10 1053@j Seminoncol 2016 10 001

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Seminars in Oncology ] (2016) ]]]–]]]

Contents lists available at ScienceDirect

Seminars in Oncology
journal homepage: www.elsevier.com/locate/ysonc

Updates on breast cancer genetics: Clinical implications of detecting


syndromes of inherited increased susceptibility to breast cancer
Erin F. Cobaina,b, Kara J. Millirona,b, Sofia D. Merajvera,b,c,n
a
Department of Internal Medicine, University of Michigan Medical School, Ann Arbor, MI
b
University of Michigan Comprehensive Cancer Center, Ann Arbor, MI
c
Department of Epidemiology, University of Michigan School of Public Health , Ann Arbor, MI

a r t i c l e in fo a b s t r a c t

Since the initial discovery that pathogenic germline alterations in BRCA 1/2 increase susceptibility to
Keywords: breast and ovarian cancer, many additional genes have now been discovered that also increase breast
Hereditary breast cancer cancer risk. Given that several more genes have now been implicated in hereditary breast cancer
Genetic testing syndromes, there is increased clinical use of multigene panel testing to evaluate patients with a
Panel testing suspected genetic predisposition to breast cancer. While this is most certainly a cost-effective approach,
Variant of uncertain significance broader testing strategies have resulted in a higher likelihood of identifying moderate-penetrance genes,
BRCA1/2 for which management guidelines regarding breast cancer risk reduction have not been firmly
TP53
established. In addition, the testing of more genes has led to increased detection of variants of uncertain
significance. We review the current knowledge regarding both high- and moderate-risk hereditary breast
cancer syndromes, as well as additional genes implicated in hereditary breast cancer for which there is
limited data. Furthermore, strategies for cancer risk reduction in mutation carriers as well as therapeutic
implications for those patients who harbor pathogenic germline alterations are discussed.
& 2016 Elsevier Inc. All rights reserved.

1. Introduction mutations [7]. Interestingly, whereas somatic de novo mutations


were identified in sporadic ovarian cancers [8] they were relatively
1.1. History of BRCA1 and BRCA2 gene mapping and discovery rare in breast cancers, suggesting that BRCA1 differed from
classic tumor suppressors such as WT1 or TP53 in its spectrum
In 1990, Mary-Claire King and colleagues reported genetic of somatic changes. Finding germline mutations in families affected
linkage studies in families with early-onset breast cancer which with breast and ovarian cancer set the stage for the integration of
implicated a relatively narrow region in chromosome 17q [1]. The genetic counseling and predisposition testing into clinical care for
previous year, Narod and colleagues had identified linkage in the cancer risk management of high-risk patients and families [9].
same region in families with breast and ovarian cancer [2]. These
seminal reports spurred a competitive race between several
international collaborations to isolate the BRCA1 gene, which was 1.2. Early efforts to incorporate genetic counseling and testing into
finally cloned in 1994 [3] and mutations were identified in tumors the clinic
belonging to the linked kindreds [4]. BRCA1, and BRCA2 isolated in
1995 [5], are essential to the process of repairing double strand The discovery of the breast cancer susceptibility genes BRCA1
breaks in DNA via homologous recombination [6]. In this form of and BRCA2 made predictive genetic testing possible. Academic
repair, the DNA is restored to its original sequence; however, when centers initially took the lead in offering integrated risk evalua-
this process is deficient, other forms of less conservative DNA tions, including counseling before and after genetic testing and
repair take place. One such mechanism is non-homologous end personalized risk management plans tailored to patients’ age,
joining, which simply joins two ends of broken DNA without using health status, reproductive plans, and cultural background. Since
a template sequencing, often resulting in the introduction of then, genetic risk assessment for breast cancer risk has been
integrated into clinical care in a number of different clinical
settings (primary care, obstetrics and gynecology, oncology, and
n
screening mammography) and the emergence of patient advocacy
Corresponding author. Sofia D. Merajver, Division of Hematology/Oncology,
Molecular Medicine & Genetics, Department of Internal Medicine, Room 7303
groups and increasing interest about genetic medicine on the part
Cancer Center , 1500 E Medical Center Dr, Ann Arbor, MI, 48109. of patients and their providers has significantly increased referrals
E-mail address: smerajve@umich.edu (S.D. Merajver). for genetic risk evaluation.

http://dx.doi.org/10.1053/j.seminoncol.2016.10.001
0093-7754/& 2016 Elsevier Inc. All rights reserved.
2 E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]]

2. Hereditary breast cancer syndromes autosomal dominant fashion. Individuals with this syndrome are
at very high risk of developing multiple malignancies in childhood
Most women who present with breast cancer do not have an or early adulthood, including but not limited to: breast cancer,
inherited predisposition to developing the disease. However, up to sarcomas, brain cancer, leukemias, lung cancers, and adrenocort-
20% of patients with a family history of breast cancer may harbor a ical cancers [24,25]. For females, the risk of developing breast
germline mutation associated with increased cancer risk [10]. cancer by the age of 60 is approximately 50% and the median age
Inherited susceptibility to developing breast cancer most com- at diagnosis is under 35 years [25–28]. Breast cancer arising in
monly results from mutations in BRCA1 or BRCA2; however, addi- women with a germline TP53 mutation is more likely to have
tional genes have been implicated in genetic predisposition to amplification of the human epidermal growth factor receptor 2
breast cancer. Below, these syndromes are classified into high-risk (HER2) [29]. In general, mastectomy is usually recommended for
and moderate-risk of developing breast cancer (Table 1). surgical management of breast cancers in patients with LFS, as
radiation exposure can markedly increase the risk for developing a
2.1. High-risk syndromes secondary radiation-associated malignancy.

