miércoles, 5 de febrero de 2020

Genetics of Breast and Gynecologic Cancers (PDQ®)–Health Professional Version - National Cancer Institute

Genetics of Breast and Gynecologic Cancers (PDQ®)–Health Professional Version - National Cancer Institute

National Cancer Institute



Genetics of Breast and Gynecologic Cancers (PDQ®)–Health Professional Version



Psychosocial Issues in Inherited Breast and Ovarian Cancer Syndromes



Introduction

Psychosocial research in the context of cancer genetic testing helps to define psychological outcomes, interpersonal and familial effects, and cultural and community responses. This type of research also identifies behavioral factors that encourage or impede screening and other health behaviors. It can enhance decision making about risk-reduction interventions, evaluate psychosocial interventions to reduce distress and/or other negative sequelae related to risk notification and genetic testing, provide data to help resolve ethical concerns, and predict the interest in testing of various groups.
This section addresses psychosocial issues in hereditary breast and ovarian cancer syndromes. Psychosocial and screening issues related to gynecologic cancers associated with Lynch syndrome are discussed in the Psychosocial Issues in Hereditary Colon Cancer Syndromes section in the PDQ summary on Genetics of Colorectal Cancer.

Uptake of Genetic Counseling and Genetic Testing

Degree of uptake of genetic counseling and genetic testing

Comparison of uptake rates among studies in which counseling and testing were offered is challenging because of differences in methodologies, including the sampling strategy used, the recruitment setting, and testing through a research protocol with high-risk cohorts or kindreds. In a systematic review of 40 studies conducted before 2002 that had assessed genetic testing utilization, uptake rates varied widely and ranged from 25% to 96%, with an average uptake rate of 59%.[1] Results of multivariate analysis found that BRCA1/BRCA2 genetic testing uptake was associated with having a personal or family history of breast or ovarian cancer, and with methodological features of the studies, including sampling strategies, recruitment settings, and how studies defined actual uptake versus the intention to have testing.
Other factors have been positively correlated with uptake of BRCA1/BRCA2 genetic testing, although these findings are not consistent across all studies. Psychological factors that have been positively correlated with testing uptake include greater cancer-specific distress and greater perceived risk of developing breast or ovarian cancer. Having more cancer-affected relatives also has been correlated with greater testing uptake.
Table 14 summarizes the uptake of genetic testing in clinical and research cohorts in the United States.
Table 14. Predictors Associated with Uptake of Genetic Testing (GT)
ENLARGE
Study CitationStudy PopulationSample Size (N)Uptake of GTPredictors Associated With Uptake of GTComments
GC = genetic counseling; HMO = health maintenance organization.
aSelf-report as data source.
bMedical records as data source.
Schwartz et al. (2005) [2]Newly diagnosed and locally untreated breast cancer patients with ≥10% risk of having a BRCA1/BRCA2 pathogenic variant a231177/231 (77%) underwent GTHaving decided on definitive local treatment. Women who were undecided on a definitive local treatment were more likely to be tested.Testing was offered free of charge.
34/231 (15%) had baseline interview but declined GT
Physician recommendation for testing. Women whose physician had recommended GT were more likely to be tested.38/177 chose to proceed with treatment before receiving test results.
20/231 declined baseline interview
Kieran et al. (2007) [3]Women who received GC between 2002 and 2004a25088/250 (35%) underwent GTAbility to pay for GT (entire cost or cost not covered by insurance). Nonuptake was 5.5 times more likely in women who could not afford testing.450 women received GC for breast and ovarian cancer risk during study period. 250 women were retrospectively identified as eligible and were mailed a study questionnaire.
36/88 returned surveys
Ability to recall risk estimates that were provided post-GC. Nonuptake was 15.5 times more likely in women who could not recall their risk estimates.All women had some form of insurance.
162/250 (65%) eligible
65/162 returned surveys
Susswein et al. (2008) [4]African American women and white women with breast cancerb768529/768 (69%) underwent GTRace/ethnicity. African American women were less likely to be tested than were white women.Sample obtained from a clinical database. Testing was offered free of charge when it was not covered by insurance. This effect for time of diagnosis was significant in the African American, but not white, subgroup.
African American women: 77/132 (58%) underwent GT
Recent diagnosis. African American women who were recently diagnosed were more likely to be tested.
White women: 452/636 (71%) underwent GT
Olaya et al. (2009) [5]Patients referred for GT between 2001 and 2008b213111/213 (52%) underwent GTPersonal history of breast cancer. Having a personal history was associated with 3 times greater odds of being tested.Insurance coverage for testing was available for 91.1% (175/213) of patients. Of those who had coverage for GT, 51.4% underwent testing and 48.6% did not. Of those without coverage, 41.2% had GT and 58.9% did not.
102/213 (48%) declined GTHigher level of education. Those with a high school education or less had one-third the odds of being tested, compared with those with at least some college.
Levy et al. (2010) [6]Women aged 20–40 y with newly diagnosed early-onset breast cancer.b1,474446/1,474 (30%) underwent GTRace/ethnicity. Women of Jewish ethnicity were 3 times more likely to be tested than were non-Jewish white women. African American and Hispanic women were significantly less likely to receive testing than were non-Jewish white women.Sample obtained from a national database of commercially insured individuals.
Jewish women: 18/32 (56%) underwent GTHome location. Women living in the south were more likely to be tested than were women living in the northeast.
African American women: 10/82 (12%) underwent GTInsurance type. Women with point-of-service plans were more likely to be tested than were women with HMO plans.
Recent diagnosis. Women diagnosed in 2007 were 3.8 times more likely to be tested than were women diagnosed in 2004.
Several studies conducted in non-U.S. settings have examined the uptake of genetic testing.[7-11] In studies examining the uptake of testing among at-risk relatives of carriers of BRCA1/BRCA2 pathogenic variants, uptake rates have averaged below 50% (range, 36%–48%), with higher uptake reported among female relatives than in male relatives. Other factors associated with higher uptake of testing were not consistently reported among studies but have most commonly included being a parent and wanting to learn information about a child’s risk.

Factors influencing uptake of genetic counseling and genetic testing

In reviews that have examined the cumulative evidence concerning the predictors of uptake of BRCA1/BRCA2 genetic testing, important predictors of testing uptake include older age, Ashkenazi Jewish (AJ) heritage, unmarried status, a personal history of breast cancer, and a family history of breast cancer. Studies recruiting participants in hospital settings had significantly higher recruitment rates than did studies recruiting participants in community settings. Studies that required an immediate decision to test, rather than allowing delayed decision making, tended to report higher uptake rates.[1] However, there is evidence that women diagnosed with breast cancer are equally satisfied with genetic counseling (including information received and strength and timing of physician recommendations for counseling), whether they received genetic counseling before or after their definitive surgery for breast cancer.[12] Another review [13] found that uptake of genetic testing for BRCA1/BRCA2 pathogenic variants was related to psychological factors (e.g., anxiety about breast cancer and perceived risk of breast cancer) and demographic and medical factors (e.g., history of breast cancer or ovarian cancer, presence of children, and higher number of affected first-degree relatives [FDRs]). Family members with a known BRCA1/BRCA2 pathogenic variant were more likely to pursue testing; those with more extensive knowledge of BRCA1/BRCA2 testing, heightened risk perceptions, beliefs that mammography would promote health benefit, and high intentions to undergo testing were more likely to follow through with testing.[14]
In a review of racial/ethnic differences that affect uptake of BRCA1/BRCA2 testing, intention to undergo genetic testing in African American women was related to having at least one FDR with breast cancer or ovarian cancer, higher perceived risk of being a carrier, and less anticipatory guilt about the possibility of being a gene carrier.[15] A systematic review found that certain ethnic minority groups including African Americans and Hispanics had more negative views and greater concerns about genetic counseling and testing when compared with whites. African Americans and Hispanics were more likely to believe genetic testing could be used to show their ethnic group was inferior to other groups. Additionally, African Americans and Hispanics were found to have low awareness and knowledge about the importance of genetics in cancer, BRCA status, and genetic testing.[16]
Reasons cited for following through with testing included a desire to learn about a child's risk, to feel relief from uncertainty, to inform screening or risk-reducing surgery decisions, and to inform important life decisions such as marriage and childbearing.[14,17] Among African American women, the most important reason for testing included motivation to help other relatives decide on genetic testing.[15]
Physician recommendation may be another motivator for testing. In a retrospective study of 335 women considering genetic testing, 77% reported that they wanted the opinion of a genetics physician about whether they should be tested, and 49% wanted the opinion of their primary care provider.[18] However, there is some evidence of referral bias favoring those with a maternal family history of breast cancer or ovarian cancer. In a Canadian retrospective review of 315 patients, those with a maternal family history of breast cancer or ovarian cancer were 4.9 times (95% confidence interval, 3.6–6.7) more likely to be referred for a cancer genetics consultation by their physician than were those with a paternal family history (P < .001).[19] Studies have found that physicians may not adequately assess paternal family history [20] or may underestimate the significance of a paternal family history for genetic risk.[20-22] Other studies have shown that physician referral of patients who meet U.S. Preventive Services Task Force guidelines for BRCA genetic counseling has been suboptimal.[23]
The uptake of BRCA testing to inform surgical treatment decisions when offered appears to be high in research cohorts;[2,24] however, findings from other studies suggest that testing is underutilized in clinical practice to inform breast cancer treatment decisions.[6,25,26] Barriers to the use of BRCA testing to inform surgical treatment decisions, including lack of physician referral of newly diagnosed patients for genetic counseling, type of insurance coverage (such as Medicare or Medicaid), and challenges in the timing and coordination of testing, have been reported.[6,27-30]
Insurance coverage
Insurance coverage is an important consideration for individuals deciding whether to undergo genetic testing. (Refer to the Insurance coverage section in the PDQ summary on Cancer Genetics Risk Assessment and Counseling for more information.)

Uptake of genetic counseling and genetic testing in diverse populations

Degree of uptake of genetic counseling and genetic testing in diverse populations
There are limited data on uptake of genetic counseling and testing among nonwhite populations, and further research will be needed to define factors influencing uptake in these populations.[31] The uptake of BRCA testing appears to vary across some racial/ethnic groups. A few studies have compared uptake rates between African American and white women.[4,32] In a case-control study of women who had been seen in a university-based primary care system, African American women with family histories of breast cancer or ovarian cancer were less likely to undergo BRCA1/BRCA2 testing than were white women who had similar histories.[32] In another study among breast cancer patients who were counseled about BRCA1/BRCA2 risk in a clinical setting, lower uptake was reported among African American women than among white women.[4]
Notably, the racial differences observed in these studies do not appear to be explained by factors related to cost, access to care, risk factors for carrying a BRCA1 or BRCA2 pathogenic variant, or differences in psychosocial factors, including risk perceptions, worry, or attitudes toward testing.
Factors influencing uptake of genetic counseling and genetic testing in diverse populations
Several studies have examined uptake or “acceptance” of BRCA testing among African Americans enrolled in genetic research programs. Among study enrollees from an African American kindred in Utah, 83% underwent BRCA1 testing.[33] Age, perceived risk of being a carrier, and more extensive cancer knowledge predicted testing acceptance. Another study that recruited African American women through physician and community referrals reported a BRCA1/BRCA2 testing acceptance rate of 22%.[34] Predictors of test acceptance included having a higher probability of having a pathogenic variant, being married, and being less certain about one’s cancer risk. Finally, a third study that recruited at-risk African American women from an urban cancer screening clinic found that acceptors of BRCA testing were more knowledgeable about breast cancer genetics and perceived fewer barriers to testing, including negative emotional reactions, stigmatization concerns, and family-related guilt.[35] While these are independent predictors of genetic testing uptake, they do not explain the disparities in testing uptake across different ethnic groups. What may explain these differences are several attitudes and beliefs held about testing by individuals from diverse populations.
Work examining attitudes toward breast cancer genetic testing in Latino and African American populations indicates limited knowledge and awareness about testing but a generally receptive view once they are informed; in comparison with whites, Latino and African American populations have relatively more concerns about testing.
For example, in a qualitative study with 51 Latino individuals unselected for risk status, important findings included the fact that participants were highly interested in genetic testing for inherited cancer susceptibility, despite very limited knowledge about genetics. One important barrier involved secrecy or embarrassment about family discussions of cancer and genetics, which could be addressed in intervention strategies.[36] Another qualitative study with 54 Latina women at risk of hereditary breast cancer showed that knowledge about BRCA1/BRCA2 counseling was low, although the women were interested in learning more about counseling to gain risk information for family members. Barriers to counseling included life demands, cost, and language issues.[37]
A telephone survey of 314 patients from an inner-city network of Pittsburgh, Pennsylvania, health centers, 50% of whom were African American, found that most participants (57%) (both African Americans and whites) felt that genetic testing to evaluate disease risk was a good idea; however, more African Americans than whites thought that genetic testing would lead to racial discrimination (37% vs. 22%, respectively) and that genetics research was unethical and tampered with nature (20% vs. 11%, respectively).[38] Finally, in a study of 222 women in Savannah, Georgia, where most had neither a personal history (70%) nor a family history (60%) of breast cancer, African American women (who comprised 26% of the sample) were less likely to be aware of breast cancer genes and genetic testing. Awareness was also related to higher income, higher education level, and having a family breast cancer history. However, 74% of the entire sample expressed willingness to be tested for breast cancer susceptibility.[39]
In a sample of 146 African American women meeting criteria for BRCA1/BRCA2 pathogenic variant testing, women born outside the United States reported higher levels of anticipated negative emotional reactions (e.g., fear, hopelessness, and lack of confidence that they could emotionally handle testing). Higher levels of breast cancer–specific distress were associated with anticipated negative emotional reactions, confidentiality concerns, and anticipated guilt regarding the family impact of breast cancer genetic testing.[40] A future orientation (e.g., "I often think about how my actions today will affect my health when I am older") was associated with overall perceived benefits of breast cancer genetic testing in this population (n = 140); however, future orientation was also found to be positively associated with family-related cons of testing, including family guilt and worry regarding the impact of testing on the family.[41]
There are racial differences in provider discussion and patient uptake of genetic testing for variants in BRCA1/BRCA2. A study of women aged 18 to 64 years and diagnosed with invasive breast cancer between 2007 and 2009 found that, even after adjusting for pathogenic variant risk, African American women were less likely to report having received a physician recommendation for genetic testing. There was no difference across all races in concerns that BRCA1/BRCA2 testing was too expensive and only minimal differences in testing attitudes or insurance concerns were found, none of which influenced testing uptake.[42] A study of breast or ovarian cancer survivors (N = 50) eligible for BRCA1/BRCA2 genetic testing found that 48% were referred for genetic counseling and testing and/or had undergone genetic testing. Individuals with higher breast cancer genetics knowledge and higher self-efficacy were more likely to have engaged in genetic counseling and testing.[43] In a study of women with invasive breast cancer diagnosed before age 50 years between 2009 and 2012 who were identified through the Florida Cancer Data System state registry and eligible for BRCA1/BRCA2 genetic testing on the basis of existing guidelines, African Americans were less likely to report a discussion with their health care provider and undergo genetic testing.[44] The same study found similar overall testing rates in Hispanic (61%) and non-Hispanic (65%) whites. However, testing rates were lower among Hispanics who spoke primarily Spanish at home (50% Spanish speaking vs. 69% English speaking; P = .0009), and in general, Hispanics were less likely to have been referred for genetic testing.[45] However, this finding is not consistent across all studies. In a study of women aged 20 to 79 years with ductal carcinoma in situ or invasive breast cancer identified through the Surveillance, Epidemiology, and End Results (SEER) registry in Georgia and Los Angeles County, all eligible for BRCA1/BRCA2 genetic testing on the basis of existing guidelines, no ethnic differences were detected in receipt of genetic counseling or physician-directed discussion about genetic testing.[30]

Factors associated with declining genetic counseling and testing

There is evidence that primary reasons for declining testing involves being childless, which reduces any family motivations for testing; and concerns about the negative ramifications of testing, including difficulty retaining insurance or concerns about personal health.
Limited data are available about the characteristics of at-risk individuals who decline to be tested or have never been tested. It is difficult to access samples of test decliners because they may be reluctant to participate in research studies. Studies of genetic testing uptake are difficult to compare because people may decline at different points and with different amounts of pretest education and counseling. One study found that 43% of affected and unaffected individuals from hereditary breast/ovarian cancer families who completed a baseline interview regarding testing declined to be tested. Most individuals who declined testing chose not to participate in educational sessions. Decliners were more likely to be male and be unmarried, and have fewer relatives with breast cancer. Decliners who had high levels of cancer-related stress had higher levels of depression. Decliners lost to follow-up were significantly more likely to be affected with cancer.[46]
Another study looked at a small number (n = 13) of women decliners who carried a 25% to 50% probability of harboring a BRCA pathogenic variant; these nontested women were more likely to be childless and to have higher levels of education. This study showed that most women decided not to undergo the test after serious deliberation about the risks and benefits. Satisfaction with frequent surveillance was given as one reason for nontesting by most of these women.[47] Other reasons for declining included having no children and becoming acquainted with breast/ovarian cancer in the family relatively early in their lives.[46,47]
A third study evaluated characteristics of 34 individuals who declined BRCA1/BRCA2 testing in a large multicenter study in the United Kingdom. Decliners were younger than a national sample of test acceptors, and female decliners had lower mean scores on a measure of cancer worry. Although 78% of test decliners/deferrers felt that their health was at risk, they reported that learning about their BRCA1/BRCA2 pathogenic variant status would cause them to worry about the following:
  • Their children's health (76%).
  • Their life insurance (60%).
  • Their own health (56%).
  • Loss of their job (5%).
  • Receiving less screening if they did not carry a BRCA1/BRCA2 pathogenic variant (62%).
Apprehension about the impact of the test result was a more important factor in the decision to decline testing than were concrete burdens such as time required to travel to a genetics clinic and time spent away from work, family, and social obligations.[48] In 15% (n = 31) of individuals from 13 hereditary breast and ovarian cancer families who underwent genetic education and counseling and declined testing for a documented pathogenic variant in the family, positive changes in family relationships were reported—specifically, greater expressiveness and cohesion—compared with those who pursued testing.[49]

Genetic counseling and testing in children

Testing for BRCA1/BRCA2 pathogenic variants has been almost universally limited to adults older than 18 years. The risks of testing children for adult-onset disorders, such as breast and ovarian cancers, as inferred from developmental data on children’s medical understanding and ability to provide informed consent, have been outlined in several reports.[50-53]
Studies suggest that persons who have undergone BRCA1/BRCA2 genetic testing or who are adult offspring of persons who have had testing are generally not in favor of testing minors.[54,55] Although the data are limited, research suggests that males, pathogenic variant noncarriers, and those whose mothers did not have personal histories of breast cancer may be more likely to favor genetic testing in minors in general.[54] Of those who had minor children at the time the study was conducted, only 17% stated a preference for having their own children tested. Concerns regarding testing of minors included psychological risks and insufficient maturity. Potential benefits included the ability to influence health behaviors.[55]
No data exist on the testing of children for BRCA1/BRCA2 pathogenic variants, although some researchers believe it is necessary to test the validity of assumptions underlying the general prohibition of testing children for genetic variants associated with breast and ovarian cancers and other adult-onset diseases.[56-58] In one study, 20 children (aged 11–17 y) of a selected group of mothers undergoing genetic testing (80% of whom previously had breast cancer and all of whom had discussed BRCA1/BRCA2 testing with their children) completed self-report questionnaires on their health beliefs and attitudes toward cancer, feelings related to cancer, and behavioral problems.[59] Ninety percent of children thought they would want cancer risk information as adults; half worried about themselves or a family member developing cancer. There was no evidence of emotional distress or behavioral problems.

