NSGC

  • About Genetic Counselors
  • NSGC Leadership
  • Reach NSGC Members
  • Job Connection
  • Awards & Grants
  • In Remembrance
  • Contact NSGC
  • Member Benefits
  • Military SIG
  • State Chapters
  • Student Corner
  • Registration
  • Hotel & Travel
  • Sponsors & Exhibitors
  • NSGC Online Education Center
  • Genetic Counselor Awareness Day
  • NSGC Member Webinar Series
  • NSGC Podcast Series
  • Past and Future Annual Conferences
  • Continuing Education Unit (CEU) Information
  • Request a Speaker
  • Advocacy Documents
  • Send a letter to your Member of Congress
  • State Licensure for Genetic Counselors
  • Justice, Equity, Diversity and Inclusion (J.E.D.I.)
  • Disciplinary Procedures
  • Position Statements
  • Practice Guidelines
  • Code of Ethics & Conflict of Interest
  • Genetic Counselor Workforce
  • NSGC Endorsements
  • Professional Status Survey
  • Genetic Counseling Literature Repository
  • Perspectives in Genetic Counseling

Journal of Genetic Counseling

  • Expert Media Panel

Search NSGC

SUBMIT AN ARTICLE EDITORIAL BOARD JOURNAL CEUs

The Journal of Genetic Counseling , the official journal of the National Society of Genetic Counselors, is a timely, international forum addressing all aspects of the discipline and practice of genetic counseling. The journal focuses on the critical questions and problems that arise at the interface between rapidly advancing technological developments and the concerns of individuals and communities at genetic risk. The publication provides genetic counselors, other clinicians and health educators, laboratory geneticists, bioethicists, legal scholars, social scientists, and other researchers with a premier resource on genetic counseling topics in national, international, and cross-national contexts. Starting in 2019, the journal is published by Wiley.

NSGC Member Access

NSGC members receive free access to the Journal of Genetic Counseling through Wiley. In order to access the journal you must be logged in .

Please note that NSGC Full, Emeritus, New Genetic Counselor, and Associate Members should access the journal through the member link above to ensure full access. Student Members do not receive access to the Journal of Genetic Counseling .

  • Genetic Counseling Theory, Methods and Practice
  • Public Health, Public Policy, and Access and Genetics Service Delivery
  • Practice Guideline
  • Practice Resource
  • Education and Genetics Professional Workforce Issues
  • Ethical, Legal, Psychological, and Social Issues
  • Minority and Health Disparities
  • Risk Assessment
  • Special Issue

Genetic Counseling Theory, Methods and Practice | 9.26.2023

Patients' attitudes regarding genetic counseling before germline brca1/2 pathogenic variants testing in taiwan: a single-country, multi-center, patient-reported outcome study.

research about genetic counselling

What knowledge is required for an informed choice related to non-invasive prenatal screening?

Ethical, legal, psychological, and social issues | 9.26.2023, current attitudes toward carrier screening for spinal muscular atrophy among pregnant women in eastern china, prenatal patient perceptions of receiving difficult news over the telephone.

  • Search Menu
  • Browse content in Arts and Humanities
  • Browse content in Archaeology
  • Anglo-Saxon and Medieval Archaeology
  • Archaeological Methodology and Techniques
  • Archaeology by Region
  • Archaeology of Religion
  • Archaeology of Trade and Exchange
  • Biblical Archaeology
  • Contemporary and Public Archaeology
  • Environmental Archaeology
  • Historical Archaeology
  • History and Theory of Archaeology
  • Industrial Archaeology
  • Landscape Archaeology
  • Mortuary Archaeology
  • Prehistoric Archaeology
  • Underwater Archaeology
  • Urban Archaeology
  • Zooarchaeology
  • Browse content in Architecture
  • Architectural Structure and Design
  • History of Architecture
  • Residential and Domestic Buildings
  • Theory of Architecture
  • Browse content in Art
  • Art Subjects and Themes
  • History of Art
  • Industrial and Commercial Art
  • Theory of Art
  • Biographical Studies
  • Byzantine Studies
  • Browse content in Classical Studies
  • Classical History
  • Classical Philosophy
  • Classical Mythology
  • Classical Literature
  • Classical Reception
  • Classical Art and Architecture
  • Classical Oratory and Rhetoric
  • Greek and Roman Epigraphy
  • Greek and Roman Law
  • Greek and Roman Papyrology
  • Greek and Roman Archaeology
  • Late Antiquity
  • Religion in the Ancient World
  • Digital Humanities
  • Browse content in History
  • Colonialism and Imperialism
  • Diplomatic History
  • Environmental History
  • Genealogy, Heraldry, Names, and Honours
  • Genocide and Ethnic Cleansing
  • Historical Geography
  • History by Period
  • History of Emotions
  • History of Agriculture
  • History of Education
  • History of Gender and Sexuality
  • Industrial History
  • Intellectual History
  • International History
  • Labour History
  • Legal and Constitutional History
  • Local and Family History
  • Maritime History
  • Military History
  • National Liberation and Post-Colonialism
  • Oral History
  • Political History
  • Public History
  • Regional and National History
  • Revolutions and Rebellions
  • Slavery and Abolition of Slavery
  • Social and Cultural History
  • Theory, Methods, and Historiography
  • Urban History
  • World History
  • Browse content in Language Teaching and Learning
  • Language Learning (Specific Skills)
  • Language Teaching Theory and Methods
  • Browse content in Linguistics
  • Applied Linguistics
  • Cognitive Linguistics
  • Computational Linguistics
  • Forensic Linguistics
  • Grammar, Syntax and Morphology
  • Historical and Diachronic Linguistics
  • History of English
  • Language Acquisition
  • Language Evolution
  • Language Reference
  • Language Variation
  • Language Families
  • Lexicography
  • Linguistic Anthropology
  • Linguistic Theories
  • Linguistic Typology
  • Phonetics and Phonology
  • Psycholinguistics
  • Sociolinguistics
  • Translation and Interpretation
  • Writing Systems
  • Browse content in Literature
  • Bibliography
  • Children's Literature Studies
  • Literary Studies (Asian)
  • Literary Studies (European)
  • Literary Studies (Eco-criticism)
  • Literary Studies (Romanticism)
  • Literary Studies (American)
  • Literary Studies (Modernism)
  • Literary Studies - World
  • Literary Studies (1500 to 1800)
  • Literary Studies (19th Century)
  • Literary Studies (20th Century onwards)
  • Literary Studies (African American Literature)
  • Literary Studies (British and Irish)
  • Literary Studies (Early and Medieval)
  • Literary Studies (Fiction, Novelists, and Prose Writers)
  • Literary Studies (Gender Studies)
  • Literary Studies (Graphic Novels)
  • Literary Studies (History of the Book)
  • Literary Studies (Plays and Playwrights)
  • Literary Studies (Poetry and Poets)
  • Literary Studies (Postcolonial Literature)
  • Literary Studies (Queer Studies)
  • Literary Studies (Science Fiction)
  • Literary Studies (Travel Literature)
  • Literary Studies (War Literature)
  • Literary Studies (Women's Writing)
  • Literary Theory and Cultural Studies
  • Mythology and Folklore
  • Shakespeare Studies and Criticism
  • Browse content in Media Studies
  • Browse content in Music
  • Applied Music
  • Dance and Music
  • Ethics in Music
  • Ethnomusicology
  • Gender and Sexuality in Music
  • Medicine and Music
  • Music Cultures
  • Music and Religion
  • Music and Media
  • Music and Culture
  • Music Education and Pedagogy
  • Music Theory and Analysis
  • Musical Scores, Lyrics, and Libretti
  • Musical Structures, Styles, and Techniques
  • Musicology and Music History
  • Performance Practice and Studies
  • Race and Ethnicity in Music
  • Sound Studies
  • Browse content in Performing Arts
  • Browse content in Philosophy
  • Aesthetics and Philosophy of Art
  • Epistemology
  • Feminist Philosophy
  • History of Western Philosophy
  • Metaphysics
  • Moral Philosophy
  • Non-Western Philosophy
  • Philosophy of Science
  • Philosophy of Language
  • Philosophy of Mind
  • Philosophy of Perception
  • Philosophy of Action
  • Philosophy of Law
  • Philosophy of Religion
  • Philosophy of Mathematics and Logic
  • Practical Ethics
  • Social and Political Philosophy
  • Browse content in Religion
  • Biblical Studies
  • Christianity
  • East Asian Religions
  • History of Religion
  • Judaism and Jewish Studies
  • Qumran Studies
  • Religion and Education
  • Religion and Health
  • Religion and Politics
  • Religion and Science
  • Religion and Law
  • Religion and Art, Literature, and Music
  • Religious Studies
  • Browse content in Society and Culture
  • Cookery, Food, and Drink
  • Cultural Studies
  • Customs and Traditions
  • Ethical Issues and Debates
  • Hobbies, Games, Arts and Crafts
  • Lifestyle, Home, and Garden
  • Natural world, Country Life, and Pets
  • Popular Beliefs and Controversial Knowledge
  • Sports and Outdoor Recreation
  • Technology and Society
  • Travel and Holiday
  • Visual Culture
  • Browse content in Law
  • Arbitration
  • Browse content in Company and Commercial Law
  • Commercial Law
  • Company Law
  • Browse content in Comparative Law
  • Systems of Law
  • Competition Law
  • Browse content in Constitutional and Administrative Law
  • Government Powers
  • Judicial Review
  • Local Government Law
  • Military and Defence Law
  • Parliamentary and Legislative Practice
  • Construction Law
  • Contract Law
  • Browse content in Criminal Law
  • Criminal Procedure
  • Criminal Evidence Law
  • Sentencing and Punishment
  • Employment and Labour Law
  • Environment and Energy Law
  • Browse content in Financial Law
  • Banking Law
  • Insolvency Law
  • History of Law
  • Human Rights and Immigration
  • Intellectual Property Law
  • Browse content in International Law
  • Private International Law and Conflict of Laws
  • Public International Law
  • IT and Communications Law
  • Jurisprudence and Philosophy of Law
  • Law and Politics
  • Law and Society
  • Browse content in Legal System and Practice
  • Courts and Procedure
  • Legal Skills and Practice
  • Primary Sources of Law
  • Regulation of Legal Profession
  • Medical and Healthcare Law
  • Browse content in Policing
  • Criminal Investigation and Detection
  • Police and Security Services
  • Police Procedure and Law
  • Police Regional Planning
  • Browse content in Property Law
  • Personal Property Law
  • Study and Revision
  • Terrorism and National Security Law
  • Browse content in Trusts Law
  • Wills and Probate or Succession
  • Browse content in Medicine and Health
  • Browse content in Allied Health Professions
  • Arts Therapies
  • Clinical Science
  • Dietetics and Nutrition
  • Occupational Therapy
  • Operating Department Practice
  • Physiotherapy
  • Radiography
  • Speech and Language Therapy
  • Browse content in Anaesthetics
  • General Anaesthesia
  • Neuroanaesthesia
  • Browse content in Clinical Medicine
  • Acute Medicine
  • Cardiovascular Medicine
  • Clinical Genetics
  • Clinical Pharmacology and Therapeutics
  • Dermatology
  • Endocrinology and Diabetes
  • Gastroenterology
  • Genito-urinary Medicine
  • Geriatric Medicine
  • Infectious Diseases
  • Medical Toxicology
  • Medical Oncology
  • Pain Medicine
  • Palliative Medicine
  • Rehabilitation Medicine
  • Respiratory Medicine and Pulmonology
  • Rheumatology
  • Sleep Medicine
  • Sports and Exercise Medicine
  • Clinical Neuroscience
  • Community Medical Services
  • Critical Care
  • Emergency Medicine
  • Forensic Medicine
  • Haematology
  • History of Medicine
  • Browse content in Medical Dentistry
  • Oral and Maxillofacial Surgery
  • Paediatric Dentistry
  • Restorative Dentistry and Orthodontics
  • Surgical Dentistry
  • Browse content in Medical Skills
  • Clinical Skills
  • Communication Skills
  • Nursing Skills
  • Surgical Skills
  • Medical Ethics
  • Medical Statistics and Methodology
  • Browse content in Neurology
  • Clinical Neurophysiology
  • Neuropathology
  • Nursing Studies
  • Browse content in Obstetrics and Gynaecology
  • Gynaecology
  • Occupational Medicine
  • Ophthalmology
  • Otolaryngology (ENT)
  • Browse content in Paediatrics
  • Neonatology
  • Browse content in Pathology
  • Chemical Pathology
  • Clinical Cytogenetics and Molecular Genetics
  • Histopathology
  • Medical Microbiology and Virology
  • Patient Education and Information
  • Browse content in Pharmacology
  • Psychopharmacology
  • Browse content in Popular Health
  • Caring for Others
  • Complementary and Alternative Medicine
  • Self-help and Personal Development
  • Browse content in Preclinical Medicine
  • Cell Biology
  • Molecular Biology and Genetics
  • Reproduction, Growth and Development
  • Primary Care
  • Professional Development in Medicine
  • Browse content in Psychiatry
  • Addiction Medicine
  • Child and Adolescent Psychiatry
  • Forensic Psychiatry
  • Learning Disabilities
  • Old Age Psychiatry
  • Psychotherapy
  • Browse content in Public Health and Epidemiology
  • Epidemiology
  • Public Health
  • Browse content in Radiology
  • Clinical Radiology
  • Interventional Radiology
  • Nuclear Medicine
  • Radiation Oncology
  • Reproductive Medicine
  • Browse content in Surgery
  • Cardiothoracic Surgery
  • Gastro-intestinal and Colorectal Surgery
  • General Surgery
  • Neurosurgery
  • Paediatric Surgery
  • Peri-operative Care
  • Plastic and Reconstructive Surgery
  • Surgical Oncology
  • Transplant Surgery
  • Trauma and Orthopaedic Surgery
  • Vascular Surgery
  • Browse content in Science and Mathematics
  • Browse content in Biological Sciences
  • Aquatic Biology
  • Biochemistry
  • Bioinformatics and Computational Biology
  • Developmental Biology
  • Ecology and Conservation
  • Evolutionary Biology
  • Genetics and Genomics
  • Microbiology
  • Molecular and Cell Biology
  • Natural History
  • Plant Sciences and Forestry
  • Research Methods in Life Sciences
  • Structural Biology
  • Systems Biology
  • Zoology and Animal Sciences
  • Browse content in Chemistry
  • Analytical Chemistry
  • Computational Chemistry
  • Crystallography
  • Environmental Chemistry
  • Industrial Chemistry
  • Inorganic Chemistry
  • Materials Chemistry
  • Medicinal Chemistry
  • Mineralogy and Gems
  • Organic Chemistry
  • Physical Chemistry
  • Polymer Chemistry
  • Study and Communication Skills in Chemistry
  • Theoretical Chemistry
  • Browse content in Computer Science
  • Artificial Intelligence
  • Computer Architecture and Logic Design
  • Game Studies
  • Human-Computer Interaction
  • Mathematical Theory of Computation
  • Programming Languages
  • Software Engineering
  • Systems Analysis and Design
  • Virtual Reality
  • Browse content in Computing
  • Business Applications
  • Computer Security
  • Computer Games
  • Computer Networking and Communications
  • Digital Lifestyle
  • Graphical and Digital Media Applications
  • Operating Systems
  • Browse content in Earth Sciences and Geography
  • Atmospheric Sciences
  • Environmental Geography
  • Geology and the Lithosphere
  • Maps and Map-making
  • Meteorology and Climatology
  • Oceanography and Hydrology
  • Palaeontology
  • Physical Geography and Topography
  • Regional Geography
  • Soil Science
  • Urban Geography
  • Browse content in Engineering and Technology
  • Agriculture and Farming
  • Biological Engineering
  • Civil Engineering, Surveying, and Building
  • Electronics and Communications Engineering
  • Energy Technology
  • Engineering (General)
  • Environmental Science, Engineering, and Technology
  • History of Engineering and Technology
  • Mechanical Engineering and Materials
  • Technology of Industrial Chemistry
  • Transport Technology and Trades
  • Browse content in Environmental Science
  • Applied Ecology (Environmental Science)
  • Conservation of the Environment (Environmental Science)
  • Environmental Sustainability
  • Environmentalist Thought and Ideology (Environmental Science)
  • Management of Land and Natural Resources (Environmental Science)
  • Natural Disasters (Environmental Science)
  • Nuclear Issues (Environmental Science)
  • Pollution and Threats to the Environment (Environmental Science)
  • Social Impact of Environmental Issues (Environmental Science)
  • History of Science and Technology
  • Browse content in Materials Science
  • Ceramics and Glasses
  • Composite Materials
  • Metals, Alloying, and Corrosion
  • Nanotechnology
  • Browse content in Mathematics
  • Applied Mathematics
  • Biomathematics and Statistics
  • History of Mathematics
  • Mathematical Education
  • Mathematical Finance
  • Mathematical Analysis
  • Numerical and Computational Mathematics
  • Probability and Statistics
  • Pure Mathematics
  • Browse content in Neuroscience
  • Cognition and Behavioural Neuroscience
  • Development of the Nervous System
  • Disorders of the Nervous System
  • History of Neuroscience
  • Invertebrate Neurobiology
  • Molecular and Cellular Systems
  • Neuroendocrinology and Autonomic Nervous System
  • Neuroscientific Techniques
  • Sensory and Motor Systems
  • Browse content in Physics
  • Astronomy and Astrophysics
  • Atomic, Molecular, and Optical Physics
  • Biological and Medical Physics
  • Classical Mechanics
  • Computational Physics
  • Condensed Matter Physics
  • Electromagnetism, Optics, and Acoustics
  • History of Physics
  • Mathematical and Statistical Physics
  • Measurement Science
  • Nuclear Physics
  • Particles and Fields
  • Plasma Physics
  • Quantum Physics
  • Relativity and Gravitation
  • Semiconductor and Mesoscopic Physics
  • Browse content in Psychology
  • Affective Sciences
  • Clinical Psychology
  • Cognitive Psychology
  • Cognitive Neuroscience
  • Criminal and Forensic Psychology
  • Developmental Psychology
  • Educational Psychology
  • Evolutionary Psychology
  • Health Psychology
  • History and Systems in Psychology
  • Music Psychology
  • Neuropsychology
  • Organizational Psychology
  • Psychological Assessment and Testing
  • Psychology of Human-Technology Interaction
  • Psychology Professional Development and Training
  • Research Methods in Psychology
  • Social Psychology
  • Browse content in Social Sciences
  • Browse content in Anthropology
  • Anthropology of Religion
  • Human Evolution
  • Medical Anthropology
  • Physical Anthropology
  • Regional Anthropology
  • Social and Cultural Anthropology
  • Theory and Practice of Anthropology
  • Browse content in Business and Management
  • Business Strategy
  • Business Ethics
  • Business History
  • Business and Government
  • Business and Technology
  • Business and the Environment
  • Comparative Management
  • Corporate Governance
  • Corporate Social Responsibility
  • Entrepreneurship
  • Health Management
  • Human Resource Management
  • Industrial and Employment Relations
  • Industry Studies
  • Information and Communication Technologies
  • International Business
  • Knowledge Management
  • Management and Management Techniques
  • Operations Management
  • Organizational Theory and Behaviour
  • Pensions and Pension Management
  • Public and Nonprofit Management
  • Strategic Management
  • Supply Chain Management
  • Browse content in Criminology and Criminal Justice
  • Criminal Justice
  • Criminology
  • Forms of Crime
  • International and Comparative Criminology
  • Youth Violence and Juvenile Justice
  • Development Studies
  • Browse content in Economics
  • Agricultural, Environmental, and Natural Resource Economics
  • Asian Economics
  • Behavioural Finance
  • Behavioural Economics and Neuroeconomics
  • Econometrics and Mathematical Economics
  • Economic Systems
  • Economic History
  • Economic Methodology
  • Economic Development and Growth
  • Financial Markets
  • Financial Institutions and Services
  • General Economics and Teaching
  • Health, Education, and Welfare
  • History of Economic Thought
  • International Economics
  • Labour and Demographic Economics
  • Law and Economics
  • Macroeconomics and Monetary Economics
  • Microeconomics
  • Public Economics
  • Urban, Rural, and Regional Economics
  • Welfare Economics
  • Browse content in Education
  • Adult Education and Continuous Learning
  • Care and Counselling of Students
  • Early Childhood and Elementary Education
  • Educational Equipment and Technology
  • Educational Strategies and Policy
  • Higher and Further Education
  • Organization and Management of Education
  • Philosophy and Theory of Education
  • Schools Studies
  • Secondary Education
  • Teaching of a Specific Subject
  • Teaching of Specific Groups and Special Educational Needs
  • Teaching Skills and Techniques
  • Browse content in Environment
  • Applied Ecology (Social Science)
  • Climate Change
  • Conservation of the Environment (Social Science)
  • Environmentalist Thought and Ideology (Social Science)
  • Natural Disasters (Environment)
  • Social Impact of Environmental Issues (Social Science)
  • Browse content in Human Geography
  • Cultural Geography
  • Economic Geography
  • Political Geography
  • Browse content in Interdisciplinary Studies
  • Communication Studies
  • Museums, Libraries, and Information Sciences
  • Browse content in Politics
  • African Politics
  • Asian Politics
  • Chinese Politics
  • Comparative Politics
  • Conflict Politics
  • Elections and Electoral Studies
  • Environmental Politics
  • European Union
  • Foreign Policy
  • Gender and Politics
  • Human Rights and Politics
  • Indian Politics
  • International Relations
  • International Organization (Politics)
  • International Political Economy
  • Irish Politics
  • Latin American Politics
  • Middle Eastern Politics
  • Political Methodology
  • Political Communication
  • Political Philosophy
  • Political Sociology
  • Political Behaviour
  • Political Economy
  • Political Institutions
  • Political Theory
  • Politics and Law
  • Public Administration
  • Public Policy
  • Quantitative Political Methodology
  • Regional Political Studies
  • Russian Politics
  • Security Studies
  • State and Local Government
  • UK Politics
  • US Politics
  • Browse content in Regional and Area Studies
  • African Studies
  • Asian Studies
  • East Asian Studies
  • Japanese Studies
  • Latin American Studies
  • Middle Eastern Studies
  • Native American Studies
  • Scottish Studies
  • Browse content in Research and Information
  • Research Methods
  • Browse content in Social Work
  • Addictions and Substance Misuse
  • Adoption and Fostering
  • Care of the Elderly
  • Child and Adolescent Social Work
  • Couple and Family Social Work
  • Developmental and Physical Disabilities Social Work
  • Direct Practice and Clinical Social Work
  • Emergency Services
  • Human Behaviour and the Social Environment
  • International and Global Issues in Social Work
  • Mental and Behavioural Health
  • Social Justice and Human Rights
  • Social Policy and Advocacy
  • Social Work and Crime and Justice
  • Social Work Macro Practice
  • Social Work Practice Settings
  • Social Work Research and Evidence-based Practice
  • Welfare and Benefit Systems
  • Browse content in Sociology
  • Childhood Studies
  • Community Development
  • Comparative and Historical Sociology
  • Economic Sociology
  • Gender and Sexuality
  • Gerontology and Ageing
  • Health, Illness, and Medicine
  • Marriage and the Family
  • Migration Studies
  • Occupations, Professions, and Work
  • Organizations
  • Population and Demography
  • Race and Ethnicity
  • Social Theory
  • Social Movements and Social Change
  • Social Research and Statistics
  • Social Stratification, Inequality, and Mobility
  • Sociology of Religion
  • Sociology of Education
  • Sport and Leisure
  • Urban and Rural Studies
  • Browse content in Warfare and Defence
  • Defence Strategy, Planning, and Research
  • Land Forces and Warfare
  • Military Administration
  • Military Life and Institutions
  • Naval Forces and Warfare
  • Other Warfare and Defence Issues
  • Peace Studies and Conflict Resolution
  • Weapons and Equipment

Genetic Counseling Research: A Practical Guide

Genetic Counseling Research: A Practical Guide

Assistant Professor of Psychology

Professor of Educational Psychology

Professor and Director of the Graduate Program of Study in Genetic Counseling

  • Cite Icon Cite
  • Permissions Icon Permissions

Genetic Counseling Research: A Practical Guide is an online resource devoted to research methodology in genetic counseling. It offers step-by-step guidance for conducting research, from the development of a question to the publication of findings. Genetic counseling examples and practical tips guide readers through the research and publication processes. With a highly accessible, pedagogical approach, this resource will help promote quality research by genetic counselors and research supervisors—and in turn, increase the knowledge base for genetic counseling practice, other aspects of genetic counseling service delivery, and professional education.

Signed in as

Institutional accounts.

  • Google Scholar Indexing
  • GoogleCrawler [DO NOT DELETE]

Personal account

  • Sign in with email/username & password
  • Get email alerts
  • Save searches
  • Purchase content
  • Activate your purchase/trial code

Institutional access

  • Sign in with a library card Sign in with username/password Recommend to your librarian
  • Institutional account management
  • Get help with access

Access to content on Oxford Academic is often provided through institutional subscriptions and purchases. If you are a member of an institution with an active account, you may be able to access content in one of the following ways:

IP based access

Typically, access is provided across an institutional network to a range of IP addresses. This authentication occurs automatically, and it is not possible to sign out of an IP authenticated account.

Sign in through your institution

Choose this option to get remote access when outside your institution. Shibboleth/Open Athens technology is used to provide single sign-on between your institution’s website and Oxford Academic.

  • Click Sign in through your institution.
  • Select your institution from the list provided, which will take you to your institution's website to sign in.
  • When on the institution site, please use the credentials provided by your institution. Do not use an Oxford Academic personal account.
  • Following successful sign in, you will be returned to Oxford Academic.

If your institution is not listed or you cannot sign in to your institution’s website, please contact your librarian or administrator.

Sign in with a library card

Enter your library card number to sign in. If you cannot sign in, please contact your librarian.

Society Members

Society member access to a journal is achieved in one of the following ways:

Sign in through society site

Many societies offer single sign-on between the society website and Oxford Academic. If you see ‘Sign in through society site’ in the sign in pane within a journal:

  • Click Sign in through society site.
  • When on the society site, please use the credentials provided by that society. Do not use an Oxford Academic personal account.

If you do not have a society account or have forgotten your username or password, please contact your society.

Sign in using a personal account

Some societies use Oxford Academic personal accounts to provide access to their members. See below.

A personal account can be used to get email alerts, save searches, purchase content, and activate subscriptions.

Some societies use Oxford Academic personal accounts to provide access to their members.

Viewing your signed in accounts

Click the account icon in the top right to:

  • View your signed in personal account and access account management features.
  • View the institutional accounts that are providing access.

Signed in but can't access content

Oxford Academic is home to a wide variety of products. The institutional subscription may not cover the content that you are trying to access. If you believe you should have access to that content, please contact your librarian.

For librarians and administrators, your personal account also provides access to institutional account management. Here you will find options to view and activate subscriptions, manage institutional settings and access options, access usage statistics, and more.

Our books are available by subscription or purchase to libraries and institutions.

  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Rights and permissions
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

ASHG

Realizing the benefits of human genetics and genomics research for people everywhere.

Genetic Counseling Researchers Shape the Future of Precision Medicine 

Guest Post By: Heather A. Zierhut, PhD, MS, CGC; and Adam H. Buchanan, MS, MPH, LGC 

Genetic counselors’ advanced training in – you guessed it – genetics and counseling, plays a key role in healthcare. Genetic counselors can guide and support people seeking more information about how inherited diseases and conditions might affect them or their families, and help interpret test results. While many genetic counselors work directly with patients, others focus on research to help in the development of new or improved treatment or care for people with genetic conditions. In honor of  Genetic Counselor Awareness Day  on Nov. 8, National Society of Genetic Counselors (NSGC) members Heather A. Zierhut, PhD, MS, CGC, and Adam H. Buchanan, MS, MPH, LGC, share how their research is shaping the future of genetic counseling and precision medicine.

NSGC: What inspired you to be a genetic counselor?  

Heather:  I think there are two main categories of people who become genetic counselors: those who have a long exploration into the profession or those who have a light bulb moment. My entrance into genetic counseling was like a laser light show. My undergraduate genetics professor briefly mentioned the career. Like any good researcher, I went to my tiny matchbook size dorm room and looked up everything I could possibly find on the topic. Fireworks were going off in my mind. Genetic counseling was love at first literature search.

Adam:  I’m one of those long exploration types. In college I cycled through several possible career choices. My career research was a great source of stories about the (sometimes gross) lives of a veterinary technician or hospital employee who cleans up after surgeries. But, nothing grabbed me until I worked with a genetic counselor while doing my thesis in a master’s of public health program. I loved the way genetic counseling combined the communication skills I was learning as a public health educator with the connections you make while interacting with patients.

NSGC: How did you first get involved in genetic counseling research?  

Heather:  Like many aspiring Principal Investigators, my gateway to research was undergraduate summer programs. I loved the idea of being able to set up experiments, ask questions, and get answers all in a three-month time span. But after experimenting in yeast and fruit fly labs, I found myself wanting to research something different. I switched gears to my ultimate organism of interest, humans. It all started with two genetics icons, Janet and Marc Williams, giving me a small research project looking at outcomes of early hearing screening tests. Even after making the decision to pursue genetic counseling, I never lost my desire for research. I searched for a training program where I could find mentors to support my interest in clinical care and research. The desire for research mentors was first and foremost in my decision to go to the University of Minnesota.

Adam:  When my kids were little and their favorite question was “why?”, I knew they came by it honestly – it’s always been one of my favorite questions, too. As a genetic counselor I sought out ways to answer the whys that arose in my clinical work. Like Heather, I was fortunate to have colleagues willing to indulge those questions and a mentor who looked for opportunities to involve me in all phases of research during my research apprenticeship – including writing and reviewing manuscripts and learning the ins and outs of grant proposal preparation. So, when I was driving home from a rural cancer genetics clinic one day and had a light bulb moment about how to more efficiently improve access to care, I had the tools to turn that spark into a viable research project.

NSGC: What do you enjoy about genetic counseling research?  

Heather:  You either love and appreciate the ups and downs of research or you gut out the experimentation period in hope and anticipation of the day you get those final results. As much as I wish I could say that I love to stop and smell the research roses, I do not. I am a destination data person. Getting results is my favorite part of research. It’s like the best holidays all wrapped into one called Data Day. Knowing this about myself, I chose to pursue a doctoral degree in genetics and epidemiology with research projects that give me data sooner rather than later.

Adam:  I’m totally with Heather on her love of Data Day. I also love thinking about the research process as crafting a story. If everything falls into place, you develop this beautiful narrative that arcs from why your research topic is important to patients all the way to how the study you’ve proposed will generate data that will help those patients. It’s a challenging process that takes a lot of tinkering and being able to communicate the big picture while having your details in order. That’s why it’s such great fun when it works.

NSGC:   What advice do you have for aspiring genetic counseling researchers? 

Heather:  I shifted research directions completely after completing my PhD. The shift was the right direction for my newly developed skill set, but I needed help getting started. I reached out to people with expertise and asked them to assist me. I applied for numerous early investigator grants and was told repeatedly that I needed more experience in the field. Not letting it slow me down too much, I started collecting pilot data to build my case. Others in the field started to take notice and I was asked to be on several national committees. These combined efforts led to funding through the Jane Engelberg Memorial Fellowship, a grant intended to promote the professional development of individual counselors and to improve the practice of genetic counseling. This took persistence and a following a slightly different path than I had envisioned. My advice is don’t be afraid to take risks and get out of your comfort zone.

Adam:  All research starts with a good idea. And our clinical experience can provide a wealth of research ideas. Pay attention to the whys that linger in your mind after seeing patients. Also, researchers fail. Sometimes spectacularly – unfunded grant proposals, rejected manuscripts, and studies that go off the rails. This is a particular challenge for the over-achievers who make up the ranks of genetics professionals. But, it’s an inevitable part of doing research that moves the field forward. Keep your head up, rely on the mentors Heather mentioned – they’re out there and they really do want to help – and learn from everything. Finally, find a research focus that gets you jazzed. It’s a lot easier to be persistent when you’re excited about your work.