2.1.1. BRCA1- and BRCA2-associated hereditary breast cancer 2.1.3. Peutz-Jeghers syndrome
Approximately 40% of hereditary breast and ovarian cancer (HBOC) Peutz-Jeghers Syndrome (PJS) results from germline alterations
cases are associated with pathogenic germline alterations in BRCA1 or in the STK11 gene and is inherited in an autosomal dominant
BRCA2, which are inherited in an autosomal dominant fashion. fashion. Patients with PJS often present with intestinal obstruc-
Families that harbor these pathogenic mutations are often character- tions as a result of small bowel polyps; however, they also have
ized by multiple females affected by breast cancer at an early age very elevated risks of gastrointestinal cancers [30], and women are
(premenopausal). While tumors associated with germline alterations at increased risk for breast and ovarian cancer. Lifetime risk for
in BRCA1 or BRCA2 are more likely to be triple negative [11], other breast cancer is approximately 55% and median age at diagnosis is
histologic subtypes can be seen. Estimates of lifetime risk of develop- in the late 30s [30,31]. Mucocutaneous pigmentation and or
ing breast cancer for patients with BRCA1 and BRCA2 mutations varies gastrointestinal hamartomatous polyps are noted in 95% of indi-
across studies; however, pooled results suggest cumulative risk of viduals with PJS [30].
developing breast cancer by age 70 in BRCA1 mutations carriers is
approximately 65% compared to 45% in BRCA2 mutation carriers [12].
In addition to a higher lifetime risk of developing breast cancer with a 2.1.4. PTEN hamartoma tumor syndrome (Cowden syndrome)
BRCA1 mutation, these patients also tend to have earlier onset of PTEN hamartoma tumor syndrome (PTHS) results from germ-
disease, particularly before age of 50 [13–15]. BRCA1 and BRCA2 line alterations in the tumor-suppressor gene PTEN. For patients
germline mutations are also found at higher frequency among certain with this condition, the lifetime risk of developing breast cancer is
ethnic groups. For example, approximately 1 in 40 individuals of approximately 85% [32]. Most women are diagnosed with preme-
Ashkenazi Jewish decent unselected for personal or family history of nopausal breast cancer [33] and these patients also have increased
malignancy carry one of three founder mutations: 187delAG or risk of benign breast diseases such as fibroadenomas [34]. Women
5385insC in BRCA1 or 6174delT in BRCA2 [16]. Founder mutations in with this syndrome also have an increased risk for endometrial
BRCA1 and BRCA2 have also been reported in the Netherlands, Sweden, cancer. Men and women with a PTEN gene mutation also have an
Hungary, Iceland, Italy, France, South Africa, Pakistan, Asia, French increased risk for thyroid cancer (particularly follicular type), renal
Canadians, Hispanics, and African Americans [12,17–21]. Men with cancer, and colon cancer. Additional non-cancer manifestations
BRCA2 mutations have an approximately 7%–8% lifetime risk of include autism spectrum disorder, macrocephaly, hamartomas,
developing breast cancer (a risk comparable to women who are at trichilemmomas, and plantar keratosis.
lower than average risk) versus 1% lifetime risk for those with BRCA1
mutations [22,23]. 2.1.5. Hereditary diffuse gastric cancer syndrome
This syndrome is characterized by genetic susceptibility to
2.1.2. Li-Fraumeni syndrome diffuse gastric cancer, associated with germline cadherin-1
Li-Fraumeni Syndrome (LFS) is associated with germline alter- (CDH1) mutations [35,36]; however, CDH1 mutations are also
ations in the tumor-suppressor gene TP53 and is inherited in an associated with increased risk of lobular breast cancer in women,

Table 1
Summary of high- and moderate-penetrance genes implicated in genetic predisposition to breast cancer.

Germline alteration Lifetime breast cancer risk Most common associated Other associated malignancies
breast cancer pathology

BRCA1 65% (by age 70) Triple Negative Ovarian, fallopian tube, primary peritoneal, pancreas, prostate
BRCA2 45% (by age 70) Triple-negative or estrogen Ovarian, melanoma, fallopian tube, primary peritoneal,
receptor–positive pancreas, prostate. stomach, gallbladder/biliary
TP53 50% (by age 60) Her2-amplified Sarcoma, CNS malignancies, adrenocortical carcinoma,
gastrointestinal, radiation-associated cancers
STK11 30%–50% (lifetime) N/A Colorectal, stomach, small bowel, pancreas, ovarian, cervical,
Sertoli cell testicular
PTEN 85% (lifetime) N/A Thyroid, endometrial, colorectal, renal
CDH1 60% (lifetime) Invasive lobular carcinoma Stomach (diffuse)
MSH1, MSH2, MLH1, Unknown N/A Colorectal, endometrial, ovary, stomach, small bowel,
PMS2, EPCAM hepatobiliary, CNS
CHEK2 20%–25% (lifetime) Estrogen receptor–positive Colorectal (1100delC mutation), stomach, prostate, kidney,
thyroid, sarcoma
PALB2 33% (by age 70 without family history), 58% (by N/A Pancreas
age 70 with family history)
ATM 20% (lifetime) N/A Pancreas
E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]] 3

with a lifetime risk of developing breast cancer of approximately are heterozygous carriers of deleterious ATM gene mutations
60% [35–37]. Of note, germline CDH1 mutations have been appear to have a risk of breast cancer that is approximately
identified in patients with invasive lobular breast cancer in the twofold higher than that of the general population [57–59]. ATM
absence of diffuse gastric cancer, implying that development of germline mutations are relatively common in the general popula-
gastric cancer is not an obligatory component of the syndrome tion, with approximately 3% of Caucasians in the United States are
[38]. While guidelines suggest that all patients with a confirmed ATM heterozygotes [60].
pathogenic CDH1 mutation be counseled regarding prophylactic
gastrectomy to mitigate risk of developing diffuse gastric cancer, 2.3. Additional genes implicated in hereditary breast cancer
whether this recommendation should be modified for individuals
with no family history of gastric cancer remains controversial. It Several other genes have been identified that may increase the
should be noted, however, that individuals who present with risk of developing breast cancer; however, additional studies are
lobular breast cancer may harbor a de novo mutation, and lack needed to better characterize their specific risk. These genes are
of family history does not necessarily indicate a low-penetrance being more frequently identified in clinical practice, particularly
mutation [39,40]. with the increase in use of multigene panel testing. Kurian and
colleagues noted in a study of 198 women referred for clinical
2.1.6. Lynch syndrome BRCA1/2 testing that approximately 11% of patients who tested
Lynch syndrome, otherwise known as hereditary nonpolyposis negative for mutations in BRCA1/2 tested positive for pathogenic
colon cancer (HNPCC), results from germline alterations in genes variants in other genes [61]. Below is a brief summary of the
involving mismatch repair (MSH2, MLH1, MSH6, and PMS2) or a available data regarding genes that have been implicated in
mutation in the epithelial cell adhesion molecule (EPCAM) gene hereditary breast cancer.
[41,42]. While risks for colon, endometrial, ovarian and stomach
cancers are markedly increased, some studies have suggested risk 2.3.1. BRCA1-associated RING domain protein (BARD1)
for female breast cancer may also be higher [41,43,44]; however, BARD1 shares strong structural homology with BRCA1, and has
the magnitude of risk may vary by the specific gene affected [45]. been demonstrated to be involved with the cellular DNA repair
process [62]. Several studies have identified germline BARD1
2.2. Moderate-risk syndromes mutations resulting in protein truncation in breast cancer families
that do not harbor germline alterations in BRCA1/2 [63–65]. As
2.2.1. CHEK2 these reported germline alterations occur at low frequency even in
The checkpoint kinase 2 gene (CHEK2) is involved in DNA high-risk populations, information regarding specific breast cancer
damage repair via the Fanconi anemia (FA)–BRCA pathway [46]. risk and penetrance is still lacking.
There are many variants of CHEK2 that have been identified;
however, the 1100delC polymorphism has been associated with a 2.3.2. BRCA1 interacting protein C-terminal helicase 1 (BRIP1)
two- to threefold increased risk of breast cancer, particularly in BRIP1 encodes a helicase-like protein that directly binds to
Caucasian women [47–50]. The lifetime risk of breast cancer is BRCA1 and is involved in homologous recombination [66]. Inher-
approximately 20%–25% among CHEK2 mutations carriers who itance of two pathogenic germline alterations in BRIP1 has also
have a positive family history of breast cancer [51,52]. When been demonstrated to be causative of FA [67]. Several studies have
compared with breast cancer patients that do not harbor a germ- reported that truncating germline BRIP1 mutations confer
line alteration in CHEK2, mutation carriers are more likely to be increased susceptibility to breast cancer [68–71]; however, a
younger at diagnosis, have a family history of breast cancer, have recently reported large study from Europe reported that genotyp-
estrogen receptor–positive breast cancer, develop a second pri- ing from over 48,000 breast cancer cases and 43,000 controls
mary breast cancer, and appear to have higher risk of cancer- found the number of truncating variants of BRIP1 to be similar
specific death [50]. Male CHEK2 carriers also appear to have a between the two groups, suggesting that germline BRIP1 muta-
higher risk of developing male breast cancer [53]. Chek2 gene tions may not substantially increase breast cancer risk [72]. BRIP1
mutations also cause an increased risk for thyroid cancer, colon gene mutations may also cause an increased risk for developing
cancer, and perhaps ovarian cancer. ovarian cancer, although the exact risk data is in flux.

