N. Iarc Unclassified Genetic for the IARC Unclassified Genetic Variants Working Group Assessment of functional effects of unclassified genetic variants. Hum. Mutat. 2008, 29, 1314–1326. [Google Scholar] [CrossRef] [Green Version]
  110. Farrugia, D.J.; Agarwal, M.K.; Pankratz, V.S.; Deffenbaugh, A.M.; Pruss, D.; Frye, C.; Wadum, L.; Johnson, K.; Mentlick, J.; Tavtigian, S.V.; et al. Functional Assays for Classification of BRCA2 Variants of Uncertain Significance. Cancer Res. 2008, 68, 3523–3531. [Google Scholar] [CrossRef] [Green Version]
  111. Kuznetsov, S.G.; Liu, P.; Sharan, S.K. Mouse embryonic stem cell–based functional assay to evaluate mutations in BRCA2. Nat. Med. 2008, 14, 875–881. [Google Scholar] [CrossRef] [PubMed]
  112. Guidugli, L.; Pankratz, V.S.; Singh, N.; Thompson, J.; Erding, C.A.; Engel, C.; Schmutzler, R.; Domchek, S.; Nathanson, K.; Radice, P.; et al. A Classification Model for BRCA2 DNA Binding Domain Missense Variants Based on Homology-Directed Repair Activity. Cancer Res. 2012, 73, 265–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Guidugli, L.; Shimelis, H.; Masica, D.L.; Pankratz, V.S.; Lipton, G.B.; Singh, N.; Hu, C.; Monteiro, A.N.A.; Lindor, N.M.; Goldgar, D.E.; et al. Assessment of the clinical relevance of brca2 missense variants by functional and computational approaches. Am. J. Hum. Genet 2018, 102, 233–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Hendriks, G.; Morolli, B.; Calléja, F.M.; Plomp, A.; Mesman, R.L.; Meijers, M.; Sharan, S.K.; Vreeswijk, M.P.; Vrieling, H. An efficient pipeline for the generation and functional analysis of human BRCA2 variants of uncertain significance. Hum. Mutat. 2014, 35, 1382–1391. [Google Scholar] [CrossRef] [PubMed]
  115. Mesman, R.; Msc, F.M.G.R.C.; Hendriks, G.; Morolli, B.; Misovic, B.; Devilee, P.; Van Asperen, C.J.; Vrieling, H.; Vreeswijk, M.P.G. The functional impact of variants of uncertain significance in BRCA2. Genet. Med. 2019, 21, 293–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Richardson, M.E.; Hu, C.; Lee, K.Y.; LaDuca, H.; Fulk, K.; Durda, K.M.; Deckman, A.M.; Goldgar, D.E.; Monteiro, A.N.; Gnanaolivu, R.; et al. Strong functional data for pathogenicity or neutrality classify BRCA2 DNA-binding-domain variants of uncertain significance. Am. J. Hum. Genet. 2021, 108, 458–468. [Google Scholar] [CrossRef]
  117. Shimelis, H.; Mesman, R.; Von Nicolai, C.; Ehlen, A.; Guidugli, L.; Martin, C.; Calléja, F.M.; Meeks, H.; Hallberg, E.; Hinton, J.; et al. BRCA2 Hypomorphic Missense Variants Confer Moderate Risks of Breast Cancer. Cancer Res. 2017, 77, 2789–2799. [Google Scholar] [CrossRef] [Green Version]
  118. Balìa, C.; Galli, A.; Caligo, M.A. Effect of the overexpression of BRCA2 unclassified missense variants on spontaneous homologous recombination in human cells. Breast Cancer Res. Treat. 2011, 129, 1001–1009. [Google Scholar] [CrossRef]
  119. Nakanishi, A.; Han, X.; Saito, H.; Taguchi, K.; Ohta, Y.; Imajoh-Ohmi, S.; Miki, Y. Interference with BRCA2, which localizes to the centrosome during S and early M phase, leads to abnormal nuclear division. Biochem. Biophys. Res. Commun. 2007, 355, 34–40. [Google Scholar] [CrossRef]
  120. Spugnesi, L.; Balia, C.; Collavoli, A.; Falaschi, E.; Quercioli, V.; Caligo, M.A.; Galli, A. Effect of the expression of BRCA2 on spontaneous homologous recombination and DNA damage-induced nuclear foci in Saccharomyces cerevisiae. Mutagenesis 2013, 28, 187–195. [Google Scholar] [CrossRef] [Green Version]
  121. Thomassen, M.; Blanco, A.; Montagna, M.; Hansen, T.V.O.; Pedersen, I.S.; Gutiérrez-Enríquez, S.; Menéndez, M.; Fachal, L.; Santamariña, M.; Steffensen, A.Y.; et al. Characterization of BRCA1 and BRCA2 splicing variants: A collaborative report by ENIGMA consortium members. Breast Cancer Res. Treat. 2011, 132, 1009–1023. [Google Scholar] [CrossRef] [PubMed]
  122. Wu, K.; Hinson, S.R.; Ohashi, A.; Farrugia, D.; Wendt, P.; Tavtigian, S.V.; Deffenbaugh, A.; Goldgar, D.; Couch, F.J. Functional evaluation and cancer risk assessment of BRCA2 unclassified variants. Cancer Res. 2005, 65, 417–426. [Google Scholar] [PubMed]
  123. Fraile-Bethencourt, E.; Díez-Gómez, B.; Velásquez-Zapata, V.; Acedo, A.; Sanz, D.J.; Velasco, E.A. Functional classification of DNA variants by hybrid minigenes: Identification of 30 spliceogenic variants of BRCA2 exons 17 and 18. PLoS Genet. 2017, 13, e1006691. [Google Scholar] [CrossRef] [PubMed]