What People Bring to Genetic Testing: Impact of Risk Perception, Health Beliefs, and Personality Characteristics

The emerging literature in this area suggests that risk perceptions, health beliefs, psychological status, and personality characteristics are important factors in decision making about breast/ovarian cancer genetic testing. Many women presenting at academic centers for BRCA1/BRCA2 testing arrive with a strong belief that they have a pathogenic variant, having decided they want genetic testing, but possessing little information about the risks or limitations of testing.[60] Most mean scores of psychological functioning at baseline for subjects in genetic counseling studies were within normal limits.[61] Nonetheless, a subset of subjects in many genetic counseling studies present with elevated anxiety, depression, or cancer worry.[62,63] Identification of these individuals is essential to prevent adverse outcomes. In a study of 205 women pursuing genetic counseling, interactions among cancer worry, breast cancer risk perception, and perceived severity of having a breast cancer genetic variant were found such that those with high worry, high breast cancer risk perception, and low perceived severity were twice as likely to follow through with BRCA1/BRCA2 testing than others.[64]
A general tendency to overestimate inherited risk of breast and ovarian cancer has been noted in at-risk populations,[65-68] in cancer patients,[66,69,70] in spouses of breast and ovarian cancer patients,[71] and among women in the general population.[72-74] but underestimation of breast cancer risk in higher-risk and average-risk women also has been reported.[75] This overestimation may encourage a belief that BRCA1/BRCA2 genetic testing will be more informative than it is currently thought to be. Some evidence exists that even counseling does not dissuade women at low to moderate risk from the belief that BRCA1 testing could be valuable.[31] Overestimation of both breast and ovarian cancer risk has been associated with nonadherence to physician-recommended screening practices.[76,77] A meta-analysis of 12 studies of outcomes of genetic counseling for breast/ovarian cancer showed that counseling improved the accuracy of risk perception.[78]
Women appear to be the prime communicators within families about the family history of breast cancer.[79] Higher numbers of maternal versus paternal transmission cases are reported,[80] likely due to family communication patterns, to the misconception that breast cancer risk can only be transmitted through the mother, and to the greater difficulty in recognizing paternal family histories because of the need to identify more distant relatives with cancer. In an analysis of 2,505 women participating in the Family Healthware Impact Trial,[81] not only was evidence of underreporting of paternal family history identified, but also women reported a lower level of perceived breast cancer risk with a paternal versus maternal breast cancer family history.[82] Physicians and counselors taking a family history are encouraged to elicit paternal and maternal family histories of breast, ovarian, or other associated cancers.[79]
The accuracy of reported family history of breast or ovarian cancer varies; some studies found levels of accuracy above 90%,[83,84] with others finding more errors in the reporting of cancer in second-degree or more distant relatives [85] or in age of onset of cancer.[86] Less accuracy has been found in the reporting of cancers other than breast cancer. Ovarian cancer history was reported with 60% accuracy in one study compared with 83% accuracy in breast cancer history.[87] Providers should be aware that there are a few published cases of Munchausen syndrome in reporting of false family breast cancer history.[88] Much more common is erroneous reporting of family cancer history due to unintentional errors or gaps in knowledge, related in some cases to the early death of potential maternal informants about cancer family history.[79] (Refer to the Taking a Family History section of the Cancer Genetics Risk Assessment and Counseling summary for more information.)
Targeted written,[89,90] video, CD-ROM, interactive computer programs and websites,[91-98] and culturally targeted educational materials [99-101] may be effective and efficient methods of increasing knowledge about the pros and cons of genetic testing. Such supplemental materials may allow more efficient use of the time allotted for pretest education and counseling by genetics and primary care providers and may discourage individuals without appropriate indication of risk from seeking genetic testing.[89]

Genetic Counseling for Hereditary Predisposition to Breast Cancer

Counseling for breast cancer risk typically involves individuals with family histories that are potentially attributable to BRCA1 or BRCA2. It also, however, may include individuals with family histories of Li-Fraumeni syndrome, ataxia-telangiectasia, Cowden syndrome, or Peutz-Jeghers syndrome.[102] (Refer to the High-Penetrance Breast and/or Gynecologic Cancer Susceptibility Genes section of this summary for more information.)
Management strategies for carriers may involve decisions about the nature, frequency, and timing of screening and surveillance procedures, chemoprevention, risk-reducing surgery, and use of hormone replacement therapy (HRT). The utilization of breast conservation and radiation as cancer therapy for women who are carriers may be influenced by knowledge of pathogenic variant status. (Refer to the Clinical Management of Carriers of BRCA Pathogenic Variants section of this summary for more information.)
Counseling also includes consideration of related psychosocial concerns and discussion of planned family communication and the responsibility to warn other family members about the possibility of having an increased risk of breast, ovarian, and other cancers. Data suggest that individual responses to being tested as adults are influenced by the results status of other family members.[103,104] Management of anxiety and distress are important not only as quality-of-life factors, but also because high anxiety may interfere with the understanding and integration of complex genetic and medical information and adherence to screening.[105-107] Formal, objective evaluation of these outcomes are well documented. (Refer to the Emotional Outcomes and Behavioral Outcomes sections of this summary for more information.)
Published descriptions of counseling programs for BRCA1 (and subsequently for BRCA2) testing include strategies for gathering a family history, assessing eligibility for testing, communicating the considerable volume of relevant information about breast/ovarian cancer genetics and associated medical and psychosocial risks and benefits, and discussion of specialized ethical considerations about confidentiality and family communication.[108-115] Participant distress, intrusive thoughts about cancer, coping style, and social support were assessed in many prospective testing candidates. The psychosocial outcomes evaluated in these programs have included changes in knowledge about the genetics of breast/ovarian cancer after counseling, risk comprehension, psychological adjustment, family and social functioning, and reproductive and health behaviors.[116] A Dutch study of communication processes and satisfaction levels of counselees going through cancer genetic counseling for inherited cancer syndromes indicated that asking more medical questions (by the counselor), providing more psychosocial information, and longer eye contact by the counselor were associated with lower satisfaction levels. The provision of medical information by the counselor was most highly related to satisfaction and perception that needs have been fulfilled.[117]
Many of the psychosocial outcome studies involve specialized, highly selected research populations, some of which were utilized to map and clone BRCA1 and BRCA2. One such example is K2082, an extensively studied kindred of more than 800 members of a Utah Mormon family in which a BRCA1 pathogenic variant accounts for the observed increased rates of breast and ovarian cancer. A study of the understanding that members of this kindred have about breast/ovarian cancer genetics found that, even in breast cancer research populations, there was incomplete knowledge about associated risks of colon and prostate cancer, the existence of options for RRM and RRSO, and the complexity of existing psychosocial risks.[108] A meta-analysis of 21 studies found that genetic counseling was effective in increasing knowledge and improved the accuracy of perceived risk. Genetic counseling did not have a statistically significant long-term impact on affective outcomes including anxiety, distress, or cancer-specific worry and the behavioral outcome of cancer surveillance activities.[61] These prospective studies, however, were characterized by a heterogeneity of measures of cancer-specific worry and inconsistent findings in effects of change from baseline.[61]

Emotional Outcomes

Although there were initial concerns about the possibility of adverse emotional consequences from BRCA testing, most studies conducted over the years have shown low levels of psychological distress among both carriers and noncarriers, particularly over the longer term.[118-120] In a meta-analysis examining cancer-specific distress over short (0–4 weeks), moderate (5–24 weeks), and long (25–52 weeks) periods of time since the receipt of testing results, carriers were found to demonstrate increased levels of distress shortly after receiving results, with levels returning to baseline within moderate and long periods of time.[118] In contrast, noncarriers and those with inconclusive results showed reduced levels of distress over time.[118,121] Psychological distress patterns were found to vary as a function of several factors, including the cancer history of the individual and the country within which the study was conducted. Carriers with a personal history of cancer experienced small decreases in distress over time, whereas no changes were observed among carriers without a personal history of cancer. Among individuals with inconclusive results, greater decreases in distress were observed among those without a cancer history than among those with a cancer history. Among noncarriers, those in the United States experienced significantly greater decreases in psychological distress than noncarriers from Europe and Australia. A study conducted in Austria noted that certain subgroups of counselees experienced greater distress, including those who were older, had a more recent cancer diagnosis, or those who had received counseling but declined BRCA testing.[122]
Several studies have reported on emotional outcomes over longer follow-up periods (i.e., greater than 12 months after disclosure) than those reported in the meta-analysis described above.[118] In a U.K. study, cancer-related worry did not differ between carriers and noncarriers at 3 years of follow-up.[123] Two U.S.-based studies published since the meta-analytic review [118] have reported similar findings among women who were surveyed more than 3 years after receipt of BRCA test results.[124,125] In a cross-sectional study,[124] 167 women who were surveyed more than 4 years after receiving BRCA test results reported low levels of genetic testing–specific concerns, as measured using the Multidimensional Impact of Cancer Risk Assessment Scale.[126] In multivariate regression models, carriers of pathogenic variants were significantly more likely to experience distress than were noncarriers. In a second study,[125] 464 women were followed prospectively for a median of 5 years (range, 3.4–9.1 y) after testing. Among both affected and unaffected participants, BRCA carriers reported significantly higher levels of distress, uncertainty (affected only), perceived stress (affected only), and lower positive testing experiences (unaffected only) than women who received negative results for a known pathogenic variant in the family.[125] Although both studies [124,125] reported greater distress among BRCA carriers than among noncarriers, the level of distress was not reflective of clinically significant dysfunction.
Although most studies have reported that a positive BRCA test result has a relatively minimal impact on psychological distress, many of these studies were conducted among families with a strong family history of breast or ovarian cancer who underwent extensive pretest genetic counseling. Therefore, emotional responses may not generalize to individuals who test under different contexts. For example, individuals who are tested with population BRCA screening may not have a family history of cancer.[127-129] Although pretest genetic counseling is recommended, this is not always done when genetic testing is ordered by nongenetic providers [130] or directly through commercial companies.[131,132]
For example, in a Canadian study of 2,080 Jewish women who participated in a population-based genetic screening study to test for three BRCA pathogenic variants common in families of Jewish heritage, women were not offered in-person genetic counseling but were given a pamphlet on genetic testing for BRCA1/BRCA2 before they provided a DNA sample. One year after genetic testing, women who were positive for a pathogenic variant (n = 18) showed significant increases in cancer-specific distress, whereas no changes in distress were observed among women who were negative for a pathogenic variant.[128] The mean distress score on the Impact of Event Scale for the 18 women with a known pathogenic variant was 25.3 (range, 2–51); 10 of 18 women (56%) scored within moderate (26–43) (n = 7) or severe (44+) (n = 3) ranges. It is unclear from this study whether the increase in distress observed at 1 year of follow-up was due to the lack of in-person genetic counseling, or whether the lower levels of distress at baseline observed were because the women in the study were low risk but eligible for testing because of their ancestry. A follow-up study with this cohort found that distress decreased between 1 to 2 years after testing and that changes in distress varied by risk-reduction options undertaken by carriers. Specifically, those who had undergone risk-reducing mastectomy or oophorectomy experienced significant decreases in distress compared with those who did not have either surgery.[129] Another smaller qualitative study also supports these findings.[133]
Similarly, the impact of direct-to-consumer (DTC) BRCA testing through commercial companies requires further evaluation. Case studies have reported adverse emotional responses after receipt of a positive BRCA result from DTC genetic testing, suggesting the need for further evaluation of the emotional outcomes of women undergoing genetic testing through commercial companies.[131,132] Only one study, conducted by a commercial company, has attempted to evaluate the impact of BRCA testing in this context.[134] A total of 32 individuals (16 women and 16 men) who tested positive for one of three BRCA founder pathogenic variants common in Ashkenazi Jews completed semi-structured interviews. None of the carriers reported extreme anxiety, although some experienced moderate anxiety (13%) or initial disappointment and anxiety that dissipated over time (28%). These findings should be interpreted with caution given that only 24% (32 of 136) of invited carriers of BRCA pathogenic variants participated in the study, raising concerns about selection bias.
Despite evidence of a short-term increase in distress after the receipt of genetic testing results, any adverse responses to a positive carrier status dissipate within 12 months.[118] Additional research is needed to examine emotional outcomes for those who are not provided genetic counseling before testing.[130]

Emotional outcomes in newly diagnosed breast cancer patients

It is increasingly common for women with breast cancer to pursue genetic counseling and testing at the time of diagnosis to assist with treatment decision making. (Refer to the Benefits of offering genetic testing at the time of cancer diagnosis section in the Introduction section of this summary for more information.) Although concerns have been raised about the adverse psychological implications of offering rapid genetic counseling and testing between diagnosis and surgery,[135,136] other studies,[137-139] including a randomized trial,[140] have provided evidence indicating no additional adverse psychological effects in newly diagnosed breast cancer patients. One randomized controlled trial found that patients undergoing rapid genetic counseling and testing felt more actively involved in treatment decision making than those receiving standard care.[141] However, qualitative research on 20 newly diagnosed breast cancer patients found that some subgroups of these patients may have more difficulty coping with BRCA test results, such as carriers who have no family history of cancer; those who do not have an affected relative with whom they can identify; and higher risk women who receive uninformative negative BRCA results.[135]

Family Effects

Family communication about genetic testing and hereditary risk

Family communication about genetic testing for cancer susceptibility, and specifically about the results of BRCA1/BRCA2 genetic testing, is complex. Gender appears to be an important variable in family communication and psychological outcomes. Studies have documented that female carriers are more likely to disclose their status to other family members (especially sisters and children aged 14–18 y) [142] than are male carriers.[142,143] Among males, noncarriers were more likely than carriers to tell their sisters and children the results of their tests. BRCA1/BRCA2 carriers who disclosed their results to sisters had a slight decrease in psychological distress, compared with a slight increase in distress for carriers who chose not to tell their sisters. One study found that men reported greater difficulty disclosing a known pathogenic variant to family members than women (90% vs. 70%).[144]
Family communication of BRCA1/BRCA2 test results to relatives is another factor affecting participation in testing. There have been more studies of communication with FDRs and second-degree relatives than with more distant family members. Studies have investigated the process and content of communication among sisters about BRCA1/BRCA2 test results.[145,146] Study results suggest that both carriers of pathogenic variants [145] and women with uninformative results [145,146] communicate with sisters to provide them with genetic risk information. Similar findings were reported in women with uninformative results disclosing test results to their daughters.[146] Among relatives with whom genetic test results were not discussed, the most important reason given was that the affected women were not close to their relatives [145] or had a poor relationship with them.[146] Studies found that women with a BRCA pathogenic variant more often shared their results with their mother and adult sisters and daughters than with their father and adult brothers and sons.[79,147-150] A study that evaluated communication of test results to FDRs at 4 months postdisclosure found that women aged 40 years or older were more likely to inform their parents of test results compared with younger women. Participants also were more likely to inform brothers of their results if the BRCA pathogenic variant was inherited through the paternal line.[148] Another study found that disclosure was limited mainly to FDRs, and dissemination of information to distant relatives was problematic.[151] Age was a significant factor in informing distant relatives with younger patients being more willing to communicate their genetic test result.[145,147,151] Additionally, one study found that lower genetic worry, higher interest in genomic information, carrying a BRCA1 or BRCA2 pathogenic variant, or having never been married was associated with communication to more family members.[143] In contrast, a longer time interval since diagnosis was associated with communication to fewer family members.[143]
A few in-depth qualitative studies have looked at issues associated with family communication about genetic testing. Although the findings from these studies may not be generalizable to the larger population of at-risk persons, they illustrate the complexity of issues involved in conveying hereditary cancer risk information in families.[152] On the basis of 15 interviews conducted with women attending a familial cancer genetics clinic, the authors concluded that while women felt a sense of duty to discuss genetic testing with their relatives, they also experienced conflicting feelings of uncertainty, respect, and isolation. Decisions about whom in the family to inform and how to inform them about hereditary cancer and genetic testing may be influenced by tensions between women's need to fulfill social roles and their responsibilities toward themselves and others.[152] Another qualitative study of 21 women who attended a familial breast and ovarian cancer genetics clinic suggested that some women may find it difficult to communicate about inherited cancer risk with their partners and with certain relatives, especially brothers, because of those persons’ own fears and worries about cancer.[150] This study also suggested that how genetic risk information is shared within families may depend on the existing norms for communicating about cancer in general. For example, family members may be generally open to sharing information about cancer with each other, may selectively avoid discussing cancer information with certain family members to protect themselves or other relatives from negative emotional reactions, or may ask a specific relative to act as an intermediary to disclosure of information to other family members.[153] The potential importance of persons outside the family, such as friends, as both confidantes about inherited cancer risk information and as sources of support for coping with this information was also noted in the study.[150]
A study of 31 mothers with a documented BRCA pathogenic variant explored patterns of dissemination to children.[154] Of those who chose to disclose test results to their children, age of offspring was the most important factor. Fifty percent of the children who were told were aged 20 to 29 years and slightly more than 25% of the children were aged 19 years or younger. Sons and daughters were notified in equal numbers. More than 70% of mothers informed their children within a week of learning their test result. Ninety-three percent of mothers who chose not to share their results with their children indicated that it was because their children were too young. These findings were consistent with three other studies showing that children younger than 13 years were less likely to be informed about test results compared with older children.[148,155,156] Another study of 187 mothers undergoing BRCA1/BRCA2 testing evaluated their need for resources to prepare for a facilitated conversation about sharing their BRCA1/BRCA2 testing results with their children. Seventy-eight percent of mothers were interested in three or more resources, including literature (93%), family counseling (86%), talk to prior participants (79%), and support groups (54%).[155]
A longitudinal study of 153 women self-referred for genetic testing for BRCA1 and BRCA2 pathogenic variants and 118 of their partners evaluated communication about genetic testing and distress before testing and at 6 months posttesting.[157] The study found that most couples discussed the decision to undergo testing (98%), most test participants felt their partners were supportive, and most women disclosed test results to their partners (97%, n = 148). Test participants who felt their partners were supportive during pretest discussions experienced less distress after disclosure, and partners who felt more comfortable sharing concerns with test participants pretest experienced less distress after disclosure. Six-month follow-up revealed that 22% of participants felt the need to talk about the testing experience with their partners in the week before the interview. Most participants (72%, n = 107) reported comfort in sharing concerns with their partners, and 5% (n = 7) reported relationship strain as a result of genetic testing. In couples in which the woman had a positive genetic test result, more relationship strain, more protective buffering of their partners, and more discussion of related concerns were reported than in couples in which the woman had a true-negative or uninformative result.[157]
A study of 561 FDRs of women who had undergone BRCA1/BRCA2 genetic testing found that 22% of FDRs did not recall being informed of the genetic test results despite the women reporting that the results had been shared.[158] Men were less likely to recall receiving the results (P > .001). Of those with recall about receiving the test results, 10.5% of FDRs did not recall the findings. For those with recall of the results, 17.9% of FDRs had an interpretation that was discordant with the correct results. Accuracy of test results recall was greater for informative test results (those that were either true positive or true negative) (P = .029). However, regardless of the test results, FDRs perceived the cancer risk to be higher before they learned of the findings than after (74% and 53% of FDRs reported that they believed their risk for cancer was greater than average before and after hearing test results, respectively).
There is a small but growing body of literature regarding psychological effects in men who have a family history of breast cancer and who are considering or have had BRCA testing. A qualitative study of 22 men from 16 high-risk families in Ireland revealed that more men in the study with daughters were tested than men without daughters. These men reported little communication with relatives about the illness, with some men reporting being excluded from discussion about cancer among female family members. Some men in the study also reported actively avoiding open discussion with daughters and other relatives.[159] In contrast, a study of 59 men testing positive for a BRCA1/BRCA2 pathogenic variant found that most men participated in family discussions about breast and/or ovarian cancer. However, fewer than half of the men participated in family discussions about risk-reducing surgery. The main reason given for having BRCA testing was concern for their children and a need for certainty about whether they could have transmitted the pathogenic variant to their children. In this study, 79% of participating men had at least one daughter. Most of these men described how their relationships had been strengthened after receipt of BRCA results, helping communication in the family and greater understanding.[160] Men in both studies expressed fears of developing cancer themselves. Irish men especially reported fear of cancer in sexual organs.

Family functioning

One study assessed 212 individuals from 13 hereditary breast and ovarian cancer families who received genetic counseling and were offered BRCA1/BRCA2 testing for documented pathogenic variants in the family. Individuals who were not tested were found 6 to 9 months later to have significantly greater increases in family expressiveness and cohesiveness compared with those who were tested. Persons who were randomly assigned to a client-centered versus problem-solving genetic counseling intervention had a significantly greater reduction in conflict, regardless of the test decision.[49]

Partners of high-risk women

Many studies have looked at the psychological effects in women of having a high risk of developing cancer, either on the basis of carrying a BRCA1/BRCA2 pathogenic variant or having a strong family history of cancer. Some studies have also examined the effects on the partners of such women.
A Canadian study assessed 59 spouses of women found to have a BRCA1/BRCA2 pathogenic variant. All were supportive of their spouses’ decision to undergo genetic testing and 17% wished they had been more involved in the genetic testing process. Spouses who reported that genetic testing had no impact on their relationship had long-term relationships (mean duration 27 years). Forty-six percent of spouses reported that their major concern was of their partner dying of cancer. Nineteen percent were concerned their spouse would develop cancer and 14% were concerned their children would also be carriers of BRCA1/BRCA2 pathogenic variants.[161]
In a U.S. study, 118 partners of women who underwent genetic testing for pathogenic variants in BRCA1 and BRCA2 completed a survey before testing and then again 6 months after result disclosure. At 6 months, only 10 partners reported that they had not been told of the test result. Ninety-one percent reported that the testing had not caused strain on their relationship. Partners who were comfortable sharing concerns before testing experienced less distress after testing. Protective buffering was not found to impact distress levels of partners.[157]
An Australian study of 95 unaffected women at high risk of developing breast and/or ovarian cancer (13 carriers of pathogenic variants and 82 with unknown variant status) and their partners showed that although the majority of male partners had distress levels comparable to a normative population sample, 10% had significant levels of distress that indicated the need for further clinical intervention. Men with a high monitoring coping style and greater perceived breast cancer risk for their wives reported higher levels of distress. Open communication between the men and their partners and the occurrence of a cancer-related event in the wife’s family in the last year were associated with lower distress levels. When men were asked what kind of information and support they would like for themselves and their partners, 57.9% reported that they would like more information about breast and ovarian cancer, and 32.6% said they would like more support in dealing with their partner's risk. Twenty-five percent of men had suggestions on how to improve services for partners of high-risk women, including strategies on how to best support their partner, greater encouragement from health care professionals to attend appointments, and meeting with other partners.[162]
A review of this literature reported that the BRCA testing process may be distressing for male partners, particularly for those with spouses identified as carriers. Male partner distress appears to be associated with their beliefs about the woman’s breast cancer risk, lack of couple communication, and feelings of alienation from the testing process.[163]
At-risk males
A review of the literature on the experiences of males in families with a known BRCA1 and BRCA2 pathogenic variant reported that while the data are limited, men from variant-positive families are less likely than females to participate in communication regarding genetics at every level, including the counseling and testing process. Men are less likely to be informed of genetic test results received by female relatives, and most men from these families do not pursue their own genetic testing.[164]
A study of Dutch men at increased risk of having inherited a BRCA1 pathogenic variant reported a tendency for the men to deny or minimize the emotional effects of their risk status, and to focus on medical implications for their female relatives. Men in these families, however, also reported considerable distress in relation to their female relatives.[165] In another study of male psychological functioning during breast cancer testing, 28 men belonging to 18 different high-risk families (with a 25% or 50% risk of having inherited a BRCA1/BRCA2 pathogenic variant) participated. The study purpose was to analyze distress in males at risk of carrying a BRCA1/BRCA2 pathogenic variant who applied for genetic testing. Of the men studied, most had low pretest distress; scores were lowest for men who were optimistic or who did not have daughters. Most carriers of pathogenic variants had normal levels of anxiety and depression and reported no guilt, though some anticipated increased distress and feelings of responsibility if their daughters developed breast or ovarian cancer. None of the noncarriers reported feeling guilty.[166] In one study,[160] adherence to recommended screening guidelines after testing was analyzed. In this study, more than half of male carriers of pathogenic variants did not adhere to the screening guidelines recommended after disclosure of genetic test results. These findings are consistent with those for female carriers of BRCA1/BRCA2 pathogenic variants.[160,167]
A multicenter U.K. cohort study examined prospective outcomes of BRCA1/BRCA2 testing in 193 individuals, of which 20% were men aged 28 to 86 years. Men’s distress levels were low, did not differ among carriers and noncarriers, and did not change from baseline (before genetic testing) to the 3-year follow-up. Twenty-two percent of male carriers of pathogenic variants received colorectal cancer screening and 44% received prostate cancer screening;[123] however, it is unclear whether men in this study were following age-appropriate screening guidelines.