2.2.2. PALB2 2.3.3. Mre11 (MRN) complex (includes MRE11, RAD50, and NBS1
Partner and localizer of BRCA2 (PALB2) is a gene that is integral genes)
in mediating the BRCA2 DNA damage response. Biallelic germline In response to DNA double strand breaks, the MRN complex
mutations in PALB2 have been identified in a subset of patients genes sense DNA damage and aid in the recruitment of the DNA
with of FA and monoallelic deleterious germline alterations in repair machinery [73]. In a study that mutation screened the MRN
PALB2 have been identified in approximately 2% of familial breast genes in 1,313 early-onset breast cancer cases and 1,123 heathy
cancer cases [54]. Carriers of PALB2 germline mutations have a risk controls, germline alterations in the MRN genes occurred at higher
of developing breast cancer that is similar to patients with frequency in those with breast cancer, suggesting that these may
pathogenic BRCA2 alterations [55,56], although risk for those be intermediate-risk breast cancer susceptibility genes [74].
patients who also have a positive family history of breast cancer
may be higher [54]. 2.3.4. RAD51 paralogs (RAD 51 C and D)
The RAD51C and D genes encodes proteins that are essential for
2.2.3. ATM homologous recombination repair. Similar to BRIP1 above, biallelic
Ataxia telangiectasia mutated (ATM) is involved in DNA repair. inheritance of missence mutations in RAD51 can cause a FA-like
While carriers of biallelic germline mutations develop the autoso- syndrome [75]. In a study reported by Meindl and colleagues, six
mal recessive disorder ataxia-telangiectasia (AT) [57], monoallelic monoallelic pathogenic mutations were identified in RAD51C in
germline mutations have been identified at increased frequency more than 1,000 German families with gynecologic malignancies,
among patients with various cancer types including pancreatic and suggesting that these mutations confer an increased risk for breast
breast cancer, implicating it in cancer susceptibility. Females who and ovarian cancer [76]. RAD51D was first implicated as a potential
4 E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]]

risk gene for the development of ovarian cancer [77]; however, its 3.3. Options for genetic testing (single gene versus multigene panel
correlation with breast cancer risk is less clear. A recent case report testing)
featured identification of a pathogenic RAD51D germline mutation
in a young female with triple-negative breast cancer and family The use of genetic testing in suspected HBOC has expanded
history of male breast cancer [78]. well beyond evaluating for mutations in BRCA1/2 since these tests
became commercially available in the 1990s. Next-generation
sequencing technologies (NGS) have made it possible to sequence
2.3.5. MUTYH multiple genes simultaneously at a cost that is often lower
The MUTYH gene is an important component of the DNA base compared to Sanger sequencing for single genes. Currently, many
excision repair mechanism. Bi-allelic MUTYH mutation carriers have a patients undergoing evaluation for HBOC ultimately have multi-
28-fold increase in colorectal cancer risk, while mono-allelic carriers gene panel testing. Selection of patients appropriate for single
have a moderately increased colorectal cancer risk [79]. Mono-allelic gene (ie, BRCA 1/2 sequencing only) versus multigene panel testing
MUTYH mutations are relatively common (carrier rate is 1% in the in suspected HBOC is currently a topic of debate within the field.
population) and it remains controversial whether or these confer an Many laboratories are marketing multigene panel testing as a
increased risk of breast cancer, although studies have reported “more is better” approach; however, paucity of data regarding
increased risk in both bi-allelic [80] and mono-allelic [81] carriers. cancer risks and lack of specific management recommendations
Case control studies, however, have failed to demonstrate a clear for carriers of moderate penetrance genes has introduced numer-
association between mono-allelic MUTYH carriers and increased ous challenges when counseling patients [84]. Furthermore,
susceptibility to breast cancer [82]. increased identification of variants of undetermined significance
(VUS) can increase anxiety among patients.
3. Strategies for genetic testing for genetic predisposition to It is important to consider, however, the potential advantages of
breast cancer multigene panel tests. Aside from being cost-effective, panel tests
are particularly helpful if there are features in the family history
3.1. Criteria for breast cancer genetic risk evaluation suggestive of multiple hereditary cancer syndromes. For example,
for patients diagnosed with breast cancer at age o 40, the differ-
Guidelines for genetic counseling and possible testing for ential diagnosis includes TP53 or PTEN mutations, as well as BRCA1
genetic predisposition to breast cancer have been established by and BRCA2. Genetic testing laboratories offer a wide variety of
the National Comprehensive Cancer Network [83]. Key criteria in cancer gene panel tests; while some include only selected high-
these guidelines are summarized in Table 2. risk/highly penetrant genes, others offer expanded gene lists that
includes dozens of genes associated moderate cancer risk, for
3.2. Collection of family history in breast cancer risk evaluation which clinical implications and management recommendations
remain uncertain [85]. The decision whether to use a comprehen-
Family history is a complex tool for assessing disease risk that can sive versus more targeted gene panel should consider each
patient’s personal and family history, as well as the clinical impact
offer insight into the multifactorial risk for many conditions (especially
of findings of alterations in moderate/low-penetrance genes and/
cancer) and can clarify a patient’s potential risk and can play an
important role in increasing uptake and effectiveness of preventative or variants of uncertain significance.
services.
A detailed family history, comprising at least three generations, if 3.4. Variants of undetermined significance in breast cancer risk genes
available, is a highly useful first step in cancer risk evaluation. If
possible, confirmation of cancer diagnosis should be made via medical Complete gene sequencing can identify changes in the
records or pathology reports. Information on gynecologic malignancies sequence of a gene that are known to be pathogenic, as well as
affecting a family member can be especially unreliable, and obtaining variants of unknown or uncertain significance (VUS), for which the
pathology confirmation becomes quite important in these cases to effect of the sequence change on gene function is not known.
prevent inaccurate risk calculations. Important components of family When a VUS is found during a test for inherited susceptibility to
history are summarized below in Table 3. cancer, the diagnosis of a hereditary cancer syndrome often cannot

Table 2
NCCN criteria for HBOC risk evaluation.