NSGC: What genetic counseling research are you working on now?

Heather:  My current research is on the vast under-diagnosis of familial hypercholesterolemia (FH) and issues that arise in communication of genetic information in FH families. There are 1 in 250 people out there with a treatable genetic cause of high cholesterol. It’s my research and my job to make sure that people with FH know the risks and have opportunities to prevent heart disease in their families. But as with anything having to do with families and communication, it’s complex. Our research team chips away at finding ways to screen families and help them become aware of their heart disease risks. We hope to give an opportunity to end the cycle of heart disease in their family.

Adam:  I’m leading a group that was just funded by the NIH to study how children and their parents react to receiving genomic results for conditions that do not occur until adulthood (such as hereditary breast and ovarian cancer syndrome). There are lots of intriguing questions we’re hoping to answer in this study, particularly about the psychosocial impact of this information and how family members use it to guide their healthcare. It’s a topic that has been written about a bunch, but there’s barely any empirical data to guide whether and how to provide this information to families in practice. We’re looking forward to adding our story to the mix and helping families in the process.

Heather Zierhut is the associate director of the University of Minnesota Graduate Program of Study in Genetic Counseling and assistant professor in Genetics, Cell Biology, and Development. Current areas of research include the psychosocial and public implications involved with the provision of genetic counseling services, implementations of whole genome sequencing, and outcomes of genetic counseling. 

Adam Buchanan is an assistant professor in the Genomic Medicine Institute at Geisinger, Co-Director of the MyCode Genomic Screening and Counseling program, and a member of the American Board of Genetic Counseling. His research focuses on access to genetic counseling and assessing behavioral and psychosocial outcomes of genetic counseling. 

ASHG uses cookies to provide you with a secure and custom web experience. Privacy Policy

Characterization of genetic counselor practices in inpatient care settings

Affiliations.

  • 1 Department of Pediatrics, Baylor College of Medicine, San Antonio, TX, USA.
  • 2 Department of Molecular and Human Genetics, Baylor College of Medicine, Houston, TX, USA.
  • 3 Stanford Center for Inherited Cardiovascular Disease, Stanford, CA, USA.
  • 4 Department of Pediatrics, Division of Medical Genetics, Stanford University, Stanford, CA, USA.
  • 5 Smidt Heart Institute Cedars Sinai Medical Center, Los Angeles, CA, USA.
  • 6 Department of Medical & Molecular Genetics, Indiana University School of Medicine, Indianapolis, IN, USA.
  • 7 Department of Epidemiology, Indiana University Fairbanks School of Public Health, Indianapolis, IN, USA.
  • PMID: 33713511
  • DOI: 10.1002/jgc4.1401

Rapid genomic testing is increasingly used in inpatient settings for diagnostic and treatment purposes. With the expansion of genetic testing in this setting, requests for inpatient genetics consultations have increased. There have been reports of genetic counselors working in inpatient care, though their specific roles are not well described. In this study, we characterized the roles of genetic counselors practicing in inpatient care settings in the United States and Canada. Genetic counselors were recruited via professional organization listservs to complete an online survey. The survey gathered information on participants' roles and workflow of inpatient genetics consultation services at their institution. Responses from 132 participants demonstrate that 50.4% of genetic counselors cover genetics consultations as needed or on a rotating schedule (34.6%). They practice in general pediatric (59.1%), neonatal (42.5%), cancer (28.3%), and/or prenatal (18.9%) specialties, among others. Participants reported working independently (16.1%) or with other providers (54.8%), including geneticists and other attending physicians. The workflow of genetics consultation services varies between institutions in the delivery of consults, members of the inpatient genetics consultation care team, and administrative support. Fifty percent of participants reported having no exposure to inpatients during graduate training, and 87.3% of participants reported receiving no institutional training for their inpatient role. This is the first study to describe roles of genetic counselors in inpatient care. It establishes a foundation for future research on inpatient genetic counseling and genetic counseling outcomes in inpatient services. As demand for genetics expertise in inpatient care grows, genetic counselors can be hired to serve inpatient populations alongside genetics and non-genetics providers.

Keywords: genetic counselors; genetics services; inpatient genetics; professional development; workforce.

© 2021 National Society of Genetic Counselors.

Publication types

  • Research Support, Non-U.S. Gov't
  • Counselors*
  • Genetic Counseling
  • Genetic Testing
  • Infant, Newborn
  • Surveys and Questionnaires
  • United States

Genetic Counseling

genetic counslor talking to a couple

What is Genetic Counseling?

Genetic counseling gives you information about how genetic conditions might affect you or your family. The genetic counselor or other healthcare professional will collect your personal and family health history . They can use this information to determine how likely it is that you or your family member has a genetic condition. Based on this information, the genetic counselor can help you decide whether a genetic test might be right for you or your relative.

Learn more about genetic counseling in the time of COVID-19 .

Reasons for Genetic Counseling

Based on your personal and family health history, your doctor can refer you for genetic counseling. There are different stages in your life when you might be referred for genetic counseling:

  • Genetic conditions that run in your family or your partner’s family
  • History of infertility, multiple miscarriages, or stillbirth
  • Previous pregnancy or child affected by a birth defect  or genetic condition
  • Assisted Reproductive Technology (ART) options
  • Previous pregnancy or child affected by a birth defect or genetic condition
  • Abnormal test results, such as a blood test, ultrasound, Chorionic Villus Sampling (CVS) , or amniocentesis
  • Maternal infections, such as Cytomegalovirus (CMV) , and other exposures such as medicines , drugs, chemicals, and x-rays
  • Genetic screening that is recommended for all pregnant women, which includes cystic fibrosis , sickle cell disease , and any conditions that run in your family or your partner’s family
  • Abnormal newborn screening  results
  • Birth defects
  • Intellectual disability or developmental disabilities
  • Autism spectrum disorders (ASD)
  • Vision or hearing  problems
  • Hereditary breast and ovarian cancer (HBOC) syndrome
  • Lynch syndrome (hereditary colorectal and other cancers)
  • Familial hypercholesterolemia
  • Muscular dystrophy and other muscle diseases
  • Inherited movement disorders such as Huntington’s disease
  • Inherited blood disorders such as sickle cell disease

Following your genetic counseling session, you might decide to have genetic testing. Genetic counseling after testing can help you better understand your test results and treatment options, help you deal with emotional concerns, and refer you to other healthcare providers and advocacy and support groups.

There are various ways to access genetic counseling services, including in person, by phone, and by video conference.

Find a genetic counselor using the National Society of Genetic Counselors directory.

Find a genetics clinic using the American College of Medical Genetics and Genomics Genetics Clinics Database.

Disclaimer: The Centers for Disease Control and Prevention (CDC) does not provide counseling, diagnoses, or personal medical advice. Please see your healthcare provider for these needs. Linking to a non-federal site does not constitute an endorsement by CDC or any of its employees of the sponsors or the information and products presented on the site.

Exit Notification / Disclaimer Policy

  • The Centers for Disease Control and Prevention (CDC) cannot attest to the accuracy of a non-federal website.
  • Linking to a non-federal website does not constitute an endorsement by CDC or any of its employees of the sponsors or the information and products presented on the website.
  • You will be subject to the destination website's privacy policy when you follow the link.
  • CDC is not responsible for Section 508 compliance (accessibility) on other federal or private website.
  • Introduction to Genomics
  • Educational Resources
  • Policy Issues in Genomics
  • The Human Genome Project
  • Funding Opportunities
  • Funded Programs & Projects
  • Division and Program Directors
  • Scientific Program Analysts
  • Contact by Research Area
  • News & Events
  • Research Areas
  • Research investigators
  • Research Projects
  • Clinical Research
  • Data Tools & Resources
  • Genomics & Medicine
  • Family Health History
  • For Patients & Families
  • For Health Professionals
  • Jobs at NHGRI
  • Training at NHGRI
  • Funding for Research Training
  • Professional Development Programs
  • NHGRI Culture
  • Social Media
  • Broadcast Media
  • Image Gallery
  • Press Resources
  • Organization
  • NHGRI Director
  • Mission & Vision
  • Policies & Guidance
  • Institute Advisors
  • Strategic Vision
  • Leadership Initiatives
  • Diversity, Equity, and Inclusion
  • Partner with NHGRI
  • Staff Search

JHU/NIH Genetic Counseling Training Program

Drawing on resources from three outstanding research institutions, the National Human Genome Research Institute (NHGRI) and National Cancer Institute (NCI) at the National Institutes of Health (NIH) and the Department of Health, Behavior and Society at the Johns Hopkins University (JHU) Bloomberg School of Public Health have collaborated to develop and support the JHU/NIH Genetic Counseling Training Program (GCTP), a competitive graduate program that addresses the growing need for genetic counseling services.

More about GCTP

​GCTP Clinical and Research Training | NHGRI

Introduction

The accelerated discovery of disease and susceptibility genes made possible by the sequencing of the human genome has brought new and exciting challenges to the field of genetic counseling. The traditional emphasis of genetic counseling has been on providing information coupled with supportive counseling, primarily for people facing reproductive decisions. The field has broadened dramatically to address a multiplicity of emerging needs, ranging from clients seeking disease susceptibility testing to those wanting to know if a therapeutic treatment option is right for them. Therefore, genetic counselors must not only convey to these individuals and their families information about risks but also the consequences of testing and the potential for therapeutic intervention. These choices are laden with uncertainty and raise difficult ethical, legal and social issues.

Genetic information can have profound psychological meaning for clients, particularly for members of families affected by a genetic condition or risk. Decisions about whether to use genetic tests require that clients evaluate scientific information in the context of their personal values and beliefs. Genetic counselors are trained to facilitate decision-making to promote informed choices. When there is no genetic test or therapeutic option to offer, genetic counselors help family members to adapt to the condition or risk, often under conditions of uncertainty.

As the scope of genetic counseling expands and evolves, patient, professional and community education will be imperative. Increasingly, primary care practitioners are providing aspects of genetic counseling and other genetic services, resulting in a need to educate nurses, social workers and physicians. Genetic counselors play a key role in educating these providers and developing standards of practice. Trained genetic counselors also provide a means for health professionals and patients to communicate with policy makers, the media and the public about new and emerging genetic technologies and services.

Graduate Program Overview

Mission Statement

The JHU/NIH Genetic Counseling Training Program shapes genetic counseling services through student and faculty research and develops outstanding genetic counselors who are innovators and leaders in: 

  • psychotherapeutic genetic counseling,
  • genetic counseling research and scholarship, 
  • applications of genomics and precision health 
  • transdisciplinary learning and practice, incorporating perspectives from public health, policy, ethics, and advocacy  

Vision 

Genetic counseling clinician scholars transform evidence-based genomic healthcare.

Read the  JHU/NIH GCTP's entire strategic plan , along with our program’s objectives. 

View our  Frequently Asked Questions  about the training program.

The GCTP distinguishes itself in offering extensive interactive coursework to support completion of high quality, publishable thesis studies. Since its inception, the GCTP has produced a cadre of genetic counselors who broaden the scope of genetic counseling by contributing to a growing research literature that critically examines a variety of aspects of the profession and shapes future directions in the field.

Additionally, program faculty provide students with one-on-one supervision for an hour each week throughout their graduate studies. These sessions offer students feedback based on audiotaped sessions with clients and on interventions consistent with development of counseling expertise.

Johns Hopkins Bloomberg School of Public Health provides a strong academic home for the GCTP, while NHGRI and NCI provide funding, instruction and leadership. This collaborative program, which represents the first allocation of federal funds to support graduate education in genetic counseling, is regarded as an important effort to address new challenges resulting from genomics research.

Our financial package for all admitted students consists of three parts:

  • A 30% tuition reduction granted by The Johns Hopkins Bloomberg School of Public Health: The formula dictates that Johns Hopkins pays for the first $3000 of tuition and then pays for 30% of the remainder.  This applies for the duration of the 2.5 years of the program. For the 2021-22 academic year, full tuition is $59,184 for the year. This means that Johns Hopkins will pay $19,855 toward the first-year tuition, and each student’s tuition bill would be $39,329. See JHU School of Public Health Tuition and Fees .  
  • A $10,000 first year scholarship paid by the NIH  This reduces the cost of the first year of tuition to $29,329. A predoctoral training stipend given by the NIH to all US citizens or permanent residents. See NIH Predoctoral IRTA and Visiting Fellow Stipend Levels .

group of people

An applicant's potential to become a skilled, compassionate, and self-aware counselor is an absolute criterion for admission. The greatest consideration will be given to applicants whom we consider best able to advance the profession of genetic counseling through their leadership in research and their interface with the fields of public health, public policy and health education.

An admissions committee, composed of the  Executive Committee , reviews the entire application package and considers the overall balance of each applicant's qualifications.

Unfortunately this year we are only able to take applicants who are U.S. citizens or permanent residents. Exceptions may be made for  international (non-U.S. citizen) applicants with a prior doctoral level degree, such as an M.D. or Ph.D.  If you think you may meet the advanced degree criteria and wish to discuss further, please contact the program admin at  [email protected] for more information. 

Requirements

  • Completion of an undergraduate degree
  • Completion of undergraduate level courses in biochemistry and genetics.
  • Counseling experience, either paid or voluntary.  This experience should provide the applicant with an opportunity to work with individuals in emotional distress and to reflect on that process.   

In addition, prior coursework in psychology and statistics is strongly recommended.

The Johns Hopkins Bloomberg School of Public Health requires applicants from countries where English is not the primary, official language to submit scores on an English Language Proficiency Test . 

The NIH/JHU GCTP does NOT require GRE scores or other standardized test scores.  However, we will consider GRE scores if they are submitted as part of our holistic admissions review process. 

Our admissions committee relies on a holistic review process, taking into account academic performance, counseling experience, other relevant work and life experiences, and the match between the  applicant’s training goals and our program’s goals and objectives. Given this process, we do not have a specific GPA requirement.  However, a strong academic background does provide the admissions committee with one helpful indicator of future success. 

Applications must be submitted and completed by the December 1 deadline to be considered.

Invitations for interviews will be presented in February, with final decisions and offers made in April.

The JHU/NIHProgram supports diversification of the genetic counseling profession. A diverse profession leads to a richer professional dialogue and enhances genetic counseling for minority communities. Therefore, candidates from a wide variety of backgrounds and experiences are welcomed. We especially encourage candidates from groups currently under-represented in the genetic counseling profession, including people with disabilities, men, and people from ethnic and racial minorities.

The Maryland and DC Society of Genetic Counselors (MDCGC) may have application fee reimbursements available for some prospective students. For more information and to apply for a reimbursement, please visit the Student Application Fee Reimbursement .

How to Apply

You should use the  standard application form  for the Johns Hopkins Bloomberg School of Public Health. You will be following instructions and deadlines for the ScM degree in the Department of Health, Behavior and Society. This application website will provide instructions for submitting all supporting documents, which include:

  • Official transcripts from their undergraduate institution(s)
  • curriculum vitae or resume.
  • Three letters of recommendation (at least one of which should be an academic recommendation from an instructor or advisor).
  • A personal statement.
  • When appropriate, scores from an English proficiency exam.
  • Official scores from the GRE exam (optional).

The Johns Hopkins/National Institutes of Health Genetic Counseling Training Program is participating in the Genetic Counseling Admissions Match through National Matching Services (NMS) beginning with admissions for Fall 2018. The GC Admissions Match has been established to enhance the process of placing applicants into positions in masters-level genetic counseling programs that are accredited by the Accreditation Council for Genetic Counseling (ACGC). The Match uses a process that takes into account both applicants' and programs' preferences. All applicants must first register for the Match with NMS before applying to participating genetic counseling graduate programs. At the conclusion of all program interviews, both applicants and programs will submit ranked lists of preferred placements to NMS according to deadlines posted on the NMS website. The binding results of the Match will be released to both applicants and programs simultaneously in late April.

Please  visit the NMS website  to register for the match, review detailed information about the matching process, and to view a demonstration of how the matching algorithm works.

For Fall 2019 admissions, there is not a formal place for entering the NMS match number into the online Johns Hopkins application.

*Please put your NMS match number in the header of your personal statement. *

Additional Inquiries

Specific inquiries regarding the program should be addressed to:

Lori Hamby Erby, ScM, Ph.D., C.G.C. Phone: 301-443-2635 Email:  [email protected]

Or to the Program Coordinators:

Ellie Younger Phone: 301-439-6639 Email: [email protected]

Match Fee Waiver Process

The Association of Genetic Counseling Program Directors/Genetic Counselor Educations Association (AGCP has established a waiver for the $100 fee associated with the Genetic Counseling Admissions Match. Prospective students who demonstrate financial need AND have a cumulative GPA of 3.0 or greater can apply for a Match fee waiver. Prospective students should apply for a fee waiver BEFORE registering for a Match Code Number.

Applications for the Match fee waiver are due by October 4, 2023. Applicants will be notified by October 30, 2023 as to whether they have received a waiver. There are a limited number of waivers, so not all applicants will receive a waiver. Instructions for how to register for the GC Admissions Match will be provided to those who receive a waiver.

Waivers are non-transferable to future Match cycles and cannot be transferred by the recipient to another prospective student. Prospective students who are granted fee waivers must register for the Match by January 1, 2024 . Any waiver not used by January 1, 2024 will be rescinded and may be granted by AGCPD to a different prospective student.

Prospective students who are applying for a Match fee waiver will be required to write a short essay and to upload at least one of the following documents demonstrating financial need:

  • A copy of a letter verifying unemployment benefits received within the past two years.
  • A copy of a letter on official letterhead from a government agency verifying that you or your family have qualified for public assistance based on low income criteria anytime within the past 5 years
  • A copy of your financial aid award letter from another / previous institution dated within 5 the past years
  • A copy of your approved GRE fee waiver (for tests taken within 5 years of the application deadline)
  • A copy of your federal Student Aid Report (SAR) that verifies you qualified for financial aid within the past 5 years based on a family contribution of:| o Not more than $1,500 if the student is a dependent student o Not more than $1,900 if the student is an independent student  

To apply, register for the match .

Please direct any questions about the Match Fee Waiver process or requirements to [email protected]

Faculty and Clinical Supervisors

The faculty consists of tenured investigators, scientists and health care providers at the National Human Genome Research Institute (NHGRI) at the National Institutes of Health (NIH) and the Johns Hopkins University (JHU). Leadership for this joint program is provided by the faculty who serve on the  Executive Committee .

The genetic counseling coursework, student supervision, coordination of clinical rotations and Accreditation Council for Genetic Counseling (ACGC) accreditation are overseen by:

  • Program Director Lori Erby, Sc.M., Ph.D. , a genetic counselor and scientist in the NHGRI Medical Genomics and Metabolic Genetics Branch   
  • Associate Program Director Megan Cho, ScM.    
  • Associate Director of Cancer Genomics Leila Jamal, Sc.M., Ph.D.    
  • Charles Venditti, M.D., Ph.D., senior investigator, Medical Genomics and Metabolic Genetics Branch, NHGRI, serves as medical director of the program.   

Program coordination is provided by Ellie Younger at NIH.

Past Meetings and Workshops

  • (Factor) Analyze This: PCA or EFA   July 31, 2015: Sam Wolford, Ph.D., PSTAT®, C.Q.E, Professor of Statistics and Director, Center for Quantitative Analysis, Bentley University explains the similarities and differences between Principal Components Analysis (PCA) and Exploratory Factor Analysis (EFA).  View slides from the workshop    
  • 2014 GCTP Alumni Research Symposium   August 1, 2014: The GCTP hosted it's second full-day alumni research symposium, which was attended by approximately half of the alumni. During the symposium, alumni were invited to give presentations about current research projects. All alumni participated in seeking feedback from the group about potential future research endeavors.

Genetic Counseling Resources

To learn more about genetics, the field of genetic counseling and organizations that work in these areas, click on the links below.

  • National Society of Genetic Counselors - About Genetic Counselors
  • National Society of Genetic Counselors - Student Corner
  • American Society of Human Genetics
  • American Board of Genetic Counseling
  • Information for Genetic Professionals
  • American Board of Medical Genetics and Genomics
  • American College of Medical Genetics and Genomics
  • National Organization for Rare Disorders
  • Genetic Alliance

Last updated: March 6, 2024

  • Program Finder
  • Admissions Services
  • Course Directory
  • Academic Calendar
  • Hybrid Campus
  • Lecture Series
  • Convocation
  • Strategy and Development
  • Implementation and Impact
  • Integrity and Oversight
  • In the School
  • In the Field
  • In Baltimore
  • Resources for Practitioners
  • Articles & News Releases
  • In The News
  • Statements & Announcements
  • At a Glance
  • Student Life
  • Strategic Priorities
  • Inclusion, Diversity, Anti-Racism, and Equity (IDARE)
  • What is Public Health?

Master of Science (ScM) in Genetic Counseling

Offered By: Department of Health, Behavior and Society

Onsite | Full-Time | 2.5 years

  • MAS Application Fee Waiver Requirements
  • Master of Arts (MA) in Geography and Environmental Engineering
  • Master of Arts and Master of Science in Public Health (MA/MSPH)
  • Master of Arts in Public Health Biology (MAPHB)
  • Master of Bioethics (MBE)
  • MHA Frequently Asked Questions
  • Mission, Vision, and Values
  • MHA Executive in Residence and Alumni
  • Student Experience
  • Program Outcomes
  • Bachelor's/MHA Program
  • Master of Health Science (MHS) - Department of Biochemistry and Molecular Biology
  • Master of Health Science (MHS) - Department of Epidemiology
  • Alumni Update
  • MHS Combined with a Certificate Program
  • Master of Health Science (MHS) - Department of Molecular Microbiology and Immunology
  • Alumni Highlights
  • Post-Baccalaureate Program in Environmental Health for Pre-Medicine Students
  • Bachelor's/MHS in Health Economics and Outcomes Research
  • MHS HEOR Careers
  • Frequently Asked Questions
  • Master of Health Science (MHS)
  • Concurrent School-Wide Master of Health Science Program in Biostatistics
  • Master of Health Science - Department of Population, Family and Reproductive Health
  • Master of Health Science Online (MHS) - Department of Population, Family and Reproductive Health
  • Careers in Health Economics
  • Core Competencies
  • Meet the Director
  • What is Health Economics
  • MPH Capstone Schedule
  • Concentrations
  • Online/Part-Time Format
  • Requirements

Tuition and Funding

  • Executive Board Faculty
  • Master of Science (MS) in Geography and Environmental Engineering
  • Independent Professional Project and Final Essay 
  • Program Objectives and Outcomes
  • Internships
  • Master of Science (ScM) - Department of Biochemistry and Molecular Biology
  • Master of Science (ScM) - Department of Biostatistics
  • Master of Science (ScM) - Department of Epidemiology
  • Master of Science (ScM) - Department of Molecular Microbiology and Immunology
  • ScM Faculty Advisers
  • Master of Science in Engineering (MSE) in Geography and Environmental Engineering
  • Bachelor's/MSPH in Health Policy
  • FAQ for MSPH in Health Policy
  • Field Placement Experience
  • MSPH Capstone
  • MSPH Practicum
  • Required and Elective Courses
  • Student Timeline
  • Career Opportunities
  • 38-Week Dietetics Practicum
  • Completion Requirements
  • MSPH/RD Program FAQ
  • Program Goals
  • Master's Essay Titles
  • Application Fee Waiver Requirements
  • Doctor of Philosophy (PhD) - Department of Biostatistics
  • Doctor of Philosophy (PhD) - Department of Epidemiology
  • Program Goals and Expectations
  • Doctor of Philosophy (PhD) - Department of Molecular Microbiology and Immunology
  • Doctor of Philosophy (PhD) - Department of Population, Family and Reproductive Health
  • Doctor of Philosophy (PhD) in Clinical Investigation
  • Track in Environmental Sustainability, Resilience, and Health
  • Track in Exposure Sciences and Environmental Epidemiology
  • Track in Health Security
  • Track in Toxicology, Physiology and Molecular Mechanisms
  • PhD in Geography and Environmental Engineering Faculty Advisers
  • Recent Graduates and Dissertation Titles
  • PhD Funding
  • PhD TA Requirement
  • Recent Dissertation Titles
  • JHU-Tsinghua Doctor of Public Health
  • Core Course Requirements
  • Concentration in Women’s and Reproductive Health
  • Custom Track
  • Concentration in Environmental Health
  • Concentration in Global Health: Policy and Evaluation
  • Concentration in Health Equity and Social Justice
  • Concentration in Health Policy and Management
  • Concentration in Implementation Science
  • Meet Current Students
  • Combined Bachelor's / Master's Programs
  • Concurrent MHS Option for BSPH Doctoral Students
  • Concurrent MSPH Option for JHSPH Doctoral students
  • Doctor of Medicine and Doctor of Philosophy (MD/PhD)
  • Adolescent Health Certificate Program
  • Bioethics Certificate Program
  • Climate and Health Certificate Program
  • Clinical Trials Certificate Program
  • Community- Based Public Health Certificate Program
  • Demographic Methods Certificate Program
  • Environmental and Occupational Health Certificate Program
  • Epidemiology for Public Health Professionals Certificate Program
  • Evaluation: International Health Programs Certificate Program
  • Food Systems, the Environment and Public Health Certificate Program
  • Frequently Asked Questions for Certificate Programs
  • Gender and Health Certificate Program
  • Gerontology Certificate Program
  • Global Digital Health Certificate Program
  • Global Health Certificate Program
  • Global Health Practice Certificate Program
  • Health Communication Certificate Program
  • Health Disparities and Health Inequality Certificate Program
  • Health Education Certificate Program
  • Health Finance and Management Certificate Program
  • Health and Human Rights Certificate Program
  • Healthcare Epidemiology and Infection Prevention and Control Certificate Program
  • Humane Sciences and Toxicology Policy Certificate Program
  • Humanitarian Health Certificate Program
  • Implementation Science and Research Practice Certificate Program
  • Injury and Violence Prevention Certificate Program
  • International Healthcare Management and Leadership Certificate Program
  • Leadership for Public Health and Healthcare Certificate Program
  • Lesbian, Gay, Bisexual, Transgender, and Queer (LGBTQ) Public Health Certificate Program
  • Maternal and Child Health Certificate Program
  • Mental Health Policy, Economics and Services Certificate Program
  • Non-Degree Students General Admissions Info
  • Pharmacoepidemiology and Drug Safety Certificate Program
  • Population Health Management Certificate Program
  • Population and Health Certificate Program
  • Product Stewardship for Sustainability Certificate Program
  • Public Health Advocacy Certificate Program
  • Public Health Economics Certificate Program
  • Public Health Informatics Certificate Program
  • Public Health Practice Certificate Program
  • Declaration of Intent - Public Health Preparedness
  • Public Health Training Certificate for American Indian Health Professionals
  • Public Mental Health Research Certificate Program
  • Quality, Patient Safety and Outcomes Research Certificate Program
  • Quantitative Methods in Public Health Certificate Program
  • Requirements for Successful Completion of a Certificate Program
  • Rigor, Reproducibility, and Responsibility in Scientific Practice Certificate Program
  • Risk Sciences and Public Policy Certificate Program
  • Spatial Analysis for Public Health Certificate Program
  • Training Certificate in Public Health
  • Tropical Medicine Certificate Program
  • Tuition for Certificate Programs
  • Vaccine Science and Policy Certificate Program
  • Online Student Experience
  • Online Programs for Applied Learning
  • Barcelona Information
  • Fall Institute Housing Accommodations
  • Participating Centers
  • Registration, Tuition, and Fees
  • Agency Scholarship Application
  • General Scholarship Application
  • UPF Scholarship Application
  • Course Evaluations
  • Online Courses
  • Registration
  • General Institute Tuition Information
  • International Students
  • Directions to the Bloomberg School
  • All Courses
  • Important Guidance for ONSITE Students
  • D.C. Courses
  • Registration and Fees
  • Cancellation and Closure Policies
  • Application Procedures
  • Career Search
  • Current Activities
  • Current Trainees
  • Related Links
  • Process for Appointing Postdoctoral Fellows
  • Message from the Director
  • Program Details
  • Admissions FAQ
  • Current Residents
  • Elective Opportunities for Visiting Trainees
  • What is Occupational and Environmental Medicine?
  • Admissions Info
  • Graduates by Year
  • Compensation and Benefits
  • How to Apply
  • Academic Committee
  • Course Details and Registration
  • Tuition and Fees
  • ONLINE SOCI PROGRAM
  • Principal Faculty
  • Johns Hopkins RAPID Psychological First Aid
  • General Application
  • JHHS Application
  • Areas of Study
  • Important Dates
  • Our Faculty
  • Welcome Letter
  • Descripción los Cursos
  • Programa en Epidemiología para Gestores de Salud, Basado en Internet
  • Consultants
  • Britt Dahlberg, PhD
  • Joke Bradt, PhD, MT-BC
  • Mark R. Luborsky, PhD
  • Marsha Wittink, PhD
  • Rebekka Lee, ScD
  • Su Yeon Lee-Tauler, PhD
  • Theresa Hoeft, PhD
  • Vicki L. Plano Clark, PhD
  • Program Retreat
  • Mixed Methods Applications: Illustrations
  • Announcements
  • 2023 Call for Applications
  • Jennifer I Manuel, PhD, MSW
  • Joke Bradt, PhD
  • Josiemer Mattei, PhD, MPH
  • Justin Sanders, MD, MSc
  • Linda Charmaran, PhD
  • Nao Hagiwara, PhD
  • Nynikka R. A. Palmer, DrPH, MPH
  • Olayinka O. Shiyanbola, BPharm, PhD
  • Sarah Ronis, MD, MPH
  • Susan D. Brown, PhD
  • Tara Lagu, MD, MPH
  • Theresa Hoft, PhD
  • Wynne E. Norton, PhD
  • Yvonne Mensa-Wilmot, PhD, MPH
  • A. Susana Ramírez, PhD, MPH
  • Animesh Sabnis, MD, MSHS
  • Autumn Kieber-Emmons, MD, MPH
  • Benjamin Han, MD, MPH
  • Brooke A. Levandowski, PhD, MPA
  • Camille R. Quinn, PhD, AM, LCSW
  • Justine Wu, MD, MPH
  • Kelly Aschbrenner, PhD
  • Kim N. Danforth, ScD, MPH
  • Loreto Leiva, PhD
  • Marie Brault, PhD
  • Mary E. Cooley, PhD, RN, FAAN
  • Meganne K. Masko, PhD, MT-BC/L
  • PhuongThao D. Le, PhD, MPH
  • Rebecca Lobb, ScD, MPH
  • Allegra R. Gordon, ScD MPH
  • Anita Misra-Hebert, MD MPH FACP
  • Arden M. Morris, MD, MPH
  • Caroline Silva, PhD
  • Danielle Davidov, PhD
  • Hans Oh, PhD
  • J. Nicholas Dionne-Odom, PhD RN ACHPN
  • Jacqueline Mogle, PhD
  • Jammie Hopkins, DrPH, MS
  • Joe Glass, PhD MSW
  • Karen Whiteman, PhD MSW
  • Katie Schultz, PhD MSW
  • Rose Molina, MD
  • Uriyoán Colón-Ramos, ScD MPA
  • Andrew Riley, PhD
  • Byron J. Powell, PhD, LCSW
  • Carrie Nieman MD, MPH
  • Charles R. Rogers, PhD, MPH, MS, CHES®
  • Emily E. Haroz, PhD
  • Jennifer Tsui, Ph.D., M.P.H.
  • Jessica Magidson, PhD
  • Katherine Sanchez, PhD, LCSW
  • Kelly Doran, MD, MHS
  • Kiara Alvarez, PhD
  • LaPrincess C. Brewer, MD, MPH
  • Melissa Radey, PhD, MA, MSSW
  • Sophia L. Johnson, PharmD, MPH, PhD
  • Supriya Gupta Mohile, MD, MS
  • Virginia McKay, PhD
  • Andrew Cohen, MD, PhD
  • Angela Chen, PhD, PMHNP-BC, RN
  • Christopher Salas-Wright, PhD, MSW
  • Eliza Park MD, MS
  • Jaime M. Hughes, PhD, MPH, MSW
  • Johanne Eliacin, PhD, HSPP
  • Lingrui Liu ScD MS
  • Meaghan Kennedy, MD
  • Nicole Stadnick, PhD, MPH
  • Paula Aristizabal, MD
  • Radhika Sundararajan, MD
  • Sara Mamo, AuD, PhD
  • Tullika Garg, MD MPH FACS
  • Allison Magnuson, DO
  • Ariel Williamson PhD, DBSM
  • Benita Bamgbade, PharmD, PhD
  • Christopher Woodrell MD
  • Hung-Jui (Ray) Tan, MD, MSHPM
  • Jasmine Abrams, PhD
  • Jose Alejandro Rauh-Hain, MD
  • Karen Flórez, DrPH, MPH
  • Lavanya Vasudevan, PhD, MPH, CPH
  • Maria Garcia, MD, MPH
  • Robert Brady, PhD
  • Saria Hassan, MD
  • Scherezade Mama, DrPH
  • Yuan Lu, ScD
  • 2021 Scholars
  • Sign Up for Our Email List
  • Workforce Training
  • Cells-to-Society Courses
  • Course/Section Numbers Explained
  • Pathway Program with Goucher College
  • The George G. Graham Lecture

About the ScM in Genetic Counseling Program

The Johns Hopkins/National Institutes of Health Genetic Counseling Training Program is a joint effort of the Department of Health, Behavior and Society and the National Institutes of Health through the National Human Genome Research Institute (NHGRI) and the National Cancer Institute (NCI). The ScM in Genetic Counseling prepares graduates for a career in genetic counseling with an emphasis on clients’ psychological and educational needs. The program provides a solid foundation for conducting social and behavioral research related to genetic counseling and equips students with the necessary skills to educate health care providers, policymakers, and the public about genetics and related health and social issues.