Children

Several studies have explored communication of BRCA test results to at-risk children. Across all studies, the rate of disclosure to children ranging in age from 4 to 25 years is approximately 50%.[147,148,151,155,168-171] In general, age of offspring was the most important factor in deciding whether to disclose test results. In one study of 31 mothers disclosing their BRCA test results, 50% of the children who were informed of the results were aged 20 to 29 years and slightly more than 25% of the children were aged 19 years or younger. Sons and daughters were notified in equal numbers.[154] Similarly, in another study of 42 female carriers of BRCA pathogenic variants, 83% of offspring older than age 18 years were told of the results, while only 21% of offspring aged 13 years or younger were told.[155]
Several studies have also looked at the timing of disclosure to children after parents receive their test results. Although the majority of children were told within a week to several months after results disclosure,[148,154,155] some parents chose to delay disclosure.[155] Reasons for delaying disclosure included waiting for the child to get older, allowing time for the parent to adjust to the information, and waiting until results could be shared in person (in the case of adult children living away from home).[155]
In one study, participants who told children younger than 13 years about their carrier status had increased distress, and those who did not tell their young children experienced a slight decrease in distress. Communication with young children was found to be influenced by developmental variables such as age and style of parent/child communication.[170]
One study looked at the reaction of children to results disclosure or the effect on the parent-child relationship of communicating the results.[155] With regard to offspring’s understanding of the information, almost half of parents from one study reported that their child did not appear to understand the significance of a positive test result, although older children were reported to have a better understanding. This same study also showed that 48% of parents reported at least one negative reaction in their child, ranging from anxiety or concern (22%) to crying and fear (26%). It should be noted, however, that in this study children's level of understanding and reactions to the test result were measured qualitatively and based only on the parents' perception. Also, given the retrospective design of the study, there was a potential for recall bias. There were no significant differences in emotional reaction depending on age or gender of the child. Lastly, 65% of parents reported no change in their relationship with their child, while 5 parents (22%) reported a strengthening of their relationship.
Interestingly, a large multicenter study of 869 mother-daughter pairs (the daughters were aged 6 to 13 y) found that girls with a family history of breast cancer or a familial BRCA1/BRCA2 pathogenic variant (BCFH+) compared with those without such family histories had better psychosocial adjustment by maternal report.[172] However, based on a combination of maternal report and direct assessment of girls aged 10 to 13 years, BCFH+ girls experienced greater breast cancer–specific distress and a higher perceived risk of breast cancer than their peers without such family histories. Moreover, higher daughter distress was associated with higher maternal distress. A similarly designed study in older girls, aged 11 to 19 years, found that higher breast cancer–specific distress in daughters was associated with perceived risk and maternal distress. This older age group had higher self-esteem than did their peers without a family history of breast cancer.[173]
Another study of 187 mothers undergoing BRCA1/BRCA2 testing evaluated their need for resources to prepare for a facilitated conversation about sharing their BRCA1/BRCA2 testing results with their children. Seventy-eight percent of mothers were interested in three or more resources, including literature (93%), family counseling (86%), talking to prior participants (79%), and support groups (54%).[174]
Testing for BRCA1/BRCA2 has been almost universally limited to adults older than 18 years. The risks of testing children for adult-onset disorders (such as breast and ovarian cancer), as inferred from developmental data on children’s medical understanding and ability to provide informed consent, have been outlined in several reports.[50-53] Surveys of parental interest in testing children for adult-onset hereditary cancers suggest that parents are more eager to test their children than to be tested themselves for a breast cancer gene, suggesting potential conflicts for providers.[175,176] In a general population survey in the United States, 71% of parents said that it was moderately, very, or extremely likely that if they carried a breast-cancer predisposing pathogenic variant, they would test a 13-year-old daughter now to determine her breast cancer gene status.[175] To date, no data exist on the testing of children for BRCA1/BRCA2, though some researchers believe it is necessary to test the validity of assumptions underlying the general prohibition of testing of children for breast/ovarian cancer and other adult-onset disease genes.[56-58] In one study, 20 children (aged 11–17 y) of a selected group of mothers undergoing genetic testing (80% of whom previously had breast cancer and all of whom had discussed BRCA1/BRCA2 testing with their children) completed self-report questionnaires on their health beliefs and attitudes toward cancer, feelings related to cancer, and behavioral problems.[59] Ninety percent of children thought they would want cancer risk information as adults; half worried about themselves or a family member developing cancer. There was no evidence of emotional distress or behavioral problems. Another study by this group [170] found that 1 month after disclosure of BRCA1/BRCA2 genetic test results, 53% of 42 enrolled mothers of children aged 8 to 17 years had discussed their result with one or more of their children. Age of the child rather than pathogenic variant status of the mother influenced whether they were told, as did family health communication style.

Prenatal diagnosis and preimplantation genetic testing

The possibility of transmitting a pathogenic variant to a child may pose a concern to families affected by hereditary breast and ovarian cancer (HBOC),[177] perhaps to the extent that some carriers may avoid childbearing.[178,179] These concerns also may prompt women to consider using prenatal testing methods to help reduce the risk of transmission.[177,180] Prenatal diagnosis is an encompassing term used to refer to any medical procedure conducted to assess the presence of a genetic disorder in a fetus. Methods include amniocentesis and chorionic villous sampling (CVS).[181,182] Both procedures carry some risk of miscarriage and some evidence suggests fetal defects may result from using these tests.[181,182] Moreover, discovering the fetus is a carrier for a genetic defect may impose a difficult decision for couples regarding pregnancy continuation or termination. An alternative to these tests is preimplantation genetic testing (PGT), a procedure used to test fertilized embryos for genetic disorders before uterine implantation,[177,183,184] thereby avoiding the potential dangers associated with amniocentesis and CVS and the decision to terminate a pregnancy. Using the information obtained from the genetic testing, potential parents can decide whether or not to implant. PGT can be used to detect pathogenic variants in hereditary cancer predisposing genes, including BRCA.[177,180]
In the United States, a series of studies has evaluated awareness, interest (e.g., would consider using PGT), and attitudes related to PGT among members of Facing Our Risk of Cancer Empowered (FORCE), an advocacy organization focused on persons at increased risk of HBOC.[177,180,185] The first study was a Web-based survey of 283 members,[177] the second included 205 attendees of the 2007 annual FORCE conference,[180] and the third was a Web-based survey of 962 members.[185,186] These studies have documented low levels of awareness, with 20% to 32% of study respondents reporting having heard of PGT before study participation.[180,185] With respect to interest in PGT, the first study [177] found only 13% of women would be likely to use PGT, whereas, 33% of respondents in the subsequent FORCE studies reported that they would consider using PGT.[180,185] In the third FORCE-based study (n = 962),[185] multivariable analysis revealed PGT interest was associated with the desire to have more children, having previously had any prenatal genetic test, and previous awareness of PGT. Attitudinal predictors of interest in PGT included agreement that others at risk of HBOC should be offered PGT; the belief that PGT is acceptable for persons at risk of HBOC; the belief that PGT information should be given to individuals at risk of HBOC; and endorsement of PGT benefits of having children without genetic variants and eliminating genetic diseases. Conversely, those who indicated that PGT was “too much like playing God” and reported that they considered PGT in the context of religion, had less interest in PGT.
It is unknown whether the attitudes of FORCE members toward PGT are representative of the majority of BRCA carriers. A cross-sectional study of 1,081 BRCA carriers, 65% of whom were recruited through FORCE and the remainder by the University of Pennsylvania, revealed that a majority of carriers were in favor of offering PGT and prenatal diagnosis to carriers (59% for PGT and 55.5% for prenatal diagnosis).[187] Of those who indicated that their families were not complete, 41% of BRCA carriers reported that their carrier status impacted their decision about future biological children. This study also revealed that 21.5% of unpartnered BRCA carriers felt more pressure to get married.
The U.K. Human Fertilization and Embryology authority has approved the use of PGT for hereditary breast and ovarian cancer. In a sample of 102 women with a BRCA pathogenic variant, most were supportive of PGT but only 38% of the women who had completed their families would consider it for themselves had PGT been available, and only 14% of women who were contemplating a future pregnancy would consider PGT.[188] In a study of 77 individuals undergoing BRCA testing as part of a multicenter cohort study in Spain, 61% of respondents reported they would consider PGT. Factors associated with PGT interest were age 40 years and older and had a prior cancer diagnosis.[189]
In France, couples who obtain authorization from a multidisciplinary prenatal diagnosis team may access PGT free of charge as a benefit of their national health care system. However, no BRCA carriers have been authorized to use PGT. In a national study of 490 unaffected carriers of BRCA pathogenic variants of childbearing age (women aged 18–49 y; men aged 18–69 y), 16% stated that BRCA test results had altered their ongoing plans for childbearing.[190] Upon qualitative analysis of written comments provided by some respondents, the primary impact was related to accelerating the timing of pregnancy, feelings of guilt about possibly passing on the pathogenic variant to offspring, and having future children. In response to a hypothetical scenario in which PGT was readily available, 33% of participants reported that they would undergo PGT. Factors associated with this intention were having no future reproductive plans at the time of the survey, feeling pregnancy termination was an acceptable option in the context of identifying a BRCA pathogenic variant, and having fewer cases of breast and/or ovarian cancer in the family. When presented with questions about expectations about delivery of PGT or prenatal diagnosis (PND) information, 85% of respondents felt it should be provided along with BRCA test results; 45% felt that it should be provided when carriers decide to have children. Respondents stated that they would expect this information to be delivered by cancer geneticists (92%), obstetrician/gynecologists (76%), and general practitioners (48%).
A small (N = 25) qualitative study of women of reproductive age positive for a BRCA pathogenic variant who underwent genetic testing before having children evaluated how their BRCA status influenced their attitudes about reproductive genetic testing (both PGT and PND) and decisions about having children.[191] In this study, the decision to undergo BRCA testing was primarily motivated by the desire to manage one’s personal cancer risk, rather than a desire to inform future reproductive decisions. The perceived severity of HBOC influenced concerns about passing on a BRCA pathogenic variant to children and also influenced willingness to consider PGT or PND and varied based on personal experience. Most did not believe that a known BRCA pathogenic variant was a reason to terminate a pregnancy. As observed in prior studies, knowledge of reproductive options varied; however, there was a tendency among participants to view PGT as more acceptable than PND with regard to termination of pregnancy. Decisions regarding the pros and cons of PGT versus PND with termination of pregnancy were driven primarily by personal preferences and experiences, rather than by morality judgments. For example, women were deterred from PGT based on the need to undergo in vitro fertilization and to take hormones that might increase cancer risk and based on the observed experiences of others who underwent this procedure.
One study has examined these issues among high-risk men recruited from FORCE and Craigslist (a bulletin board website) (N = 228).[192] Similar to the previous studies of women, only 20% of men were aware of PGT before survey participation. In a multivariate analysis, those who selected the “other” option for possible benefits of PGT compared with those who selected from several predetermined options (e.g., having children without genetic variants) and those who considered PGT in the context of religion (as opposed to health and safety) were less likely to report that they would ever consider using PGT.

Cultural/Community Effects

The recognition that BRCA1/BRCA2 pathogenic variants are prevalent, not only in breast/ovarian cancer families but also in some ethnic groups,[193] has led to considerable discussion of the ethical, psychological, and other implications of having one’s ethnicity be a factor in determination of disease predisposition. Concerns that people will think everything is solely determined by genetic factors and the creation of a genetic underclass [194] have been voiced. Questions about the impact on the group of being singled out as having genetic vulnerability to breast cancer have been raised. There is also confusion about who gives or withholds permission for the group to be involved in studies of their genetic identity. These issues challenge traditional views on informed consent as a function of individual autonomy.[195]
A growing literature on the unique factors influencing a variety of cultural subgroups suggests the importance of developing culturally specific genetic counseling and educational approaches.[99,196-200] The inclusion of members within the community of interest (e.g., breast cancer survivors, advocates, and community leaders) may enhance the development of culturally tailored genetic counseling materials.[100] One study showed that participation in any genetic counseling (culturally mediated or standard approaches) reduced perceived risk of developing breast cancer.[201]

Ethical Concerns

The human implications of the ethical issues raised by the advent of genetic testing for breast/ovarian cancer susceptibility are described in case studies,[202] essays,[203,204] and research reports. Issues about rights and responsibilities in families concerning the spread of information about genetic risk promise to be major ethical and legal dilemmas in the coming decades.
Studies have shown that 62% of studied family members were aware of the family history and that 88% of hereditary breast/ovarian cancer family members surveyed have significant concerns about privacy and confidentiality. Expressed concern about cancer in third-degree relatives, or relatives farther removed, was about the same as that for first- or second-degree relatives of the proband.[205] Only half of surveyed FDRs of women with breast or ovarian cancer felt that written permission should be required to disclose BRCA1/BRCA2 test results to a spouse or immediate family member. Attitudes toward testing varied by ethnicity, previous exposure to genetic information, age, optimism, and information style. Altruism is a factor motivating genetic testing in some people.[206] Many professional groups have made recommendations regarding informed consent.[112,206-209] There is some evidence that not all practitioners are aware of or follow these guidelines.[210] Research shows that many BRCA1/BRCA2 genetic testing consent forms do not fulfill recommendations by professional groups about the 11 areas that should be addressed,[211] and they omit highly relevant points of information.[210] In a study of women with a history of breast or ovarian cancer, the interviews yielded that the women reported feeling inadequately prepared for the ethical dilemmas they encountered when imparting genetic information to family members.[212] These data suggest that more preparation about disclosure to family members before testing reduces the emotional burden of disseminating genetic information to family members. Patients and health care providers would benefit from enhanced consideration of the ethical issues of warning family members about hereditary cancer risk. (Refer to the PDQ summaries Cancer Genetics Risk Assessment and Counseling and Cancer Genetics Overview for more information about the ethics of cancer genetics and genetic testing.)

Psychosocial Aspects of Cancer Risk Management for Hereditary Breast and Ovarian Cancer

Decision aids for persons considering risk management options for hereditary breast and ovarian cancer

There is a small but growing body of literature on the use of decision aids as an adjunct to standard genetic counseling to assist patients in making informed decisions about cancer risk management.[213-217] One study showed that the use of a decision aid consisting of individualized value assessment and cancer risk management information after receiving positive BRCA1/BRCA2 test results was associated with fewer intrusive thoughts and lower levels of depression at the 6-month follow-up in unaffected women. Use of the decision aid did not alter cancer risk management intentions and behaviors. Slightly detrimental effects on well-being and several decision-related outcomes, however, were noted among affected women.[215] Another study compared responses to a tailored decision aid (including a values-clarification exercise) versus a general information pamphlet intended for women making decisions about ovarian cancer risk management. In the short term, the women receiving the tailored decision aid showed a decrease in decisional conflict and increased knowledge compared with women receiving the pamphlet, but no differences in decisional outcomes were found between the two groups. In addition, the decision aid did not appear to alter the participant’s baseline cancer risk management decisions.[214] A multisite randomized trial of 150 unaffected women with BRCA1/BRCA2 pathogenic variants assessed the effect of a decision aid on breast cancer risk management decisions and psychosocial outcomes. At 6-month and 12-month follow-up, women randomly assigned to the decision aid had lower levels of cancer-related distress (P = .01 at 6 months and P = .05 at 12 months) than did the control group.[217] Decisional conflict scores were relatively low at baseline and declined over time in both groups; the scores between the two groups were not statistically different.