Risk evaluation for HBOC should be performed in patients who meet one or more of the following criteria:
1. Female breast cancer diagnosed r 50 years of age
2. Triple-negative breast cancer diagnosed at age r 60 years
3. Invasive ovarian, fallopian tube, or primary peritoneal cancer
4. Male breast cance
5. Any HBOC-associated cancers, regardless of age at diagnosis, and of Ashkenazi Jewish Ancestry
6. Two or more primary breast cancers
Risk evaluation should be performed in patients with breast cancer and first-, second-, or third-degree relatives with any of the following:
1. Breast cancer diagnosed at age r50 years in one or more relatives
2. Invasive ovarian, fallopian tube, or primary peritoneal cancer in one or more relatives
3. Breast, prostate, and/or pancreatic cancer diagnosed at any age in two or more relatives
E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]] 5

Table 3
Important components of family history (including relatives at least to third degree)

Individual data Notes

Identification First and last name for patient, but no identifiers for relatives should be recorded to preserve privacy of the family members
Current age Current age of patient and family members, with a date stamp of when this was recorded so that the family history information can be
updated
Age of death If exact age is not known, approximations of age of death (eg, 50s) is appropriate
Cause of death Unknown cause of death should always be noted
Ethnicity/race Will inform risk assessment
Biological sex
Twin/multiple births Identical/Fraternal status must be noted
Biological parents identified Allows for pedigree construction
Consanguinity
Adoptive status
Other genetic disorders Requires documentation of diagnosis type, age of diagnosis, and age of death (if appropriate)
Relevant genetic/genomic In following of HIPPA rules
test
Certainty of data Noted by provider

Adapted from Feero and colleagues [108].

be confirmed or excluded, and patients should be counseled process, may have therapeutic implications for the patient.
accordingly about their own risks as well as potential risks for Although not approved by the US Food and Drug Administration
their family members. When a VUS genetic test result is identified, (FDA) for treatment of breast cancer, studies of inhibitors of poly
there are several additional pieces of data that can help assess the adenosine diphosphate-ribose polymerase (PARP) have shown
likelihood that the VUS is associated with an increased risk for promise in the treatment of patients with metastatic BRCA1/2-
cancer, including the following: associated breast cancer [87–90]. For example, in a study by Tutt
and colleagues, it was demonstrated that olaparib administered to
(1) Does the VUS track with cancer in the family? If yes, then women with advanced BRCA1/2-associated breast cancer resulted
this is additional information that the VUS may be associated in an overall response rate of 41% and progression-free survival of
with an increased risk for developing cancer, if not, than this is 5.7 months [87]. Of note, olaparib does not appear to have activity
additional information that the VUS may not be associated in breast cancer outside of BRCA1/2 mutation carriers; this is in
with an increased risk for developing cancer. contrast to epithelial ovarian cancers, which appear to respond to
(2) Is the VUS seen in both individuals with cancer and without PARP inhibition independent of germline BRCA1/2 status [89].
cancer? If yes, then the VUS may not be associated with an There are numerous ongoing clinical trials looking at the use of
increased risk for developing cancer. PARP inhibitors the neoadjuvant, adjuvant and metastatic treat-
(3) Has the VUS been seen in association with a known gene ment settings for patients with pathogenic germline alterations in
mutation in a gene that causes an increased risk for BRCA1/2. The American Society of Clinical Oncology (ASCO) Tar-
developing cancer? If yes, then the VUS may not be associated geted Agent Profiling and Utilization Registry (TAPUR) study is a
with an increased risk for developing cancer. prospective, non-randomized trial that aims to describe the safety
(4) Is the area of the gene where the VUS is located in an area of and efficacy of commercially available, targeted anti-cancer drugs
the gene that appears to be important for all species? If yes, prescribed for treatment of patients with advanced cancer that
then the VUS may be associated with an increased risk for have a potentially actionable genomic variant revealed by a
developing cancer. genomic test. Under this protocol, patients with either germline
(5) Is the gene change anticipated to affect how the gene or somatic bi-allelic loss of BRCA1/2 or ATM identified in a
functions in the body? If yes, then the VUS may be associated metastatic tumor sample are eligible to receive a PARP inhibitor
with an increased risk for developing cancer. as therapy.
(6) Is there a loss of the normal gene in tumor analysis? If yes,
then the VUS may be associated with an increased risk for
developing cancer. 5. Management strategies for cancer risk reduction
(7) Does the protein seem to function normally in laboratory
studies? If yes, then the VUS may not be associated with an Strategies for breast cancer risk reduction for individuals at
increased risk for developing cancer. high–moderate risk of developing breast cancer include surveil-
lance, risk-reducing surgery, chemoprevention, and to some
There are many databases in the public domain that collect extent, lifestyle modifications. Below, these options are reviewed
information on VUS. These include but are not limited to the Breast in detail with regard to potential magnitude of benefit in specific
Information Core (BIC), ClinVar, Prospective Registry of Multiplex mutation carriers as well as potential impact on patient quality
Testing (PROMPT), International Agency of Cancer Research (IARC), of life.
International Society for Inherited Gastrointestinal Tumours
(InSight), and Mismatch Repair Genes Variant Database [86]. 5.1. Risk-reducing surgery

National Comprehensive Cancer Network (NCCN) guidelines


4. Implications of genetic diagnosis for treatment of breast recommend that prophylactic mastectomy be offered to all
cancer patients who carry a germline BRCA1 or BRCA2 mutation, as both
retrospective and prospective studies have demonstrated that the
Identification of a pathogenic germline BRCA1/2 mutation, or incidence of breast cancer can be reduced by as much as 90% in
pathogenic mutation of other genes involved in the DNA repair those that undergo surgery [91–98]. It should be noted that for
6 E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]]

Table 4
NCCN guidelines: breast cancer screening in BRCA1/2 mutation carriers

Periodic self-breast examination beginning at age 18


Clinical breast examination every 6–12 months beginning at age 25
Annual mammography beginning at age 30 (or individualized if earliest onset of cancer in the family is under age 25)
Annual breast magnetic resonance imaging beginning at age 25
Discussion of ovarian cancer screening in those whom have not undergone BSO to include concurrent transvaginal ultrasound and CA-125 measurement every 6 months
beginning at age 30 (or 5–10 years prior to the age of first diagnosis of ovarian cancer in the family)

