We use cookies to enhance our website for you. Proceed if you agree to this policy or learn more about it.

  • Essay Database >
  • Essay Examples >
  • Essays Topics >
  • Essay on Health

Example Of Thesis Statement On Genetic Testing

Type of paper: Thesis Statement

Topic: Health , Testing , Genetics , Disease , Disorders , Education , Medicine , Family

Published: 03/24/2020

ORDER PAPER LIKE THIS

INTRODUCTION

Genetic testing is the complex process in which the DNA of the individual is analysed for concerned traits. It helps in detecting certain diseases which are inherited from either of the parents. Many people are now choosing to undergo these tests as this helps them to make them aware of many dangerous genetic disorders like Hemophilia, Cystic Fibrosis, Klinefelter Syndrome, Down syndrome, Turner Syndrome, Sickle-Cell disease, etc. There are many methods of genetic testing. For example, in buccal smear, a small brush is used to collect the samples of the cells present inside the surface of cheek. Other methods include taking samples from blood, saliva, hair, skin, amniotic fluid, etc. All the methods have their own indications. When the samples are received by the laboratory, these are tested depending on the disorder which needs to be detected. Many methods are employed to process these samples, namely, DNA sequencing, Microarrays (to detect mutations in pathologies like breast cancer). After performing all the desired tests, the reports are delivered to the patients and then they are systematically counseled about the preventive measures they need to take as per the results.

There are many advantages from genetic testing. The most important advantage is the detection of many potential genetic disorders and thus the patients can therefore organize their regular medical checkups that need to be done so as to keep the track of the genetic disorders they are suffering from. This allows them to prevent that particular disorder from advancing and therefore it helps in preventing these dangerous diseases. On the other hand, it also prevents these diseases to pass to the next generation. Prenatal genetic testing ensures that many genetic disorders are detected beforehand and thus the appropriate actions can be taken to prevent these disorders. Many types of Cancers can also be detected and immediate actions can be taken. Preventive surgeries can be planned to decrease the incidence and mortality caused due to Cancer. Sometimes, certain persons do not show any sign or symptom of the particular disease running in their family but those mutated genes are always present in the dormant form. Such persons are known as “CARRIERS” and the persons with the active mutated genes are commonly known as “SUFFERERS”. The carriers often feel relaxed or are less tensed owing to the fact that they do not notice any sign or symptom of that particular disease but there are very good chances that they can pass these ill fated genes to their young ones. And these newborns will then have to pay the price. The genetic testing thus will ensure that these genes are detected and necessary steps are implemented.

DISADVANTAGES

As each and every coin has two faces, similarly Genetic testing also comes with both, advantages and disadvantages. The most important disadvantage is the distress that may be caused among the family members due to detection of certain altered genes. Family relations may get disrupted in certain cases like if a certain individual is not at all interested in undergoing a certain genetic test, then, he/she may face difficult time as there can be the presence of immense pressure from the other family members. And if the test comes out to be positive, then, some persons may feel very guilty of themselves because they will think that they are the only reason for the sufferings of their young ones. On the other hand, if the genetic test results come out to be negative, then this can also be a very big emotional calamity for some individuals as these people may undergo the guilt of surviving the risk of having the fatally dangerous disease but their siblings turn out to be positive. Another important drawback is the limited answers about some inherited diseases (like colon cancer) form the tests. For example, if the test turns out to be positive, it will only mean that the concerned disease may happen. And if the test is negative, that also do not mean that one cannot have that disease ever. If the predisposing or secondary factors are present, then that certain disease can occur.

So, after discussing the positive and negative aspects of the genetic testing, it becomes quite clear that, Genetic Testing has more good hidden in it, than the bad, as negative aspects mainly include only emotional or social or financial issues and on the contrast, the positive aspects are helpful enough to outweigh these disadvantages. Genetic testing, obviously, is the most practical and sincerest approach to one’s well being. These tests should be implemented as much as possible in the lives of people as this will make them aware of the potential dangers coming ahead of them. And then after knowing about these potential dangers, they can prepare themselves for a healthier future. This can also help certain couples with their decisions to have children so they can conduct prenatal genetic tests so to ensure the safety of their young ones.

double-banner

Cite this page

Share with friends using:

Removal Request

Removal Request

Finished papers: 690

This paper is created by writer with

ID 253537933

If you want your paper to be:

Well-researched, fact-checked, and accurate

Original, fresh, based on current data

Eloquently written and immaculately formatted

275 words = 1 page double-spaced

submit your paper

Get your papers done by pros!

Other Pages

Politics thesis proposals, communication thesis proposals, relationships thesis proposals, investment thesis proposals, children thesis proposals, teaching thesis proposals, computers thesis proposals, crack research proposals, sensor research proposals, cook research proposals, recruit research proposals, controversy research proposals, graphic research proposals, tourist research proposals, advert research proposals, volleyball college essays, example of card counting research paper, free essay on organizational theory and organizational effectiveness, example of essay on the higher education of women in america prior to 1860, visit to norton simon museum essay sample, essay on the criminal justice system, accounting for decision making case study sample, example of essay on paper chromatography experiment report, the case against marijuana legalization essay sample, essay on the implications of class and power in, prospect and rational expectation theory course work examples, and give order to order entry research paper, five guys burgers and fries research paper, free essay on ethics paper, cooperative security research paper examples, motility essays, gateway drug essays, polygamist essays, fleet essays, bad weather essays, death rate essays, implantation essays, pragmatics essays, flight deck essays, iteration essays, transient essays, visual aspect essays, transvaal essays.

Password recovery email has been sent to [email protected]

Use your new password to log in

You are not register!

By clicking Register, you agree to our Terms of Service and that you have read our Privacy Policy .

Now you can download documents directly to your device!

Check your email! An email with your password has already been sent to you! Now you can download documents directly to your device.

or Use the QR code to Save this Paper to Your Phone

The sample is NOT original!

Short on a deadline?

Don't waste time. Get help with 11% off using code - GETWOWED

No, thanks! I'm fine with missing my deadline

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here .

Loading metrics

Open Access

Peer-reviewed

Research Article

Attitudes towards genetic testing: The role of genetic literacy, motivated cognition, and socio-demographic characteristics

Contributed equally to this work with: Maxim Likhanov, Ilya Zakharov

Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Visualization, Writing – original draft

Affiliation State Key Laboratory of Cognitive Neuroscience and Learning, Beijing Normal University, Beijing, China

Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Resources, Visualization, Writing – original draft

Affiliations Ural Federal University Named after the First President of Russia B. N. Yeltsin, Yekaterinburg, Russia, Psychological Institute of Russian Academy of Education, Moscow, Russia

Roles Investigation, Writing – review & editing

Affiliation Department of Biological Sciences, Tai Solarin University of Education, Ijebu-Ode, Nigeria

ORCID logo

Roles Data curation, Writing – review & editing

Roles Writing – review & editing

Affiliation Department of Psychology, Goldsmiths, University of London, London, United Kingdom

Roles Conceptualization, Writing – review & editing

Roles Conceptualization, Data curation, Project administration, Writing – review & editing

* E-mail: [email protected]

  • Maxim Likhanov, 
  • Ilya Zakharov, 
  • Adeyemi Awofala, 
  • Olusegun Ogundele, 
  • Fatos Selita, 
  • Yulia Kovas, 
  • Robert Chapman

PLOS

  • Published: November 15, 2023
  • https://doi.org/10.1371/journal.pone.0293187
  • Reader Comments

Fig 1

Understanding reasons for why people choose to have or not to have a genetic test is essential given the ever-increasing use of genetic technologies in everyday life. The present study explored the multiple drivers of people’s attitudes towards genetic testing. Using the International Genetic Literacy and Attitudes Survey (iGLAS), we collected data on: (1) willingness to undergo testing; (2) genetic literacy; (3) motivated cognition; and (4) demographic and cultural characteristics. The 37 variables were explored in the largest to-date sample of 4311 participants from diverse demographic and cultural backgrounds. The results showed that 82% of participants were willing to undergo genetic testing for improved treatment; and over 73%—for research. The 35 predictor variables together explained only a small proportion of variance: 7%—in the willingness to test for Treatment; and 6%—for Research. The strongest predictors of willingness to undergo genetic testing were genetic knowledge and deterministic beliefs. Concerns about data misuse and about finding out unwanted health-related information were weakly negatively associated with willingness to undergo genetic testing. We also found some differences in factors linked to attitudes towards genetic testing across the countries included in this study. Our study demonstrates that decision-making regarding genetic testing is influenced by a large number of potentially interacting factors. Further research into these factors may help consumers to make decisions regarding genetic testing that are right for their specific circumstances.

Citation: Likhanov M, Zakharov I, Awofala A, Ogundele O, Selita F, Kovas Y, et al. (2023) Attitudes towards genetic testing: The role of genetic literacy, motivated cognition, and socio-demographic characteristics. PLoS ONE 18(11): e0293187. https://doi.org/10.1371/journal.pone.0293187

Editor: Muhammad Shoaib Akhtar, School of Medicine, University of California-Davis, UNITED STATES

Received: April 25, 2023; Accepted: September 26, 2023; Published: November 15, 2023

Copyright: © 2023 Likhanov et al. This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Data Availability: The script used for analyses is available on osf ( https://osf.io/nex76/ ) (DOI: 10.17605/OSF.IO/NEX76 ).

Funding: The author(s) received no specific funding for this work.

Competing interests: The authors have declared that no competing interests exist.

Introduction

Since the completion of the Human Genome Project in 2003 humanity has entered the Era of the Genome [ 1 ]. This new period is associated with the extensive use and development of genetic technologies, including genetic testing. For example, many specialists in the area predict that by 2030 in some countries the DNA of every newborn will be sequenced at birth [ 2 ]. This technology already exists and is becoming increasingly inexpensive: costing about $1000 for each full human genome sequence in 2020 [ 3 ], about $600 in 2023, and an anticipated $200 soon [ 4 ]. In addition, the time needed for genome sequencing has dramatically reduced from many days to around 7 hours [ 3 , 5 ]. These trends mean that genetic testing, including direct-to-consumer, is becoming increasingly accessible.

Widespread availability of genetic testing may speed up the development of personalized medicine—optimized prophylactic or therapeutic solutions based on individuals’ genetic make-up [ 6 ], see e.g. a Precision Medicine Initiative [ 7 ]. Beyond medicine, genetic information can be applied in many contexts, including sports, education and the justice system [ 8 , 9 ]. A growing body of research has begun to examine potential benefits and risks associated with diagnostic and predictive genetic testing [ 10 – 13 ].

Willingness to undergo genetic testing

Relatedly, research has begun to examine people’s attitudes and views about genetic testing, indicating a trend from resistance to greater acceptance (see e.g. [ 14 ]). Several recent studies indicate that most people accept genetic testing for medical purposes. For example, in 2010 one study showed that 85% of 2000 respondents from a Russian urban population expressed positivity towards undergoing predictive genetic testing for preventable health conditions [ 15 ]. Similar results were found in another recent study, with almost 90% of participants from general populations in the UK, the USA and Russia expressing willingness to undergo genetic testing for improved treatment [ 16 ]. Another study with Romanian justice stakeholders, found that most judges and lawyers in the study expressed willingness to undergo genetic testing for medical purposes [ 17 ]. Yet another study showed that 71% of participants from a representative sample of 837 adult Qataris were willing to undergo genetic testing [ 18 ]. A high endorsement was also found in a large sample of 1500 Korean individuals from general public, 1500 cancer patients, 113 clinicians, and 413 researchers, with the majority of participants being positive towards genetic testing (from 88.5% among clinicians to 94.3% among patients [ 19 ]. A somewhat lower endorsement (63.8%) was shown by a study of Chinese individuals at high risk of breast cancer [ 20 ].

Research has also indicated that willingness to test for medical purposes is higher for treatable conditions and conditions with a clear family transmission patterns [ 17 , 21 , 22 ], and can be quite low in other cases [ 23 ]. Findings regarding attitudes towards genetic testing in non-medical contexts are much more mixed. For example, in one study, most legal professionals expressed willingness to provide DNA samples for research purposes [ 17 ]. In contrast, a study with university students in the USA found that only 11% were willing to donate DNA to research without reward, increasing to 50%—for payment [ 24 ]. People’s willingness to undergo genetic testing for other purposes is largely unexplored, including for family planning, career planning [e.g. taking up professional sports], and insurance decisions. Available literature suggests wide variability in such views [ 25 – 28 ].

Genetic knowledge and willingness to undergo testing

Understanding factors that shape attitudes towards genetic testing is an important agenda for the Genomic Era. Several studies found that people’s views on genetic testing are related to their genetic literacy [ 18 , 29 , 30 ]. For example, one recent study, with more than 5400 participants from several countries, showed that willingness to undergo genetic testing was positively correlated with genetic knowledge (B = .18; [ 16 ]). The same study also found that the general population has relatively low genetic knowledge and held some common striking misconceptions. For example, only two thirds (68%) of participants were aware of the polygenic (many genes involved) nature of complex traits (autism spectrum disorder and schizophrenia). This finding is in line with several recent studies that showed quite low knowledge in different samples, including pharmacy students [ 21 , 31 ]. Low knowledge of basic genetic concepts might lead to under-informed choices in relation to undergoing genetic testing, receiving consultation on genetic-related matters, or using off-the-shelf genetic services.

Genetic knowledge, however, does not explain much of the variance in people’s decision regarding genetic testing, as evidenced in the low correlation between them (e.g., [ 16 ]). This is because people are not passive recipients of scientific knowledge regarding genetics, but rather engage with it in a motivated fashion (see [ 32 ], for review). In other words, people’s judgments about genetic testing are based on rational considerations (’cognition’), as well as on their beliefs, attitudes and values (’motivated cognition’) [ 33 ]. It has been shown that motivated cognition factors can affect legal judgments in courts [ 33 ], political judgments [ 34 , 35 ], and the use and interpretation of empirical research itself [ 36 ]. In the case of genetic testing, considerations can include risks related to access to health insurance and privacy, suspicions about hidden political/economic agenda behind genetic studies, concerns about misuses of genetic information, etc. (e.g. [ 16 , 37 ]).

Motivated cognition and willingness to undergo testing

Motivated cognitions relevant to genetic testing can also include beliefs regarding the malleability of different traits and an individual’s control over them. This is because such beliefs may influence one’s evaluations of usefulness of genetic information. For example, people’s beliefs regarding ‘free will’ have been linked to such phenomena as pursuit of self-directed goals, level of prosocial and aggressive behavior, autonomy and conformity, self-efficacy and perceived capacity [ 38 ]. In public health, deterministic beliefs were shown to impact perceptions of disease risk and inclination to engage in medical evaluations, prophylactics, and treatments (see [ 39 ] for review). For example, participants’ perception of a condition as being genetic was linked to: greater expectations on the effectiveness of genetic testing and related technologies [ 40 ]; reduced optimism for treatment and more willingness to seek medically intensive treatments [ 41 , 42 ]; and decreased efforts to manage diseases such as diabetes with lifestyle changes [ 43 ]. It is possible that people who believe they have a conscious control of their behavior (’free will’), may be more willing to undergo genetic testing—so that they can act on this information by introducing changes to lifestyle and other prophylactics. For example, research has shown that people who knew that type 2 diabetes is preventable by means of life style changes had higher inclination to undergo genetic testing [ 44 ].

The attitudes towards genetic testing may also depend on whether people believe that genes are involved in traits (see [ 45 ] for a thorough discussion). Recent advances in behavioral genetics have shown at least moderate effects of genes on practically all human traits [ 46 , 47 ]. However, many people still hold misconceptions, such as believing that genes are not important for human behaviour, that genes are important only for some traits, or that only genes are important for behaviour. For example, one study has shown that 25% of participants believed that their destiny is written in their genes [ 16 ].

Demographic characteristics and willingness to undergo testing

Research has also shown that attitudes towards genetic testing are influenced by demographic characteristics. For example, some studies show that younger people have on average a more positive attitude towards genetic testing and demonstrated higher interest in it [ 15 , 48 , 49 ]. In addition, males demonstrated slightly more positive attitudes towards testing [ 15 , 50 ]. Attitudes towards genetics can also be culturally informed [ 51 ]. Some cross-cultural differences in attitudes towards genetics have been found, including in views regarding general moral issues of self- and society-responsibility [ 52 ]; and specific questions of prenatal diagnostics and reproductive technologies [ 53 ]. For example, one recent study [ 54 ] has shown cross-country differences in endorsement of genetic testing before pregnancy by medical students: 96% of participants in Israel and 40% in Croatia disagreed with the idea that "Screening for reproductive risks in prospective parents is wrong". Another study found differences in awareness and attitudes concerning genetic testing across regions within one country (USA), suggesting some influences from local sociocultural environments [ 55 ]. Yet another study found that African Americans, compared with White Americans, anticipated fewer negative consequences of genetic testing identifying potential health-related problems [ 56 ]. A cross-cultural approach towards genetic testing and attitudes towards it allows to shed light on cultural/historical specifics that potentially might affects willingness to undergo testing. For example, consanguinity is common in south Asian and Arab societies (e.g. in Pakistan prevalence of consanguinity has been reported as high as 80% due to marriages within caste groups), which increases risk of genetic diseases (e.g. colour vision impairment [ 57 ] or β-thalassemia [ 58 ]). This may increase willingness to undergo genetic testing in these populations.

The interrelations between genetic knowledge, concerns about genetics, demographic characteristics, and willingness to undergo testing appear to be complex. For example, several studies reported gender differences in genetic knowledge, but results differed depending on the sample. In some studies, lower genetic knowledge levels were found in males [ 59 , 60 ]; whereas other studies found lower levels in females [ 16 , 61 ]. Moreover, some research suggested greater knowledge about biotechnology is associated with lower pessimism about biotechnology for men and with greater pessimism for women [ 61 ]. The age-related effects on genetics knowledge are also mixed. Numerous studies found that higher levels of genetic knowledge, especially among young adults with higher education levels, are associated with more favorable attitudes towards genetic testing, e.g. for chronic disease [ 62 , 63 ]. However, this association was not found in some studies [ 48 , 50 ]. For example, in one study, people aged 45–60 years and with less education showed most interest in genetic testing for heart disease [ 64 ].

Moreover, greater genetic knowledge may result in more concerns in some contexts. For example, in one study, women who disagreed with statements like “genetic information should be used to enable parents to choose physical and mental characteristics of their children” had higher genetic knowledge [ 65 ]. One qualitative study with 4 focus groups showed that after an open discussion of potential positive and negative implications of predictive genetic testing, a quarter of the participants initially interested in having a test changed their mind [ 66 ]. Another study [ 67 ] showed that well-informed participants had more critical attitudes towards morally or socially sensitive applications of genetics (e.g., genetic engineering). Yet another study showed that results of genetic testing and information provided after it might affect attitudes towards testing, with 91.6% of patients with negative results of BRCA 1/2 gene test and 100% of patients with positive results were willing to recommend family members to participate in the cancer screening program [ 20 ].

The current study

Gaining further insights into sources of attitudes towards genetic testing will help people to make informed decisions regarding genetic testing in specific circumstances [ 32 ]. The present study explores the multiple drivers of people’s attitudes towards genetic testing, as well as interrelations among these factors. Using the International Genetic Literacy and Attitudes Survey (iGLAS) [ 68 ], we collected data on: 1. willingness to undergo testing (for improved treatment and for research); 2. genetic literacy—knowledge about the basic principles and current state of genetics; 3. motivated cognition—personal concerns about the benefits of testing and future applications of data, general trust in governmental research institutions or insurance companies; and 4. demographic and cultural characteristics—gender, age, level of education, country of residence, country of secondary education, occupation, and religiosity. These variables are explored in the largest to-date sample (N = 4311) of participants from diverse demographic and cultural backgrounds.

Participants

The total sample included 5238 participants from 86 different countries. Participants had to be 18 or older, with no upper age limit. Eighty-six percent of the participants had either completed or were working towards an undergraduate degree or higher. The number of participants varied across different analyses, as not all participants answered all questions. After all data exclusions, including outliers deletion, and list-wise deletion (needed for factor analysis of Opinions), data from 4311 participants were analyzed. As most of participants came from Russia, Nigeria, USA and UK (more than 295 participants in each; See Fig 1 ), we run an additional analysis for these 4 countries, exploring differences in genetic testing attitudes across them. Other countries had smaller samples, with 6 countries having a sample size from 48 to 256 participants; 35 countries–from 2 to 39, and 15 countries–only 1 participant (see S1 Table for Ns of individual countries)

thumbnail

  • PPT PowerPoint slide
  • PNG larger image
  • TIFF original image

Participants were reached through social media and Reddit AMA. Targeted collections were also carried out through higher education institutions in Nigeria, Russia, and the UK. Participants were recruited on a rolling basis over a period of 3 years, largely before Covid-19 pandemic. The targeted collections also happened before the pandemic.

https://doi.org/10.1371/journal.pone.0293187.g001

We used the International Genetic Literacy and Attitudes Survey (iGLAS). The current version of iGLAS is available in 9 languages (Albanian, English, French, Italian, Japanese, Persian, Romanian, Russian, and Spanish). A short sample of the English language version of iGLAS can be found at https://goldpsych.eu.qualtrics.com/jfe/form/SV_9zOfCcGhht7qwy9 . Information on the validation of iGLAS can be found in [ 68 ]. The latest version of iGLAS can be found at http://tagc.world/iglas/ .

In the present study we used 2 items to explore willingness to undergo genetic testing and 37 items—as potential predictors of the willingness to undergo genetic testing. The predictors represented three groups of potential sources of individual differences: 1) genetic literacy; 2) motivated cognition; and 3) demographic characteristics.

Willingness to undergo genetic testing . Two questions explored participants’ willingness to undergo genetic testing, using 7-point Likert scales (Response options: 1- “Very unlikely”, 2 - “Unlikely”, 3 - “Somewhat unlikely”, 4 - “Neutral”, 5 - “Somewhat likely”, 6 - “Likely”, and 7 –“Very likely”):

  • “ Would you take a genetic test if it allowed you to have improved treatment (for example , medication with fewer side effects) ” (Henceforth Test for treatment );
  • “ Would you be willing to give a sample of your DNA for scientific research if your data are stored anonymously ?” (Henceforth Test for science ).

Predictors.

  • Genetic literacy was measured with 20 items. The questions were developed to assess a basic functional level of genetic knowledge. An example item with 4 response options being: “ Which of the following 4 letter groups represent the base units of DNA : 1) GHPO; 2) HTPR; 3) GCTA; and 4) LFWE ?”. More details about the genetic knowledge items can be found in [ 16 ]. In the present study the total Genetic Knowledge score was obtained by summing up correct responses for each item.
  • Motivated cognition was assessed with 10 items. 8 of the items tapped into participants’ concerns about genetic testing. Participants were asked to answer whether they have any of the 8 concerns in response to the following question: “ In deciding whether to take a genetic test , which of the considerations below apply to you” (See Table 1 ). Multiple responses could be chosen. There was an additional– 9th, free text option in this question: “Other”. More than 95% of the participants did not state any “other” consideration, thus, this question was not included into further analysis.

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t001

Prior to the main analysis, a factor analysis was conducted on 8 concern items. Scree plot and the eigenvalues have shown that there are two concerns factors. Exploratory factor analysis with 2 factors and Oblimin rotation have shown that Q1, Q2, Q7 and Q8 loaded on a single Data security factor; and Q3, Q5 and Q6 loaded onto Health issues factor (see Table 1 ). Q4 did not load on any of the factors and was also excluded. Data security and Health issues variables were used as predictors in further analysis.

1 item tapped into views on genetic influences on behavior: “Destiny is written in our genes” ( Destiny is written ). 1 item tapped into views on potential data usage violations Mistrust in research : “I do not trust research institutions in my country because they might misuse data obtained from participants”. A 7-level Likert scale (from “Strongly disagree” to “Strongly agree”) was used for both questions.

3. Demographics included the following 5 characteristics: Gender, Age, Country of residence, Level of education (on a self-report 7-point scale from “no school certification” through to “post-doctoral”), and level of Religiosity (on a self-report 10-point scale, varying from “Not at all” to “Very religious”).

The study was approved by the Goldsmiths Department of Psychology Ethics Committee (ref: PS101016RCS). All the analyses presented in this paper were based on data accessed after ethical approval for the study was provided by the ethics committee of the psychology department at Goldsmiths, University of London on 10/10/2016. Participants answered iGLAS items online, in their own time and place; or at their University. The data were fully anonymous. Informed consent was implemented at the beginning of the survey. Participants were asked to accept the following statements: My participation in this study is voluntary; I am over 18; I may withdraw from this research at any time and for any reason; I may omit any questions I do not wish to answer; All data will be treated with full confidentiality and, if published, it will not be identifiable as mine. For those who did not endorse any of the statements, the survey discontinued.

Statistical approach.

The data were analyzed with R language for statistical programming and R studio platform (R Core Team, 2017). The regression and correlation (Pearson) analyses were performed with in-built R functions “lm” and “cor”, the factor analysis was performed with the ‘psych’ package ( https://personality-project.org/r ). Significance level was set to p < 0.05. Mixed-effect modelling was performed with lme4 package [ 69 ]. Compliance with normality and linearity assumptions was ensured for all analysis. Individual answers that exceeded a ± 3 SD level were considered outliers and deleted. Significance level was set at p < .05 threshold. The R code used for all analysis is available at: https://osf.io/nex76/

4.1 Descriptive statistics for the full sample

The descriptive statistics for the study variables are presented in Table 2 . 2545 females (59.03%) and 1606 males participated in the study. 30 participants identified as being non-binary, 29 participants decided not to disclose their gender and 101 participants did not provide information on their gender. Only data from male and female groups were used in regression analyses, as the other groups were too small. Most of the sample were working towards or had at least an undergraduate degree (70.5%) at the time of completing iGLAS, with extra 8.7% working towards or holding an MSc and 6.9%—a PhD degree. Age for some participants exceeded +3 SD threshold (~ 62.09), however we decided to include them in further analysis too. Frequencies for Country of residence can be found in SOM Table 1 .

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t002

4.2 Willingness to undergo genetic testing

As can be seen from Table 2 , the mean response for both willingness items was above 5 (out of 7), indicating considerable willingness to undergo genetic testing. A paired sample t-test showed that willingness to undergo Test for treatment was slightly higher than Test for science (Cohen’s d equals 0.21).

We also computed percentages for each response option, showing that 81.64% of the respondents chose “Very likely”, “Likely”, and “Somewhat likely” for Test for treatment and 72.78%—for Test for science questions.

4.3 Associations among the study variables

Zero-order correlations (Pearson) among all study variables are presented in Table 3 . The two outcome variables were moderately correlated (r = .42), suggesting that participants who were willing to undergo genetic testing for improved medical treatment were also on average more likely to undergo genetic testing for research purposes.

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t003

Willingness to undergo genetic testing in both contexts was positively correlated with Genetic knowledge (r = ~.20), as well as with Age (r = ~.10), Education level (r = ~.06) and deterministic views ( Destiny is written ; r = ~.10); and negatively correlated with Religiosity (r = ~.11) and Mistrust in research (r = ~.08). Data security and Health issues concerns showed only negligible correlations with willingness to undergo testing. Educational level did not correlate with willingness to undergo Test for treatment and only negligibly with Test for science .

4.4 Statistical predictors of willingness to undergo genetic testing in the full sample

We tested two step-wise regression models, exploring predictors of willingness to undergo Test for treatment and Test for science . Genetic knowledge , Destiny is written, Religiosity, Mistrust in Research and Gender were positively associated with Test for treatment , together explaining approximately 7% of the variance. The same set of significant predictors were shown for willingness to undergo Test for science , together explaining 6.3% of the variance ( Table 4 ).

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t004

To control for effect of Country of residence , we ran a similar model (predicting both outcomes from all study predictors) but with a random intercept for Country of residence –random slope model did not fit because of small sample sizes for some countries (15 countries with N = 1). The mixed model analysis showed a result similar to stepwise regression (i.e. same predictors being significant), explaining 6.2% in Test for treatment with extra 0.3% when random effects were accounted for. For Test for science the mixed model analysis showed an almost similar result, with fixed effects of Genetic Knowledge, Destiny is written, Religiosity, Mistrust in research and Gender being significant. The only extra predictor was Data security (with very small effect– 0.02). This set of predictors explained 7.1% in Test for treatment with no extra addition from random effects. Extra details on the analysis are available from authors on request.

4.5 Country differences in willingness to undergo genetic testing

We next examined potential cross-cultural differences, re-running the analyses for four countries with the biggest samples separately: participants from Russia (N = 1358), Nigeria (N = 1020), the USA (N = 340) and the UK (N = 295). Table 5 presents descriptive statistics and ANOVA results for all variables divided by country. Post-hoc comparisons for all variables are available in S2 Table in SOM.

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t005

Our data showed that frequencies for willingness to undergo testing were quite different across the 4 countries, with small effect. For Test for treatment , the highest endorsement of testing (participants opting for “Very likely”, “Likely”, and “Somewhat likely”) was shown for UK (87.51%) and USA (86.46%), followed by Russia (78.69%) and Nigeria (74.11%). The pattern was the same for the overall endorsement of Test for science , with the highest endorsement in the UK (82.42%) and the USA (81.18%), followed by Russia (67.36%) and Nigeria (67.48%).

Of all predictors, three showed substantial average differences across the samples: for Age , UK and USA participants were considerably older than those from Nigeria and Russia; the overall Genetic knowledge was higher in UK and USA samples, compared to those from Nigeria and Russia; and Religiosity was higher in Nigeria than in the other 3 countries. There were some differences across the 4 countries in the Educational level, with Educational level for Russia being lower when in Nigeria, USA and UK; and no differences across the three (See S2 Table ).

4.6 Statistical predictors of willingness to undergo genetic testing by country

The results of the separate step-wise regression analysis for each group are presented in Tables 6 and 7 .

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t006

thumbnail

https://doi.org/10.1371/journal.pone.0293187.t007

None of the predictors of Test for treatment replicated across all 4 samples. Two predictors were significant in 3 samples (excluding UK): level of Genetic knowledge and the deterministic views ( Destiny is written ). For other predictors, there were differences across the countries. For example, greater Mistrust in research institutions and Health issues (greater concern regarding medical information) were moderately associated with less willingness to undergo genetic testing for treatment ( Test for treatment ) only in the UK sample. More concerns regarding Data security were associated with less willingness to undertake Test for treatment in the USA sample, and, surprisingly, with more willingness in the Nigerian sample. Greater religiosity was associated with less willingness to undergo Test for treatment only in the Russian sample. Overall, more variance was explained by the predictors in the UK and the USA (17 and 13%), than in Nigeria and Russia (~6%).