The curriculum consists of coursework in the areas of human genetics, genetic counseling, health education, communication, ethics, public policy, and research methodology. The program also requires a minimum of 600+ contact hours of supervised clinical rotations in a variety of settings across more than 25 sites and a thesis study worthy of publication.

Accreditation: The ScM in Genetic Counseling is accredited by the Accreditation Council for Genetic Counseling, and graduates are eligible to sit for the American Board of Genetic Counseling certification examination after completion of the program and a clinical log book demonstrating significant involvement in the evaluation and counseling of patients seen in approved clinical rotation sites.

ScM in Genetic Counseling Program Highlights

Psychotherapeutic counseling training model.

with 600+ contact hours of supervised clinical rotations

Genetic Counseling Research and Scholarship

requiring completion of a publishable thesis

Emphasis on Clinical Genomics and Precision Health

with a focus on critical thinking and applications to healthcare

Transdisciplinary Learning and Practice

incorporating perspectives from public health, policy, ethics, and advocacy

What Can You Do With a Graduate Degree In Genetic Counseling?

Graduates of the JHU/NIH Genetic Counseling Training Program are qualified to work as genetic counselors in many different settings. Many will provide clinical genetic counseling services, but genetic counselors are increasingly working with genetic testing laboratories, advocacy organizations, or insurers. In addition, our graduates are well-positioned to work as genetic counselors in research settings, and almost a third of our alumni currently have a research element to their positions. For more information about genetic counseling careers, visit www.nsgc.org .

Curriculum for the ScM in Genetic Counseling

The curriculum for the master's degree consists of at least 149 credit hours of didactic coursework in the areas of genetic counseling, human genetics, public health, public policy, research methodology, bioethics and health communication and decision-making. For a list of courses and links to course descriptions, please click on the links below.

Browse an overview of the requirements for this master's program in the JHU  Academic Catalogue , explore all course offerings in the Bloomberg School  Course Directory , and find many more details in the program's Student Handbook .

Admissions Requirements

For general admissions requirements, please visit the How to Apply page. This specific program also requires:

Prior Work Experience

Prior counseling experience (paid or voluntary) working with individuals in emotional distress

Prior Coursework

College-level courses in biochemistry and genetics; psychology and statistics strongly recommended

Important Note

We regret that we will be unable to consider applications this year from individuals who are not U.S. citizens or permanent residents unless the applicant has already received a doctoral degree. Please contact us if you have questions.

Standardized Test Scores

Standardized test scores (GRE) are  optional  for this program. The admissions committee will make no assumptions if a standardized test score is omitted from an application, but will require evidence of quantitative/analytical ability through other application components such as academic transcripts and/or supplemental questions.  Applications will be reviewed holistically based on all application components.

Program Faculty Spotlight

Debra Roter

Debra Roter

Debra Roter, DrPH '77, MPH '75, is an expert on patient-provider communication and author of the widely used Roter Interaction Analysis System (RIAS).

Lori Erby

Lori H. Erby

research about genetic counselling

Leila Jamal

Leila Jamal, ScM '12, PhD '17, is a genetic counselor/bioethicist who studies ethical issues raised by the implementation of genomic medicine. Her primary appointment is at the NCI.

Funding is determined annually by an agreement with NIH. Typically, we are able to offer the following:

30% reduction in tuition at the Johns Hopkins Bloomberg School of Public Health; $10,000 scholarship in the first year; Citizens and permanent residents of the US, an NIH intramural research fellowship with amounts dictated by the following federal program: https://www.training.nih.gov/predoctoral_irta_stipend_levels

Information regarding the cost of tuition and fees can be found on the Bloomberg School's Tuition & Fees page.

Questions about the program? We're happy to help.

Application and Admissions Procedural Questions

Please direct questions about application and admissions procedures to the BSPH Admissions Office.

Email: [email protected] Phone: 410-955-3543

Program-Specific Questions

Please direct program-specific questions to [email protected] .

General Academic Questions

For general academic questions, please contact our Department's academic program administrator, L. Robin Newcomb.

Email: [email protected]

Additional Program Information

For additional information about the program, please visit:  www.genome.gov/gctp .

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
  • My Account Login
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 21 December 2022

Research priorities in psychiatric genetic counselling: how to talk to children and adolescents about genetics and psychiatric disorders

  • Jessica Mundy 1   na1 ,
  • Helena L. Davies 1   na1 ,
  • Mădălina Radu 2 ,
  • Jehannine Austin   ORCID: orcid.org/0000-0003-0338-7055 3 , 4 ,
  • Evangelos Vassos   ORCID: orcid.org/0000-0001-6363-0438 1 ,
  • Thalia C. Eley   ORCID: orcid.org/0000-0001-6458-0700 1 , 5 ,
  • Gerome Breen   ORCID: orcid.org/0000-0003-2053-1792 1 , 5 &
  • Ramona Moldovan   ORCID: orcid.org/0000-0002-6013-5076 2 , 6 , 7  

European Journal of Human Genetics volume  31 ,  pages 262–264 ( 2023 ) Cite this article

1908 Accesses

27 Altmetric

Metrics details

  • Health policy
  • Patient education
  • Public health
  • Social sciences

It is now well established that mental health disorders are heritable [ 1 ]. Genetic counselling is a process through which a trained professional helps an individual to better understand and adapt to the medical, psychological, and familial implications of genetic contributions to disease [ 2 ]. As of yet, no genetic tests to confirm a psychiatric diagnosis exist and alone they are unlikely to be sufficient for diagnosis. However, a formal genetic test is not required for the delivery of genetic counselling and for the benefits to be realised [ 3 ]. Psychiatric genetic counselling is conceptually identical to genetic counselling for other types of disorders and its efficacy for adults diagnosed with psychiatric disorders and their family members has been documented [ 4 ]. We are now entering an era in which genetic information is likely to become far more available as it is gradually integrated into healthcare. Thus, how to effectively communicate complex genetic risk information is of high research priority.

Communicating genetic risk for psychiatric disorders comes with many challenges due to their complex and multifactorial nature. Nonetheless, psychiatric genetic counselling is associated with a number of positive immediate and long-term outcomes [ 4 ]. For instance, psychiatric genetic counselling can tackle misconceptions about causes of illness, address genetic and/or environmental determinism, empower and reduce shame and/or guilt, change one’s approach to treatment, and enable more informed decision-making regarding major life decisions, such as having children [ 4 ]. Such established benefits suggest that psychiatric genetic counselling will become an important part of clinical care for psychiatric patients in the future.

Half of mental health disorders start before the age of 14 [ 5 ], with 1 in 7 young people between 10 to 19 years old experiencing mental ill health [ 6 ]. Thus, childhood or adolescence could be a particularly suitable window within which to receive psychiatric genetic counselling. This may prevent misconceptions about the causes of one’s mental illness, manage stigmatising beliefs related to personal or family history of mental health problems [ 7 ], and encourage risk-reducing behaviours [ 8 , 9 ]. Psychiatric genetic counselling could also have a positive impact on parents and caregivers, who often feel responsible for their child’s mental health and may experience feelings of guilt, shame, or a heavy burden of responsibility [ 10 ]. Such feelings may be partially rooted in a limited understanding of the contributions of genetic and environmental factors to mental disorders [ 11 ]. This can have a variety of negative behavioural consequences such as not seeking out suitable support for their child or potentially limiting the number of their future children [ 12 , 13 ].

Lack of research into the communication of genetic risk information to children and adolescents

Research exploring physical health conditions in young people highlights a strong desire to better understand the cause/s of their diagnosed disorder [ 14 , 15 ]. Given the complex nature of psychiatric disorders, genetic counsellors recognise that children and adolescents require specific tools and techniques to adequately comprehend information about the aetiology and management of their condition. For example, adapting the wording of medical terminology or using metaphors and visual aids may prove beneficial [ 16 ]. A mixed-method systematic review analysing the literature on genetic counselling for children and adolescents identified a dearth of publications on how best to communicate this type of information to children and adolescents (Radu M, Moldovan R. Genetic counseling for children and adolescents: A mixed-method systematic review. 2022. Manuscript in preparation.). The authors noted the stark contrast between the large number of guidelines and recommendations available and the relatively scarce number of empirical research studies. The picture is even less clear for psychiatric disorders. Furthermore, to this day, studies investigating the efficacy of psychiatric genetic counselling have focussed on adults only [ 17 ].

Research directions and practical recommendations

We propose the following research priorities:

What is the most effective way to communicate personal genetic risk to children and adolescents?

Young people form a distinct patient group and may need specific strategies or tools to explain the concept of inherited genetic risk for psychiatric disorders. For instance, the ‘mental health jar’ has proved to be extremely successful within adults with psychiatric disorders [ 8 ]. However, traditional communication styles used with adults may not be directly translatable to young people. Research now needs to investigate whether this tool (and others) are similarly successful in young populations. Furthermore, given that genetic information may have implications for other family members, genetic counsellors can support individuals when they are adapting to the medical, psychological, and familial implications of their condition [ 17 ].

When is the most appropriate time to discuss risk for psychiatric disorders with young people?

Young people want to learn about their physical disorders at a much younger age than when genetic counselling is routinely offered [ 14 , 15 ]. As most research has focused on physical health conditions or on adults with psychiatric conditions, the effectiveness of psychiatric genetic counselling in younger age groups is still unclear. Further, c ommunication requirements may differ depending on the specific developmental stage within childhood and adolescence, or indeed the psychiatric problems experienced. A recent qualitative study explored the opinions of ten parents of children with the genetic condition 22q11.2 deletion syndrome and highlighted that the child’s developmental age and thus ability to comprehend complex information influenced when they decided to initiate discussions about the cause of the disorder [ 18 ]. Similar types of research are required to establish the optimal age period within childhood and/or adolescence during which psychiatric genetic counselling is most effective for a range of disorders.

Who is best placed to offer this service?

Given the prevalence of mental illnesses and the relative dearth of genetic counsellors (which falls far below the recommendations by the WHO), a focused effort to train other healthcare professionals may reduce reliance on genetic counsellors. A meta-analysis found that the type of professional, i.e., genetic counsellors versus other specialists, had no significant impact on the association between psychiatric genetic counselling and positive outcomes in adults, such as reductions in anxiety or guilt and increase in empowerment or self-efficacy [ 4 ]. However, more research is needed to understand the ability of other specialists to fulfil this role effectively within child and adolescent services after adequate training. Some argue that all medical professionals who interact with psychiatric patients may be required to initiate discussions about genetic and environmental risk. As such, there is an argument for mainstreaming psychiatric genetic counselling into healthcare. This has garnered more attention over the years as genetic testing becomes more commonplace but, regardless of the setting in which it is offered, it will be important for psychiatric genetic counselling to be delivered in an evidence-based manner [ 11 ] and will require integration into their professional training.

Psychiatric genetic counselling holds great potential for benefiting children and adolescents experiencing mental health disorders, as well as their parents/caregivers. While children and adolescents want to better understand the cause of their disorder/s, little empirical research has been conducted on the communication of genetic information to this distinct patient group. Therefore, we propose that investigating how to best address psychiatric disorders and deliver genetic information to children and adolescents must become a research priority in order to support and inform evidence-based practice and guidelines.

Polderman TJC, Benyamin B, de Leeuw CA, Sullivan PF, van Bochoven A, Visscher PM, et al. Meta-analysis of the heritability of human traits based on fifty years of twin studies. Nat Genet. 2015;47:702–9.

Article   CAS   PubMed   Google Scholar  

Resta R, Biesecker BB, Bennett RL, Blum S, Hahn SE, Strecker MN, et al. A new definition of genetic counseling: National Society of Genetic Counselors’ task force report. J Genet Couns. 2006;15:77–83.

Article   PubMed   Google Scholar  

Moldovan R, McGhee KA, Coviello D, Hamang A, Inglis A, Ingvoldstad Malmgren C, et al. Psychiatric genetic counseling: A mapping exercise. Am J Med Genet B Neuropsychiatr Genet. 2019;180:523–32.

Moldovan R, Pintea S, Austin J. The efficacy of genetic counseling for psychiatric disorders: a meta-analysis. J Genet Couns. 2017;26:1341–7.

Kessler RC, Berglund P, Demler O, Jin R, Merikangas KR, Walters EE. Lifetime prevalence and age-of-onset distributions of DSM-IV disorders in the National Comorbidity Survey Replication. Arch Gen Psychiatry. 2005;62:593–602.

World Health Organisation. Adolescent mental health. World Health Organisation 2021. https://www.who.int/news-room/fact-sheets/detail/adolescent-mental-health (accessed April 14 2022).

Hinshaw SP, Stier A. Stigma as related to mental disorders. Annu Rev Clin Psychol. 2008;4:367–93.

Semaka A, Austin J. Patient perspectives on the process and outcomes of psychiatric genetic counseling: An “Empowering Encounter”. J Genet Couns. 2019;28:856–68.

PubMed   Google Scholar  

Sabatello M, Chen Y, Herrera CF, Brockhoff E, Austin J, Appelbaum PS. Teenagers and precision psychiatry: a window of opportunity. Public Health Genomics. 2021;24:14–25.

Radu M, Ciucă A, Crișan C-A, Pintea S, Predescu E, Șipos R, et al. The impact of psychiatric disorders on caregivers: An integrative predictive model of burden, stigma, and well-being. Perspect Psychiatr Care 2022. https://doi.org/10.1111/ppc.13071 .

Austin JC. Evidence-based genetic counseling for psychiatric disorders: a road map. Cold Spring Harb Perspect Med. 2020;10. https://doi.org/10.1101/cshperspect.a036608 .

Meiser B, Mitchell PB, Kasparian NA, Strong K, Simpson JM, Mireskandari S, et al. Attitudes towards childbearing, causal attributions for bipolar disorder and psychological distress: a study of families with multiple cases of bipolar disorder. Psychol Med. 2007;37:1601–11.

Austin JC, Hippman C, Honer WG. Descriptive and numeric estimation of risk for psychotic disorders among affected individuals and relatives: implications for clinical practice. Psychiatry Res. 2012;196:52–6.

Article   PubMed   PubMed Central   Google Scholar  

Pichini A, Shuman C, Sappleton K, Kaufman M, Chitayat D, Babul-Hirji R. Experience with genetic counseling: the adolescent perspective. J Genet Couns. 2016;25:583–95.

Szybowska M, Hewson S, Antle BJ, Babul-Hirji R. Assessing the informational needs of adolescents with a genetic condition: what do they want to know? J Genet Couns. 2007;16:201–10.

Duncan RE, Young M-A. Tricky teens: are they really tricky or do genetic health professionals simply require more training in adolescent health? Per Med. 2013;10:589–600.

Ryan J, Virani A, Austin JC. Ethical issues associated with genetic counseling in the context of adolescent psychiatry. Appl Transl Genom. 2015;5:23–9.

PubMed   PubMed Central   Google Scholar  

Cook CB, Slomp C, Austin J. Parents’ perspectives, experiences, and need for support when communicating with their children about the psychiatric manifestations of 22q11.2 deletion syndrome (22q11DS). J Community Genet. 2022;13:91–101.

Download references

Jessica Mundy acknowledges funding from the Lord Leverhulme charitable grant as part of a Ph.D. studentship. Helena L Davies acknowledges funding from the Economic and Social Research Council (ESRC) as part of a Ph.D. studentship. Professor Eley and Professor Breen are part-funded by a program grant from the UK Medical Research Council (MR/V012878/1) and the NIHR Biomedical Research Centre at South London and Maudsley NHS Foundation Trust. Dr. Vassos also acknowledges funding from the NIHR Biomedical Research Centre at South London and Maudsley NHS Foundation Trust. The views expressed in this commentary are those of the author(s) and not necessarily those of the NHS, the NIHR, or the Department of Health and Social Care. Professor Moldovan acknowledges funding from NHS North West Genomic Medicine Service Alliance, Health Education England and IS-BRC-1215-20007/Manchester NIHR BRC.No other authors acknowledge sources of funding.

Author information

These authors contributed equally: Jessica Mundy, Helena L. Davies.

Authors and Affiliations

Social, Genetic and Developmental Psychiatry Centre, Institute of Psychiatry, Psychology and Neuroscience, King’s College London, London, UK

Jessica Mundy, Helena L. Davies, Evangelos Vassos, Thalia C. Eley & Gerome Breen

Department of Psychology, Babeş-Bolyai University, Cluj-Napoca, Romania

Mădălina Radu & Ramona Moldovan

Departments of Psychiatry and Medical Genetics, The University of British Columbia, Vancouver, BC, Canada

Jehannine Austin

BC Mental Health and Substance Use Services Research Institute, Vancouver, BC, Canada

UK National Institute for Health Research (NIHR) Biomedical Research Centre for Mental Health, South London and Maudsley Hospital, London, UK

Thalia C. Eley & Gerome Breen

Division of Evolution and Genomic Sciences, School of Biological Science, University of Manchester, Manchester, UK

Ramona Moldovan

Manchester Centre for Genomic Medicine, St Mary’s Hospital, Manchester University Hospitals NHS Foundation Trust, Manchester Academic Health Science Centre, Manchester, UK

You can also search for this author in PubMed   Google Scholar

Contributions

Conceptualization: JM, HLD, TCE, RM. Investigation: JM, HLD, RM, Project administration: JM, HLD, RM, Supervision: EV, TCE, GB, RM Writing – original draft: JM, HLD, Writing—review & editing: JM, HLD, MR, JA, EV, TCE, GB, RM.

Corresponding author

Correspondence to Ramona Moldovan .

Ethics declarations

Competing interests.

Prof. Breen has received honoraria, research or conference grants, and consulting fees from Illumina, Otsuka, and COMPASS Pathfinder Ltd. No. other authors declare 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

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as 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 images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Mundy, J., Davies, H.L., Radu, M. et al. Research priorities in psychiatric genetic counselling: how to talk to children and adolescents about genetics and psychiatric disorders. Eur J Hum Genet 31 , 262–264 (2023). https://doi.org/10.1038/s41431-022-01253-0

Download citation

Received : 18 August 2022

Accepted : 21 November 2022

Published : 21 December 2022

Issue Date : March 2023

DOI : https://doi.org/10.1038/s41431-022-01253-0

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

Quick links

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

research about genetic counselling

Recent Blogs

Tammy LeDoux

Determine your risk of developing certain diseases, and get screening and preventive treatment.

Many health conditions have a genetic link. A gene is like an instruction manual for your body that tells your body how to function, develop and stay healthy. Some health conditions are almost completely determined by your genes, while others you may develop because of your environment.

Empowered to live a healthy life

Our genetic counselors work with you from start to finish, navigating the complexity of genetic testing by discussing genetic testing options that may be helpful for you and helping you understand what your results mean for you and your family — empowering you to live your best life.

Learn more:

What is genetic counseling, what is genetic testing, genetics faq, making an appointment, preparing for your appointment.

Genetic counseling seeks to help you understand complex genetic information and the options available for genetic testing. During the counseling appointment, a review of your personal medical history and family health history will be conducted. A family tree, called a pedigree, will be created to document your family history.

Identifying genetic variants that may indicate your risk of inheriting or developing certain illnesses or diseases allows you to make informed health decisions:

  • Lifestyle or environmental changes Some health conditions are almost completely determined by your genes, while others are more influenced by your environment. While you can't change your genes, you can control some aspects of your environment, such as diet, physical activity and tobacco use.
  • Preventive treatment Knowing your risk of developing a condition allows for preventive care tailored to you, such as starting screenings earlier than the recommended age or more frequently.
  • Family planning To define the risk of having a child with an inherited condition, you and your partner may choose to have genetic testing before having children if you have a high risk of a specific genetic disorder to determine if you are carriers for the same condition.

Based on your family history, a genetic counselor will discuss:

  • Your risk of having certain genetic changes or conditions
  • Review genetic tests to consider, insurance coverage and cost
  • What the testing may or may not tell you
  • What the chances are of passing the genetic disorder onto your children

Get the information you need to make informed decisions about your health. From the explanation of test results to providing an initial emotional and psychosocial assessment, a genetic counselor will be there with you to navigate genetic disorders.

Back to top

Genetic testing involves examining your DNA — the chemical database that carries instructions for your body's functions. Genetic testing can reveal changes, or mutations, in your genes that may cause illness or disease.

Although genetic testing can provide important information for diagnosing, treating and preventing illness, there are limitations. For example, if you're a healthy person, a positive result from genetic testing doesn't always mean you will develop a disease. On the other hand, in some situations, a negative result doesn't guarantee that you won't have a certain disorder.

Talking to your health care professional, medical geneticist or genetic counselor about what you will do with the results is an important step in the process of genetic testing.

Genetic testing is available for conditions in these specialties, as well as medications:

  • Cancer or oncology A cancer syndrome is a genetic disorder in which inherited genetic mutations in one or more genes puts the person at higher risk for developing cancer during their lifetime. Different hereditary cancer syndromes can be associated with different cancers and having a hereditary cancer syndrome does not necessarily guarantee someone will develop cancer in their lifetime. There are known syndromes associated with female and male breast cancer, ovarian cancer, uterine cancer, pancreatic cancer, prostate cancer, melanoma, colorectal cancer, colorectal polyps, gastric cancer, renal cancer, thyroid cancer, sarcoma, cancer of the brain and nervous system, and others.
  • Cardiology Many cardiac conditions can be found to have a genetic component to them. These genetic components, also called inherited genetic mutations, can predispose an individual to developing a cardiac condition. Cardiomyopathy, high cholesterol/coronary artery disease, aortic aneurysm and/or dissection, and arrhythmia all can have genetic risk factors associated with the condition.
  • Family planning Knowing your genetic risk can help determine risk for your future children to have the same condition. Genetic testing can be performed on a patient and their partner prior to or during pregnancy to see if there are hidden inherited conditions that run in their family and could be passed onto their children, such as Cystic Fibrosis and Spinal Muscular Atrophy.
  • Prenatal or pregnant Genetic testing can also be performed during pregnancy to screen for common conditions that occur randomly and are not passed on in families such as Down syndrome. Many individuals utilize prenatal genetic testing to determine risk for these genetic conditions and prepare medically and emotionally to have a child with a genetic condition. Not all prenatal testing contributes physical risk to an individual or their pregnancy.
  • Medications Pharmacogenomics determines the right drug at the right dose based on your genes. Your genes may affect your body's response to, and interaction with, some medications. While this is not a service genetic counseling addresses, it may be a tool your provider uses to select a medication for you. Learn more about pharmacogenomics .

If you have a specific family history concern you would like to address, discuss with your primary care health care professional. He or she may refer you to meet with a genetic counselor for further discussion.

Genetic counseling visits are available via:

  • Video appointment Appointments with our genetic counselors are available virtually from any location for Mayo Clinic Health System patients using the Mayo Clinic app or your online patient portal with Patient Online Services. If genetic testing is ordered after your virtual appointment, a testing kit will be mailed to you to complete in the comfort of your home and returned by mail.
  • In-person appointment Our genetic counselors offer in-person appointments in Eau Claire  and La Crosse, Wisconsin , and Mankato, Minnesota . If genetic testing is ordered after your appointment, it may be completed on campus or a testing kit may be mailed to you to complete from the comfort of your home.

Referring providers

To refer a patient, external health care professionals should call 1-855-392-8400.

Before meeting with our genetic counselor, gather as much information as possible about your family's medical history. Both your mother's and father's side of your family are relevant for you, regardless of your biological sex or whether you more closely resemble one side of the family more than the other. The more distantly related a person is, the less his or her health history is expected to affect you.

It still may help to gather information from as many relatives as you can, including:

  • First-degree relatives — Parents, full siblings, children
  • Second-degree relatives — Half-siblings, grandparents, aunts and uncles, nieces and nephews, grandchildren
  • Third-degree relatives — First cousins

You can print a family health history form or use your own method of recording the information you gather.

It may be challenging to capture a complete health history due to adoption, estrangement, or lack of details known or willing to be shared within the family. It's OK to simply record the information that you can gather and know that it can be updated if more information becomes available in the future.

You may receive a phone call one week prior to your appointment to review your family history or be asked to complete a family history questionnaire at the time of your visit.

What to ask about

In general, you should try to gather information about whether relatives are living, their current age or age at which they passed away, and their medical history.

Relevant details about health history include:

  • Any major health issues and age of diagnosis
  • Chronic conditions
  • Conditions that required them to see a specialist
  • Conditions that required them to take medication
  • Conditions they were born with or developed unexpectedly at a young age
  • Known familial mutation in a family member
  • Country of origin

It's helpful to know about your relatives' environmental risk factors, such as history of tobacco use, alcohol abuse, drug abuse, obesity or exposure to radiation or chemicals. Also, if you have relatives who have visited with a genetic counselor before, they may have already had a visual representation of the family health history — like a family tree — created, which also is called a "pedigree." It may help to ask if they'd be willing to share their pedigree, as this can save you a lot of time and work when gathering family history details.

Does insurance cover genetic counseling and testing?

Prior authorizations are not needed for the genetic counseling appointment. However, we encourage you to confirm that Mayo Clinic Health System is considered "in network" by your insurance provider.

Genetic testing is billed separately from the genetic counseling appointment. During the genetic counseling appointment, if you proceed with genetic testing, the genetic counselor will talk about genetic testing costs, options, insurance coverage and prior authorizations.

For more information about insurance coverage prior to a genetic counseling appointment, contact the financial navigator at 715-838-1747 .

Are there risks to genetic tests?

Generally, genetic tests have little physical risk. Blood and saliva (cheek swab) tests have almost no risk. There are two main types of prenatal testing: screening and diagnostic tests. Screening tests during pregnancy include blood tests, a specific type of ultrasound and prenatal cell-free DNA screening that are no risk to the pregnancy. If results of these screening tests indicate an increased risk for a genetic disorder, diagnostic tests may be considered, such as amniocentesis and chorionic villus sampling which carry a slight risk of miscarriage (pregnancy loss).

Genetic testing can have emotional, social and financial risks as well. Discuss all risks and benefits of genetic testing with your health care professional, medical geneticist or genetic counselor before you have a genetic test.

Can genetic information be used against me for health insurance or employment?

No. The Genetic Information Nondiscrimination Act of 2008 is a federal law that protects people from genetic discrimination, specifically in the setting of health insurance and employment. This means your genetic information cannot be used against you when it comes to health insurance and employment.

Genetic information includes family health history genetic testing results; use of genetic services, such as genetic counseling; and participating in genetic research. Under the act, your family health history and genetic testing results cannot be considered a pre-existing condition.

It's important to know that there are some exceptions to the act, and it does not have protections in place for life insurance, long-term care insurance or disability insurance.

These locations offer genetic counseling:

Providers view more.

Rebecca Jirik, CGC

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

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • J Community Genet
  • v.10(1); 2019 Jan

Logo of commgen

Ethics in genetic counselling

Angus j. clarke.

1 Institute of Medical Genetics, Division of Cancer & Genetics, School of Medicine, Cardiff University, Heath Park, Cardiff, Wales CF14 4XN UK

Carina Wallgren-Pettersson

2 The Folkhaelsan Department of Medical Genetics, Topeliusgatan, 20 00250 Helsinki, Finland

3 The Folkhaelsan Institute of Genetics and the Department of Medical and Clinical Genetics, University of Helsinki, Helsinki, Finland

Difficult ethical issues arise for patients and professionals in medical genetics, and often relate to the patient’s family or their social context. Tackling these issues requires sensitivity to nuances of communication and a commitment to clarity and consistency. It also benefits from an awareness of different approaches to ethical theory. Many of the ethical problems encountered in genetics relate to tensions between the wishes or interests of different people, sometimes even people who do not (yet) exist or exist as embryos, either in an established pregnancy or in vitro. Concern for the long-term welfare of a child or young person, or possible future children, or for other members of the family, may lead to tensions felt by the patient (client) in genetic counselling. Differences in perspective may also arise between the patient and professional when the latter recommends disclosure of information to relatives and the patient finds that too difficult, or when the professional considers the genetic testing of a child, sought by parents, to be inappropriate. The expectations of a patient’s community may also lead to the differences in perspective between patient and counsellor. Recent developments of genetic technology permit genome-wide investigations. These have generated additional and more complex data that amplify and exacerbate some pre-existing ethical problems, including those presented by incidental (additional sought and secondary) findings and the recognition of variants currently of uncertain significance, so that reports of genomic investigations may often be provisional rather than definitive. Experience is being gained with these problems but substantial challenges are likely to persist in the long term.

Introduction

This paper outlines the issues and the material discussed in a module on “ethical issues in genetic counselling” that is a key feature of the Cardiff University MSc course in Genetic Counselling. It cannot be complete and exhaustive and focuses on genetic counselling practice in UK and Europe, but we hope it may be of interest also to those who work elsewhere.

Reflecting on the past

Ethics is important in the practice of every area of medicine, but this is nowhere more true than in attending to the genetic aspects of disease. Why is this so?

In part, this arises from the history of genetics. Public discussion about the inheritance of disease and non-disease traits in humans began in the nineteenth century England, led by the polymath Francis Galton. From its beginning, it was associated with attempts to improve society through selective breeding, as if mankind was another farmyard animal with some strains worthy of encouragement and others not. The discoveries of Mendel in his investigation of plants had not then been widely recognised and Galton approached inheritance predominantly from a population perspective rather than the more focused clinical perspective of the individual. Galton’s approach led to attempts to improve the genetic health of the population, which he termed ‘eugenics’. This population perspective was taken up by philanthropists and politicians and led to the ideas of positive eugenics (encouraging some to breed) and negative eugenics (discouraging others). Such ideas were seen as eminently respectable across Europe and North America at the turn of the 19th into the twentieth century. The forced sterilisation of those deemed unfit to reproduce followed from this logic in many countries, including Sweden and USA. In Nazi Germany, this approach was developed further into a policy of killing those patients considered unfit to live. The medical profession and associated scientists, especially anthropologists, often colluded with these activities in a devastating Faustian pact.