those patients who choose to have surgery, a bilateral total observed toward increased risk of breast cancer in those to
mastectomy or skin-sparing mastectomy is recommended as use OCPs.
opposed to a subcutaneous mastectomy. The latter surgical
approach leaves behind glandular tissue which can subsequently 5.4. Breast cancer risk reduction strategies for other gene mutation
develop into breast cancer [92,99]. carriers
NCCN guidelines also recommend that bilateral salpingo-
oophorectomy (BSO) be offered to BRCA mutation carriers once Strategies to reduce the risk of developing breast cancer in
they have completed childbearing, and is generally performed patients with germline mutations in other high-penetrance cancer
between the ages of 35–40. In BRCA mutation carriers, BSO predisposition genes associated with lifetime breast cancer risks of
has been demonstrated to reduce the risk of ovarian cancer and 4 40% (TP53, PTEN, STK11, CDH1) are similar to those described
breast cancer, as well as mortality [97,100,101]. Following BSO, above for BRCA mutation carriers; however, management of those
there is a theoretical concern that reduction in breast cancer risk patients with genetic mutations associated with more moderate
could be negated by the use of hormone replacement therapy cancer risk (CHEK2, PALB2 and ATM) is more complex. In general,
(HRT) in patients that are symptomatic following surgery. breast cancer surveillance recommendations are similar to those
Several studies have demonstrated, however, that this is not the made for carriers of highly penetrant genes; however, some
case [97,102]. Given this, it is reasonable to offer women who are guidelines bodies, such as the United Kingdom’s National Health
symptomatic following BSO HRT to mitigate these symptoms, Service, have recommended that patients with germline ATM
particularly in those that have also undergone prophylactic mutations try to minimize exposure to radiation, and undergo
mastectomy. surveillance with annual breast magnetic resonance imaging
beginning at age 25 in addition to mammography every 18 months
5.2. Surveillance strategies from ages 40–49, followed by every 3 years after age 50 [106].
Decisions regarding risk reduction surgery (such as prophylactic
For those women who harbor a BRCA1 or BRCA2 mutation and mastectomy) must be highly individualized in these patients,
do not pursue risk-reducing surgeries, breast cancer surveillance taking into account the patient’s family history of breast cancer
should be offered and screening for ovarian cancer can be in addition to their mutation status. In addition, patients with
considered. The highlights of the NCCN guidelines regarding breast CHEK2 mutations may be candidates for chemoprevention with
cancer screening in BRCA mutations carriers are listed in Table 4 tamoxifen, as they are considered to be more likely to develop
[83]. estrogen receptor–positive breast cancers.

5.3. Chemoprevention 6. Conclusions

For BRCA1 or BRCA2 mutation carriers who decline prophylactic The demand for clinical evaluations of individuals for genetic
surgery, the use of chemoprevention agents can be considered; susceptibility to breast cancer is increasing. Amidst changes in
however, data are limited regarding the effectiveness of selective technology and the advent of more data on the significance of gene
estrogen receptor modulators (ie, tamoxifen or raloxifene) for variants, ASCO provides guidelines for genetic evaluation for
preventing breast cancer in patients with BRCA1 or BRCA2 muta- cancer susceptibility, highlighting the challenges of multigene
tions. The most robust evidence for use of tamoxifen in this setting panel testing, the need to ensure quality assurance in genetic
comes from a subset analysis of National Surgical Adjuvant Breast testing laboratories, promote education among oncology profes-
and Bowel Project (NSABP) Breast Cancer Prevention Trial (P-1 sionals that are ordering, interpreting and discussing genetic
trial). In this study, use of tamoxifen in BRCA2 mutations carriers testing results with patients, and providing access to cancer
decreased the risk of developing breast cancer by 62%; however, genetics services [107].
risk of breast cancer was not reduced in BRCA1 mutation carriers Pre- and post-genetic test counseling by healthcare providers
[103]. This study was limited by small numbers; however, the with genetics expertise leads to proven improvements in
difference in risk reduction observed between BRCA1 and BRCA2 patients’ knowledge about genetics, increased satisfaction with
mutation carriers may be attributable to the higher incidence of decision-making about risk management options, and risk appro-
estrogen receptor–positive breast cancers in the BRCA2 group. priate access to screening. There is a significant disparity in access
While aromatase inhibitors and raloxifene have been demon- to specialized genetic counseling for HBOC yndromes by tradi-
strated to decrease the risk of developing breast cancer in other tionally underserved populations based on ethnicity and socio-
high-risk populations, there are no studies that have looked at the economic status. Such disparities ought to be high priority
potential use of these agents in patients with BRCA1 or BRCA2 concerns in public health in the United States. Risk management
mutations. plans for increased cancer susceptibility are dynamic strategies
Use of oral contraceptives (OCPs), primarily for reduction in risk that should be revisited intermittently throughout the lifespan and
of developing ovarian cancer, can be considered in those BRCA as the person ages. The rapid advances in science and technology
mutation carriers who decline BSO. While studies have demon- provide excellent new tools that need to be harnessed for the
strated a benefit for this purpose [104,105], there has been a trend benefit of society, for improving cancer outcomes and decreasing
E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]] 7