For the Test for science , the only significant predictor for all countries was Genetic Knowledge . Deterministic views ( Destiny is written ) was positively linked to Test for science in three countries (except the UK). Mistrust in research was negatively linked to Test for science in the three countries (except Nigeria). The pattern of results was somewhat different for other predictors. For example, Data security was negatively associated with Test for science only in the USA sample. Health concerns were negatively associated with willingness to Test for science in the UK; and positively–in Russia. Religiosity was negatively associated with Test for science only in the Russian sample. Overall, more variance was explained by the predictors in the USA and the UK (13 and 14%), than in Nigeria and Russia (2 and 5%).

The current study explored attitudes towards genetic testing and the sources of individual differences in such attitudes in the largest to-date sample of participants from diverse demographic and cultural backgrounds.

Our data showed that 82% of the 4311 participants were willing to undergo genetic testing for improved treatment; and 73%—for research. This finding is in line with previously shown positive attitudes towards genetic testing in samples from different countries [ 15 , 16 , 51 , 70 – 72 ]. Consistent with previous findings, participants were more willing to test for treatment than for science. One explanation for higher endorsement of testing for improved treatment might be that people perceive testing for improved treatment as beneficial for them and their relatives (“benevolent”); compared with “abstract greater good” (“altruistic”) testing for science. A similar pattern of results was shown in a study that investigated motivation to donate blood [ 73 ]. Blood donors were more willing to donate blood when exposed to a benevolent message (selfish + societal benefits) rather than an altruistic one (societal benefits only). The differences may also reflect a preference for immediate benefit (improved treatment) over a “delayed gratification” [ 74 ]–willingness to contribute to a later reward (potential better diagnostics and treatments in the future). However, differences in percentages between the two questions are quite small, suggesting that people view both testing options as potentially useful. Interestingly, the correlation between the two purposes of testing was only moderate [ 42 ], suggesting that many people endorse only one of the contexts and not another, or endorse them to a different extent. As a whole, these results highlight complexities of the processes underlying public understanding of genetic science and reasons for why people opt for genetic testing, which may be of interest to stakeholders and policymakers.

Analysis of willingness to undergo testing in separate countries with the largest samples in the current study (Russia, Nigeria, the USA and the UK) showed lower willingness in Russia and Nigeria. Several sample characteristics could have contributed to the observed differences. For example, USA and UK samples were on average older (Partial eta squared equals .52) and had higher genetic literacy (.33). In addition, USA and UK have higher GDP [ 75 ], which may also be relevant in terms of genetic testing advancements and accessibility. Differences in educational systems, mass-media coverage of genetic discoveries and overall scientific literacy levels across countries may all also contribute to the differences in genetic testing attitudes and warrant further investigation.

This study also explored factors that might explain individual differences in people’s willingness to undergo genetic testing in the overall sample and within the four countries. We examined genetic knowledge, motivated cognition (beliefs and concerns), and demographic characteristics—as potential predictors. Overall, the investigated factors explained little variance in willingness to undergo genetic testing: around 7% in testing for treatment and 6% for research. The effects of individual predictors were weak, with genetic knowledge being the strongest positive predictor for both treatment and research. Contribution of genetic knowledge to willingness to undergo testing is expected and in line with previous research [ 16 , 18 ]. Our results support suggestions that a lack of knowledge is one of the most important barriers to the acceptance of genetic technologies [ 76 ], potentially because it prevents comprehension of potential testing results [ 77 ].

On the other hand, genetic knowledge explained only a little variance in willingness to undergo genetic testing, which indicates the complex nature of this association. Some research suggested that greater genetic literacy is associated with greater reported fear and uncertainty about the implications of genetic testing; and, in contrast, lack of understanding and unreasonable expectations may lead to unwarranted enthusiasm about testing and unrealistic hopes about its results [ 65 , 78 ]. Our data showed that, on average, participants correctly answered 60.4% of the genetic knowledge questions, which indicates insufficient knowledge given that the questionnaire tapped into basic knowledge regarding genetics. More nuanced research is needed to understand whether greater genetic knowledge may allow people to assess pros and cons of testing in every situation and undergo testing when it can actually be beneficial to them.

In line with our prediction, the belief that ’one’s destiny is written in one’s genes’ was positively associated with willingness to undergo genetic testing (.08-.12). Given that genetic testing can provide probabilistic prediction of future life outcomes (e.g. educational achievement; [ 79 ]), it seems natural for people who believe that genes are important for different traits and behaviours to want to know what is ’written there’ and potentially to alter ‘destiny’. People can introduce various changes to their lifestyles to ameliorate the potential negative effects of genes on their behaviour and health. For example, they can choose to get some extra reading classes and other educational interventions for a child, if a predisposition to reading problems is identified via genetic testing [ 80 ]. In the future, gene editing advances may also enable prevention of diseases or improved treatments, with the first cases of genetic therapies already at the clinical trials stage [ 81 – 83 ].

Further analysis of predictors for willingness to undergo genetic testing in UK, USA, Nigeria and Russia separately showed both similarities and differences across the four countries. Genetic knowledge and belief in ‘genetic destiny’ predicted willingness to test for treatment in 3 of the 4 countries. The absence of these effects in the UK can be explained by specific characteristics of the UK sample in this study. For example, this sample was the oldest out of the four (Mage = 43.73) and had the highest average education level (4.37). However, the overall effect of Level of education was small (eta squared equal to .09). Moreover, the UK participants showed the highest overall willingness to undergo test for treatment and the narrowest distribution of scores around the mean. This restricted variance may have led to the observed lack of prediction.

Several predictors that were not present in the overall sample emerged in the analysis of data for separate countries. For Test for treatment, data security was a significant predictor in Nigeria and the USA; while health-related concerns—in Russia and the UK. One of the strongest predictors in the UK sample was mistrust of research institutions (-.33), which is in line with some previous studies [ 44 ]. Such mistrust may stem both from people’s personal experiences, as well as awareness of infamous violations of research ethics in previous studies. For example, studies that might have negatively affected attitudes towards research institutions may include the “Tuskegee Syphilis Study”, “Stanford Prison experiment”, and many other across the world (e.g. [ 84 ]). Beyond mistrust of research institutions, people’s mistrust of various kinds of official institutions (police, courts, the government) may influence the decisions to undergo testing [ 85 ]. Quite a similar pattern emerged for the willingness to test for science, with genetic knowledge being a significant predictor for all countries; and belief in ‘genetic destiny’–for three out of four countries (except UK). Mistrust of research institutions was also a significant predictor in three out of four countries (except Nigeria).

Data security concern was a significant negative predictor of willingness to test for science only in the USA. This negative effect on a decision to undergo genetic testing is expected, as recently there have been many public outcries regarding data protection and storage, including public cases of companies selling genomic data for profit. For example, 23andMe sold the data of participants to GlaxoSmithKline for $USD300 million [ 86 ]. Given that genomic data can reveal a person’s life history, ancestry and genetic predispositions, using such data poses serious ethical, social and legal concerns, including in the sectors of employment, health insurance and justice [ 87 , 88 ]. A potential genomic data breach might cause serious harm, including discrimination in health insurance [ 89 ]. It is not clear why Data security concerns were not a significant predictor in the other samples, as data security risks are well known. For example, one recent study that analyzed the NHS Digital audits found that in the past year 33 UK organisations, including GlaxoSmithKline and Imperial College London, were audited and every one had breached data sharing agreements, with hundreds more inspected and found in breach since audits began in 2015 [ 90 ]. In our study, fewer data security concerns were reported by participants from Russia, Nigeria and UK, compared with the USA (who also showed the highest genetic knowledge). It might be that data security concerns are not linked to willingness to undergo testing in Russia, Nigeria and UK as participants are underinformed regarding consequences, including negative ones, of genetic testing. This explanation is in line with one previous study that showed well-informed participants had more critical attitudes towards applications of genetics [ 67 ]; and is indirectly supported by a relatively high correlation between genetic knowledge and data security concerns (.24). Further research is needed to investigate complex interactions across knowledge, concerns, and attitudes towards testing.

Contrary to the pattern obtained for the whole sample, health-related concerns were a significant negative predictor in the UK; and a significant positive predictor in Russia for testing for treatment and science. A negative effect of health-related concerns is in line with previous studies that showed people worrying about gene-based health information. For example, US citizens were found to be afraid that genetic test results could be used to discriminate against those with a genetic predisposition for illness [ 37 ]; or that genetic testing could result in people being labeled as having “good” or “bad” genes [ 76 ]. Previous studies also found negative associations between attitudes towards genetic testing and fear that genetic information might encourage lifestyle changes that are difficult to sustain [ 91 ]. In addition, several previous studies showed higher stigma being associated with those diseases that were considered genetic [ 41 , 42 ]. However, other studies suggest that strong negative long-term emotional responses to test results are rare and that people are generally poor at anticipating their emotional responses to future events (i.e., ‘affective forecasting’; [ 92 ]). In fact, the association in our study was only weak in individual countries and absent in the overall sample, suggesting that concern of finding out unwanted information is not a major factor in decisions about genetic testing.

Moreover, in the Russian sample the effect was weak and positive, suggesting that people with more concerns were actually more likely to undergo testing for research. Though small, this effect might be reflective of a general willingness to obtain some valid medical information regarding one’s health-related concerns irrespective of the source of such knowledge. This result again stresses the need for improvement of health-related literacy in the general population, including through genetic counselling. In the Era of Genome genetic counselling should be more readily available as it allows to support and educate people about medical, psychological and familial aspects of heritable diseases [ 93 ]. For example, there is evidence that genetic counselling can not only provide knowledge regarding a condition and potentially help to avoid a negative outcome (e.g. [ 57 ]) but also can improve perceived personal control and anxiety symptoms [ 94 ]. It is also worth noting that there was somewhat reduced variance for health-related concerns in the overall sample, as 90% of the sample answered “Yes” to one of the three questions that comprised this variable–“ I would rather not know of any potential debilitating diseases in my future ’.

The current study has a number of limitations. Firstly, the examined predictors explained only a small proportion of the overall variance in willingness to undergo genetic testing. This means that further research must explore additional factors and potential interactions among them. Secondly, some effects varied as a function of the sample. These differences across countries are likely to result from multiple factors that were beyond the current research, including differences in educational systems, socio-economic status, and media coverage of genetic-related questions (e.g. novel findings or regulations introduced). Some of these differences might stem from differences in recruitment strategies (e.g. targeting mostly student populations in Nigeria and Russia during data collection, which resulted in slightly larger samples for these two countries). Further, Nigerian and Russian sample were on average younger compared to UK and USA. However, the effect of age on willingness was very small and was not significant in a step-wise regression. Thirdly, the current study was correlational and did not allow causal inferences. Fourthly, only 4 out of the 86 countries were represented by substantial numbers of participants. As such, a truly cross-cultural study was beyond the scope of this paper. Another limitation might be posed by lengthy data collection, as it might introduce some uncontrolled confounds (e.g. media-related changes in attitudes or effects of Covid-19 pandemic). However, data in different countries were collected in parallel and most collection was made prior to the pandemic. Finally, the current study did not investigate willingness to undergo genetic testing beyond medical and research contexts, such as for obtaining information on non-medical traits, kinship and ancestry [ 95 ], and criminal investigations -–factors that will have increasing relevance as we progress further into the Genomic Era. Further research in this area would aid the development of new approaches to improve decision making about genetic testing [ 96 ].

Genetic testing is at the forefront of discussions about healthcare, health insurance and disease prevention. Thus, it is important to investigate factors affecting decision-making regarding the willingness to undergo genetic testing. Our data showed positive attitudes towards genetic testing in the full sample and in separate countries. Genetic literacy, data protection and general trust in research institutions were among the robust predictors of willingness to undergo genetic testing. However, despite the large number of variables explored in the study, only a small proportion of variance in willingness to undergo genetic testing was explained. This suggests that decision making in this field is a product of many, potentially interacting, factors. Better understanding of the reasons for decisions regarding testing would help people to make decisions that are right for their specific circumstances.

Supporting information

S1 table. proportion of the total n and number of participants from a country after data cleaning..

https://doi.org/10.1371/journal.pone.0293187.s001

S2 Table. Post hoc comparisons for all study variables for Russia, Nigeria, USA and UK.

https://doi.org/10.1371/journal.pone.0293187.s002

  • View Article
  • Google Scholar
  • PubMed/NCBI
  • 3. Wetterstrand K. DNA Sequencing Costs: Data from the NHGRI Genome Sequencing Program (GSP). [Internet]. 2021 [cited 2023 Feb 17]. Available from: https://www.genome.gov/about-genomics/fact-sheets/DNA-Sequencing-Costs-Data .
  • 4. Shin Y. Whole genome sequencing cost in 2023 [Internet]. 2022 [cited 2023 Feb 17]. Available from: https://3billion.io/blog/whole-genome-sequencing-cost-2023/
  • 6. Guttinger S, Dupré J. Genomics and Postgenomics. In: Zalta EN, editor. The Stanford Encyclopedia of Philosophy [Internet]. Winter 2016 Edition. Metaphysics Research Lab, Stanford University; 2016. Available from: https://plato.stanford.edu/archives/win2016/entries/genomics/
  • 8. Kovas Y, Selita F. Oedipus Rex in the Genomic Era: Human Behaviour, Law and Society [Internet]. London: Palgrave Macmillan UK; 2021 [cited 2023 Feb 17]. Available from: https://link.springer.com/10.1057/978-1-349-96048-4
  • 9. Vallance P. Genomics Beyond Health [Internet]. Government Office for Science; 2022 Jan [cited 2023 Feb 20]. Available from: https://www.gov.uk/government/publications/genomics-beyond-health
  • 14. Davies S. Annual Report of the Chief Medical Officer 2016, Generation Genome [Internet]. London: Department of Health; 2017 [cited 2023 Feb 17]. Available from: https://www.gov.uk/government/publications/chief-medical-officer-annual-report-2016-generation-genome
  • 45. Knopik VS, Neiderhiser JM, DeFries JC, Plomin R. Behavioral genetics. Seventh edition. New York: Worth Publishers, Macmillan Learning; 2017. 508 p.
  • 75. World Bank. GDP Ranked by Country. [Internet]. Worldometer; 2017 [cited 2022 Aug 26]. Available from: https://www.worldometers.info/gdp/gdp-by-country/
  • 87. Jones NL, Smith A. Genetic Information: Legal Issues Relating to Discrimination and Privacy [Internet]. Washington D.C.: University of North Texas Libraries; 2005 Feb [cited 2023 Mar 6]. Available from: https://digital.library.unt.edu/ark:/67531/metadc810705/

Scholar Commons

Home > USC Columbia > Medicine, School of > Genetic Counseling > Genetic Counseling Theses and Dissertations

Genetic Counseling Theses and Dissertations

Theses/dissertations from 2023 2023.

Face Validation of a Spanish Non-Invasive Prenatal Screening Knowledge Scale , Kenya Michelle De Leon

Genetic Testing for Autism: The Autistic Adult Perspective , Thomas Scott Dent

Comparing Efficiency, Empowerment, and Satisfaction Between Individual and Group Genetic Counseling for Prostate Cancer , Sarah Marie Dickman

Evaluating Health Awareness in Cancer Genetics Amongst the Black and African American Community in South Carolina , Annika Jaliya Gadson

Jewish Genetic Diseases: Knowledge of Reproductive Risk and Cancer Predisposition Among Young Adults of Ashkenazi Descent , Hayley Kathleen Granger

Experiences of Parents of Children With Late-Onset Pompe Disease Diagnosed By Newborn Screening , Allison Marie Paltzer

Preferences of Adults With Turner Syndrome Regarding Disclosure of Potential Neurodevelopmental And Psychiatric Features , Elizabeth Pancake

Assessing the Gaps and Potential Solutions to Improve Access to Genetic Testing for Autistic Individuals , Nisha Dhiren Pandya

Understanding the Spectrum of CACNA1A Related Disorders: An Analysis of Genotypes and Phenotypes in 416 Individuals , Olivia Jane Wilmarth

Theses/Dissertations from 2022 2022

Evaluation of the Perceived Benefit of a Psychiatric Resource for Parents of Children With 22Q11.2 Deletion Syndrome , Kayla Blankenship

Exploring Genetic Counselors’ Experiences, Language, and Discussion of Consanguinity in Clinical Practice: A Multinational Perspective , Romy Isabel Fawaz

The Utilization of Healthcare Chaplains by Genetic Counselors , Elizabeth G. Hollingsworth

Describing the Experiences of Canadian Genetic Counseling Students Studying in the United States , Madeline Emma Ladouceur

Assessing Identification of Newly Diagnosed Breast Cancer Patients for Referral to Genetic Counseling , Corinne Marie Locke

Assessing Parental Satisfaction of Current Fragile X Syndrome Resources Provided at Diagnosis , Megan Michalski

Examining Parental Disclosure to Unaffected Siblings of Children Diagnosed With Rett Syndrome , Rachel Nicely

Sexual Health Education for Children with Neurodevelopmental Disorders: Genetic Counselor and Pediatrician Perspectives , Mary Catherine Smith

Infertility Education for Men With Cystic Fibrosis and Its Effects: Outlooks on Fertility and Birth Control , Allison Elizabeth Perez Szczepanski

Theses/Dissertations from 2021 2021

Assessing the Anticipated Needs of Transgender Patients In Cancer Genetic Counseling , Jacqueline Baquet

An Exploration of the Genetic Counselor’s Role in The Individualized Education Program , Naomi Jean Barker

Exploration of Patient Communication Preference Regarding Reclassified Genetic Test Results , Cooper Nicole Hall

Language Interpreters’ Perspective of the Interpreter-Genetic Counselor Working Alliance , Dacia Lipkea

Efficacy of Telegenetics: A Diagnostic Yield Comparison Between In-Person and Telemedicine Pediatric Genetic Evaluations , Allie Merrihew

Developmental Regression Analysis and Investigation of Genotype Correlations in Individuals With Classic Rett Syndrome , Aubrey Lynn Rose

Previvor and High-Risk Breast Cancer Patients’ Opinions on A Specialized Management Clinic , Madeleine Nicole Tjoelker

Essential Informational Needs of Parents Receiving a Turner Syndrome Diagnosis: Parent and Genetic Counselor Perspectives , Jewel Lynne Wasson

Theses/Dissertations from 2020 2020

The Psychosocial Burden of LI-Fraumeni Syndrome Tumor Surveillance on Mutation and Non-Mutation Carriers Within Families , Emily Anne Berenson

Evaluating Pregnancy Outcomes of Abnormal Non-Invasive Prenatal Screening Results in a High Risk Obstetrics Practice , Olivia Kesler

Experiences With and Knowledge of Genetics in Families Affected by Congenital Adrenal Hyperplasia: The Parent Perspective , Christine Maccia

Assessing Social Media for Themes of Trisomy 18 and 13 , Falecia Metcalf

The Utility of Whole Exome Sequencing in Patients With Intellectual Disability and Developmental Delay as a First-Tier Diagnostic Testing Strategy , Ellen Richardson

Checkmate: Exploring Father-Son Communication Regarding Reproduction and Sexual Health in Males With Cystic Fibrosis , Dianna C. Sanderson

Quality of Life of Children With Spinal Muscular Atrophy: Parents’ Perspectives in Light of New Treatments , Analyssa R. Tallas

Revisiting the Essential Informational Needs of Parents Receiving a Diagnosis of Down Syndrome , Margaret Jean Wilkes

Theses/Dissertations from 2019 2019

The Perspectives of Emerging Adults with Hereditary Diffuse Gastric Cancer , Carrie Anderson

Reproductive-Aged Adults Diagnosed with Tuberous Sclerosis Complex (TSC): Understanding of Clinical Variability, Perceived Disease Burden, and Reproductive Decision-Making , Diane L. Biederman

Discussing History of Mental Illness in a General Genetic Counseling Setting: Patient and Caregiver Interest and Comfort , Alena Faulkner

Impact of Service Delivery Model on Patient Perceptions and Utility of Genetic Counseling for Hereditary Breast and Ovarian Cancer: An Exploration of Group Genetic Counseling , Alyssa M. Gates

BRCA1 and BRCA2 Mutation Carrier Perspectives on Direct-To- Consumer Genetic Testing for Brca Mutations , Caitlyn E. Mitchell

Amish Perspectives of the Genetic Counseling Process , Brianna Teapole

Evaluating the Social Informational Needs of Emerging Adults with Genetic Conditions , Courtney Whitmore

Theses/Dissertations from 2018 2018

Integrating Genetic Counseling And Testing In The Pediatric Oncology Setting: Parental Attitudes And Influencing Factors , Lauren Renee Desrosiers

Parental experience with whole exome sequencing reanalysis and its impact on the diagnostic odyssey , Nicole D. Lucas

Assessing the Barriers to Cardiac Care in Carriers of Duchenne and Becker Muscular Dystrophy , Lauren Renae Eekhoff

Understanding Barriers To Genetic Testing For Sickle Cell Trait: The African-American Male Perspective , Shandrea Foster

Family Planning Decisions After A Child’s Diagnosis Of Rett Syndrome: A Pilot Study , Erin E. Huggins

Discussing History of Mental Illness During Prenatal Genetic Counseling: Patient Interest and Comfort , Sarah Nimrichter

Is Current Fragile X Syndrome Counseling Enough? Expanding the Clinical Phenotype of Fragile X in Premutation And Intermediate Allele Carriers , Zahra Saadat Girnary

Spiritual Care in Cancer Genetic Counseling: Patient Perceptions of Methods , Christopher Michael Spencer

The Impact of Communication Deficits on Puberty and Sexual Development in Adolescents on the Autism Spectrum , Ashton Wolfe

Theses/Dissertations from 2017 2017

Assessing Women's attitudes Towards Genetic Testing For Hereditary Breast Cancer , Taylor Jane Apostolico

Exercise Recommendations for Active Adults at Risk for Sudden Cardiac Death: “Can I Continue to Exercise?” , Kacie Lynn Baker

Adoptees’ Experiences with Direct-to-Consumer Genetic Testing: Emotions, Satisfaction, and Motivating Factors , Anna Childers

Pediatric Genetic Counselor Perspective on Serving the Foster Care Population and the Integration of Genetic Information within the Health Passport , Angela Rose Douglas

Assessment of Patient Satisfaction with the Provision of Fertility Information in Women with Lynch Syndrome , Rachel Elizabeth Hickey

Recipients’ Perspectives Regarding Expanded Carrier Screening of Gamete Donors , Erika Kristy Jackson

The Perceived Utility of Personalized Genomic Medicine in Individuals with a Family History of Heart Disease: A Pilot Study , Dana Margaret Mittag

The Decision-Making Process for Individuals at Risk for Hereditary Diffuse Gastric Cancer , Alexa Prose

Theses/Dissertations from 2016 2016

Parents’ Understanding of Sensory Processing in their Child with Autism Spectrum Disorder , Katelynn M. Anderson

Parental Experience Of Divulging a Diagnosis Of Fragile X Syndrome To Their Affected Child , Barbara Alyxandra Athens

Unaffected Women’s Decisions to Have Prophylactic Risk-Reducing Mastectomies , Stephanie N. Galloway

Genetic Counseling for Alcohol Use Disorder: Assessment Of Need In Affected And At-Risk Populations , Fayth Michelle Kalb

The Informational and Emotional Support Needs of Grandparents of Children with Pompe Disease , Natasha Lousie Rudy

Re-contacting Cancer Genetic Counseling Patients: Expectations of Patients and Physicians , Zoe Elizabeth Siegel

Reflections On The Current State Of Healthcare Transition for Young Adult Women With Turner Syndrome: Strategies For Facilitating Autonomy and Self-Management , Molly Elizabeth Snyder

Theses/Dissertations from 2015 2015

Evaluating Changes in Patient Anxiety Regarding Classic Cancer Genetic Testing Versus Expanded Multiplex Cancer Genetic Testing , Andrew Todd Alfonso

An Exploration of the Genetic Counselor-Patient Relationship Following a Life-limiting Prenatal Diagnosis , Sabrina Anderson

Exploring Communication Patterns in the Discussion of Maternal PKU Syndrome Between Parents and Daughters , Hannah Beth Andrews

Evaluating the “Family-Centered” Approach of Pediatric Multidisciplinary Down Syndrome Clinics: A Parents’ Perspective , Devon A. Haynes

Goal Achievement in Young Adults with Asperger Syndrome and High Functioning Autism , Melissa Racobaldo

Exploring Birthparent’s Experiences of Creating an Adoption Plan for Their Children with Special Needs , Sanjukta Tawde

Variant Reclassification in Cancer Genetic Testing: Are Genetic Counselors Prepared? A Review of Current Practices , Niamh Gemma White

Exploring How the Risk of Sudden Cardiac Death is Discussed in Families with a Diagnosis of a SADS Condition , Kristin Anne Wiley

Theses/Dissertations from 2014 2014

Walking the Edge with Controversial Use of Preimplantation Genetic Diagnosis (PGD): Opinions and Attitudes of Genetic Counselors , Kristen Everton

The Use of Social Media and the Impact of Support on the Well-Being of Adult Cystic Fibrosis Patients , Margaret Anne Faust

Phenotypes and Variants in Cases Submitted for X-Linked Intellectual Disability (XLID) Gene Panel Testing , Michael J. Friez

The Paternal Age Effect: A Preliminary Study of Current Challenges for Prenatal Genetics Care , Andrew Tyler Gunter

Parents' Dreams for Their Young Adults with Down syndrome: What Resources are Needed to Achieve Them? , Julianna Elise Hudnall

I Wish I Had Known This!: Impact of Age on Life Choices and Testing Satisfaction for BRCA1/2 Mutation Carriers who Underwent Genetic Testing By Age 25 , Sarah Elaine King

The Impact of Culture & Ethnicity on the Counseling Process: Perspectives of Genetic Counselors from Minority Ethnic Groups , Brittanie Morris

Parental Satisfaction and Teacher Perspectives on Inclusive Education of Students with Asperger Syndrome: An Educational Tool , Hannah Warren

Theses/Dissertations from 2013 2013

Examining the Differences in Rapport between Male and Female Cancer Genetic Counselors and Female Clients , John Abernethy

A Qualitative Study on How Prenatal Genetic Counselors Prioritize Cultural Issues When Counseling Patients , Darcy Katherine Berry

Perceptions of Breast Cancer-Related Stigma and Genetic Knowledge Among Latina Women: El Mejor Entendimiento Del Miedo. , Jade Cognetti

Use of Social Media as a Support Network in Families with a Child Diagnosed with Trisomy 13, 18, or 21 , Ginger Elizabeth Edwardsen

Prenatal Decision-Making Process of Patients in Three Cities in South Carolina , Kimberly Marie Hamann

Unique Perspectives and Struggles of Parents Rearing Children with Phenylketonuria with Unaffected Siblings , Cassandra Nicole Hollinger

Working with Patients at risk for Hereditary Degenerative Brain Disorders , Stephen John White

Theses/Dissertations from 2012 2012

Variants of Uncertain Clinical Significance In Pediatric Microarray: Parent Perspectives , Kalina M. Benedict

Assessing Genetics Providers' Perspectives of and Experiences with DTC Genetic Testing: Creation of an Educational Module , Audra Lea Bettinelli

Risk Perception Among BRCA1 and BRCA2 Mutation Negative Patients , Kelsey Johnson

Interpreting and Delivering Microarray Results of `Variants of Unknown Significance': Genetic Counselors' Perspectives , Vruti Mehta

Referral Patterns and Utilization of Clinical Genetic Testing For Patients With Inherited Cardiac Disease , Olivia Myers

Anxiety and Life-stressors Surrounding the Telemedicine Versus Traditional Counseling Experience , Elizabeth Hyden Ziglinski

Theses/Dissertations from 2011 2011

Pediatric Microarray Testing: The Process of Informing and Consenting , Courtney Ann Downtain

Knowledge and Experience With Genetics and Genetic Counseling Among Sign Language Interpreters , Madison Foster

Medical Interpreters' Knowledge of Key Terminology and Principles of Genetic Counseling , Lindsay Diaz Langford

Communication In Family Members With A Rare APC Mutation , Megan Michelle Mann

Termination Rates Following Prenatal Diagnosis For Down Syndrome: A Systematic Review , Jaime Lynn Natoli

Identifying the Needs of Fathers Raising Children With Autism and Children With Down Syndrome , Emily Olivia Rosebrough

Advanced Search

  • Notify me via email or RSS
  • Collections
  • Disciplines

Submissions

  • University Libraries

Home | About | FAQ | My Account | Accessibility Statement

Privacy Copyright

DigitalCommons@SarahLawrence

Home > Graduate Studies > Human Genetics > Human Genetics Theses

Human Genetics Theses

As of May 2015, all Sarah Lawrence College Master’s theses are available digitally. They are made accessible in one of three ways:

1. “Thesis – Open Access” – If the document type in the metadata reads “Thesis – Open Access,” the thesis is available to download immediately via the Download button on the right. NOTE: If a thesis is embargoed for a specific length of time, the date after which the item will be available via open access is noted under the Download button on the right.

2. “Thesis – Campus Access Only” – If the document type in the metadata reads “Thesis – Campus Access Only,” the thesis is available only on the campus of Sarah Lawrence College via IP address or via the Off-Campus Download button if you are a current SLC faculty, staff, or student. You will need your MySLC login and password to access from off-campus.

3. “Thesis – Closed Access” – If the document type in the metadata reads “Thesis – Closed Access,” access to the thesis requires written permission from the author.

If you are not a member of the Sarah Lawrence College Community and you are interested in access to a “Campus Access Only” thesis, please contact the author (if the Contact Author button is available on the right) or request the item through Interlibrary Loan from your local librarian. For additional questions, please contact [email protected].

Copyright for theses is held by the authors.