These events occurred in societies that viewed populations as both inherently distinct and of unequal value. Ideas about a natural heirarchy of racial types coalesced with ideas about the incipent degeneration of any population if the less worthy types were not stopped from breeding to excess. The effort to maintain the purity and quality of the race—racial hygiene—became especially closely entangled in Nazi ideology with their parallel beliefs about racial superiority, leading to the murder of much larger numbers of those deemed inferior because of their assigned race. However, what is common to the two activities is the assessment of the worth of an individual from a population perspective: what good to the nation or the state is this individual with a malformation or a cognitive impairment? Of what use is this man or woman from an ‘inferior race’?

One use that could be made of inferior types in Nazi Germany was a role in human experimentation: the name of Mengele remains infamous to this day. Some of his medical colleagues gained high repute from their pathological studies conducted post mortem on their patients, whom they had selected to be killed for this purpose (Müller-Hill 1988 ; Harper 1996 ).

Apartheid can be seen as another application of this ideology, with the mixing of ‘races’ seen as a contamination that would lead to a degeneration of the true stock. The language of genetics could just as easily have promoted such mixing of populations through citing the benefits of ‘hybrid vigour’ but that perspective is not appealing to the eugenicists.

The central lesson for us from this experience must be that any benefits to a population or nation from clinical genetic services should arise only as a secondary consequence of the primary benefits to patients and their families. Benefits to the population should never be used as the principal goal of clinical genetic services, which should always give priority to the interests and wishes of the individuals involved. Similarly, any experimentation or research involving humans must be conducted with the consent of the research participants, who must be fully aware of any potential risks.

Since the end of World War II, the medical profession has largely accepted these lessons, drawn from reflection on the abuse of patients under the Nazi regime, and they are applied far beyond the scope of genetics. Thus, the requirement of informed consent for research is a central principle of human rights under the terms of the World Medical Association’s Declaration of Helsinki in 1964 (updated most recently in 2017). We are reminded of the need for this when we consider some of the flagrant counter examples, such as the Tuskegee syphilis study, which abused research subjects for a decade after ‘Helsinki’. Despite these grounds for caution, enthusiasts for population screening programmes have sometimes strayed close to the eugenic motivation of the Nazi race hygienists, justifying their interventions by the reduced birth rate of children born with malformation or cognitive impairment and/or the financial savings to the state. Indeed, it is sometimes unclear whether the motivation of enthusiasts for antenatal screening is the ‘health’ of the population, the promotion of the informed choices of parents, concern about the financial pressures on the state or (sometimes) their own personal financial gain through their role in providing screening tests. Conflicts of interest abound, and mixed motives may apply. It is to be noted that according to the 2017 update of the Declaration, the Physician explicitly pledges not to use her medical knowledge to violate human rights and civil liberties, even under threat.

While few western governments would now espouse an explicitly eugenic approach to genetic population health, the underlying motivation for some antenatal screening programmes may be very ‘traditional’ in this respect, despite a veneer of politically correct rhetoric. In a time of financial austerity, and with fading confidence in the equitable, state-led provision of health care, social care and special education for those with physical and/or cognitive impairment, the choices being made by families in antenatal screening are not made in a vacuum. Their decisions may be their own, but they may be tightly constrained by circumstance and will not always reflect their own wishes and values.

Reason, emotion and experience

Another reason for those who work in clinical genetics to consider the ethics of their profession is that this topic area engages us all, both patients and professionals, in a very deep and personal sense. We are engaged in our heads and in our hearts, both rationally/cognitively and emotionally, and the two do not always mesh well together. It is possible to make a decision objectively, in a cognitively ‘detached’ fashion, and then find that it conflicts with one’s feelings. A logical process of reasoning may lead to a decision that one finds unacceptable or even repugnant. Our task is to help our patients understand their position biologically but then make personal decisions that they can live with in the long term. The decisions must be theirs, but we can help them to consider the various factors that may be relevant, including factors they had not thought of. The potential for a gulf between a detached and objective decision and one’s feelings will often arise in decisions about the sharing of information within the family, about predictive genetic testing where there are no available treatments, and about prenatal genetic testing where the only medical intervention available would be a termination of pregnancy.

The assessment of the value of our work is much more complex and nuanced than in many other areas of medicine. Whereas there may be simple and objective measures of outcome for the success of an orthopaedic procedure or the treatment of organ failure, our criteria for the success of a consultation are set by the patient and, therefore, will vary from one consultation to the next, even for the same condition. We have to ask our patients what they want from a consultation and then do our best to achieve it. It is so patient-led that measures of outcome, that would be valid across consultations and conditions and between patients, are difficult to develop. The most satisfactory measure of outcome is one based on the concept of patient empowerment (McAllister et al. 2011 ).

In practice, we have to ask each patient about their particular concerns and decide with them what we can do to support them. The core activity of clinical genetics and genetic counselling is therefore communication, most especially listening to our patients. However, we may then have to help our patients reassess their situation and revise their decisions in the light of an enhanced understanding of ‘the facts’ and also of their social and family context. So, we need to listen, to assess the biological facts as far as we are able, to provide the relevant ‘medical’ information to the patient and family in terms that can be digested and applied by them, and then help them adjust to their situation, so that they make decisions that are true to their nature while grounded in as accurate an understanding of the biological facts as possible.

Our support may involve helping a patient come to a decision in the light of both ‘the facts of the matter’ and of an awareness of their own likely response to the situation, which will combine cognitive, rational elements with sometimes powerful emotions. The patient has to integrate the facts with their feelings.

The term ‘non-directiveness’ is often used to convey the ethos of the profession, but this can be misleading: it may suggest that we merely provide information and then leave decisions to the patient. If true, that would, as we discuss further below, be a superficial approach to genetic counselling, akin to the abandonment of our patients when they most need our support. We would propose an active and engaged form of non-directiveness, in which it may be our role to challenge our patients, asking them to consider important factors that they may have chosen to avoid. We will ask them to consider a number of ‘what if ….?’ scenarios, so that they do not undertake a genetic test without considering the full range of options in front of them (including a decision not to have the test) and the full range of potential outcomes of testing (including an unclear result or an important but unanticipated finding, incidental to the reason for having the test).

A framework for ethics in the clinical genetics consultation

The first element of ethics is to recognise a problem as being ethical, i.e. that there is a question about what one ought to do. This may appear to be obvious but there are several ways in which the obvious can be obscured. First, it may appear that the ethical aspect of a situation—what ought to happen—applies to someone else, such as the patient or a member of their family or perhaps society’s arrangements for funding health care. While the political context may on occasion be relevant to the problems arising within the clinic, especially if resources are limited, and one may feel obligations as a citizen to address such questions, this does not resolve the ethical issues that still face us, day to day, in the meantime.

Secondly, problematic aspects of genetic consultations may be framed as ethical (for the patient or for the professional) but may also be open to framing in other ways, perhaps as an issue of counselling practice or as a question of the social consequences of a patient’s decision for others in their family, rather than a question of professional ethics. It may be possible to frame the same problem in any or all of these ways.

Having recognised the ‘ought’ element in the setting, especially as it applies to oneself (the professional), one then needs a framework through which to approach it. We need to consider who is involved and what their interests are (or might be). This could entail considering the interests of an unborn child or of future people (not yet conceived), whose interests could not be considered in a court of law but who must be considered from the broader, more inclusive perspective of ethics. Listing those involved and their potential interests can be helpful in clarifying what is at stake and for whom.

The third element of this framework is to consider potential solutions, noting whose interests they take fully and fairly into account. Then, the final element is to formulate a coherent argument about which of these potential solutions to work towards (i.e. making a decision about what to do).

This fourfold framework may seem rather insubstantial, but it is a helpful approach in addressing even the most complex problems. It is not a theory in its own right, however, and it will need to be supplemented by other insights drawn from logic and moral philosophy. In describing the ethical aspects of a problem and in formulating an argument, it is essential that any underlying assumptions are explicit and that the steps in the process of reasoning are clear and valid (Flew 1975 ; Holm 2004 ): we will not go far into this here, but it is important to consider the consistency of an argument as it comes to be applied to similar but non-identical instances, and to approach any explicit discussion of ethics with patients with great care (Kaldjian 2013 ) as such discussions could easily become manipulative. A broad notion of the other ‘interested parties’ whose interests are to be considered must be used, so that the opportunity costs of a course of action are included in the assessment, along with wider issues of social justice. And the steps that may be needed to resolve a difference of opinion between individuals or groups—the procedures involved—may need to be made explicit, as in Habermas’s discourse ethics. Such an approach places emphasis on the value of following a process or procedure impartially and with consistency.

An appreciation of the major schools of moral philosophy is very helpful (Blackburn 2001 ). Those who examine questions of ethics from a specific philosophical or theological position may adopt different approaches to thinking through these questions, so that some familiarity with important schools of thought may be useful.

Consequentialism assesses a course of action by weighing the outcomes it leads to, the best known consequentialist approach being utilitarianism. Deontology draws on rules, including ethical principles, rights and obligations, to determine the right course of action. Virtue-based ethics is grounded in ideas of human nature and the ends that we should pursue, individually and collectively within society. If we do not maintain this broader awareness of ethical theory, it becomes all too easy for medical ethics to reduce to the mechanical application, as if by rote, of the well-known Four Principles set out by Beauchamp and Childress (autonomy, beneficence, non-maleficence and justice) (Beauchamp and Childress 2013 ).

In considering eugenics, as above, one could say that Galton’s eugenics is utilitarian in its attempt to maximise the overall welfare (or happiness) of society. Objections to eugenics will often be Kantian, as Kant was famous for his categorical imperative, one formulation of which specifies that a person should never use other people as mere means to her own chosen end(s).

What is needed to address the ethical issues is a full engagement of the professional as a person on every level: the rational and intellectual, indeed, but also the interpersonal awareness needed to understand the perspectives of others and to recognise the subtleties of meaning that may be expressed obliquely in communication, and the creative imagination to consider likely emotional responses in different scenarios and to generate a range of possible solutions. A good place to start practising these skills can be the consideration of ‘a case’ from the perspective of those most closely involved (e.g. Thiele 2010 ; Wilkinson 2010 ).

In the rest of this chapter, we turn to consider the major, recurrent issues that arise in genetic counselling practice and, more generally, in medicine as it deals with genetic disorders. A fuller account of these areas is given by Parker ( 2012 ) and a survey of the issues emerging as genomic technologies enter clinical practice is given in Clarke ( 2014 ).

Family matters: secrecy, disclosure, communication and testing

A recurrent question in genetic counselling practice is, ‘When—and how—should we help, encourage and perhaps persuade patients to share their personal medical and genetic information with relatives and those close to them?’ Important medical information about one person may be relevant to their family members for many reasons. If one person in a family is found to be at risk of developing a cancer or a serious cardiac problem, for example, then medical surveillance may be helpful to reduce morbidity and/or mortality in others as well as in them. A relative may also want to know if they are at risk of developing a late-onset neurodegenerative condition, such as Huntington’s disease, or if their children may be at risk of a serious disease or a genetic condition that can cause a disease, a cognitive or sensory impairment or a malformation that has already affected one or more members of the family.

Of course, such information, when it may be relevant to family members and when this is explained to the patient, will usually be passed to relatives. This is especially true if the unaffected but at-risk relatives can take action to avoid at least some of the likely problems. They may be able to have screening for malignancies of the breast or bowel, for example, or for a disorder of cardiac muscle or rhythm, or they may watch for a potentially silent complication such as hypertension at an unusually early age, as in polycystic kidney disease or neurofibromatosis type 1. Awareness of such problems can substantially improve the person’s likely prognosis.

Passing information to relatives can be experienced as more difficult if there are no interventions known to improve the outlook for those who have inherited the condition; this applies (at least for now) in families affected by Huntington disease (HD) and other neurodegenerative disorders. Telling a relative that they may be at risk of such a condition but that there is no treatment, so that ‘nothing can be done about it’, may be a thankless task, with the relatives possibly becoming angry, frightened and resentful. The family informant may therefore decide to wait ‘until the time is right’ before passing on the unwelcome information. But there may never be a ‘right’ time. Family gatherings may provide an opportunity for such discussions, but it may feel inappropriate to raise such questions at a wedding, or a funeral, or at Christmas. The disclosure of genetic information by an at-risk adult to their partner, before marriage or reproduction, can also be a very difficult decision (Keenan et al. 2013 ).

Conversely, when there is a helpful intervention, the motivation to pass on information is stronger. Families are often willing to let health services make contact directly with their relatives when the medical benefit is clear, as in familial hypercholesterolaemia (Newson and Humphries 2005 ). On the other hand, when such a condition is not seen as very strongly inherited but rather as a routine question for anyone of managing their own cholesterol level, the drive to pass information to relatives may be less (Weiner 2011 ).

There have been numerous studies, often conducted by social scientists, of communication about difficult genetic information within families. Such studies can be referred to as research in ‘empirical ethics’ as the findings help ethicists to remain grounded in their discussions in the concrete and often messy reality of everyday life. Important and helpful studies of this type are presented by Forrest et al. ( 2003 ) and Featherstone et al. ( 2006 ). They discuss the barriers that can obstruct the passing of information, such as geographical distance, family estrangement, concern about the emotional readiness of a relative to be given unwelcome information and concern about the possible emotional response of the individual, who may become distressed or angry. Professionals can draw on this work to help their patients plan how best to disclose information to relatives. These studies remind professionals of the difference between ‘giving’ information and ‘receiving’ it: one member of a family may consider that they have passed on the information but the ‘recipient’ may never have ‘received’ it. Important information should be given in a very clear and explicit fashion, and it may be helpful for professionals to back this up with written information such as letters ‘To Whom It May Concern’ that can be passed by patients to their relatives.

The studies of both Forrest and Featherstone explore the dynamics of disclosure within families, showing how they are shaped by many factors including gender roles, with women often assuming the management of information within a family, even when the relevant inheritance relates to their partner’s family (d’Agincourt-Canning 2001 ). It is difficult to address these factors realistically without colluding with the (western) stereotypes of the silent male and the busybody female. However, what is crucial for the clinician is to make no assumptions about how the particular individuals in a family will be inclined to behave. The importance of an accurate understanding of inheritance cannot be over-emphasised, especially when gender roles and the biological facts combine to downplay the involvement of males, so that the family may fail to appreciate that fathers can transmit a risk of breast cancer to their daughters (Hallowell et al. 2006 ).

Even testing for carrier status in a family affected by an autosomal recessive disorder may have personal repercussions for family members, both the affected individual and the unaffected but possibly carrier siblings. Testing for carrier status puts at risk the sense of family unity that is triggered by the disease in question. A choice to be tested may suggest disloyalty to the affected family member(s), as by implication it suggests that the patient’s brother or sister would not want to have a child like their affected sibling, while finding that one is not a carrier may lead the healthy sibling to feel isolated from the family, as no longer involved with the disease in the way that the others are (Fanos and Johnson 1995a , b ).

The feelings raised by carrier testing in the context of sex-linked disorders and chromosome rearrangements may be rather stronger for two reasons: that the ‘carrier’ may sometimes be affected, at least to some extent, and that the condition is transmitted to the child solely by them, although the decision to have children will usually be a shared decision with their partner. In terms of the weight or burden of being a carrier, carrier status for these disorders may be seen as intermediate between the situation in autosomal dominant and autosomal recessive disorders.

Genetics health professionals usually handle difficult family situations in practice by maintaining links with the patient and attempting to persuade them to disclose the relevant information to their relatives. It is exceptional for professionals to force disclosure of information about one person to another against their wishes (Clarke et al. 2005 ). There are important procedures to follow before a practitioner should implement such a disclosure, differing between jurisdictions. In UK, the General Medical Council has issued guidance on this. Recent changes to the law in France have emphasised the role of the family member in passing on information but this may then leave an obligation on the professional to ensure that this has happened.

A particular issue faced occasionally by health professionals is that of misattributed paternity that comes to light through genetic testing. While the ‘abstract’ ethical arguments weigh heavily in favour of disclosing this to the ‘social father’ as well as to the mother in a couple who have come for testing (Lucassen and Parker 2001 ), in practice professionals usually try to avoid precipitating confrontation and conflict within a couple. One lesson from such experiences is to make clear in advance to those involved when a genetic test is being discussed that might reveal misattributed paternity. This issue can also arise when the daughter of a man with a sex-linked disorder requests confirmation of her obligate carrier status.

In addition to gender, ‘culture’ may be seen as an important factor in family communication within some communities. If consanguinity is customary within a community, and if carrier status for a genetic disorder may attract stigma and lead to difficulty in arranging marriage for oneself or one’s kin, a family may wish to conceal their genetic condition from everyone else, and especially from other members of the family (Shaw and Hurst 2009 ).

Difficult situations can also arise in the reverse situation, when a relative feels obliged to undergo genetic testing so as to pass on the information obtained to others in the family, even when s/he would prefer not to be tested. This can arise in the context of familial breast cancer, for example, when testing the BRCA1 and BRCA2 genes in an affected individual might allow the family’s mutation to be identified. This may be very helpful for the at-risk relatives but the affected person, who already has a cancer, may not wish to face the prospect of the higher risk of a second breast cancer or of developing ovarian cancer (Hallowell et al. 2003 ; Foster et al. 2004 ). It may be difficult to determine whether a competent patient is ‘taking the interests of relatives into account’ in an appropriate way, as an expression of relational autonomy, or whether they are being subjected to ‘undue pressure’ by family members. The elucidation of the ethical considerations in this genetics context are interestingly different from the issues raised by family members who seek to ‘persuade’ a patient to make specific decisions about their own healthcare, when the interests of family members may differ (Ho 2008 ). Relational autonomy is of broad applicability to many genetic counselling contexts because it is a perspective from which all our lives are seen as socially embedded, and within which family relationships are seen as foundational to the rest. A strong version of relational autonomy also emphasises that our very ability to form human relationships arises out of our social origins and the care we all receive from infancy onwards (Scully 2008 , pp. 160–163).

Other difficulties are faced by those patients in whom a sequence variant of uncertain significance (a VUS) has been found. How helpful will that be to relatives, even if their involvement might clarify the interpretation of the variant? (Vos et al. 2011 ).

In HD families, too, this sense of being pressed or even coerced by family pressure to be tested can arise when a young adult at 1 in 4 risk of the disease wishes to clarify their status while their at-risk parent prefers not to do so. The young adult may wish to make reproductive decisions in the light of their risk status while their parent, closer to the likely age of onset, may prefer to avoid confronting the possibility that they may soon develop the condition. Testing the young adult may reveal that their (so far unaffected) parent carries the HD gene expansion; it can be extremely difficult for such information to be kept secret within a family.

In such situations, we would encourage family members to meet to discuss the issues before any of the individuals are tested. Attending a genetic consultation together can sometimes be helpful in clarifying what is to be gained or lost by any particular course of action.

A thoughtful approach to weighing up the competing issues around personal genetic information in these and other contexts is set out in Inside Information (Human Genetics Commission 2002 ).

Predictive genetic testing consent, competence and (non)directiveness

As discussed above, a predictive genetic test may be helpful in a clinical, medical sense because it allows optimal management of a patient at risk of an inherited disorder. When there is no medical intervention to recommend, however, the benefits of testing may be less clear so that considerations of the likely personal impact of test results on the individual and those close to them may be the key factor in whether to take the test. There may be a strong tension between wanting to know (that the result is favourable) and the fear of finding out (an unfavourable result). In those who know they are at risk of HD, fewer than 20% have a predictive test to see if they will become affected as, for many, the desire to resolve the uncertainty is outweighed by the wish to retain hope (that one will not be affected) (Tassicker et al. 2009 ; Baig et al. 2016 ).

In contrast to predictive testing for HD and most other neurodegenerative disorders, testing for many other conditions, such as the familial cancers and some of the inherited cardiac conditions, has the potential to confer some potential medical benefit. Once that becomes true for HD, the counselling in such contexts will change its nature completely and become much more like counselling for the BRCA genes or Lynch syndrome. Problems can arise in relation to predictive testing for cardiac- and cancer-related disorders when there is a mismatch between, on the one hand, how the preventive or therapeutic possibilities are presented by clinicians and understood by families in advance of testing and how, on the other hand, they are delivered by health services and experienced by families afterwards.

Constraints of space prevent a full exploration of predictive testing across the many different settings in which it occurs. Here we will focus on the context of testing for HD but we will mention just two other disorders where experience with predictive testing demonstrates the dificulties that can arise if the testing is approached as a straightforward, medically ‘obvious’ procedure, a matter for purely ‘clinical’ judgement, without attention to the counselling aspects of the decision. First, in the setting of hypertrophic cardiomyopathy (HCM), parents can arrange predictive genetic testing for their children and then regret their decision as the subsequent impact upon their children’s lives becomes apparent. Ethnographic studies that track families over some years are difficult to resource but can be very valuable (Geelen et al. 2011 ). In the very different context of predictive genetic testing for the Li-Fraumeni syndrome, a scheme for tumour surveillance has been proposed (Villani et al. 2011 ) but proof that it is of value at any age—and especially in childhood—is lacking. What stance should clinicians adopt towards predictive testing of young children? There is no single, clear answer (see discussion in British Society for Human Genetics 2010 , page 21).

Returning to the non-therapeutic context of HD, perhaps the main question for the clinician is how firmly to challenge the patient’s request for testing. One does not want to be unhelpful and obstructive to those who wish to be tested but it is important to ensure that an irrevocable step is not taken before it has been considered carefully and from multiple perspectives. Predictive testing in the absence of any useful medical intervention—‘for information only’—is therefore one of the paradigmatic settings in which the ‘counselling’ aspects of genetic counselling come to the fore, and we therefore devote a considerable portion of this article to testing in this context, especially in relation to Huntington’s disease. HD was historically the first disorder for which predictive genetic testing became widespread and much experience has been gained in this setting. Testing for HD is also much commoner than for the other autosomal dominant Mendelian neurodegenerative disorders.

In this context of counselling for predictive testing without clear medical benefit, there is a blurring of the distinctions between factors to be considered by the professionals. There are clinical factors, issues of counselling and communication, and ethical considerations. These three categories merge and become very difficult to keep distinct. In this section, therefore, we discuss the clinical process in some detail.

Here, we will consider three more focused questions:

  • (i) How to respond to those who request testing but are unwilling to engage in the counselling?
  • (ii) How to respond to requests for testing made by adolescents and young people (ages 16–24). Can it be appropriately paternalistic to seek to delay the making of decisions by at least some patients?
  • (iii) How can we challenge patients—gently, with consideration and respect—to help them arrive at their decision in a more integrated or authentic manner?

We wish to influence not the decisions made by our patients but the way in which they make their decisions, so that we can all be more confident that taking the test—if that is what they choose to do—will on balance be helpful, whatever the result.

Talking about testing—with Huntington’s disease as a model condition

Some individuals at risk of HD come to clinic convinced that they should have testing at once, without any discussion, and that they will be able to deal with whatever results emerge. After all, it is their ‘right’.

When it is explained that there is no medical benefit from testing and that we have a professional obligation to ensure that they have given valid consent before being tested, then most of these individuals agree to go through with the counselling process as recommended by international guidelines (MacLeod et al. 2013 ). Even if they feel they are humouring or tolerating us professionals, they will usually be willing to oblige. If they are not willing to do this, we sometimes fear that that their hold on calm consideration is fragile and likely to dissolve. Taking short cuts because of pressure to test rapidly is usually not helpful, although (very occasionally) it may be appropriate. It may, for example, be appropriate to accelerate the predictive testing process in a pregnancy. Even then, however, caution is important. Our experience leads us to believe that rushed testing in a pregnancy can have a grave and continuing impact on the individuals involved, including the next generation.

While the best practice guidelines cited above draw on evidence and experience, it is difficult to claim that they are fully justified on the evidence. Indeed, it would be difficult to know what evidence could be provided to confirm that a particular tradition of clinical practice was the optimum approach. Longitudinal studies could demonstrate that a certain proportion of the patients were adequately satisfied (especially those given a favourable result), but some patients would inevitably drop out from the study (especially those given an unfavourable result). Furthermore, it would be difficult to compare the outcomes of one clinical approach with those of another, if they differed substantially, as many clinicians would be unwilling to follow the full range of available policies. We, for instance, would not be willing to test those who request a predictive test without any counselling: we feel a responsibility to be as sure as we can be that everyone tested understands the range of possible test results and has had a real opportunity to reflect on the consequences, for them and others, for each result. In addition, evidence on some points could only be gathered by performing tests that we would consider unethical, and then following up those involved for several decades: as with the predictive testing of young children for HD. This topic will be explored further in the ‘ Genetic testing of (young) children ’ section.

In what follows, therefore, we present a composite account of our preferred clinical approach (even when it is supported by our experience rather than a body of research evidence) along with a discussion of some of the relevant research.

When a patient finds it emotionally difficult to engage with the discussion, it can be very helpful to reassure them that the decision is theirs and that we will carry out the test if they persist in requesting it. This can help overcome the sense that they must always present a strong, competent face to us in case our suspicion of their ambivalence or weakness means that we will deny them the test. In fact, of course, we feel much readier to test someone who can honestly discuss their ambivalence than someone who cannot admit such feelings even to her- or him-self, but an honest reassurance that we will really give away control completely can remove an important block that prevents some patients from engaging with the counselling. This reassurance, of course, should not be given if we in fact have reservations, as when we sense the features of depression or another mental illness.

An imbalance of power is an inevitable part of a medical consultation—if the professionals did not have greater knowledge and skill, then there would be no point in people seeking our services—but the imbalance can be destructive and, in these circumstances, we should work to minimise it. At the same time, we can acknowledge that the patient will be expert in their own situation and we need to work with them in partnership to achieve the best outcome possible.

When a patient appears unable or reluctant to engage with the counselling around genetic testing, the question may arise as to their ability to give consent to testing. In addition to having the information needed to make a decision and to not being subject to coercion (either blatant coercion or more subtle emotional manipulation), they should have the cognitive capacity to consent. For the professionals to have confidence that a patient has the capacity to consent under English law, professionals may need to weigh up this question: ‘Does this patient have a disturbance or impairment of the mind or brain and, if so, can this patient understand and retain and weigh up and communicate the information needed to make this decision?’ If the answer is ‘No’, then a decision about testing may need to be taken after discussion at a ‘best interests’ meeting. If there are real doubts about the patient’s capacity, or about their best interests, then a process of assessment may be required, potentially involving the courts or a legally appointed advocate.

When a competent patient argues that s/he should have testing without any discussion, and cannot be persuaded that the professionals have the responsibility to comply with standards of practice, an impasse may be reached. This is unusual but does—infrequently—occur. In such circumstances, the professional’s responsibility to provide good quality care may take precedence over the patient’s wishes. This can be framed as the professional over-riding the patient’s autonomy, but is perfectly legitimate in its own terms (Lantos et al. 2011 ). However, there is a separate discussion to be had about whether such a patient’s stated wishes do express their autonomous decision. From a Kantian perspective, which sets a high standard as to what counts as autonomy, this could be contested if the patient has failed to seek full information or has not given due weight to relevant and important considerations.

Our clinical approach follows that recommended in the consensus policy (MacLeod et al. 2013 ). We almost always have at least two meetings with the at-risk patient, sometimes more, with an interval of a month or so between each meeting. A genetic counsellor usually has a preliminary meeting to gather background information about the at-risk patient and their family, to seek confirmation of the precise diagnosis in an affected relative, to identify the wishes and expectations of the patient and to explain to them how the clinic functions. In the next consultation, we would usually discuss how they became aware of their risk of HD and how long ago. If this is recent, within the last year, that may be a reason to proceed more slowly with the testing. We would wish to know what they have learned about HD, through experience or in other ways, and perhaps provide additional information about the variability in age of onset and in the pattern of disease. Crucially, we will ask what it is that has triggered the request for testing now rather than last year or next year.

We would ask the patient how they feel they would cope with an adverse result, although of course no-one can be sure about this, and whether there is anything they or we could do in advance to help prepare for a period of real distress. It may be helpful to consider how they have coped with distress in the past, whether they rely on alcohol or ‘street drugs’, and whether they have a history of psychiatric disease or psychotropic medication. Such factors need not block access to testing but do help them to reflect on their request and whether it may be better to defer testing until they feel more resilient, their circumstances are more stable or they are better supported.

Other topics to discuss in advance of testing include who knows about their risk of HD (in the family, among their friends, their employer, etc.) and who knows that they are coming to clinic to discuss testing. To whom would they pass on news of their result? If they have children, are the children aware of their own risk? If the children are adolescent or adult, have they discussed the decision to be tested with them? If not, we would usually encourage them to do so, especially if a major reason for testing is so that they can tell their children. If an at-risk adult defers any discussion with their at-risk children (whether still minors or already adult) until their own risk is known, there is a trade-off between difficult scenarios. Either they must discuss the question of HD with their children when the children are still at 1 in 4 risk (when some of these difficult discussions will prove to have been unnecessary) or they must have a still more difficult discussion when their own decision to have predictive testing may have placed their children at 1 in 2 risk before they (the children) have any knowledge of their potential risk. We would generally suggest that the discussion is to be preferred earlier, as we have often seen the latter approach lead to resentment in the children at having been kept in the dark for too long. Breaking the family secret about HD at that point—once the child has been put at 1 in 2 risk—may also prove too difficult for some parents, who are still adjusting to their own positive result. This then leads to further secrecy within the family and a greater potential for emotionally destructive disclosures in the future.

This question of an at-risk parent discussing with their children—or other important family members—their own decision in advance of the testing is one of the points about which gentle challenging may be appropriate. Indeed, the impact of the test on the family as a system is a useful focus for discussion (Sobel and Cowan 2000 ). The question of ‘challenging’ about the impact on the family also arises in the context of testing those at 1 in 4 prior risk of disease (Maat-Kievit et al. 1999 ).

Following these discussions, the testing would then usually take place at a third appointment. This allows the patient some further opportunity for reflection and also for other questions to be raised, especially if the patient’s companion in the clinic—often a partner or spouse—has less knowledge and experience of HD.

An adverse result will often cause distress but may nevertheless be very helpful. However, how does the patient think that the quality of her life will be altered by this knowledge, between the test result and the onset of disease? In many ways, this is the crucial question for someone at risk to consider. It must be disentangled from the practical question of coping with the disease once it has begun. In addition, and also very practically, what might the implications of testing be for employment or career, for health and life insurance, and for driving? These topics are the aspects of HD most likely to trigger institutional discrimination and this often arises as a topic of discussion in the clinic (Erwin et al. 2010 ). Fortunately, in many countries, those at risk of HD now have a degree of legal protection, although that does not prevent stigmatisaton at the level of personal behaviour and intimate relationships in HD and other neurodegenerative disorders (Bombard et al. 2012 ; Mendes et al. 2017 ).

In the event of a favourable result, some patients whose burden of risk is lifted from them may nevertheless experience some very negative feelings (Huggins et al. 1992 ). They may feel excluded from their family, as no longer sharing in this key aspect of family life, and may find it hard to communicate with those siblings still at risk. Others may regret decisions they have made in the past (concerning marriage, career, reproduction). Such possible consequences of a ‘good’ result may be unanticipated by many if they are not raised in the discussion before testing.

Explaining the limitations of testing is also most important: the possibility of a grey zone result; the possibility of a misleading result if molecular confirmation of the diagnosis in the family has not been achieved because no sample of DNA is available from an affected family member; and the impossibility of predicting with any accuracy when or how the condition might begin to manifest. This can lead into a discussion of the unexpected outcomes of testing outlined above, especially the often unanticipated negative impact of a favourable result.