cancer incidence, through the provision of specialized services to [28] Olivier M, Goldgar DE, Sodha N, et al. Li-Fraumeni and related syndromes:
individuals at increased risk in an equitable and cost-effective correlation between tumor type, family structure, and TP53 genotype.
Cancer Res 2003;63(20):6643–50.
manner. [29] Wilson JR, Bateman AC, Hanson H, et al. A novel HER2-positive breast cancer
phenotype arising from germline TP53 mutations. J Med Genet 2010;47
(11):771–4.
[30] Beggs AD, Latchford AR, Vasen HF, et al. Peutz-Jeghers syndrome: a
Conflicts of interest systematic review and recommendations for management. Gut 2010;59
(7):975–86.
None [31] Giardiello FM, Brensinger JD, Tersmette AC, et al. Very high risk of cancer in
familial Peutz-Jeghers syndrome. Gastroenterology 2000;119(6):1447–53.
[32] Tan MH, Mester JL, Ngeow J, Rybicki LA, Orloff MS, Eng C. Lifetime cancer
risks in individuals with germline PTEN mutations. Clin Cancer Res 2012;18
References
(2):400–7.
[33] Starink TM, van der Veen JP, Arwert F, et al. The Cowden syndrome: a clinical
[1] Hall JM, Lee MK, Newman B, et al. Linkage of early-onset familial breast and genetic study in 21 patients. Clin Genet 1986;29(3):222–33.
cancer to chromosome 17q21. Science 1990;250(4988):1684–9. [34] Pilarski R, Burt R, Kohlman W, Pho L, Shannon KM, Swisher E. Cowden
[2] Narod SA, Sobol H, Schuffenecker I, Ezekowitz RA, Lenoir GM. Early detection syndrome and the PTEN hamartoma tumor syndrome: systematic review
of hereditary medullary thyroid cancer with polymorphic DNA probes. and revised diagnostic criteria. J Natl Cancer Inst 2013;105(21):1607–16.
Groupe d'Etude des Tumeurs a Calcitonine. Henry Ford Hosp Med J [35] Fitzgerald RC, Hardwick R, Huntsman D, et al. Hereditary diffuse gastric
1989;37(3-4):106–8. cancer: updated consensus guidelines for clinical management and direc-
[3] Miki Y, Swensen J, Shattuck-Eidens D, et al. A strong candidate for the breast tions for future research. Journal of medical genetics 2010;47(7):436–44.
and ovarian cancer susceptibility gene BRCA1. Science 1994;266(5182): [36] Guilford P, Humar B, Blair V. Hereditary diffuse gastric cancer: translation of
66–71. CDH1 germline mutations into clinical practice. Gastric Cancer 2010;13
[4] Futreal PA, Liu Q, Shattuck-Eidens D, et al. BRCA1 mutations in primary (1):1–10.
breast and ovarian carcinomas. Science 1994;266(5182):120–2. [37] Hansford S, Kaurah P, Li-Chang H, et al. Hereditary diffuse gastric cancer
[5] Tavtigian SV, Simard J, Rommens J, et al. The complete BRCA2 gene and syndrome: CDH1 mutations and beyond. JAMA Oncol 2015;1(1):23–32.
mutations in chromosome 13q-linked kindreds. Nat Genet 1996;12(3): [38] Xie ZM, Li LS, Laquet C, Penault-Llorca F, Uhrhammer N, Xie XM, Bignon YJ.
333–7. Germline mutations of the E-cadherin gene in families with inherited
[6] Venkitaraman AR. Cancer suppression by the chromosome custodians, invasive lobular breast carcinoma but no diffuse gastric cancer. Cancer
BRCA1 and BRCA2. Science 2014;343(6178):1470–5. 2011;117(14):3112–7.
[7] Lord CJ, Ashworth A. The DNA damage response and cancer therapy. Nature [39] Shah MA, Salo-Mullen E, Stadler Z, et al. De novo CDH1 mutation in a family
2012;481(7381):287–94. presenting with early-onset diffuse gastric cancer. Clin Genet 2012;82
[8] Merajver SD, Pham TM, Caduff RF, et al. Somatic mutations in the BRCA1 (3):283–7.
gene in sporadic ovarian tumours. Nat Genet 1995;9(4):439–43. [40] Sugimoto S, Yamada H, Takahashi M, et al. Early-onset diffuse gastric cancer
[9] Hoskins KF, Stopfer JE, Calzone KA, et al. Assessment and counseling for associated with a de novo large genomic deletion of CDH1 gene. Gastric
women with a family history of breast cancer. A guide for clinicians. JAMA Cancer 2014;17(4):745–9, http://dx.doi.org/10.1007/s10120-013-0278-2.
1995;273(7):577–85. [41] Walsh MD, Buchanan DD, Cummings MC, et al. Lynch syndrome-associated
[10] Olopade OI, Grushko TA, Nanda R, Huo D. Advances in breast cancer: breast cancers: clinicopathologic characteristics of a case series from the
pathways to personalized medicine. Clin Cancer Res 2008;14(24):7988–99. colon cancer family registry. Clin Cancer Res 2010;16(7):2214–24.
[11] Couch FJ, Hart SN, Sharma P, et al. Inherited mutations in 17 breast cancer [42] Shulman LP. Hereditary breast and ovarian cancer (HBOC): clinical features
susceptibility genes among a large triple-negative breast cancer cohort and counseling for BRCA1 and BRCA2, Lynch syndrome, Cowden syndrome,
unselected for family history of breast cancer. J Clin Oncol 2015;33 and Li-Fraumeni syndrome. Obstet Gynecol Clin North Am 2010;37
(4):304–11. (1):109–33.
[12] Antoniou A, Pharoah PD, Narod S, et al. Average risks of breast and ovarian [43] Buerki N, Gautier L, Kovac M, et al. Evidence for breast cancer as an integral
cancer associated with BRCA1 or BRCA2 mutations detected in case Series part of Lynch syndrome. Genes Chromosomes Cancer 2012;51(1):83–91.