Uptake of cancer risk management options

An increasing number of studies have examined uptake and adherence to cancer risk management options among individuals who have undergone genetic counseling and testing for BRCA1 and BRCA2 pathogenic variants. Findings from these studies are reported in Table 15 and Table 16. Outcomes vary across studies and include uptake or adherence to screening (mammography, magnetic resonance imaging [MRI], cancer antigen [CA] 125, transvaginal ultrasound [TVUS]) and selection of RRM and RRSO. Studies generally report outcomes by pathogenic variant carrier or testing status (e.g., positive for pathogenic variants, negative for pathogenic variants, or declined genetic testing). Follow-up time after notification of genetic risk status also varied across studies, ranging from 12 months up to several years.
Findings from these studies suggest that breast screening often improves after notification of BRCA1/BRCA2 pathogenic variant carrier status; nonetheless, screening remains suboptimal. Fewer studies have examined adoption of MRI as a screening modality, probably due to the recent availability of efficacy data. Screening for ovarian cancer varied widely across studies, and also varied based on type of screening test (i.e., CA-125 serum testing vs. TVUS screening). However, ovarian cancer screening does not appear to be widely adopted by carriers of BRCA1/BRCA2 pathogenic variants. Uptake of RRM varied widely across studies, and may be influenced by personal factors (such as younger age or having a family history of breast cancer), psychosocial factors (such as a desire for reduction of cancer-related distress), recommendations of the health care provider, and cultural or health care system factors. An individual’s choice to have a bilateral mastectomy also appears to be influenced by pretreatment genetic education and counseling regardless of the genetic test results.[218] Similarly, uptake of RRSO also varied across studies, and may be influenced by similar factors, including older age, personal history of breast cancer, perceived risk of ovarian cancer, cultural factors (i.e., country), and the recommendations of the health care provider.
Table 15. Uptake of Risk-reducing Mastectomy (RRM) and/or Breast Screening Among Carriers of BRCA1 and BRCA2 Pathogenic Variants
ENLARGE
Study CitationStudy PopulationUptake of RRMUptake of Breast Screening Mammography and/or Breast MRILength of Follow-upComments
MRI = magnetic resonance imaging; RRSO = risk-reducing salpingo-oophorectomy.
aSelf-report as data source.
bMedical records as data source.
United States
Botkin et al. (2003) [219]Carriers (n = 37)aCarriers 0%Mammography24 mo
– Carriers 57%
Noncarriers (n = 92)aNoncarriers 0%– Noncarriers 49%
– Declined test 20%
Declined testing (n = 15)aMRI
– Not evaluated
Beattie et al. (2009) [220]Carriers (n = 237)bCarriers 23%Not applicableMean, 3.7 yWomen opting for RRM were younger than 60 y, had a prior diagnosis of breast cancer, and also underwent RRSO.
Median time to RRM: 124 days from receiving results.
O’Neill et al. (2010) [221]Carriers (n = 146)aCarriers 13%Not applicable12 moIntentions at test result disclosure predicted RRM decisions.
Schwartz et al. (2012) [222]Carriers (n = 108)aCarriers 37%MammographyMean, 5.3 yPredictors of RRM were younger age, higher precounseling cancer distress, more recent diagnosis of breast or ovarian cancer, and intact ovaries.
– Carriers affected 92%
– Carriers unaffected 82%
Noncarriers (n = 60)aNoncarriers 0%– Noncarriers 66%
– Uninformative affected 89%
MRI
Uninformative (n = 206)aUninformative 6.8%– Carriers affected 51%
– Carriers unaffected 46%
– Noncarriers 11%
– Uninformative 27%
Garcia et al. (2013) [223]Carriers (n = 250)bCarriers 44%Excluding women post RRM:41 months; range, 26–66 moBreast surveillance decreased significantly from y 1–5 of follow-up: Mammography 43% to 7%; MRI 35% to 3%.
Mammography:
– Carriers 43%
MRI:
– Carriers 35%
Singh et al. (2013) [224]Carriers (n = 136)bCarriers 42%Not applicableRange, 1–11 yPredictors of RRM were first- or second-degree relative diseased from breast cancer, having had at least one childbirth, and having undergone testing after 2005.
International
Phillips et al. (2006) [225]Carriers (n = 70)aCarriers 11%Mammography3 y
– Carriers 89%
MRI
– Not evaluated
Metcalfe et al. (2008) [226]Carriers (N = 2,677)aCarriers 18% (unaffected)Mammography3.9 y; range, 1.5–10.3 yLarge differences in uptake of risk management options by country.
– Carriers 87%
MRI1,294 participants had a personal history of breast cancer.
– Carriers 31%
Julian-Reynier et al. (2011) [227]Carriers (n = 101)aCarriers 6.9%Mammography5 yNoncarriers often continued screening.
– Carriers 59%
– Noncarriers aged 30–39 y 53%
Noncarriers (n = 145)aNoncarriers 0%MRI
– Carriers 31%
– Noncarriers 4.8%
Table 16. Uptake of Risk-reducing Salpingo-oophorectomy (RRSO) and/or Gynecologic Screening Among Carriers of BRCA1 and BRCA2 Pathogenic Variants
ENLARGE
Study CitationStudy PopulationUptake of RRSOUptake of Gynecologic ScreeningLength of Follow-upComments
CA-125 = cancer antigen 125; RRM = risk-reducing mastectomy; TVUS = transvaginal ultrasound.
aSelf-report as data source.
bMedical records as data source.
cData source not specified.
United States
Scheuer et al. (2002) [228]Carriers (n = 179)aCarriers 50.3%CA-125Mean, 24.8 mo; range, 1.6–66.0 moWomen undergoing RRSO were older and more likely to have a personal history of breast cancer.
– Carriers 67.6%
TVUS
– Carriers 72.9%
Beattie et al. (2009) [220]Carriers (n = 240)bCarriers 51%Not applicableMean, 3.7 yWomen opting for RRSO <60 y had a prior diagnosis of breast cancer and also underwent RRM.
Median time to RRSO: 123 days from receiving results.
O'Neill et al. (2010) [221]Carriers (n = 146)aCarriers 32%Not applicable12 mo
Schwartz et al. (2012) [222]Carriers (n = 100)aCarriers 65%CA-125Mean, 5.3 yPredictors of RRSO were being ≥40 y and having received a diagnosis of breast cancer more than 10 y ago.
Noncarriers (n = 52)aNoncarriers 1.9%– Carriers 56%
– Noncarriers 12%
– Uninformative 33%
Uninformative (n = 203)aUninformative 13.3%TVUS
– Carriers 42%
– Noncarriers 20%
– Uninformative 26%
Garcia et al. (2013) [223]Carriers (n = 305)bCarriers 74%Excluding women post-RRSO:41 mo; range, 26–66 moOvarian surveillance decreased significantly from years 1–5 of follow-up; CA-125: 47% to 2%; TVUS: 45% to 2.3%
CA-125
– Carriers 47%
TVUS
– Carriers 45%
Mannis et al. (2013) [229]Carriers (n = 201)aCarriers 69.6%CA-125Median, 3.7 yPredictors of RRSO and screening included being a carrier of a BRCA pathogenic variant, age 40–49 y, having a higher income, ≥2 children, a personal history of breast cancer, and a first-degree relative with ovarian cancer.
– 26.3%
TVUS
– 26.3%
Noncarriers (n = 103)aNoncarriers 2.0%Not reported
Uninformative (n = 773)a; 59/773 with a variant of uncertain significanceUninformative 12.3%CA-125
– 10.4%
TVUS
– 6.5%
Singh et al. (2013) [224]Carriers (n = 136)bCarriers 52%Not applicableRange, 1–11 yPredictors of RRSO were first- or second-degree relative with breast cancer, a mother lost to pelvic cancer, having had ≥1 childbirths, age ≥50 y, and having undergone testing after 2005.
International
Phillips et al. (2006) [225]Carriers (n = 70)aCarriers 29%CA-1253 y
– Carriers 0%
TVUS
– Carriers 67%
Friebel et al. (2007) [230]Carriers (N = 537)cCarriers 55%Not applicableMinimum 6 mo; median 36 moRRSO greatest in parous women >40 y.
Madalinska et al. (2007) [231]Carriers (n = 160)a, bCarriers 74%Carriers 26%12 moWomen who underwent RRSO had lower education levels, viewed ovarian cancer as incurable, and believed strongly in the benefits of RRSO.
Specific method(s) of gynecological screening not reported.
Metcalfe et al. (2008) [226]Carriers (N = 2,677)aCarriers 57%Not applicable3.9 y; range, 1.5–10.3 yLarge differences in uptake of risk management options by country.
Julian-Reynier et al. (2011) [227]Carriers (n = 101)aCarriers 42.6%TVUS5 yRRSO uptake increased with age. Having undergone RRSO did not alter breast cancer risk perception. Noncarriers often continued screening.
Noncarriers (n = 145)aNoncarriers 2%– Noncarriers 43.2%
Rhiem et al. (2011) [232]Carriers (N = 306)bCarriers 57%Not evaluatedMean, 47.8 mo post-oophorectomyMedian age at time of RRSO = 47 y. One occult fallopian tube cancer was detected at the time of RRSO. One peritoneal carcinoma was diagnosed 26 mo post-RRSO.
Sidon et al. (2012) [233]Carriers (N = 700)a; 386/700 with personal history of breast cancerBRCA1 carriers:Not evaluatedAffected with breast cancerUptake of RRSO was lower in women >60 y (22% uptake at 5 y). None of the women >70 y had a RRSO performed.
– 54.5%
BRCA2 carriers:– BRCA1: Mean, 2.29; range, 0.1–11.45 y
– 45.5%
All carriers with no personal history of breast cancer– BRCA2: Mean, 1.77; range, 0.1–11.1 y
Not affected with breast cancer
– 54.2%
All carriers with personal history of breast cancer– BRCA1: Mean, 1.63; range, 0.1–11.28 y
– 43.2%– BRCA2: Mean, 1.75; range, 0.1–8.98 y
On the other hand, many women found to be pathogenic variant carriers express interest in RRM in hopes of minimizing their risk of breast cancer. In one study of a number of unaffected women with no previous risk-reducing surgery who received results of BRCA1 testing after genetic counseling, 17% of carriers (2 of 12) intended to have mastectomies and 33% (4 of 12) intended to have oophorectomies.[234] In a later study of the same population, RRM was considered an important option by 35% of women who tested positive, whereas risk-reducing oophorectomy was considered an important option by 76%. A prospective study assessed the stability of risk management preferences over five time points (pre-BRCA testing to 9 months after results disclosure) among 80 Dutch women with a documented BRCA pathogenic variant. Forty-six participants indicated a preference for screening at baseline. Of 25 women who preferred RRM at baseline, 22 indicated the same preference 9 months after test results disclosure; however, it was not reported how many women actually had RRM.[235]
Initial interest does not always translate into the decision for surgery. Two different studies found low rates of RRM among carriers of pathogenic variants in the year after result disclosure, one showing 3% (1 of 29) of carriers and the other 9% (3 of 34) of carriers having had this surgery.[167,236] Among members from a large BRCA1 kindred, utilization of cancer screening and/or risk-reducing surgeries was assessed at baseline (before disclosure of results), and at 1 year and 2 years after disclosure of BRCA1 test results. Of the 269 men and women who participated, complete data were obtained on 37 female carriers and 92 female noncarriers, all aged 25 years or older. At 2 years after disclosure of test results, none of the women had undergone RRM, although 4 of the 37 carriers (10.8%) said they were considering the procedure. In contrast, of the 26 women who had not had an oophorectomy before baseline, 46% (12 of 26) had obtained an oophorectomy by 2 years after testing. Of those carriers aged 25 to 39 years, 29% (5 of 17) underwent oophorectomy, while 78% (7 of 9) of the carriers aged 40 years and older had this procedure.[219] In a study assessing uptake of risk-reducing surgery 3 months after BRCA result disclosure, 7 of 62 women had undergone RRM and 13 of 62 women had undergone RRSO. Intent to undergo RRSO before testing correlated with procedure uptake. In contrast, intent to undergo RRM did not correlate with uptake. Overall, reasons given for indecision about risk-reducing surgery included complex testing factors such as the significance of family history in the absence of a pathogenic variant, concerns over the surgical procedure, and time and uncertainty regarding early menopause and the use of HRT.[237] In a U.K. study, data were collected during observations of genetic consultations and in semistructured interviews with 41 women after they received genetic counseling.[238] The option of risk-reducing surgery was raised in 29 consultations and discussed in 35 of the postclinic interviews. Fifteen women said they would consider having an oophorectomy in the future, and nine said they would consider having a mastectomy. The implications of undergoing oophorectomy and mastectomy were discussed in postclinic interviews. Risk-reducing surgery was described by the counselees as providing individuals with a means to (a) fulfill their obligations to other family members and (b) reduce risk and contain their fear of cancer. The costs of this form of risk management were described by the respondents as follows:
  • Compromising social obligations.
  • Upsetting the natural balance of the body.
  • Not receiving protection from cancer.
  • Operative and postoperative complications.
  • The onset of menopause.
  • The effects of body image, gender, and personal identity.
  • Potential effects on sexual relationships.[238]
A number of women choose to undergo RRM and RRSO without genetic testing because of the following:
  • Testing is not readily accessible.
  • They do not wish exposure to the psychosocial risks of genetic testing.
  • They do not trust that a negative genetic test result means they are not at increased risk.
  • They find any level of risk, even baseline population risk, unacceptable.[239,240]
Among FDRs of breast cancer patients attending a surveillance clinic, women who expressed an interest in RRM and/or had undergone surgery were found to have significantly more breast cancer biopsies (P < .05) and higher subjective 10-year breast cancer risk estimates (P < .05) than women not interested in RRM. Cancer worry at the time of entry into the clinic was highest among women who subsequently underwent RRM compared with women who expressed interest but had not yet had surgery and women who did not intend to have surgery (P < .001).[241]
BRCA testing, when offered to women newly diagnosed with breast cancer, has been shown to influence surgical decision making in that carriers are more likely to opt for bilateral mastectomy compared with noncarriers.[218,242,243] A study that evaluated predictors of contralateral RRM among 435 breast cancer survivors found that 16% had undergone contralateral RRM (in conjunction with mastectomy of the affected breast) before referral for genetic counseling and BRCA1/BRCA2 genetic testing.[244] Predictors of contralateral RRM before genetic counseling and testing included younger age at breast cancer diagnosis, more time since diagnosis, having at least one affected FDR, and not being employed full-time. In the year after disclosure of test results, 18% of women who tested positive for a BRCA1/BRCA2 pathogenic variant and 2% of those whose test results were uninformative underwent contralateral RRM. Predictors of contralateral RRM after genetic testing included younger age at breast cancer diagnosis, higher cancer-specific distress before genetic counseling, and having a positive BRCA1/BRCA2 test result. In this study, contralateral RRM was not associated with distress at 1 year after disclosure of genetic test results. A retrospective chart review evaluated uptake of bilateral mastectomies in 110 women who underwent BRCA1/BRCA2 genetic testing before making surgical decisions about the treatment of newly diagnosed breast cancer. Carriers of BRCA pathogenic variants were more likely to undergo bilateral mastectomies than were women in whom no variant was detected (83% vs. 37%; P = .046).[245] The only predictor of contralateral RRM in women without a pathogenic variant was being married (P = .03). Age, race, parity, disease stage and biomarkers, increased mammographic breast density, and breast MRI did not influence contralateral RRM decisions at the time of primary surgical treatment.
A study conducted from 2006 to 2014 in 11 U.S. academic and community centers of 897 women, aged 40 years and younger at breast cancer diagnosis, found that rates of BRCA genetic testing have increased over time.[243] Within 1 year after diagnosis, 87% of the sample underwent BRCA testing. The rate increased from 77% of newly diagnosed women tested in 2006 to 95% of women tested in 2013. Among women who tested positive for a pathogenic BRCA variant and stated that testing affected their surgery decisions (n = 88), 86% underwent bilateral mastectomy compared with 51% of noncarriers (P < .001). Among untested women, about one-third reported that they were not told by a health care provider that they were candidates for BRCA testing; yet, according to national guidelines, all were eligible for testing solely on the basis of their age at diagnosis.
Dutch women (N = 114) who had undergone unilateral or bilateral RRM with breast reconstruction between 1994 and 2002 were retrospectively surveyed to determine their satisfaction with the procedure.[246] Sixty-eight percent were either unaffected carriers of BRCA pathogenic variants or at a 50% risk of having a BRCA pathogenic variant in their family. Sixty percent of respondents indicated that they were satisfied with the procedure, 95% would opt for RRM again, and 80% would opt for the same reconstruction procedure. Less than half reported some perioperative or postoperative complications, ongoing physical complaints, or some physical limitations. Twenty-nine percent reported altered feelings of femininity after the procedure, 44% reported adverse changes in their sexual relationships, and 35% indicated that they believed their partners experienced adverse changes in their sexual relationship. Ten percent of women, however, reported positive changes in their sexual relationship after the procedure. Compared with patients who indicated satisfaction with this procedure, nonsatisfied patients were more likely to feel less informed about the procedure and its consequences, report more complications and physical complaints, feel that their breasts did not belong to their body, and indicate that they would not opt for reconstruction again. Those who reported a negative effect on their sexual relationship were more likely to:
  • Feel less informed.
  • Experience more physical complaints and limitations.
  • Express that their breasts did not feel like their own.
  • Be disinclined to opt for reconstruction again.
  • State that the surgery had not met their expectations.
  • Experience altered feelings of femininity and perceived adverse changes in their partner’s view of their femininity and their sexual relationship.
Ninety Swedish women who had undergone RRM between 1997 and 2005 were surveyed before surgery, 6 months after surgery, and 1 year after surgery to evaluate changes in health-related quality of life, depression, anxiety, sexuality, and body image. There were no significant changes in health-related quality of life or depression at the three time points; anxiety decreased over time (P = .0004). More than 80% of women reported having an intimate relationship at all three time points. Women who reported being sexually active were asked to respond to questions about sexual pleasure, discomfort, habit, and frequency of activity. There were no statistically significant differences related to frequency, habit, or discomfort. However, pleasure significantly decreased between baseline and 1 year after surgery (P = .005). At 1 year after surgery, 48% of women reported feeling less attractive, 48% reported feeling self-conscious, and 44% reported dissatisfaction with surgical scars.[247]
Discussion of risk-reducing surgical options may not consistently occur during pretest genetic counseling. In one multi-institutional study, only one-half of genetics specialists discussed RRM and RRSO in consultations with women from high-risk breast cancer families,[248,249] despite the fact that discussion of surgical options was significantly associated with meeting counselees’ expectations, and that such information was not associated with increased anxiety.[250]
Given the increased risk of ovarian cancer faced by women with a BRCA1 or BRCA2 pathogenic variant, those who do receive information about RRSO show wide variations in surgery uptake (27%–72%).[9,123,228,231,251,252] A study showed that clinical factors related to choosing RRSO versus surveillance alone are older age, parity of one or more, and a prior breast cancer diagnosis.[253] In this study, the choice of RRSO was not related to family history of breast or ovarian cancer. Hysterectomy was presented as an option during genetic counseling, and 80% of women who underwent RRSO also elected to have a hysterectomy.

Cancer screening and risk-reducing behaviors

Data are now emerging regarding uptake and adherence to cancer risk management recommendations such as screening and risk-reducing interventions. Cancer screening adherence and risk-reduction behaviors as defined by the National Comprehensive Cancer Network Guidelines were assessed in a cross-sectional study of 214 women with a personal history (n = 134) or family history (n = 80) of breast or ovarian cancer. Among unaffected women older than 40 years, 10% had not had a mammogram or clinical breast examination (CBE) in the previous year and 46% did not practice breast self-examination (BSE). Among women previously affected with breast or ovarian cancer, 21% had not had a mammogram, 32% had not had a CBE, and 39% did not practice BSE.[254]
Three hundred and twelve women who were counseled and tested for BRCA pathogenic variants between 1997 and 2005 responded to a survey regarding their perception of genetic testing for hereditary breast and ovarian cancer. The survey included questions on risk reduction options, including screening and risk-reducing surgeries. Two hundred and seventeen (70%) of the women had been diagnosed with breast cancer, and 86 (28%) tested positive for a pathogenic variant in either the BRCA1 or BRCA2 gene. None of the BRCA-positive women agreed that mammograms are difficult procedures because of the discomfort, while 11 (5.4%) of the BRCA-negative women did agree with this statement. Both groups (BRCA-positive and BRCA-negative) agreed that risk-reducing surgeries provide the best means for lowering cancer risk and worry, and most patients in both groups expressed the belief that risk-reducing mastectomy is not too drastic, too scary, or too disfiguring.[255]
A prospective study from the United Kingdom examined the psychological impact of mammographic screening in 1,286 women aged 35 to 49 years who have a family history of breast cancer and were participants in a multicenter screening program. Mammographic abnormalities that required additional evaluation were detected in 112 women. These women, however, did not show a statistically significant increase in cancer worry or negative psychological consequences as a result of these findings. The 1,174 women who had no mammographic abnormality detected experienced a decrease in cancer worry and a decrease in negative psychological consequences compared with baseline after receipt of their results. At 6 months, the entire cohort had experienced a decrease in measures of cancer worry and psychological consequences of breast screening.[256]
A qualitative study explored health care professionals’ views regarding the provision of information about health protective behaviors (e.g., exercise and diet). Seven medical specialists and ten genetic counselors were interviewed during a focus group or individually. The study reported wide variation in the content and extent of information provided about health-protective behaviors and in general, participants did not consider it their role to promote such behaviors in the context of a genetic counseling session. There was agreement, however, about the need to form consensus about provision of such information both within and across risk assessment clinics.[257]
Not all studies specify whether screening uptake rates fall within recommended guidelines for the targeted population or the specific clinical scenario, nor do they report on other variables that may influence cancer screening recommendations. For example, women who have a history of atypical ductal hyperplasia would be advised to follow screening recommendations that may differ from those of the general population.

Psychosocial Outcome Studies

Risk-reducing mastectomy

A prospective study conducted in the Netherlands found that among 26 carriers of BRCA1/BRCA2 pathogenic variants, the 14 women who chose mastectomy had higher distress both before test result disclosure and 6 and 12 months later, compared with the 12 carriers who chose surveillance and compared with 53 women negative for a pathogenic variant. Overall, however, anxiety declined in women undergoing risk-reducing mastectomy (RRM); at 1 year, their anxiety scores were closer to those of women choosing surveillance and to the scores of women negative for a pathogenic variant.[258] Interestingly, women opting for RRM had lower pretest satisfaction with their breasts and general body image than carriers who opted for surveillance or noncarriers of BRCA1/BRCA2 pathogenic variants. Of the women who had an RRM, all but one did not regret the decision at 1 year posttest disclosure, but many had difficulties with body image, sexual interest and functioning, and self-esteem. The perception that doctors had inadequately informed them about the consequences of RRM was associated with regret.[258] At the 5-year follow-up, women who had undergone RRM had less favorable body image and changes in sexual relationships, but also had a significant reduction in the fear of developing cancer.[259] In a study of 78 women who underwent risk-reducing surgery (including BRCA1/BRCA2 carriers and women who were from high-risk families with no detectable BRCA1/BRCA2 pathogenic variant), cancer-specific and general distress were assessed 2 weeks before surgery and at 6 and 12 months postsurgery.[260] The sample included women who had RRM and RRSO alone and women who had both surgeries. There was no observable increase in distress over the 1-year period.
Mixed psychosocial outcomes were reported in a follow-up study (mean 14 years) of 609 women who received RRM at the Mayo Clinic. Seventy percent were satisfied with RRM, 11% were neutral, and 19% were dissatisfied. Eighteen percent believed that if they had the choice to make again, they probably or definitely would not have an RRM. About three-quarters said their worry about cancer was diminished by surgery. One-half reported no change in their satisfaction with body image; 16% reported improved body image after surgery. Thirty-six percent said they were dissatisfied with their body image after RRM. About one-quarter of the women reported adverse impact of RRM on their sexual relationships and sense of femininity, and 18% had diminished self-esteem. Factors most strongly associated with satisfaction with RRM were postsurgical satisfaction with appearance, reduced stress, no reconstruction or lack of problems with implants, and no change or improvement in sexual relationships. Women who cited physician advice as the primary reason for choosing RRM tended to be dissatisfied after RRM.[261]
A study of 60 healthy women who underwent RRM measured levels of satisfaction, body image, sexual functioning, intrusion and avoidance, and current psychological status at a mean of 4 years and 4 months postsurgery. Of this group, 76.7% had either a strong family history (21.7%) or carried a BRCA1 or BRCA2 pathogenic variant (55%). Overall, 97% of the women surveyed were either satisfied (17%) or extremely satisfied (80%) with their decision to have RRM, and all but one participant would recommend this procedure to other women. Most women (66.7%) reported that surgery had no impact on their sexual life, although 31.7% reported a worsening sexual life, and 76.6% reported either no change in body image or an improvement in body image, regardless of whether reconstruction was performed. Worsening self-image was reported by 23.3% of women after surgery. Women’s mean distress levels after surgery were only slightly above normal levels, although those women who continued to perceive their postsurgery breast cancer risk as high had higher mean levels of global and cancer-related distress than those who perceived their risk as low. Additionally, carriers of BRCA1 and BRCA2 pathogenic variants and women with a strong family history of breast and/or ovarian cancer had higher mean levels of cancer-related distress than women with a limited family history.[262]
Very little is known about how the results of genetic testing affect treatment decisions at the time of cancer diagnosis. Two studies explored genetic counseling and BRCA1/BRCA2 genetic testing at the time of breast cancer diagnosis.[24,218] One of these studies found that genetic testing at the time of diagnosis significantly altered surgical decision making, with more pathogenic variant carriers than noncarriers opting for bilateral mastectomy. Bilateral RRM was chosen by 48% of women with a known pathogenic variant [218] and by 100% of women with a known pathogenic variant in a smaller series [24] of women undergoing testing at the time of diagnosis. Of women in whom no pathogenic variant was found, 24% also opted for bilateral RRM. Four percent of the test decliners also underwent bilateral RRM. Among carriers of pathogenic variants, predictors of bilateral RRM included whether patients reported that their physicians had recommended BRCA1/BRCA2 testing and bilateral RRM before testing, and whether they received a positive test result.[218] Data are lacking on quality-of-life outcomes for women who undergo RRM after genetic testing that is performed at the time of diagnosis.
A prospective study from the Netherlands evaluated long-term psychological outcomes of offering women with breast cancer genetic counseling and, if indicated, genetic testing at the onset of breast radiation for treatment of their primary breast cancer. Of those who were approached for counseling, some underwent genetic testing and chose to receive their result (n = 58), some were approached but did not fulfill referral criteria (n = 118), and some declined the option of counseling/testing (n = 44). Another subset of women undergoing radiation therapy was not approached for counseling (n = 182) but was followed using the same measures. Psychological distress was measured at baseline and at 4, 11, 27, and 43 weeks after initial consultation for radiation therapy. No differences were detected in general anxiety, depression or breast cancer–specific distress across all four groups.[263]
A retrospective questionnaire study of 583 women with a personal and family history of breast cancer and who underwent contralateral RRM between 1960 and 1993 measured overall satisfaction after mastectomy and factors influencing satisfaction and dissatisfaction with this procedure.[264] The mean time of follow-up was 10.3 years after risk-reducing surgery. Overall, 83% of all participants stated they were satisfied or very satisfied, 8% were neutral, and 9% were dissatisfied with contralateral RRM. Most women also reported favorable effects or no change in their self-esteem, level of stress, and emotional stability after surgery (88%, 83%, and 88%, respectively). Despite the high levels of overall satisfaction, 33% reported negative body image, 26% reported a reduced sense of femininity, and 23% reported a negative effect on sexual relationships. The type of surgical procedure also affected levels of satisfaction. The authors attributed this difference to the high rate of unanticipated reoperations in the group of women having subcutaneous mastectomy (43%) versus the group having simple mastectomy (15%) (P < .0001). Limitations to this study are mostly related to the time period during which participants had their surgery (i.e., availability of surgical reconstructive option).[264,265] None of these women had genetic testing for pathogenic variants in the BRCA1/BRCA2 genes. Nevertheless, this study shows that while most women in this group were satisfied with contralateral RRM, all women reported at least one adverse outcome.
A retrospective survey of 137 BRCA carriers examined the psychosocial impact of preserving the nipple-areolar complex (NAC) in women with bilateral RRM.[266] The study found that body image and sexual well-being differed significantly based on the type of RRM the women underwent. Women with NAC preservation were more satisfied with their breasts (72% vs. 61%), were more satisfied with the surgical outcome (85% vs. 74%), and had greater sexual well-being (68% vs. 52%) than women without NAC preservation. No differences in cancer-related distress, anxiety, depression, or risk perceptions were observed between the two groups. Oncologic outcomes of nipple-sparing mastectomy in BRCA carriers have not been inferior to RRM without NAC preservation.[267] (Refer to the RRM section of this summary for more information.)
Another study compared long-term quality-of-life outcomes in 195 women after bilateral RRM performed between 1979 and 1999 versus 117 women at high risk of breast cancer opting for screening. No statistically significant differences were detected between the groups for psychosocial outcomes. Eighty-four percent of those opting for surgery reported satisfaction with their decision. Sixty-one percent of women from both the surgery and screening groups reported being very much or quite a bit contented with their quality of life.[268]
In a study of psychosocial outcomes associated with RRM and immediate reconstruction, 61 high-risk women (27 carriers of pathogenic variants, others with high-risk family history), 31 of whom had a prior history of breast cancer, were evaluated on average 3 to 4 years after surgery.[269] The study utilized questions designed to elicit yes versus no responses and found that the surgery was well-tolerated with 83% of participants reporting that the results of their reconstructive surgery were as they expected, 90% reporting that they had received adequate preoperative information, none reporting that they regretted the surgery, and all reporting that they would choose the same route if they had to do it again. Satisfaction with the results ranged from 74% satisfied with the shape of their breasts to 89% satisfied with the appearance of the scarring. Comparison of this group to normative samples on quality-of-life indicators (Short Form 36 Health Survey Questionnaire [SF-36]; Hospital Anxiety and Depression Scale questionnaire scores) indicated no reductions in quality of life in these women.
A qualitative study examining material on the FORCEExit Disclaimer website posted by 21 high-risk women (BRCA1/BRCA2 positive) undergoing RRM showed that these women anticipated and received negative reactions from friends and family regarding the surgery, and that they managed disclosure in ways to maintain emotional support and self-protection for their decision. Many of the women expressed a relief from intrusive breast cancer thoughts and worry, and were satisfied with the cosmetic result of their surgery.[270]
In contrast, another study examined long-term psychosocial outcomes in 684 women who had had bilateral or contralateral RRM on average 9 years before assessment.[271] A majority of women (59%) also had reconstructive surgery. Interestingly, based on a Likert scale, 85% of women reported that they were satisfied or very satisfied with their decision to have an RRM. However, in qualitative interviews, a large number of women went on to describe dissatisfaction or negative psychosocial outcomes associated with surgery. The authors coded the responses as negative when women reported still being anxious about their breast cancer risk and/or reported negative feelings about their body image, pain, and sexuality. Seventy-nine percent of the women providing negative comments and 84% of those making mixed comments (mixture of satisfaction and dissatisfaction) responded that they were either satisfied or very satisfied with their decision. Twice as many women with bilateral mastectomy made negative and mixed comments than did women with contralateral mastectomy. The areas of most concern were body image, problems with breast implants, pain after surgery, and sexuality. The authors proposed that those who had undergone contralateral procedures had already been treated for cancer, while those who had undergone bilateral procedures had not been treated previously, and this may partially account for the differences in satisfaction between the two groups. These findings suggest that women's satisfaction with RRM may be tempered by their complex reactions over time.
In a qualitative study of 108 women who underwent or were considering RRM, more than half of those who had RRM felt that presurgical consultation with a psychologist was advisable; nearly two-thirds thought that postsurgical consultation was also appropriate. All of the women who were considering RRM believed that psychological consultation before surgery would facilitate decision-making.[272]