Having gestured towards the complexity of factors to be considered in coming to a decision about testing, we can now return to the patient who is reluctant to engage in the discussion. Some who want the test without the talk have clearly brittle defences, so that it would be irresponsible to test them at first meeting. Others, however, find it difficult to consider the hypothetical scenarios (‘How would you feel if ….?’) that comprise the core of pre-test counselling (Sarangi et al. 2005 ). This may be expressed as, ‘I’ll deal with it when it happens’. This may be a question of personality, of coping style, of psychological defence, of cognitive limitation, or of a failure of the professional and patient jointly to establish a connection—a bond of trust. Whatever the reason, the failure to engage can obstruct the counselling. This, on its own, is not a reason for refusing to perform the test, however, but an indication to the professional that an adverse result in such a patient may lead to greater distress as there has been less opportunity for them to rehearse and prepare for an adverse outcome.

While a reluctance to engage with the genetic counselling process will often need to be discussed and to become a topic within the consultation, there may be other, practical barriers that should be considered carefully, especially if they could result in social or geograhical inequity of access to genetic sevices for HD (as for other disorders). Thus, in countries with sparse and dispersed populations, geography may be a major barrier to accessing such services (Hawkins et al. 2013 ). Creative solutions may then be very appropriate, such as outreach clinics or tele-consultations, as long as the quality of service provided does not suffer. Simply providing the predictive test without the usual high (we would hope) standard of counselling and support would not be professionally acceptable.

Patients who are ‘young’

As experience of predictive genetic testing for HD and other disorders ‘without treatment’ has accumulated over two decades, there has been a greater professional willingness to consider requests for testing from young people. While the international guidelines have suggested deferring a request for testing until at least 18 years but providing information, counselling and support to those younger than 18 years (MacLeod et al. 2013 ), others point out that there is no evidence of harm from testing children or young people (Michie 1996 ) and advocate for making predictive testing available to those younger than 18 (Duncan et al. 2008 ; Mand et al. 2013 ). However, the absence of evidence of harm is far from being evidence of the absence of harm. Consideration about the nature of the evidence that would be required to establish that harm had occurred is required; it may be impractical to gather such evidence, especially if the timescale involved might extend to decades.

Any age cut-off for predictive testing will be arbitrary, and this could indeed be counter-productive if those who cross the age limit then assume that they can have a test on request and without discussion. An age cut-off, however, may serve a useful function of protecting some very vulnerable individuals. Researchers who are enthusiasts for testing younger individuals sometimes report their research findings on those tested for a variety of disorders without distinguishing carefully between the conditions. Disorders with very different implications may be reported together, where some do and others do not have useful medical interventions. Down-playing these differences is unhelpful, and so is the lack of emphasis on the bias, among those tested for HD, who are willing to participate in follow-up research interviews towards those who test negative (a bias not so apparent for other disorders, where testing is clinically indicated). Clearly distinguishing tests for HD from those where the test is clinically recommended is crucial as the context of the decision to test is so different. There is a more general discussion about the genetic testing of children, and several of the international policy guidelines, in the ‘ Genetic testing of (young) children ’ section.

For the practitioner, there are competing tensions. We do not want to deny young people the potential benefits of testing, whether the test result is negative or positive, but nor do we want a young person to make a decision that they may soon regret. While at least some young people at risk of HD have an excellent understanding of the condition in context, others have been impacted by difficult family circumstances and their ability to make a good decision or their family’s ability to provide support afterwards may be impaired. Professionals will be concerned that factors such as denial, an adolescent sense of invulnerability linked to risk taking, an impulse to rebellion and a desire to separate from their family may all lead to the hasty making of decisions, if the young person has a limited ability to imagine the full range of possible test outcomes and how they may play out in their lives.

Some young people simply ‘wish to know’ without there being a specific decision that depends upon the result. We would suggest caution in such circumstances as the wish to know might apply only to the wish to have a good result. It may be difficult for any 18 year old to imagine their life at 30 years. How will decisions about relationships, courtship and reproduction work out in the light of different test results, as compared with a decision not to be tested (for the moment)? When and how will they raise the question of HD for discussion with a (potential) partner? Will it be easier to do that in the knowledge of their test result, or will that make it more difficult? On the other hand, they may be able to adapt well to their long-term prognosis at this age, as it may seem remote. Further research in this area, involving patient support groups and the accumulation of longitudinal case series, will be important.

Very helpful research into the experiences of young people at risk of HD draws attention both to the potential benefits of testing and the challenging circumstances faced by some young people (Forrest Keenan et al. 2015 ). Testing can be empowering and support the development of the young person’s identity. Factors that can complicate discussions with young people about testing for HD include the lack of family support that many experience, often indeed their sense of isolation. The response to test results can be especially difficult if the young person has only recently become aware of their risk or if the result differs from what they had expected. The impact of testing on subsequent family relationships may also throw up unanticipated difficulties.

In striving to develop a constructive approach to predictive testing in young people, it is very helpful to draw upon frameworks for understanding the normal processes of personality development. McConkie-Rosell and Spiridigliozzi ( 2004 ) set out a helpful framework while Richards ( 2006 ) examines the needs of adolescents in relation to genetic testing. Binedell et al. ( 1996 ) set out an approach to assessing the ‘maturity’ of adolescents seeking testing for HD. Some very helpful reports of testing young people for HD and other disorders have appeared, giving constructive suggestions as to how the genetic counselling process may provide improved support for this group. Thus Gaff et al. ( 2006 ) reflect upon their experience of genetic testing for familial bowel cancer in adolescents, while MacLeod et al. ( 2014 ) report the accounts and suggestions of young people after the process of testing for several different disorders (including HD) had been completed. There are good grounds for taking the patient’s age into account, through being especially open to flexibility with adolescents, but also being concerned to ensure that a request for testing is based on a considered and sustained wish, that the patient has been helped to think through the potential implications for relationships within the family, that the support available to the patient has been addressed carefully, and that the testing is not being sought as an act of defiance or self-assertion or as an automatic rite of passage for the 18th birthday.

Active non-directiveness

Returning to ‘non-directiveness’, it is necessary to distinguish the professional-patient relationship from a commercial transaction with a customer. If ‘non-directiveness’ simply meant doing what the customer requested, then, in our opinion, that would be an unacceptably shallow conception and a failure to fulfil our role. Simply following the patient’s instructions could amount to the professional acting on the basis of a patient’s whim or their ill-considered initial impulse in a difficult personal situation, which would be an abandonment by the professional of not only that particular patient but also their professional responsibilities. Rather than this, a practitioner has to provide the relevant information in an accessible way and raise factors for consideration that the patient may have not yet considered, as set out above.

Non-directiveness, rather contrary to its label, aspires to be an active process rather than a passive state. It can be difficult and challenging for the professional as well as for the patient; it demands full engagement with both the facts and interactionally with the patient in the consultation. There are many decisions about genetic testing that are much more routine because there is a clear medical recommendation to make, as in testing for certain susceptibilities to malignancy (Elwyn et al. 2000 ). It is still good practice to discuss the emotions that may be faced in those different circumstances but the intent in raising such issues is not to challenge a patient’s decision whether to be tested but to help improve their preparation for facing the test results. In contrast, in the context of HD, one is wanting to engage the patient much more in thinking//feeling her way through the decision and, potentially, to change her mind. We would want to help her to imagine her potential responses to the different outcome scenarios, so that she can effectively self-select as to whether she feels strong enough to go ahead with testing or not (yet).

In essence, active non-directiveness recognises that professionals will influence their patients, and that a process of influence is actually desirable. One can then work with this positively rather than simply denying or attempting to minimise any influence, as a shallow non-directiveness would do. Training for the genetic counsellor will aim to develop insight about the influence that is exerted and how to make this influence appropriate and helpful, to enhance the decision making process of the patient (Kessler 1992 ; Wolff and Jung 1995 ; Clarke 1997 ). This approach can be framed in the language of counselling or psychotherapy, or it can be viewed as grounded within a philosophical tradition. John Stuart Mill’s sense of autonomy incorporates questioning and challenging that promotes the other person’s autonomy (discussed in Lantos et al. 2011 ), but it can also very readily be considered from within a Kantian perspective or from within virtue ethics.

In the next section, we will move from predictive genetic testing to the question of pregnancy and prenatal diagnostic testing. There is a strong continuity between these two sections as the prenatal context is the other setting within clinical genetics in which the question of non-directiveness has particular salience and force.

Prenatal diagnosis, prenatal screening and disability

The use of genetic testing in various forms of prenatal diagnosis raises many ethical issues. These include questions about whether, and under what circumstances, it can be acceptable or even the best course of action to terminate a pregnancy. Here, we will focus on the questions that arise when the selective termination of pregnancy is being considered. Given that a woman and her partner wish to have a baby, under what circumstances is it acceptable or proper (even ‘to be recommended’) to investigate the embryo or foetus so as to make a decision about whether to continue the pregnancy or to end it, and perhaps try again. The key question is the process of deciding between a current conception and possible future conceptions: which are to be accepted and welcomed, and which are to be rejected and terminated?

Here, we consider this by looking at three subsidiary questions:

  • (i) How can prenatal screening or diagnosis be offered without the offer in itself conveying a strong recommendation (to accept the test and terminate the pregnancy, if the foetus is affected)? Is this best approached from the perspective of population health? Or should we concentrate our attention and efforts on implementing an actively non-directive approach in prenatal clinic discussions? Furthermore, what are the social conditions under which antenatal screening for genetic conditions could be offered without the social circumstances being, in effect, coercive?
  • (ii) How do we maintain respect for those affected by genetic conditions and congenital defects while offering couples such reproductive decisions?
  • (iii) What difference will non-invasive prenatal testing (whether diagnostic or screening) and population screening for recessive disease carrier status make (a) to the experience of pregnancy? and (b) to the experience of having an affected child?

Does offer = recommendation?

In the setting of the routine antenatal clinic, in which the offer of antenatal screening is made and considered, there are many processes that could routinise patient participation and thereby undermine the achievement of non-directiveness. Some of these processes depend upon the personalities and motivations of the professionals involved, but many will be impersonal and more structural, relating to clinic processes rather than the behaviour of individual al staff. It may require hard, active work on the part of staff to make it apparent to patients that they are being offered a screening programme and that they need to consider it carefully before climbing aboard the conveyor belt (Clarke 1991 ). However, the language used by practitioners may exert a strong influence which they (we) should learn to check and control. Examples of language use that needs to be watched (and challenged) include ‘This test is to provide you with reassurance’ and ‘Most people do this …’ (Pilnick 2002 ). These statements can be uttered in many different ways and with quite contrary intentions but they may be understood (‘received’) as being coercive, with the intention of directing patients to accept the ‘offer’ of routinised screening programmes.

In this context, it is helpful to reflect on the differences between the three main approaches to ethics. While the consequentialists look at the outcomes of reproductive decisions—often adopting a perspective from which they are confident in knowing a good decision when they see one—others prefer to be guided by principles (especially the autonomy of the pregnant woman), while virtue ethicists consider what it means to be a good (enough) parent or professional, and what decisions or actions such a parent or professional would be likely to make or to undertake.

Savulescu and Kahne ( 2009 ) urge us all to bring into this world the best quality infants that we can and to use all available technology to achieve this. Judgements about the ‘quality’ of infants, however, may be both highly contested and insuperably complex; is a foetus with cystic fibrosis one to be avoided? (Boslett 2011 ) There are questions of disease severity and treability to be considered. In addition, the contemporary focus on rights, especially the autonomy of the pregnant woman, could lead to a supine acceptance of any request made by a woman or couple but could also lead to discussions about the place of women in society, which we will return to in a later section. If women experience their lives as a burden because they are powerless in a patriarchy, does that make it acceptable to collude with the patriarchy and terminate female foetuses so as to save them from the same difficult lives as their oppressed mothers have lived? Finally, at what point can or does a woman become unconditionally committed to her foetus or child? Such commitment is regarded as virtuous, as an integral part of the nurturing that a good mother will provide (McDougall 2007 ). But does this begin at conception? at quickening? at birth? or, as in some ancient societies, once the child has survived the high mortality of early childhood? Or, in some more contemporary circumstances, the checks on ‘foetal quality’?

The consideration of a mother’s commitment to her infant reminds one of a mother’s deferred commitment to her foetus that Katz ( 1986 ) described so eloquently as a consequence of the introduction of antenatal screening by amniocentesis: commitment to the foetus//baby was delayed until the results of chromosome testing were available. This deferral of commitment, however, came at a social and personal cost. The very experience of choice about participation in screening can be experienced as a burden (van Berkel and van der Weele 1999 ) and it is difficult, although perhaps not impossible, for this choice to be offered in a ‘neutral’ or non-directive fashion (Clarke 1991 ; Rapp 2000 ; Williams et al. 2002 ). In the routine antenatal clinic, discussions about screening for genetic disorders often concentrate on the facts and not on the meaning and weight of the decision to be made and its implications (Hodgson et al. 2010 ). The difficulty of the decisions that mothers are asked to make has been explored in several studies. The notion that the pregnant woman may feel a conflict between different ethical principles in coming to her decision may not be the most helpful way to understand her position. Instead, it may be recognised that she experiences a conflict of interests between her foetus and other members of the family, perhaps especially her other children (García et al. 2009 ).

For a couple to terminate a wanted pregnancy is always going to be difficult and distressing. The long-term sequelae of such terminations on the grounds of foetal abnormality have not been much studied and the work of van Mourik remains valuable (White-van Mourik et al. 1992 ; White-van Mourik 1994 ). The support needs of the women and couples involved are substantial and the recognition that the mother sometimes has the moral role of taking on herself the suffering of ‘the child who would have been born’ can be important.

A singularly helpful study of the decisions made by pregnant women (and their partners) about antenatal screening, as well as other types of genetic testing, is that of Scully et al. ( 2007 ), in which it is shown how women will often make a series of microdecisions in a pregnancy, deferring any important decision to the last possible moment. It can be too difficult and challenging to confront their situation in its full complexity and so they may focus on only what is immediately in front of them. It will not usually be our role to confront those who respond to events in this way in any but the most gentle fashion.

The situation of women considering prenatal diagnosis because of a known high risk of disease in their foetus, arising from a family history of a rare disease, is very different from the experience of women in routine antenatal screening. The complex and agonising decisions to be made when a prospective parent is at risk of Huntington’s disease, for example, are well recognised (Downing 2005 ; Klitzman et al. 2007 ).

When we turn to a population focus on a single disease, and look at the collective consequences of decisions made by many individuals, it becomes clear how difficult it can be to decide when a collectivist attempt to give families genetic information and choices becomes an inappropriate attempt to lead them to make specific decisions. Thus, the paper by Helderman-van den Ende et al. ( 2013 ) on pregnancies at risk of Duchenne muscular dystrophy (DMD) shows that many female carriers who could have been identified as carriers in advance of the pregnancy had not in fact been identified. Using a package of measures, including mutation testing on chorionic biopsies found to be female at prenatal diagnoses for risk of DMD and more effective family cascade testing, it would be possible to recognise more female carriers before they have affected sons. These carriers could then be offered the opportunity to have genetic counselling. The stated concern in this paper is to achieve the prevention of more cases of DMD, while fully respecting the rights of the women involved to make their own decisions. What perspective should one adopt on this issue? Is the population-level concern appropriate, or is it likely in practice to prove coercive, or at least manipulative and directive? If we assert that genetic services are merely working to ensure that prospective parents are able to make informed reproductive decisions, is our response adequate or are we evading the moral question at the heart of this issue?

One recommendation of the paper (Helderman-van den Ende et al. 2013 ) would be to adopt a different approach to the genetic testing of children for carrier status. If we do this, are we being eugenicist, i.e. protecting ‘Society’ from the consequences of decisions made by family members who have not faced up to their responsibilities, whether the decisions are made actively or by default? Or are we simply improving the services available to DMD families? How should we discuss and evaluate the cumulative population consequences of multiple decisions made by individuals?

In the context of countries that are poorer and less developed, how do the arguments play out in relation to genetic testing or population screening for carriers of beta-thalassaemia? (see also the “ Ethical issues in our multi-cultural UK society ” section).

Does it devalue an individual with a genetic condition if society chooses to establish antenatal screening programmes to reduce the numbers of individuals born with the same condition? Do they have a right to object that this is disrespectful and promotes discrimination against them? This concern has been labelled the expressivist objection to antenatal screening programmes for genetic disorders.

One approach to this is to talk with those affected by Down syndrome, for example, and see how the question of antenatal screening ‘for’ (‘ against ’, perhaps? ) their condition impacts on them. As anyway feeling devalued, it is no surprise that antenatal screening can add to their sadness (Alderson 2001 ; Barter et al. 2017 ). This does not give affected individuals a veto over such policy questions (Shakespeare 1998 ), but it certainly reinforces the need to be as sensitive and respectful as possible in the implementation of such a programme and the need to take great care in producing information about screening. What is one to say about Down syndrome to inform prospective parents who have no personal experience of individuals with the condition? How can information be provided in a ‘balanced’ manner, when one is proposing to eliminate such individuals? (Hippman et al. 2012 ). The ‘twice-told tales’ element—the difference in information provided during pregnancy or after birth—was pointed out by Lippman and Wilfond ( 1992 ) and applies as much today as ever before. If we shift the provision of information to the preconception setting, could that give prospective parents a chance to reflect more calmly and come to a decision with which they can subsequently feel confident (Schoonen et al. 2012 )?

If we shift to consider disorders for which prenatal diagnosis may be sought by those with a family history, but which are not generally included in antenatal screening, how are the considerations different? What can we learn from reflecting on them? In contrast to population screening programmes, there will often be much less need to provide information about the disorder, as it will often be well known to members of the family. However, the factors that shape the decisions about reproduction may have no less impact on affected members of the family, who may be greatly saddened that a close relative—a sister or brother, a son or daughter—has decided not to risk having a child like they are. This is not the invariable response—some affected individuals would encourage relatives not to transmit their condition—but it is a possible response. It can be thought of as the ‘expressivist objection’ to antenatal screening but at a much more personal and intimate level, within the family. This has been studied recently within families affected by X-linked hypohidrotic ectodermal dysplasia (XHED) (Clarke 2013 , 2016 ) and spinal muscular atrophy (Boardman 2014a , b ) and it is to be noted that the different modes of inheritance (sex-linked and autosomal recessive, respectively) alter the dynamics of decision-making within families.

One aspect of coping with a genetic disorder, which also often influences the making of reproductive decisions, concerns the stigma attached to the specific condition. In XHED, for example, the stigmatisation that affects many affected males can be a major determinant of a female carrier’s decisions about bearing children. She will often have seen her male relatives struggle with the stigma and cope with varying degrees of success. This may be at least as important a factor in her decisions as the physical and clinical, medical problems of life as an affected male. One wonders how this plays out in other disorders, where there is a combination of physical anomalies with unusual appearance and also, perhaps, some cognitive impairment. Which aspects of these disorders is most important for the affected person and for his or her relatives? How major a contribution is made by the prospect of stigmatisation? One study that gives some insight into the strength of feeling around these issues is the interview-based study of Kelly ( 2009 ), which shows how the parents of children with serious disorders often prefer to avoid making decisions about prenatal diagnosis that would cause inner conflict; they may avoid conceiving, or avoid prenatal diagnosis, or at least avoid terminations of pregnancy. They choose not to make choices.

Would we be comfortable with a decision about prenatal diagnosis and the termination of an affected pregnancy on the grounds of ‘the cost of care’? If we mean the cost of care borne by the state, then we are back in the days of state-sponsored eugenics. However, in these times of state retrenchment and public austerity, what about concerns by families that they would find it difficult to fund the facilities that their affected child would be likely to require? One example put forward has also related to XHED, where the costs of dental care may be substantial (if dental implants are preferred to the much cheaper dentures)? (Aldred et al. 2003 ). To paraphrase, concisely but perhaps unfairly: ‘The termination costs less than dental implants and we would not want our child to make do with dentures’. How do we respond to that?

More generally, we can consider how—in what terms—to discuss normality and disability. This maps across onto the concepts of social inclusion or exclusion, normalisation or marginalisation. There is a rich literature here that deserves to be explored by those of us in clinical genetics, whose daily work involves us in using but often not examining the concepts of disability and normality (Bridgens 2009 ). This exploration can usefully involve developing personal relationships with disability rights groups to explore the issues, as well as a purely theoretical engagement (Peterson 2012 ). Engagement with the experience of those living with impairments can bring valuable insights, both through personal relationships and from second-hand encounters in the literature. Important studies and narrative accounts are available of life with numerous genetic conditions, such as achondroplasia (Ablon 1984 ; Adelson 2005 ), neurofibromatosis type 1 (Ablon 2013 ) and HD, including the dimension of life with the risk of HD (Konrad 2005 ; Browner and Preloran 2010 ). The accounts of parents whose children have neurodevelopmental difficulties from infancy, for whatever reason, have demonstrated the recalibration of expectations of their child and of their role as parent (Voysey 1975 ; Landsman 1998 , 2003 ) that is familiar also in ‘quality of life’ assessments of adults with a range of impairments: those affected are much less dissatisfied with their lives than ‘objective observers’ think would be appropriate (Scully 2008 ). The implications of this can be read in different ways. This question of how to read such different perspectives is at the heart of the disability rights movement, which is not prepared to concede that the lives of disabled individuals are not as worthwhile as those of anyone else.

Studies of illness that are grounded in a phenomenological perspective are bringing deep insights but have usually addressed acquired or later-onset disease (e.g. Carel 2016 ) rather than congenital disorders manifest from birth: phenomenological studies that address these conditions would be most valuable.

The field of disability studies has flourished for some decades. This has led to the firmer underpinning of ‘advocacy’ and a strong link between the experience of disability and its analysis within academia (Parens and Asch 2000 ; Ablon op cit). There has been debate about whether it makes sense to speak of a collective disability identity, or whether the different types of disability and impairment are too disparate for this to be coherent (Scully 2008 ). This includes efforts to anchor the experience of physical and sensory impairment in the sociology of the body. The field of disability studies has also retained its sense of the social roots of disability and the context from within which disablement arises (McLaughlin et al. 2008 ), and has absorbed much from feminist theory (Woodin 2012 ). Through activism and advocacy, it has also contributed to the welfare of many with motor impairments, so that the facilities for wheelchair users in the UK have improved very substantially over the past four decades. The representation of the experiences of those with cognitive impairment is less developed; those with severe cognitive impairment may be especially vulnerable to having their status as fully human persons undermined (Edwards 2006 ).

One particular area in which further empirical investigation and conceptual clarification would be expected to be immensely helpful is that of the disorders of sexual development and gender assignment decisions in infancy, where ‘treatment’ and even surgery may be driven by parents’ confusion and their understandable distress but where such pressures may in the long run work to the detriment of the patient (Feder and Karkazis 2008 ).

Reproduction into the future

Finally, in this section, we turn to consider some of the new technologies and how their application to reproduction may change the experience of pregnancy and raise new issues or raise familiar issues but more powerfully.

With the advent of non-invasive prenatal testing (NIPT) through the sequencing of cell-free DNA in the maternal plasma, a modest fraction of which derives from the foetus (technically, the chorion), several questions arise. At present, women who are reluctant to undergo any sort of prenatal genetic screening can explain that they wish to protect the pregnancy from the risk of miscarriage associated with invasive procedures. This account of why she declines a screening test may be used to justify herself to the professionals or to critics within the family. The advent of NIPT undermines this ‘excuse’ and may make it harder for her to sustain her avoidance of testing. To what extent this is an important consideration—how widespread this practice has been—is unknown but it is clear that NIPT has the potential to become even more routinised than the current types of antenatal screening, with consequences for the validity of consent for such tests (van den Heuvel et al. 2010 ).

While NIPT may complicate the decisions of those reluctant to use prenatal diagnosis, it will in other ways bring some very clear benefits in that fewer miscarriages are likely to result from invasive procedures as fewer will be performed. NIPT can operate as either a first-tier screening test or an intermediate tier (between the first tier of serum-plus-nuchal-translucency screening and invasive procedures). NIPT is substantially more sensitive and specific than current first tier screening so there will be many fewer amniocenteses and fewer miscarriages. However, it is important to remember that the positive predictive value (PPV) of NIPT for Down syndrome (DS) varies with the chance of DS in that pregnancy and is far short of 100%; for a standard risk pregnancy (not already known to be at higher risk) the PPV is about 80%; as a second tier test, its PPV is at least 90% (Taylor-Phillips et al. 2016 ; Nuffield Council 2017 ).

If a majority of pregnant women continue to use antenatal screening, the proportion of DS foetuses that are identified through screening and then terminated will increase, so the number of infants with DS at birth will decrease. How will this be experienced by current and future persons with DS? And by their families and the professionals who care for and support them? There will certainly be some sadness (Skotko 2009 ). How do we place a value on this, when we see DS as another way of being human rather than a disorder? This is where the question of ‘balance’ in representation becomes so important but, ultimately, cannot be achieved. On the one hand, DS may be seen as a ‘disorder’, as it can have grave medical consequences as well as sometimes a very severe impact on cognitive development. On the other hand, many with DS experience fulfilling lives and make a positive contribution to the lives of those around them. The problem is that the different ways in which DS is being examined may be incommensurable.

As screening by NIPT is implemented more widely, the range of disorders it is being used to identify is also broadening out to include many recognised chromosome microdeletion syndromes. The decision as to what conditions to include in screening seems at present to be driven largely by considerations of marketing, with companies seeking to claim that their NIPT test “covers more” than their competitors. However, the performance of the test changes as new conditions are added to the testing package, and the PPV for additional tests is usually lower than for DS, so that the advantage of NIPT over other approaches to screening for the autosomal trisomies may be lost: the number of invasive tests performed, and so also the number of procedure-related miscarriages, would both increase. These and related issues are considered with care in a recent report (Nuffield Council 2017 ).

What other conditions may be included in screening, beyond chromosome deletions? Women should be aware that incidental findings may emerge, including (rarely) evidence of malignancy. Will parents pay to find out about traits or predispositions for which there may be no clinical utility (and would therefore not be included in evidence-based health care)? It may also be possible to assess the foetus for non-medical traits or for adult-onset disorders, for which it would usually be regarded as inappropriate to test young children (Deans et al. 2014 ). Where should society draw limits as to what is permitted, or should this be left to each individual in the marketplace?

Foetal sex is another trait that parents may be interested to determine early in a pregnancy. While superficially harmless, this may be used to terminate pregnancies carrying a female foetus in societies that devalue women. Indeed, late terminations of female pregnancies (after ultrasound scans) have very substantially skewed the sex ratio at birth in parts of India and China. As an act of solidarity, we believe that foetal sex should not be determined by NIPT except when it is clinically relevant in relation to a sex-linked disease (Nuffield Council 2017 ).

The combination of NIPT and carrier detection by high-throughput DNA sequencing has the potential to eliminate not only Down syndrome but also, in western countries and wealthy countries elsewhere, all chromosome copy number anomalies (at least all recognised pathogenic deletions and duplications) and virtually all cases of autosomal recessive disease (Edwards et al. 2015 ; Human Genetics Commission 2011 ). There would clearly be difficulties with the implementation of such a programme, especially the problem of sequence variants of uncertain pathogenicity that are detected in autosomal recessive loci. However, there would in addition be profound consequences if ‘society’ were either to make a collective decision to prevent the birth of people affected by these conditions, or allow social pressures to work to the same end for those who could afford the technology. Society in general, and health care in particular, would look very different. The argument against screening would lead us to remain in our present situation, where prenatal diagnosis to avoid the birth of a child with a serious recessive disorder is only available once a couple has had at least one affected child. Should prenatal diagnosis and the selective termination of an affected foetus only be available under those circumstances? (Why?)

This brave new future, with few children affected by chromosomal or autosomal recessive disorders, could be represented either as a major triumph or as an inhuman dystopia, in which something vital has been lost. How are we to make up our minds?

Another challenge that will soon arrive in the clinic is that of rational, gene-based therapies that can be applied in utero. The circumstances in which such treatments may be preferred to the current use of prenatal diagnosis and the selective termination of affected pregnancies will have to be considered and defined. Cost will be a major issue: will private or state-backed health insurance schemes or national health care schemes be willing to provide such treatments if they are both costly and of uncertain or incomplete effectiveness? (The field is reviewed from a multidisciplinary perspective in Schmitz et al. 2018 ).

Genetic testing of (young) children

We have already discussed some of the issues that are raised by the genetic testing of young people in the ‘ Predictive genetic testing consent, competence and (non)directiveness ’ section. Here, we turn to consider decisions about genetic testing of young children, who are unable to participate in the discussions and decisions about this. What genetic information is it appropriate to generate about young children?

We take for granted that diagnostic testing is (almost) always appropriate, if the cause is being sought for a clinical condition currently affecting a patient of any age. When a sick child is being investigated, it may be wise to raise in advance with the parents the possibility that the disorder may be genetic, perhaps with familial implications, but that is not a reason for not performing the investigation. What must be considered is when it would be appropriate to perform tests of no relevance to the child before s/he reaches maturity. This includes predictive genetic testing for adult-onset disorders and carrier testing for autosomal recessive disorders, or for sex-linked disorders and chromosome rearrangements.

While reflecting on this, there are some subsidiary questions to consider.

  • How do children find out about the problem in their family?
  • What are the factors to be considered in relation to the predictive testing of young children for an adult-onset condition, or carrier status testing for any disorder, when the condition is already known to run in their family (when the child is too young to participate in the decision-making)?
  • Does the weight of these arguments change when incidental findings emerge from testing a young child, of which the family would otherwise be unaware?
  • When is it reasonable to test a young child as a preliminary to arranging their adoption?

Finding out

Perhaps the first question to consider, before thinking about genetic testing, is how children come to find out about a genetic condition in the family. In discussing with parents in clinic ‘how to tell the children’, it would be common to encourage open disclosure but in a gentle and age-appropriate manner, in contrast to the idea of keeping the problem secret and then planning a disclosure session as a single, discrete event. Such an event could be daunting for the parent and emotionally traumatic for the child, if they are old enough to appreciate what is being said. The decision not to mislead or deceive but to give the information sought by a child over time, as they mature, is a policy that is simple to recommend but there have been few studies of how parents do in fact communicate difficult genetic information to their children.

The first study to address this issue is probably that of Manjoney and McKegnay ( 1978 ), which described the cycles of non- or mis-communication between generations in families affected by polycystic kidney disease. A failure to pass on information in a timely and supportive manner condemns the next generation to discover their diagnosis in the same difficult, and sometimes catastrophic, fashion as their parent had done. A few studies have reported the experience of communication within families affected by Huntington’s disease. Holt ( 2006 ) reported the experience of two contrasting families, and came down strongly in favour of a drip-drip trickle of information: a gentle, supported and slow-paced disclosure of information to children. Etchegary ( 2006 ) reported a typology of patterns of discovery, with four trajectories of finding out: (i) ‘something’ is wrong; (ii) out of the blue, (iii) knowing but dismissing, and (iv) growing up with HD. It is the period of ignorance before finding out that shapes the account given by the ‘child’ (perhaps by now an adult) about their recognition of their situation of risk.

One large study has looked at these issues in the context of several disorders and found that children prefer the slow and gentle drip feed of information over time (Metcalfe et al. 2011 ). Further research in this area, in other disease contexts, would be important and worthwhile.

Current policy

Numerous professional bodies have recommended caution, at least, in performing genetic tests on children unless it is clearly a diagnostic test (to explain a problem affecting the child now) or is needed to arrange treatment, prevention or health surveillance during childhood (British Society for Human Genetics 2010 ; Borry et al. 2009 for the European Society of Human Genetics ESHG; Botkin et al. ( 2015 ) for the American Society of Human Genetics ASHG; and Ross et al. 2013 for the American Academy of Pediatrics AAP and the American College of Medical Genetics and Genomics ACMG). These position statements are very similar, except that the AAP/ACMG statement, but not the ASHG policy, has been disappointingly weakened from its previous joint position with the ASHG from 1995. These more cautious policies are based on two considerations (i) that it is important to preserve the autonomy of the child so that, as a future adult, she is able to make her own decision about testing. This in effect applies one understanding of Feinberg’s concept of an Open Future to this area of competing rights and interests, and (ii) that testing the child as a child results in the loss of privacy of their test result and may harm the child if it influences parental or institutional behaviour towards her, with scope for (perhaps subtle) discrimination between siblings whose test results differ.