  124. Fraile-Bethencourt, E.; Valenzuela-Palomo, A.; Díez-Gómez, B.; Acedo, A.; Velasco, E.A. Identification of Eight Spliceogenic Variants in BRCA2 Exon 16 by Minigene Assays. Front. Genet. 2018, 9, 188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Sanz, D.J.; Acedo, A.; Infante, M.; Durán, M.; Pérez-Cabornero, L.; Esteban-Cardeñosa, E.; Lastra, E.; Pagani, F.; Miner, C.; Velasco, E.A. A High Proportion of DNA Variants of BRCA1 and BRCA2 Is Associated with Aberrant Splicing in Breast/Ovarian Cancer Patients. Clin. Cancer Res. 2010, 16, 1957–1967. [Google Scholar] [CrossRef] [Green Version]
  126. Wangensteen, T.; Felde, C.N.; Ahmed, D.; Mæhle, L.; Ariansen, S.L. Diagnostic mRNA splicing assay for variants in BRCA1 and BRCA2 identified two novel pathogenic splicing aberrations. Hered. Cancer Clin. Pr. 2019, 17, 14. [Google Scholar] [CrossRef] [PubMed]
  127. Ikegami, M.; Kohsaka, S.; Ueno, T.; Momozawa, Y.; Inoue, S.; Tamura, K.; Shimomura, A.; Hosoya, N.; Kobayashi, H.; Tanaka, S.; et al. High-throughput functional evaluation of BRCA2 variants of unknown significance. Nat. Commun. 2020, 11, 2573. [Google Scholar] [CrossRef]
  128. Findlay, G.M.; Daza, R.M.; Martin, B.; Zhang, M.D.; Leith, A.P.; Gasperini, M.; Janizek, J.; Huang, X.; Starita, L.M.; Shendure, J. Accurate classification of BRCA1 variants with saturation genome editing. Nat. Cell Biol. 2018, 562, 217–222. [Google Scholar] [CrossRef]
Figure 1. Percentages of lifetime breast and ovarian cancer risks in BRCA germline carriers.
Figure 1. Percentages of lifetime breast and ovarian cancer risks in BRCA germline carriers.
Genes 12 00780 g001
Figure 2. Schematic of BRCA2 domains. DBD: DNA binding domain; CTD: C-terminal domain. DOG illustrator of protein domain structure [50].
Figure 2. Schematic of BRCA2 domains. DBD: DNA binding domain; CTD: C-terminal domain. DOG illustrator of protein domain structure [50].
Genes 12 00780 g002
Figure 3. Pie charts representing the percentages of BRCA2 germline and somatic variants grouped into clinical significance (left) or molecular consequences (right) as reported in the ClinVar database (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/clinvar/?term=BRCA2%5Bgene%5D, accessed on 10 April 2021).
Figure 3. Pie charts representing the percentages of BRCA2 germline and somatic variants grouped into clinical significance (left) or molecular consequences (right) as reported in the ClinVar database (https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/clinvar/?term=BRCA2%5Bgene%5D, accessed on 10 April 2021).
Genes 12 00780 g003
Figure 4. cBioPortal for Cancer Genomics display of the (A) percentage of BRCA2 germline and somatic alterations and (B) spectrum of BRCA2 germline and somatic mutations in the different protein domains. The curated set of non-redundant studies of cBioPortal for Cancer Genomics was used. Combined study with 48,081 samples and 184 studies. Graphs display BRCA2 somatic mutation frequency of 3% and BRCA2 germline mutation frequency of <0.1.
Figure 4. cBioPortal for Cancer Genomics display of the (A) percentage of BRCA2 germline and somatic alterations and (B) spectrum of BRCA2 germline and somatic mutations in the different protein domains. The curated set of non-redundant studies of cBioPortal for Cancer Genomics was used. Combined study with 48,081 samples and 184 studies. Graphs display BRCA2 somatic mutation frequency of 3% and BRCA2 germline mutation frequency of <0.1.
Genes 12 00780 g004
Figure 5. Pie charts representing the percentages of BRCA2 germline and somatic genetic variants grouped into driver mutations (left) and VUS (right) as reported in cBioPortal for Cancer Genomics database (https://www.cbioportal.org/results/cancerTypesSummary?tab_index=tab_visualize&Action=Submit&session_id=607f022de4b0242bd5d49461, accessed on 10 April 2021). The curated set of non-redundant studies of cBioPortal for Cancer Genomics was used. A total of 48,081 samples and 184 studies. The total number of driver mutations and variants of uncertain significance (1934) corresponds to 3% of the samples with BRCA2 somatic mutations and <0.1% of the samples with BRCA2 germline mutations.