unselected for family history: a combined analysis of 22 studies. Am J Hum
[44] Win AK, Young JP, Lindor NM, et al. Colorectal and other cancer risks for
Genet 2003;72(5):1117–30.
carriers and noncarriers from families with a DNA mismatch repair gene
[13] Ford D, Easton DF, Stratton M, et al. Genetic heterogeneity and penetrance
mutation: a prospective cohort study. J Clin Oncol 2012;30(9):958–64.
analysis of the BRCA1 and BRCA2 genes in breast cancer families. The Breast
[45] Win AK, Lindor NM, Jenkins MA. Risk of breast cancer in Lynch syndrome: a
Cancer Linkage Consortium. Am J Hum Genet 1998;62(3):676–89.
systematic review. Breast Cancer Res 2013;15(2):R27.
[14] van der Kolk DM, de Bock GH, Leegte BK, et al. Penetrance of breast cancer,
[46] Moldovan GL, D'Andrea AD. How the fanconi anemia pathway guards the
ovarian cancer and contralateral breast cancer in BRCA1 and BRCA2 families:
genome. Annu Rev Genet 2009;43:223–49.
high cancer incidence at older age. Breast Cancer Res Treat 2010;124(3):
[47] Naslund-Koch C, Nordestgaard BG, Bojesen SE. Increased risk for other
643–51.
cancers in addition to breast cancer for CHEK2*1100delC heterozygotes
[15] Meijers-Heijboer EJ, Verhoog LC, Brekelmans CT, et al. Presymptomatic DNA
estimated from the Copenhagen General Population Study. J Clin Oncol
testing and prophylactic surgery in families with a BRCA1 or BRCA2
2016;34(11):1208–16.
mutation. Lancet 2000;355(9220):2015–20.
[48] Schmidt MK, Tollenaar RA, de Kemp SR, et al. Breast cancer survival and
[16] Roa BB, Boyd AA, Volcik K, Richards CS. Ashkenazi Jewish population
tumor characteristics in premenopausal women carrying the CHEK2*1100-
frequencies for common mutations in BRCA1 and BRCA2. Nat Genet
delC germline mutation. J Clin Oncol 2007;25(1):64–9.
1996;14(2):185–7.
[17] Ferla R, Calo V, Cascio S, et al. Founder mutations in BRCA1 and BRCA2 genes. [49] Guenard F, Pedneault CS, Ouellette G, et al. Evaluation of the contribution of
Ann Oncol 2007;18(Suppl 6):vi93–8. the three breast cancer susceptibility genes CHEK2, STK11, and PALB2 in non-
[18] Szabo CI, King MC. Population genetics of BRCA1 and BRCA2. Am J Hum BRCA1/2 French Canadian families with high risk of breast cancer. Genet Test
Genet 1997;60(5):1013–20. Mol Biomarkers 2010;14(4):515–26.
[19] Begg CB, Haile RW, Borg A, et al. Variation of breast cancer risk among [50] Weischer M, Nordestgaard BG, Pharoah P, et al. CHEK2*1100delC hetero-
BRCA1/2 carriers. JAMA 2008;299(2):194–201. zygosity in women with breast cancer associated with early death, breast
[20] Wang F, Fang Q, Ge Z, Yu N, Xu S, Fan X. Common BRCA1 and BRCA2 cancer-specific death, and increased risk of a second breast cancer. J Clin
mutations in breast cancer families: a meta-analysis from systematic review. Oncol 2012;30(35):4308–16.
Mol Biol Rep 2012;39(3):2109–18. [51] Narod SA. Testing for CHEK2 in the cancer genetics clinic: ready for prime
[21] Judkins T, Rosenthal E, Arnell C, et al. Clinical significance of large rearrange- time? Clin Genet 2010;78(1):1–7.
ments in BRCA1 and BRCA2. Cancer 2012;118(21):5210–6. [52] Cybulski C, Wokolorczyk D, Jakubowska A, et al. Risk of breast cancer in
[22] Tai YC, Domchek S, Parmigiani G, Chen S. Breast cancer risk among male women with a CHEK2 mutation with and without a family history of breast
BRCA1 and BRCA2 mutation carriers. J Natl Cancer Inst 2007;99(23):1811–4. cancer. J Clin Oncol 2011;29(28):3747–52.
[23] Evans DG, Susnerwala I, Dawson J, Woodward E, Maher ER, Lalloo F. Risk of [53] Meijers-Heijboer H, van den Ouweland A, Klijn J, et al. Low-penetrance
breast cancer in male BRCA2 carriers. J Med Genet 2010;47(10):710–1. susceptibility to breast cancer due to CHEK2(*)1100delC in noncarriers of
[24] Malkin D. Li-Fraumeni syndrome. Genes Cancer 2011;2(4):475–84. BRCA1 or BRCA2 mutations. Nat Genet 2002;31(1):55–9.
[25] Hisada M, Garber JE, Fung CY, Fraumeni JF Jr, Li FP. Multiple primary cancers [54] Antoniou AC, Casadei S, Heikkinen T, et al. Breast-cancer risk in families with
in families with Li-Fraumeni syndrome. J Natl Cancer Inst 1998;90(8): mutations in PALB2. N Engl J Med 2014;371(6):497–506.
606–11. [55] Catucci I, Milgrom R, Kushnir A, et al. Germline mutations in BRIP1 and
[26] Lammens CR, Bleiker EM, Aaronson NK, et al. Regular surveillance for Li- PALB2 in Jewish high cancer risk families. Fam Cancer 2012;11(3):483–91.
Fraumeni Syndrome: advice, adherence and perceived benefits. Fam Cancer [56] Tischkowitz M, Xia B. PALB2/FANCN: recombining cancer and Fanconi
2010;9(4):647–54. anemia. Cancer Res 2010;70(19):7353–9.
[27] Masciari S, Dillon DA, Rath M, et al. Breast cancer phenotype in women with [57] Renwick A, Thompson D, Seal S, et al. ATM mutations that cause ataxia-
TP53 germline mutations: a Li-Fraumeni syndrome consortium effort. Breast telangiectasia are breast cancer susceptibility alleles. Nat Genet 2006;38
Cancer Res Treat 2012;133(3):1125–30. (8):873–5.
8 E.F. Cobain et al. / Seminars in Oncology ] (2016) ]]]–]]]