A libertarian or an enthusiast for testing children might argue that such decisions should be left to parents, who often make proxy decisions for their children, and this decision should be treated as no different. The drawback of such an approach is that many parents will want to know the genetic status of their child for any disorder that runs in the family. Once they have tested the child, that information will be known within the family but it removes the possibility of the child playing a key part in discovering the information. This may be especially important if adjustment to an unfavourable result is substantially better when the at-risk individual has made their own active decision about testing as a ‘mature enough’ adult and has not had testing imposed on them when younger.

The shift in policy between the previous ASHG/ACMG policy and the current AAP/ACMG document (Ross et al. 2013 ) is contained in this sentence, ‘After careful genetic counselling, it may be ethically acceptable to proceed with predictive genetic testing to resolve disabling parental anxiety or to support life-planning decisions that parents sincerely believe to be in the child’s best interests’. This could be seen as an invitation to any parents focused on their ‘right to know their child’s genetics’ to claim that they are disabled by anxiety, or that they need to take the child’s genetic status into account in making important decisions (such as, presumably, investment in their education or health). It is unclear to us that testing would in fact be justified under such circumstances and this policy formulation appears too open to abuse, especially in a country with a largely private system of health care.

In one ‘ideal’ scenario, information that is not clinically indicated or useful until the child is mature would not be generated until then. The ‘child’—by then a Young Person or perhaps a mature adult—makes the decision to be tested. Their involvement in the decision, with appropriate counselling support, gives them a choice and makes it easier for them to accept an adverse result whereas, for a parent to give them an adverse result generated years before may be very difficult for the parent and may lead to anger, resentment, disbelief or rejection on the part of the ‘child’. The testing will have been performed in childhood at the request of the parent(s) and for their own reasons. The child’s interests have been overridden by the parents’ (often misplaced) concern or by curiosity labelled as “disabling anxiety”. If the professionals comply with the inappropriate parental request, they are very likely to have been negligent of the best interests of the child. This sets up problems for the future, with perhaps a failure ever to disclose the result to the child, or a disclosure made in difficult, perhaps damaging, circumstances.

We know that carrier information generated about young children or a foetus is often not transmitted to the child (Jolly et al. 1998 ; Jarvinen et al. 1999 ); the same question has not been studied in relation to ‘clinically inappropriate’ predictive testing, but our experience of inappropriate parental requests leads us to believe that they are likely to find it even more difficult to pass on that type of information.

Our ‘ideal scenario’ has the advantage of demonstrating trust in the child to make a wise decision about testing, which creates an opportunity for personal growth in both parent and child.

The case against testing a child without a clear medical indication will not always be so strong. In carrier testing for autosomal recessive disease there may be much less at stake than in predictive testing for HD. If families can accept such carrier information and pass it to their children in an open and constructive way, then much may be gained and little lost by the early testing (Vears and Metcalfe 2015 ). In testing for carriers of sex-linked and chromosomal disorders, however, the chance of unhelpful consequences is greater, such as the stigmatisation of carriers and altered expectations of their future roles and relationships. Such problems are still more likely to occur in late-onset dominant disorders where there is no health benefit from testing in childhood. Professional discussion of many varied scenarios is presented in Arribas-Ayllon et al. ( 2009 ) and Parker ( 2012 ). Problems may arise when family loyalty to an affected sibling is in tension with the welfare of future children, as with carrier testing in the unaffected sibs of those with X-linked disease (James et al. 2003 ) or autosomal recessive disease (Fanos and Johnson 1995a , b ).

Other difficulties arise with predictive genetic testing if it is difficult for the parents or the children to live with the results. Although they have existed for some years, we still find some predictive information difficult to handle and to explain to parents. Tests that indicate a substantial risk of behaviour problems in childhood and of schizophrenia in adolescence and adult life are very difficult for parents to manage (Hercher and Bruenner 2008 ). How does one respond to the difficult behaviour of a child who (one fears) may be destined to develop psychosis? Does one give in to all challenges so as not to cause frustration and resentment? Or adopt a ‘zero tolerance’ approach of very firm boundaries? Or try to treat the child as if one had not been given this information? And if your child is found to be at risk of sudden cardiac death (Hendriks et al. 2005 ), how does one decide what pattern of exercise and social life to recommend? And how can this be enforced? As discussed in the ‘ Predictive genetic testing consent, competence and (non)directiveness ’ section 3, some families in which children were tested for their cardiomyopathy have regretted the decision to arrange the testing (Geelen et al. 2011 ). It will be important for researchers to build long-term relationships with families in which these issues arise if we are to gain insight into responses to different patterns of professional practice.

Incidental findings

An additional consideration about testing children for late-onset genetic disorders has arisen in the setting of genome-wide testing, especially exome and whole genome sequencing. The incidental identification of a (probably) pathogenic variant, likely to lead to a late-onset disorder such as cardiac disease or a malignancy, may occur in a child who is being tested to determine the cause of a different condition, perhaps a complex neurodevelopmental disorder. If the family has so far been unaware of this risk, then this incidental recognition of a disorder may, in some circumstances, be very helpful as one of the parents is likely to carry the same pathogenic variant. Recognising this may be to the parent’s great clinical benefit, and hence to the benefit of the child too.

This approach was proposed by a working group of the American College of Medical Genetics and Genomics (ACMG), being proposed as a requirement for any genome sequencing of children, whether in research or diagnosis (Green et al. 2013 ). A list of 56 genes was to be examined with the disclosure of pathogenic variants, whether or not any prior consent had been obtained. This policy was subsequently revised by the ACMG Board in 2014 with the concession that disclosure should not be required without prior discussion and parental consent, which brings the ACMG recommendations more into line with those of other bodies such as the ESHG.

While the ACMG policy has some strong arguments to recommend it, the provocative wording of the initial policy and the rush to see it implemented betrays the True Believer enthusiasm of some members of the working group. There is a vision of a person’s genome sequence being available throughout their life as a resource for use in making any health care decision. However, there are both technical and ethical problems with this. What we accept as a genome sequence will not be the same in 10 or even 5 years as is current today and the practicalities of data storage are not simple, so that repeating the genome sequencing when required may be a better solution than long term data storage with its difficulties of perpetually needing to update the IT systems, both software and hardware, and of ensuring privacy and data security (Chadwick et al. 2013 ; Clarke 2014 ). There are also political difficulties with this model of genetics in health care, including the inappropriate individualisation of health policy and the neglect of equity. Despite these quibbles about the details, however, the ACMG report did set out a convincing case that the incidental recognition of a late-onset genetic disorder in a child should generally be disclosed to the family, unless it is known that the family is already aware of the problem identified. A decision not to seek genetic testing by one who knows s/he is at risk is altogether different from their failure to seek testing because they are ignorant of the risk.

When we turn to adoption, there are other challenges. Health professionals are usually reluctant to carry out genetic testing on a child being considered for adoption if the test would not usually be performed on that child under different social circumstances. In contrast, those from a social work background may see the long-term advantages to the child of being adopted as crucial to the child’s future. If that takes one or two additional genetic tests, then a successful adoption would be well worth the price. How do we resolve this difference in perspective, when both paediatricians and social workers are involved in the adoption process?

The restrictive view of genetic testing in adoption has been put forward by several authors (Morris et al. 1988 ; Newson and Leonard 2010 ) and challenged by others, who see that the interests of a child may indeed be bound up in the outcome of a decision about adoption (Jansen and Ross 2001 ). It has also been argued that genetic testing, even a genome sequence, may be appropriate in children being considered for adoption because the family history information available to the prospective adopters is often poor and the genome sequence will compensate for this (May et al. 2015 ). Counter arguments to this last claim are numerous and persuasive, for the present, as the interpretation of a genome sequence is still far from straightforward and the introduction of yet more genomic uncertainty into the portfolio of information about the child is unlikely to help the prospects of her or him being adopted.

The debate so far has largely related to information about a known risk, most often Huntington’s disease (HD). While health professionals may be more open to persuasion in relation to genetic testing for some cancer predispositions, such flexibility is much less likely in relation to HD given the small number of those at risk who seek testing as adults, the great potential for stigmatisation and discrimination, the altered expectations likely to be held by parents about the child with an adverse result, and fear for the child’s future (including the prospects for being adopted) if s/he tests positive. Improving prospects for treatment may alter this but, as yet, they are still too remote and uncertain.

The force of the ‘ideal’ setting discussed above is still applicable in the setting of a possible adoption, with the implication being that the ideal adoptive parents are those who can accept that the child has a risk of future disease but are willing to adopt in any case, without insisting on certainty either way. If there were an excess of prospective adopters, then a ‘Judgement of Solomon’ solution might apply. Whether or not that surfeit of adopters is to be found for any particular child will vary between communities and countries.

The other setting in which genetic tests are considered in relation to adoption is chromosome evaluation for neurodevelopmental difficulties, with the method now in use being array comparative genomic hybridisation (aCGH). This has largely replaced chromosome analysis but the paediatricians engaged in adoption work are still often finding (at least in the UK) that the greater diagnostic yield of aCGH is accompanied by a challenging injection of uncertainties from the variants of uncertain significance (VUSs) in the result and, less often, the incidental findings of disease risk for some other problem in the birth family. Of these difficulties of interpretation, it is the VUSs that impact on the adoption process. Prospective adoptive parents may be willing to adopt a child with known learning difficulties, whether or not there is an explanation apparent on chromosome studies, but less willing to accept a child where there is a cloud of uncertainty around the interpretation of the aCGH report.

One can foresee the time when genome-wide investigation of children to be adopted might become standard practice, at least in some jurisdictions, although we would prefer to continue without generating such clinically inappropriate datasets. The notion that genomic investigation will compensate for the lack of a detailed family history in a child being adopted is not scientifically plausible and certainly not a convincing reason to override the usual considerations of medical ethics. Furthermore, if genomic investigations are performed on a child being considered for adoption, or already adopted, a plan must be in place as to how to pass potentially important findings to the birth family, if they arise.

High-throughput testing for genetic disorders

High-throughput investigations have transformed biology as a science. Instead of experimental design for the testing of hypotheses, the frontier of research is now the development of tools for data interpretation: making sense from the data. The same change has impacted on clinical investigations: it is no longer so difficult to generate sequence information, but it can be very challenging to make useful diagnostic sense of it for the individual patient. The history of human laboratory genetics was the focusing onto progressively smaller details—starting with the chromosome and working through linkage studies with close markers, to the specific gene and then its sequence, looking for point mutations. Over the last decade, however, that level of detail has become available across the whole genome at a single step, the whole genome sequence (WGS).

Given this fundamental shift in the methods of laboratory genetics (Clarke et al. 2012 ), it is now necessary to choose what information to attend to and to pass along the chain from laboratory scientist to bioinformaticist to clinician, and then further, to the patient and family. While it is possible to pass the entire WGS to the patient as a data file, this would miss the point: patients look to health professionals for more than that. They want and require an interpretation and recommendations that are supported by the available evidence.

What difficulties are raised by these new circumstances? The two major problems to address have already been met above: (i) variants of uncertain significance (VUSs) and (ii) incidental findings (IFs), also known by other terms such as ‘off-target’, ‘additional’ or ‘secondary’ findings.

Incidental findings and variants of uncertain significance

VUSs are not entirely new to genetics, of course. They have been recognised for as long as laboratory genetics has existed, but it is the scale of the problem that is new. The principal current approach to VUSs is that advocated by Berg et al. ( 2013 ) and subsequently adapted by ACMG (Richards et al. 2015 ) of a systematic approach to classifying each VUS as being more or less likely to result in disease, using five ‘bins’ to one of which each variant is assigned (not pathogenic; unlikely to be pathogenic; of uncertain pathogenicity; likely pathogenic; definitely pathogenic). The grounds on which these assignments are made are well recognised and will not be rehearsed here except to note that surprises can occur: a likely pathogenic variant may be reassigned to the ‘definitely not pathogenic’ bin, as new information and experience accumulates, or a variant thought likely not to be pathogenic may turn out to be pathogenic. The process of making these assignments is described by Timmermans ( 2015 ) in a study that examines the trust placed by laboratory scientists in their laboratory and bioinformatic processes and in each other and in their clinicial colleagues. They then make judgements as to which variants, in which genes, should be reported to the clinician who has requested the investigation. The clinician will then pass to the patient or parents the clinically important, ‘definite’ results that are relevant to the patient’s problem, but then has to decide when to report any of the other results, especially if a VUS has been found in a gene thought likely to be relevant to the disease in the family. If this VUS may be pathogenic and further investigation may clarify this, then of course it must be disclosed, although that will sometimes lead to families misunderstanding this; they may believe that the VUS must be pathogenic and act on that misunderstanding.

The variants found may be sequence variants or copy number variants (CNVs). CNVs are detected by array comparative genomic hybridisation (aCGH) as well as by sequencing methods, especially WGS (Boone et al. 2013 ). There is more experience with reporting CNVs to families than sequence variants but they still generate many difficulties for both the clinician and the patient or family: how does one ‘act on’ or understand a CNV result that indicates a modest susceptibility to attention deficit disorder, autistic traits, schizophrenia or intellectual disability? Information has to be imparted in a very tentative manner to help prevent families jumping to unwarranted conclusions.

Addressing only the question of IFs, and restricting this to only the 56 genes that ACMG initially recommended for assessment for possible IFs, one recent study found that the mean number of likely pathogenic variants was 1.69 per patient (Jurgens et al. 2015 ). Broadening out to WGS as a whole, there are even more serious problems of interpretation. The level of uncertainty of interpretation of many of the variants found when performing WGS on healthy volunteers, the incomplete sequencing coverage (even of known disease genes), and the inadequate knowledge base for deciding when to trigger further diagnostic investigations (Dewey et al. 2014 ) suggest that WGS may be best regarded as a research investigation rather than a regular clinical tool, except as a last resort in the face of serious disease without a diagnosis, and when the clinician has made her best efforts using currently established methods.

Consent and Recontacting

When explaining WGS, aCGH, clinical exome analysis or even a large gene panel to a patient or family, it is essential that the issues of VUSs and IFs are discussed. Without that, valid consent cannot have been given to the investigation. However, it is here important to take a short detour into the question of consent more generally. ‘Taking consent’ and, ‘giving’ it are the two sides of one activity, a social encounter. What is going on for the two (or more) participants in this encounter? The professional will want to put on a convincing performance as a competent, caring and ethical professional but, at the same time, s/he will want to ‘get the box ticked’ and move on to the next task. Similarly, the patient or parent will (usually) wish to perform as a sensible and responsible patient or parent but may not want to read through pages of typescript in a small font and, in any case, may be too anxious really to consider the details of what is being proposed. This will often be a recipe for collusion between both parties to have a brief but not challenging or intense discussion, so that neither party has to dwell on it. They can hurry through the consent step as not much more than a formality. Further work is needed to examine ‘what is going on for the participants’ in the setting of consent for diagnostic genomics, along the lines of work by Corrigan ( 2003 ) on clinical trials and Shipman et al. ( 2014 ) on biobanking.

Consent for genomic investigations has to be ‘broad’ (Helgesson 2012 ). The line between diagnostic and research procedures may (understandably) be blurred—as in the 100,000 Genomes Project of NHS England - so the need to avoid the misunderstanding that participation in the research will have therapeutic value (i.e. the ‘therapeutic misconception’ familiar from drug trials) is clear (Halverson and Ross 2012 ).

The initial suggestion of ACMG (Green et al. 2013 ) that information be returned even without prior consent was flawed but useful in at least three ways: (i) it stimulated interest and debate, (ii) it set a defensible limit on the list of loci about which IFs should be returned, thereby reining in the potential competition between research groups and among commercial companies in what IFs would be sought and disclosed, and (iii) it drew attention to the difference between the usual ‘Genetic Testing of Children’ scenario, where there is a family history of a specific disorder, and this very different situation in which there is often no prior awareness of a genetic diagnosis in the family so that there could well be further casualties of the condition (perhaps the parents or other relatives) if the incidental finding were not acted upon. The detailed composition of the ACMG list can be debated, and has been revised (with four additions to, and one removal from, the list that now totals 59 genes: Kalia et al. 2017 ) but the principle has been established of reporting back to families if a clinically important and actionable IF is found for the benefit of the family as a whole (not only the child). Previously known disorders within a family are of course excluded from this, and disclosure of these additional findings to the parents must have been raised with them as part of the process of ‘information and consent’, so that they should have agreed to this possibility in advance. This respects the (contested) ‘right not to know’. However, the potential social costs of this ‘right’—in terms of either failing to respect considered personal decisions or denying their relatives access to potentially important health care—are clearly overwhelming. The case for exempting information of clinical utility generated about children from the ‘right not to know’ is very strong. A parental decision not to consider information of likely practical clinical benefit to their child, or perhaps to their parents or siblings, could hardly have been guided by the overall best interests of the others involved.

It will also be necessary to address two other areas in the consent process: (i) which results should be returned to the patient (or parents)? and (ii) when or how often should findings be reassessed? The question of what results to return has generated a lot of debate in the clinical and bioethics literatures. Research is simpler—the obligations on researchers are less complex—if only aggregated results are given back to participants (Beskow et al. 2012 ). Indeed, that was all that would have been sensible when GWAS results were at issue. With sequencing results (exome or WGS), however, the findings of each individual are of much greater potential significance. This alters the situation dramatically (Tabor et al. 2011 ; Kaye et al. 2014 ). The PHG Foundation (Hall et al. 2013 ) and Knoppers et al. ( 2014 , 2015 ) have proposed helpful frameworks for those generating sequence information in the research setting.

In clinical practice, the process of consent for paediatric genome-wide investigation has been studied in several settings. Lessons have been learned about the minimum information that should be provided and discussed with parents (Burke and Clarke 2016 ). In a research setting in Toronto, the attitudes of the parents of children having genomic investigations for the purpose of a diagnostic assessment have been assessed and are of real interest (Anderson et al. 2017 ). The participants, a small and somewhat atypical subset of those whose children were being investigated in the hospital’s Genome Clinic, described their sense of having difficult and even distressing findings ‘inflicted’ on them but feeling unable to opt out of receiving these additional//incidental findings because of their sense of a burdensome duty to know all the potential problems that might face their child (an ‘inflicted ought’). In an insightful commentary, Newson reminds us that the parents’ response to IFs is specific to the setting of this particular clinic, where at least some parents felt that the unwanted information had indeed been inflicted upon them, when they were at best ambivalent about receiving it (Newson 2017 ). This arose out of the requirement to agree to receive IFs as a condition of accessing the diagnostic application of the WGS.

In relation to consent for genome investigations in children, the decisions will usually be made by their parents ‘in the child’s best interests’ (Ross, 2013 ). In the UK, there is no arbitrary line drawn at a specific age, before which it must be the parent who gives consent for medical investigations and treatments. Where a child of less than 16 years has some capacity to be involved in the discussion, this will be encouraged and they may give their views. The patient may be able to assent to medical procedures but the parent(s) will usually take responsibility for formally providing consent. Even young children can make important contributions to the decisions made about them, and research into consent to elective surgery in childhood has demonstrated this with great effect (Alderson 1993 ). Between 16 and 18 years, either the child or the parents can give consent to treatment. From the age of 18 years, a patient will be presumed to have the capacity to give consent unless there are grounds for considering that s/he may lack capacity, in which case their capacity has to be assessed. As mentioned earlier in the context of predictive testing, young people of less than 16 years can give consent, and in limited circumstances even refuse consent, as long as they can demonstrate adequate maturity of judgement. The essential test for those of less than 16 years is ‘Gillick competence’, as established by a court ruling under the Common Law of England & Wales, in which the test is whether the child has achieved, ‘sufficient understanding and intelligence to understand fully what is proposed’.

For adults, capacity is framed as the ability, with support if necessary, ‘to understand the information they are given; retain that information for long enough to make a decision; weigh up the information; and communicate their decision’. It is important to appreciate that capacity is not a single attribute but varies, being specific to the particular question at issue, and even fluctuating with the same person over time.

The question of when (and how often) to reinterpret any VUS results has also been widely discussed. With the rapidly increasing numbers of VUSs and the accumulation of information and experience about variants, the status of VUSs can change, with their status as pathogenic or benign usually becoming clearer. That, however, raises the question of how genetic services could periodically re-evaluate all the VUSs they have identified, and then how they could contact the patients whose variants had changed in a clinically meaningful fashion. What obligation has a service to take on this challenge? How can the costs of this additional service be met, when these are likely to escalate exponentially for a decade or more before any sort of stability of variant interpretation is reached? These issues remain challenging although constructive work to find solutions is in process. One difficulty with some proposals is that they rely on a pro-active engagement with the patients and families involved (Pyeritz 2011 ; Dheensa et al. 2017 ). This would certainly mean that only some of those who could benefit from ‘recontact’ will do so, is highly likely to be socially inequitable, and could well drive an increase in such inequity (Tudor Hart 1971 ).

What about investigations whose results are of little or no validity or applicability? The giving of information of no clinical utility, such as GWAS-based risk modifications, would not be regarded as equivalent to the disclosure of IFs. However, such results may be of interest in one way: if an increased risk of disease is assigned to the patient based on GWAS-related risks, does it lead to the modification of behaviour that would be recommended in the light of that genetic susceptibility? This type of personally assessed disease risk does not seem to lead to the ‘appropriate’ behaviour change: it does not tap into a strong, additional motivator to encourage compliance (Marteau et al. 2010 ; McBride et al. 2010 ). In any case, such ‘advice’ is consistent neither over time, as more or different SNPs are utilised, nor between test providers. What is of much more ethical concern is the inequity of access to genome-based and DNA sequencing-based investigations of proven value, which comes down to a question of the political system within which health care is being provided (McClellan et al. 2013 ).

An overview of the issues raised by the ‘genomic approach’ to clinical genetics is given by Clarke ( 2014 ), following a broader assessment of the science as well as the clinical implications (Clarke et al. 2012 ). A very helpful ethics-based framework for assessing the applicability of genome-based investigations in paediatric practice has been formulated by McCullough et al. ( 2015 ). A likewise constructive suggestion from Newson and colleagues is that genomic investigations should be approached in professional training and in discussions between patients and professionals as an exercise of uncertainty management: the ethos of genomics is uncertainty and this should be acknowledged and embraced (Newson et al. 2016 ).

Newborn metabolic screening and whole genome sequencing in the newborn and the foetus

As with the shift from focused to genome-wide investigations in laboratory genetics, biochemical screening of newborn infants for metabolic disorders has changed fundamentally with the introduction of tandem mass spectrometry (TMS). This has greatly increased the capacity of screening in terms of the volume of samples analysed in one laboratory and also the scope of what is screened for: TMS is able to identify a much wider range of different molecules than the previous, chromatography-based methods. This technological shift has greatly expanded the number of metabolic disorders that can be detected by screening.

This in turn makes it possible to identify newborns affected by (or destined to become affected by) disorders that do not fulfil the conventional, 1968 WHO criteria for screening devised by Wilson and Jungner. The key mismatch between these criteria and the new possibilities is that early diagnosis often fails to lead to improved outcomes for the infant because there is no effective treatment.

If the inclusion of a disorder in a newborn screening programme fails to meet the established criteria for screening, why might it nevertheless be a reasonable course of action? One response is that the parents become aware of the child’s diagnosis sooner than if they waited for the child to present with symptoms. They avoid the extended diagnostic process that might otherwise result, which can be an ordeal reminiscent of the Odyssey. They are spared this process, reported vividly in the case of Duchenne muscular dystrophy (Firth et al. 1983 ). The average time lag between the family seeking medical advice and achieving a diagnosis was 23 months in one study; we have no reason to think it has improved. Late diagnosis is a serious problem in many disorders (e.g. Bouwman et al. 2013 ).

The other major factor is that this earlier diagnosis enables parents to take account of the diagnosis in their plans for (further) children. They may then be able to make future reproductive decisions in the light of their risk of having affected children (Bombard et al. 2009 ). Newborn screening for DMD has shown that (most) families appreciate the early diagnosis, if they have made an active decision for the screen for DMD to be included in their infant’s newborn blood spot test (Parsons et al. 2002 ). This provides both the potential benefits: avoiding the distressing diagnostic Odyssey and enabling informed reproductive decisions. Will this also be true for other untreatable genetic disorders?

The question of whether to require specific and distinct consent for disorders like DMD or SMA (Swoboda 2010 ), for which treatments are still experimental and not yet routinely available, is important (Ross 2006 ; Ross and Clarke 2017 ). It would require a two-tier consent process for newborn screening, with the traditional screen being heavily recommended for the direct benefit of the infant and the second-level, opt-in, tests being for broader benefits to the family. It may be helpful to ‘tweak’ the information-and-consent process to emphasise the distinction between the two types of test (Parsons et al. 2000 ), thereby improving the quality of the consent and the satisfaction of patients and professionals. It may further be argued that a different set of criteria should be developed to help think through which disorders could legitimately be included in extended newborn screening (Petros 2012 ).

A further concern about the extended panel approach to newborn screening is that it identifies as patients many who are far from developing symptoms, some of whom will never become unwell (e.g. half of those with medium chain acyl coA dehydrogenase deficiency, MCADD). This leads to a new category of the ‘patient-in-waiting’ (Timmermans and Buchbinder 2010 ). There may be additional categories of patients where even early treatment for the affected child will not lead to a ‘cure’ but to a long and difficult illness; a family’s choices in future pregnancies may need to be based on a realistic assessment of treatment outcomes, which may not be available until long-term research programmes have reported their results. It should also be acknowledged that the natural history of some of these biochemical disorders has actually been clarified by newborn screening, which is an unbiased method of case ascertainment. The decision to include some disorders in screening may not always have met the Wilson and Jungner criterion of an ‘adequate’ knowledge of the natural history (Timmermans and Buchbinder 2012 ) but has led to an improved understanding of the penetrance of the condition.

Newborn screening may lead to the identification of other patients in the family, in addition to the diagnosed infant, perhaps an older sibling or a parent. A health care system that identifies such cases but fails to support the families in meeting the needs of these additional patients is clearly failing to discharge its responsibilities (Buchbinder and Timmermans 2011 ).

Another category of newborn screening which has been introduced, often for a genetic condition, is screening for impaired hearing. This gives access to cochlear implantation for those with severe deafness. In those for whom cochlear implants do not permit highly effective communication, the use of sign language may have been preferable. The decision to use sign language rather than cochlear implants may also be preferable to the community of The Deaf, which is formed around and through sign language. How will this community fare if so many potential recruits are now treated with implant surgery? How can we be confident that we have chosen the best treatment to be selected for each child?

Finally, in this section, we turn to consider WGS as a routine for newborn infants. Some useful diagnoses would emerge earlier than when waiting for a symptomatic presentation, although that is not always an advantage if it generates distress but there are no effective interventions to alter the natural history of the condition. This would enormously amplify the ‘patient-in-waiting’ category, i.e. the ‘worried well’ of asymptomatic individuals who have been found to carry sometimes pathogenic genetic variants in genes associated with late onset or incomplete penetrance. This would apply, for example, to those carrying a variant in a gene known to make a contribution to the risk of a cardiac dysrhythmia or cardiomyopathy. VUSs may also be found in such genes. The difficulties of VUS interpretation would loom large and the handling of information about adult-onset disorders and carrier status of largely reproductive significance. The disadvantages of WGS performed on newborn infants as a routine screen (i.e. unless used to achieve a diagnosis for a severe and complex condition) are serious (Howard et al. 2015 ).

It must not be forgotten that newborn screening by WGS could never replace the more traditional metabolic screen, both because the prognostic value of the metabolic screen has great power and because the commonest disorder identified, congenital hypothyroidism, often has no specific genetic cause (it most often arises from a failure of thyroid gland development and migration). Furthermore, this model of newborn WGS, both as a form of newborn screening and as a lifetime resource, depends on long-term data storage as already discussed, with the attendant difficulties and costs (Clarke 2014 ).

In the present state of our ability to interpret genome sequence data, the generation of WGS on newborns as a routine - rather than its infrequent use in attempting to achieve the diagnosis of rare and complex disorders, which would be its ‘obvious’ (clinically warranted) application—appears to be a wolf in sheep’s clothing: a massive research programme in the guise of a population screening programme. However, while it would be too vast to secure research funding, it could not possibly warrant funding as a population screening programme because it lacks any evidence of clinical utility and the opportunity costs would be immense. Given the extent of the unmet health needs of the population of the wealthiest country on the planet—the country from which the proposal for newborn screening by WGS emanate—we cannot imagine such a programme being justifiable in the foreseeable future, even if it were entirely benign in its consequences.

A recent proposal to go beyond neonatal to routine, unrestricted prenatal WGS by non-invasive prenatal testing (NIPWGS) appears to be a still less reasonable proposition, given that it addresses no specific clinical goal. The proponents appear oblivious to the confusion, anxiety and distress that would inevitably accompany its implementation (Chen and Wasserman 2017 ). While it is suggested that there would be an intensive programme of parental education about genome-based testing and an opportunity for moral reflection, there seems to be little awareness of the challenges this would entail at multiple levels. Furthermore, it is difficult to see what is driving the programme, or what it is hoping to achieve, apart from an increase in genome corporation profits at the expense of much potential family misery.

This proposal was accompanied in the same journal issue by a set of commentaries, some of which elaborated on the problems likely to be caused if this rash proposal were implemented, while others were supportive. The argument of Chen and Wasserman that they are not drawing arbitrary, ‘eugenic’ lines between foetuses, but instead empowering parents to make their own decisions, seems wilfully to ignore that this reproductive empowerment is taking the shape of consumerist eugenics, open to manipulation by Big Data Corporations, health insurers and fashion, and fits neatly into the account of contemporary ‘backdoor’ eugenics by Duster ( 2003 ).

Kaposy ( 2017 ) reminds us that the offer of choice in a pregnancy will not always be experienced as liberating but can be burdensome and disturbing, and that there is no call or demand for prenatal NIPWGS in apparently healthy pregnacies. (While it may be very possible for marketing agencies to create a demand, this would most likely be achieved by a re-framing of NIPWGS as the virtuous expression of parental responsibility and/or an enviable privilege of wealth). Furthermore, Kaposy (correctly) sees values as constructed and context-responsive rather than monolithic and fixed, so that the parents’ values may be moulded by the process of seeking information through foetal WGS. We might add that the process is likely both to reflect and to reinforce an inappropriately strong form of genetic determinism.

Ravitsky et al. ( 2017 ) reject the claim that NIPWGS would increase parental autonomy. They point out that the prior risk of disease in a healthy foetus is low and this impacts on the interpretation of the findings. In addition, the quantity of data that could be imparted is massive and the demands on members of the public to make sense of this are utterly unrealistic. The apparent belief of Chen and Wasserman that efficient information transfer and a limited period of reflection will allow couples to march happily on with their lives suggests that we encounter very different people in our clinics from those they meet. We do not recognise in their fantasy the world that we inhabit. As Mazersky and Sankar ( 2017 ) explain, the making of decisions is in reality far removed from the rationalistic account of Chen and Wasserman. The direction of inference from genotype to phenotype is an entirely different sport from explanation of phenotype in terms of genotype, as indicated by Botkin et al. ( 2017 ), Chen and Wasserman seem grossly to exaggerate the capacity of the former. Botkin et al. also make the point that the troubling aspects of the selective termination of pregnancies on the basis of prenatal diagnosis are not removed but exacerbated by extending the range of ‘abnormalities’ considered to be acceptable grounds for the termination of an otherwise wanted pregnancy, especially when the genotype=>phenotype link may be insecure. We are not writing from an anti-abortion position fixed in principle but from a recognition of the human pain and suffering involved in the termination of wanted (and perhaps perfectly healthy) pregnancies.