Figure 5. Pie charts representing the percentages of BRCA2 germline and somatic genetic variants grouped into driver mutations (left) and VUS (right) as reported in cBioPortal for Cancer Genomics database (https://www.cbioportal.org/results/cancerTypesSummary?tab_index=tab_visualize&Action=Submit&session_id=607f022de4b0242bd5d49461, accessed on 10 April 2021). The curated set of non-redundant studies of cBioPortal for Cancer Genomics was used. A total of 48,081 samples and 184 studies. The total number of driver mutations and variants of uncertain significance (1934) corresponds to 3% of the samples with BRCA2 somatic mutations and <0.1% of the samples with BRCA2 germline mutations.
Genes 12 00780 g005
Figure 6. BRCA2 Management Strategies.
Figure 6. BRCA2 Management Strategies.
Genes 12 00780 g006
Figure 7. IARC proposed a classification system for BRCA2 sequence variants identified by genetic testing.
Figure 7. IARC proposed a classification system for BRCA2 sequence variants identified by genetic testing.
Genes 12 00780 g007
Figure 8. Workflow to analyze BRCA2 variants of uncertain significance.
Figure 8. Workflow to analyze BRCA2 variants of uncertain significance.
Genes 12 00780 g008
Table 1. International groups investigating BRCA mutations in various populations.
Table 1. International groups investigating BRCA mutations in various populations.
Study/ConsortiumDescription
CIMBAConsortium of investigators of modifiers of BRCA1/2 was established in 2005
PROSEConsortium in the prevention and observation of surgical endpoints [81]
ENIGMAInternational consortium of investigators to determine the clinical significance of sequence variants in BRCA1, BRCA2 and other known or suspected breast cancer genes (www.enigmaconsortium.org, accessed on 15 April 2021) [82]
EMBRACEStudy aims to create a register of BRCA1 and BRCA2 families with defects in other genes to find out the cancer risk in these associations [10]
IBCCSInternational BRCA1/2 carrier cohort study: purpose, rationale, and study design (https://0-breast--cancer--research-biomedcentral-com.brum.beds.ac.uk/articles/10.1186/bcr93, accessed on 15 April 2021)
Breast Cancer Linkage ConsortiumInternational data sharing platform established in 1989 to collect families with breast cancer by linkage analysis
PROMPTProspective Registry of Multiplex Testing for patients and their families (www.promptstudy.org, accessed on 15 April 2021)
Table 2. Clinical Significance Databases with BRCA2 mutations (Updated April 2021).
Table 2. Clinical Significance Databases with BRCA2 mutations (Updated April 2021).
BRCA2 DatabasesTotal VariantsVariants of Uncertain SignificanceMissense VariantsPathogenic Variants
Breast Cancer Information Core (BIC)
(https://research.nhgri.nih.gov/bic/, accessed on 15 April 2021)
14,914 Genes 12 00780 i0017156 (48%) Genes 12 00780 i001
Leiden Open Variant Database (LOVD)
(https://www.lovd.nl, accessed on 15 April 2021)
920 Genes 12 00780 i001871 (94.6%) Genes 12 00780 i001
Catalogue of Somatic Mutation in Cancer (COSMIC)
(https://cancer.sanger.ac.uk/cosmic, accessed on 15 April 2021)
2729 Genes 12 00780 i0011548 (56.72%) Genes 12 00780 i001
BRCA Exchange
(https://brcaexchange.org, accessed on 15 April 2021)
20,75116,649 (83%) Genes 12 00780 i0012672 (13%)
ClinVar
(https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/clinvar/, accessed on 15 April 2021)
13,1095381 (41%)5756 (44%)3780 (29%)
Genes 12 00780 i001 Data unavailable.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jimenez-Sainz, J.; Jensen, R.B. Imprecise Medicine: BRCA2 Variants of Uncertain Significance (VUS), the Challenges and Benefits to Integrate a Functional Assay Workflow with Clinical Decision Rules. Genes 2021, 12, 780. https://0-doi-org.brum.beds.ac.uk/10.3390/genes12050780

AMA Style

Jimenez-Sainz J, Jensen RB. Imprecise Medicine: BRCA2 Variants of Uncertain Significance (VUS), the Challenges and Benefits to Integrate a Functional Assay Workflow with Clinical Decision Rules. Genes. 2021; 12(5):780. https://0-doi-org.brum.beds.ac.uk/10.3390/genes12050780

Chicago/Turabian Style

Jimenez-Sainz, Judit, and Ryan B. Jensen. 2021. "Imprecise Medicine: BRCA2 Variants of Uncertain Significance (VUS), the Challenges and Benefits to Integrate a Functional Assay Workflow with Clinical Decision Rules" Genes 12, no. 5: 780. https://0-doi-org.brum.beds.ac.uk/10.3390/genes12050780

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