Theses from 2024 2024

Carrier Screening in the Middle Eastern Population: A Study of the Qatar Genome Programme , Hajer Almulla and Esther Nkrumah

PGx for Psychiatric Conditions: Establishing Best Practices for PGx Return of Results and Evaluating Commercial Test Reports , Mark Carman and Beatriz Diaz

Active-Duty U.S. Army Soldiers’ Awareness and Opinions of Mandatory Sickle Cell Trait Screening , Colleen Castellani

Prenatal Genetic Counselors' Perceptions of Non-Invasive Prenatal Testing (NIPT): A Look at the Informed Consent Process and Common Patient Misconceptions , Emma Duarte and Alaina Swendseid

Characterizing the genetic counselor's role in hematology , Luise Johannes and Kira Dzedzy

Development and Piloting of Patient Education Material on Pharmacogenomic Testing , Louise Limoges

Experiences of Patients with Vascular Connective Tissue Disorders on the Family Planning Option of Surrogacy , Aditi Makaram and Abigail Flores

Exploring The Impact of Chromosome 18 Anomalies on Sibling Mental Health and Family Dynamics: Insights and Implications for Support , Sienna Miley and Rebecca Abel

Examining Rare Instances of Vexas Syndrome in Females , Emily Nasho and Kelsey Crocker

Perceptions of Hyflex in Genetic Counseling Training , Andy Peralta

Exploring Diversity in Genetic Counseling: Experiences and Perspectives of BIPOC Faculty and Leadership , Dessiah Phillips

Further understanding atypical Panorama non-invasive prenatal testing results via pregnancy outcomes: a retrospective study , Shannon Reinhard and Sabrina Wiesmann

An Assessment of Genetics Providers’ Attitudes and Expectations for a Reverse Phenotyping Clinical Decision Aid , Ariella Reiter and Kristina Inman

Evaluating Natural Language Processing Algorithms for the Phenotype-Guided Genomic Diagnosis Platform, GenomeDiver , Bailey Rutter, Maggie Kan, and Kathy Rosales

Evaluation of Patient and Provider Satisfaction with a Point of Care Genetic Testing Model for Cancer Patients , Andrea Verdes and Molli S. Goodman

Theses from 2023 2023

Implementation of a Patient Decision Aid to Facilitate the Selection of a Pre-Test Genetic Counseling Model , Taylor Berry, Valerie Schimmenti, and Magan Trottier

Psychiatrists' Percepitions of and Reactions to a Simulated Psychiatric Genetic Counseling Session , Katharine Cardiff

“People Don't Even Know to Test for It”: Fragile X Community Views on Fragile X Testing, Guidelines, and Impacts , Mugdha Devalkar and Maysen Bratbo

Current Practice Behavior Surrounding Recurrent Pregnancy Loss , Samantha Estin and Katherine Stephens

Analyzing the Physician Experience of Incorporating Population Genomic Screening Results in the Continuity of Care for Patients , Annie C. Foley and Paige Z. Springmann

Evaluation of Targeted On-Demand Genetic Counseling Model for Familial Variant Cascade Testing , Elizabeth Hain and Jonathan Ploeger

Elucidating Reason(s) Low Ses/Underrepresented Patients Who Qualify for Cancer Genetic Testing Decline Testing , Nadiyah Priasti and Amy Huang

Reflections of the Pioneers: An Oral History of the Early Years of Genetic Counseling , Talia K. Sanford and Danielle J. Clynes

Investigating Perceived Barriers and Challenges to Using Virtual Genetic Assistants Among Genetic Counselors , Yongsik Shin

Patient-Reported Barriers & Outcomes of Cardiovascular Genetic Counseling in Diverse Populations of New York City , Marisa M. Thornburg and Alynn M. Kruse

Transgender and Gender Diverse Individuals’ Perspectives on Discussion of Fetal Sex Chromosomes in Obstetrics Care , Dana Tyrie and Alejandra Oliva

Exploring the perspectives of genetics providers in Latin America on genetic counseling and the barriers to establishing the field in the region , Rebeca B. Venezia and Christina M. Sanders

Legal Implications of Documentation in the Electronic Medical Record for Prenatal Genetic Counselors Post-Roe V. Wade , Cassidy Welsh and Cameron Soriano

Simulated Patient Case Development for Cultural Competency Training of Genetic Counseling Students , Makenzie Woltz

Impact of the Sarah Lawrence Genetic Counseling Introductory Fieldwork Simulation on the Self-Efficacy and Training of Students from the Classes of 2021-2023 , Meghan Zadorsky and Skye Wurmbrand

Theses from 2022 2022

Script Concordance Testing in Genetic Counseling Training: A Pilot Study , Yakira Begun and Lila Rae Stange

Clinical psychology trainees' perspectives on psychiatric genetic counseling and the genetic contribution to psychiatric disorders , Talia Belcher and Celine Aziz

An Assessment of the Sexual and Mental Health Needs of Transition Age Foster Youth in New York City , Elizabeth Claessens

Child Welfare Professionals' Perceptions of Referring Foster Children to the Genetics Clinic , Laura Cooper-Hastings

The Role of High Throughput Functional Evidence In Reducing-Population Specific Differences In The Quality of Variant Interpretation , Makenzie Fong and Taylor Silkey

Restrictions on Abortion Affect Genetic Counseling Practice: Genetic Counselors in Abortion Unfriendly States Reflect on Current and Impending Challenges , Grace Amelia Getchell and Sofia Angela Horan

A Systematic Review of Conceptualization of Trust in Precision Medicine Research from Historically Underserved Racial and Ethnic Groups and Populations with Disabilities , Chanyong Tina Kim and Revathy Suresh

Assessing a Diagnostic Tool for Hypermobile Ehlers-Danlos Syndrome: Helping Users Find Appropriate Care For a Relatively Common But Often Unfamiliar Genetic Disease , Abby Kosmin and Natalie Laible

Missing Heritability in Congenital Disorders of Glycosylation (CDG) IIn: A Case Report , Tari Little

The Conversion to Telegenetics Through the COVID-19 Pandemic Impact on the Accessibility to Genetic Counseling for Remote Indigenous Communities , Kaitlyn Mowat and Caroline A. McCrae

Primary Care Providers' Perspectives of Precision Medicine in a Tribal Healthcare System , Katie Neimeyer

A Sharp Increase In The Use Of PGT To Prevent Mendelian Disorders: A Review of Patient Data From 2004-2021 In A Single New York City Center For Reproductive Medicine , Emily Oddo and Hallie Metzger

The Stream Study: Streamlining The Genetic Counseling Process For Newly Diagnosed Breast Cancer Patients , Erika M. Renkes and Tina N. Tran

Contemporary Initiatives to Enhance Racial Diversity in Genetic Counseling Programs: A Snapshot of Current Efforts , Mirriam Sang and Chelsea Miller

Exploring Parental Attitudes on Autism Genetic Testing After Receiving Non-Pathogenic Results , Christina Szi

Return of Results from Genetic Research: A Study Of Nephrologists , Robyn Weiss

Theses from 2021 2021

Genetic Counseling Around the Globe: Prenatal Screening Practices During the First Trimester , Anna Bauer, Dhart Adhia, Rachel K. Lanning, and Jenny G. Zhang

The Impact of a Lynch Syndrome Diagnosis by Population Genomic Screening on Family Communication, Medical Management, and Lifestyle Changes , Reem Ibrahim Bux and Brooke Nicole Delehoy

Utilization and Perceived Value of Genetic Counselors Within U.S. Hemophilia Treatment Centers , Caylynn Carls

Exploring Attitudes Towards Newly Approved Therapeutics in Prenatal Genetic Counseling Practice , Charlotte Close

Patient and Provider Response to a Prenatal Pre-Visit Education Chatbot , Sabina Gudmundsson, Chantal Marie Muyalde, and Catherine Urbina

On the Development of Psychosocial Skills in Genetic Counseling Programs: Perspectives from Recent Graduates , Emily Johnson and Daniel Morrice

Investigating the Experience of Underrepresented Groups Applying to Genetic Counseling Training Programs , Enas Louzy-Hanna

Carrier Screening for Women Undergoing Elective Oocyte Cryopreservation (EOC): A Look at Practice Among Reproductive Endocrinologists at an Academic Fertility Clinic , Melissa Manuelli

Psychiatric Professionals Believe Polygenic Risk Scores for Schizophrenia Have Future Clinical Utility: Reservations about Current Integration , Tiahna Moorthy and Huyen Nguyen

Patient and Provider Response to a Prenatal Pre-Visit Education Chatbot , Chantal Marie Muyalde, Catherine Urbina, and Sabina Gudmundsson

Attitudes and Actions of Genetic Counseling Program Directors Regarding Standards for Accreditation of Genetic Counseling Training Programs , Ben Newsum

Towards The Growing Edge: Integrating Simulation In A Genetic Counseling Graduate Program , Cassandra Pisieczko

Lessons Learned from COVID-19: Genetic Counseling Adaptations and Challenges to Alternate Service Delivery Models , Aisha Rekab and Angel Nguyen

“Did You Know Before?”: Documenting the Lived Experiences of Parents of Children with Down Syndrome in an Era of Widespread Prenatal Testing , Samantha Riddell and Amandeep Pabla

Relationship Between Outcomes of Psychiatric Genetic Counseling and Time Since Onset and Diagnosis of Psychiatric Illness , Sarah Saxton

Patient and Provider Response To A Prenatal Pre-Visit Chatbot , Catherine Urbina, Chantal Marie Marie Muyalde, and Sabina Gudmundsson

How Genetic Findings Associated with Autism are Used: A Study of Parents and Caregivers , Veronica Sue Yamane and Tzofia Nechama Drori

Theses from 2020 2020

Exploration Of Patient Attitudes Toward Receiving Incidental Diagnoses Of Lysosomal Storage Disorders Through Expanded Carrier Screening , Ally Abbott and Xindi Song

Post Mortem Genetic Testing in Sudden Unexpected Death in Epilepsy [SUDEP]: A Pilot Story , Yu An and Alejandra M. Cantú Villarreal

The Efficacy of Whole-Genome Sequencing in the Diagnosis Of Complex Neurological Phenotypes , Bianca Blake

Test Ordering Practices in Cancer Genetic Counseling , Lauren Costantin and Tiana Grgas

Impact of Psychiatric Genetic Data on Tort Litigation and Its Relationship With Stigma , Ashlyn Enokian and Kira Dineen

The Utility Of Adding Key Phenotypic Criteria Refinement To ACMG Guidelines , Lauren Frank

Current Practices in Post-Mortem Cariogeentic Test-Ordering an Genetic Counseling In Cases of Sudden Cardiac Death (SCD) , Kiran Gangwani and Miranda Di Biase

A Survey Of Autistic Adults' Opinions On Data Security and Privacy In Precision Medicine Research , Bianca Haser

Safety and Feasibility of Early Prenatal Diagnosis Via Celocentesis , Elisheva Kleinman

An Analysis of Fitness Affiliated Direct-To-Consumer Genetic Tests , Karl N. Krahn and Nicole D. Wengrofsky

Primary Care For Disease Patients: Exploring Services Received, Healthcare Providers Involved, And Patient Satisfaction , Catherine Mayo, Eliana Kahan, and Jasmine Chao

Clinical Implementation of Polygenic Risk Scores For Cardiovascular Disease: An Assessment of Cardiovascular Genetic Counselors' Knowledge and Opinions , Maryam Nazir and Maria van Noordenne

The Career Arc Of Genetic Counselors: Trends, Transitions, And Motivations , Michael Peneycad, Wanchun Janice Smires, Kara Williams, and Jovanni Cuevas

Genetic Counselors' Views Of Traits That Make For Effective Leaders , Jennifer Rand

Psychosocial Outcomes Associated With ctNA-Based Cancer Screening Test , Gabrielle Shermanski and Simone Biggers

An Investigation Into The Reasons Behind Quality Of Life Perception In Individuals With HNPP As Compared To Individuals With CMT1A , Caitlin Walsh and Sophia Rodriguez

Theses from 2019 2019

Descriptive Analysis of the Testing Outcome Populations of a Highly Facilitated Cascade Genetic Testing Framework for Cancer Predisposition , Samantha R. Anderson

Adolescents’ Attitudes Towards Direct-to-Consumer Genetic Testing , Carli Andrews, J. Fitzpatrick Doyle, Rebekah Hutchins, and Katherine Orr

Disability Service Learning: A Study on the Potential Impact of an Educational Intervention on the Attitudes and Biases of Genetic Counseling Students Toward Disability , Michelle Bina and Lucas Hollifield

What do Teenagers Think about Precision Psychiatry? , Erika Brockhoff and Fiorella Herrera

Recall Of Informed Consent For Prenatal Aneuploidy Screening , Taylor Cain, Michelle Kao, Elena Cothalis, and Pranali Shingala

Exploring Perceptions of What Genetic Counseling is Amongst Families Affected by Genetic Conditions, Who Have Not Yet Had Genetic Counseling Themselves , Stephanie Briana Cordeiro

Exploring the Implementation of Non-Traditional / Expanded Training for Genetic Counselors , Katia Dergham

Assessing The Impact Of Predictive Testing Protocols On Provider Burden For Huntington's Disease , Paige Ernste and Abigail Patenaude

Genetic Counseling Students’ Attitudes Towards Psychiatric Illness , Rebecca Haegedorn and Patricia Thompson

Group Genetic Counseling for Low-Risk Prenatal Patients , Sarah Hopkins

Genetic Counseling Referrals and Genetic Profiles of Male and Young Female Breast Cancer Populations , Ji-Sun Kim

Motivations, Barriers, and Interests in Genetic Testing for Patients with Pancreatic Cancer , Anna Kolbuszewska

Sun exposure as a risk factor for precipitating vision loss for individuals with LHON mitochondrial variants , Michelle Kowanda

Attitudes of Healthcare Professionals Towards the Utilization of Genetics Professionals Following the Diagnosis of Autism Spectrum Disorder , Sydney Alexandra Lau and Tova Lejtman Wagner

Parental perception of pediatric clinical exome sequencing in a Latino population , Daniel Luksic and Radha Sukhu

Experiences of Adults at 50% Risk of Inheriting a Genetic Mutation for Early-Onset Dementia , Margalit E. Rosenblatt

Genetic Counseling Approaches To Moderate-Penetrance Breast Cancer , Tanaya Shroff and Elizabeth Del Buono

Should Genetic Information be Used to Determine Special Education Eligibility and Other Educational Services? , Nicholas Staropoli

Theses from 2018 2018

Cardiovascular Genetics: Getting a “Pulse” on How Cardiologists Assess and Act on Cardiogenetic findings that May Lead to Sudden Cardiac Death , Tamar Ailenberg, Anna (Schon) Levy, and Samone Schneider

Current Attitudes and Practices in Genetic Counseling Concerning Noninvasive Prenatal Screening – A Follow Up Study , Carla Bennett and Abigail Whiting

Genetic Counselors’ Preparedness for Incidental Findings from Non-Invasive Prenatal Testing , Janel Case and Paige Hazelton

  • Collections
  • Disciplines
  • Sarah Lawrence Library
  • Terms of Use

Advanced Search

  • Notify me via email or RSS

Author Corner

  • FAQs for Grad Students
  • Submission Guidelines
  • Submit Your Thesis

Home | About | FAQ | My Account | Accessibility Statement

Privacy Copyright

  • Research article
  • Open access
  • Published: 25 May 2017

Awareness, knowledge, perceptions, and attitudes towards genetic testing for cancer risk among ethnic minority groups: a systematic review

  • Katie E. J. Hann 1 ,
  • Madeleine Freeman 2 ,
  • Lindsay Fraser 1 ,
  • Jo Waller 2 ,
  • Saskia C. Sanderson 2 ,
  • Belinda Rahman 1 ,
  • Lucy Side 1 ,
  • Sue Gessler 1 &
  • Anne Lanceley 1

for the PROMISE study team

BMC Public Health volume  17 , Article number:  503 ( 2017 ) Cite this article

19k Accesses

196 Citations

4 Altmetric

Metrics details

Genetic testing for risk of hereditary cancer can help patients to make important decisions about prevention or early detection. US and UK studies show that people from ethnic minority groups are less likely to receive genetic testing. It is important to understand various groups’ awareness of genetic testing and its acceptability to avoid further disparities in health care. This review aims to identify and detail awareness, knowledge, perceptions, and attitudes towards genetic counselling/testing for cancer risk prediction in ethnic minority groups.

A search was carried out in PsycInfo, CINAHL, Embase and MEDLINE. Search terms referred to ethnicity, genetic testing/counselling, cancer, awareness, knowledge, attitudes, and perceptions. Quantitative and qualitative studies, written in English, and published between 2000 and 2015, were included.

Forty-one studies were selected for review: 39 from the US, and two from Australia. Results revealed low awareness and knowledge of genetic counselling/testing for cancer susceptibility amongst ethnic minority groups including African Americans, Asian Americans, and Hispanics. Attitudes towards genetic testing were generally positive; perceived benefits included positive implications for personal health and being able to inform family. However, negative attitudes were also evident, particularly the anticipated emotional impact of test results, and concerns about confidentiality, stigma, and discrimination. Chinese Australian groups were less studied, but of interest was a finding from qualitative research indicating that different views of who close family members are could impact on reported family history of cancer, which could in turn impact a risk assessment.

Interventions are needed to increase awareness and knowledge of genetic testing for cancer risk and to reduce the perceived stigma and taboo surrounding the topic of cancer in ethnic minority groups. More detailed research is needed in countries other than the US and across a broader spectrum of ethnic minority groups to develop effective culturally sensitive approaches for cancer prevention.

Peer Review reports

Several types of cancer have been found to be associated with hereditary gene mutations [ 1 , 2 ]. Specifically, mutations in the BRCA1 and BRCA2 genes are known to be linked to breast, ovarian [ 3 , 4 ], and prostate cancer [ 5 ]. Over the past 20 years, the scientific understanding of cancer related genetics has greatly improved. Within several European countries and the US, patients diagnosed with a potentially hereditary cancer or with a strong family history can receive genetic counselling and testing to establish whether they have an inherited cancer gene mutation. Knowledge about personal cancer risk can help currently healthy individuals to make health care decisions, such as whether to attend regular screening or opt for surgery, in order to help reduce the risk of developing cancer [ 6 ].

Public attitudes towards genetic testing for the risk of diseases, including cancer, have been found to be generally positive [ 7 , 8 , 9 ]. In a US study, 97% of participants indicated that they were at least somewhat interested in the topic of genetic testing and the majority had positive attitudes about genetic research and approved of the use of genetic testing in the detection of diseases [ 9 ]. Positive attitudes towards genetic testing are also reported in a Dutch survey study that found that 64% of participants believed genetic testing would help people to live longer [ 8 ]. However, there are also concerns amongst the general public that genetic test results could be used to discriminate against those with a genetic predisposition for illness [ 9 ] and that genetic testing could result in people being labelled as having “good” or “bad” genes [ 8 ].

The incidence and burden of cancer varies between ethnic groups. In the UK, incidence of breast cancer is higher amongst White women compared to all other ethnic groups [ 10 ], but Black men have higher rates of prostate cancer than White men [ 11 ]. Similarly, in the US overall cancer incidence and mortality has been found to be highest in Black men compared to other ethnic groups, and whilst Black women have a lower incidence of breast cancer than White women they have a worse mortality rate [ 12 ]. In the US Hispanics are reported to have lower incidence and mortality rates of cancer than White Americans [ 13 ], and evidence suggests that whilst Asian Americans also tend to have lower cancer rates than Whites, increasing incidence rates have been detected between 1990 and 2008 for breast, colorectal, and uterine cancer amongst several subgroups of Asian American women [ 14 ].

Despite the potential health benefits of cancer susceptibility testing, ethnic minority groups have been found to be underserved by genetic services and underrepresented in research in Europe and the US [ 15 , 16 , 17 , 18 ]. In a UK study, only 3% of patients referred to 22 regional genetics services were from an ethnic minority group [ 19 ]. Armstrong et al. [ 20 ] also found that women pursuing BRCA1/2 genetic testing in a US study were significantly more likely to be White, and Levy et al. [ 21 ] report that significantly fewer Black and Hispanic women with a new diagnosis of breast cancer and at risk of carrying a BRCA gene mutation had genetic testing.

A previously published review investigating what may hinder African, White Irish, and South Asian ethnic minority groups’ access to cancer genetic services found potential barriers included low awareness and knowledge of genetic testing and available services, language barriers, stigma associated with being at risk, fatalistic views of cancer, anticipation of negative emotions, uncertainty about the information provided, and mistrust of how data would be used [ 22 ]. The current review aims to investigate factors that might act as barriers or facilitators to the uptake of genetic testing across diverse ethnic minority groups. The review fills a critical knowledge gap by focusing on awareness, knowledge, perceptions, and attitudes towards genetic testing for cancer susceptibility, and reasoning for and against testing by people from Black and ethnic minority backgrounds.

The systematic review followed PRISMA guidelines and an a priori published protocol ( https://www.crd.york.ac.uk/PROSPERO/ CRD42016033485).

Eligibility

The review includes primary studies that aimed specifically to investigate ethnic minority groups’ awareness, knowledge, perceptions, and attitudes, or provide information on participants’ reasons for/against interest, intentions or actual uptake of genetic testing/counselling for personal cancer risk. Studies with a focus on genetic testing or genetic counselling were included as currently these services often go hand-in-hand and we wanted to see if there were important attitudes affecting uptake. We included quantitative and qualitative studies conducted in Europe, the US or Australia, and published in English in a peer reviewed journal from the year 2000 onwards.

Research that investigated awareness/attitudes/perceptions of direct-to-consumer genetic testing, or that only investigated participants’ perceptions of cancer or their experience of receiving genetic information were excluded from the review. Articles that evaluated recruitment methods, or interventions to raise awareness/knowledge of genetic testing amongst ethnic minority groups were also excluded. Studies focusing mainly on Ashkenazi Jewish participants were not included as this group is known to have an increased risk of carrying BRCA1/2 gene mutations which might uniquely influence their knowledge/interest/attitudes to genetic testing. Ashkenazi Jewish women also have a high level of acceptance of genetic testing [ 23 , 24 ]. Unpublished research was not sought for this review.

Study search

Four databases were searched in December 2015: PsycInfo (OVID interface), CINAHL (EBSCO interface), Embase (OVID interface) and MEDLINE (OVID interface). The search terms included a combination of thesaurus (MeSH) terms and keywords such as: ‘ethnic’ or ‘minorit*’ or ‘African’ or ‘Asian’, combined with ‘know*’ or ‘aware*‘or ‘attitude*’ or ‘perception*’ and ‘cancer’ and ‘genetic testing’ or ‘genetic counselling’. [See Additional file 1 for full lists of search terms.]

Study selection

Figure 1 presents the article selection process. Study articles were independently selected by two researchers [KH & MF]. Articles were initially identified by the titles and abstracts and those which appeared eligible were reviewed in full. A reference list search was also conducted to identify additional articles. Any disagreements or uncertainties on the inclusion of studies were brought to a third researcher [AL] to make a final decision.

Study inclusion flow diagram. The flow diagram presents the processes of inclusion and exclusion to identify the final sample of articles to be included in the review and reasons for excluding articles

Data extraction and analysis

A data extraction tool was designed to gather relevant information on the study design, methods, and results on awareness, knowledge, attitudes, and perceptions relating to genetic testing/counselling for cancer risk.

A thematic synthesis of the qualitative research was carried out in line with recommendations by Thomas and Harden [ 25 ]. The synthesis involved extracting and analysing each article’s results section, including direct quotes. An iterative process of re-reading and coding the text was used, and a coding manual was produced based on the themes identified. Once coding was completed a second researcher [MF] independently analysed 20% of the qualitative data using the coding manual. The analysis was organised using the software package NVivo (version10).

Quality assessment

The quality of each study was assessed using tried and tested tools designed by Kmet et al. [ 26 ] for quantitative and qualitative research. The assessment tool criteria included an assessment of the description of the study objective, design, methods, data analysis, and conclusions. Criteria are scored using a 3-point Likert scale: 0 (criteria not fulfilled); 1(partial fulfilment); 2 (criteria fulfilled). A final score is calculated by adding all relevant criteria scores and dividing by the total possible score for each study. The included articles were independently quality assessed by two researchers [KH & MF]. When disagreements arose they were discussed until consensus was reached.

For the purpose of the review we have standardised the terms used to refer to the groups included, see Table 1 for a key and frequencies of groups included across the studies. Table 2 summarises the 31 quantitative studies and the results on ethnic minority groups’ awareness, knowledge, attitudes, and perceptions in regards to genetic counselling ( n  = 2), genetic testing ( n  = 27), or both ( n  = 2) for cancer risk. Measures used in the studies were often non validated study specific instruments and due to their heterogeneity across different ethnic groups and in reference to different cancers a meta-analysis of the results was not possible [see Additional file 2 for a list of the measures used]. Table 3 presents a summary of the 10 qualitative and mixed methods studies, noting themes as originally identified. Of these 10 qualitative studies, 5 involved interviews, 4 involved focus groups and 1 used both focus groups and in-depth interviews. The tables have been arranged by ethnicity and cancer type.

A total of 82,432 individuals from African American, Hispanic, Asian American, White, and Chinese Australian ethnic groups took part in the 41 studies reviewed. Thirty-nine of the forty-one studies were conducted in the US and two were conducted in Australia. Breast and ovarian cancer or BRCA genetic testing were most frequently referred to within the reviewed research, twelve referred to a number of cancers or cancer in general, three referred to prostate cancer, two to colon/colorectal cancer, and one to lung cancer.

Quantitative studies

Awareness was measured in two ways across 15 studies; either a dichotomous Yes/No question [ 27 , 28 , 29 , 30 , 31 , 32 , 33 , 34 , 35 ] or a measure asking participants how much they had heard/read about genetic testing [ 36 , 37 , 38 , 39 , 40 , 41 ]. Figure 2 presents the percentage of participants who indicated awareness or having heard/read a fair amount/a lot about genetic testing for cancer risk by study and ethnic group. Within general population samples of Hispanics, awareness of genetic testing for cancer risk ranged from 7.7% of Spanish only speaking Hispanics [ 28 ] to 27.9% of internet users [ 30 ]. Across two samples of Hispanics with a high risk of cancer due to family history or a personal experience of cancer, 46.7% [ 38 ] and 47.1% [ 41 ] were aware of genetic counselling for cancer risk in general; awareness of genetic counselling for colon, breast, and ovarian cancer varied. Whilst Gammon et al. [ 27 ] found 43.1% of a Hispanic sample were aware of BRCA1/2 testing, this sample included only high cancer risk participants of whom several had already had genetic testing.

Awareness of genetic testing/counselling for cancer risk across ethnic groups. The bar graph presents the percentage of participants aware of genetic testing or counselling for cancer risk across ethnic groups. *Indicates percentage of participants that reported having read/heard a fair amount/a lot about genetic testing. BC/OC = breast/ovarian cancer; GT = genetic testing; GC = genetic counselling

Amongst African Americans, awareness of genetic testing for cancer risk ranged from 29% [ 30 ] to 54% in general population samples [ 35 ]. Awareness of the BRCA1/2 gene and mutation testing appears to be particularly low, one study [ 35 ] reported that only 12% were aware and another [ 33 ] reported that 25% of African American participants were aware. Two studies reported that around 25% of Asian Americans were aware of genetic testing for cancer risk [ 31 , 32 ].

Five studies reported that awareness significantly differed by ethnicity, with more White participants being aware of genetic testing for cancer risk than Hispanics, African Americans, and Asian Americans [ 29 , 30 , 32 , 33 , 40 ]. Vadaparampil et al. [ 34 ] also reported that whilst 20% of their Hispanic sample were aware of genetic testing for cancer risk, awareness varied by ethnic subgroups.

Only one study assessed whether awareness was associated with intentions to have genetic counselling, and found no significant association [ 41 ].

Eight of the included studies measured knowledge of hereditary cancer genetics within samples at an increased risk for cancer based on family history and including individuals with a personal cancer diagnosis [ 41 , 42 , 43 , 44 , 45 , 46 , 47 , 48 ]. Figure 3 presents the average percentage of correctly answered knowledge questions across the studies. Findings suggest that knowledge was varied but limited amongst African Americans, with between 30% and 70% of questions answered correctly on average across five studies [ 42 , 43 , 44 , 45 , 46 ]. Two studies suggested that Hispanic participants also had limited knowledge, with scores of approximately 5 out of 11 for hereditary breast cancer knowledge [ 47 , 48 ], although a high average score of 43.8 out of 55 was observed for genetic counselling knowledge in a third [ 41 ].

Percentage of correctly answered knowledge questions by ethnicity. The bar graph presents the average percentage of correctly answered knowledge questions across the studies and ethnic groups

Only Donovan and Tucker [ 42 ] found a significant difference between African American and White participants’ scores on cancer genetics knowledge, indicating lower knowledge in the African American group. However the overall difference between scores was small, with a mean score of 7.7 out of 14 for Whites and 7.0 for African American.

High cancer genetics knowledge was found to be a significant predictor of genetic testing uptake in one study with African Americans [ 44 ]. In another study, those who participated in genetic counselling and testing had significantly higher scores for knowledge of cancer genetics than those who accepted neither, however cancer genetics knowledge did not reach significance as a predictor of genetic testing uptake [ 45 ]. Three other studies found no associations between cancer genetics knowledge and interest in or intentions to have genetic testing [ 41 , 42 , 43 ].

Attitudes and perceptions

Several different measures of attitudes/beliefs/perceptions regarding genetic testing or counselling were used, although similar items were used across these. Some studies reported attitude scores and others reported the number or percentage of participants who agreed with each statement. Ten studies reported that African American and Hispanic participants highly endorsed statements about the benefits of genetic counselling/testing for cancer risk, whilst endorsing limitations to a lesser extent [ 36 , 37 , 38 , 41 , 42 , 45 , 47 , 49 , 50 , 51 ]. The results indicate that overall participants had positive attitudes and perceived several benefits of genetic testing.

Highly endorsed benefits of genetic counselling/testing for cancer risk included: to help make decisions on (enhanced) screening (endorsed by 81–100%) [ 36 , 42 , 43 , 45 , 49 , 50 ]; to motivate self-examination (endorsed by 90–92%)[ 43 , 45 , 49 ]; receipt of information for family/being able to help family and children (endorsed by 38–99%) [ 27 , 36 , 38 , 41 , 43 , 45 , 47 , 49 , 52 ]; to reduce concern about cancer (endorsed by 60–90%) [ 38 , 41 , 43 , 45 , 49 ]; to reduce uncertainty (endorsed by 68–100%) [ 42 , 43 , 50 ]; to provide a sense of personal control (endorsed by 67–79%) [ 37 , 43 , 45 , 49 ]; to help plan for the future (endorsed by 66%) [ 36 , 52 ]. More variation was seen in attitudes towards benefits such as: to help make important life decisions (endorsed by 21–88%) [ 36 , 41 , 43 , 45 , 49 , 50 ]; to provide reassurance (endorsed by 42–80% of participants) [ 38 , 50 , 52 , 53 ]; to help with cancer prevention (endorsed by 31–88%) [ 27 , 33 , 50 ]; and to help make decisions about preventative surgery (endorsed by 44–77%) [ 41 , 43 , 45 , 50 ]. Similarly, Kinney et al. [ 44 ] report that, when asked for reasons why they enrolled for genetic testing, participants cited family/personal motives (62%), information (28%), and society (9%).

Some of the most frequently endorsed limitations or barriers to genetic testing/counselling included: anticipated increased worry about offspring/relatives if test result is positive (endorsed by 53–95%) [ 38 , 41 , 43 , 45 , 49 ]; anticipated personal emotional reaction if test result is positive e.g. worry, fear, anger (endorsed 31–75%) [ 27 , 36 , 37 , 38 , 43 , 45 , 49 , 52 , 53 ]; concern about family’s reaction or impact on family (endorsed by 27–52%) [ 36 , 49 , 50 , 52 ]; concerns about confidentiality (endorsed by 12–72%) [ 40 , 42 , 45 , 49 , 50 ]; concern about jeopardising/losing insurance (endorsed by 11–58%) [ 33 , 36 , 38 , 41 , 43 , 45 , 47 , 49 , 50 ]; cost (endorsed by 32–40%) [ 27 , 47 ]; and feeling unable to handle the test emotionally (endorsed by 3–40%) [ 40 , 42 , 49 , 50 ]. Reported reasons for a lack of interest/intention to pursue testing included cost, time, worry that others would find out, belief the results could be wrong, worry about increased risk, concern about discomfort, concern about discrimination, not wanting blood taken, logistical reasons, personal reasons, and having heard negative experiences of others who had undergone testing [ 44 , 54 , 55 ].