Munthe ( 2017 ) makes the additional point that Chen and Wasserman ignore the distinction between the technology perhaps having some acceptable applications, while at the same time being completely unacceptable as a population screening programme.

Some of the supporters of NIPWGS focus attention away from the foetal decisions being made to the management of information about the future child (if s/he survives to term). There are two particular problems with this. First, the fact that NIPWGS would of necessity be offered during pregnancy means that it cannot be aimed primarily at post-natal interventions for the benefit of the child, as WGS would be simpler and cheaper in the neonatal period or later in childhood. The proposal of Chen & Wasserman must therefore be intended primarily to lead to decisions about terminating apparently healthy pregnancies; otherwise, why not wait? The other possible explanation - aiming at treatments applicable in utero—is still premature even for most of the malformations that may be amenable to maternal-foetal surgery; treatments in utero for genetic disorders are still more remote. Secondly, the argument that it should be up to parents to decide whether to generate ‘difficult’ predictive information about neurodegenerative disease in their children (Rhodes 2017 ) ignores the fact that most adults at high risk of such conditions choose not to find out, and that this is for very good reasons given the realities of human psychology and family relationships. (They may not be pure, autonomous Kantians but can still be good people!)

Ethical issues in our multi-cultural UK society

How does one address ‘culture’ or ‘cultural difference’ in genetic counselling? Like Odysseus, one has to steer one’s craft between two waiting perils, the Scylla of cultural stereotypes and false homogenisation on the one side and the Charybdis of vain attempts to be blind to cultural difference on the other. Avoiding both errors is important and requires insight. This is a matter for training in counselling skills, personal development and supervision rather than chapters on ethics; we do not have space to treat these aspects further here. But some topics can—must—be broached.

One of the settings that most often raises ‘culture’ as an issue in the context of genetic counselling in Britain is consanguinity. What does the professional need to be aware of when addressing this? First, that the topic is, or may be, highly charged. Within the UK, many of those from communities in which consanguineous marriage is customary and often preferred will have been treated badly by health professionals, and perhaps others, when this topic has been raised in the past. If they have a child with autosomal recessive disease, they may have been blamed for this, perhaps quite explicitly, and may even have been insulted. More likely, however, the blame will have been implied, in the tone of voice perhaps. Such behaviour by professionals is unacceptable and unprofessional. While it may occur less frequently than in the past, we believe that it still happens all too often.

In the UK, at least, the sense of blame may have been compounded by statements from Members of Parliament and public health officials, who sometimes talk about consanguinity as a major social problem, especially among Muslims drawn from South Asian populations, and suggest that this community should change its long-standing practices. (They tend not to mention other communities that favour this practice). The consanguineous family may have felt that their community was under assault for this mark of difference, as well as for other religious and political factors. It should not be forgotten that the anti-consanguinity message of the health professionals can find a target elsewhere too. There are other consanguineous marriages in the UK, both within other ethnic minorities and also within the white British majority. Such statements can impact on their people’s (often already strong) sense of shame and guilt for any genetic problem in their family.

From the perspective of the community practising consanguineous marriage, it will usually be seen as entirely natural and a force for stability in the family, which enhances the status of women and which is not usually associated with genetic disorders in the community’s children. Professionals must not forget the positive features of consanguinity and should not use genetics as a route through which to voice cultural prejudice. Furthermore, it is unhelpful, unfeeling and unprofessional to blame parents for having children with a serious inherited condition. The incidence of recessive disorders is increased in consanguineous families and communities, and needs to be raised in a respectful manner in any discussion about reproductive risks, but the global context within which such marriages constitute a widespread social tradition must not be forgotten (Modell and Darr 2002 ; Hamamy et al. 2011 ; Bittles 2013 ).

Moving to consider lay and folk beliefs about the causes of malformation and other birth defects, a wide-ranging review of such beliefs by Kenen ( 1980 ) is enlightening. It shows the recurrent patterns in such beliefs from around the world and reminds us that these have been common in Europe until quite recently; superstition may still play a role in the thinking of people from many parts of the world. The Urdu-speaking community from South Asia resident in UK has a wide range of beliefs about birth defects and other paediatric conditions that fit into Kenen’s scheme, and a range of attitudes to the use of genetic investigations and, in pregnancies, to the use of antenatal screening that might lead to the offer of a termination of pregnancy (Shaw and Hurst 2008 ; Bryant et al. 2011 ; Shaw 2012 ). It is most important to remember that ‘religion’, whether Christianity, Islam or any other, is not monolithic and homogeneous but diverse and often negotiable. Professionals must make no assumptions about how a patient thinks or how they will respond in a given circumstance simply because they come from a specific community or religious group. Many of the attitudes and practices in minority ethnic groups are shared with those from the majority white British ethnic group.

Particular solutions have sometimes been devised to meet the needs of specific communities, whom one might expect to be reluctant to use some aspects of genetic technology. Some Orthodox Jewish communities use an anonymous programme of carrier screening with the results disclosed to the matchmaker (not to the couple) indicating that a particular couple would not make a good match (i.e. both are carriers of the same autosomal recessive disorder) (Raz and Vizner 2008 ). This approach may lead western clinicians to feel horror at the loss of autonomy this entails but, in the context, it may at least allow such communities to begin to engage with this aspect of modern health care. However, it is important for the ‘western’ genetic counsellor (i.e. from a western family and community background) to appreciate that many from ethnic minority groups will have their own views that are shaped by their experience of disease in the family and community as much as by the beliefs and values of their religion (Atkin et al. 2008 ). Make no assumptions!

When consanguinity is preferred within a community, it is possible to maintain this custom but reduce the risks of serious genetic disorders. One effective approach is to marry more distant cousins. Another approach is to make carrier screening available, either before marriage or in the antenatal clinic. A number of countries in North Africa and the Persian//Arab Gulf are using next generation sequencing to define the recessive gene mutations present in their populations, often unique to small population isolates or endogamous groups within a country. Schemes differ in whether each individual tested is given their own results and whether the results influence decisions about marriage or decisions about pregnancy, prenatal diagnosis and children. They also differ in whether they are mandatory or optional, and the extent of the pressure to comply with screening.

Several countries have developed a range of compulsory interventions with the stated goal of reducing the birth incidence of genetic disorders. One approach has been to carry out tests for carrier state of genetic disorders, sometimes of specific disorders such as beta-thalassaemia, either before marriage or conception or in the antenatal clinic. A different approach in China entails a (fairly) set and structured approach to assessment of a married couple. The relevant law has changed recently but before that the potential parents were judged on criteria familiar from Nazi Eugenics, with an interview and examination aimed to support the doctor making decisions about the desirability of two persons having children together (Hesketh 2003 ).

Infertility and childlessness are problematic within all societies, and the form taken by this problem will of course be shaped by the culture and its expectations. In the lives of South Asians who have settled in UK, there may be a tug between the influence of tradition and the influence of the individuals’ new experiences and acculturation into local life (Hampshire et al. 2012 ). Secrecy around infertility and its investigation and treatment is common in many communities. Infertility seems to be particularly wounding to the male ego, perhaps from some conflation of infertility and impotence. Adoption can also be a very difficult idea to accept within some communities, so that neither the admission of infertility nor adoption as a constructive ‘remedy’ may be recognised as an available option for the infertile couple (Bharadwaj 2003 , a study set in India). This of course increases the pressure on couples to use infertility services—but in secret.

Another potential conflict of perspectives arises in relation to the preference for sons that is stronger within some communities than others (although it was very powerful in UK until not so long ago). The sex ratio at birth is heavily distorted in some Indian states and in parts of China, with the sex ratio at birth being around 900 females per 1000 male births across India and below 800 females per 1000 males in some states (Madan and Breuning 2014 ). A comparable distortion of the sex ratio is also found within China and can be seen in some western countries with communities of South Asian origin when there is no legislation to restrict foetal sex selection. The long-term social consequences of such ‘unbalanced’ sex ratios are unpredictable and may change over time but will surely generate major social problems. This imbalance of the sexes results in both India and China from the widespread (although illegal) use of medical technology to determine foetal sex and terminate the female foetuses, and is reinforced in India by the (also illegal) system of dowry payments. The same would doubtless have happened in UK in previous centuries if the technology had been available (i.e. if witchcraft, magic or prayer had been more effective). In UK, where foetal sex selection is not permitted, professionals can become concerned when the talk or behaviour of patients in clinic suggests that the family may be motivated to choose a ‘social’ termination of pregnancy after foetal sexing performed as part of prenatal diagnosis for other reasons. Foetal sex selection has been shown to occur in some South Asian groups resident in Canada (Urquia et al. 2016 ).

Our western and feminist response to this scenario is horror at the systematic devaluation of women that this implies. Where women from communities that practice foetal sex selection support this approach, this adds to the horror as it supports the sense that life is hard for the women in these communities. A contrasting feminist response has also been proposed, reflecting a position of support for women’s decisions even when a rejection of sex selection and strong support for women’s position in society would seem more constructive (Moazam ( 2004 ). The sense that women’s social status needs support in some of these societies is reinforced by a sad tale of stigmatisation within UK of a woman of Bangladeshi origin who was affected by neurofibromatosis type 1 (Rozario 2007 ). There are also unsettling parallels between foetal sex selection and the low social status of women in some communities, on the one hand, and antenatal screening ‘for’ Down syndrome and the social status of affected individuals in many western countries on the other (Alderson 2001 ).

Decisions about what disorders are sufficiently serious to warrant a termination of pregnancy have a cross-cultural dimension. However, it would clearly be difficult to disentangle the many factors that may contribute to this. Thus the stage of economic development within a society, confidence in support from the family, community or state, intrinsic beliefs about what makes life worthwhile and the moral status of the embryo are all important but unquantifiable and often negotiable factors. One example is the striking diversity of views about termination of pregnancy for deafness, with the lack of confidence of support from the community in raising a child with deafness given as one factor that may contribute to a decision to terminate in some countries (Nahar et al. 2013 ). For a child born into a Deaf family, of course, rearing a Deaf child may be the preferred option and sign language will often be chosen over cochlear implantation. Personal experiences and the degree of acculturation are likely also to be relevant factors.

We present this whistle-stop tour through ‘the ethics of genetics in medicine’ as an educational resource. It has evolved since the turn of the century and will doubtless continue to change and develop as the issues that arise in our clinical practice also change. We have only been able to cover some aspects of this broad topic but we hope that the general approach adopted here proves useful and that it can be adapted to the particular circumstances in which you are working.

Acknowledgements

This chapter has been based largely on the 3–4 day residential course on ‘Ethics in Genetic Counselling’ held annually since 2001 in Neuadd Gregynog (Gregynog Hall) in Wales. This course was established for the students on the Cardiff University MSc Course in Genetic Counselling but each year a number of others have also attended. The course has changed over the years, being shaped by many factors including input from colleagues and students too numerous to list here. However, AC would like to thank three colleagues in particular who have supported the course in different ways and at different times: Drs Tara Clancy, Ainsley Newson, and Heather Strange. He would also like to thank the students who have participated over the years and the many other colleagues who have supported him and helped to shape the course.