Risk-reducing salpingo-oophorectomy

A retrospective self-administered survey of 40 women aged 35 to 74 years at time of RRSO (57.5% were younger than 50 y), who had undergone the procedure through the Ontario Ministry of Health due to a family history of ovarian cancer, found that RRSO resulted in a significant reduction in perceived ovarian cancer risk. Fifty-seven percent identified a decrease in perceived risk as a benefit of RRSO (35% did not comment on RRSO benefits) and 49% reported that they would repeat RRSO to decrease cancer risk. The overall quality-of-life scores were consistent with those published for women who are menopausal or participating in hormone studies.[273] Quality of life in 59 women who underwent RRSO was assessed at 24 months postprocedure.[274] Overall quality of life was similar to the general population and breast cancer survivors, with approximately 20% reporting depression. The 30% of subjects reporting vaginal dryness and dyspareunia were more likely to report dissatisfaction with the procedure.
A Canadian prospective study examined the impact of RRSO on menopausal symptoms and sexual functioning before surgery and then 1 year later in a sample of 114 women with known BRCA1/BRCA2 pathogenic variants.[275] Satisfaction with the decision to undergo RRSO was high regardless of symptoms reported. Those who were premenopausal at the time of surgery (n = 75) experienced a worsening of symptoms and a decline in sexual functioning. HRT addressed vaginal dryness and dyspareunia but not declines in sexual pleasure. HRT also resulted in fewer moderate to severe hot flashes.
Additional work reported by this group found that the majority of the 127 women who had undergone RRSO 1 year previously (75 with BRCA1 pathogenic variants; 52 with BRCA2 pathogenic variants) felt that RRSO reduced their risk of both breast and ovarian cancer.[276] There was a wide range of risk perceptions for ovarian cancer noted in the group. Twenty percent of carriers of BRCA1 and BRCA2 pathogenic variants thought that their risk of ovarian cancer was completely eliminated; others had an inflated perception of their ovarian cancer risk both before and after surgery. A small group of these women were further surveyed at about 3 years postsurgery, and their risk perceptions did not change significantly during this extended time period. These findings suggest that important misperceptions about ovarian cancer risk may persist after RRSO. Additional genetic education and counseling may be warranted.
A larger study assessed quality of life in women at high risk of ovarian cancer who opted for periodic gynecologic screening (GS) versus those who underwent RRSO. Eight hundred forty-six high-risk women, 44% of whom underwent RRSO and 56% of whom chose GS, completed questionnaires evaluating quality of life, cancer-specific distress, endocrine symptoms, and sexual functioning.[277] Women in the RRSO group were a mean of 2.8 ±1.9 years from surgery and women in the GS group were a mean of 4.3 years from their first visit to a gynecologist for high-risk management. No statistical difference in overall quality of life was detected between the RRSO and GS groups. When compared with the GS group, women who underwent RRSO had poorer sexual functioning and more endocrine symptoms such as vaginal dryness, dyspareunia, and hot flashes. Women who underwent RRSO experienced lower levels of breast and ovarian cancer distress and had a more favorable perception of cancer risk.
Women (N = 182) at risk of hereditary breast and ovarian cancer referred for genetic counseling were surveyed concerning their satisfaction with their choice of either RRSO or periodic screening (PS) (biannual pelvic examination with TVUS and CA-125 determination) to manage their ovarian cancer risk.[278] Overall satisfaction with both options was extremely high, but highest among those who chose RRSO over PS. There were no other demographic or clinical factors that distinguished satisfaction level. There was higher decisional ambivalence among those who chose PS.
A retrospective study assessed 98 carriers of BRCA pathogenic variants who underwent RRSO about their preoperative counseling regarding symptoms to expect after surgery.[279] The mean age at RRSO was 45.5 years (range, 32–63 y). Eighty-five percent pursued RRSO after learning that they harbored a BRCA pathogenic variant, and 48.0% were premenopausal at the time of surgery. Participants reported ‘‘frequent’’ or ‘‘very frequent’’ postsurgical symptoms of vaginal dryness (52.1%), changes in interest in sex (50.0%), sleep disturbances (46.7%), changes in sex life (43.9%), and hot flashes (42.9%). Only vaginal dryness and hot flashes were commonly recalled to have been addressed preoperatively. While 96% would have the surgery again, participants reported that the discussion of the impact of surgery on their sex life (59.2%), risk of coronary heart disease (57.1%), and the availability of sex counseling (57.1%) would have been helpful.

Behavioral Outcomes

A study [280] of screening behaviors of 216 self-referred, high-risk (>10% risk of carrying a BRCA1/BRCA2 pathogenic variant) women who are members of hereditary breast cancer families found a range of screening practices. Even the presence of known pathogenic variants in their families was not associated with good adherence to recommended screening practices. Sixty-nine percent of women aged 50 to 64 years and 83% of women aged 40 to 49 years had had a screening mammogram in the previous year. Twenty percent of participants had ever had a CA-125 test and 31% had ever had a pelvic ultrasound or TVUS. Further analysis of this study population [280] looking specifically at 107 women with informative BRCA test results found good use of breast cancer screening, though the uptake rate in younger carriers is lower. The reason for the lower uptake rate was not explored in this study.[281] One survey of screening behaviors among women at increased risk of breast/ovarian cancer identified physician recommendations as a significant factor in adherence to screening.[282]
While motivations cited for pursuing genetic testing often include the expectation that carriers of pathogenic variants will be more compliant with breast and/or ovarian screening recommendations,[280,283-285] limited data exist about whether participants in genetic testing alter their screening behaviors over time and about other variables that may influence those behaviors, such as insurance coverage and physician recommendations or attitudes. The impact of cancer genetic counseling on screening behaviors was assessed in a U.K. study of 293 women followed for 12 months postcounseling at four cancer genetics clinics.[286] BSE, CBE, and mammography were significantly increased after counseling; however, gaps in adherence to recommendations were noted: 38% of women aged 35 to 49 years had not had a mammogram by 12 months postcounseling. BSE was not done by most women at the recommended time and frequency.
This is a critical issue not only for women testing positive, but also for adherence to screening for those testing negative and those who have received indeterminate results or choose not to receive their results. It is possible that adherence actually diminishes with a decrease in the perceived risk that may result from a negative genetic test result.
In addition, while there is still some question regarding the link between cancer-related worry and breast cancer screening behavior, accumulating evidence appears to support a linear rather than a curvilinear relationship. That is, for some time, the data were not consistent; some data supported the hypothesis that mild-to-moderate worry may increase adherence, while excessive worry may actually decrease the utilization of recommended screening practices. Other reports support the notion that a linear relationship is more likely; that is, more worry increases adherence to screening recommendations. Few studies, however, have followed women to assess their health behaviors after genetic testing. Thus, a negative test result leading to decreased worry could theoretically result in decreased screening adherence. A large study found that patient compliance with screening practices was not related to general or screening-specific anxiety—with the exception of BSE, for which compliance is negatively associated with procedure-specific anxiety.[76] Further research designed to clarify this potential concern would highlight the need for comprehensive genetic counseling to discuss the need for follow-up screening.
Further complicating this area of research are issues such as the baseline rate of mammography adherence among women older than 40 or 50 years before genetic testing. More specifically, the ability to note a significant difference in adherence on this measure may be affected by the high adherence rate to this screening behavior before genetic testing by women undergoing such testing. It may be easier to find significant changes in mammography use among women with a family history of breast cancer who test positive. Finally, adherence over time will likely be affected by how women undergoing genetic testing and their caregivers perceive the efficacy of many of the screening options in question, such as mammography for younger women, BSE, and ovarian cancer screening (periodic vaginal ultrasound and serum CA-125 measurements), along with the value of preventive interventions.
The issue of screening decision-making and adherence among women undergoing genetic testing for breast and ovarian cancer is the subject of several ongoing trials, and an area of much needed ongoing study.