In studies comparing ethnic groups, results suggest that those from ethnic minority groups may hold less positive views or have greater concerns about genetic counselling/testing compared to White participants. Peters et al. [ 33 ] found that compared to White participants, African Americans were less likely to endorse health benefits of genetic testing and were more likely to believe that the government would use test results to label groups as inferior, however there were no other significant differences between the groups on attitudes towards insurance and job discrimination due to genetic test results. Sussner et al. [ 56 ] also reported that foreign born African American participants anticipated a more negative emotional reaction to genetic testing than those born in the US. Thompson et al. [ 40 ] found that African Americans and Latinas had significantly higher medical mistrust than Whites, and were more likely to believe that genetic test results would be used to show that their ethnic group is not as good as others, or to interfere with the natural order of things. Latinas with lower levels of acculturation and heightened medical mistrust also cited more barriers to genetic testing [ 39 ]. However, whilst Armstrong et al. [ 57 ] found that significantly more African Americans had high healthcare mistrust, there was no significant difference between African American and White participants’ willingness to undergo cancer risk genetic testing across several hypothetical scenarios.

Attitudes and perceptions were associated with uptake, interest, or intentions to have genetic testing in ten studies. Thompson et al. [ 45 ] reported that those who attended genetic counselling/testing perceived significantly fewer barriers and had more intrusive thoughts of breast cancer than those who did not. Similarly, Donovan and Tucker [ 42 ] reported that being at least somewhat interested in genetic testing was associated with more positive beliefs. Seven studies found that perceived high risk of cancer or of being a cancer gene mutation carrier was associated with acceptance of genetic testing or interest/intentions to test [ 41 , 42 , 43 , 44 , 50 , 53 , 58 ]; however, one study reported that perceived cancer susceptibility was negatively associated with intentions [ 54 ]. Whilst Hughes et al. [ 59 ] found higher fatalistic beliefs about cancer in genetic test accepters than decliners, Myers et al. [ 54 ] indicated a negative association between fatalism about prostate cancer prevention and intentions to receive prostate cancer testing to find out about personal risk. Other attitudes and perceptions positively associated with genetic testing uptake or interest/intentions included the absence of the belief that testing leads to discrimination, reassurance from testing [ 53 ], belief in efficacy of screening [ 54 ], perceiving that it is good to know future risk, and believing in control through knowledge [ 37 ].

Qualitative studies

The 10 qualitative studies included 31 Chinese Australians, 17 Asian Americans, 120 African Americans, and 163 Hispanics. The majority of participants were female and were at high risk of cancer based on family history or had a diagnosis of cancer, one study used a sample of the general population [ 35 ]. The thematic synthesis of the qualitative studies identified five broad themes: information deficits; cancer related anxiety; positive and negative attitudes and perceptions; family; and service provision and access.

Information deficits

Although some studies indicated that participants had good awareness and knowledge of hereditary cancer and the involvement of genetics, these participants tended to be those who had experienced cancer themselves or in a close family member [ 60 , 61 ], or had previously attended genetic counselling [ 62 , 63 ].

“Genes have a big part in us having cancer. My family has a higher chance of getting cancer compared to other families as many of our family members had cancer.” Chinese Australian, Eisenbruch et al. [ 62 ]

Several studies including Chinese Australian, Hispanic, and African American samples reported that the majority of participants had low awareness and knowledge about the importance of genetics in cancer, BRCA , and genetic testing [ 52 , 60 , 61 , 63 , 64 , 65 , 66 , 67 ]. Some confusion on the difference between genetic counselling and genetic testing was identified [ 61 , 65 ]. Generational differences in awareness, knowledge, and beliefs were observed in studies with Hispanic and Chinese Australian samples, indicating that older generations were less aware and perhaps less open to genetic testing for cancer risk [ 61 , 62 , 63 , 64 ]. Some Chinese Australian participants reported that older generations and those with traditional beliefs would not accept that mutated genes can be passed through generations, attributing the development of illness instead to bad luck or past bad deeds [ 62 , 64 ]. The lack of understanding about hereditary cancer and unfamiliarity with genetic testing could act as a barrier to testing. Participants across several studies expressed a need for interventions, campaigns, and the provision of information to increase awareness and knowledge of cancer and genetic testing [ 35 , 52 , 60 , 61 , 65 ].

“Nobody knows it [genetic testing]. They are so far away [from Western medicine]. The mammogram is coming out but this is the first time I heard about the gene test.” Asian American female, Glenn et al. [ 63 ]
“Black people aren’t into that yet. We need to know more about it so it will not seem so farfetched.” African American female, Matthews et al. [ 52 ]

Closely linked to participants’ low awareness and knowledge of cancer and genetic testing were the misconceptions held, including believing that genetic counselling prevents cancer [ 65 ], stress causes cancer [ 52 , 63 ], cancer is contagious [ 62 ], injuries cause cancer [ 67 ], and focusing or worrying about cancer causes it to develop [ 52 , 62 , 64 ]. It is highlighted that some of these beliefs may make people less willing to discuss cancer or attend cancer services.

“I have heard that if you hit your breasts, it could cause cancer—I saw a woman who got cancer after being hit with a bottle on her breast. My mother hit her stomach and then she got cancer in her uterus. The underwires in bras are not good and they harm your breast tissue and cause cancer. I take them out.” Hispanic female, Vadaparampil et al. [ 67 ]

Cancer related anxiety

Four studies found that African American and Hispanic participants had fatalistic views of cancer, which they perceived to be an illness that cannot be cured and is associated with death [ 52 , 61 , 65 , 66 ]. Across all the included ethnic groups participants associated emotions such as worry, anxiety, and fear with cancer and genetic testing [ 35 , 52 , 60 , 61 , 63 , 64 , 65 , 66 , 67 ]. Participants expressed that fear of cancer and knowing their cancer risk was a deterrent to genetic testing [ 35 , 52 , 60 , 61 , 63 , 65 , 66 , 67 ]. Participants anticipated experiencing anxiety whilst undergoing genetic counselling/testing [ 52 , 64 ] and worried about the potential emotional impact that a positive test result would have on them [ 52 , 67 ]. Some Hispanics reported that at times embarrassment prevented them from seeking medical help and attending cancer screening [ 63 , 66 ]. Sheppard et al. [ 60 ] also found that women were fearful of making decisions about managing cancer risk after testing.

“… I think it will be harder for them to find out that they have a gene or a mutation that may cause...their body to develop cancer...in the future, and so I think for a lot of people just knowing that they have the mutation will be a lot more of anxiety source than actually helpful…” Hispanic, Kinney et al. [ 66 ]

However, within a Hispanic community, Sussner et al. [ 61 ] found that younger generations were less fatalistic about cancer and were more enthusiastic towards genetic counselling.

“But now the word cancer is different from before…before people were ashamed, now they believe they will beat it, it was once a taboo.” Hispanic female, Sussner et al. [ 61 ]

Positive and negative attitudes and perceptions

Interest in and positive attitudes towards genetic testing for cancer susceptibility were identified [ 60 , 63 , 66 ]. Personal health benefits, such as cancer prevention, early detection, and risk management, were highlighted as main benefits of genetic counselling/testing by African American, Hispanic, and Asian American participants [ 35 , 52 , 63 , 65 , 66 ]. Some participants felt that genetic testing would help to motivate them to be proactive towards their health and take action by making positive behavioural changes [ 35 , 65 ] and others considered the knowledge gained from genetic testing to be empowering [ 61 , 65 ].

Benefits of genetic testing included “the opportunity for early detection, instead of waiting until the cancer develops.” African American female, Adams et al. [ 35 ]

Some African American and Hispanic participants discussed their spirituality and God, this was not presented as a barrier to genetic testing but as a way of seeking guidance and coping [ 60 , 61 , 67 ]. Some participants felt that their spirituality was a motivator for finding out about their cancer risk [ 65 ].

“I do believe that everything is in God’s hands, regardless. But, I don’t think that my spiritual beliefs would prevent me from going. If anything, they would motivate me to go…” African American female, Ford et al. [ 65 ]

Several negative attitudes and perceptions were held by participants in relation to cancer and genetic testing. Discussing illnesses such as cancer was described to be taboo by Chinese Australian, African American, and Hispanic participants [ 61 , 62 , 64 , 66 ]. Participants explained that in their culture there is secrecy around cancer and individuals might not tell family if they had it [ 64 , 66 ]. This taboo is closely linked to the perceived stigma of having an illness, feeling shame and not wanting to be treated differently. Fear of discrimination by employers and insurers, and stigma was discussed across several studies and highlighted as a deterrent to testing [ 35 , 60 , 61 , 63 , 65 , 66 ]. Among Asian American and Chinese Australian participants, there were also concerns that receiving a positive genetic test result would create stigma that they have “bad genes”, which could impact on marriage prospects [ 63 , 64 ].

“The stigma [would stop me from getting tested] because you always have a concern that somewhere this information is gonna reside on a computer somewhere; it may prevent you from getting employment, future insurance, or any number of things so that’s [my] concern. That is one thing that has stopped me from going ahead with testing.” African American female, Sheppard et al. [ 61 ]

Five of the eight qualitative studies conducted in the US reported that participants, particularly African Americans, discussed mistrust and disillusionment in the health care system [ 35 , 52 , 61 , 63 , 65 ]. Participants expressed concerns that their genetic data would be misused [ 63 ], that they would be experimented on [ 52 ], and that tests were offered to make money rather than out of necessity [ 52 , 61 ]. Medical mistrust among some African Americans was reported to stem from negative events in history [ 52 ] and the “Tuskegee effect” relating to the mistreatment of African American men in the Tuskegee Syphilis study [ 35 ]. Concerns regarding confidentiality, what would happen with genetic test results after testing, and concerns about private information being discovered by others without the donor’s permission were also highlighted as important issues [ 52 , 61 , 65 ].

“…I felt in the past that doctors have sent me for tests just to get money. I feel that I was put through something…really bad, going you know, put fear in me for something…just so the doctor could put the claim in and I would never trust. It was terrible…” Hispanic female, Sussner et al. [ 61 ]
“One of my girlfriends who is so narrow-minded … would say ‘I wouldn’t let those people experiment [referring to genetic testing] on me’ …. Some of my other friends, who are not as narrowminded would think, ‘It’s best to find out all you can’… ‘Go for all the tests you can’ ” African American female, Glenn et al. [ 63 ]

Only one study, involving Asian American participants, suggested that a cultural mismatch exists between genetic testing and their culture’s traditional medicine [ 63 ]. However, scepticism of genetic testing was identified amongst Hispanic participants who were unsure of the need for genetic testing and screening without having symptoms [ 66 ], and others who felt there was no point in knowing [ 67 ]. Some African Americans were unsure of the test accuracy or reliability for their ethnic group [ 35 , 52 ]. Ford et al. [ 65 ] also reported that African American women who did not attend genetic counselling perceived few benefits.

The theme of Family was raised in discussion by many participants and was described as both a motivator and a barrier to testing. Two studies reported that Hispanics identified a culture of prioritising family needs over their own, resulting in a lack of time to attend health appointments [ 61 , 63 ] and therefore a potential barrier to attending genetic counselling or testing.

“The woman is meant to do everything, take care of the house, take care of this, take care of that, so health does go on the background…That’s how we were raised. You last, everybody else first” Hispanic female, Sussner et al. [ 61 ]

However, helping others in society [ 60 , 63 ] and family [ 35 , 52 , 60 , 61 , 63 , 66 ] was cited as an important benefit and motivator for genetic testing across most groups. Some Hispanic participants also felt it was important that they act as role models by making health more of a priority and encouraging older women to attend genetic counselling [ 61 ].

“I was the first one diagnosed with breast cancer in my family, so I would be concerned about it for their sake to find out what was going on…so that’s what would motivate me, family.” African American female, Sheppard et al. [ 60 ]

Only two studies referred to family support in relation to genetic testing [ 61 , 65 ]. Ford et al. [ 65 ] reported that some African American women felt they lacked family support to attend genetic counselling for cancer risk due to the potential negative impact of finding out they were high risk. Sussner et al. [ 61 ] reported that Hispanic participants felt family support would motivate them to attend genetic counselling. Both groups acknowledged that the decision was ultimately theirs; therefore, a lack of family support might not prevent individuals from having genetic counselling or testing.

“[Family] wouldn’t keep me from going. If I wanted to go, I would go” African American female, Ford et al. [ 65 ]

A further issue specific to some Chinese Australians was their understanding or definition of ‘close relatives’ [ 62 , 64 ]. These studies highlighted that, due to the importance placed on the paternal line, maternal family members might not always be considered ‘close relatives.’ Furthermore, when female relatives previously considered close relatives marry into other families, they may no longer be considered as such. Some indicated that relatives would be considered ‘close’ depending on how much they interacted with them rather than by bloodline [ 64 ]. These issues could lead to omissions in the assessment of cancer family history, resulting in a less accurate perception of risk.

Service provision and access

Several practical issues were highlighted in relation to uptake of cancer genetic testing. Ease of access to services would motivate attendance for genetic testing [ 60 ]. This could be achieved by providing services in local communities and during afternoons or weekends [ 35 ]. Physician recommendation was also an important factor in accessing services. Participants in four studies including African Americans, Hispanics, and Chinese Australians indicated that physician recommendation or referral would be a motivator to testing [ 35 , 61 , 64 , 67 ], whilst needing yet not receiving a primary care referral would be a barrier [ 60 , 61 , 65 , 67 ].

“The doctor mentioned it during my last visit but he didn’t do anything about it so I didn’t do anything.” Hispanic female, Vadaparampil et al. [ 67 ]

As the majority of the qualitative studies took place in the US, cost and insurance were often discussed [ 35 , 52 , 60 , 61 , 63 , 65 , 66 , 67 ]. Participants were unsure whether insurance would cover the costs of testing [ 67 ] and were unwilling or unable to pay themselves if testing was expensive or not covered by their insurance [ 35 ]. Both African Americans [ 65 ] and Hispanics voiced concerns that the cost of testing would be a particular barrier within their ethnic group due to low wages and not having health insurance [ 66 ]. Some African American participants who were affected by cancer indicated that genetic testing was too important not to have due to financial cost [ 60 ].

“I think if I had to incur the fees myself, I wouldn’t think of going…I can’t even imagine having to pay out of my pocket for tests.” African American female, Ford et al. [ 65 ]

Language was identified as a barrier to using genetic services by Hispanics in two studies [ 61 , 66 ]. Participants were concerned that poor communication, due also to physicians’ use of complex medical language, could have been a reason for their lack of a referral. However, some indicated that the use of a translator, including family members, had been helpful [ 61 ].

“But if you are let’s say Spanish and you don’t know the terminology or you don’t know, they (the doctors) think the person is ignorant. That’s not the case because just because you have a language barrier does not make you ignorant.” Hispanic female, Sussner et al. [ 61 ]

All studies scored quite highly in the quality assessment, receiving scores of 0.75 or above (range 0.75–1.0, see Tables 1 and 2 ). [Additional file 3 presents the included studies in order of quality with comments on where the study failed to meet the quality criteria.] The quantitative studies scored well for sufficiently describing the study objective, design, method of subject selection, and sample characteristics. Where issues did arise, these were mainly due to shortcomings in analytic methods, for example failing to report sufficient estimates of variance, and having missing data that was not sufficiently acknowledged. Six of the included quantitative studies involved pilot research [ 36 , 38 , 46 , 47 , 48 , 58 ]; however, this did not necessarily effect the study quality score, for example if it was stated that the analyses were exploratory the criteria of needing to have controlled for confounding was not applicable. The qualitative studies scored highly for sufficiently describing the study objective, design and context. However, none reflected on how the personal characteristics of the researchers may have impacted upon the analysis and results.

The current review supports and adds to the findings of previous work by Allford et al. [ 22 ], particularly due to the inclusion of several studies with Hispanic participants. The results highlight that only a small proportion of African American, Hispanic, and Asian ethnic minority groups are aware and knowledgeable about genetic testing for cancer susceptibility, and whilst attitudes are quite positive, concerns and negative perceptions exist. Practical issues are also highlighted as potential barriers to obtaining genetic counselling or testing.

Awareness of genetic testing for cancer risk varied across the reviewed studies; however, as recently as 2014, less than 30% of a general population sample of African Americans and Hispanics were aware [ 30 ]. The review also found some evidence that more White Americans are aware of genetic testing than Asian American, African American, and Hispanics, and this is further supported by research which did not primarily aim to investigate ethnic minority groups [ 68 , 69 ]. Importantly, overall awareness of genetic testing for cancer risk was low across all ethnic groups, including White participants, highlighting that within the general population it is not a well-known health service [ 29 , 30 , 32 ].

Ethnic minority groups’ knowledge of genetic risk of cancer also varied across the studies reviewed, indicating low to moderate levels of knowledge. Among samples of participants at high risk for cancer, often less than 50% of knowledge questions were answered correctly. Although knowledge about cancer genetics may not relate to intentions to test, individuals who lack understanding about the hereditary nature of cancer (e.g. that BRCA gene mutations can be passed on by male as well as female relatives) may not pursue genetic testing that is relevant to them. Low awareness and knowledge was also highlighted in qualitative research as a barrier to attending genetic counselling or undergoing genetic testing, and participants suggested a need for awareness and educational interventions especially emphasising the importance of family history of cancer. Qualitative studies also identified several misconceptions about cancer that could deter patients from attending cancer genetic services. However, it is difficult to make generalisations based on the results from qualitative research which may only represent the views of a few individuals. Furthermore, these misconceptions do not appear to be unique to those from ethnic minority groups. For example, Ford et al. [ 65 ] report that White women who did not attend genetic counselling discussed the belief that talking about breast cancer could increase their risk; however, this misconception does not appear to have been referred to by African American women in the study. A study of White and Hispanic women found that both groups held misconceptions, including that consuming sugar substitutes and bruising from being hit can cause cancer; but such beliefs were found to be significantly more prevalent in Hispanic women [ 70 ]. Similarly, in a more recent UK study, whilst both White and South Asian participants believed that stress causes cancer, significantly more South Asians held this belief and a small minority (4.3%) also believed that cancer is contagious, a misconception that no White participants held [ 71 ].

Personal health benefits and the ability to provide information about cancer risk to the family were recognised as important benefits of genetic susceptibility testing. Family support was also discussed as a potential motivator to testing for some Hispanics, suggesting that interventions encouraging Hispanics to attend cancer genetic services could benefit from involving family. Finally, easy access to genetic services is important to facilitate uptake of genetic counselling/testing, and several qualitative studies found that a physician recommendation was a motivator for testing. Physician education on the hereditary basis of some cancers and the importance of referring people with a family history of cancer for genetic counselling is also important to ensure thorough personal risk assessment and that all patients who could benefit from testing are offered this.

Several negative attitudes and perceptions that might influence uptake of genetic susceptibility testing were found to exist among ethnic minority groups, including reluctance to talk about cancer amongst family, concerns about stigma, and concerns about the emotional reaction of undergoing genetic testing and receiving a high risk result. Whilst cancer stigma is not unique to ethnic minorities [ 72 ], it is unclear to what extent such perceived stigma and secrecy varies across different ethnic groups. Research has also found that cancer stigma varies in magnitude depending on the type of cancer [ 73 ]. As cancer stigma not only has negative consequences for patients’ psychological well-being, and can also discourage individuals from seeking help from healthcare services [ 74 ], it seems important that efforts are made to reduce stigma to encourage individuals to attend cancer genetic services.

Other perceptions and issues identified in this review appear to be more specific to certain groups. For example, perceptions of ‘close relatives’ among Chinese Australian communities could hinder risk assessment based on descriptions of family history of cancer, and Hispanic women’s culture of prioritising family over their own health could influence their behaviours relating to genetic testing and cancer risk. These various attitudes and perceptions highlight the importance of identifying the barriers and issues that are specific to different ethnic groups and the need to tailor interventions to address them.

It is particularly important to recognise that historical experiences of ethnic minority groups in different countries may result in varied attitudes towards healthcare and services such as genetic testing. There is well documented mistrust of medical research and the healthcare system among some African Americans, influenced, for example, by the Tuskegee Syphilis Study in which African American males with Syphilis were misled by researchers and refused treatment in order to study the progression of the illness [ 75 ]. Practical issues, including cost and the need for health insurance to receive testing, and concerns about insurance discrimination as a result of testing are also more relevant to some American ethnic minorities. The Genetic Information Non-discrimination Act (GINA) was signed into US law in 2008 making it illegal for health insurers and employers to discriminate against individuals on the basis of genetic information and prohibiting insurers from requiring a person to have testing or to provide genetic information. However, the act is limited as it does not cover disability, long-term care, or life insurance. Furthermore, two surveys carried out in 2009/2010 in the US, one with women at increased risk of hereditary breast/ovarian cancer [ 76 ] and one with family physicians [ 77 ], found that over half had not been aware of GINA before taking part in the survey. Considering the lack of awareness of GINA, even amongst those that might be expected to have an interest in it, and due to the limitations of the act, it is perhaps unsurprising that concerns of discrimination by insurers continues to be cited as a limitation or barrier to genetic testing in the US. In addition, several of the studies included in this review were conducted before GINA was brought into law.

Genetic testing may identify a gene change whose association with disease is not known, a ‘variant of unknown significance’ (VUS). Unlike a pathogenic high risk gene change, the association between a VUS and increased risk of cancer is unclear, leading to issues for providers (counselling, risk management and preventative care) [ 78 ] and patients (confusion, psychological impact). It has previously been estimated that approximately 10% of Caucasians undergoing BRCA genetic testing receive a VUS result [ 79 ], with a much higher proportion in Hispanics (9–23%) and African Americans (16–24%) [ 80 , 81 , 82 ]. Different prevalence rates of VUS according to ethnic group lead to differences in the ease of interpretation of results between groups. That public databases of US gene testing will have far fewer non-Whites on them than Whites contributes to the problem of interpretation. Many VUS in non-Whites will be poorly characterised compared with Whites simply because less data is available on them. The greater likelihood of receiving an uninformative VUS result in African ethnic groups might potentially contribute to a lack of understanding about genetic testing or its perceived benefits, but our review provides no evidence to support this hypothesis.

The majority of the reviewed studies were conducted in the US which limits their generalisability to other countries, particularly in relation to levels of awareness and knowledge. Nevertheless, lack of awareness, the taboo of discussing cancer, and language difficulties, have also been reported in a UK genetics service evaluation [ 83 ] and a Genetics Alliance UK report involving interviews with individuals from ethnic minorities [ 84 ]. The evaluation of a UK pilot genetic service within a culturally diverse society by Atkin et al. [ 83 ] reports that poor communication was a barrier for both White and South Asian patients; however, communication issues were most difficult for those whose first language was not English and language support was not always viewed positively. Furthermore, some female patients reported that discussing breast cancer with male practitioners was embarrassing and even felt shameful. The article highlights the benefit of employing multilingual, culturally sensitive workers within the clinic [ 83 ]. In other research, cancer fatalism has been found to be higher amongst ethnic minority women in the UK than White women, whilst cancer fear appears to vary by ethnic group [ 85 ]. Further research into the attitudes and perceptions of ethnic minority groups in the UK and other European countries is needed in order to gain a better understanding of specific cultural barriers to cancer susceptibility testing.

The review has a number of limitations including the omission of intervention studies aiming to increase participant knowledge of cancer genetics from which we may have been able to extract baseline data. We did not include unpublished literature and our experience of using the quality assessment tool, specifically chosen for its ability to assess both qualitative and quantitative studies, indicated that it lacked a finer discriminatory capacity. The heterogeneous nature of the measures of awareness, knowledge, and attitudes used across various ethnic groups and different cancer types in the quantitative studies, made a meta-analysis impossible and interpretation of results difficult. Furthermore, the majority of included studies measuring knowledge and attitudes included only at risk patients or those already with a cancer diagnosis; therefore, results may not represent the knowledge or attitudes of the general population of ethnic minorities. Whilst still in the early stages of conception, due to an increasing understanding of cancer genetics and decreases in the cost of genetic testing, population based risk stratified interventions for cancer including genetic susceptibility testing are likely to be introduced in the future. Such interventions would provide people with information on their risk of cancer, and screening or treatment would be recommended based on their estimated level of risk. Before such interventions become a reality it is necessary to gain an understanding of the attitudes of the wider population, including those who are currently underserved such as ethnic minority groups. Meisel et al. [ 86 ] found positive attitudes towards population-based genetic testing and risk stratified screening for ovarian cancer amongst a sample of women from the UK general population. However, just 34% of the sample interviewed were non-White participants and cultural/ethnic differences in attitudes and perceptions were not discussed in detail.

Conclusions

Widening public participation amongst ethnic minority groups in future cancer risk prediction programmes may be enabled by culturally sensitive knowledge and awareness raising interventions that decrease the stigma and taboo of cancer. For this, further international research is needed to provide clearer information of ethnic minorities’ attitudes, and information and support needs so that cancer risk prediction programmes are inclusive and effective in the prevention and early detection of cancer.

Abbreviations

Genetic Information Non-discrimination Act

United Kingdom

United States

Variant of unknown significance

Garber JE, Offit K. Hereditary cancer predisposition syndromes. J Clin Oncol. 2005;23(2):276–92.

Article   PubMed   Google Scholar  

Jasperson KW, Tuohy TM, Neklason DW, Burt RW. Hereditary and familial colon cancer. Gastroenterology. 2010;138(6):2044–58.

Article   CAS   PubMed   PubMed Central   Google Scholar  

McPhearson K, Steel C, Dixon J. Breast cancer-epidemiology, risk factors and genetics. Br Med J. 2000;321:624–8.

Article   Google Scholar  

Miki Y, Swensen J, Shattuck-Eidens D, Futreal PA, Harshman K, Tavtigian S, Liu Q, Cochran C, Bennett LM, Ding W. A strong candidate for the breast and ovarian cancer susceptibility gene BRCA1. Science. 1994;266(5182):66–71.

Article   CAS   PubMed   Google Scholar  

Edwards SM, Kote-Jarai Z, Meitz J, Hamoudi R, Hope Q, Osin P, Jackson R, Southgate C, Singh R, Falconer A, et al. Two percent of men with early-onset prostate cancer harbor germline mutations in the BRCA2 gene. Am J Human Genet. 2003;72(1):1–12.

Article   CAS   Google Scholar  

Axilbund JE, Peshkin BN. Hereditary cancer risk. US: Springer; 2010.

Book   Google Scholar  

Etchegary H. Public attitudes toward genetic risk testing and its role in healthcare. Personalized Med. 2014;11(5):509–22.

Henneman L, Vermeulen E, van El CG, Claassen L, Timmermans DR, Cornel MC. Public attitudes towards genetic testing revisited: comparing opinions between 2002 and 2010. Eur J Hum Genet. 2013;21(8):793–9.

Haga SB, Barry WT, Mills R, Ginsburg GS, Svetkey L, Sullivan J, Willard HF. Public knowledge of and attitudes toward genetics and genetic testing. Genet Test Mol Biomarkers. 2013;17(4):327–35.

Article   PubMed   PubMed Central   Google Scholar  

Jack RH, Davies EA, Møller H. Breast cancer incidence, stage, treatment and survival in ethnic groups in South East England. Br J Cancer. 2009;100(3):545–50.

Odedina FT, Akinremi TO, Chinegwundoh F, Roberts R, Yu D, Reams RR, Freedman ML, Rivers B, Green BL, Kumar N. Prostate cancer disparities in black men of African descent: a comparative literature review of prostate cancer burden among black men in the United States, Caribbean, United Kingdom, and West Africa. Infect Agents Cancer. 2009;4(1):S2.

Siegel RL, Miller KD, Jemal A. Cancer statistics, 2015. CA Cancer J Clin. 2015;65(1):5–29.

Siegel RL, Fedewa SA, Miller KD, Goding-Sauer A, Pinheiro PS, Martinez-Tyson D, Jemal A. Cancer statistics for Hispanics/Latinos, 2015. CA Cancer J Clin. 2015;65(6):457–80.

Gomez SL, Noone AM, Lichtensztajn DY, Scoppa S, Gibson JT, Liu L, Morris C, Kwong S, Fish K, Wilkens LR, et al. Cancer incidence trends among Asian American populations in the United States, 1990–2008. J Nat Cancer Inst. 2013;105(15):1096–110.

Godard B, Kaariainen H, Kristoffersson U, Tranebjaerg L, Coviello D, Ayme S. Provision of genetic services in Europe: current practices and issues. Eur J Hum Genet. 2003;11(Suppl 2):S13–48.

Hall MJ, Olopade OI. Disparities in genetic testing: thinking outside the BRCA box. J Clin Oncol. 2006;24(14):2197–203.

Hensley Alford S, McBride CM, Reid RJ, Larson EB, Baxevanis AD, Brody LC. Participation in genetic testing research varies by social group. Public Health Genomics. 2011;14(2):85–93.

van Riel E, van Dulmen S, Ausems MGEM. Who is being referred to cancer genetic counseling? Characteristics of counselees and their referral. J Community Genet. 2012;3(4):265–74.

Wonderling D, Hopwood P, Cull A, Douglas F, Watson M, DBurn J, McPherson K. A descriptive study of UK cancer genetics services: an emerging clinical response to the new genetics. Br J Cancer. 2001;85(2):166–70.

Armstrong K, Weber B, Stopfer J, Calzone K, Putt M, Coyne J, Schwartz JS. Early use of clinical BRCA1/2 testing: associations with race and breast cancer risk. Am J Med Genet. 2003;117A(2):154–60.

Levy DE, Byfield SD, Comstock CB, Garber JE, Syngal S, Crown WH, Shields AE. Underutilization of BRCA1/2 testing to guide breast cancer treatment: black and Hispanic women particularly at risk. Genet Med. 2011;13(4):349–55.

Allford A, Qureshi N, Barwell J, Lewis C, Kai J. What hinders minority ethnic access to cancer genetics services and what may help? Eur J Hum Genet. 2014;22(7):866–74.

Lee SC, Bernhardt BA, Helzlsouer KJ. Utilization of BRCA1/2 genetic testing in the clinical setting: report from a single institution. Cancer. 2002;94(6):1876–85.

Armstrong K, Calzone K, Stopfer J, Fitzgerald G, Coyne J, Weber B. Factors associated with decisions about clinical BRCA1/2 testing. Cancer Epidemiol Biomark Prev. 2000;9(11):1251–4.