[58] Thompson D, Duedal S, Kirner J, et al. Cancer risks and mortality in [85] Doherty J, Bonadies DC, Matloff ET. Testing for hereditary breast cancer:
heterozygous ATM mutation carriers. J Natl Cancer Inst 2005;97(11):813–22. panel or targeted testing? Experience from a clinical cancer genetics
[59] Paglia LL, Lauge A, Weber J, et al. ATM germline mutations in women with practice. J Genet Couns 2015;24(4):683–7.
familial breast cancer and a relative with haematological malignancy. Breast [86] Domchek S, Weber BL. Genetic variants of uncertain significance: flies in the
Cancer Res Treat 2010;119(2):443–52. ointment. J Clin Oncol 2008;26(1):16–7.
[60] Swift M, Morrell D, Cromartie E, Chamberlin AR, Skolnick MH, Bishop DT. [87] Tutt A, Robson M, Garber JE, et al. Oral poly(ADP-ribose) polymerase
The incidence and gene frequency of ataxia-telangiectasia in the United inhibitor olaparib in patients with BRCA1 or BRCA2 mutations and advanced
States. Am J Hum Genet 1986;39(5):573–83. breast cancer: a proof-of-concept trial. Lancet 2010;376(9737):235–44.
[61] Kurian AW, Hare EE, Mills MA, et al. Clinical evaluation of a multiple-gene [88] Tutt ARM, Garber JE, et al. Phase II trial of the oral PARP inhibitor olaparib in
sequencing panel for hereditary cancer risk assessment. J Clin Oncol 2014;32 BRCA-deficient advanced breast cancer. J Clin Oncol 2009;27(18s): Abstract
(19):2001–9. CRA501.
[62] Wu LC, Wang ZW, Tsan JT, et al. Identification of a RING protein that can [89] Gelmon KA, Herte HW, Robidoux A, et al. Can we define tumors that will
interact in vivo with the BRCA1 gene product. Nat Genet 1996;14(4):430–40. respond to PARP inhibitors? A phase II correlative study of olaparib in
[63] Thai TH, Du F, Tsan JT, et al. Mutations in the BRCA1-associated RING domain advanced serous ovarian cancer and triple-negative breast cancer. J Clin
(BARD1) gene in primary breast, ovarian and uterine cancers. Hum Mol Oncol 2010;28(15s).
Genet 1998;7(2):195–202. [90] Isakoff SJ, Overmoyer B, Tung NM, et al. A phase II trial of the PARP inhibitor
[64] Ghimenti C, Sensi E, Presciuttini S, et al. Germline mutations of the BRCA1- veliparib (ABT888) and temozolomide for metastatic breast cancer. J Clin
associated ring domain (BARD1) gene in breast and breast/ovarian families Oncol 2010;28(15s):abstract 3002.
negative for BRCA1 and BRCA2 alterations. Genes Chromosomes Cancer [91] Meijers-Heijboer H, van Geel B, van Putten WL, et al. Breast cancer after
2002;33(3):235–42. prophylactic bilateral mastectomy in women with a BRCA1 or BRCA2
[65] De Brakeleer S, De Greve J, Loris R, et al. Cancer predisposing missense and mutation. N Engl J Med 2001;345(3):159–64.
protein truncating BARD1 mutations in non-BRCA1 or BRCA2 breast cancer [92] Rebbeck TR, Friebel T, Lynch HT, et al. Bilateral prophylactic mastectomy
families. Hum Mutat 2010;31(3):E1175–85. reduces breast cancer risk in BRCA1 and BRCA2 mutation carriers: the PROSE
[66] Cantor SB, Bell DW, Ganesan S, et al. BACH1, a novel helicase-like protein, Study Group. J Clin Oncol 2004;22(6):1055–62.
interacts directly with BRCA1 and contributes to its DNA repair function. Cell [93] Hartmann LC, Schaid DJ, Woods JE, et al. Efficacy of bilateral prophylactic
2001;105(1):149–60. mastectomy in women with a family history of breast cancer. N Engl J Med
[67] Levitus M, Waisfisz Q, Godthelp BC, et al. The DNA helicase BRIP1 is defective 1999;340(2):77–84.
in Fanconi anemia complementation group. J Nat Genet 2005;37(9):934–5. [94] Hartmann LC, Sellers TA, Schaid DJ, et al. Efficacy of bilateral prophylactic
[68] Frank B, Hemminki K, Meindl A, et al. BRIP1 (BACH1) variants and familial mastectomy in BRCA1 and BRCA2 gene mutation carriers. J Natl Cancer Inst
breast cancer risk: a case-control study. BMC Cancer 2007;7:83. 2001;93(21):1633–7.
[69] Wong MW, Nordfors C, Mossman D, et al. BRIP1, PALB2, and RAD51C [95] Heemskerk-Gerritsen BA, Brekelmans CT, Menke-Pluymers MB, et al. Pro-
phylactic mastectomy in BRCA1/2 mutation carriers and women at risk of
mutation analysis reveals their relative importance as genetic susceptibility
hereditary breast cancer: long-term experiences at the Rotterdam Family
factors for breast cancer. Breast Cancer Res Treat 2011;127(3):853–9.
Cancer Clinic. Ann Surg Oncol 2007;14(12):3335–44.
[70] Guenard F, Labrie Y, Ouellette G, et al. Mutational analysis of the breast
[96] Geiger AM, Yu O, Herrinton LJ, et al. A population-based study of bilateral
cancer susceptibility gene BRIP1 /BACH1/FANCJ in high-risk non-BRCA1/
prophylactic mastectomy efficacy in women at elevated risk for breast cancer
BRCA2 breast cancer families. J Hum Genet 2008;53(7):579–91.
in community practices. Arch Intern Med 2005;165(5):516–20.
[71] Seal S, Thompson D, Renwick A, et al. Truncating mutations in the Fanconi
[97] Domchek SM, Friebel TM, Singer CF, et al. Association of risk-reducing
anemia J gene BRIP1 are low-penetrance breast cancer susceptibility alleles.
surgery in BRCA1 or BRCA2 mutation carriers with cancer risk and mortality.
Nat Genet 2006;38(11):1239–41.
JAMA 2010;304(9):967–75.
[72] Easton DF, Lesueur F, Decker B, et al. No evidence that protein truncating
[98] Ingham SL, Sperrin M, Baildam A, et al. Risk-reducing surgery increases
variants in BRIP1 are associated with breast cancer risk: implications for
survival in BRCA1/2 mutation carriers unaffected at time of family referral.
gene panel testing. J Med Genet 2016;53(5):298–309, http://dx.doi.org/
Breast Cancer Res Treat 2013;142(3):611–8.
10.1136/jmedgenet-2015-103529 PubMed PMID: 26921362; PubMed Central
[99] Guillem JG, Wood WC, Moley JF, et al. ASCO/SSO review of current role of
PMCID: PMCPMC4938802.
risk-reducing surgery in common hereditary cancer syndromes. J Clin Oncol
[73] Lavin MF. ATM and the Mre11 complex combine to recognize and signal DNA
2006;24(28):4642–60.
double-strand breaks. Oncogene 2007;26(56):7749–58. [100] Rebbeck TR, Lynch HT, Neuhausen SL, et al. Prophylactic oophorectomy in
[74] Damiola F, Pertesi M, Oliver J, et al. Rare key functional domain missense carriers of BRCA1 or BRCA2 mutations. N Engl J Med 2002;346(21):1616–22.
substitutions in MRE11A, RAD50, and NBN contribute to breast cancer [101] Kauff ND, Satagopan JM, Robson ME, et al. Risk-reducing salpingo-oopho-
susceptibility: results from a Breast Cancer Family Registry case-control rectomy in women with a BRCA1 or BRCA2 mutation. N Engl J Med 2002;346
mutation-screening study. Breast Cancer Res 2014;16(3):R58. (21):1609–15.
[75] Vaz F, Hanenberg H, Schuster B, et al. Mutation of the RAD51C gene in a [102] Rebbeck TR, Friebel T, Wagner T, et al. Effect of short-term hormone
Fanconi anemia-like disorder. Nat Genet 2010;42(5):406–9. replacement therapy on breast cancer risk reduction after bilateral prophy-
[76] Meindl A, Hellebrand H, Wiek C, et al. Germline mutations in breast and lactic oophorectomy in BRCA1 and BRCA2 mutation carriers: the PROSE
ovarian cancer pedigrees establish RAD51C as a human cancer susceptibility Study Group. J Clin Oncol 2005;23(31):7804–10.
gene. Nat Genet 2010;42(5):410–4. [103] King MC, Wieand S, Hale K, et al. Tamoxifen and breast cancer incidence
[77] Loveday C, Turnbull C, Ramsay E, et al. Germline mutations in RAD51D confer among women with inherited mutations in BRCA1 and BRCA2: National
susceptibility to ovarian cancer. Nat Genet 2011;43(9):879–82. Surgical Adjuvant Breast and Bowel Project (NSABP-P1) Breast Cancer
[78] Baker JL, Schwab RB, Wallace AM, Madlensky L. Breast cancer in a RAD51D Prevention Trial. JAMA 2001;286(18):2251–6.
mutation carrier: case report and review of the literature. Clin Breast Cancer [104] Iodice S, Barile M, Rotmensz N, et al. Oral contraceptive use and breast or
2015;15(1):e71–5. ovarian cancer risk in BRCA1/2 carriers: a meta-analysis. Eur J Cancer
[79] Theodoratou E, Campbell H, Tenesa A, et al. A large-scale meta-analysis to 2010;46(12):2275–84.
refine colorectal cancer risk estimates associated with MUTYH variants. Br J [105] Moorman PG, Havrilesky LJ, Gierisch JM, et al. Oral contraceptives and risk of
Cancer 2010;103(12):1875–84. ovarian cancer and breast cancer among high-risk women: a systematic
[80] Nielsen M, Franken PF, Reinards TH, et al. Multiplicity in polyp count and review and meta-analysis. J Clin Oncol 2013;31(33):4188–98.
extracolonic manifestations in 40 Dutch patients with MYH associated [106] Programme NBCS. NHSBSP Publication no., December 2012. Available from:
polyposis coli (MAP). J Med Genet 2005;42(9):e54. https://www.gov.uk/government/uploads/system/uploads/attachment_data/
[81] Wasielewski M, Out AA, Vermeulen J, et al. Increased MUTYH mutation file/439601/nhsbsp68.pdf. Accessed [July 22, 2016].
frequency among Dutch families with breast cancer and colorectal cancer. [107] Robson ME, Bradbury AR, Arun B, et al. American Society of Clinical Oncology
Breast Cancer Res Treat 2010;124(3):635–41. Policy Statement update: genetic and genomic testing for cancer suscepti-
[82] Out AA, Wasielewski M, Huijts PE, et al. MUTYH gene variants and breast bility. J Clin Oncol 2015;33(31):3660–7.
cancer in a Dutch case-control study. Breast Cancer Res Treat 2012;134 [108] Feero WG, Bigley MB, Brinner KM. Family Health History Multi-Stakeholder
(1):219–27. Workgroup of the American Health Information Community. New standards
[83] Network NCC. Version 2.2016 [Available from: https://www.nccn.org/profes and enhanced utility for family health history information in the electronic
sionals/physician_gls/pdf/genetics_screening.pdf. Accessed [July 22, 2016]. health record: an update from the American Health Information Commun-
[84] Domchek SM. Evolution of genetic testing for inherited susceptibility to ity's Family Health History Multi-Stakeholder Workgroup. J Am Med Inform
breast cancer. J Clin Oncol 2015;33(4):295–6. Assoc 2008;15(6):723–8.

You might also like