References
  1. Ropka ME, Wenzel J, Phillips EK, et al.: Uptake rates for breast cancer genetic testing: a systematic review. Cancer Epidemiol Biomarkers Prev 15 (5): 840-55, 2006. [PUBMED Abstract]
  2. Schwartz MD, Lerman C, Brogan B, et al.: Utilization of BRCA1/BRCA2 mutation testing in newly diagnosed breast cancer patients. Cancer Epidemiol Biomarkers Prev 14 (4): 1003-7, 2005. [PUBMED Abstract]
  3. Kieran S, Loescher LJ, Lim KH: The role of financial factors in acceptance of clinical BRCA genetic testing. Genet Test 11 (1): 101-10, 2007. [PUBMED Abstract]
  4. Susswein LR, Skrzynia C, Lange LA, et al.: Increased uptake of BRCA1/2 genetic testing among African American women with a recent diagnosis of breast cancer. J Clin Oncol 26 (1): 32-6, 2008. [PUBMED Abstract]
  5. Olaya W, Esquivel P, Wong JH, et al.: Disparities in BRCA testing: when insurance coverage is not a barrier. Am J Surg 198 (4): 562-5, 2009. [PUBMED Abstract]
  6. Levy DE, Byfield SD, Comstock CB, et al.: Underutilization of BRCA1/2 testing to guide breast cancer treatment: black and Hispanic women particularly at risk. Genet Med 13 (4): 349-55, 2011. [PUBMED Abstract]
  7. Landsbergen K, Verhaak C, Kraaimaat F, et al.: Genetic uptake in BRCA-mutation families is related to emotional and behavioral communication characteristics of index patients. Fam Cancer 4 (2): 115-9, 2005. [PUBMED Abstract]
  8. Denayer L, Boogaerts A, Philippe K, et al.: BRCA1/2 predictive testing and gender: uptake, motivation and psychological characteristics. Genet Couns 20 (4): 293-305, 2009. [PUBMED Abstract]
  9. Meijers-Heijboer EJ, Verhoog LC, Brekelmans CT, et al.: Presymptomatic DNA testing and prophylactic surgery in families with a BRCA1 or BRCA2 mutation. Lancet 355 (9220): 2015-20, 2000. [PUBMED Abstract]
  10. Wakefield CE, Ratnayake P, Meiser B, et al.: "For all my family's sake, I should go and find out": an Australian report on genetic counseling and testing uptake in individuals at high risk of breast and/or ovarian cancer. Genet Test Mol Biomarkers 15 (6): 379-85, 2011. [PUBMED Abstract]
  11. Schlich-Bakker KJ, ten Kroode HF, Wárlám-Rodenhuis CC, et al.: Barriers to participating in genetic counseling and BRCA testing during primary treatment for breast cancer. Genet Med 9 (11): 766-77, 2007. [PUBMED Abstract]
  12. Vadaparampil ST, McIntyre J, Quinn GP: Awareness, perceptions, and provider recommendation related to genetic testing for hereditary breast cancer risk among at-risk Hispanic women: similarities and variations by sub-ethnicity. J Genet Couns 19 (6): 618-29, 2010. [PUBMED Abstract]
  13. Meiser B: Psychological impact of genetic testing for cancer susceptibility: an update of the literature. Psychooncology 14 (12): 1060-74, 2005. [PUBMED Abstract]
  14. Pasacreta JV: Psychosocial issues associated with genetic testing for breast and ovarian cancer risk: an integrative review. Cancer Invest 21 (4): 588-623, 2003. [PUBMED Abstract]
  15. Simon MS, Petrucelli N: Hereditary breast and ovarian cancer syndrome : the impact of race on uptake of genetic counseling and testing. Methods Mol Biol 471: 487-500, 2009. [PUBMED Abstract]
  16. Hann KEJ, Freeman M, Fraser L, et al.: Awareness, knowledge, perceptions, and attitudes towards genetic testing for cancer risk among ethnic minority groups: a systematic review. BMC Public Health 17 (1): 503, 2017. [PUBMED Abstract]
  17. Meiser B, Gaff C, Julian-Reynier C, et al.: International perspectives on genetic counseling and testing for breast cancer risk. Breast Dis 27: 109-25, 2006-2007. [PUBMED Abstract]
  18. Armstrong K, Stopfer J, Calzone K, et al.: What does my doctor think? Preferences for knowing the doctor's opinion among women considering clinical testing for BRCA1/2 mutations. Genet Test 6 (2): 115-8, 2002 Summer. [PUBMED Abstract]
  19. McCuaig JM, Greenwood CM, Shuman C, et al.: Breast and ovarian cancer: the forgotten paternal contribution. J Genet Couns 20 (5): 442-9, 2011. [PUBMED Abstract]
  20. Burke W, Culver J, Pinsky L, et al.: Genetic assessment of breast cancer risk in primary care practice. Am J Med Genet A 149A (3): 349-56, 2009. [PUBMED Abstract]
  21. Mouchawar J, Klein CE, Mullineaux L: Colorado family physicians' knowledge of hereditary breast cancer and related practice. J Cancer Educ 16 (1): 33-7, 2001. [PUBMED Abstract]
  22. Yong MC, Zhou XJ, Lee SC: The importance of paternal family history in hereditary breast cancer is underappreciated by health care professionals. Oncology 64 (3): 220-6, 2003. [PUBMED Abstract]
  23. Bellcross CA, Leadbetter S, Alford SH, et al.: Prevalence and healthcare actions of women in a large health system with a family history meeting the 2005 USPSTF recommendation for BRCA genetic counseling referral. Cancer Epidemiol Biomarkers Prev 22 (4): 728-35, 2013. [PUBMED Abstract]
  24. Weitzel JN, McCaffrey SM, Nedelcu R, et al.: Effect of genetic cancer risk assessment on surgical decisions at breast cancer diagnosis. Arch Surg 138 (12): 1323-8; discussion 1329, 2003. [PUBMED Abstract]
  25. Kurian AW, Griffith KA, Hamilton AS, et al.: Genetic Testing and Counseling Among Patients With Newly Diagnosed Breast Cancer . JAMA 317 (5): 531-534, 2017. [PUBMED Abstract]
  26. Childers CP, Childers KK, Maggard-Gibbons M, et al.: National Estimates of Genetic Testing in Women With a History of Breast or Ovarian Cancer. J Clin Oncol 35 (34): 3800-3806, 2017. [PUBMED Abstract]
  27. Pal T, Bonner D, Kim J, et al.: Early onset breast cancer in a registry-based sample of African-american women: BRCA mutation prevalence, and other personal and system-level clinical characteristics. Breast J 19 (2): 189-92, 2013 Mar-Apr. [PUBMED Abstract]
  28. Weldon CB, Trosman JR, Gradishar WJ, et al.: Barriers to the use of personalized medicine in breast cancer. J Oncol Pract 8 (4): e24-31, 2012. [PUBMED Abstract]
  29. Hafertepen L, Pastorino A, Morman N, et al.: Barriers to genetic testing in newly diagnosed breast cancer patients: Do surgeons limit testing? Am J Surg 214 (1): 105-110, 2017. [PUBMED Abstract]
  30. Katz SJ, Ward KC, Hamilton AS, et al.: Gaps in Receipt of Clinically Indicated Genetic Counseling After Diagnosis of Breast Cancer. J Clin Oncol 36 (12): 1218-1224, 2018. [PUBMED Abstract]
  31. Lerman C, Hughes C, Benkendorf JL, et al.: Racial differences in testing motivation and psychological distress following pretest education for BRCA1 gene testing. Cancer Epidemiol Biomarkers Prev 8 (4 Pt 2): 361-7, 1999. [PUBMED Abstract]
  32. Armstrong K, Micco E, Carney A, et al.: Racial differences in the use of BRCA1/2 testing among women with a family history of breast or ovarian cancer. JAMA 293 (14): 1729-36, 2005. [PUBMED Abstract]
  33. Kinney AY, Simonsen SE, Baty BJ, et al.: Acceptance of genetic testing for hereditary breast ovarian cancer among study enrollees from an African American kindred. Am J Med Genet A 140 (8): 813-26, 2006. [PUBMED Abstract]
  34. Halbert CH, Kessler L, Stopfer JE, et al.: Low rates of acceptance of BRCA1 and BRCA2 test results among African American women at increased risk for hereditary breast-ovarian cancer. Genet Med 8 (9): 576-82, 2006. [PUBMED Abstract]
  35. Thompson HS, Valdimarsdottir HB, Duteau-Buck C, et al.: Psychosocial predictors of BRCA counseling and testing decisions among urban African-American women. Cancer Epidemiol Biomarkers Prev 11 (12): 1579-85, 2002. [PUBMED Abstract]
  36. Kinney AY, Gammon A, Coxworth J, et al.: Exploring attitudes, beliefs, and communication preferences of Latino community members regarding BRCA1/2 mutation testing and preventive strategies. Genet Med 12 (2): 105-15, 2010. [PUBMED Abstract]
  37. Sussner KM, Edwards T, Villagra C, et al.: BRCA genetic counseling among at-risk Latinas in New York City: new beliefs shape new generation. J Genet Couns 24 (1): 134-48, 2015. [PUBMED Abstract]
  38. Zimmerman RK, Tabbarah M, Nowalk MP, et al.: Racial differences in beliefs about genetic screening among patients at inner-city neighborhood health centers. J Natl Med Assoc 98 (3): 370-7, 2006. [PUBMED Abstract]
  39. MacNew HG, Rudolph R, Brower ST, et al.: Assessing the knowledge and attitudes regarding genetic testing for breast cancer risk in our region of southeastern Georgia. Breast J 16 (2): 189-92, 2010. [PUBMED Abstract]
  40. Sussner KM, Thompson HS, Jandorf L, et al.: The influence of acculturation and breast cancer-specific distress on perceived barriers to genetic testing for breast cancer among women of African descent. Psychooncology 18 (9): 945-55, 2009. [PUBMED Abstract]
  41. Edwards TA, Thompson HS, Kwate NO, et al.: Association between temporal orientation and attitudes about BRCA1/2 testing among women of African descent with family histories of breast cancer. Patient Educ Couns 72 (2): 276-82, 2008. [PUBMED Abstract]
  42. McCarthy AM, Bristol M, Domchek SM, et al.: Health Care Segregation, Physician Recommendation, and Racial Disparities in BRCA1/2 Testing Among Women With Breast Cancer. J Clin Oncol 34 (22): 2610-8, 2016. [PUBMED Abstract]
  43. Hurtado-de-Mendoza A, Jackson MC, Anderson L, et al.: The Role of Knowledge on Genetic Counseling and Testing in Black Cancer Survivors at Increased Risk of Carrying a BRCA1/2 Mutation. J Genet Couns 26 (1): 113-121, 2017. [PUBMED Abstract]
  44. Cragun D, Weidner A, Lewis C, et al.: Racial disparities in BRCA testing and cancer risk management across a population-based sample of young breast cancer survivors. Cancer 123 (13): 2497-2505, 2017. [PUBMED Abstract]
  45. Cragun D, Weidner A, Kechik J, et al.: Genetic Testing Across Young Hispanic and Non-Hispanic White Breast Cancer Survivors: Facilitators, Barriers, and Awareness of the Genetic Information Nondiscrimination Act. Genet Test Mol Biomarkers 23 (2): 75-83, 2019. [PUBMED Abstract]
  46. Lerman C, Hughes C, Lemon SJ, et al.: What you don't know can hurt you: adverse psychologic effects in members of BRCA1-linked and BRCA2-linked families who decline genetic testing. J Clin Oncol 16 (5): 1650-4, 1998. [PUBMED Abstract]
  47. Lodder L, Frets PG, Trijsburg RW, et al.: Attitudes and distress levels in women at risk to carry a BRCA1/BRCA2 gene mutation who decline genetic testing. Am J Med Genet 119A (3): 266-72, 2003. [PUBMED Abstract]
  48. Foster C, Evans DG, Eeles R, et al.: Non-uptake of predictive genetic testing for BRCA1/2 among relatives of known carriers: attributes, cancer worry, and barriers to testing in a multicenter clinical cohort. Genet Test 8 (1): 23-9, 2004. [PUBMED Abstract]
  49. McInerney-Leo A, Biesecker BB, Hadley DW, et al.: BRCA1/2 testing in hereditary breast and ovarian cancer families II: impact on relationships. Am J Med Genet A 133 (2): 165-9, 2005. [PUBMED Abstract]
  50. Patenaude AF: Cancer susceptibility testing: risks, benefits, and personal beliefs. In: Clarke A, ed.: The Genetic Testing of Children. Oxford, England: BIOS Scientific, 1998, pp 145-156.
  51. Richards M: The genetic testing of children: adult attitude's and children's understanding. In: Clarke A, ed.: The Genetic Testing of Children. Oxford, England: BIOS Scientific, 1998, pp 169-179.
  52. Wertz DC, Fanos JH, Reilly PR: Genetic testing for children and adolescents. Who decides? JAMA 272 (11): 875-81, 1994. [PUBMED Abstract]
  53. Borry P, Stultiëns L, Nys H, et al.: Attitudes towards predictive genetic testing in minors for familial breast cancer: a systematic review. Crit Rev Oncol Hematol 64 (3): 173-81, 2007. [PUBMED Abstract]
  54. Hamann HA, Croyle RT, Venne VL, et al.: Attitudes toward the genetic testing of children among adults in a Utah-based kindred tested for a BRCA1 mutation. Am J Med Genet 92 (1): 25-32, 2000. [PUBMED Abstract]
  55. Bradbury AR, Patrick-Miller L, Pawlowski K, et al.: Should genetic testing for BRCA1/2 be permitted for minors? Opinions of BRCA mutation carriers and their adult offspring. Am J Med Genet C Semin Med Genet 148C (1): 70-7, 2008. [PUBMED Abstract]
  56. Points to consider: ethical, legal, and psychosocial implications of genetic testing in children and adolescents. American Society of Human Genetics Board of Directors, American College of Medical Genetics Board of Directors. Am J Hum Genet 57 (5): 1233-41, 1995. [PUBMED Abstract]
  57. Michie S, Marteau TM: Predictive genetic testing in children: the need for psychological research. In: Clarke A, ed.: The Genetic Testing of Children. Oxford, England: BIOS Scientific, 1998, pp 169-182.
  58. MacDonald DJ, Lessick M: Hereditary cancers in children and ethical and psychosocial implications. J Pediatr Nurs 15 (4): 217-25, 2000. [PUBMED Abstract]
  59. Tercyak KP, Peshkin BN, Streisand R, et al.: Psychological issues among children of hereditary breast cancer gene (BRCA1/2) testing participants. Psychooncology 10 (4): 336-46, 2001 Jul-Aug. [PUBMED Abstract]
  60. Winer E, Winer N, Bluman L, et al.: Attitudes and risk perceptions of women with breast cancer considering testing for BRCA1/2. [Abstract] Proceedings of the American Society of Clinical Oncology 16: A1937, 537a, 1997.
  61. Braithwaite D, Emery J, Walter F, et al.: Psychological impact of genetic counseling for familial cancer: a systematic review and meta-analysis. J Natl Cancer Inst 96 (2): 122-33, 2004. [PUBMED Abstract]
  62. Mikkelsen EM, Sunde L, Johansen C, et al.: Psychosocial conditions of women awaiting genetic counseling: a population-based study. J Genet Couns 17 (3): 242-51, 2008. [PUBMED Abstract]
  63. Dorval M, Bouchard K, Maunsell E, et al.: Health behaviors and psychological distress in women initiating BRCA1/2 genetic testing: comparison with control population. J Genet Couns 17 (4): 314-26, 2008. [PUBMED Abstract]
  64. Wang C, Gonzalez R, Janz N, et al.: The role of cognitive appraisal and worry in BRCA1/2 testing decisions among a clinic population. Psychol Health 22 (6): 719-36, 2007.
  65. Hallowell N, Statham H, Murton F: Women's understanding of their risk of developing breast/ovarian cancer before and after genetic counseling. J Genet Couns 7 (4): 345-64, 1998.
  66. MacDonald DJ, Choi J, Ferrell B, et al.: Concerns of women presenting to a comprehensive cancer centre for genetic cancer risk assessment. J Med Genet 39 (7): 526-30, 2002. [PUBMED Abstract]
  67. Matloff ET, Moyer A, Shannon KM, et al.: Healthy women with a family history of breast cancer: impact of a tailored genetic counseling intervention on risk perception, knowledge, and menopausal therapy decision making. J Womens Health (Larchmt) 15 (7): 843-56, 2006. [PUBMED Abstract]
  68. Meiser B, Price MA, Butow PN, et al.: Misperceptions of ovarian cancer risk in women at increased risk for hereditary ovarian cancer. Fam Cancer 13 (2): 153-62, 2014. [PUBMED Abstract]
  69. Bluman LG, Rimer BK, Berry DA, et al.: Attitudes, knowledge, and risk perceptions of women with breast and/or ovarian cancer considering testing for BRCA1 and BRCA2. J Clin Oncol 17 (3): 1040-6, 1999. [PUBMED Abstract]
  70. Iglehart JD, Miron A, Rimer BK, et al.: Overestimation of hereditary breast cancer risk. Ann Surg 228 (3): 375-84, 1998. [PUBMED Abstract]
  71. Bluman LG, Rimer BK, Regan Sterba K, et al.: Attitudes, knowledge, risk perceptions and decision-making among women with breast and/or ovarian cancer considering testing for BRCA1 and BRCA2 and their spouses. Psychooncology 12 (5): 410-27, 2003 Jul-Aug. [PUBMED Abstract]
  72. McCaul KD, O'Donnell SM: Naive beliefs about breast cancer risk. Womens Health 4 (1): 93-101, 1998 Spring. [PUBMED Abstract]
  73. Huiart L, Eisinger F, Stoppa-Lyonnet D, et al.: Effects of genetic consultation on perception of a family risk of breast/ovarian cancer and determinants of inaccurate perception after the consultation. J Clin Epidemiol 55 (7): 665-75, 2002. [PUBMED Abstract]
  74. Davis S, Stewart S, Bloom J: Increasing the accuracy of perceived breast cancer risk: results from a randomized trial with Cancer Information Service callers. Prev Med 39 (1): 64-73, 2004. [PUBMED Abstract]
  75. Katapodi MC, Dodd MJ, Lee KA, et al.: Underestimation of breast cancer risk: influence on screening behavior. Oncol Nurs Forum 36 (3): 306-14, 2009. [PUBMED Abstract]
  76. Lindberg NM, Wellisch D: Anxiety and compliance among women at high risk for breast cancer. Ann Behav Med 23 (4): 298-303, 2001 Fall. [PUBMED Abstract]
  77. Ritvo P, Irvine J, Robinson G, et al.: Psychological adjustment to familial-genetic risk assessment for ovarian cancer: predictors of nonadherence to surveillance recommendations. Gynecol Oncol 84 (1): 72-80, 2002. [PUBMED Abstract]
  78. Meiser B, Halliday JL: What is the impact of genetic counselling in women at increased risk of developing hereditary breast cancer? A meta-analytic review. Soc Sci Med 54 (10): 1463-70, 2002. [PUBMED Abstract]
  79. Green J, Richards M, Murton F, et al.: Family communication and genetic counseling: the case of hereditary breast and ovarian cancer. J Genet Couns 6 (1): 45-60, 1997.
  80. Quillin JM, Ramakrishnan V, Borzelleca J, et al.: Paternal relatives and family history of breast cancer. Am J Prev Med 31 (3): 265-8, 2006. [PUBMED Abstract]
  81. O'Neill SM, Rubinstein WS, Wang C, et al.: Familial risk for common diseases in primary care: the Family Healthware Impact Trial. Am J Prev Med 36 (6): 506-14, 2009. [PUBMED Abstract]
  82. Rubinstein WS, O'neill SM, Rothrock N, et al.: Components of family history associated with women's disease perceptions for cancer: a report from the Family Healthware™ Impact Trial. Genet Med 13 (1): 52-62, 2011. [PUBMED Abstract]
  83. Theis B, Boyd N, Lockwood G, et al.: Accuracy of family cancer history in breast cancer patients. Eur J Cancer Prev 3 (4): 321-7, 1994. [PUBMED Abstract]
  84. Breuer B, Kash KM, Rosenthal G, et al.: Reporting bilaterality status in first-degree relatives with breast cancer: a validity study. Genet Epidemiol 10 (4): 245-56, 1993. [PUBMED Abstract]
  85. Parent ME, Ghadirian P, Lacroix A, et al.: The reliability of recollections of family history: implications for the medical provider. J Cancer Educ 12 (2): 114-20, 1997 Summer. [PUBMED Abstract]
  86. Kelly KM, Shedlosky-Shoemaker R, Porter K, et al.: Cancer family history reporting: impact of method and psychosocial factors. J Genet Couns 16 (3): 373-82, 2007. [PUBMED Abstract]
  87. Kerber RA, Slattery ML: Comparison of self-reported and database-linked family history of cancer data in a case-control study. Am J Epidemiol 146 (3): 244-8, 1997. [PUBMED Abstract]
  88. Kerr B, Foulkes WD, Cade D, et al.: False family history of breast cancer in the family cancer clinic. Eur J Surg Oncol 24 (4): 275-9, 1998. [PUBMED Abstract]
  89. Schwartz MD, Peshkin BN, Hughes C, et al.: Impact of BRCA1/BRCA2 mutation testing on psychologic distress in a clinic-based sample. J Clin Oncol 20 (2): 514-20, 2002. [PUBMED Abstract]
  90. Mancini J, Noguès C, Adenis C, et al.: Impact of an information booklet on satisfaction and decision-making about BRCA genetic testing. Eur J Cancer 42 (7): 871-81, 2006. [PUBMED Abstract]
  91. Green MJ, Biesecker BB, McInerney AM, et al.: An interactive computer program can effectively educate patients about genetic testing for breast cancer susceptibility. Am J Med Genet 103 (1): 16-23, 2001. [PUBMED Abstract]
  92. Green MJ, Peterson SK, Baker MW, et al.: Effect of a computer-based decision aid on knowledge, perceptions, and intentions about genetic testing for breast cancer susceptibility: a randomized controlled trial. JAMA 292 (4): 442-52, 2004. [PUBMED Abstract]
  93. Green MJ, McInerney AM, Biesecker BB, et al.: Education about genetic testing for breast cancer susceptibility: patient preferences for a computer program or genetic counselor. Am J Med Genet 103 (1): 24-31, 2001. [PUBMED Abstract]
  94. Dabney MK, Huelsman K: Counseling by computer: breast cancer risk and genetic testing. Developed by the University of Wisconsin-Madison Department of Medicine and the Program in Medical Ethics. Genet Test 4 (1): 43-4, 2000. [PUBMED Abstract]
  95. Green MJ, Peterson SK, Baker MW, et al.: Use of an educational computer program before genetic counseling for breast cancer susceptibility: effects on duration and content of counseling sessions. Genet Med 7 (4): 221-9, 2005. [PUBMED Abstract]
  96. Wang C, Gonzalez R, Milliron KJ, et al.: Genetic counseling for BRCA1/2: a randomized controlled trial of two strategies to facilitate the education and counseling process. Am J Med Genet A 134 (1): 66-73, 2005. [PUBMED Abstract]
  97. Joseph G, Beattie MS, Lee R, et al.: Pre-counseling education for low literacy women at risk of Hereditary Breast and Ovarian Cancer (HBOC): patient experiences using the Cancer Risk Education Intervention Tool (CREdIT). J Genet Couns 19 (5): 447-62, 2010. [PUBMED Abstract]
  98. Albada A, van Dulmen S, Lindhout D, et al.: A pre-visit tailored website enhances counselees' realistic expectations and knowledge and fulfils information needs for breast cancer genetic counselling. Fam Cancer 11 (1): 85-95, 2012. [PUBMED Abstract]
  99. Baty BJ, Kinney AY, Ellis SM: Developing culturally sensitive cancer genetics communication aids for African Americans. Am J Med Genet 118A (2): 146-55, 2003. [PUBMED Abstract]
  100. Permuth-Wey J, Vadaparampil S, Rumphs A, et al.: Development of a culturally tailored genetic counseling booklet about hereditary breast and ovarian cancer for Black women. Am J Med Genet A 152A (4): 836-45, 2010. [PUBMED Abstract]
  101. Pal T, Stowe C, Cole A, et al.: Evaluation of phone-based genetic counselling in African American women using culturally tailored visual aids. Clin Genet 78 (2): 124-31, 2010. [PUBMED Abstract]
  102. Calzone KA: Predisposition testing for breast and ovarian cancer susceptibility. Semin Oncol Nurs 13 (2): 82-90, 1997. [PUBMED Abstract]
  103. Smith KR, West JA, Croyle RT, et al.: Familial context of genetic testing for cancer susceptibility: moderating effect of siblings' test results on psychological distress one to two weeks after BRCA1 mutation testing. Cancer Epidemiol Biomarkers Prev 8 (4 Pt 2): 385-92, 1999. [PUBMED Abstract]
  104. Wylie JE, Smith KR, Botkin JR: Effects of spouses on distress experienced by BRCA1 mutation carriers over time. Am J Med Genet 119C (1): 35-44, 2003. [PUBMED Abstract]
  105. Kash KM, Holland JC, Halper MS, et al.: Psychological distress and surveillance behaviors of women with a family history of breast cancer. J Natl Cancer Inst 84 (1): 24-30, 1992. [PUBMED Abstract]
  106. Lerman C, Schwartz M: Adherence and psychological adjustment among women at high risk for breast cancer. Breast Cancer Res Treat 28 (2): 145-55, 1993. [PUBMED Abstract]
  107. Kelly PT: Understanding Breast Cancer Risk. Philadelphia, Pa: Temple University Press, 1991.
  108. Baty BJ, Venne VL, McDonald J, et al.: BRCA1 testing: genetic counseling protocol development and counseling issues. J Genet Couns 6 (2): 223-44, 1997.
  109. Richards MP, Hallowell N, Green JM, et al.: Counseling families with hereditary breast and ovarian cancer: a psychosocial perspective. J Genet Couns 4 (3): 219-33, 1995.
  110. Hoskins KF, Stopfer JE, Calzone KA, et al.: Assessment and counseling for women with a family history of breast cancer. A guide for clinicians. JAMA 273 (7): 577-85, 1995. [PUBMED Abstract]
  111. Schneider KA: Genetic counseling for BRCA1/BRCA2 testing. Genet Test 1 (2): 91-8, 1997. [PUBMED Abstract]
  112. McKinnon WC, Baty BJ, Bennett RL, et al.: Predisposition genetic testing for late-onset disorders in adults. A position paper of the National Society of Genetic Counselors. JAMA 278 (15): 1217-20, 1997. [PUBMED Abstract]
  113. Cummings S, Olopade O: Predisposition testing for inherited breast cancer. Oncology (Huntingt) 12 (8): 1227-41; discussion 1241-2, 1998. [PUBMED Abstract]
  114. Lipkus IM, Klein WM, Rimer BK: Communicating breast cancer risks to women using different formats. Cancer Epidemiol Biomarkers Prev 10 (8): 895-8, 2001. [PUBMED Abstract]
  115. Butow PN, Lobb EA: Analyzing the process and content of genetic counseling in familial breast cancer consultations. J Genet Couns 13 (5): 403-24, 2004. [PUBMED Abstract]
  116. Lerman C, Audrain J, Croyle RT: DNA-testing for heritable breast cancer risks: lessons from traditional genetic counseling. Ann Behav Med 16(4): 327-333, 1994.
  117. Pieterse AH, van Dulmen AM, Beemer FA, et al.: Cancer genetic counseling: communication and counselees' post-visit satisfaction, cognitions, anxiety, and needs fulfillment. J Genet Couns 16 (1): 85-96, 2007. [PUBMED Abstract]
  118. Hamilton JG, Lobel M, Moyer A: Emotional distress following genetic testing for hereditary breast and ovarian cancer: a meta-analytic review. Health Psychol 28 (4): 510-8, 2009. [PUBMED Abstract]
  119. Beran TM, Stanton AL, Kwan L, et al.: The trajectory of psychological impact in BRCA1/2 genetic testing: does time heal? Ann Behav Med 36 (2): 107-16, 2008. [PUBMED Abstract]
  120. Bosch N, Junyent N, Gadea N, et al.: What factors may influence psychological well being at three months and one year post BRCA genetic result disclosure? Breast 21 (6): 755-60, 2012. [PUBMED Abstract]
  121. O'Neill SC, Rini C, Goldsmith RE, et al.: Distress among women receiving uninformative BRCA1/2 results: 12-month outcomes. Psychooncology 18 (10): 1088-96, 2009. [PUBMED Abstract]
  122. Oberguggenberger A, Sztankay M, Morscher RJ, et al.: Psychosocial outcomes and counselee satisfaction following genetic counseling for hereditary breast and ovarian cancer: A patient-reported outcome study. J Psychosom Res 89: 39-45, 2016. [PUBMED Abstract]
  123. Foster C, Watson M, Eeles R, et al.: Predictive genetic testing for BRCA1/2 in a UK clinical cohort: three-year follow-up. Br J Cancer 96 (5): 718-24, 2007. [PUBMED Abstract]
  124. Halbert CH, Stopfer JE, McDonald J, et al.: Long-term reactions to genetic testing for BRCA1 and BRCA2 mutations: does time heal women's concerns? J Clin Oncol 29 (32): 4302-6, 2011. [PUBMED Abstract]
  125. Graves KD, Vegella P, Poggi EA, et al.: Long-term psychosocial outcomes of BRCA1/BRCA2 testing: differences across affected status and risk-reducing surgery choice. Cancer Epidemiol Biomarkers Prev 21 (3): 445-55, 2012. [PUBMED Abstract]
  126. Cella D, Hughes C, Peterman A, et al.: A brief assessment of concerns associated with genetic testing for cancer: the Multidimensional Impact of Cancer Risk Assessment (MICRA) questionnaire. Health Psychol 21 (6): 564-72, 2002. [PUBMED Abstract]
  127. Shkedi-Rafid S, Gabai-Kapara E, Grinshpun-Cohen J, et al.: BRCA genetic testing of individuals from families with low prevalence of cancer: experiences of carriers and implications for population screening. Genet Med 14 (7): 688-94, 2012. [PUBMED Abstract]
  128. Metcalfe KA, Poll A, Llacuachaqui M, et al.: Patient satisfaction and cancer-related distress among unselected Jewish women undergoing genetic testing for BRCA1 and BRCA2. Clin Genet 78 (5): 411-7, 2010. [PUBMED Abstract]
  129. Metcalfe KA, Mian N, Enmore M, et al.: Long-term follow-up of Jewish women with a BRCA1 and BRCA2 mutation who underwent population genetic screening. Breast Cancer Res Treat 133 (2): 735-40, 2012. [PUBMED Abstract]
  130. Armstrong J, Toscano M, Kotchko N, et al.: Utilization and Outcomes of BRCA Genetic Testing and Counseling in a National Commercially Insured Population: The ABOUT Study. JAMA Oncol 1 (9): 1251-60, 2015. [PUBMED Abstract]
  131. Dohany L, Gustafson S, Ducaine W, et al.: Psychological distress with direct-to-consumer genetic testing: a case report of an unexpected BRCA positive test result. J Genet Couns 21 (3): 399-401, 2012. [PUBMED Abstract]
  132. Mahon SM: Impact of direct-to-consumer genetic testing. J Oncol Pract 8 (4): 260, 2012. [PUBMED Abstract]
  133. Jeffers L, Morrison PJ, McCaughan E, et al.: Maximising survival: the main concern of women with hereditary breast and ovarian cancer who undergo genetic testing for BRCA1/2. Eur J Oncol Nurs 18 (4): 411-8, 2014. [PUBMED Abstract]
  134. Francke U, Dijamco C, Kiefer AK, et al.: Dealing with the unexpected: consumer responses to direct-access BRCA mutation testing. PeerJ 1: e8, 2013. [PUBMED Abstract]
  135. Meiser B, Quinn VF, Gleeson M, et al.: When knowledge of a heritable gene mutation comes out of the blue: treatment-focused genetic testing in women newly diagnosed with breast cancer. Eur J Hum Genet 24 (11): 1517-1523, 2016. [PUBMED Abstract]
  136. Ardern-Jones A, Kenen R, Eeles R: Too much, too soon? Patients and health professionals' views concerning the impact of genetic testing at the time of breast cancer diagnosis in women under the age of 40. Eur J Cancer Care (Engl) 14 (3): 272-81, 2005. [PUBMED Abstract]
  137. Wevers MR, Hahn DE, Verhoef S, et al.: Breast cancer genetic counseling after diagnosis but before treatment: a pilot study on treatment consequences and psychological impact. Patient Educ Couns 89 (1): 89-95, 2012. [PUBMED Abstract]
  138. Zilliacus E, Meiser B, Gleeson M, et al.: Are we being overly cautious? A qualitative inquiry into the experiences and perceptions of treatment-focused germline BRCA genetic testing amongst women recently diagnosed with breast cancer. Support Care Cancer 20 (11): 2949-58, 2012. [PUBMED Abstract]
  139. Tercyak KP, Peshkin BN, Brogan BM, et al.: Quality of life after contralateral prophylactic mastectomy in newly diagnosed high-risk breast cancer patients who underwent BRCA1/2 gene testing. J Clin Oncol 25 (3): 285-91, 2007. [PUBMED Abstract]
  140. Wevers MR, Ausems MG, Verhoef S, et al.: Does rapid genetic counseling and testing in newly diagnosed breast cancer patients cause additional psychosocial distress? results from a randomized clinical trial. Genet Med 18 (2): 137-44, 2016. [PUBMED Abstract]
  141. Wevers MR, Aaronson NK, Bleiker EMA, et al.: Rapid genetic counseling and testing in newly diagnosed breast cancer: Patients' and health professionals' attitudes, experiences, and evaluation of effects on treatment decision making. J Surg Oncol 116 (8): 1029-1039, 2017. [PUBMED Abstract]
  142. Lerman C, Peshkin BN, Hughes C, et al.: Family disclosure in genetic testing for cancer susceptibility: determinants and consequences. Journal of Health Care Law and Policy 1 (2): 353-73, 1998.
  143. Elrick A, Ashida S, Ivanovich J, et al.: Psychosocial and Clinical Factors Associated with Family Communication of Cancer Genetic Test Results among Women Diagnosed with Breast Cancer at a Young Age. J Genet Couns 26 (1): 173-181, 2017. [PUBMED Abstract]