CAS   Google Scholar  

Thomas J, Harden A. Methods for the thematic synthesis of qualitative research in systematic reviews. BMC Med Res Methodol. 2008;8:45.

Kmet LM, Lee RC, Cook LS. Standard quality assessment criteria for evaluating primary research papers from a variety of fields. 2004:1-22.

Gammon AD, Rothwell E, Simmons R, Lowery JT, Ballinger L, Hill DA, Boucher KM, Kinney AY. Awareness and preferences regarding BRCA1/2 genetic counseling and testing among Latinas and non-Latina white women at increased risk for hereditary breast and ovarian cancer. J Genet Couns. 2011;20(6):625–38.

Heck JE, Franco R, Jurkowski JM, Sheinfeld Gorin S. Awareness of genetic testing for cancer among United States Hispanics: the role of acculturation. J Community Genet. 2008;11(1):36–42.

Google Scholar  

Honda K. Who gets the information about genetic testing for cancer risk? The role of race/ethnicity, immigration status, and primary care clinicians. Clin Genet. 2003;64:131–6.

Huang H, Apouey B, Andrews J. Racial and ethnic disparities in awareness of cancer genetic testing among online users: internet use, health knowledge, and socio-demographic correlates. J Consum Health Internet. 2014;18(1):15–30.

Kaplan CP, Haas JS, Perez-Stable EJ, Gregorich SE, Somkin C, Des Jarlais G, Kerlikowske K. Breast cancer risk reduction options: awareness, discussion, and use among women from four ethnic groups. Cancer Epidemiol Biomark Prev. 2006;15(1):162–6.

Pagan JA, Su D, Li L, Armstrong K, Asch DA. Racial and ethnic disparities in awareness of genetic testing for cancer risk. Am J Prev Med. 2009;37(6):524–30.

Peters N, Rose A, Armstrong K. The association between race and attitudes about predictive genetic testing. Cancer Epidemiol Biomark Prev. 2004;13(3):361–5.

Vadaparampil ST, Wideroff L, Breen N, Trapido E. The impact of acculturation on awareness of genetic testing for increased cancer risk among Hispanics in the year 2000 National Health Interview Survey. Cancer Epidemiol Biomark Prev. 2006;15(4):618–23.

Adams I, Christopher J, Williams KP, Sheppard VB. What black women know and want to know about counseling and testing for BRCA1/2. J Cancer Educ. 2015;30(2):344–52.

Ramirez AG, Aparicio-Ting FE, de Majors SS, Miller AR. Interest, awareness, and perceptions of genetic testing among Hispanic family members of breast cancer survivors. Ethn Dis. 2006;16(2):398–403.

PubMed   Google Scholar  

Satia JA, McRitchie S, Kupper LL, Halbert CH. Genetic testing for colon cancer among African-Americans in North Carolina. Prev Med. 2006;42(1):51–9.

Sussner KM, Jandorf L, Thompson HS, Valdimarsdottir HB. Interest and beliefs about BRCA genetic counseling among at-risk Latinas in new York City. J Genet Couns. 2010;19(3):255–68.

Sussner KM, Thompson HS, Valdimarsdottir HB, Redd WH, Jandorf L. Acculturation and familiarity with, attitudes towards and beliefs about genetic testing for cancer risk within Latinas in East Harlem, new York City. J Genet Couns. 2009;18(1):60–71.

Thompson HS, Valdimarsdottir HB, Jandorf L, Redd W. Perceived disadvantages and concerns about abuses of genetic testing for cancer risk: differences across African American, Latina and Caucasian women. Patient Educ Couns. 2003;51(3):217–27.

Sussner KM, Jandorf L, Thompson HS, Valdimarsdottir HB. Barriers and facilitators to BRCA genetic counseling among at-risk Latinas in new York City. Psycho-Oncology. 2013;22(7):1594–604.

Donovan KA, Tucker DC. Knowledge about genetic risk for breast cancer and perceptions of genetic testing in a sociodemographically diverse sample. J Behav Med. 2000;23(3):15–36.

Kinney AY, Croyle RT, Dudley WN, Bailey CA, Pelias MK, Neuhausen SL. Knowledge, attitudes, and interest in breast-ovarian cancer gene testing: a survey of a large African-American kindred with a BRCA1 mutation. Prev Med. 2001;33(6):543–51.

Kinney AY, Simonsen SE, Baty BJ, Mandal D, Neuhausen SL, Seggar K, Holubkov R, Smith K. Acceptance of genetic testing for hereditary breast ovarian cancer among study enrollees from an African American kindred. Am J Med Genet . 2006;140(8):813–26.

Thompson HS, Valdimarsdottir HB, Duteau-Buck C, Guevarra J, Dana H, Bovbjerg DC, Richmond-Avellaneda C, Amarel D, Godfrey D, Brown K, Offit K. Psychosocial predictors of BRCA counseling and testing decisions among urban African-American women. Cancer Epidemiol Biomark Prev. 2002;11:1579–85.

Weinrich S, Vijayakumar S, Powell IJ, Priest J, Hamner CA, McCloud L, Pettaway C. Knowledge of hereditary prostate cancer among high-risk African American men. Oncol Nurs Forum. 2007;34(4):845.

Vadaparampil ST, Quinn GP, Dutil J, Puig M, Malo TL, McIntyre J, Perales R, August EM, Closser Z. A pilot study of knowledge and interest of genetic counseling and testing for hereditary breast and ovarian cancer syndrome among Puerto Rican women. J Community Genet. 2011;2(4):211–21.

Vadaparampil ST, Quinn GP, Small BJ, McIntyre J, Loi CA, Closser Z, Gwede CK. 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. 2010;14(1):99–106.

Edwards TA, Thompson HS, Kwate NO, Brown K, McGovern MM, Forman A, Kapil-Pair N, Jandorf L, Bovbjerg DH, Valdimarsdottir HB. Association between temporal orientation and attitudes about BRCA1/2 testing among women of African descent with family histories of breast cancer. Patient Educ Couns. 2008;72(2):276–82.

Kessler L, Collier A, Brewster K, Smith C, Weathers B, Wileyto EP, Halbert CH. Attitudes about genetic testing and genetic testing intentions in African American women at increased risk for hereditary breast cancer. Genet Med. 2005;7(4):230–8.

Sussner KM, Edwards TA, Thompson HS, Jandorf L, Kwate NO, Forman A, Brown K, Kapil-Pair N, Bovbjerg DH, Schwartz MD, et al. Ethnic, racial and cultural identity and perceived benefits and barriers related to genetic testing for breast cancer among at-risk women of African descent in new York City. Public Health Genomics. 2011;14(6):356–70.

Matthews AK, Cummings S, Thompson S, Wohl V, List M, Olopade OI. Genetic testing of African Americans for susceptibility to inherited cancers: use of focus groups to determine factors contributing to participation. J Psychosoc Oncol. 2000;18(2):1–19.

Armstrong K, Micco E, Carney A, Stopfer J, Putt M. Racial differences in the use of BRCA1/2 testing among women with a family history of breast or ovarian cancer. J Am Med Assoc. 2005;293(14):1729–36.

Myers RE, Hyslop T, Jennings-Dozier K, Wolf TA, Burgh DY, Diehl JA, Caryn Lerman C, Chodak GW. Intention to be tested for prostate cancer risk among African-American men. Cancer Epidemiol Biomark Prev. 2000;9:1323–8.

Weinrich S, Royal C, Pettaway CA, Dunston G, Faison-Smith L, Priest JH, Roberson-Smith P, Frost J, Jenkins J, Brooks KA, et al. Interest in genetic prostate cancer susceptibility testing among African American men. Cancer Nurs. 2002;25(1):28–34.

Sussner KM, Thompson HS, Jandorf L, Edwards TA, Forman A, Brown K, Kapil-Pair N, Bovbjerg DH, Schwartz MD, Valdimarsdottir HB. The influence of acculturation and breast cancer-specific distress on perceived barriers to genetic testing for breast cancer among women of African descent. Psycho Oncol. 2009;18(9):945–55.

Armstrong K, Putt M, Halbert CH, Grande D, Schwartz JS, Liao K, Marcus N, Demeter MB, Shea J. The influence of health care policies and health care system distrust on willingness to undergo genetic testing. Med Care. 2012;50(5):381–7.

McBride CM, Lipkus IM, Jolly D, Lyna P. Interest in testing for genetic susceptibility to lung cancer among black college students “at risk” of becoming cigarette smokers. Cancer Epidemiol Biomark Prev. 2005;14(12):2978–81.

Hughes C, Fasaye G-A, LaSalle VH, Finch C. Sociocultural influences on participation in genetic risk assessment and testing among African American women. Patient Educ Couns. 2003;51(2):107–14.

Sheppard VB, Graves KD, Christopher J, Hurtado-de-Mendoza A, Talley C, Williams KP. African American women's limited knowledge and experiences with genetic counseling for hereditary breast cancer. J Genet Couns. 2014;23(3):311–22.

Sussner KM, Edwards T, Villagra C, Rodriguez MC, Thompson HS, Jandorf L, Valdimarsdottir HB. BRCA genetic counseling among at-risk Latinas in new York City: new beliefs shape new generation. J Genet Couns. 2015;24(1):134–48.

Eisenbruch M, Yeo SS, Meiser B, Goldstein D, Tucker K, Barlow-Stewart K. Optimising clinical practice in cancer genetics with cultural competence: lessons to be learned from ethnographic research with Chinese-Australians. Soc Sci Med. 2004;59(2):235–48.

Glenn BA, Chawla N, Bastani R. Barriers to genetic testing for breast cancer risk among ethnic minority women: an exploratory study. Ethn Dis. 2012;22(3):267–73.

Barlow-Stewart K, Yeo SS, Meiser B, Goldstein D, Tucker K, Eisenbruch M. Toward cultural competence in cancer genetic counseling and genetics education: lessons learned from Chinese-Australians. Genet Med. 2006;8(8):24–32.

Ford ME, Alford SH, Britton D, McClary B, Gordon HS. Factors influencing perceptions of breast cancer genetic counseling among women in an urban health care system. J Genet Couns. 2007;16(6):735–53.

Kinney AY, Gammon A, Coxworth J, Simonsen SE, Arce-Laretta M. Exploring attitudes, beliefs, and communication preferences of Latino community members regarding BRCA1/2 mutation testing and preventive strategies. Genet Med. 2010;12(2):105–15.

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. 2010;19(6):618–29.

Lacour RA, Daniels MS, Westin SN, Meyer LA, Burke CC, Burns KA, Kurian S, Webb NF, Pustilnik TB, Lu KH. What women with ovarian cancer think and know about genetic testing. Gynecol Oncol. 2008;111(1):132–6.

Mai PL, Vadaparampil ST, Breen N, McNeel TS, Wideroff L, Graubard BI. Awareness of cancer susceptibility genetic testing: the 2000, 2005, and 2010 National Health Interview Surveys. Am J Prev Med. 2014;46(5):440–8.

Pérez-Stable EJ, Sabogal F, Otero-Sabogal R, Hiatt RA, McPhee SJ. Misconceptions about cancer among Latinos and Anglos. JAMA. 1992;268(22):3219–23.

Lord K, Mitchell AJ, Ibrahim K, Kumar S, Rudd N, Symonds P. The beliefs and knowledge of patients newly diagnosed with cancer in a UK ethnically diverse population. Clin Oncol. 2012;24(1):4–12.

Daher M. Cultural beliefs and values in cancer patients. Ann Oncol. 2012;23(Suppl 3):66–9.

Marlow LAV, Waller J, Wardle J. Does lung cancer attract greater stigma than other cancer types? Lung Cancer. 2015;88(1):104–7.

Fujisawa D, Hagiwara N. Cancer stigma and its health consequences. Current Breast Cancer Reports. 2015;7(3):143–50.

Gamble VN. Under the shadow of Tuskegee: African Americans and health care. Am J Public Health. 1997;87(11):1773–8.

Allain DC, Friedman S, Senter L. Consumer awareness and attitudes about insurance discrimination post enactment of the genetic information nondiscrimination act. Familial Cancer. 2012;11:637–44.

Laedtke AL, O’Neill SM, Rubinstein WS, Vogel KJ. Family physicians’ awareness and knowledge of the genetic information non-discrimination act (GINA ) . J Genet Couns. 2012;21:345–52.

Eccles D, Mitchell G, Monteiro ANA, Schmutzler R, Couch FJ, Spurdle AB, Gómez-García EB, Driessen R, Lindor NM, Blok MJ, et al. BRCA1 and BRCA2 genetic testing–pitfalls and recommendations for managing variants of uncertain clinical significance. Ann Oncol. 2015;10:2057–65. doi: 10.1093/annonc/mdv278 .

Frank TS, Critchfield GC. Hereditary risk of women’s cancers. Best Practice Res Clin Obstet Gynaecol. 2002;16(5):703–13.

Kurian AW. BRCA1 and BRCA2 mutations across race and ethnicity: distribution and clinical implications. Curr Opin Obstet Gynecol. 2010;22(1):72–8.

Nanda R, Schumm LP, Cummings S, Fackenthal JD, Sveen L, Ademuyiwa F, Cobleigh M, Esserman L, Lindor NM, Neuhausen SL, et al. Genetic testing in an ethnically diverse cohort of high-risk women a comparative analysis of BRCA1 and BRCA2 mutations in American families of European and African ancestry. J Am Med Assoc. 2005;294(15):1925–33.

Weitzel JN, Lagos V, Blazer KR, Nelson R, Ricker C, Herzog J, McGuire C, Neuhausen S. Prevalence of BRCA mutations and founder effect in high-risk Hispanic families. Cancer Epidemiol Biomark Prev. 2005;14(7):1666–71.

Atkin K, Ali N, Chu CE. The politics of difference? Providing a cancer genetics service in a culturally and linguistically diverse society. Divers Health Care. 2009;6(3):149–57.

Genetic Alliance UK. Identifying family risk of cancer: Why is this much more difficult for ethnic minority communities and what would help. 2012.

Vrinten C, Wardle J, Marlow LA. Cancer fear and fatalism among ethnic minority women in the United Kingdom. Br J Cancer. 2016;114(5):597–604.

Meisel SF, Side L, Fraser L, Gessler S, Wardle J, Lanceley A. Population-based, risk-stratified genetic testing for ovarian cancer risk: a focus group study. Public Health Genomics. 2013;16:184–91.

Download references

Acknowledgements

This work was carried out at UCLH/UCL within the Cancer Theme of the NIHR UCLH/UCL Comprehensive Biomedical Research Centre supported by the UK Department of Health for the PROMISE study team.

Completion of this project was funded by the Eve Appeal and Cancer Research (grant code: UKC1005/A12677). The funders had no role in the study design; collection, management, analysis, or interpretation of data; writing of the report; or the decision to submit the report for publicationThis work was supported by.

Availability of data and materials

The data is included in the manuscript and tables.

Authors’ contribution

AL and KH designed the study. LF, JW and SS contributed to the design of the study. LS and SG provided design implementation support. KH conducted the literature search and carried out study selection with BR, MF and AL. KH and MF extracted the data and conducted the quality assessment of studies. KH drafted the manuscript which all authors critically reviewed and approved. All listed authors meet the criteria for authorship and no individual meeting these criteria has been omitted.

Authors’ information

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Ethics approval and consent to participate, publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Author information

Authors and affiliations.

Department of Women’s Cancer, EGA UCL Institute for Women’s Health, University College London, London, UK

Katie E. J. Hann, Lindsay Fraser, Belinda Rahman, Lucy Side, Sue Gessler & Anne Lanceley

Department of Behavioural Science and Health, Institute of Epidemiology and Health Care, University College London, London, UK

Madeleine Freeman, Jo Waller & Saskia C. Sanderson

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Anne Lanceley .

Additional files

Additional file 1:.

Review search terms. Full search terms for PsycInfo, CINAHL, Embase and MEDLINE used to find studies for the review. (DOCX 18 kb)

Additional file 2:

Study specific measures: Awareness, knowledge, and attitudes. The table presents descriptions of the main measures used to investigate awareness, knowledge and attitudes in relation to genetic counselling/testing. (DOCX 17 kb)

Additional file 3:

Rank order of included studies based on quality assessment. The table presents the quality assessments for each study and comments on why studies were marked down. (DOCX 19 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License ( http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Cite this article.

Hann, K.E.J., 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 , 503 (2017). https://doi.org/10.1186/s12889-017-4375-8

Download citation

Received : 23 August 2016

Accepted : 07 May 2017

Published : 25 May 2017

DOI : https://doi.org/10.1186/s12889-017-4375-8

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Genetic testing
  • Cancer risk
  • Ethnic minorities

BMC Public Health

ISSN: 1471-2458

thesis statement for genetic testing

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • Published: 21 March 2019

Rethinking the “open future” argument against predictive genetic testing of children

  • Jeremy R. Garrett PhD 1 , 2 ,
  • John D. Lantos MD 1 , 2 ,
  • Leslie G. Biesecker MD 3 ,
  • Janet E. Childerhose PhD 4 ,
  • Wendy K. Chung MD, PhD 5 ,
  • Ingrid A. Holm MD, MPH 6 ,
  • Barbara A. Koenig PhD 7 ,
  • Jean E. McEwen JD, PhD 3 ,
  • Benjamin S. Wilfond MD 8 , 9 &
  • Kyle Brothers MD, PhD 10

on behalf of the Clinical Sequencing Exploratory Research (CSER) Consortium Pediatrics Working Group

Genetics in Medicine volume  21 ,  pages 2190–2198 ( 2019 ) Cite this article

3167 Accesses

44 Citations

14 Altmetric

Metrics details

Professional consensus has traditionally discouraged predictive genetic testing when no childhood interventions can reduce future morbidity or mortality. However, advances in genome sequencing and accumulating evidence that children and families cope adequately with predictive genetic information have weakened this consensus. The primary argument remaining against testing appeals to children’s “right to an open future.” It claims that the autonomy of the future adult is violated when others make an irreversible choice to obtain or disclose predictive genetic information during childhood. We evaluate this argument and conclude that children’s interest in an open future should not be understood as a right . Rather an open future is one significant interest to weigh against other important interests when evaluating decisions. Thus, predictive genetic testing is ethically permissible in principle, as long as the interests promoted outweigh potential harms. We conclude by offering an expanded model of children’s interests that might be considered in such circumstances, and present two case analyses to illustrate how this framework better guides decisions about predictive genetic testing in pediatrics.

You have full access to this article via your institution.

Similar content being viewed by others

thesis statement for genetic testing

Parents’ attitudes towards research involving genome sequencing of their healthy children: a qualitative study

Toward the diagnosis of rare childhood genetic diseases: what do parents value most, parents, their children, whole exome sequencing and unsolicited findings: growing towards the child’s future autonomy, introduction.

For more than two decades, professional consensus has discouraged predictive genetic testing of children for adult-onset conditions (hereafter, predictive genetic testing) when no interventions in childhood might reduce morbidity or mortality. 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 More recently, this position has been extended to discourage disclosure of results produced by genome sequencing performed for other purposes. These professional recommendations are now being questioned. The ethical framework originally developed for single-gene testing requires re-evaluation given the ready availability of genome sequencing and the wealth of information it produces. Unlike single-gene testing, the use of genome sequencing to answer a clinical question can generate hundreds or even thousands of genetic findings. Many of these findings would not alter the care of the patient, and most fall well outside the indication for a test. Genome sequencing also has spurred studies of the psychosocial impacts of genetic information on pediatric patients and families that have deepened our understanding of those impacts. The time is ripe to re-examine the prevailing wisdom on this matter.

Historically, two primary ethical arguments have underpinned the consensus against predictive genetic testing in the position statements of leading professional societies. 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 One is empirically grounded and focused on the consequences of the decision. It warns of potential psychosocial harms resulting from children or parents learning about future risks for adult-onset conditions. However, accumulating evidence indicates such psychosocial harms are less common and less impactful than originally feared. Although important work is still ongoing, these preliminary data have begun to reduce concerns about the presumed psychological harms of this information. 9 , 10

The second argument focuses on the special moral status of children as future adults. It invokes children’s “right to an open future,” which purportedly is violated when irreversible choices are made for them during childhood. Numerous literature reviews and qualitative surveys indicate that this argument is perhaps the most frequently cited objection to predictive genetic testing among clinicians and bioethicists. 11 , 12 , 13 , 14 Moreover, unlike the concern about immediate psychosocial harms, the violation of a moral right is a value claim that cannot be refuted by empirical evidence. 13 Hence, this argument against predictive genetic testing likely will become even more important as the evidence base develops and shifts.

Given this evolution in justification for restricting predictive genetic testing, it is imperative to evaluate critically the conceptual basis for the right to an open future. We undertake this task by analyzing the concept of an open future and the nature of moral rights, both in theory and in the context of two clinical case scenarios. We argue that children’s interest in an open future can be protected adequately without imposing a strict ethical obligation to refrain from infringing that interest (i.e., a right ). Rather, an open future should be regarded as one significant interest to weigh against other important interests when determining whether testing or disclosure would provide more benefits than harm. 15 Considered this way, predictive genetic testing is ethically permissible in principle, as long as multiple important interests are considered and balanced. We conclude that this shift in ethical frameworks better guides decision-making about both genome sequencing and standard single-gene testing for children.

HISTORICAL BACKGROUND

The philosopher Joel Feinberg is generally credited with coining the phrase, “the child’s right to an open future.” He did so in the context of a legal case, Wisconsin v. Yoder , 16 which considered whether Amish communities should be exempted from compulsory school attendance laws. 17 In his argument, Feinberg highlighted a category of rights held primarily by children—so-called “rights-in-trust,” which “look like adult autonomy rights” but cannot yet be exercised in childhood. 17 Describing this class of rights further, Feinberg stated:

When sophisticated autonomy rights are attributed to children who are clearly not yet capable of exercising them, their names refer to rights that are to be saved for the child until he is an adult, but which can be violated “in advance,” so to speak, before the child is even in a position to exercise them… His right while he is still a child is to have these future options kept open until he is a fully formed, self-determining adult capable of deciding among them. 17

For Feinberg, the Amish violated the rights-in-trust of the children whose formal education was cut short. He argued that this practice left children prepared for few careers or lifestyles outside of an Amish farm and, thus, curtailed many potentially desirable future options. Importantly, this argument did not carry the day in Yoder . The court decided in favor of the Amish. Nonetheless, the right to an open future took hold in numerous ethical contexts. 18 , 19 , 20 , 21 , 22 , 23 , 24 , 25 , 26 , 27 , 28 , 29 , 30 , 31 , 32 , 33 , 34 , 35 , 36 , 37 , 38 , 39 , 40 , 41 , 42 , 43 , 44 , 45 , 46

The consensus against predictive genetic testing based on this right took shape in the early to mid-1990s. 47 , 48 Within the roughly 20-year period that followed, the right to an open future became prominently ensconced in the official position statements of many leading professional societies around the world 49 (Table  1 ). This restrictive consensus was endorsed and reinforced in the work of bioethicists as well. For example, Dena Davis, in both the 2001 and 2010 editions of her influential book, Genetic Dilemmas , argues that “[predictive genetic testing] is a decision each individual can make only for herself. Thus respect for the child’s right to an open future supports the growing consensus in the United States against allowing parents to choose such testing for their children.” 20 This interpretation of the right to an open future has, in turn, shaped many arguments against predictive genetic testing in the medical and bioethics literature and in clinical settings over the past two decades. 11 , 14 , 35 , 36 , 50 , 51 , 52 , 53 , 54 , 55 , 56 , 57 , 58 For example, in a 2017 article written for an audience of genetic counselors, Fenwick and colleagues summarize the status quo as follows: “The recommendations from these guidelines are well-established and have not changed significantly over time. Their primary message is that unless testing has current medical benefit, it should be deferred until a child is old enough to make her/his own decision protecting what Feinberg called the child’s right to an open future.” 14

CONCEPTUAL AND ETHICAL ANALYSIS

While the right to an open future has influenced pediatric bioethics significantly, we believe it is a principle that is apt to sow confusion and ambiguity rather than clarity.

First, what it means for a child’s future to be “open” or “closed” is not intuitively clear. Must future options be permanently inaccessible to be considered “closed”? Or is it sufficient that certain options are simply more difficult to access ? Few decisions made during childhood permanently foreclose some future possibility. However, many decisions (including many standard parenting decisions) make certain futures more or less difficult to realize. It is, to take Feinberg’s example, not literally impossible for Amish youth with abbreviated formal education to realize futures outside their community farm. 59 Indeed, a small proportion demonstrate this following their period of Rumspringa (the rite of passage for Amish teenagers preceding the choice between baptism within and separation from the Amish church), by deciding to leave the community and begin a new life. This first ambiguity, then, has real significance. If “closing” a child’s future requires making some future state of affairs literally impossible , then the right to an open future will have few applications within pediatrics or parenting more broadly. But if closing a child’s future refers to any decision that makes certain futures significantly more difficult , then this right will apply too widely, prohibiting many ordinary and even unavoidable parenting practices. 60

A second ambiguity arises regarding what the right to an open future is intended to protect precisely. We might view this right as requiring fiduciaries to ensure children have the fundamental resources to make decisions as adults. This could include basic capacities like the ability to reason from means to ends, as well as basic preferences , since adult decision-making requires mature desires and values. 61 Alternatively, we might determine specific skills , like learning a second language, to be vital to an open future. 61 Finally, we might instead focus on preserving specific options and opportunities , like becoming an Amish farmer or attending college. 61 Here again the choice of target matters. If the right to an open future merely requires preserving and developing children’s basic capacities , then few medical decisions, including predictive genetic testing, will seriously threaten this right. However, if the right forbids foreclosing specific options and opportunities , then it will be far too strong, implying that parents routinely violate their children’s open future by encouraging certain opportunities and discouraging others.

A third ambiguity relates to whether we should understand an open future quantitatively or qualitatively . Defenders of the right to an open future sometimes speak as if children have the right to “reach maturity with as many open options, opportunities, and advantages as possible” 17 and at other times as merely the right to reach maturity without a “radical narrowing” 18 of their options, opportunities, and advantages. However, defenders also sometimes focus less on the quantity of choices and more on their quality , entailing a right for children to reach maturity with certain vital options, opportunities, and advantages left open. Here again the choice of interpretation has important implications. A maximal ideal seems impossible to satisfy; parents make future-effecting choices for their children every day and cannot avoid doing so. A minimal threshold, on the other hand, is more defensible on its face, but also less relevant to pediatrics: few single decisions made for children, including predictive genetic testing, “radically narrow” their future. Focusing on the moral quality of specific choices seems more promising, but leaves us with the task of determining, in a nonarbitrary manner, which choices are vital to the adults whom children will become. And even there it is unclear whether parents can, or should, avoid shaping such choices for their children, especially when doing so serves other important interests. 62 , 63 , 64

Within the context of predictive genetic testing, these questions generate significant ambiguity. Obviously, any decision made for children in this context (to pursue testing or not) will affect their future. These decisions may even permanently close off certain options. Disclosing information about future risk does rule out the recipient not knowing that information. However, delaying testing/disclosure until adulthood is also a decision made for the child in childhood . Moreover, this decision to defer may close off important opportunities. We should emphasize that many test results alleviate anxiety and uncertainty, inform planning for a future health condition, enable children to begin—while in a stable and supportive environment—incorporating information into their developing identity and autonomy, or live out a limited lifespan in a way that prioritizes and maximizes what matters most to the child (i.e., meaningful time with family and friends, personal adventures, spiritual/religious pursuits, advocacy for political change or medical progress, and so on). 53 , 63 , 64 Either way, a decision must be made for the child in childhood and each opens some future options and closes others. 63 , 65 , 66

In light of these concerns, it seems inappropriate for health professionals to appeal to a “right” to an open future to encourage some decisions about predictive genetic testing, while fervently discouraging others. The right to an open future is traditionally regarded as a “negative claim-right.” 67 Within our common morality, negative claim-rights constitute the strongest ethical constraint on the actions of others. 68 This type of right protects an interest—in this case, the interest in having decisions deferred to make for and by one’s future self—by placing all moral agents, including parents and clinicians, under a strict obligation to refrain from infringing that interest. 67 Violations of a negative moral right are permitted rarely and only under strict conditions. The 1994 Institute of Medicine report, Assessing Genetic Risks , proposes four conditions that would need to be met to justifiably override an individual’s autonomy in the context of predictive genetic testing (Table  2 ) (ref. 1 ). Crucially, all four of these conditions will be met only in very rare circumstances. To assert that children have a negative moral “right” against such testing in childhood is to make a forceful moral claim, one that will override all other interests in nearly all realistic circumstances .

Even setting aside these more theoretical concerns, the vast potential for misinterpretation and misuse by itself provides sufficient reason to refrain from rights language here. In our experience, many practicing clinicians and scientists have felt compelled to adopt the right to an open future but lack a nuanced understanding of the concept; for them, it is no more permissible to violate this “right” than to violate a child’s right to life or bodily integrity. Indeed, in at least two significant cases, recent updates to professional guidelines noted that many clinicians have interpreted prior guidance as more prohibitive of predictive genetic testing than was intended or explicitly stated. 69 , 70 A significant factor contributing to this widespread misinterpretation is precisely the way in which concerns about the open future became ensconced within the language of rights .

Fortunately, the ethical construct of a “right” is unnecessary to identify and adequately protect the primary concern here—the child’s interest in an open future. Children are more likely, in general, to flourish when certain future options are left open. Most parents understand this. The notion that children have interests in an open future fits well with standard ethical guidance in pediatric bioethics. 71 The child’s interest in an open future is one important, but not automatically the most important, interest to consider and balance in the process of shared decision-making. 63 In other words, an open future is best understood not as a separate principle of pediatric bioethics, but instead as one component of its traditional focus on interests and balancing benefits and harms to children and families . 62 , 72 , 73 There is a robust literature on the interests of children in pediatric ethics. While a comprehensive examination of this literature is beyond the scope of this paper, examining one promising account of children’s interests helps demonstrate how an open future interest can be weighed alongside other interests. This account, developed in 2009 by Malek, 74 utilizes extant statements about children’s needs and interests to propose a list of 13 important interests that should be considered in pediatric decision-making (Table  3 ). Each interest is a capacity, activity, or state of affairs that contributes to the well-being of children, and most medical decisions will involve some tradeoffs among these interests. While an interest in an open future is not explicitly included in Malek’s list, we propose that it should be added, perhaps as one component of the interest in autonomy, i.e., an interest in preserving future autonomy.

A nuanced interest-based framework like this better serves pediatric clinical ethics than a rights-based approach for two important reasons (Table  4 ). First, unlike a rights-based approach, an interest-based framework is not rigidly committed to a predetermined conclusion and can nimbly incorporate and respond to an ever-evolving evidence base. Second, it enables a comprehensive and systematic, but also flexible and balanced, assessment of children’s many diverse interests. In particular, this framework facilitates better decisions about whether the full range of interests for any particular child are promoted more by opening or closing certain futures. The interest in preserving future autonomy is weighed alongside other interests identified by Malek, potentially supporting different conclusions in different circumstances. Foreclosing the opportunity to make certain choices later may promote a child’s overall interests in some circumstances. For example, many parental and pediatric decisions—like disclosing to an 11-year-old child that he was adopted or removing an infant’s supernumerary digit— could be delayed until the child reaches adulthood. However, while such delays might preserve one particular opportunity for autonomous choice, foreclosing those later decisions and proceeding with disclosure or surgery now will often open other opportunities that better serve the child’s overall interests. 64 In other circumstances, though, producing or preserving openness may be most compelling. Decisions like career choice are frequently viewed this way: parents typically take measures to keep their child’s future career options open, even though it might serve other interests to radically narrow this range earlier (e.g., strongly funneling them into a lucrative family business). Individual families working with their chosen care providers are best positioned to identify and balance competing interests in particular circumstances. 53 , 60 , 66 , 75 , 76 The presumed validity of a right to an open future has impeded that ideal approach for too long in too many pediatric contexts, including decisions about predictive genetic testing.

THE ADVANTAGES OF AN INTEREST-BASED FRAMEWORK

In what follows, we use case-based reasoning to contrast rights-based and interest-based approaches to children’s open future. Admittedly, we cannot consider all factors that are ethically relevant to decisions about predictive genetic testing, such as the developmental status of the child, medical conditions, or clinical contexts. Nonetheless, the two vignettes we analyze represent commonly-encountered cases and, together, illustrate the value of an interests-based framework.

Ariel and Lynch syndrome

Ariel is a ten-year-old girl with seizures but otherwise normal neurological development. Her neurologist specializes in the genetic basis of neurological conditions, and decided, after discussions with Ariel and her mother, to use genome sequencing to identify the cause of a seizure disorder, even though gene panels more narrowly focused on epilepsy are available. In analyzing the sequencing data, the neurologist determines that Ariel has a pathogenic variant in the MLH1 gene, which causes Lynch syndrome. Ariel has no contact with her father, and her mother was not sequenced. It is therefore not known from whom this variant was inherited or if it is de novo. When she reaches 20–25 years of age, Ariel should begin biennial colonoscopies to screen for cancer. Ariel does not have a family history of colon cancer or other Lynch syndrome cancers, but her mother was adopted and has no information about her own parents. If Ariel’s mother knew that Ariel had a pathogenic variant in MLH1 , she could seek genetic testing for herself to determine whether she has that same variant, and thus pursue preventive measures that could provide substantial benefit. In fact, both Ariel and her mother might benefit if her mother’s risk of dying from early-onset colon cancer is decreased, which would decrease the chances of Ariel losing the care, attention, and financial stability her mother provides for her .

In this case, a clinician must decide whether to disclose a secondary (or “additional” 77 ) finding that identifies a child’s predisposition for developing colon cancer. Although this information would not be clinically actionable for Ariel until adulthood, the clinician knows that Ariel’s positive finding means her mother may have the same variant. Given that her mother is unaware of her family medical history and adults do not routinely undergo predispositional genetic testing, this “incidental” piece of information might provide her only warning.

If Ariel is understood to have a negative moral “right” to an open future, it would be irrelevant that her mother (and indirectly Ariel) might benefit from this information. After all, “rights” function as moral trump cards, overriding other competing interests. 63 , 78 An interest-based standard, however, provides a framework within which multiple competing interests can be weighed. In deciding whether to share this information, the clinician should consider potential benefits to Ariel’s mother as a relevant interest, as well as potential benefits Ariel might receive from her mother avoiding morbidity and mortality. 79 Keeping Ariel’s mother healthy could potentially support Ariel’s interest in having her basic needs met (#3), having support for emotional development (#5), and having a relationship with her parent (#10) (Table  3 ). Ariel’s interest in deferring the decision whether to receive this information until she can decide for herself as an adult is relevant, but it should be weighed alongside these other important interests.

Ariel’s case also highlights an important asymmetry when the right to an open future is applied to secondary genomic findings. Typically, claims about this right assume that the decision to seek or disclose genetic information can be made now , while the child is a minor, or later , when the child has reached adulthood. The options are not always so clear, however. If Ariel is still receiving care from the same physician when she reaches adulthood, then she may have an opportunity to decide for herself whether to receive secondary genetic results, including variants in the MLH1 gene. However, Ariel will not transition to adulthood for another eight years. In this time it is likely that Ariel’s family will have moved, their contact information will have changed, or her neurologist will have relocated or retired. Even if Ariel’s interests are, in principle, better served by deferring the decision until she reaches adulthood, nothing guarantees that she actually will be afforded this opportunity. An interest-based framework, however, can account for this possibility, weighing Ariel’s interest in deciding later with her interest in not missing an opportunity to receive important genetic information that otherwise may be unavailable.

Byron and Huntington disease

Byron is an 11-year-old male who has no chronic medical conditions. Byron’s paternal grandfather recently was diagnosed with Huntington disease. After discussing the issue with a genetic counselor and his family, Byron’s father decided to have a genetic test to determine whether he had inherited this condition. The test came back positive. Byron’s parents explained his father’s result to him, and all three decided to discuss genetic testing with his pediatrician and a genetic counselor. Although there were no pressing medical reasons for Byron to undergo testing, he and his parents agreed, after multiple conversations with his pediatrician and genetic counselor, that Byron should undergo genetic testing soon. The uncertainty was not causing psychosocial problems; indeed, all three were doing well in handling the uncertainty about Byron’s risk. However, all three saw value in not waiting until Byron reached adulthood to learn this information. Byron had decided that resolving this uncertainty would help him think about his future career, mainly because that could affect his choice of high school magnet programs. His parents, for their part, were already planning for his father’s long-term care, and wanted to understand how best to include Byron’s future health in that planning .

In contrast to Ariel’s case, neither Byron nor his parents have a time-sensitive medical interest in Byron undergoing genetic testing for Huntington disease as a minor. There is a strong argument for delaying testing until a person can decide for himself to undergo testing, especially since 85% of adults at risk for Huntington disease choose not to pursue testing. 80 , 81 On the other hand, Byron already has decided he wants to pursue testing and no evidence indicates that he is likely to regret that decision later. Both Byron and his parents want to plan for the future, and these choices could be informed by genetic test results that establish whether Byron is likely to develop Huntington disease as an adult. Given the many other contingencies in life, it is debatable whether a possible diagnosis far in the future should be a determining factor in decisions such as which high school to attend. But on the other hand, the desire to plan ahead in this way is well within the bounds of reasonable (even laudable) behavior for parents and adolescents.

In theory, a rights-based approach is capable of accounting for the circumstance where a relatively mature minor and his parents agree about wanting testing. As the policy statements of the professional organizations referenced above demonstrate, the right to an open future has explicitly been construed as a right to have decisions delayed until the young person is capable of making autonomous decisions (an ethical concept), not necessarily until the young person gains the legal authority to make medical decisions when they reach the age of majority (a legal concept). Assent, for example, provides one strategy for legal minors to express their developing autonomy in medical decisions. 82 , 83 In this case, Byron’s assent to Huntington testing would further justify the decision to obtain testing, while his parents’ permission would render it legally effective.

In practice, however, this is not how the right to an open future has been applied in clinical and research genetics. While strategies like assent can encourage minors to contribute meaningfully to medical decisions, the rights framing fosters the assumption that it is unnecessary, or even inappropriate, for parents to influence these decisions. If children have a right for decisions to be delayed until they can contribute to such choices, then children must also have a right for these decisions to be delayed even longer until no legal obligation requires including parents in the decision.

In contrast to the rights framing, interest-based approaches benefit from development in a range of applications in pediatric ethics, 26 , 62 and thus provide a robust framework for balancing the developing autonomy of children with the authority of parents. In Byron’s case, an interest-based framework incorporates and weighs multiple factors, even interests that are not “clinically actionable,” such as Byron’s emotional development (#5) and sense of self and identity (#11 and #12), his desire to plan for his education and career (#7 and #13), his relationship with his parents (#10), and his parents’ desire to plan for long-term care (#2) (Table  3 ).

Within this framework, deliberation on multiple interests will not always generate identical guidance. In some cases, the child and his family may have compelling interests that override the child’s interest in preserving his future autonomy. On the other hand, there will be cases when the interests served by testing a minor will not be particularly compelling, and the child’s interest in an open future will remain the overriding interest. Consider what would happen if we changed the circumstances surrounding Byron’s case. If Byron were seven years old and perceived no personal utility in learning about his risk for developing Huntington disease, and his parents were simply curious or nervous about what his results might reveal, then the balance of interests could look quite different. His pediatrician or geneticist might be well-justified in declining the parents’ request for immediate genetic testing if the parents’ reason for wanting testing are outweighed by Byron’s interest in later making a decision for himself. This ability to discriminate among dissimilar cases is one notable advantage of the interests approach, and coheres with the intuition that compelling circumstances can override the child’s interest in an open future. 60 , 63 , 64 , 69

In 2013, the American Academy of Pediatrics (AAP) and American College of Medical Genetics and Genomics (ACMG) issued a joint policy statement on genetic testing in children. They recommended against predictive genetic testing unless interventions in childhood are likely to decrease morbidity and mortality. Departing from the former policy, however, they suggested that exceptions might be valid, and gave the example of “families for whom diagnostic uncertainty poses a significant psychosocial burden.” 84 Our approach is consistent with the AAP/ACMG position, as well as a 2015 statement by the American Society of Human Genetics (ASHG) that makes a similar point. 70 We are providing a more substantive framework than the concise policy statements could offer for considering when exceptions might be made.

SUMMARY AND IMPLICATIONS FOR THE FUTURE

In this paper, we noted how new understandings of predictive genetic information raise questions about the ethical foundations of predictive genetic testing. We think those questions are best answered by a shift in the basic approach to children’s open future from a rights-based to an interest-based framework. Childhood is inescapably subject to parental decisions that curtail and shape future choices in diverse and important ways. The idea of a right to an open future is thus impractical, ambiguous, and ripe for misinterpretation. Children need fiduciaries to make choices on their behalf, and nearly all such choices constrain a child’s future in some way.

To be clear, while we have argued that our interest-based framework improves upon the status quo, we recognize that it requires further development. Future research and collaboration should aim to:

Develop a compelling account of how to evaluate, balance, and prioritize the interests on an unranked list like Malek’s (expanded to include a [future] autonomy interest)

Translate this more detailed framework into specific professional guidelines that address the full spectrum of ethically challenging cases related to predictive genetic testing

Apply this framework in other domains where the right to an open future has been evoked, including within the field of genetics (germline modification of the mitochondrial genome and reproductive cloning) and beyond (the sterilization of minors, growth attenuation in children with developmental delay, and many others)

Still, our proposed framework provides a more fruitful and nuanced approach to the complicated ethical issues surrounding predictive genetic testing. We hope it will help guide those who must make these difficult decisions.

Institute of Medicine. Assessing genetic risks: implications for health and social policy. Washington, DC: The National Academies Press; 1994.

Clarke A, Working Party of the Clinical Genetics Society (UK). The genetic testing of children. J Med Genet. 1994;31:785–797.

Article   CAS   PubMed   PubMed Central   Google Scholar  

American Society of Human Genetics Board of Directors, American College of Medical Genetics Board of Directors. Points to consider: ethical, legal, and psychosocial implications of genetic testing in children and adolescents. Am J Hum Genet. 1995;57:1233–1241.

Google Scholar  

American Academy of Pediatrics Committee on Bioethics. Ethical issues with genetic testing in pediatrics. Pediatrics. 2001;107:1451–1455.

Article   Google Scholar  

Arbour L, Canadian Paediatric Society, Bioethics Committee. Guidelines for genetic testing of healthy children. Paediatr Child Health. 2003;8:42–45.

European Society of Human Genetics. Genetic testing in asymptomatic minors: recommendations of the European Society of Human Genetics. Eur J Hum Genet. 2009;17:720–721.

National Society of Genetic Counselors. NSGC position statement: genetic testing of minors for adult-onset conditions. 2017. https://www.nsgc.org/p/bl/et/blogaid=860 . Accessed 20 January 2019.

Human Genetics Society of Australasia. Presymptomatic and predictive testing for children and young adults. 2014. https://www.hgsa.org.au/documents/item/272 . Accessed 20 January 2019.

Wade CH, Wilfond BS, McBride CM. Effects of genetic risk information on children’s psychosocial wellbeing: a systematic review of the literature. Genet Med. 2010;12:317–326.

Article   PubMed   Google Scholar  

Wakefield CE, Hanlon LV, Tucker KM, et al. The psychological impact of genetic information on children: a systematic review. Genet Med. 2016;18:755–762.

Fryer A. Inappropriate genetic testing of children. Arch Dis Child. 2000;83:283–285.

Duncan R, Savulescu J, Gillam L, et al. An international survey of predictive genetic testing in children for adult onset conditions. Genet Med. 2005;7:390–396.

Mand C, Gillam L, Delatycki MB, Duncan RE. Predictive genetic testing in minors for late-onset conditions: a chronological and analytical review of the ethical arguments. J Med Ethics. 2012;38:519–524.

Fenwick A, Plantinga M, Dheensa S, Lucassen A. Predictive genetic testing of children for adult-onset conditions: negotiating requests with parents. J Genet Couns. 2017;26:244–250.

McCullough LB, Brothers KB, Chung WK, et al. Professionally responsible disclosure of genomic sequencing results in pediatric practice. Pediatrics. 2015;136:e974–982.

Article   PubMed   PubMed Central   Google Scholar  

Wisconsin v. Yoder, et al. 406 U.S. 205 (1972).

Feinberg J. The child’s right to an open future. In: Feinberg J, ed. Freedom and fulfillment: philosophical essays. Princeton, NJ: Princeton University Press; 1992. p. 76–97.

Davis DS. Genetic dilemmas and the child’s right to an open future. Hastings Cent Rep. 1997;27:7–15.

Article   CAS   PubMed   Google Scholar  

Davis DS. Genetic dilemmas and the child’s right to an open future. Rutgers Law J. 1997;28:549–592.

CAS   PubMed   Google Scholar  

Davis DS. Genetic dilemmas: reproductive technology, parental choices, and children’s futures. 2nd ed. New York: Oxford University Press; 2010.

Gooding HC, Wilfond B, Boehm K, Biesecker BB. Unintended messages: the ethics of teaching genetic dilemmas. Hastings Cent Rep. 2002;32:37–39.

Hercher L, Uhlmann WR, Hoffman EP, Gustafson S, Chen KM, Public Policy Committee of NSGC. Prenatal testing for adult-onset conditions: the position of the National Society of Genetic Counselors. J Gen Couns. 2016;25:1139–1145.

Mameli M. Reproductive cloning, genetic engineering and the autonomy of the child: the moral agent and the open future. J Med Ethics. 2007;33:87–93.

Davis DS. The parental investment factor and the child’s right to an open future. Hastings Cent Rep. 2009;39:24–27.

de Souza M. Regulating preimplantation genetic diagnosis in Australia: disability and parental choice. J Law Med. 2015;22:915–933.

PubMed   Google Scholar  

Wilfond BS, Miller PS, Korfiatis C, et al. Navigating growth attenuation in children with profound disabilities: children’s interests, family decision-making, and community concerns. Hastings Cent Rep. 2010;40:27–40.

Bredenoord AL, Dondorp W, Pennings G, De Wert G. Ethics of modifying the mitochondrial genome. J Med Ethics. 2011;37:97–100.

Darby RJ. The child’s right to an open future: is the principle applicable to non-therapeutic circumcision? J Med Ethics. 2013;39:463–468.

Hainz T. The enhancement of children versus circumcision: a case of double moral standards? Bioethics. 2015;29:507–515.

Camporesi S. Bend it like Beckham! The ethics of genetically testing children for athletic potential. Sports Ethics Philos. 2013;7:175–185.

Loland S. Against genetic tests for athletic talent: the primacy of the phenotype. Sports Med. 2015;45:1229–1233.

Camporesi S, McNamee MJ. Ethics, genetic testing, and athletic talent: children’s best interests, and the right to an open (athletic) future. Physiol Genomics. 2016;48:191–195.

Graf WD, Nagel SK, Epstein LG, Miller G, Nass R, Larriviere D. Pediatric neuroenhancement: ethical, legal, social, and neurodevelopmental implications. Neurology. 2013;80:1251–1260.

Krutzinna JI. Beyond an open future. Camb Q Healthc Ethics. 2017;26:313–325.

Yu JH, Jamal SM, Tabor HK, Bamshad MJ. Self-guided management of exome and whole-genome sequencing results: changing the results return model. Genet Med. 2013;15:684–690.

Bredenoord AL, de Vries MC, van Delden JJ. Next-generation sequencing: does the next generation still have a right to an open future? Nat Rev Genet. 2013;14:306.

Bredenoord AL, de Vries MC, van Delden H. The right to an open future concerning genetic information. Am J Bioeth. 2014;14:21–23.

Hofmann B. Incidental findings of uncertain significance: to know or not to know—that is not the question. BMC Med Ethics. 2016;17:13.

Sundby A, Boolsen MW, Burgdorf KS, et al. Attitudes of stakeholders in psychiatry towards the inclusion of children in genomic research. Hum Genomics. 2018;12:12.

Kon AA. Ethical issues in decision-making for infants with disorders of sex development. Horm Metab Res. 2015;47:340–343.

Cutas D, Hens K. Preserving children’s fertility: two tales about children’s right to an open future and the margins of parental obligations. Med Health Care Philos. 2015;18:253–260.

Kranendork EJ, Ploem MC, Hennekam RC. Regulating biobanking with children’s tissue: a legal analysis and the experts’ view. Eur J Hum Genet. 2016;24:30–36.

Mason PH. Personal genomic testing, genetic inheritance, and uncertainty. J Bioeth Inq. 2017;14:583–584.

Morissey C, Walker RL. The ethics of general population preventive genomic sequencing: rights and social justice. J Med Philos. 2018;43:22–43.

Mintz RL, Loike JD, Fischbach RL Will CRISPR germline engineering close the door to an open future. Sci Eng Ethics 2018. https://doi.org/10.1007/s11948-018-0069-6 . pp. 1–15. [Epub ahead of print].

Sziron M, Hildt E. Digital media, the right to an open future, and children 0-5. Front Psychol. 2018;9:2137.

Bloch M, Hayden MR. Opinion: predictive testing for Huntington disease in childhood: challenges and implications. Am J Hum Genet. 1990;46:1–4.

CAS   PubMed   PubMed Central   Google Scholar  

Harper PS, Clarke A. Should we test children for ‘adult’ genetic diseases? Lancet. 1990;335:1205–1206.

Borry P, Stultiens L, Nys H, et al. Presymptomatic and predictive genetic testing in minors: a systematic review of guidelines and position papers. Clin Genet. 2006;70:374–381.

Duncan RE. Predictive genetic testing in young people: when is it appropriate? J Paediatr Child Health. 2004;40:593–595.

Duncan RE, Delatycki MB. Predictive genetic testing in young people for adult-onset conditions: where is the empirical evidence? Clin Genet. 2006;69:8–16.

Borry P, Goffin T, Nys H, Dierickx K. Predictive genetic testing in minors for adult‐onset genetic diseases. Mt Sinai J Med. 2008;75:287–296.

Wilfond B, Ross LF. From genetics to genomics: ethics, policy, and parental decision-making. J Pediatr Psychol. 2009;34:639–647.

Parker M. Genetic testing in children and young people. Fam Cancer. 2010;9:15–18.

Clarke A. What is at stake in the predictive genetic testing of children? Fam Cancer. 2010;9:19–22.

Fenwick J. Are guidelines for genetic testing of children necessary? Fam Cancer. 2010;9:23–25.

Gilbar R. Genetic testing of children for familial cancers: a comparative legal perspective on consent, communication of information and confidentiality. Fam Cancer. 2010;9:75–87.

Anderson JA, Hayeems RZ, Shuman C, et al. Predictive genetic testing for adult-onset disorders in minors: a critical analysis of the arguments for and against the 2013 ACMG guidelines. Clin Genet. 2015;87:301–310.

Mills C. The child’s right to an open future? J Soc Philos. 2003;34:499–509.

Cohen CB. Wrestling with the future: should we test children for adult-onset genetic conditions? Kennedy Inst Ethics J. 1998;8:111–130.

Millum J. The foundation of the child’s right to an open future. J Soc Philos. 2014;45:522–538.

Ross LF. Children, families, and health care decision making. New York: Clarendon Press; 1998.

Harris J, Keywood K. Ignorance, information and autonomy. Theor Med Bioeth. 2001;22:415–436.

Malpas P. Predictive genetic testing on children and respect for autonomy. Int J Child Right. 2005;13:273–285.

Robertson S, Savulescu J. Is there a case in favour of predictive genetic testing in young children? Bioethics. 2001;15:26–49.

Rhodes R. Why test children for adult-onset genetic diseases? Mt Sinai J Med. 2006;73:609–616.

Häyry M, Takala T. Genetic information, rights, and autonomy. Theor Med Bioeth. 2001;22:403–414.

Wenar L. Rights. Stanford Encyclopedia of Philosophy is Edward N. Zalta. Stanford, California: Center for the Study of Language and Information (CSLI), Stanford University; 2015.

British Society of Human Genetics (BSHG). Report on the genetic testing of children. 2010. https://www.bsgm.org.uk/media/678741/gtoc_booklet_final_new.pdf . Accessed 20 January 2019.

American Society of Human Genetics Board of Directors, American College of Medical Genetics Board of Directors. Points to consider: ethical, legal, and psychosocial implications of genetic testing in children and adolescents. Am J Hum Genet. 2015;97:6–21.

Ross LF. Theory and practice of pediatric bioethics. Perspect Biol Med. 2016;58:267–280.

Diekema DS. Parental refusals of medical treatment: the harm principle as threshold for state intervention. Theor Med Bioeth. 2004;25:243–264.

Gillam L. Children’s bioethics and the zone of parental discretion. Monash Bioeth Rev. 2010;20:01–03.

Malek J. What really is in a child’s best interest? Toward a more precise picture of the interests of children. J Clin Ethics. 2009;20:175–182.

Clayton EW. Genetic testing in children. J Med Philos. 1997;22:233–251.

Lucassen A, Fenwick A. Testing children for adult onset conditions: the importance of contextual clinical judgment. J Med Ethics. 2012;38:531–532.

Tan N, Amendola LM, O’Daniel JM, et al. Is “incidental finding” the best term?: a study of patients’ preferences. Genet Med. 2017;19:176–181.

Dworkin R. Rights as trumps. In: Waldron J, ed. Theories of rights. Oxford: Oxford University Press; 1984. p. 153–167.

Wilfond BS, Fernandez CV, Green RC. Disclosing secondary findings from pediatric sequencing to families: considering the “benefit to families”. J Law Med Ethics. 2015;43:552–558.

PubMed   PubMed Central   Google Scholar  

Creighton S, Almqvist EW, MacGregor D, et al. Predictive, pre-natal and diagnostic genetic testing for Huntington’s disease: the experience in Canada from 1987 to 2000. Clin Genet. 2003;63:462–475.

Morrison PJ, Harding-Lester S, Bradley A. Uptake of Huntington disease predictive testing in a complete population. Clin Genet. 2011;80:281–286.

Wilfond BS, Diekema DS. Engaging children in genomics research: decoding the meaning of assent in research. Genet Med. 2012;14:437–443.

Brothers KB, Lynch JA, Aufox SA, et al. Practical guidance on informed consent for pediatric participants in a biorepository. Mayo Clin Proc. 2014;89:1471–1480.

AAP Committee on Bioethics, AAP Committee on Genetics, ACMG Social Ethical and Legal Issues Committee. Ethical and policy issues in genetic testing and screening of children. Pediatrics. 2013;131:620–622.

Download references

Acknowledgements

The CSER Consortium is funded by National Human Genome Research Institute (NHGRI) and National Cancer Institute (NCI) (U01 HG006485 [Baylor College of Medicine], U01 HG006500 [Brigham and Women’s Hospital], U01 HG006546 [Children’s Hospital of Philadelphia], U01 HG006492 [Dana-Farber Cancer Institute], UM1 HG007301 [HudsonAlpha Institute], UM1 HG007292 [Kaiser Permanente], UM1 HG006508 [University of Michigan], U01 HG006487 [University of North Carolina], U01 HG006507 [University of Washington], R01 HG006615 [Boston Children’s Hospital], R21 HG006596 [Columbia University], R01 HG006600 [Columbia University], R21 HG006613 [Children’s Mercy Hospital], R21 HG006594 [Johns Hopkins University], R01 HG004500 [Mayo Clinic], R01 HG006618 [Seattle Children’s Hospital], R01 CA154517 [UC–San Francisco, Mayo College of Medicine, and University of Minnesota], R21 HG006612 [Vanderbilt University and McGill University], U01 HG007307 [University of Washington serving as the Coordinating Center]). ClinSeq (ZIA HG200387) is supported by the NHGRI Intramural Research Program. The authors thank the coordinating center of the Clinical Sequencing Exploratory Research (CSER) Consortium (University of Washington) for their support, and in particular Jeffrey Ou, who provided substantive support for this effort. Lucia Hindorff with NHGRI provided program staff support for this project. The authors also thank the following members of the CSER Consortium Pediatrics Working Group for participating in meetings to discuss the article as it was conceptualized and written: Benjamin Berkman (National Institutes of Health, Department of Bioethics), Barbara Bernhardt (University of Pennsylvania), Charlisse Caga-Anan (NCI), Ellen Wright Clayton (Vanderbilt University Medical Center), Aaron Goldenberg (Case Western Reserve University), Sara Chandros Hull (National Institutes of Health, Department of Bioethics), Steve Joffe (University of Pennsylvania), Ian Krantz (Children’s Hospital of Philadelphia), Michelle Lewis (John Hopkins Berman Institute of Bioethics), Wayne Liang (University of Alabama at Birmingham), Nicole Lockhart (NHGRI), Susana McCollum, Larry McCullough (Baylor College of Medicine), Amy McGuire (Baylor College of Medicine), Ali Noorbaksh (Baylor College of Medicine), Sarita Panchang (Baylor College of Medicine), D. Will Parsons (Baylor College of Medicine), Jacob Reiss (Kaiser Permanente Northwest), Myra Roche (University of North Carolina at Chapel Hill), Laura Rodriguez (NHGRI), Edward Romasko (Children’s Hospital of Philadelphia), Lainie Friedman Ross (University of Chicago), Richard Sharp (Mayo Clinic), Debra Skinner (University of North Carolina at Chapel Hill), Melody Slashinski (University of Massachusetts Amherst), Holly Tabor (Stanford University), Ashley Tomlinson (University of Pennsylvania), Susan Wolf (University of Minnesota), and Joon-Ho Yu (University of Washington). The authors also thank Leslie Ann McNolty (Center for Practical Bioethics) for helpful suggestions on revisions at several stages of the paper’s development, including the final draft. We would also like to thank Dena Davis (Lehigh University) who graciously served as a guest discussant for one of these meetings.

Author information

Authors and affiliations.

Children’s Mercy Bioethics Center, Children’s Mercy Kansas City, Kansas City, MO, USA

Jeremy R. Garrett PhD & John D. Lantos MD

Department of Pediatrics, University of Missouri–Kansas City, Kansas City, MO, USA

National Human Genome Research Institute, National Institutes of Health, Bethesda, MD, USA

Leslie G. Biesecker MD & Jean E. McEwen JD, PhD

Division of Pediatric Clinical and Translational Research, University of Louisville School of Medicine, Louisville, KY, USA

Janet E. Childerhose PhD

Departments of Medicine and Pediatrics, Columbia University, New York, NY, USA

Wendy K. Chung MD, PhD

Division of Genetics and Genomics and the Manton Center for Orphan Diseases Research, and Department of Pediatrics, Harvard Medical School, Boston Children’s Hospital, Boston, MA, USA

Ingrid A. Holm MD, MPH

UCSF Bioethics, University of California San Francisco, San Francisco, CA, USA

Barbara A. Koenig PhD

Treuman Katz Bioethics Center, Seattle Children’s Research Institute, Seattle, WA, USA

Benjamin S. Wilfond MD

Department of Pediatrics, University of Washington, Seattle, WA, USA

Department of Pediatrics, University of Louisville, Louisville, KY, USA

Kyle Brothers MD, PhD

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Jeremy R. Garrett PhD .

Ethics declarations

The authors declare no conflicts of interest.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Garrett, J.R., Lantos, J.D., Biesecker, L.G. et al. Rethinking the “open future” argument against predictive genetic testing of children. Genet Med 21 , 2190–2198 (2019). https://doi.org/10.1038/s41436-019-0483-4

Download citation

Received : 20 November 2018

Accepted : 27 February 2019

Published : 21 March 2019

Issue Date : October 2019

DOI : https://doi.org/10.1038/s41436-019-0483-4

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • adult-onset conditions
  • children’s interests
  • pediatric genetic testing
  • right to an open future
  • secondary findings

This article is cited by

Puberty suppression for pediatric gender dysphoria and the child’s right to an open future.

  • Sarah C. J. Jorgensen
  • Nicole Athéa
  • Céline Masson

Archives of Sexual Behavior (2024)

The Right Not to Be Subjected to AI Profiling Based on Publicly Available Data—Privacy and the Exceptionalism of AI Profiling

  • Thomas Ploug

Philosophy & Technology (2023)

Return of non-ACMG recommended incidental genetic findings to pediatric patients: considerations and opportunities from experiences in genomic sequencing

  • Kevin M. Bowling
  • Michelle L. Thompson
  • Anna C. E. Hurst

Genome Medicine (2022)

Die Rechte zukünftiger Kinder im Kontext pränataler Diagnostik

  • Dagmar Schmitz
  • Marcus Düwell

Ethik in der Medizin (2022)

Polygenic risk scores in the clinic: new perspectives needed on familiar ethical issues

  • Anna C. F. Lewis
  • Robert C. Green

Genome Medicine (2021)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

thesis statement for genetic testing

Thesis Statement Examples

Caleb S.

Practical Thesis Statement Examples That Will Transform Your Writing

10 min read

Published on: Sep 3, 2024

Last updated on: Sep 3, 2024

thesis statement examples

People also read

Writing a strong thesis statement is key to a great essay, but coming up with the perfect one can be tricky. 

No matter if you're working on a personal essay, an argumentative paper, or something else, having clear examples can really help.

In this blog, we’ll explore a variety of thesis statement examples for different types of essays. This will help you understand how to create strong statements that guide your writing and keep your readers engaged. 

Let’s find the right thesis for your next essay!

What Is A Thesis Statement?

A thesis statement is a key part of your essay that tells readers what your essay is about. It’s a clear sentence that sums up the main point or argument you’re going to make. Think of it as a guide that helps your readers understand your main idea and what to expect from your essay.