  144. Finlay E, Stopfer JE, Burlingame E, et al.: Factors determining dissemination of results and uptake of genetic testing in families with known BRCA1/2 mutations. Genet Test 12 (1): 81-91, 2008. [PUBMED Abstract]
  145. Hughes C, Lerman C, Schwartz M, et al.: All in the family: evaluation of the process and content of sisters' communication about BRCA1 and BRCA2 genetic test results. Am J Med Genet 107 (2): 143-50, 2002. [PUBMED Abstract]
  146. Baars JE, Ausems MG, van Riel E, et al.: Communication Between Breast Cancer Patients Who Received Inconclusive Genetic Test Results and Their Daughters and Sisters Years After Testing. J Genet Couns 25 (3): 461-71, 2016. [PUBMED Abstract]
  147. Wagner Costalas J, Itzen M, Malick J, et al.: Communication of BRCA1 and BRCA2 results to at-risk relatives: a cancer risk assessment program's experience. Am J Med Genet 119C (1): 11-8, 2003. [PUBMED Abstract]
  148. Patenaude AF, Dorval M, DiGianni LS, et al.: Sharing BRCA1/2 test results with first-degree relatives: factors predicting who women tell. J Clin Oncol 24 (4): 700-6, 2006. [PUBMED Abstract]
  149. MacDonald DJ, Sarna L, van Servellen G, et al.: Selection of family members for communication of cancer risk and barriers to this communication before and after genetic cancer risk assessment. Genet Med 9 (5): 275-82, 2007. [PUBMED Abstract]
  150. Kenen R, Arden-Jones A, Eeles R: Healthy women from suspected hereditary breast and ovarian cancer families: the significant others in their lives. Eur J Cancer Care (Engl) 13 (2): 169-79, 2004. [PUBMED Abstract]
  151. Claes E, Evers-Kiebooms G, Boogaerts A, et al.: Communication with close and distant relatives in the context of genetic testing for hereditary breast and ovarian cancer in cancer patients. Am J Med Genet 116A (1): 11-9, 2003. [PUBMED Abstract]
  152. Foster C, Eeles R, Ardern-Jones A, et al.: Juggling roles and expectations: dilemmas faced by women talking to relatives about cancer and genetic testing. Psychol Health 19 (4): 439-55, 2004.
  153. Kenen R, Arden-Jones A, Eeles R: We are talking, but are they listening? Communication patterns in families with a history of breast/ovarian cancer (HBOC). Psychooncology 13 (5): 335-45, 2004. [PUBMED Abstract]
  154. Segal J, Esplen MJ, Toner B, et al.: An investigation of the disclosure process and support needs of BRCA1 and BRCA2 carriers. Am J Med Genet A 125 (3): 267-72, 2004. [PUBMED Abstract]
  155. Bradbury AR, Dignam JJ, Ibe CN, et al.: How often do BRCA mutation carriers tell their young children of the family's risk for cancer? A study of parental disclosure of BRCA mutations to minors and young adults. J Clin Oncol 25 (24): 3705-11, 2007. [PUBMED Abstract]
  156. Bradbury AR, Patrick-Miller L, Pawlowski K, et al.: Learning of your parent's BRCA mutation during adolescence or early adulthood: a study of offspring experiences. Psychooncology 18 (2): 200-8, 2009. [PUBMED Abstract]
  157. Manne S, Audrain J, Schwartz M, et al.: Associations between relationship support and psychological reactions of participants and partners to BRCA1 and BRCA2 testing in a clinic-based sample. Ann Behav Med 28 (3): 211-25, 2004. [PUBMED Abstract]
  158. Daly MB, Montgomery S, Bingler R, et al.: Communicating genetic test results within the family: Is it lost in translation? A survey of relatives in the randomized six-step study. Fam Cancer 15 (4): 697-706, 2016. [PUBMED Abstract]
  159. McAllister MF, Evans DG, Ormiston W, et al.: Men in breast cancer families: a preliminary qualitative study of awareness and experience. J Med Genet 35 (9): 739-44, 1998. [PUBMED Abstract]
  160. Liede A, Metcalfe K, Hanna D, et al.: Evaluation of the needs of male carriers of mutations in BRCA1 or BRCA2 who have undergone genetic counseling. Am J Hum Genet 67 (6): 1494-504, 2000. [PUBMED Abstract]
  161. Metcalfe KA, Liede A, Trinkaus M, et al.: Evaluation of the needs of spouses of female carriers of mutations in BRCA1 and BRCA2. Clin Genet 62 (6): 464-9, 2002. [PUBMED Abstract]
  162. Mireskandari S, Sherman KA, Meiser B, et al.: Psychological adjustment among partners of women at high risk of developing breast/ovarian cancer. Genet Med 9 (5): 311-20, 2007. [PUBMED Abstract]
  163. Sherman KA, Kasparian NA, Mireskandari S: Psychological adjustment among male partners in response to women's breast/ovarian cancer risk: a theoretical review of the literature. Psychooncology 19 (1): 1-11, 2010. [PUBMED Abstract]
  164. Daly MB: The impact of social roles on the experience of men in BRCA1/2 families: implications for counseling. J Genet Couns 18 (1): 42-8, 2009. [PUBMED Abstract]
  165. DudokdeWit AC, Tibben A, Frets PG, et al.: Males at-risk for the BRCA1 gene, the psychological impact. Psychooncology 5(3): 251-257, 1996.
  166. Lodder L, Frets PG, Trijsburg RW, et al.: Men at risk of being a mutation carrier for hereditary breast/ovarian cancer: an exploration of attitudes and psychological functioning during genetic testing. Eur J Hum Genet 9 (7): 492-500, 2001. [PUBMED Abstract]
  167. Lerman C, Hughes C, Croyle RT, et al.: Prophylactic surgery decisions and surveillance practices one year following BRCA1/2 testing. Prev Med 31 (1): 75-80, 2000. [PUBMED Abstract]
  168. Hughes C, Lynch H, Durham C, et al.: Communication of BRCA1/2 Test Results in Hereditary Breast Cancer Families. Cancer Research in Therapy and Control Vol. 8, 1999, pp. 51-59.
  169. Tercyak KP, Hughes C, Main D, et al.: Parental communication of BRCA1/2 genetic test results to children. Patient Educ Couns 42 (3): 213-24, 2001. [PUBMED Abstract]
  170. Tercyak KP, Peshkin BN, DeMarco TA, et al.: Parent-child factors and their effect on communicating BRCA1/2 test results to children. Patient Educ Couns 47 (2): 145-53, 2002. [PUBMED Abstract]
  171. McGivern B, Everett J, Yager GG, et al.: Family communication about positive BRCA1 and BRCA2 genetic test results. Genet Med 6 (6): 503-9, 2004 Nov-Dec. [PUBMED Abstract]
  172. Bradbury AR, Patrick-Miller L, Schwartz L, et al.: Psychosocial Adjustment in School-age Girls With a Family History of Breast Cancer. Pediatrics 136 (5): 927-37, 2015. [PUBMED Abstract]
  173. Bradbury AR, Patrick-Miller L, Schwartz LA, et al.: Psychosocial Adjustment and Perceived Risk Among Adolescent Girls From Families With BRCA1/2 or Breast Cancer History. J Clin Oncol 34 (28): 3409-16, 2016. [PUBMED Abstract]
  174. Tercyak KP, Peshkin BN, Demarco TA, et al.: Information needs of mothers regarding communicating BRCA1/2 cancer genetic test results to their children. Genet Test 11 (3): 249-55, 2007. [PUBMED Abstract]
  175. Wertz DC: International perspectives. In: Clarke A, ed.: The Genetic Testing of Children. Oxford, England: BIOS Scientific, 1998, pp 271-287.
  176. Benkendorf JL, Reutenauer JE, Hughes CA, et al.: Patients' attitudes about autonomy and confidentiality in genetic testing for breast-ovarian cancer susceptibility. Am J Med Genet 73 (3): 296-303, 1997. [PUBMED Abstract]
  177. Staton AD, Kurian AW, Cobb K, et al.: Cancer risk reduction and reproductive concerns in female BRCA1/2 mutation carriers. Fam Cancer 7 (2): 179-86, 2008. [PUBMED Abstract]
  178. Friedman LC, Kramer RM: Reproductive issues for women with BRCA mutations. J Natl Cancer Inst Monogr (34): 83-6, 2005. [PUBMED Abstract]
  179. Smith KR, Ellington L, Chan AY, et al.: Fertility intentions following testing for a BRCA1 gene mutation. Cancer Epidemiol Biomarkers Prev 13 (5): 733-40, 2004. [PUBMED Abstract]
  180. Quinn G, Vadaparampil S, Wilson C, et al.: Attitudes of high-risk women toward preimplantation genetic diagnosis. Fertil Steril 91 (6): 2361-8, 2009. [PUBMED Abstract]
  181. Cunniff C; American Academy of Pediatrics Committee on Genetics: Prenatal screening and diagnosis for pediatricians. Pediatrics 114 (3): 889-94, 2004. [PUBMED Abstract]
  182. Rappaport VJ: Prenatal diagnosis and genetic screening--integration into prenatal care. Obstet Gynecol Clin North Am 35 (3): 435-58, ix, 2008. [PUBMED Abstract]
  183. Baruch S, Kaufman D, Hudson KL: Genetic testing of embryos: practices and perspectives of US in vitro fertilization clinics. Fertil Steril 89 (5): 1053-8, 2008. [PUBMED Abstract]
  184. Ogilvie CM, Braude PR, Scriven PN: Preimplantation genetic diagnosis--an overview. J Histochem Cytochem 53 (3): 255-60, 2005. [PUBMED Abstract]
  185. Vadaparampil ST, Quinn GP, Knapp C, et al.: Factors associated with preimplantation genetic diagnosis acceptance among women concerned about hereditary breast and ovarian cancer. Genet Med 11 (10): 757-65, 2009. [PUBMED Abstract]
  186. Quinn GP, Vadaparampil ST, King LM, et al.: Conflict between values and technology: perceptions of preimplantation genetic diagnosis among women at increased risk for hereditary breast and ovarian cancer. Fam Cancer 8 (4): 441-9, 2009. [PUBMED Abstract]
  187. Chan JL, Johnson LNC, Sammel MD, et al.: Reproductive Decision-Making in Women with BRCA1/2 Mutations. J Genet Couns 26 (3): 594-603, 2017. [PUBMED Abstract]
  188. Menon U, Harper J, Sharma A, et al.: Views of BRCA gene mutation carriers on preimplantation genetic diagnosis as a reproductive option for hereditary breast and ovarian cancer. Hum Reprod 22 (6): 1573-7, 2007. [PUBMED Abstract]
  189. Fortuny D, Balmaña J, Graña B, et al.: Opinion about reproductive decision making among individuals undergoing BRCA1/2 genetic testing in a multicentre Spanish cohort. Hum Reprod 24 (4): 1000-6, 2009. [PUBMED Abstract]
  190. Julian-Reynier C, Fabre R, Coupier I, et al.: BRCA1/2 carriers: their childbearing plans and theoretical intentions about having preimplantation genetic diagnosis and prenatal diagnosis. Genet Med 14 (5): 527-34, 2012. [PUBMED Abstract]
  191. Ormondroyd E, Donnelly L, Moynihan C, et al.: Attitudes to reproductive genetic testing in women who had a positive BRCA test before having children: a qualitative analysis. Eur J Hum Genet 20 (1): 4-10, 2012. [PUBMED Abstract]
  192. Quinn GP, Vadaparampil ST, Miree CA, et al.: High risk men's perceptions of pre-implantation genetic diagnosis for hereditary breast and ovarian cancer. Hum Reprod 25 (10): 2543-50, 2010. [PUBMED Abstract]
  193. Struewing JP, Abeliovich D, Peretz T, et al.: The carrier frequency of the BRCA1 185delAG mutation is approximately 1 percent in Ashkenazi Jewish individuals. Nat Genet 11 (2): 198-200, 1995. [PUBMED Abstract]
  194. Rothenberg KH: Breast cancer, the genetic "quick fix," and the Jewish community. Ethical, legal, and social challenges. Health Matrix Clevel 7 (1): 97-124, 1997 Winter. [PUBMED Abstract]
  195. Foster MW, Bernsten D, Carter TH: A model agreement for genetic research in socially identifiable populations. Am J Hum Genet 63 (3): 696-702, 1998. [PUBMED Abstract]
  196. Burhansstipanov L, Bemis LT, Dignan MB: Native American cancer education: genetic and cultural issues. J Cancer Educ 16 (3): 142-5, 2001 Autumn. [PUBMED Abstract]
  197. Hughes C, Fasaye GA, LaSalle VH, et al.: Sociocultural influences on participation in genetic risk assessment and testing among African American women. Patient Educ Couns 51 (2): 107-14, 2003. [PUBMED Abstract]
  198. Julian-Reynier CM, Bouchard LJ, Evans DG, et al.: Women's attitudes toward preventive strategies for hereditary breast or ovarian carcinoma differ from one country to another: differences among English, French, and Canadian women. Cancer 92 (4): 959-68, 2001. [PUBMED Abstract]
  199. Phillips KA, Warner E, Meschino WS, et al.: Perceptions of Ashkenazi Jewish breast cancer patients on genetic testing for mutations in BRCA1 and BRCA2. Clin Genet 57 (5): 376-83, 2000. [PUBMED Abstract]
  200. Vadaparampil ST, Quinn GP, Small BJ, et al.: A pilot study of hereditary breast and ovarian knowledge among a multiethnic group of Hispanic women with a personal or family history of cancer. Genet Test Mol Biomarkers 14 (1): 99-106, 2010. [PUBMED Abstract]
  201. Halbert CH, Kessler L, Troxel AB, et al.: Effect of genetic counseling and testing for BRCA1 and BRCA2 mutations in African American women: a randomized trial. Public Health Genomics 13 (7-8): 440-8, 2010. [PUBMED Abstract]
  202. Freedman TG: Genetic susceptibility testing: ethical and social quandaries. Health Soc Work 23 (3): 214-22, 1998. [PUBMED Abstract]
  203. Hubbard R, Lewontin RC: Pitfalls of genetic testing. N Engl J Med 334 (18): 1192-4, 1996. [PUBMED Abstract]
  204. Parens E: Glad and terrified: on the ethics of BRACA1 and 2 testing. Cancer Invest 14 (4): 405-11, 1996. [PUBMED Abstract]
  205. Winter PR, Wiesner GL, Finnegan J, et al.: Notification of a family history of breast cancer: issues of privacy and confidentiality. Am J Med Genet 66 (1): 1-6, 1996. [PUBMED Abstract]
  206. Geller G, Doksum T, Bernhardt BA, et al.: Participation in breast cancer susceptibility testing protocols: influence of recruitment source, altruism, and family involvement on women's decisions. Cancer Epidemiol Biomarkers Prev 8 (4 Pt 2): 377-83, 1999. [PUBMED Abstract]
  207. Rimer BK, Schildkraut JM, Lerman C, et al.: Participation in a women's breast cancer risk counseling trial. Who participates? Who declines? High Risk Breast Cancer Consortium. Cancer 77 (11): 2348-55, 1996. [PUBMED Abstract]
  208. Robson ME, Storm CD, Weitzel J, et al.: American Society of Clinical Oncology policy statement update: genetic and genomic testing for cancer susceptibility. J Clin Oncol 28 (5): 893-901, 2010. [PUBMED Abstract]
  209. Burke W, Daly M, Garber J, et al.: Recommendations for follow-up care of individuals with an inherited predisposition to cancer. II. BRCA1 and BRCA2. Cancer Genetics Studies Consortium. JAMA 277 (12): 997-1003, 1997. [PUBMED Abstract]
  210. Durfy SJ, Buchanan TE, Burke W: Testing for inherited susceptibility to breast cancer: a survey of informed consent forms for BRCA1 and BRCA2 mutation testing. Am J Med Genet 75 (1): 82-7, 1998. [PUBMED Abstract]
  211. Statement of the American Society of Clinical Oncology: genetic testing for cancer susceptibility, Adopted on February 20, 1996. J Clin Oncol 14 (5): 1730-6; discussion 1737-40, 1996. [PUBMED Abstract]
  212. Hallowell N, Foster C, Eeles R, et al.: Balancing autonomy and responsibility: the ethics of generating and disclosing genetic information. J Med Ethics 29 (2): 74-9; discussion 80-3, 2003. [PUBMED Abstract]
  213. Metcalfe KA, Poll A, O'Connor A, et al.: Development and testing of a decision aid for breast cancer prevention for women with a BRCA1 or BRCA2 mutation. Clin Genet 72 (3): 208-17, 2007. [PUBMED Abstract]
  214. Tiller K, Meiser B, Gaff C, et al.: A randomized controlled trial of a decision aid for women at increased risk of ovarian cancer. Med Decis Making 26 (4): 360-72, 2006 Jul-Aug. [PUBMED Abstract]
  215. van Roosmalen MS, Stalmeier PF, Verhoef LC, et al.: Randomized trial of a shared decision-making intervention consisting of trade-offs and individualized treatment information for BRCA1/2 mutation carriers. J Clin Oncol 22 (16): 3293-301, 2004. [PUBMED Abstract]
  216. Culver JO, MacDonald DJ, Thornton AA, et al.: Development and evaluation of a decision aid for BRCA carriers with breast cancer. J Genet Couns 20 (3): 294-307, 2011. [PUBMED Abstract]
  217. Metcalfe KA, Dennis CL, Poll A, et al.: Effect of decision aid for breast cancer prevention on decisional conflict in women with a BRCA1 or BRCA2 mutation: a multisite, randomized, controlled trial. Genet Med 19 (3): 330-336, 2017. [PUBMED Abstract]
  218. Schwartz MD, Lerman C, Brogan B, et al.: Impact of BRCA1/BRCA2 counseling and testing on newly diagnosed breast cancer patients. J Clin Oncol 22 (10): 1823-9, 2004. [PUBMED Abstract]
  219. Botkin JR, Smith KR, Croyle RT, et al.: Genetic testing for a BRCA1 mutation: prophylactic surgery and screening behavior in women 2 years post testing. Am J Med Genet A 118 (3): 201-9, 2003. [PUBMED Abstract]
  220. Beattie MS, Crawford B, Lin F, et al.: Uptake, time course, and predictors of risk-reducing surgeries in BRCA carriers. Genet Test Mol Biomarkers 13 (1): 51-6, 2009. [PUBMED Abstract]
  221. O'Neill SC, Valdimarsdottir HB, Demarco TA, et al.: BRCA1/2 test results impact risk management attitudes, intentions, and uptake. Breast Cancer Res Treat 124 (3): 755-64, 2010. [PUBMED Abstract]
  222. Schwartz MD, Isaacs C, Graves KD, et al.: Long-term outcomes of BRCA1/BRCA2 testing: risk reduction and surveillance. Cancer 118 (2): 510-7, 2012. [PUBMED Abstract]
  223. Garcia C, Wendt J, Lyon L, et al.: Risk management options elected by women after testing positive for a BRCA mutation. Gynecol Oncol 132 (2): 428-33, 2014. [PUBMED Abstract]
  224. Singh K, Lester J, Karlan B, et al.: Impact of family history on choosing risk-reducing surgery among BRCA mutation carriers. Am J Obstet Gynecol 208 (4): 329.e1-6, 2013. [PUBMED Abstract]
  225. Phillips KA, Jenkins MA, Lindeman GJ, et al.: Risk-reducing surgery, screening and chemoprevention practices of BRCA1 and BRCA2 mutation carriers: a prospective cohort study. Clin Genet 70 (3): 198-206, 2006. [PUBMED Abstract]
  226. Metcalfe KA, Birenbaum-Carmeli D, Lubinski J, et al.: International variation in rates of uptake of preventive options in BRCA1 and BRCA2 mutation carriers. Int J Cancer 122 (9): 2017-22, 2008. [PUBMED Abstract]
  227. Julian-Reynier C, Mancini J, Mouret-Fourme E, et al.: Cancer risk management strategies and perceptions of unaffected women 5 years after predictive genetic testing for BRCA1/2 mutations. Eur J Hum Genet 19 (5): 500-6, 2011. [PUBMED Abstract]
  228. Scheuer L, Kauff N, Robson M, et al.: Outcome of preventive surgery and screening for breast and ovarian cancer in BRCA mutation carriers. J Clin Oncol 20 (5): 1260-8, 2002. [PUBMED Abstract]
  229. Mannis GN, Fehniger JE, Creasman JS, et al.: Risk-reducing salpingo-oophorectomy and ovarian cancer screening in 1077 women after BRCA testing. JAMA Intern Med 173 (2): 96-103, 2013. [PUBMED Abstract]
  230. Friebel TM, Domchek SM, Neuhausen SL, et al.: Bilateral prophylactic oophorectomy and bilateral prophylactic mastectomy in a prospective cohort of unaffected BRCA1 and BRCA2 mutation carriers. Clin Breast Cancer 7 (11): 875-82, 2007. [PUBMED Abstract]
  231. Madalinska JB, van Beurden M, Bleiker EM, et al.: Predictors of prophylactic bilateral salpingo-oophorectomy compared with gynecologic screening use in BRCA1/2 mutation carriers. J Clin Oncol 25 (3): 301-7, 2007. [PUBMED Abstract]
  232. Rhiem K, Foth D, Wappenschmidt B, et al.: Risk-reducing salpingo-oophorectomy in BRCA1 and BRCA2 mutation carriers. Arch Gynecol Obstet 283 (3): 623-7, 2011. [PUBMED Abstract]
  233. Sidon L, Ingham S, Clancy T, et al.: Uptake of risk-reducing salpingo-oophorectomy in women carrying a BRCA1 or BRCA2 mutation: evidence for lower uptake in women affected by breast cancer and older women. Br J Cancer 106 (4): 775-9, 2012. [PUBMED Abstract]
  234. Lerman C, Narod S, Schulman K, et al.: BRCA1 testing in families with hereditary breast-ovarian cancer. A prospective study of patient decision making and outcomes. JAMA 275 (24): 1885-92, 1996. [PUBMED Abstract]
  235. van Dijk S, van Roosmalen MS, Otten W, et al.: Decision making regarding prophylactic mastectomy: stability of preferences and the impact of anticipated feelings of regret. J Clin Oncol 26 (14): 2358-63, 2008. [PUBMED Abstract]
  236. Claes E, Evers-Kiebooms G, Decruyenaere M, et al.: Surveillance behavior and prophylactic surgery after predictive testing for hereditary breast/ovarian cancer. Behav Med 31 (3): 93-105, 2005. [PUBMED Abstract]
  237. Ray JA, Loescher LJ, Brewer M: Risk-reduction surgery decisions in high-risk women seen for genetic counseling. J Genet Couns 14 (6): 473-84, 2005. [PUBMED Abstract]
  238. Hallowell N: 'You don't want to lose your ovaries because you think 'I might become a man". Women's perceptions of prophylactic surgery as a cancer risk management option. Psychooncology 7 (3): 263-75, 1998 May-Jun. [PUBMED Abstract]
  239. Schneider KA, Stopfer JE, Peters JA, et al.: Complexities in cancer risk counseling: presentation of three cases. J Genet Couns 6 (2): 147-67, 1997.
  240. Tarkan L: My Mother's Breast: Daughters Face Their Mothers' Cancer. Dallas, TX: Taylor Publishing, 1999.
  241. Stefanek ME, Helzlsouer KJ, Wilcox PM, et al.: Predictors of and satisfaction with bilateral prophylactic mastectomy. Prev Med 24 (4): 412-9, 1995. [PUBMED Abstract]
  242. Cortesi L, Razzaboni E, Toss A, et al.: A rapid genetic counselling and testing in newly diagnosed breast cancer is associated with high rate of risk-reducing mastectomy in BRCA1/2-positive Italian women. Ann Oncol 25 (1): 57-63, 2014. [PUBMED Abstract]
  243. Rosenberg SM, Ruddy KJ, Tamimi RM, et al.: BRCA1 and BRCA2 Mutation Testing in Young Women With Breast Cancer. JAMA Oncol 2 (6): 730-6, 2016. [PUBMED Abstract]
  244. Graves KD, Peshkin BN, Halbert CH, et al.: Predictors and outcomes of contralateral prophylactic mastectomy among breast cancer survivors. Breast Cancer Res Treat 104 (3): 321-9, 2007. [PUBMED Abstract]
  245. Howard-McNatt M, Schroll RW, Hurt GJ, et al.: Contralateral prophylactic mastectomy in breast cancer patients who test negative for BRCA mutations. Am J Surg 202 (3): 298-302, 2011. [PUBMED Abstract]
  246. Bresser PJ, Seynaeve C, Van Gool AR, et al.: Satisfaction with prophylactic mastectomy and breast reconstruction in genetically predisposed women. Plast Reconstr Surg 117 (6): 1675-82; discussion 1683-4, 2006. [PUBMED Abstract]
  247. Brandberg Y, Sandelin K, Erikson S, et al.: Psychological reactions, quality of life, and body image after bilateral prophylactic mastectomy in women at high risk for breast cancer: a prospective 1-year follow-up study. J Clin Oncol 26 (24): 3943-9, 2008. [PUBMED Abstract]
  248. Lobb E, Meiser B: Genetic counselling and prophylactic surgery in women from families with hereditary breast or ovarian cancer. Lancet 363 (9424): 1841-2, 2004. [PUBMED Abstract]
  249. Lobb EA, Butow PN, Meiser B, et al.: Tailoring communication in consultations with women from high risk breast cancer families. Br J Cancer 87 (5): 502-8, 2002. [PUBMED Abstract]
  250. Lobb EA, Butow PN, Barratt A, et al.: Communication and information-giving in high-risk breast cancer consultations: influence on patient outcomes. Br J Cancer 90 (2): 321-7, 2004. [PUBMED Abstract]
  251. Schwartz MD, Kaufman E, Peshkin BN, et al.: Bilateral prophylactic oophorectomy and ovarian cancer screening following BRCA1/BRCA2 mutation testing. J Clin Oncol 21 (21): 4034-41, 2003. [PUBMED Abstract]
  252. Kauff ND, Satagopan JM, Robson ME, et al.: Risk-reducing salpingo-oophorectomy in women with a BRCA1 or BRCA2 mutation. N Engl J Med 346 (21): 1609-15, 2002. [PUBMED Abstract]
  253. Schmeler KM, Sun CC, Bodurka DC, et al.: Prophylactic bilateral salpingo-oophorectomy compared with surveillance in women with BRCA mutations. Obstet Gynecol 108 (3 Pt 1): 515-20, 2006. [PUBMED Abstract]
  254. MacDonald DJ, Sarna L, Uman GC, et al.: Cancer screening and risk-reducing behaviors of women seeking genetic cancer risk assessment for breast and ovarian cancers. Oncol Nurs Forum 33 (2): E27-35, 2006. [PUBMED Abstract]
  255. Litton JK, Westin SN, Ready K, et al.: Perception of screening and risk reduction surgeries in patients tested for a BRCA deleterious mutation. Cancer 115 (8): 1598-604, 2009. [PUBMED Abstract]
  256. Tyndel S, Austoker J, Henderson BJ, et al.: What is the psychological impact of mammographic screening on younger women with a family history of breast cancer? Findings from a prospective cohort study by the PIMMS Management Group. J Clin Oncol 25 (25): 3823-30, 2007. [PUBMED Abstract]
  257. Rees G, Young MA, Gaff C, et al.: A qualitative study of health professionals' views regarding provision of information about health-protective behaviors during genetic consultation for breast cancer. J Genet Couns 15 (2): 95-104, 2006. [PUBMED Abstract]
  258. Lodder LN, Frets PG, Trijsburg RW, et al.: One year follow-up of women opting for presymptomatic testing for BRCA1 and BRCA2: emotional impact of the test outcome and decisions on risk management (surveillance or prophylactic surgery). Breast Cancer Res Treat 73 (2): 97-112, 2002. [PUBMED Abstract]
  259. van Oostrom I, Meijers-Heijboer H, Lodder LN, et al.: Long-term psychological impact of carrying a BRCA1/2 mutation and prophylactic surgery: a 5-year follow-up study. J Clin Oncol 21 (20): 3867-74, 2003. [PUBMED Abstract]
  260. Bresser PJ, Seynaeve C, Van Gool AR, et al.: The course of distress in women at increased risk of breast and ovarian cancer due to an (identified) genetic susceptibility who opt for prophylactic mastectomy and/or salpingo-oophorectomy. Eur J Cancer 43 (1): 95-103, 2007. [PUBMED Abstract]
  261. Frost MH, Schaid DJ, Sellers TA, et al.: Long-term satisfaction and psychological and social function following bilateral prophylactic mastectomy. JAMA 284 (3): 319-24, 2000. [PUBMED Abstract]
  262. Metcalfe KA, Esplen MJ, Goel V, et al.: Psychosocial functioning in women who have undergone bilateral prophylactic mastectomy. Psychooncology 13 (1): 14-25, 2004. [PUBMED Abstract]
  263. Schlich-Bakker KJ, Ausems MG, Schipper M, et al.: BRCA1/2 mutation testing in breast cancer patients: a prospective study of the long-term psychological impact of approach during adjuvant radiotherapy. Breast Cancer Res Treat 109 (3): 507-14, 2008. [PUBMED Abstract]
  264. Frost MH, Slezak JM, Tran NV, et al.: Satisfaction after contralateral prophylactic mastectomy: the significance of mastectomy type, reconstructive complications, and body appearance. J Clin Oncol 23 (31): 7849-56, 2005. [PUBMED Abstract]
  265. Schwartz MD: Contralateral prophylactic mastectomy: efficacy, satisfaction, and regret. J Clin Oncol 23 (31): 7777-9, 2005. [PUBMED Abstract]
  266. Metcalfe KA, Cil TD, Semple JL, et al.: Long-Term Psychosocial Functioning in Women with Bilateral Prophylactic Mastectomy: Does Preservation of the Nipple-Areolar Complex Make a Difference? Ann Surg Oncol 22 (10): 3324-30, 2015. [PUBMED Abstract]
  267. Yao K, Liederbach E, Tang R, et al.: Nipple-sparing mastectomy in BRCA1/2 mutation carriers: an interim analysis and review of the literature. Ann Surg Oncol 22 (2): 370-6, 2015. [PUBMED Abstract]
  268. Geiger AM, Nekhlyudov L, Herrinton LJ, et al.: Quality of life after bilateral prophylactic mastectomy. Ann Surg Oncol 14 (2): 686-94, 2007. [PUBMED Abstract]
  269. Isern AE, Tengrup I, Loman N, et al.: Aesthetic outcome, patient satisfaction, and health-related quality of life in women at high risk undergoing prophylactic mastectomy and immediate breast reconstruction. J Plast Reconstr Aesthet Surg 61 (10): 1177-87, 2008. [PUBMED Abstract]
  270. Kenen RH, Shapiro PJ, Hantsoo L, et al.: Women with BRCA1 or BRCA2 mutations renegotiating a post-prophylactic mastectomy identity: self-image and self-disclosure. J Genet Couns 16 (6): 789-98, 2007. [PUBMED Abstract]
  271. Altschuler A, Nekhlyudov L, Rolnick SJ, et al.: Positive, negative, and disparate--women's differing long-term psychosocial experiences of bilateral or contralateral prophylactic mastectomy. Breast J 14 (1): 25-32, 2008 Jan-Feb. [PUBMED Abstract]
  272. Patenaude AF, Orozco S, Li X, et al.: Support needs and acceptability of psychological and peer consultation: attitudes of 108 women who had undergone or were considering prophylactic mastectomy. Psychooncology 17 (8): 831-43, 2008. [PUBMED Abstract]
  273. Elit L, Esplen MJ, Butler K, et al.: Quality of life and psychosexual adjustment after prophylactic oophorectomy for a family history of ovarian cancer. Fam Cancer 1 (3-4): 149-56, 2001. [PUBMED Abstract]
  274. Robson M, Hensley M, Barakat R, et al.: Quality of life in women at risk for ovarian cancer who have undergone risk-reducing oophorectomy. Gynecol Oncol 89 (2): 281-7, 2003. [PUBMED Abstract]
  275. Finch A, Metcalfe KA, Chiang JK, et al.: The impact of prophylactic salpingo-oophorectomy on menopausal symptoms and sexual function in women who carry a BRCA mutation. Gynecol Oncol 121 (1): 163-8, 2011. [PUBMED Abstract]
  276. Finch A, Metcalfe K, Lui J, et al.: Breast and ovarian cancer risk perception after prophylactic salpingo-oophorectomy due to an inherited mutation in the BRCA1 or BRCA2 gene. Clin Genet 75 (3): 220-4, 2009. [PUBMED Abstract]
  277. Madalinska JB, Hollenstein J, Bleiker E, et al.: Quality-of-life effects of prophylactic salpingo-oophorectomy versus gynecologic screening among women at increased risk of hereditary ovarian cancer. J Clin Oncol 23 (28): 6890-8, 2005. [PUBMED Abstract]
  278. Westin SN, Sun CC, Lu KH, et al.: Satisfaction with ovarian carcinoma risk-reduction strategies among women at high risk for breast and ovarian carcinoma. Cancer 117 (12): 2659-67, 2011. [PUBMED Abstract]
  279. Campfield Bonadies D, Moyer A, Matloff ET: What I wish I'd known before surgery: BRCA carriers' perspectives after bilateral salipingo-oophorectomy. Fam Cancer 10 (1): 79-85, 2011. [PUBMED Abstract]
  280. Isaacs C, Peshkin BN, Schwartz M, et al.: Breast and ovarian cancer screening practices in healthy women with a strong family history of breast or ovarian cancer. Breast Cancer Res Treat 71 (2): 103-12, 2002. [PUBMED Abstract]
  281. Peshkin BN, Schwartz MD, Isaacs C, et al.: Utilization of breast cancer screening in a clinically based sample of women after BRCA1/2 testing. Cancer Epidemiol Biomarkers Prev 11 (10 Pt 1): 1115-8, 2002. [PUBMED Abstract]
  282. Tinley ST, Houfek J, Watson P, et al.: Screening adherence in BRCA1/2 families is associated with primary physicians' behavior. Am J Med Genet A 125 (1): 5-11, 2004. [PUBMED Abstract]
  283. Lerman C, Seay J, Balshem A, et al.: Interest in genetic testing among first-degree relatives of breast cancer patients. Am J Med Genet 57 (3): 385-92, 1995. [PUBMED Abstract]
  284. Struewing JP, Lerman C, Kase RG, et al.: Anticipated uptake and impact of genetic testing in hereditary breast and ovarian cancer families. Cancer Epidemiol Biomarkers Prev 4 (2): 169-73, 1995. [PUBMED Abstract]
  285. Jacobsen PB, Valdimarsdottier HB, Brown KL, et al.: Decision-making about genetic testing among women at familial risk for breast cancer. Psychosom Med 59 (5): 459-66, 1997 Sep-Oct. [PUBMED Abstract]
  286. Watson M, Kash KM, Homewood J, et al.: Does genetic counseling have any impact on management of breast cancer risk? Genet Test 9 (2): 167-74, 2005. [PUBMED Abstract]