Your thesis statement usually goes at the end of your introduction. It gives your essay direction and helps keep everything focused on your main point.

Examples Of Thesis Statements By Essay Type

To help you see how to write effective thesis statements, here are thesis statement examples for essays of various types. Each example will give you a clearer picture of how to approach various topics.

Examples Of Thesis Statements For Personal Essays 

A personal essay thesis statement reflects your unique experiences and feelings. It shares a central idea about a personal story or insight you’re discussing.

Here are a few examples to give you an idea of how to craft your own:

  • Overcoming my fear of public speaking taught me that courage is not the absence of fear, but the ability to act despite it.
  • Traveling to a new country showed me that stepping out of my comfort zone is the best way to grow and learn.
  • My experience of balancing school and a part-time job proved that hard work and dedication can lead to personal success and self-discovery.

Thesis Statement Examples For Informative Essay 

A thesis statement for an informative essay provides a clear and specific overview of the topic you’re explaining. It helps readers understand the focus of your essay and what information they can expect to learn. 

Let’s take a look at some informative thesis statement examples :

  • The process of photosynthesis is essential for plant growth as it converts sunlight into energy and produces oxygen, which is crucial for life on Earth.
  • The history of the internet reveals how technological advancements have transformed communication, from early networks to the modern web.
  • Understanding the basics of genetic inheritance can explain how traits are passed from parents to offspring and how genetic disorders can occur.

Thesis Statement Examples For Compare And Contrast Essays

A thesis statement for a compare and contrast essay highlights the similarities and differences between two or more subjects. It helps readers understand the main points of comparison and what you will be discussing. 

Some examples are:

  • While both public and private schools aim to provide quality education, they differ significantly in terms of funding, curriculum flexibility, and student-to-teacher ratios.
  • Although electric cars and gasoline-powered cars serve the same purpose of transportation, they differ in their environmental impact, cost, and efficiency.
  • The lifestyles of city dwellers and rural residents differ greatly, with urban areas offering more job opportunities and amenities, while rural areas provide a slower pace of life and closer connection to nature.

Argumentative Essay Thesis Statement Examples

A thesis statement for an argumentative essay clearly states your opinion on a hot topic and explains why you hold that view. It shows what you believe and what you’ll be arguing for. 

Below are some argumentative thesis statement examples :

  • A universal basic income can help reduce poverty and provide financial stability, making it a crucial step toward reducing economic inequality.
  • Renewable energy is vital for fighting climate change because it cuts down on greenhouse gas emissions and supports a healthier planet.
  • Requiring school uniforms in public schools can decrease peer pressure and help students focus more on their studies.

Thesis Statement Examples For Persuasive Essay 

A thesis statement for a persuasive essay aims to convince the reader of a particular viewpoint. It presents your position and hints at the arguments you’ll use to support it. 

Some examples include: 

  • Adopting a plant-based diet is beneficial for health and the environment, as it reduces the risk of chronic diseases and decreases ecological footprints.
  • Investing in public transportation improves urban mobility and reduces traffic congestion, leading to a more efficient and eco-friendly city.
  • Banning single-use plastics is essential to protect marine life and reduce pollution, helping to preserve the environment for future generations.

Analytical Essay Thesis Statement Examples 

A thesis statement for an analytical essay breaks down a topic and examines its components. It highlights what you will analyze and what insights or conclusions you aim to provide. 

Here are some good thesis statement examples for analytical essays:

  • Analyzing Shakespeare’s use of imagery in "Macbeth" reveals how it enhances the play’s themes of ambition and guilt.
  • Examining the impact of social media on communication shows how it has changed the way we interact and perceive relationships.
  • The portrayal of leadership in “The Great Gatsby” illustrates how wealth and power can corrupt moral values and influence behavior.

Expository Essay Thesis Statement Examples 

A thesis statement for an expository essay explains a topic or idea in detail. It provides a clear summary of what the essay will cover and how it will inform the reader. 

See the examples mentioned below: 

  • The process of recycling involves several key steps, including sorting materials, processing them into raw materials, and creating new products.
  • The history of the internet highlights key milestones such as the development of early networks, the rise of the World Wide Web, and the evolution of online communication.
  • Understanding the causes of climate change requires examining factors like greenhouse gas emissions, deforestation, and industrial activities.

Process Essay Thesis Statement Examples 

A thesis statement for a process essay explains how something is done or how a process works. It gives a clear overview of the steps involved. Take a glance at these examples :

  • Making homemade pizza involves preparing the dough, adding toppings, and baking it to create a delicious and customizable meal.
  • The steps to start a small business include researching the market, creating a business plan, and securing funding to ensure a successful launch.
  • Learning a new language requires practicing speaking and listening skills, studying grammar, and immersing oneself in the language through reading and conversation.

Thesis Statement Examples According to Different Academic Levels 

Writing a thesis statement changes as you move through different study levels. Each stage has its own approach and complexity. Here’s how thesis statements might look across different levels:

Thesis Statement Examples for Kids 

At a basic level, thesis statements for kids are simple and direct. They usually focus on familiar topics and straightforward ideas. Consider these examples :

  • Dogs make great pets because they are loyal, fun, and good with kids.
  • Reading books is important because it helps you learn new things and improves your imagination.

Middle School Thesis Statement Examples 

In middle school, thesis statements start to involve more detail and support. They reflect a better understanding of how to structure arguments. Here are some examples :

  • School uniforms should be required because they promote equality, reduce distractions, and make it easier for students to focus on their studies.
  • Eating a balanced diet is crucial for maintaining good health because it provides essential nutrients, boosts energy levels, and helps prevent diseases.

High School Thesis Statement Examples 

High school thesis statements are more sophisticated, often including a clear argument and multiple supporting points. Take a look these examples :

  • The benefits of online learning outweigh the drawbacks because it offers flexibility, access to a wide range of resources, and the ability to balance education with other responsibilities.
  • Participating in extracurricular activities is important for high school students as it helps develop leadership skills, build friendships, and enhance college applications.

College Thesis Statement Examples 

At the college level, thesis statements are complex and detailed and often address more complex arguments. Examples include:

  • Implementing renewable energy solutions, such as solar and wind power, is essential for reducing our reliance on fossil fuels and mitigating the effects of climate change.
  • The rise of social media has transformed political campaigning by increasing voter engagement, spreading misinformation, and altering traditional campaign strategies.

Thesis Statement Examples for Research Papers

For research papers, thesis statements must be well-researched and specific and provide a clear direction for the study. Consider these examples :

  • Exploring the effects of childhood trauma on adult mental health reveals significant correlations between early experiences and the development of psychological disorders later in life.
  • Investigating the impact of artificial intelligence on the job market shows that while AI creates new opportunities, it also poses challenges related to job displacement and workforce adaptation.

More Examples Of Thesis Statements 

As you work on different essays and writing tasks, you’ll see that thesis statements can vary a lot. Here are some additional examples to illustrate their diversity.

Literary Analysis Thesis Statement Examples 

  • In "To Kill a Mockingbird," Harper Lee uses symbolism, such as the mockingbird, to highlight the themes of innocence and moral growth.
  • The use of unreliable narrators in Edgar Allan Poe’s “The Tell-Tale Heart” enhances the story’s exploration of guilt and madness.

Implied Thesis Statement Examples 

  • Despite its surface simplicity, “The Little Prince” offers a profound critique of adult behavior through its exploration of human nature and relationships.
  • The persistent use of color imagery in “The Great Gatsby” subtly emphasizes the theme of the American Dream and its inherent flaws.

Thesis Statement Examples For Research Papers 

  • Examining the effects of sleep deprivation on academic performance shows a direct link between lack of sleep and reduced cognitive abilities in students.
  • Research into the impact of climate change on coastal ecosystems reveals that rising sea levels and increased temperatures are threatening biodiversity and habitat stability.

Complex Thesis Statement Examples

  • While the integration of technology in education offers numerous benefits, such as personalized learning and greater accessibility, it also presents challenges related to screen time and data privacy.
  • The debate over genetic engineering in agriculture involves both potential benefits, like increased crop yields and disease resistance, and ethical concerns, such as environmental impact and genetic diversity.

In closing, 

A strong thesis statement is the backbone of any good essay. It helps guide your writing and keeps your readers focused on your main point. With the examples provided, you can see how to shape your thesis for different types of essays and academic levels.

If you want a little extra help with your thesis statements, check out the thesis statement generator from MyEssayWriter.ai . It’s a handy tool that can help you create and perfect your thesis statements quickly. 

For extra help with essay writing, check out our essay writer . It's an AI tool that can write high-quality essays for you in a breeze!

Frequently Asked Questions

How do i write my thesis statement.

To write and start a thesis statement, you should:

  • Pick Your Topic: Decide what your essay will be about.
  • Formulate Your Argument: Choose your main point or stance on the topic.
  • Be Specific: Make sure your statement clearly outlines what you'll discuss.
  • Make It Debatable: Your thesis should present an argument that can be supported with evidence.
  • Keep It Concise: Aim for one or two sentences that clearly express your main idea.

What 3 things should a thesis statement have?

Typically, a thesis statement format includes three main parts: the topic you're discussing, your main argument or viewpoint , and the reasons or evidence you'll use to back up that argument. 

What is an example of a weak and strong thesis statement?

Weak Thesis Statement: "Social media is bad."

  • It's too broad and lacks detail.

Strong Thesis Statement: "Social media platforms negatively impact mental health by increasing anxiety and depression among teenagers, and this can be addressed through improved online safety measures."

  • It's specific, takes a clear stance, and hints at the main points of the essay.

What is a thesis statement sentence?

A thesis statement sentence is a single sentence in your essay that summarizes your main point or argument. It’s usually found at the end of your introduction and guides the rest of your essay.

Caleb S. (Mass Literature and Linguistics, Masters)

Caleb S. is an accomplished author with over five years of experience and a Master's degree from Oxford University. He excels in various writing forms, including articles, press releases, blog posts, and whitepapers. As a valued author at MyEssayWriter.ai, Caleb assists students and professionals by providing practical tips on research, citation, sentence structure, and style enhancement.

On This Page On This Page

Get Access to Advanced Features with our Affordable Plans

Complimentary Trial

Continue for FREE

OFFER ENDS Today

Signup for Free Access

Access to all Tools

No Credit Card needed

Quota: 1500 Words (6 pages)

Offer ends in: 03hr 19m 8s

Already a user? Login instead

For Monthly Usage

$ 9.99 /month

Up to 2500 words/month

Access to all features

AI Essay Writer

AI Essay Outliner

Go Big Monthly!

$ 14.99 /month

Up to 100,000 words/month

Access to all existing tools

50% off on custom essay orders

500 credits for plagiarism check

Top Annual Savings!

Advanced (4 Months FREE )

$ 99.99 /year

OFFER ENDS Soon

Unlimited essays

12 Months for the price of 8

Access to all upcoming tools

thesis statement for genetic testing

Get started for free

Please enter a valid Name

Please enter a valid email address

Please enter a valid Phone Number

Please enter atleast five characters

Please enter a valid Password

Show Password

Already have an account? Sign In here.

Please enter your email address

Forgot Password?

Don’t have an account? Sign Up

Verify Your Account

Enter the verification codes to confirm your identity.

Code sent to [email protected]

Send again in seconds

Code sent to +1 302 385 6690

loader

Claim Free Essay

Your first custom essay order on our website is FREE!

Genetic Testing: Advantages and Disadvantages Essay

  • To find inspiration for your paper and overcome writer’s block
  • As a source of information (ensure proper referencing)
  • As a template for you assignment

The issue of genetic testing is a highly controversial one, as its advantages and disadvantages present various dilemmas. It is still not clear if genetic testing should or should not become a common procedure that all people undergo regularly. I believe that it is an extremely personal decision to make. There are certain limitations and concerns that a diagnosed person can face, especially when they are diagnosed with untreatable and lethal disorders (Norrgard, 2008). Knowing about conditions like that may significantly decrease the quality of life and even lead to depression and anxiety.

At the same time, I acknowledge all the benefits that genetic testing can bring in terms of diagnosing a wide range of diseases and conditions. Fearing that they might discover hereditary predispositions to some untreatable diseases, many people choose not to get tested. However, I believe that deep inside, they still think about it and have concerns; I would if my family had a history of genetic conditions. That is why some people may even feel relieved when they undergo testing and have to face difficult results. At least they can know for sure that they are predisposed to certain conditions and focus on ways to improve their lives (Kurian et al., 2019). After all, genes are believed to be malleable; a positive approach, holistic nutritional program, and avoiding environmental toxins will not harm any person whose genetic testing results show a predisposition to certain diseases.

In the end, I do not think there is a universal answer to this question. Each person’s choice has to be authentic because they are the ones who will have to live their lives with this knowledge. I do believe, though, that scientific and health communities might focus more on raising awareness about genome sequencing, with particular reference to conditions that can be caught in the early stages, such as different types of cancer.

Kurian, A. W., Ward, K. C., Howlader, N., Deapen, D., Hamilton, A. S., Mariotto, A., Miller, D., Penberthy, L. S., & Katz, S. J. (2019). Genetic testing and results in a population-based cohort of breast cancer patients and ovarian cancer patients . Journal of Clinical Oncology , 37 (15), 1305-1315.

Norrgard, K. (2008). Genetic testing and family planning | Learn science at Scitable . Scitable.

  • Discussion of Editing Human Embryos and Adults
  • Should Cloning Be 100% Legal or Illegal?
  • Next-Generation and Traditional Sequencing Methods
  • Infectious Bacterial Identification From DNA Sequencing
  • Whole-Genome Sequencing for Identification and Gene Function Prediction of Bacterial Genomes
  • Genetic Modification and Cloning
  • Gene Modification: Building Baby From the Genes Up
  • "Ghost in the Genes" Video Review
  • Editing the Genome of Human Embryos: Pros and Cons
  • Down Syndrome: Congenital Heart Disease and Prenatal Testing
  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2022, November 8). Genetic Testing: Advantages and Disadvantages. https://ivypanda.com/essays/genetic-testing-advantages-and-disadvantages/

"Genetic Testing: Advantages and Disadvantages." IvyPanda , 8 Nov. 2022, ivypanda.com/essays/genetic-testing-advantages-and-disadvantages/.

IvyPanda . (2022) 'Genetic Testing: Advantages and Disadvantages'. 8 November.

IvyPanda . 2022. "Genetic Testing: Advantages and Disadvantages." November 8, 2022. https://ivypanda.com/essays/genetic-testing-advantages-and-disadvantages/.

1. IvyPanda . "Genetic Testing: Advantages and Disadvantages." November 8, 2022. https://ivypanda.com/essays/genetic-testing-advantages-and-disadvantages/.

Bibliography

IvyPanda . "Genetic Testing: Advantages and Disadvantages." November 8, 2022. https://ivypanda.com/essays/genetic-testing-advantages-and-disadvantages/.

Warning: The NCBI web site requires JavaScript to function. more...

U.S. flag

An official website of the United States government

The .gov means it's official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you're on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • Browse Titles

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

National Academies of Sciences, Engineering, and Medicine; Health and Medicine Division; Board on Health Care Services; Board on the Health of Select Populations; Committee on the Evidence Base for Genetic Testing. An Evidence Framework for Genetic Testing. Washington (DC): National Academies Press (US); 2017 Mar 27.

Cover of An Evidence Framework for Genetic Testing

An Evidence Framework for Genetic Testing.

  • Hardcopy Version at National Academies Press

Advances in genetics and genomics are transforming medical practice, resulting in a dramatic growth of genetic testing in the health care system. The rapid development of new technologies, however, has also brought challenges, including the need for rigorous evaluation of the validity and utility of genetic tests, questions regarding the best ways to incorporate them into medical practice, and how to weigh their cost against potential short- and long-term benefits. As the availability of genetic tests increases so do concerns about the achievement of meaningful improvements in clinical outcomes, costs of testing, and the potential for accentuating medical care inequality.

  • STATEMENT OF TASK

Given the rapid pace in the development of genetic tests and new testing technologies, the Department of Defense (DoD) Office of Health Affairs asked the National Academies of Sciences, Engineering, and Medicine to convene a committee to

examine the relevant medical and scientific literature to determine the evidence base for different types of genetic tests (e.g., predictive, diagnostic, and prognostic) for patient management. The committee is to provide recommendations to advance the development of an adequate evidence base for genetic tests to improve patient care and treatment. Additionally, the committee will recommend a framework to DoD for decision making regarding the use of genetic tests in clinical care.

The DoD Office of Health Affairs assists in the development of strategies and priorities to achieve the health mission of the Military Health System (MHS) and participates in formulating, developing, overseeing, and advocating the policies of the Secretary of Defense. TRICARE is a major part of the MHS that combines the resources of military hospitals and clinics with civilian health care networks. The Office of Health Affairs is responsible for providing a cost-effective, high-quality health benefit to about 9.6 million active-duty uniformed service members, retirees, and their families. The MHS has a $50 billion annual budget and consists of a worldwide network of 59 military hospitals, 360 health clinics, private-sector health-business partners, and the Uniformed Services University of the Health Sciences.

  • APPROACH TO THE TASK

The 17-member committee met five times over the course of the study to address the clinical application and clinical utility of genetic tests as directed by the statement of task. In support of the committee's discussions and deliberations, targeted literature searches were conducted, and information from relevant scientific, professional, and federal sources was gathered. The committee quickly found that numerous ad hoc groups, regulatory agencies, organizations, and professional societies were developing approaches to assessment of genetic testing and collecting and using evidence with various methods. None of the approaches was judged to be entirely satisfactory for DoD's purposes, but generating an entirely new approach was judged both unnecessary and infeasible because of the constraints on time and resources. Thus, the committee reached consensus early in its deliberations to adapt the best of what was already available.

Because a framework aimed only at current technologies might soon be outdated, the committee's framework focuses on general principles of clinical usefulness that are relevant to any genetic technology. After conversations with DoD and internal discussion, the committee focused its framework on germline DNA-based tests, as a too-broad scope (e.g., to include RNA and other molecular-based tests) would expand the task to an unmanageable level. The committee believes, however, that the recommended framework is broadly applicable to assessment of other molecular technologies. Finally, the committee makes many suggestions about how the framework might be used in practice, but stops short of providing complete illustrative examples because it could not substitute its judgments for those that DoD alone is in a position to make.

  • USES OF GENETIC TESTING

The committee took as its starting point the definition of genetic testing by the National Institutes of Health's National Human Genome Research Institute as

an analysis of human DNA, RNA, chromosomes, proteins, and certain metabolites in order to detect heritable disease-related genotypes, mutations, phenotypes or karyotypes for clinical purposes. Such purposes include predicting risk of disease, identifying carriers and establishing prenatal and clinical diagnosis or prognosis. Prenatal, newborn and carrier screening, as well as testing in high-risk families, are included. Tests for metabolites are considered only when they are undertaken with high probability that an excess or deficiency of the metabolite indicates the presence of heritable mutations in single genes. Tests conducted purely for research are excluded from the definition, as are tests for somatic (as opposed to heritable) mutations, and testing for forensic purposes.

The committee organized its analysis around the purpose of the test, using the categories of diagnostic, predictive, and reproductive.

Diagnostic Genetic Testing

Diagnostic genetic testing is used to identify or rule out a specific genetic condition. Genetic testing is often used to confirm a diagnosis when a particular condition is suspected on the basis of physical signs and symptoms. Family history might also play an important role in identifying appropriate diagnostic tests. For example, genetic testing for inherited cancer syndromes might be performed because a family history suggests an inherited risk of early-onset cancer. The results of a diagnostic test can influence patients' and clinicians' choices about clinical management of disorders.

Predictive Genetic Testing

Predictive genetic testing identifies gene variants that increase a person's risk of developing heritable disorders, such as some types of cancer, before any signs or symptoms appear. In some cases, a diagnostic test of an affected person yields a result that can be used to recommend a predictive test for relatives. For example, if a woman who has breast cancer is found to have a BRCA1 variant, indicating a hereditary breast–ovarian cancer syndrome, her relatives can be offered the option of being tested to determine whether they carry the variant.

Similarly, predictive genetic tests can be used for population screening, notably in the case of newborn screening, which involves testing infants a few days after birth for evidence of treatable diseases that are known to be detrimental to health or development.

Pharmacogenetic testing is another important class of predictive genetic testing. It provides information about individual variation in drug pharmacodynamics (effects on drug receptors) and pharmacokinetics (uptake, distribution, and metabolism), which makes it possible to identify patients who are at increased risk for adverse effects or who are likely to be nonresponders. Pharmacogenetic testing has the potential to help health care providers tailor therapies by adjusting the dose or drug that might work best for an individual patient, to prevent adverse drug reactions, or to select persons who are likely to respond to a given drug.

Reproductive Genetic Testing

Reproductive genetic testing offers the opportunity to identify people who are at increased risk for having a child who has a genetic disease or to identify an affected embryo or fetus. Carrier or heterozygote genetic testing is used to identify people who are at increased risk for having a child who has a genetic disease. Most carrier tests identify people who are heterozygous for or “carry” one variant copy and one normal copy of a gene that is associated with a disorder transmitted as an autosomal recessive trait (that is, a trait that requires both copies of the gene to express the disorder). Carriers typically show no signs of the disease in question, but they have the ability to pass on the variant gene to their children. Carrier testing is offered either because a family history indicates the presence of a specific inherited disease or as a screening test.

Prenatal genetic testing is used to detect abnormalities in the genes or chromosomes of a fetus before birth. Current guidelines recommend that all pregnant women be offered maternal serum screening tests to identify pregnancies that are at increased risk for a trisomy disorder (such as Down syndrome) or for neural-tube defect. If a screening test is positive, a confirmatory test can be performed. Prenatal testing might also be offered if the parents are known to be at risk for having a child who has a specific genetic disorder.

Preimplantation genetic testing, or preimplantation genetic diagnosis, is a specialized technique that can reduce the risk of having a child who has a particular genetic or chromosomal disorder. It is used to detect genetic changes in embryos that were created by the use of assisted reproductive techniques, such as in vitro fertilization.

  • ETHICAL, LEGAL, AND SOCIAL IMPLICATIONS

Genetic testing has an array of personal and societal implications that require consideration in defining appropriate testing practice. Those implications are related to shared genetic risk among family members, the use of genetic testing in reproductive decision making, and the potential for genetic information to generate stigma or discrimination. In addition, the rapid development of genomic technology and the many remaining uncertainties about the health implications of genetic risk raise ethical considerations related to health care disparities, clinical data sharing, and the scope of result reporting from genome-scale testing.

  • ASSESSMENT OF GENETIC TESTS

Many of the available methods for assessing genetic tests cover the three common domains of evaluation: analytic validity, clinical validity, and clinical utility. Table S-1 provides a comparison of the different frameworks evaluated by the committee. The ACCE 1 and Fryback–Thornbury 2 models include an additional domain: societal effect of the test. Some evaluation frameworks (such as the Fryback–Thornbury hierarchy) provide general conceptual guidance, whereas analytic frameworks (such as those of the USPSTF 3 and the EGAPP Working Group 4 ) provide additional detail regarding the relevant populations, interventions, comparators, outcomes, times, and settings.

TABLE S-1. Comparison of Frameworks.

Comparison of Frameworks.

The McMaster University Evaluation Framework provides a thorough approach and rich detail for making decisions about genetic testing. The three domains of the McMaster evaluation framework (establishing evaluation criteria, determining acceptable cutoffs for each criterion, and determining conditions of coverage for “gray zones”) provide the foundation of the model recommended by the present committee.

The domains map broadly to the ACCE criteria, with the Fryback–Thornbury hierarchy representing clinical utility in three categories: patient-outcome efficacy, therapeutic efficacy, and diagnostic-thinking efficacy. The USPSTF method evaluates health care interventions in the context of preventive services in the general population and thus emphasizes patient outcomes, such as morbidity and mortality, as high-level end points. The EGAPP method organizes evidence into the ACCE categories and evaluates the chain of evidence by using a framework similar to that of USPSTF. The McMaster University evaluation framework identifies six criteria, one of which (effectiveness) depends on the intended purpose of the test; the framework also considers aggregate costs, use metrics, and cost-effectiveness criteria that are important from the health care system perspective. The Genetic testing Evidence Tracking Tool (GETT), although not explicitly intended as an evaluation method, provides a systematic model for organizing published evidence in 10 main categories. Finally, the evaluation process, the first step of which is to determine whether the purpose of genetic testing in a particular clinical scenario is clear and worthwhile, is the subject of two questions posed by Frueh and Quinn.

As with other elements of medical care, information about genetic testing is gathered and evaluated through research. The quality and quantity of that information vary, which confers some level of certainty about the resulting findings and conclusions. Additionally, the numerous public and private organizations conducting evidence reviews have differing approaches and criteria for assessing the available evidence regarding the validity and utility of genetic tests. Evidence is generally organized, summarized, and synthesized through systematic reviews or meta-analyses. Evidence summaries then inform clinical practice guidelines and policy decisions.

The evidence base on genetic testing is limited by the frequent lack of direct evidence, for example, from randomized controlled trials (RCTs). The rarity of many genetic diseases and of variants associated with genetic diseases makes it difficult for researchers to gather enough patients who have the disease or variant to conduct a traditional study; thus, the usual sources of evidence often are not available. One response to that challenge has been the establishment of databases that aid in the collection and accumulation of evidence on clinical validity and utility; another potential approach is to provide guidance through decision modeling.

Establishing analytic validity, clinical validity, and clinical utility depends on the characteristics of the genetic test, the clinical scenario, and the prevalence, severity, and nature of the disease. For example, observational studies, such as cohort studies, are appropriate for assessing clinical validity because they can accurately measure the ability to predict or detect disease. In contrast, evidence on the clinical utility of a genetic test is ideally studied by using controlled studies, especially RCTs, inasmuch as this design maximizes internal validity and addresses issues of selection bias and confounding. Although an RCT is the ideal study design for assessing many aspects of clinical utility, the level of evidence sufficient to evaluate clinical utility might depend on the specific clinical indication, the clinical setting, and the perceived value of potential outcomes of the genetic test itself.

Thus, there are challenges to evidence-based decision making around the use of genetic tests, including the paucity of the types of research studies on which evidence-based medical decisions depend, which are exacerbated by the accelerated development of new tests, variants, and technologic discoveries. RCTs to establish clinical utility are expensive and of long duration, so decisions about the clinical value of genetic testing are often based on lower levels of evidence. Policy making is often hampered by the lack of objective methods for setting decision thresholds.

  • RESEARCH RECOMMENDATION

DoD has a number of options for contributing to the evidence base for genetic testing. It could undertake or otherwise support prospective studies of clinical utility when promising tests are identified (such as tests that have potential benefit or effects). DoD could conduct prospective studies, such as RCTs, at its facilities or at affiliated institutions, or sponsor outside research.

As another type of contribution to the evidence base, DoD could implement processes for data collection and analysis of its own experience with genetic testing. Information in TRICARE on medical care and outcomes of a large population of active-duty members, retirees, and their family members could help to answer questions about clinical utility that many researchers and organizations struggle to assess. In addition, access to long-term follow-up data on a particular population and its experience with genetic tests could be a valuable resource to generate evidence of clinical utility.

Full participation of the large DoD TRICARE population in evidence repositories (such as ClinVar) offers the opportunity to advance the evidence base on clinical validity of genomic tests substantially. DoD could also contribute to the clinical validity of tests by supporting discovery efforts regarding the evidence base on next-generation sequencing, in particular the association between variants and phenotypes. That would require a large number of people to undergo testing; DoD is in a position to provide such data on its population.

The committee recommends that the Department of Defense advance the evidence base on genetic testing through independent and collaborative research, such as supporting high-quality studies of clinical validity and clinical utility (including patient-centered outcomes) of promising tests; implementing processes for data collection and analysis of its own experience with genetic testing; supporting evidence-based systematic reviews of genetic tests; contributing genetic variants and associated clinical data to public evidence repositories; and partnering with other organizations to facilitate these recommendations.
  • OVERVIEW OF THE PROPOSED FRAMEWORK

After reviewing different methods that have been developed for the evaluation of medical tests, some of which were specifically intended for use with genetic tests, the committee developed a hybrid system that incorporates elements of several of the methods described in the genetic test assessment section above. The system that the committee developed—its decision framework—which consists of seven components, is depicted in Figure S-1 , and described in the text that follows. The framework is not linear nor is it meant to be a formal algorithm, but rather to provide general guidance to the DoD.

Visual representation of the seven components of the proposed decision framework. NOTES: This framework is a guide to help an evaluator reach a decision regarding genetic tests informed by evidence. It guides a user through seven steps to (1) define the (more...)

Component 1. Genetic Test Scenario

As a first step, a “genetic test scenario” must be systematically defined to enable relevant information to be identified. The scenario should include the test being performed, the population in which testing is considered appropriate, the purpose of the test in that population (clinical scenario), the outcomes of interest, and comparable alternative methods to accomplish those tasks. Within any given genetic test scenario, the population being tested and the purpose of the test are often considered together as a “clinical scenario.” In many cases, a particular genetic test could be considered for use in one population for one purpose and in a different population for another purpose. The performance of the test and the decision about whether testing is worthwhile might vary dramatically, depending on the clinical scenario.

If a decision has already been made about the use of one genetic test in a particular clinical scenario (considering the population and the purpose of testing), decisions about the use of other genetic tests in the same test scenario, or decisions about the use of the same test in a different clinical scenario, could be streamlined.

Component 2 (A, B, C). Prioritization of Topics to Be Evaluated and Triage Process

The evaluator must prioritize the various genetic test scenarios to determine which ones to address. That determination could be made based on a number of different criteria (e.g., volume of test requests, unit cost of requested tests, stakeholder assessments, test development planning) depending on DoD's priorities.

The evaluator then carries out an initial triage step to determine if a rapid decision can be made based on existing information. This triage step actually represents three sequential questions (see Box S-1 ). First, the evaluator determines whether the purpose of the test is “worthwhile” (Component 2A). That determination could be based on whether DoD already offers coverage of other genetic or nongenetic tests used for the same purpose. If the answer is “NO” then a decision can be rendered without further evidence review. If the answer is “YES” (because of similar existing coverage decisions) or “MAYBE” (depending on the characteristics of the test or completeness of the knowledge base), the evaluator performs a streamlined “rapid review” of the available evidence for the genetic test scenario (Component 2B).

Process of Triaging Decision About Use of a Genetic Test.

If the rapid review reveals sufficient evidence (YES), or areas of missing evidence are deemed non-essential (MAYBE), the evaluator will need to consider contextual issues (Component 2C). Those might include social issues, net benefits and harms, and aggregate costs. The balance of the issues—that is, whether a test is worthwhile, the evidence, and contextual factors—should inform a triage decision. Those considerations, along with the rapid evidence review, will be brought to the decision process (Component 4). If the rapid review does not reveal sufficient evidence (NO), then a formal evidence-based review process (Component 3) would be appropriate to systematically evaluate the evidence to facilitate a decision process (Component 4) as well as identify evidence gaps (Component 7).

Component 3. Evidence Review Process

For topics that require additional evaluation, a formal evidence assessment process is needed. The first stage of the process is to identify the purpose, important outcomes of testing, and any relevant comparators, such as other genetic or medical tests already used in clinical care for the same purpose. An evidence assessment should be conducted using systematic review methodology, for which ample guidance is available.

Component 4. Decision Process

Following the formal evidence assessment process or availability of new information, DoD will need to apply a decision process for evaluating the results of the evidence review and determine whether the available evidence indicates that the genetic test scenario in question is appropriate for coverage. The decision process will incorporate the preceding evidence review and contextual factors such as social issues, potential harms, or benefit/cost considerations and result in either a “YES” or “NO” decision regarding coverage.

In the absence of direct evidence for many genetic tests, decision makers must decide the acceptable level of uncertainty for their purposes. Even when evidence is strong and clear, the judgment of whether benefits outweigh harms is subjective. Decision makers commonly consider clinical experience, expert opinion, and personal judgments regarding potential harms of the test versus not having the test—with the result that decision makers using identical data can reach different conclusions based on their values.

In addition to questions about whether the available evidence supports the use of the genetic test in a given clinical scenario, this process is likely to also include economic considerations. In that regard, it might be important to consider multiple stakeholder perspectives in the determination and the differences among them. Input from stakeholders might affect the relative weight of net benefits and harms. Ultimately, demand for genetic tests, unit costs, and expert consensus on clinical value will be considered by decision makers.

Thus, because the process of determining what constitutes “sufficient” evidence is subjective and based on value judgments, DoD must set its own standards, preferably in a clear and consistent way, to reflect its values and needs.

Component 5. Decision Repository

A repository of decisions about individual genetic test scenarios will foster the “institutional memory” of the process and facilitate future reviews. Appropriately structured, the repository will allow reviewers to rapidly evaluate new scenarios in light of previous decisions that have been made for similar tests, populations, purposes, outcomes of interest, and methodologies. In addition, the decision repository serves as a record of the value judgments made about whether particular genetic test scenarios are deemed worthwhile, allowing for stakeholders to understand the decisions and so that DoD can ensure consistency across different decisions.

The repository will inform decision makers whether prior decisions have been made for any of the genes on the panel and if coverage has already been decided for one or more of those genes (in the same clinical scenario). For genetic test scenarios that undergo a formal evidence review, the details of the process and the evidentiary thresholds required by DoD will provide transparency to the process and allow identification of evidence gaps that can be addressed by research.

Component 6. Process for Periodic Review and Revision

Given the rapid pace of technologic development and research in genomic medicine, it is critical to develop a mechanism for periodic reassessment of decisions in light of new data. The extent to which this process can be implemented might depend somewhat on the organization of the decision-making process. If designed properly, the decision repository could facilitate a systematic process by including time stamps and automated prompts for re-review of decisions on a certain schedule, or provide a process for stakeholders to request revised decisions.

Component 7. Identification of Evidence Gaps

If the available evidence suggests a potential health benefit associated with a genetic test, especially for tests related to high-priority health conditions, DoD might consider a process for evidence development. Some genetic test scenarios might benefit from traditional research studies to address evidence gaps. Evidence-gap analysis should be conducted on a regular basis as part of a process of continual quality improvement that uses a clinical implementation science process.

  • FRAMEWORK RECOMMENDATION

The committee's recommended framework is based on the numerous frameworks it reviewed and adapts the best of those for its intended purpose. The framework is neither linear nor intended to be a formal algorithm; rather, it provides a conceptual approach as general guidance to DoD. It begins with a clear definition of the topic being considered and a triage process to evaluate whether the purpose of a test is worthwhile and an expedited provisional decision can be made. For topics that need to be evaluated in more detail, the committee provides guidance for conducting an evidence review.

The committee recommends a decision framework, as described in the text above, for the use of genetic tests in clinical care: 1. Define genetic test scenarios on the basis of the clinical setting, the purpose of the test, the population, the outcomes of interest, and comparable alternative methods. 2. For each genetic test scenario, conduct an initial structured assessment to determine whether the test should be covered, denied, or subject to additional evaluation. 3. Conduct or support evidence-based systematic reviews for genetic test scenarios that require additional evaluation. 4. Conduct or support a structured decision process to produce clinical guidance for a genetic test scenario. 5. Publicly share resulting decisions and justification about evaluated genetic test scenarios, and retain decisions in a repository. 6. Implement timely review and revision of decisions on the basis of new data. 7. Identify evidence gaps to be addressed by research.

ACCE takes its name from the four main criteria for evaluating a genetic test: analytic validity, clinical validity, clinical utility, and associated ethical, legal, and social implications.

The Fryback–Thornbury model provides conceptual guidance for evaluating the efficacy of health technologies. It is a widely used general evaluation structure for medical-test assessment and for clarifying the scope of the assessment.

USPSTF is the abbreviated form for the US Preventive Services Task Force.

EGAPP is the abbreviation for the Evaluation of Genomic Applications in Practice and Prevention. The Working Group was established in 2005 to support the development of a systematic process for assessing the available evidence regarding the validity and utility of rapidly emerging genetic tests for clinical practice.

  • Cite this Page National Academies of Sciences, Engineering, and Medicine; Health and Medicine Division; Board on Health Care Services; Board on the Health of Select Populations; Committee on the Evidence Base for Genetic Testing. An Evidence Framework for Genetic Testing. Washington (DC): National Academies Press (US); 2017 Mar 27. Summary.
  • PDF version of this title (1.6M)

In this Page

Recent activity.

  • Summary - An Evidence Framework for Genetic Testing Summary - An Evidence Framework for Genetic Testing

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

Connect with NLM

National Library of Medicine 8600 Rockville Pike Bethesda, MD 20894

Web Policies FOIA HHS Vulnerability Disclosure

Help Accessibility Careers

statistics

thesis statement for genetic testing

Brace yourself, genetic testing might give you more than you bargained for

thesis statement for genetic testing

Lab Head, Metastasis Research Group, Lowy Cancer Rearch Centre, UNSW Sydney

thesis statement for genetic testing

Program Authority and Senior Lecturer, Pharmaceutical Medicine, UNSW Sydney

Disclosure statement

The authors do not work for, consult, own shares in or receive funding from any company or organisation that would benefit from this article, and have disclosed no relevant affiliations beyond their academic appointment.

UNSW Sydney provides funding as a member of The Conversation AU.

View all partners

Drink red wine to prevent cancer. But don’t drink too much! Get some exercise. But don’t overdo it. Give up, it’s all genetic anyway – think of Angelina Jolie!

We are constantly bombarded with conflicting information about our risk of developing cancer. It is difficult to know who to believe, let alone how to respond.

What if you could take a simple test that would reveal your individual risk of developing not only a range of cancers, but hundreds of other diseases? Imagine if it could also tell you which drugs would be most effective for you, if you did develop cancer or other diseases.

The rapidly reducing cost of DNA sequencing has made this one-time fantastical idea an emerging reality. Only 10 years ago it cost about US$10 million to sequence a human genome, so there was little prospect that individuals would, or could, seek out their own unique genetic maps to find out more about their ancestry or their inherited health risks.

Recent advances in genetics mean genetic sequencing is more affordable (US$1,000 to US$3,000) and already guiding treatment across a range of illnesses from cancer to degenerative brain diseases.

New unregulated direct-to-consumer businesses are emerging, making it possible for anyone to order their individual genetic profile by posting off a saliva sample taken at home. But do you really know what you are signing up for?

The age of personalised medicine

Personalised medicine means using a patient’s genome to both predict their likelihood of developing certain diseases, and to guide which treatments are most likely to be effective in a particular individual. It’s also called customised medicine, precision medicine, individualised medicine, bespoke medicine and targeted medicine.

Our genes hold our hereditary information. Every cell in the human body is made up of about 20,000 genes that are passed down from parents to child. Genes contain information that instructs the growth, development and function of the human body. Some genes control simple characteristics such as hair colour and height, others influence complex characteristics such as intelligence. Some genes control how other genes work, telling them when to switch on and off.

thesis statement for genetic testing

We all have alterations, or mutations, in our DNA. Mutations can be passed down from parents to children, or can occur spontaneously, especially as we age. Some are harmless and may determine, for example, whether our ear wax is wet or dry.

However, a mutation in an important gene that prevents it from working properly, or a gene that is missing altogether, can have serious consequences. Early genetic testing focused on debilitating inherited diseases, such as cystic fibrosis and Huntington’s disease, that are caused by mutations in single genes. Tests looked only for a known mutation in a specific gene to confirm or rule out the associated condition.

As testing has become more sophisticated, we have been able to extend this approach to more complex conditions such as cancer. Mutations in two genes called BRCA1 and BRCA2 are associated with an increased risk of developing breast and ovarian cancer, and can be inherited within families.

BRCA1 and BRCA2 normally help clean up mistakes in our DNA that our cells can make when they divide, a process called DNA repair. When either of these genes is altered or mutated, this protective function is disabled, leading to uncontrolled replication of cells with mistakes. This can lead to cancer.

The good news is that we can test for these mutations, and patients can then use the results of this test to assess their risk of developing cancer, and make informed choices. This is the same hereditary genetic mutation that prompted Angelina Jolie to have a preventative double mastectomy two years ago, and preventative surgery to remove her ovaries this year.

The other good news is that in recent years scientists have discovered that patients with mutations in BRCA1 and BRCA2 are exquisitely sensitive to some forms of chemotherapy and a second type of drug called a PARP inhibitor. The same mutation that generates the mistakes in these cells can actually make them more responsive to this drug. Decisions about treatment can then be “personalised” to the individual.

What does the future hold?

Currently, health systems in Australia and overseas do not offer patients the option of sequencing their entire genome as a means of identifying and managing future health risks. Today genetic testing is only available in Australia for specific genes, is tightly regulated and is used only when symptoms are apparent, or a genetic risk is likely, such as a close relative developing a particular cancer or condition.

In five to 10 years’ time, however, we may be facing very different choices, including the option to look for future diseases before they actually occur.

As many cancers do not appear until middle age or later, a young healthy person might discover they have various elevated risks among the many anomalies a DNA test could throw up. Such results might not be provided by a medical professional, but by a commercial operator, and without genetic counselling to explain what they mean to the individual and their family.

thesis statement for genetic testing

What might the implication be of a high-risk result? Should an individual’s relatives be informed, as their risk may also be high, or do they have a right not to know? And what about minors: will parents have the right, or even an obligation, to test babies and children for potential genetic risks, even if medical science offers no prevention or treatment options?

Are we psychologically equipped for these kinds of dilemmas and scientifically literate enough to interpret our own results?

There are currently many reasons to be cautious. First, there are potentially millions of genetic alterations. Most are still not understood. Personalised medicine cannot currently give anyone a comprehensive picture of individual risk simply because far too much remains unknown.

Second, personalised medicine can only indicate elevated risks, it cannot determine whether or not a patient will actually go on to develop a certain type of cancer. Environment and lifestyle also play a big role in our health.

Insurance companies, however, deal entirely in risk. That means genetic profiles could be used to deny higher-risk individuals various types of insurance, or increase their insurance premiums.

Third, health outcomes for some individuals may be based on the financial viability of developing drugs. Many drugs and therapies are currently used for large numbers of patients, making them financially viable for pharmaceutical companies to develop. Genetically targeted cancer drugs, suitable for much smaller groups of patients, may be extremely expensive or might not be brought onto the market at all if society is not willing or cannot afford to pay for them.

Fourth, we may be at risk of eroding our quality of life by creating a new state of “worried wellness”, waiting for disease to strike.

Finally, we may not be sufficiently savvy consumers. New commercial operators are coming onto the global market offering a range of largely unregulated services. Currently, you don’t get much more than details of your ancestry for a US$99 DNA test. But more specialised businesses are emerging that offer , for example, to “identify potential health risks that are present now or may develop in the future”.

Is this just hype, and offering unsubstantiated hope to consumers, or does this represent the first stage of patient empowerment over their own health and lifestyle choices? It will be fascinating to watch this new age of personalised medicine develop in the coming years.

  • Breast cancer
  • Ovarian cancer
  • Personalised medicines
  • Genetic mutation
  • Angelina Jolie
  • Cystic fibrosis
  • Precision medicine
  • Huntington's disease

thesis statement for genetic testing

Admissions Officer

thesis statement for genetic testing

Director of STEM

thesis statement for genetic testing

Community member - Training Delivery and Development Committee (Volunteer part-time)

thesis statement for genetic testing

Chief Executive Officer

thesis statement for genetic testing

Head of Evidence to Action

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

applsci-logo

Article Menu

thesis statement for genetic testing

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Research on ride comfort control of air suspension based on genetic algorithm optimized fuzzy pid.

thesis statement for genetic testing

1. Introduction

2. dynamic modeling of air suspension systems, 2.1. vehicle model construction, 2.2. road input model, 2.3. selection of simulation parameters, 3. control algorithm, 3.1. pid controller, 3.2. fuzzy pid controller, 3.2.1. input and output of fuzzy controller, 3.2.2. selection of input and output variable domains, 3.2.3. membership functions of fuzzy variables, 3.2.4. establishment of fuzzy control rules, 3.3. ga f-pid controller, 3.3.1. selection of objective function, 3.3.2. controller ga optimization algorithm, 4. experimental research, 4.1. air suspension system bench test, 4.2. model validation, 5. simulation analysis, 6. conclusions, author contributions, institutional review board statement, informed consent statement, data availability statement, conflicts of interest.

  • Ming, M. Intelligent Manufacturing Technology of New Energy Vehicle Lightweight Chassis Suspension System. Mach. Build. Autom. 2021 , 50 , 224–226. [ Google Scholar ]
  • Chao, L.Y. Suspension System Optimization and Study on Control of the Electric Vehicle. Master’s Thesis, Qingdao University of Science, Qingdao, China, 2018. [ Google Scholar ]
  • Lei, S.S. Research on Control Strategy of Charging and Deflating Characteristics of Electronically Controlled Active Air Suspension. Master’s Thesis, Taiyuan University of Technology, Taiyuan, China, 2019. [ Google Scholar ]
  • Rui, B. Nonlinear adaptive sliding-mode control of the electronically controlled air suspension system. Int. J. Adv. Robot. Syst. 2019 , 16 , 1729881419881527. [ Google Scholar ] [ CrossRef ]
  • Nazemian, H.; Masih-Tehrani, M. Hybrid Fuzzy-PID Control Development for a Truck Air Suspension System. SAE Int. J. Commer. Veh. 2020 , 13 , 55–69. [ Google Scholar ] [ CrossRef ]
  • Li, H.; Li, S.; Sun, W.; Wang, L.; Lv, D. The optimum matching control and dynamic analysis for air suspension of multi-axle vehicles with anti-roll hydraulically interconnected system. Mech. Syst. Signal Process. 2020 , 139 , 106605. [ Google Scholar ] [ CrossRef ]
  • Nguyen, V.; Jiao, R.; Zhang, J. Control Performance of Damping and Air Spring of Heavy Truck Air Suspension System with Optimal Fuzzy Control. SAE Int. J. Veh. Dyn. Stab. NVH 2020 , 4 , 179–194. [ Google Scholar ] [ CrossRef ]
  • Prassad, S.G.; Malar Mohan, K. A contemporary adaptive air suspension using LQR control for passenger vehicles. ISA Trans. 2019 , 93 , 244–254. [ Google Scholar ]
  • Chen, Y.; Zhang, S.; Mao, E.; Du, Y.; Chen, J.; Yang, S. Height stability control of a large sprayer body based on air suspension using the sliding mode approach. Inf. Process. Agric. 2020 , 7 , 20–29. [ Google Scholar ] [ CrossRef ]
  • Ma, X.F.; Li, K. The Simulation of Air Suspension System Based on PID Control Strategy. Autmobile Appl. Technol. 2017 , 8 , 142–145. [ Google Scholar ] [ CrossRef ]
  • Anh, N.T. Control an active suspension system by using PID and LQR controller. Int. J. Mech. Prod. Eng. Res. Dev. 2020 , 10 , 7003–7012. [ Google Scholar ]
  • Ab Talib, M.H.; Mat Darus, I.Z.; Mohd Samin, P.; Mohd Yatim, H.; Ardani, M.I.; Shaharuddin, N.M.; Hadi, M.S. Vibration control of semi-active suspension system using PID controller with advanced firefly algorithm and particle swarm optimization. J. Ambient. Intell. Humaniz. Comput. 2021 , 12 , 1119–1137. [ Google Scholar ] [ CrossRef ]
  • Tian, M.; Nguyen, V. Control performance of suspension system of cars with PID control based on 3D dynamic model. Journal of Mechanical Engineering. Autom. Control. Syst. 2020 , 1 , 1–10. [ Google Scholar ]
  • Shafiei, B. A Review on PID Control System Simulation of the Active Suspension System of a Quarter Car Model While Hitting Road Bumps. J. Inst. Eng. (India) Ser. C 2022 , 103 , 1001–1011. [ Google Scholar ] [ CrossRef ]
  • Han, S.Y.; Dong, J.F.; Zhou, J.; Chen, Y.H. Adaptive fuzzy PID control strategy for vehicle active suspension based on road evaluation. Electronics 2022 , 11 , 921. [ Google Scholar ] [ CrossRef ]
  • Bashir, A.O.; Rui, X.; Zhang, J. Ride comfort improvement of a semi-active vehicle suspension based on hybrid fuzzy and fuzzy-PID controller. Stud. Inform. Control 2019 , 28 , 421–430. [ Google Scholar ] [ CrossRef ]
  • Zhang, P.; Yue, H.; Shi, Z.; Lin, J.; Chen, Z. A fuzzy pid algorithm-based attitude control method of suspension-type small rail vehicles. J. Vib. Eng. Technol. 2022 , 10 , 111–130. [ Google Scholar ] [ CrossRef ]
  • Ilesanmi, D.; Khumbulani, M.; Adefemi, A.; Kazeem, B. The use of adaptive fuzzy-PID for vibration control in the suspension system of a railcar. In Proceedings of the 2020 IEEE 11th International Conference on Mechanical and Intelligent Manufacturing Technologies (ICMIMT), Cape Town, South Africa, 20–22 January 2020; pp. 130–134. [ Google Scholar ]
  • Chen, G.; Lv, S.; Dai, J. Study on PID control of vehicle semi-active suspension based on genetic algorithm. Int. J. Innov. Comput. Inf. Control 2019 , 15 , 1093–1114. [ Google Scholar ]
  • Wang, Y.; Chen, K.; Dong, M. Research on active suspension system of heavy commercial vehicle controlled by PID controller based on genetic algorithm. In Advances in Precision Instruments and Optical Engineering, Proceedings of the International Conference on Precision Instruments and Optical Engineering, Guangzhou, China, 20–22 August 2021 ; Springer Nature: Singapore, 2022; pp. 555–566. [ Google Scholar ]
  • Mudduluru, S.R.; Chizari, M. Quarter and full car models optimisation of passive and active suspension system using genetic algorithm. arXiv 2021 , arXiv:2101.12629. [ Google Scholar ]
  • Rajamani, R. Vehicle Dynamics and Control ; Springer: New York, NY, USA, 2011. [ Google Scholar ]
  • GB 7031-1987 ; Vehicle Vibration Input—Road Flatness Representation. China Standard Press: Beijing, China, 1987.

Click here to enlarge figure

NameNotationValue
Sprung massm /kg500
Unsprung massm /kg50
Suspension dampingc/N·s·m 1700
Tire stiffnessk /N·m 200,000
ParametersRise TimeOvershootSetting TimeSteady-State Error
K DecreaseIncreaseSlight ChangeDecrease
K DecreaseIncreaseIncreaseEliminate
K Slight ChangeDecreaseDecreaseSlight Change
eec
NBNMNSZOPSPMPB
NBPBPBPMPMPSZOZO
NMPBPBPMPSPSZONS
NSPMPMPMPSZONSNS
ZOPMPMPSZONSNMNM
PSPSPSZONSNSNMNM
PMPSZONSNMNMNMNB
PBZOZONMNMNMNBNB
eec
NBNMNSZOPSPMPB
NBNBNBNMNMNSZOZO
NMNBNBNMNSNSZOZO
NSNBNMNSNSZOPSPS
ZONMNMNSZOPSPMPM
PSNMNSZOPSPSPMPB
PMZOZOPSPSPMPBPB
PBZOZOPSPMPMPBPB
eec
NBNMNSZOPSPMPB
NBPSNSNBNBNBNMPS
NMPSNSNBNMNMNSZO
NSZONSNMNMNSNSZO
ZOZONSNSNSNSNSZO
PSZOZOZOZOZOZOZO
PMPBPSPSPSPSPSPB
PBPBPMPMPMPSPSPB
ParametersExplanationParametersExplanation
Encoding MethodBinary EncodingMutation FunctionConstrained Adaptive Mutation
Initial PopulationRandomly Generate Within Specified BoundsCrossover Probability0.9
Population Size30Mutation Function0.1
Selection FunctionRandom Uniform SelectionMaximum Evolution Generations100
Crossover FunctionDiversified CrossoverStopping Generation100
Evaluation CriteriaPassive SuspensionFuzzy PIDGA-Fuzzy PIDPercentage of Fuzzy PID OptimizationPercentage of GA-Fuzzy PID Optimization
Vehicle Vertical Acceleration
(m/s )
1.04730.81690.754122%28%
Suspension Wheel Travel
(m)
0.01560.01290.011717%25%
Dynamic Tire Load
(kN)
4.34544.12814.04125%7%
Evaluation CriteriaPassive SuspensionFuzzy PIDGA-Fuzzy PIDPercentage of Fuzzy PID OptimizationPercentage of GA-Fuzzy PID Optimization
Vehicle Vertical Acceleration
(m/s )
1.80621.39081.264323%30%
Suspension Wheel Travel
(m)
0.03370.02760.024918%26%
Dynamic Tire Load
(kN)
6.08955.72415.54146%9%
Evaluation CriteriaPassive SuspensionFuzzy PIDGA-Fuzzy PIDPercentage of Fuzzy PID OptimizationPercentage of GA-Fuzzy PID Optimization
Vehicle Vertical Acceleration
(m/s )
2.76382.07291.907025%31%
Suspension Wheel Travel
(m)
0.06110.04890.044620%27%
Dynamic Tire Load
(kN)
8.83788.39598.13085%8%
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Zhang, S.; Li, M.; Li, J.; Xu, J.; Wang, Z.; Liu, S. Research on Ride Comfort Control of Air Suspension Based on Genetic Algorithm Optimized Fuzzy PID. Appl. Sci. 2024 , 14 , 7787. https://doi.org/10.3390/app14177787

Zhang S, Li M, Li J, Xu J, Wang Z, Liu S. Research on Ride Comfort Control of Air Suspension Based on Genetic Algorithm Optimized Fuzzy PID. Applied Sciences . 2024; 14(17):7787. https://doi.org/10.3390/app14177787

Zhang, Shaobo, Mei Li, Jinsong Li, Jie Xu, Zelong Wang, and Shuaihang Liu. 2024. "Research on Ride Comfort Control of Air Suspension Based on Genetic Algorithm Optimized Fuzzy PID" Applied Sciences 14, no. 17: 7787. https://doi.org/10.3390/app14177787

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

COMMENTS

  1. Genetic Testing Thesis Statement Examples

    There are many methods of genetic testing. For example, in buccal smear, a small brush is used to collect the samples of the cells present inside the surface of cheek. Other methods include taking samples from blood, saliva, hair, skin, amniotic fluid, etc. All the methods have their own indications. When the samples are received by the ...

  2. PDF Genetic Testing, Interpretation, and Communication: Exploration Across

    testing. As genetic tests are increasingly incorporated into both clinical and research settings, it is critical that we understand the implications of such tests for both individuals and their family members. Questions of whether, when, and how to pursue genetic testing, interpret genetic test results, and communicate those results to family ...

  3. PDF The Ethics of Genetic Testing and Screening in the U.S. Military

    screening is an unnecessary invasion of privacy, breach of confidentiality, and restriction of. choice and, therefore, violates the autonomy of SM. The military should not be permitted to mandate genetic testing or screening on. prospective or current active or reserve component service members without robust genetic.

  4. Ethics of Genetic Testing

    As noted by Miller, (2007), genetic screening entails the systematic establishment of hereditary health risks that a person may have and thus may be used to discriminate people with high risk traits from the employment. Although it is true that the concept of genetic testing has many ethical concerns, the greatest concern of genetic testing at ...

  5. Attitudes towards genetic testing: The role of genetic literacy ...

    Understanding reasons for why people choose to have or not to have a genetic test is essential given the ever-increasing use of genetic technologies in everyday life. The present study explored the multiple drivers of people's attitudes towards genetic testing. Using the International Genetic Literacy and Attitudes Survey (iGLAS), we collected data on: (1) willingness to undergo testing; (2 ...

  6. PDF Direct-to-consumer Genetic Ancestry Testing: a Morally Objectionable

    Abstract. The popularity of direct-to-consumer (DTC) genetic ancestry testing has resulted in a. surge of life-changing news for consumers and their families. DTC genetic ancestry testing. companies and third-party websites have provided consumers with new ways to create family. trees and connect with unknown relatives.

  7. Ethical issues in predictive genetic testing: a public health

    Genetic testing is a relatively new and rapidly emerging field; yet, public health has been involved in public policies and practices involving genetic testing for conditions such as phenylketonuria and sickle cell disease for some time. 1 This article will address earlier and currently emerging issues related to public health, primary care practice, and genetic testing.

  8. Social, Legal, and Ethical Implications of Genetic Testing

    Each new genetic test that is developed raises serious issues for medicine, public health, and social policy regarding the circumstances under which the test should be used, how the test is implemented, and what uses are made of its results. ... The American Society of Human Genetics issued a statement on DNA banking and DNA data banking in ...

  9. Genetic Counseling Theses and Dissertations

    Variant Reclassification in Cancer Genetic Testing: Are Genetic Counselors Prepared? A Review of Current Practices, Niamh Gemma White. PDF. Exploring How the Risk of Sudden Cardiac Death is Discussed in Families with a Diagnosis of a SADS Condition, Kristin Anne Wiley. Theses/Dissertations from 2014 PDF

  10. Human Genetics Theses

    Evaluation of Patient and Provider Satisfaction with a Point of Care Genetic Testing Model for Cancer Patients, Andrea Verdes and Molli S. Goodman. Theses from 2023 PDF. Implementation of a Patient Decision Aid to Facilitate the Selection of a Pre-Test Genetic Counseling Model, Taylor Berry, Valerie Schimmenti, and Magan Trottier. PDF

  11. PDF Genetic/Genomic Research

    GENETIC/GENOMIC RESEARCH. This guidance document is intended for researchers using participants' samples (blood, saliva, or other tissue), genetic or genomic data, and/or associated health information. Should you need additional assistance, please contact OPHS at 510-642-7461 or [email protected]. Table of Contents.

  12. PDF PERSUASION TOWARDS GENETIC TESTING by Ashlyn Lisner the Department of

    This study assessed the effect of two health-information variables related to interest in a. getting a genetic test related to disease symptoms: (1) perceived risk of disease and (2) uncertainty surrounding the penetrance of the gene or the proportion of individuals with. the gene that manifest the associated trait or phenotype.

  13. PDF Technical report: ethical and policy issues in genetic testing and

    236 ACMG PoliCy StAteMent Ethical and policy issues in genetic testing and screening of children testing is likely to avert unneeded medical interventions,36 but other possible adverse effects ...

  14. Impact of Genetic Testing on Human Health:

    Applied Clinical Genomics is the application of genetic information to the clinical setting, including improved diagnosis of disease and tailored treatment efficacy and safety. By discovering and defining the genes that underlie susceptibility to disorders, genetic information can be used to identify and better define those genes that play ...

  15. Awareness, knowledge, perceptions, and attitudes towards genetic

    Background Genetic testing for risk of hereditary cancer can help patients to make important decisions about prevention or early detection. US and UK studies show that people from ethnic minority groups are less likely to receive genetic testing. It is important to understand various groups' awareness of genetic testing and its acceptability to avoid further disparities in health care. This ...

  16. Rethinking the "open future" argument against predictive genetic

    Historically, two primary ethical arguments have underpinned the consensus against predictive genetic testing in the position statements of leading professional societies. 1,2,3,4,5,6,7,8 One is ...

  17. PDF A Thesis

    example, the genetic test for breast and ovarian cancer is recommended for those with a family history of the disease. Many people have not heard of genetic testing (Mogilner 1998). If an individual has never heard of genetic testing, he or she cannot effectively evaluate the risks and benefits to decide if testing is right for them.

  18. 40+ Thesis Statement Examples for All Essay Types

    Understanding the basics of genetic inheritance can explain how traits are passed from parents to offspring and how genetic disorders can occur. Thesis Statement Examples For Compare And Contrast Essays. A thesis statement for a compare and contrast essay highlights the similarities and differences between two or more subjects. It helps readers ...

  19. Genetic Testing: Advantages and Disadvantages Essay

    Get a custom essay on Genetic Testing: Advantages and Disadvantages. At the same time, I acknowledge all the benefits that genetic testing can bring in terms of diagnosing a wide range of diseases and conditions. Fearing that they might discover hereditary predispositions to some untreatable diseases, many people choose not to get tested.

  20. Summary

    Summary. Advances in genetics and genomics are transforming medical practice, resulting in a dramatic growth of genetic testing in the health care system. The rapid development of new technologies, however, has also brought challenges, including the need for rigorous evaluation of the validity and utility of genetic tests, questions regarding ...

  21. What to say to a parent who requests genetic testing

    If parents unreasonably request or refuse testing of their child, the physician should take steps to change or, if necessary, use legal means to override the parents' choice. (g) Encourage parents to share genetic information with the child in a manner appropriate to the child's stage of development. (h) Ensure that parents/guardians are ...

  22. Thesis Statement On Genetic Testing

    910 Words. 4 Pages. Open Document. I am agreed with this statement. Generally genetic testing is to settle on medical choices is smart thought. Fundamentally Genetic testing is a method for figuring out whether you are conveying a genetic change which can possibly bring about a medical condition. The testing identifies changes in the genes ...

  23. Brace yourself, genetic testing might give you more than you bargained for

    Published: June 30, 2015 4:10pm EDT. Some genes control simple characteristics such as hair colour and height, others influence more complex characteristics like intelligence. Insurance companies ...

  24. Research on Ride Comfort Control of Air Suspension Based on Genetic

    The air suspension system's superior variable stiffness, low vibration frequency, and resistance to road impacts significantly elevate both the comfort of vehicle occupants and the overall ride quality. By effectively controlling the air suspension system, its superior characteristics can be fully exploited to enhance the overall performance of vehicles. However, the parameter tuning process ...