Changes to This Summary (01/27/2020)

The PDQ cancer information summaries are reviewed regularly and updated as new information becomes available. This section describes the latest changes made to this summary as of the date above.
Updated statistics with estimated new breast, ovarian, and endometrial cancer cases and deaths for 2020 (cited American Cancer Society as reference 1).
Added text to state that an analysis comparing the 10-year performance of the Breast and Ovarian Analysis of Disease Incidence and Carrier Estimation Algorithm (BOADICEA), BRCAPRO, the Breast Cancer Risk Assessment Tool, and the International Breast Cancer Intervention Study (IBIS) models demonstrated superiority of the models with more detailed pedigree inclusion—specifically, BOADICEA and IBIS (cited Terry et al. as reference 131).
Updated National Comprehensive Cancer Network (NCCN) as reference 151.
Added text to state that tools are available to facilitate decision making about genetic testing at the time of diagnosis (cited Gornick et al. as reference 162).
Added text about a large study utilizing multigene (panel) testing that reported that, in addition to BRCA1/BRCA2 genes, six other breast cancer susceptibility genes were also related to a higher risk of triple-negative breast cancer (cited Shimelis et al. as reference 187).
Updated NCCN as reference 97.
Added text to state that there may be differences in ovarian cancer risk depending on the Lynch syndrome–associated pathogenic variant. In PMS2-associated Lynch syndrome, one study of 284 families was unable to identify an increased risk of ovarian cancer (cited Ten Broeke et al. as reference 329). Another prospective registry of 3,119 Lynch syndrome–pathogenic variant carriers described the cumulative risk of ovarian cancer to range from 10% to 17% in MLH1MSH2, and MSH6 carriers. In contrast, 0 of 67 women with a pathogenic variant in PMS2 developed ovarian cancer in 303 follow-up years (cited Møller et al. as reference 330). Overall, there are too few cases of PMS2 pathogenic variant carriers to make definitive recommendations for ovarian cancer management.
Updated NCCN as reference 33.
Added text to state that a cohort of 17,917 women who participated in the Prospective Family Study Cohort was assessed with 1,046 women diagnosed with breast cancer and a median follow-up of 10.7 years; no association between risk-reducing salpingo-oophorectomy and breast cancer was observed (cited Terry et al. as reference 87).
Revised text to state that, beginning at age 40 years, NCCN recommends prostate cancer screening for BRCA2 carriers and the consideration of prostate cancer screening for BRCA1 carriers.
This summary is written and maintained by the PDQ Cancer Genetics Editorial Board, which is editorially independent of NCI. The summary reflects an independent review of the literature and does not represent a policy statement of NCI or NIH. More information about summary policies and the role of the PDQ Editorial Boards in maintaining the PDQ summaries can be found on the About This PDQ Summary and PDQ® - NCI's Comprehensive Cancer Database pages.

About This PDQ Summary



Purpose of This Summary

This PDQ cancer information summary for health professionals provides comprehensive, peer-reviewed, evidence-based information about the genetics of breast and gynecologic cancers. It is intended as a resource to inform and assist clinicians who care for cancer patients. It does not provide formal guidelines or recommendations for making health care decisions.

Reviewers and Updates

This summary is reviewed regularly and updated as necessary by the PDQ Cancer Genetics Editorial Board, which is editorially independent of the National Cancer Institute (NCI). The summary reflects an independent review of the literature and does not represent a policy statement of NCI or the National Institutes of Health (NIH).
Board members review recently published articles each month to determine whether an article should:
  • be discussed at a meeting,
  • be cited with text, or
  • replace or update an existing article that is already cited.
Changes to the summaries are made through a consensus process in which Board members evaluate the strength of the evidence in the published articles and determine how the article should be included in the summary.
The lead reviewers for Genetics of Breast and Gynecologic Cancers are:
  • Doreen Agnese, MD (The Ohio State University)
  • Kathleen A. Calzone, PhD, RN, AGN-BC, FAAN (National Cancer Institute)
  • Ilana Cass, MD (Dartmouth-Hitchcock Medical Center)
  • Lee-may Chen, MD (UCSF Helen Diller Family Comprehensive Cancer Center)
  • Mary B. Daly, MD, PhD (Fox Chase Cancer Center)
  • Jennifer K. Litton, MD (University of Texas, M.D. Anderson Cancer Center)
  • Suzanne M. O'Neill, MS, PhD, CGC
  • Tuya Pal, MD, FACMG, FCCMG (Vanderbilt-Ingram Cancer Center)
  • Beth N. Peshkin, MS, CGC (Lombardi Comprehensive Cancer Center at Georgetown University Medical Center)
  • Susan K. Peterson, PhD, MPH (University of Texas, M.D. Anderson Cancer Center)
  • Mary Beth Terry, PhD (Columbia University Mailman School of Public Health)
  • Susan T. Vadaparampil, PhD, MPH (H. Lee Moffitt Cancer Center & Research Institute)
  • Catharine Wang, PhD, MSc (Boston University School of Public Health)
Any comments or questions about the summary content should be submitted to Cancer.gov through the NCI website's Email Us. Do not contact the individual Board Members with questions or comments about the summaries. Board members will not respond to individual inquiries.

Levels of Evidence

Some of the reference citations in this summary are accompanied by a level-of-evidence designation. These designations are intended to help readers assess the strength of the evidence supporting the use of specific interventions or approaches. The PDQ Cancer Genetics Editorial Board uses a formal evidence ranking system in developing its level-of-evidence designations.

Permission to Use This Summary

PDQ is a registered trademark. Although the content of PDQ documents can be used freely as text, it cannot be identified as an NCI PDQ cancer information summary unless it is presented in its entirety and is regularly updated. However, an author would be permitted to write a sentence such as “NCI’s PDQ cancer information summary about breast cancer prevention states the risks succinctly: [include excerpt from the summary].”
The preferred citation for this PDQ summary is:
PDQ® Cancer Genetics Editorial Board. PDQ Genetics of Breast and Gynecologic Cancers. Bethesda, MD: National Cancer Institute. Updated <MM/DD/YYYY>. Available at: https://www.cancer.gov/types/breast/hp/breast-ovarian-genetics-pdq. Accessed <MM/DD/YYYY>. [PMID: 26389210]
Images in this summary are used with permission of the author(s), artist, and/or publisher for use within the PDQ summaries only. Permission to use images outside the context of PDQ information must be obtained from the owner(s) and cannot be granted by the National Cancer Institute. Information about using the illustrations in this summary, along with many other cancer-related images, is available in Visuals Online, a collection of over 2,000 scientific images.

Disclaimer

The information in these summaries should not be used as a basis for insurance reimbursement determinations. More information on insurance coverage is available on Cancer.gov on the Managing Cancer Care page.

Contact Us

More information about contacting us or receiving help with the Cancer.gov website can be found on our Contact Us for Help page. Questions can also be submitted to Cancer.gov through the website’s Email Us.


  • Updated: 

No hay comentarios:

Publicar un comentario