  • Ablon J. Little people in America: the social dimensions of dwarfism. New York: Praeger; 1984. [ Google Scholar ]
  • Ablon J. Social stigma in neurofibromatosis 1. Chapter 22. In: Upadhyaya M, Cooper DN, editors. Neurofibromatosis type 1: molecular and cellular biology. Berlin: Springer-Verlag; 2013. pp. 673–682. [ Google Scholar ]
  • Adelson BM. Lives of dwarfs; their journey from public curiosity toward social liberation. Piscataway: Rutgers University Press; 2005. [ Google Scholar ]
  • Alderson P. Children's consent to surgery. Milton Keynes: Open University Press; 1993. [ Google Scholar ]
  • Alderson P. Down’s syndrome: cost, quality and value of life. Soc Sci Med. 2001; 53 :627–638. [ PubMed ] [ Google Scholar ]
  • Aldred MJ, Crawford PJM, Savarirayan R, Savulescu J. It's only teeth – are there limits to genetic testing? Clin Genet. 2003; 63 :333–339. [ PubMed ] [ Google Scholar ]
  • Anderson JA, Meyn MS, Shuman C, Zlotnik Shaul R, Mantella LE, Szego MJ, Bowdin S, Monfared N, Hayeems RZ. Parents perspectives on whole genome sequencing for their children: qualified enthusiasm? J Med Ethics. 2017; 43 :535–539. [ PubMed ] [ Google Scholar ]
  • Arribas-Ayllon M, Sarangi S, Clarke A. Professional ambivalence: accounts of ethical practice in childhood genetic testing. J Genet Couns. 2009; 18 :173–184. [ PubMed ] [ Google Scholar ]
  • Atkin K, Ahmed S, Hewison J, Green JM. Decision-making and antenatal screening for sickle cell and thalassaemia disorders. Curr Sociol. 2008; 56 :77–98. [ Google Scholar ]
  • Baig SS, Strong M, Rosser E, Taverner NV, Glew R, Miedzybrodzka Z, Clarke A, Craufurd D, UK Huntington’s Disease Prediction Consortium. Quarrell OW. 22 years of predictive testing for Huntington's disease: The Experience of the UK Huntington's Prediction Consortium. Eur J Hum Genet. 2016; 24 :1396–1402. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Barter B, Hastings RP, Williams R, Huws JC (2017) Perceptions and discourses relating to genetic testing: interviews with people with Down syndrome. J Appl Res Intellect Disabil. 30:395–406 [ PubMed ]
  • Beauchamp TL, Childress JF. Principles of biomedical ethics. 7. New York: Oxford University Press USA; 2013. [ Google Scholar ]
  • Berg JS, Adams M, Nassar N, et al. An informatics approach to analyzing the incidentalome. Genet in Med. 2013; 15 :36–44. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Beskow LM, Burke W, Fullerton SM, Sharp RR (2012) Offering aggregate results to participants in genomic research: opportunities and challenges. Genet Med 14:490–496 [ PMC free article ] [ PubMed ]
  • Bharadwaj A. Why adoption is not an option in India: the visibility of infertility, the secrecy of donor insemination and other cultural complexities. Soc Sci Med. 2003; 56 :1867–1880. [ PubMed ] [ Google Scholar ]
  • Binedell J, Soldan JR, Scourfield J, Harper PS. Huntington disease predictive testing: the case for an assessment approach to requests from adolescents. J Med Genet. 1996; 33 :912–918. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bittles AH. Consanguineous marriage and congenital anomalies. Lancet. 2013; 382 :1316–1317. [ PubMed ] [ Google Scholar ]
  • Blackburn S (2001) Being good. Oxford University Press
  • Boardman F. Knowledge is power? The role of experiential knowledge in genetically ‘risky’ reproductive decisions. Sociol Health Illn. 2014; 36 :137–150. [ PubMed ] [ Google Scholar ]
  • Boardman F. The expressivist objection to prenatal testing: the experiences of families living with genetic disease. Soc Sci Med. 2014; 107 :18–25. [ PubMed ] [ Google Scholar ]
  • Bombard Y, Miller FA, Hayeems RZ, Avard D, Knoppers BM, Cornel MC, Borry P. The expansion of newborn screening: reproductive benefit an appropriate pursuit? Nat Rev Genet. 2009; 10 :686–687. [ PubMed ] [ Google Scholar ]
  • Bombard Y, Palin J, Friedman JM, Veenstra G, Creighton S, Bottorff JL, Hayden MR, The Canadian Respond-HD Collaborative Research Group Beyond the patient: the broader impact of genetic discrimination among individuals at risk of Huntington disease. Am J Med Genet B. 2012; 159B :217–226. [ PubMed ] [ Google Scholar ]
  • Boone PM, Soens ZT, Campbell IM, et al. Incidental copy-number variants identified by routine genome testing in a clinical population. Genet Med. 2013; 15 :45–54. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Borry P, Evers-Kiebooms G, Cornel MC, Clarke A, Dierickx K, Public and Professional Policy Committee (PPPC) of the European Society of Human Genetics (ESHG) Genetic testing in asymptomatic minors: background considerations towards ESHG recommendations. Eur J Hum Genet. 2009; 17 (6):711–719. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Boslett GT. Parental procreative obligation and the categorisation of disease: the case of cystic fibrosis. J Med Ethics. 2011; 37 :280–284. [ PubMed ] [ Google Scholar ]
  • Botkin JR, Belmont JW, Berg JS, Berkman BE, Bombard Y, Holm IA, Levy HP, Ormond KE, Saal HM, Spinner NB, Wilfond BS, McInerney JD. American Society of Human Genetics Position Statement. Points to Consider: Ethical, Legal and Psychosocial Implications of Genetic Testing in Children and Adolescents. AJHG. 2015; 97 :6–21. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Botkin JR, Francis LP, Rose NC. Concerns about justification for fetal genome sequencing. Am J Bioeth. 2017; 17 (1):23–25. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bouwman MG, de Ru MH, Linthorst GE, Hollak CEM, Wijburg FA, van Zwieten MCB. Fabry patient experiences with the timing of diagnosis relevant for the discussion on newborn screening. Mol Genet Metab. 2013; 109 :201–207. [ PubMed ] [ Google Scholar ]
  • Bridgens R. Disability and being `normal': a response to McLaughlin and Goodley. Sociology. 2009; 43 (4):753–761. [ Google Scholar ]
  • British Society for Human Genetics (2010) Genetic Testing of Children: Report of a working party of the British Society for Human Genetics (available from the website of the British Society of Genetic Medicine: http://www.bsgm.org.uk/media/678741/gtoc_booklet_final_new.pdf )
  • Browner CH, Preloran HM. Neurogenetic diagnoses: the power of hope and the limits of today's medicine: the power of hope and the limits of today’s medicine. Abingdon: Routledge; 2010. [ Google Scholar ]
  • Bryant L, Ahmed S, Ahmed M, Jafri H, Raashid Y. ‘All is done by Allah’. Understandings of Down syndrome and prenatal testing in Pakistan. Soc Sci Med. 2011; 72 :1393–1399. [ PubMed ] [ Google Scholar ]
  • Buchbinder M, Timmermans S. Newborn screening and maternal diagnosis: rethinking family benefit. Soc Sci Med. 2011; 73 (7):1014–1018. [ PubMed ] [ Google Scholar ]
  • Burke K, Clarke A. The challenge of consent in clinical genome-wide testing. Arch Dis Child. 2016; 101 :1048–1052. [ PubMed ] [ Google Scholar ]
  • Carel H. Phenomenology of Illness. Oxford: Oxford University Press; 2016. [ Google Scholar ]
  • Chadwick R, Capps B, Chalmers D et al (2013) Imagined futures: capturing the benefits of genome sequencing for society. Human Genome Organisation [ PMC free article ] [ PubMed ]
  • Chen SC, Wasserman DT. A framework for unrestricted prenatal whole-genome sequencing: respecting and enhancing the autonomy of prospective parents. Am J Bioeth. 2017; 17 (1):3–18. [ PubMed ] [ Google Scholar ]
  • Clarke A. Is non-directive genetic counselling possible? Lancet. 1991; 338 :998–1001. [ PubMed ] [ Google Scholar ]
  • Clarke A. The process of genetic counselling: beyond nondirectiveness. Chapter 13. In: Harper P, Clarke AJ, editors. Genetics, society and clinical practice. Oxford: Bios Scientific Publishers; 1997. pp. 179–200. [ Google Scholar ]
  • Clarke A. 2013. Stigma, self-esteem and reproduction: talking with men about life with hypohidrotic ectodermal dysplasia. Sociology 47: 975–993
  • Clarke AJ. Managing the ethical challenges of next generation sequencing in genomic medicine. Br Med Bull. 2014; 111 (1):17–30. [ PubMed ] [ Google Scholar ]
  • Clarke A. Anticipated stigma and blameless guilt: mothers' evaluation of life with the sex-linked disorder, hypohidrotic ectodermal dysplasia (hed) Soc Si Med. 2016; 158 :141–148. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Clarke A, et al. Genetic professionals’ reports of non-disclosure of genetic risk information within families. Eur J Hum Genet. 2005; 13 :556–562. [ PubMed ] [ Google Scholar ]
  • Clarke A, Cooper DN, Krawczak M, Tyler-Smith C, Wallace H, Wilkie A, Raymond L, Chadwick R, Craddock N, John R, Gallacher J, Chiano M. ‘Sifting the significance from the data' - the impact of high-throughput genomic technologies on human genetics and health care. Human Genomics. 2012; 6 :11. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Corrigan O. Empty ethics: the problem of informed consent. Sociol Health Illn. 2003; 25 :768–792. [ PubMed ] [ Google Scholar ]
  • d’Agincourt-Canning L. Experiences of genetic risk: disclosure and the gendering of responsibility. Bioethics. 2001; 15 :231–247. [ PubMed ] [ Google Scholar ]
  • Deans Z, Clarke A, Newson A. For your interest? The ethical acceptability of using non-invasive prenatal testing to test “purely for information” Bioethics. 2014; 29 :19–25. [ PubMed ] [ Google Scholar ]
  • Dewey FE, Grove ME, Pan C, Goldstein BA, Bernstein JA, Chaib H, Merker JD, Goldfeder RL, Enns GM, David SP, Pakdaman N, Ormond KE, Caleshu C, Kingham K, Klein TE, Whirl-Carrillo M, Sakamoto K, Wheeler MT, Butte AJ, Ford JM, Boxer L, Ioannidis JPA, Yeung AC, Altman RB, Assimes TL, Snyder M, Ashley EA, Quertermous T. Clinical interpretation and implications of whole-genome sequencing. JAMA. 2014; 311 (10):1035–1044. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Dheensa S, Carrieri D, Kelly S, Clarke A, Doheny S, Turnpenny P, Lucassen A (2017) A 'joint venture' model of recontacting in clinical genomics: challenges for responsible implementation. Eur J Med Genet. 60(7):403–409 [ PubMed ]
  • Downing C. Negotiating responsibility: case studies of reproductive decision-making and prenatal genetic testing in families facing HD. J Genet Couns. 2005; 14 :219–234. [ PubMed ] [ Google Scholar ]
  • Duncan RE, Gillam L, Savulescu J et al 2008. “You’re one of us now”: young people describe their experiences of predictive genetic testing for HD and FAP. Am J Med Genet 148C; 47–55 [ PubMed ]
  • Duster T. Backdoor to Eugenics (2nd Edition). Abingdon: Routledge; 2003. [ Google Scholar ]
  • Edwards SD. 2006. Disability: philosophical issues. Wiley Online Encyclopedia of the Life Sciences (eLS). 10.1002/9780470015902.a0006178
  • Edwards JG, Feldman G, Goldberg J, Gregg AR, Norton ME, Rose NC, Schneider A, Stoll K, Wapner R, Watson MS. Expanded carrier screening in reproductive medicine—points to consider. A Joint Statement of the ACMG, ACOG, NSGC, Perinatal Quality Foundation and Society for Maternal-Fetal Medicine. Obstet Gynecol. 2015; 125 (3):653–662. [ PubMed ] [ Google Scholar ]
  • Elwyn G, Gray J, Clarke A. Shared decision making and non-directiveness in genetic counselling. J Med Genet. 2000; 37 :135–138. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Erwin C, Williams JK, Juhl AR, Mengeling M, Mills JA, Bombard Y, Hayden MR, Quaid K, Shoulson I, Taylor S, Paulsen JS, the IRESPOND-HD Investigators of the Huntington Study Group Perception, experience, and response to genetic discrimination in Huntington disease: the International RESPOND-HD Study. Am J Med Genet Part B. 2010; 153B :1081–1093. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Etchegary H. Discovering the family history of Huntington disease (HD) J Genetic Couns. 2006; 15 :105–117. [ PubMed ] [ Google Scholar ]
  • Fanos J, Johnson J. Barriers to carrier testing for adult cystic fibrosis sibs: the importance of not knowing. Am J Med Genet. 1995; 1995 (59):85–91. [ PubMed ] [ Google Scholar ]
  • Fanos J, Johnson J. Perception of carrier status by cystic fibrosis siblings. Am J Hum Genet. 1995; 57 :431–438. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Featherstone K, Bharadwaj A, Clarke A, Atkinson P. Risky relations. Family and kinship in the era of new genetics. Oxford: Berg Publishers; 2006. [ Google Scholar ]
  • Feder EK, Karkazis K. What's in a name? The controversy over "Disorders of Sex Development". Hast Cent Rep. 2008; 38 (5):33–36. [ PubMed ] [ Google Scholar ]
  • Firth M, Gardner-Medwin D, Hosking G, Wilkinson E. 1983. Interviews with parents of boys suffering from Duchenne muscular dystrophy Dev Med Child Neurol 25(4):466–71 [ PubMed ]
  • Flew A (1975) Thinking about thinking. Fontana/Collins
  • Forrest Keenan K, McKee L, Miedzybrodzka Z. Help or hindrance: young people’s experiences of predictive testing for Huntington’s disease. Clin Genet. 2015; 87 :563–569. [ PubMed ] [ Google Scholar ]
  • Forrest K, Simpson SA, Wilson BJ, van Teijlingen ER, McKee L, Haites N, Matthews E. To tell or not to tell: barriers and facilitators in family communication about genetic risk. Clin Genet. 2003; 64 :317–326. [ PubMed ] [ Google Scholar ]
  • Foster C, et al. Juggling roles and expectations: dilemmas faced by women talking to relatives about cancer and genetic testing. Psychol Health. 2004; 19 :439–455. [ Google Scholar ]
  • Gaff C, Lynch E, Spencer L. Predictive testing of eighteen-year olds: counseling challenges. J Genet Couns. 2006; 15 :245–251. [ PubMed ] [ Google Scholar ]
  • García E, Timmermans DR, van Leeuwen E. Reconsidering prenatal screening: an empirical-ethical approach to understand moral dilemmas as a question of personal preferences. J Med Ethics. 2009; 35 :410–414. [ PubMed ] [ Google Scholar ]
  • Geelen E, van Hoyweghen I, Doevendans PA, Marcelis CLM, Horstman K. Constructing ‘Best Interests’: Genetic testing of children in families with hypertrophic cardiomyopathy. Am J Med Genet A. 2011; 155A (8):1930–1938. [ PubMed ] [ Google Scholar ]
  • Green RC, Berg JS, Grody WW, Kalia SS, Korf BR, Martin CL, McGuire A, Nussbaum RL, O’Daniel JM, Ormond KE, Rehm HL, Watson MS, Williams MS, Biesecker LG. ACMG Recommendations for Reporting of Incidental Findings in Clinical Exome and Genome Sequencing. Genet Med. 2013; 15 (7):565–574. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hall A, Hallowell N, Zimmern R (2013) Managing incidental and pertinent findings from WGS in the 100,000 Genomes Project. Public Health Genomics Foundation
  • Hallowell N, Foster C, Eeles R, Ardern-Jones A, Murday V, Watson M. Balancing autonomy and responsibility: the ethics of generating and disclosing genetic information. J Med Ethics. 2003; 29 :74–79. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hallowell N, Arden-Jones A, Eeles R, Foster C, Lucassen A, Moynihan C, Watson M. Guilt, blame and responsibility: men's understanding of their role in the transmission of BRCA1/2 mutations within the family. Sociol Health Illn. 2006; 28 (7):969–988. [ PubMed ] [ Google Scholar ]
  • Halverson CM, Ross LF (2012) Incidental findings of therapeutic misconception in biobank-based research. Genet Med. 14:611–615 [ PMC free article ] [ PubMed ]
  • Hamamy H, Antonarakis SE, Cavalli-Sforza LL, et al. Consanguineous marriages, pearls and perils: Geneva International Consanguinity Workshop Report. Genet in Med. 2011; 13 :841–847. [ PubMed ] [ Google Scholar ]
  • Hampshire KR, Blell MT, Simpson B. ‘Everybody is moving on’: Infertility, relationality and the aesthetics of family life among British-Pakistani Muslims. Soc Sci Med. 2012; 74 :1045–1052. [ PubMed ] [ Google Scholar ]
  • Harper PS. Naming of syndromes and unethical activities: the case of Hallervorden and Spatz. Lancet. 1996; 348 :1224–1225. [ PubMed ] [ Google Scholar ]
  • Hawkins AK, Creighton S, Hayden MR. When access is an issue: exploring barriers to predictive testing for Huntington disease in British Columbia, Canada. Eur J Hum Genet. 2013; 21 :148–153. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Helderman-van den Ende ATGM, Madan K, Breuning MH, et al. An urgent need for a change in policy revealed by a study on prenatal testing for Duchenne muscular dystrophy. EJHG. 2013; 21 :21–26. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Helgesson G. 2012. In defence of broad consent. Camb Q Healthc Ethics 21: 40–50 [ PubMed ]
  • Hendriks KSWH, Grosfeld FJM, van Tintelen JP, et al. Can parents adjust to the idea that their child is at risk of sudden death? Am J Med Genet. 2005; 138A :107–112. [ PubMed ] [ Google Scholar ]
  • Hercher L, Bruenner G. Living with a child at risk for psychotic illness. Am J Med Genet. 2008; 146A :2355–2360. [ PubMed ] [ Google Scholar ]
  • Hesketh T. Getting married in China: pass the medical first. BMJ 2003, 326: 277–9 [ PMC free article ] [ PubMed ]
  • Hippman C, Inglis A, Austin J. What is a “balanced” description? Insight from parents of individuals with Down syndrome. J Genet Couns. 2012; 21 :35–44. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Ho A. Relational autonomy or undue pressure? Family’s role in medical decision-making. Scand J Caring Sci. 2008; 22 :128–135. [ PubMed ] [ Google Scholar ]
  • Hodgson JM, Gillam LH, Sahhar MA, Metcalfe SA. "Testing times, challenging choices": an Australian study of prenatal genetic counseling. J Genet Couns. 2010; 19 :22–37. [ PubMed ] [ Google Scholar ]
  • Holm S. If you have said A, you must also say B: is this always true? Camb Q Healthc Ethics. 2004; 13 :179–184. [ PubMed ] [ Google Scholar ]
  • Holt K. What do we tell the children? Contrasting the disclosure choices of two HD families regarding risk status and predictive genetic testing. J Genet Couns. 2006; 5 :253. [ PubMed ] [ Google Scholar ]
  • Howard HC, Knoppers BM, Cornel MC, Clayton EW, Senecal K, Borry P. Whole-genome sequencing in newborn screening? A statement on the continued importance of targeted approaches in newborn screening programmes. Eur J Hum Genet. 2015; 23 (12):1593–1600. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Huggins M, Bloch M, Wiggins S, et al. Predictive testing for Huntington disease in Canada: adverse effects and unexpected results in those receiving a decreased risk. Am J Med Genet. 1992; 42 :508–515. [ PubMed ] [ Google Scholar ]
  • Human Genetics Commission . Inside information. London: Department of Health; 2002. [ Google Scholar ]
  • Human Genetics Commission . Increasing options, informing choice: a report on preconception genetic testing and screening. London: Dept of Health; 2011. [ Google Scholar ]
  • James CA et al. 2003. Perceptions of reproductive risk and carrier testing among adolescent sisters of males with chronic granulomatous disease. Am J Med Genet C Semin Med Genet 119C: 60–69 [ PubMed ]
  • Jansen L, Ross LF. The ethics of pre-adoption genetic testing. Am J Med Genet. 2001; 104 :214–228. [ PubMed ] [ Google Scholar ]
  • Jarvinen O, et al. Carrier testing of children for two X-linked diseases in a family based setting: a retrospective long term psychosocial evaluation. J Med Genet. 1999; 36 :615–620. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Jolly A, Parsons E, Clarke A. Identifying carriers of balanced chromosomal translocations: interviews with family members. In: Clarke A, editor. The Genetic Testing of Children. Oxford: Bios Scientific Publishers; 1998. pp. 61–90. [ Google Scholar ]
  • Jurgens J, Ling H, Hetrick K, Pugh E, Schiettecatte F, Doheny K, Hamosh A, Avramopoulos D, Valle D, Sobreira N. Assessment of incidental findings in 232 whole-exome sequences from the Baylor–Hopkins Center for Mendelian Genomics. Genet Med. 2015; 17 :782–788. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kaldjian LC. Communicating moral reasoning in medicine as an expression of respect for patients and integrity among professionals. Commun Med. 2013; 10 (2):177–183. [ PubMed ] [ Google Scholar ]
  • Kalia SS, Adelman K, Bale SJ, Chung WK, Eng C, Evans JP, Herman GE, Hufnagel SB, Klein TE, Korf BR, McKelvey KD, Ormond KE, Richards CS, Vlangos CN, Watson M, Martin CL, Miller DT, on behalf of the ACMG Secondary Findings Maintenance Working Group Recommendations for reporting of secondary findings in clinical exome and genome sequencing, 2017 update. Genet Med. 2017; 19 (2):249–255. [ PubMed ] [ Google Scholar ]
  • Kaposy C. Noninvasive prenatal whole-genome sequencing: a solution in search of a problem. Am J Bioeth. 2017; 17 (1):42–44. [ PubMed ] [ Google Scholar ]
  • Katz RB (1986) (and revised editons). The tentative pregnancy: amniocentesis and the sexual politics of motherhood. New York: Viking (and London: Pandora)
  • Kaye J, Hurles M, Griffin H, Grewal J, Bobrow M, Timpson N, Smee C, Bolton P, Durbin R, Dyke S, Fitzpatrick D, Kennedy K, Kent A, Muddyman D, Muntoni F, Raymond LF, Semple R, Spector T, UK10K Managing clinically significant findings in research: the UK10K example. Eur J Hum Genet. 2014; 22 :1100–1104. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Keenan KF, Simpson SA, Miedzybrodzka Z, Alexander DA, Semper J. How do partners find out about the risk of Huntington's disease in couple relationships? J Genet Couns. 2013; 22 :336–344. [ PubMed ] [ Google Scholar ]
  • Kelly SE. Choosing not to choose: reproductive responses of parents of children with genetic conditions or impairments. Sociol Health Illn. 2009; 31 :81–97. [ PubMed ] [ Google Scholar ]
  • Kenen RH. Negotiations, superstitions and the plight of individuals born with severe birth defects. Soc Sci Med. 1980; 14A :279–286. [ PubMed ] [ Google Scholar ]
  • Kessler S. Psychological aspects of genetic counseling. VII. Thoughts on directiveness. J Genet Couns. 1992; 1 :9–17. [ PubMed ] [ Google Scholar ]
  • Klitzman R, Thorne R, Williamson J, Chung W, Marder K. Decision-making about reproductive choices among individuals at risk of Huntington's disease. J Genet Couns. 2007; 16 :347–362. [ PubMed ] [ Google Scholar ]
  • Knoppers BM, Avard D, Sénécal K, Zawati MH, P3G International Paediatrics Platform Members Return of whole-genome sequencing results in paediatric research: a statement of the P3G international paediatrics platform. Eur J Hum Genet. 2014; 22 :3–5. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Knoppers BM, Zawati MH, Sénécal K. Return of genetic testing results in the era of whole-genome sequencing. Nat Rev Genet. 2015; 16 :553–559. [ PubMed ] [ Google Scholar ]
  • Konrad M. Narrating the new predictive genetics: ethics, ethnography and science. Cambridge: Cambridge University Press; 2005. [ Google Scholar ]
  • Landsman GH. Reconstructing motherhood in the age of “perfect” babies: mothers of infants and toddlers with disabilities. Signs. 1998; 24 (1):69–99. [ Google Scholar ]
  • Landsman G. Emplotting children’s lives: developmental delay vs. disability. Soc Sci Med. 2003; 56 :1947–1960. [ PubMed ] [ Google Scholar ]
  • Lantos JD, Matlock AM, Wendler D. Clinician integrity and limits to patient autonomy. JAMA. 2011; 305 (5):495–499. [ PubMed ] [ Google Scholar ]
  • Lippman A, Wilfond BS. Twice-told tales: stories about genetic disorders. Am J Hum Genet. 1992; 51 (4):936–937. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Lucassen A, Parker M. Revealing false paternity: some ethical considerations. Lancet. 2001; 357 :1033–1035. [ PubMed ] [ Google Scholar ]
  • Maat-Kievit A, Vegter-Van Der Vlis M, Zoeteweij M, Losekoot M, van Haeringen A, Roos RA. Predictive testing of 25 percent at-risk individuals for Huntington disease (1987–1997) Am J Med Genet. 1999; 88 (6):662–668. [ PubMed ] [ Google Scholar ]
  • MacLeod R, Tibben A, Frontali M, Evers-Kiebooms G, Joes A, Martinez-Descales A, Roos RA. Recommendations for the predictive genetic test in Huntington’s disease. Clin Genet. 2013; 83 (3):221–231. [ PubMed ] [ Google Scholar ]
  • MacLeod R, Beach A, Henriques S, Knopp J, Nelson K, Kerzin-Storrar L. Experiences of predictive testing in young people at risk of Huntington’s disease, familial cardiomyopathy or hereditary breast and ovarian cancer. Eur J Hum Genet. 2014; 22 (3):396–401. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Madan K, Breuning MH. Impact of prenatal technologies on the sex ratio in India: an overview. Genet Med. 2014; 16 (6):425–432. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Mand C, Gillam L, Duncan RE, Delatycki MB. “It was the missing piece”: adolescent experiences of predictive genetic testing for adult-onset conditions. Genet Med. 2013; 15 :643–649. [ PubMed ] [ Google Scholar ]
  • Manjoney DM, McKegnay FP. Individual and family coping with polycystic kidney disease: the harvest of denial. Int J Psychiatry Med. 1978; 9 :19–31. [ PubMed ] [ Google Scholar ]
  • Marteau TM, French DP, Griffin SJ, Prevost AT, Sutton S, Watkinson C, Attwood S, Hollands GJ. Effects of communicating DNA-based disease risk estimates on risk-reducing behaviours (review) Cochrane Libr. 2010; 10 :1–74. [ PubMed ] [ Google Scholar ]
  • May T, Strong KA, Khoury MJ, Evans JP. Can targeted genetic testing offer useful health information to adoptees? Genet Med. 2015; 17 :533–535. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Mazersky J, Sankar P. Mandating moral reflection? Am J Bioeth. 2017; 17 :32–34. [ PubMed ] [ Google Scholar ]
  • McAllister M, Wood A, Dunn G, Shiloh S and Todd C. 2011. The genetic counseling outcome scale: a new patient-reported outcome measure for clinical genetics services. Clin Genet 79: 413–424 [ PubMed ]
  • McBride CM, Koehly LM, Sanderson SC, Kaphingst KA. The behavioural response to personalised genetic risk information: will genetic risk profiles motivate individuals and families to choose more healthful behaviours? Annu Rev Public Health. 2010; 31 :89–103. [ PubMed ] [ Google Scholar ]
  • McClellan KA, Avard D, Simard J, Knoppers BM. Personalized medicine and access to health care: potential for inequitable access? Eur J Hum Genet. 2013; 21 :143–147. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • McConkie-Rosell A, Spiridigliozzi GA. “Family matters”: a conceptual framework for genetic testing in children. J Genet Couns. 2004; 13 :9–29. [ PubMed ] [ Google Scholar ]
  • McCullough LB, Brothers KB, Chung WK, Joffe S, Koenig BA, Wilfond B, Yu JH, on behalf of the Clinical Sequencing Exploratory Research (CSER) Consortium Pediatrics Working Group 2015 Professionally responsible disclosure of genomic sequencing results in pediatric practice. Pediatrics136(4): e974–82 [ PMC free article ] [ PubMed ]
  • McDougall R. Parental virtue: a new way of thinking about the morality of reproductive actions. Bioethics. 2007; 21 :181–189. [ PubMed ] [ Google Scholar ]
  • McLaughlin J, Goodley D, Clavering E, Fisher P (2008 Families raising disabled children. Enabling Care and Social Justice. Palgrave Macmillan
  • Mendes A, Sousa L, Sequeiros J, Clarke A. Discredited legacy: stigma and familial amyloid polyneuropathy in northwestern Portugal. Soc Sci Med. 2017; 182 :73e80. [ PubMed ] [ Google Scholar ]
  • Metcalfe A, Plumridge G, Coad J, Shanks A, Gill P. Parents’ and children’s communication about genetic risk: a qualitative study, learning from families’ experiences. Eur J Hum Genet. 2011; 19 :640–646. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Michie S. Predictive testing in children: paternalism or empiricism? In: Marteau T, Richards M, editors. The troubled helix: social and psychological implications of the new human genetics. Cambridge: Cambridge University Press; 1996. pp. 177–186. [ Google Scholar ]
  • Moazam F. Feminist discourse on sex screening and selective abortion of female fetuses. Bioethics. 2004; 18 :205–220. [ PubMed ] [ Google Scholar ]
  • Modell B, Darr A. Genetic counselling and customary consanguineous marriage. Nat Rev Genet. 2002; 3 :225–229. [ PubMed ] [ Google Scholar ]
  • Morris M, Tyler A, Harper PS. Adoption and genetic prediction for Huntington's disease. Lancet. 1988; 2 (8619):1069–1070. [ PubMed ] [ Google Scholar ]
  • Müller-Hill B. Murderous science. Oxford: Oxford University Press; 1988. [ Google Scholar ]
  • Munthe C. Permissibility or priority? Testing or screening? Essential distinctions in the ethics of prenatal testing. Am J Bioeth. 2017; 17 (1):30–32. [ PubMed ] [ Google Scholar ]
  • Nahar R, Puri RD, Saxena R, Verma IC. Do parental perceptions and motivations towards genetic testing and prenatal diagnosis for deafness vary in different cultures? Am J Med Genet A. 2013; 161A :76–81. [ PubMed ] [ Google Scholar ]
  • Newson AJ. 2017. Whole genome sequencing in children: ethics, choice and deliberation. J Med Ethics. 2017; 43 :540–542. [ PubMed ] [ Google Scholar ]
  • Newson AJ, Humphries SE (2005) Cascade testing in familial hypercholesterolaemia: how should family members be contacted? Eur J Hum Genet 2005(13):401–408 [ PubMed ]
  • Newson A, Leonard S. Childhood genetic testing for familial cancer: does adoption make a difference? Familial Cancer. 2010; 9 :37–42. [ PubMed ] [ Google Scholar ]
  • Newson AJ, Leonard SJ, Hall A, Gaff CL. Known unknowns: building an ethics of uncertainty into genomic medicine. BMC Med Genet. 2016; 9 :57. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nuffield Council on Bioethics . Non-invasive prenatal testing: ethical issues. London: Nuffield Council on Bioethics; 2017. [ Google Scholar ]
  • Parens E, Asch A. The disability rights critique of prenatal genetic testing: reflections and recommendations. In: Parens E, Asch A, editors. Prenatal testing and disability rights. Washington DC: Georgetown University Press; 2000. pp. 3–43. [ Google Scholar ]
  • Parker M. Ethical problems and genetics practice. Cambridge: Cambridge University Press; 2012. [ Google Scholar ]
  • Parsons EP, Clarke AJ, Hood K, Bradley DM. Feasibility of a change in service delivery: the case of optional newborn screening for Duchenne muscular dystrophy. Community Genet. 2000; 3 (1):17–23. [ Google Scholar ]
  • Parsons EP, Clarke AJ, Hood K, Lycett E, Bradley DM. Newborn screening for Duchenne muscular dystrophy: a psychosocial study. Arch Dis Child Fetal Neonatal Ed. 2002; 86 :F91–F95. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Peterson M. 2012. Disability advocacy and reproductive choice: engaging with the expressivist objection. J Genet Couns 21: 13–6 [ PubMed ]
  • Petros M. Revisting the Wilson-Jungner criteria: how can supplemental criteria guide public health in the era of genetic screening? Genet Med. 2012; 14 :129–134. [ PubMed ] [ Google Scholar ]
  • Pilnick A. What ‘most people’ do: exploring the ethical implications of genetic counselling. New Genet Soc. 2002; 21 :339–350. [ Google Scholar ]
  • Pyeritz RE. The coming explosion in genetic testing--is there a duty to recontact? N Engl J Med. 2011; 365 (15):1367–1369. [ PubMed ] [ Google Scholar ]
  • Rapp R. Testing women, testing the fetus. London: Routledge; 2000. [ Google Scholar ]
  • Ravitsky V, Rousseau F, Laberge A-M. Providing unrestricted access to prenatal testing does not translate to enhanced autonomy. Am J Bioeth. 2017; 17 (1):39–41. [ PubMed ] [ Google Scholar ]
  • Raz AE, Vizner Y. Carrier matching and collective socialization in community genetics: dor Yeshorim and the reinforcement of stigma. Soc Sci Med. 2008; 67 :1361–1369. [ PubMed ] [ Google Scholar ]
  • Rhodes R. Resisting paternalism in prenatal whole-genome sequencing. Am J Bioeth. 2017; 17 (1):35–37. [ PubMed ] [ Google Scholar ]
  • Richards FH. Maturity of judgement in decision-making for predictive testing for nontreatable adult-onset neurogenetic conditions: a case against predictive testing of minors. Clin Genet. 2006; 70 :396–401. [ PubMed ] [ Google Scholar ]
  • Richards S, Aziz N, Bale S, Bick D, Das S, Gastier-Foster J, Grody WW, Hegde M, Lyon E, Spector E, Voelkerding K, Rehm HL, on behalf of the ACMG Laboratory Quality Assurance Committee Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet Med. 2015; 17 :405–424. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Ross LF. Screening for conditions that do not meet the Wilson and Jungner criteria: the case of Duchenne muscular dystrophy. Am J Med Genet. 2006; 140A :914–922. [ PubMed ] [ Google Scholar ]
  • Ross LF (2013) Predictive genetic testing of children and the role of the best interest standard. J Law Med Ethics Winter 2013:899–906 [ PubMed ]
  • Ross LF, Clarke AJ. A historical and current review of newborn screening for neuromuscular disorders from around the world: lessons for the United States. Pediatr Neurol. 2017; 77 :12–22. [ PubMed ] [ Google Scholar ]
  • Ross LF, Saal HM, David KL, Anderson RR, American Academy of Pediatrics; American College of Medical Genetics and Genomics Technical report: ethical and policy issues in genetic testing and screening of children. Genet Med. 2013; 15 (3):234–245. [ PubMed ] [ Google Scholar ]
  • Rozario S. Growing up and living with neurofibromatosis1 (NF1): A British Bangladeshi Case-study. J Genet Couns. 2007; 16 :551–560. [ PubMed ] [ Google Scholar ]
  • Sarangi S, Bennert K, Howell L, Clarke A, Harper P, Gray J. (Mis) alignments in counselling for HD predictive testing: clients’ responses to reflective frames. J Gen Couns. 2005; 14 :29–42. [ PubMed ] [ Google Scholar ]
  • Savulescu J, Kahne G. The moral obligation to create children with the best chance of the best life. Bioethics. 2009; 23 :274–290. [ PubMed ] [ Google Scholar ]
  • Schmitz D, Clarke A, Dondorp W. The fetus as a patient—a contested concept and its normative implications. London: Routledge; 2018. [ Google Scholar ]
  • Schoonen M, van der Zee B, Wildschut H, de Beaufort I, de Wert G, de Koning H, Essink-Bot ML, Steegers E. Informing on prenatal screening for down syndrome prior to conception. An empirical and ethical perspective. Am J Med Genet A. 2012; 158A (3):485–497. [ PubMed ] [ Google Scholar ]
  • Scully JL. Disability bioethics. moral bodies, moral difference. Plymouth: Rowman & Littlefield; 2008. [ Google Scholar ]
  • Scully JL, Porz R, Rehman-Sutter C. You don’t make genetic test decisions from one day to the next’—using time to preserve moral space. Bioethics. 2007; 21 :208–217. [ PubMed ] [ Google Scholar ]
  • Shakespeare T. Choices and rights: eugenics, genetics and disability equality. Disabil Soc. 1998; 13 :665–681. [ PubMed ] [ Google Scholar ]
  • Shaw A. They say Islam has a solution for everything, so why are there no guidelines for this?' Ethical dilemmas associated with the births and deaths of infants with fatal abnormalities from a small sample of Pakistani Muslim couples in Britain. Bioethics. 2012; 26 (9):485–492. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Shaw A, Hurst JA. What is this genetics, anyway? Understandings of genetics, illness causality and inheritance among British Pakistani users of genetic services. J Genet Couns. 2008; 17 (4):373–383. [ PubMed ] [ Google Scholar ]
  • Shaw A, Hurst J. 'I don't see any point in telling them': attitudes to sharing genetic information in the family and carrier testing of relatives among British Pakistani adults referred to a genetics clinic. Ethn Health. 2009; 14 :205–224. [ PubMed ] [ Google Scholar ]
  • Shipman H, Clarke AJ, Sarangi S. Accounts of consent: orienting to self–other relations regarding motivations to participate in cancer bio-banking. Commun Med. 2014; 11 (1):69–84. [ PubMed ] [ Google Scholar ]
  • Skotko BG. With new prenatal testing, will babies with Down syndrome disappear? Arch Dis Child. 2009; 94 :823–826. [ PubMed ] [ Google Scholar ]
  • Sobel S, Cowan DB. Impact of genetic testing for Huntington disease on the family system. Am J Med Genet. 2000; 90 :49–59. [ PubMed ] [ Google Scholar ]
  • Swoboda KJ. Seize the day: newborn screening for SMA. Am J Med Genet A. 2010; 152A :1605–1607. [ PubMed ] [ Google Scholar ]
  • Tabor HK, Berkman BE, Hull SC, Bamshad MJ. Genomics really gets personal: how exome and whole genome sequencing challenge the ethical framework of human genetics research. Am J Med Genet A. 2011; 155A (12):2916–2924. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Tassicker RJ, Teltscher B, Trembath MK, et al. Problems assessing uptake of Huntington disease predictive testing and a proposed solution. Eur J Hum Genet. 2009; 17 :66–70. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Taylor-Phillips S, Freeman K, Geppert J, Agbebiyi A, Uthman O, Madan J, Clarke AJ, Quenby S, Clarke AE. Accuracy of non-invasive prenatal testing using cell-free DNA for detection of Down, Edwards and Patau syndromes: a systematic review and meta-analysis. BMJ Open. 2016; 6 :e010002. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Thiele P. He was my son, not a dying baby. J Med Ethics. 2010; 36 :646–647. [ PubMed ] [ Google Scholar ]
  • Timmermans S. Trust in standards: transitioning clinical exome sequencing from bench to bedside. Soc Stud Sci. 2015; 45 (1):77–99. [ PubMed ] [ Google Scholar ]
  • Timmermans S, Buchbinder M. Patients-in-waiting: living between sickness and health in the genomics era. J Health Soc Behav. 2010; 51 (4):408–423. [ PubMed ] [ Google Scholar ]
  • Timmermans S, Buchbinder M. Expanded newborn screening: articulating the ontology of diseases with bridging work in the clinic. Sociol Health Illn. 2012; 34 (2):208–220. [ PubMed ] [ Google Scholar ]
  • Tudor Hart J. The inverse care law. Lancet. 1971; 297 (7696):405–412. [ PubMed ] [ Google Scholar ]
  • Urquia ML, Ray JG, Wanigaratne S, Moineddin R, O’Campo PJ. Variations in male–female infant ratios among births to Canadian- and Indian-born mothers, 1990–2011: a population-based register study. CMAJ Open. 2016; 4 (2):E116–E123. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • van Berkel D, van der Weele C. Norms and prenorms on prenatal diagnosis: new ways to deal with morality in counseling. Patient Educ Couns. 1999; 37 :153–163. [ PubMed ] [ Google Scholar ]
  • van den Heuvel, et al. Will the introduction of non-invasive prenatal diagnostic testing erode informed choices? Patient Educ Couns. 2010; 78 :24–28. [ PubMed ] [ Google Scholar ]
  • Vears D, Metcalfe S. Carrier testing in children and adolescents. EJMG. 2015; 58 :659–667. [ PubMed ] [ Google Scholar ]
  • Villani A, Tabori U, Schiffman J, Shlien A, Beyene J, Druker H, Novokmet A, Finlay J, Malkin D. Biochemical and imaging surveillance in germline TP53 mutation carriers with Li-Fraumeni syndrome: a prospective observational study. Lancet Oncol. 2011; 12 :559–567. [ PubMed ] [ Google Scholar ]
  • Vos J, Jansen AM, Menko F, van Asperen CJ, Stiggelbout AM, Tibben A. Family communication matters: the impact of telling relatives about unclassified variants and uninformative DNA-test results. Genet Med. 2011; 13 :333–341. [ PubMed ] [ Google Scholar ]
  • Voysey M. A constant burden. The reconstitution of family life. London: Routledge & Kegan Paul; 1975. [ Google Scholar ]
  • Weiner K. Exploring genetic responsibility for the self, family and kin in the case of hereditary raised cholesterol. Soc Sci Med. 2011; 72 :1760–1767. [ PubMed ] [ Google Scholar ]
  • White-van Mourik M. Termination of a second-trimester pregnancy for fetal abnormality. Psychosocial aspects. Chapter 5. In: Clarke A, editor. Genetic counselling: practice and principles. London: Routledge; 1994. pp. 113–132. [ Google Scholar ]
  • White-van Mourik MCA, Connor JM, Ferguson-Smith MA. The psychosocial sequelae of a second-trimester termination of pregnancy for fetal abnormality. Prenat Diagn. 1992; 12 :189–204. [ PubMed ] [ Google Scholar ]
  • Wilkinson DJC. Antenatal diagnosis of trisomy 18, harm and parental choice. J Med Ethics. 2010; 36 :644–645. [ PubMed ] [ Google Scholar ]
  • Williams C, et al. Is non-directiveness possible within the context of antenatal screening and testing? Soc Sci Med. 2002; 54 :339–347. [ PubMed ] [ Google Scholar ]
  • Wolff G, Jung C. Nondirectiveness and genetic counseling. J Gen Couns. 1995; 4 :3–25. [ PubMed ] [ Google Scholar ]
  • Woodin SL (2012) Disability: western theories. Encyclopedia of Life Sciences (eLS): Wiley, Chichester. 10.1002/9780470015902.a0005213.pub2

SUGGESTED ADDITIONAL READING

  • Arribas-Ayllon M, Srikant Sarangi S, Angus Clarke A (2011 Genetic testing. Accounts of autonomy, responsibility and blame. Routledge
  • Association of Genetic Nurses and Counsellors (AGNC). (2014) Vision statement: the future role of genetic counsellors in genomic healthcare
  • Biesecker LG, Burke W, Kohane I, Plon SE, Zimmern R. Next generation sequencing in the clinic: are we ready? Nat Rev Genet. 2012; 13 :818–824. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bosk C (1992) All god’s mistakes. Genetic counseling in a pediatric hospital. Chicago
  • Braithwaite and Baxter (2006) Engaging theories in family communication: multiple perspectives. Sage
  • Burton H, Moorthie S (2010) Expanded newborn screening: A review of the evidence. NHS NIHR, NIHR CLAHRC for South Yorkshire and PHG Foundation
  • Clarke A, Parsons EP (eds) (1997) Culture, kinship and genes. Macmillan
  • Clayton EW. State run newborn screening in the genomic era, or how to avoid drowning when drinking from a fire hose. J Law Med Ethics Fall. 2010; 2010 :697–700. [ PubMed ] [ Google Scholar ]
  • Clayton EW, Brothers K. State-offered ethnically targeted reproductive genetic testing. Genet Med. 2016; 18 (2):126–127. [ PubMed ] [ Google Scholar ]
  • Corrigan, McMillan, Liddell, Richards, Weijer (eds) (2009) The limits of consent. Oxford
  • Ebtehaj, Lindley, Richards (eds) (2006) Kinship matters. Cambridge
  • van El CG, Cornel MC, Borry P, Hastings RJ, Fellmann F, Hodgson SV, Howard HC, Cambon-Thomsen A, Knoppers BM, Meijers-Heijboer H, Scheffer H, Tranebjaerg L, Dondorp W, de Wert GM, ESHG Public and Professional Policy Committee Whole-genome sequencing in health care. Recommendations of the European Society of Human Genetics. Eur J Hum Genet. 2013; 21 (Suppl 1):S1–S5. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Evans C. Genetic counselling. A psychological approach. Cambridge: Cambridge University Press; 2006. [ Google Scholar ]
  • Gaff CL, Bylund CL. Family communication about genetics. Theory and practice. Oxford: Oxford University Press; 2010. [ Google Scholar ]
  • Goffman E. Stigma. Harmondsworth: Penguin Books; 1963. [ Google Scholar ]
  • Gould SJ. The mismeasure of man. Harmondsworth: Penguin Books; 1981. [ Google Scholar ]
  • Grody WW. Expanded carrier screening and the law of unintended consequences: from cystic fibrosis to fragile X. Genet Med. 2011; 13 :996–997. [ PubMed ] [ Google Scholar ]
  • Hall A. Realising genomics. Cambridge: PHG Foundation; 2014. [ Google Scholar ]
  • Harper PS, Clarke A. Genetics, society and clinical practice. Oxford: Bios Scientific Publishers; 1997. [ Google Scholar ]
  • Herring J, Foster C. “Please don’t tell me”: the right not to know. Camb Q Healthc Ethics. 2012; 21 :20–29. [ PubMed ] [ Google Scholar ]
  • Kessler S, Resta R (ed) (2000) Psyche and helix. Psychological aspects of genetic counselling. Wiley-Liss
  • Kevles DJ. In the name of eugenics. Genetics and the uses of human heredity. Harmondsworth: Penguin Books; 1986. [ Google Scholar ]
  • Lucassen AM, Hall A and the Joint Committee on Medical Genetics (2011) Consent and confidentiality in genetic practice: guidance on genetic testing and sharing genetic information, 2nd edn. Royal College of Physicians and Royal College of Pathologists, London
  • McDermott R. Ethics, epidemiology and the thrifty gene: biological determinism as a health hazard. Soc Sci Med. 1998; 47 :1189–1195. [ PubMed ] [ Google Scholar ]
  • Parker M, Lucassen A. Concern for families and individuals in clinical genetics. J Med Ethics. 2003; 29 :70–73. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Rehmann-Sutter C, Müller H-J. Disclosure dilemmas. Ethics of genetic prognosis after the 'Right to know/not to Know' debate. Farnham: Ashgate; 2009. [ Google Scholar ]
  • Resta R (2006) Defining and redefining the scope and goals of genetic counseling. Am J Med Genet C 142C: 269–75 [ PubMed ]
  • Rozario S (2013) Genetic disorders and Islamic identity among British Bangladeshis. University Press, Carolina
  • Shakespeare T (2006) Disability rights and wrongs. Routledge
  • Shaw A, (2009) Negotiating risk: British Pakistani experiences of genetics. Berghahn Books
  • Wiggins J, Middleton A (2013) Getting the message across. Communication with diverse populations in clinical genetics. Oxford

Skip to content

Explore VP&S

Academic programs.

Join our MD program to become a physician with compassion, a sense of self, and true grit in all medical pursuits. 

For Researchers

Areas of research.

VP&S continues to be a world leader across the entire spectrum of basic science, translational, and clinical research

  • Patient Care

Departments & Centers

Clinical departments and divisions.

Our mission is to provide world-class patient care, foster innovative research, and train the next generation of leaders in medicine.

Class of 2023

Class of 2022 

Class of 2021

IMAGES

  1. Genetic Counseling

    research about genetic counselling

  2. Genetic Screening & Counseling

    research about genetic counselling

  3. What is Genetic Counseling?

    research about genetic counselling

  4. Genetic counseling and the role of genetic counselors in the United States

    research about genetic counselling

  5. Genetic counselling

    research about genetic counselling

  6. Genetic Counseling in Practice: Genetic Counseling Research : A

    research about genetic counselling

VIDEO

  1. Genetic Counseling

  2. Cancer Genetic Testing with Kayla York, CGC

  3. Role of Genetic Counselling in Prenatal Genetic Screening & Diagnostic Tests- Dr Jaya Vyas

  4. Genetics for breeders ( 3/3 )

  5. Cancer Research Malaysia on Genetic Counselling and Asian Risk Calculator (ARiCa)

  6. Psychiatric Genetic Counselling presentation

COMMENTS

  1. Journal of Genetic Counseling

    The Journal of Genetic Counseling (JOGC), published for the National Society of Genetic Counselors, is a timely, international forum addressing all aspects of the discipline and practice of genetic counseling.The journal focuses on the critical questions and problems that arise at the interface between rapidly advancing technological developments and the concerns of individuals and communities ...

  2. Global trends and themes in genetic counseling research

    Introduction. The term "genetic counseling" was first coined by an American scientist, Sheldon C. Reed, in 1947 [].According to the Ad Hoc Committee of the American Society of Human Genetics, genetic counseling addresses human issues related to the incidence or risk of a genetic disorder in a family [].Seymour Kessler later defined genetic counseling as psychological contact or ...

  3. Genetic counselling

    Genetic counselling articles from across Nature Portfolio. Genetic counselling is an educational process that aims to inform and advise patients and relatives at risk of a genetic condition about ...

  4. Genetic Counseling and the Central Tenets of Practice

    Genetic counseling is a rapidly evolving professional practice that has kept pace with emerging genetic technologies and related ... education about inheritance, testing, management, prevention, resources and research; and counseling to promote informed choices and adaptation to the risk or condition." This definition has been cited often ...

  5. NSGC > Research and Publications > Journal of Genetic Counseling

    The Journal of Genetic Counseling, the official journal of the National Society of Genetic Counselors, is a timely, international forum addressing all aspects of the discipline and practice of genetic counseling.The journal focuses on the critical questions and problems that arise at the interface between rapidly advancing technological developments and the concerns of individuals and ...

  6. Global trends and themes in genetic counseling research

    Abstract. This article seeks to highlight the most recent trends and themes in genetic counseling that are of broad interest. A total of 3505 documents were published between 1952 and 2021, with a ...

  7. What do we do and how do we do it? Assessing genetic counselling in the

    Assessing genetic counselling in the modern era. ... LY is a recipient of a co-funded National Heart Foundation of Australia/National Health and Medical Research Council (NHMRC) PhD scholarship ...

  8. Genetic counselling in the era of genomic medicine

    Genetic counselling. There is consensus among professional bodies around the world that the act of genetic counselling is a client-centred communication process with the aim of helping patients understand, adapt and adjust to the medical or psychosocial consequences of genetic contributions to disease. 12 Definitions of genetic counselling have evolved over time, however professionals trained ...

  9. Genetic Counseling Research: A Practical Guide

    Abstract. Genetic Counseling Research: A Practical Guide is an online resource devoted to research methodology in genetic counseling. It offers step-by-step guidance for conducting research, from the development of a question to the publication of findings. Genetic counseling examples and practical tips guide readers through the research and ...

  10. Genetic counselling in the era of genomic medicine

    Genetic counselling is a communication process empowering patients and families to make autonomous decisions and effectively use new genetic information. The skills of genetic counselling and expertise of genetic counsellors are integral to the effective implementation of genomic medicine. Sources of data: Original papers, reviews, guidelines ...

  11. Genetic Counseling Researchers Shape the Future of Precision ...

    In honor of Genetic Counselor Awareness Day on Nov. 8, National Society of Genetic Counselors (NSGC) members Heather A. Zierhut, PhD, MS, CGC, and Adam H. Buchanan, MS, MPH, LGC, share how their research is shaping the future of genetic counseling and precision medicine. NSGC: What inspired you to be a genetic counselor?

  12. Recent Publications

    Michelle Florido, Jessica Giordano: Processing the process: Reflections on genetic counselor-led student supervision groups and practical tips for future supervisors Journal of Genetic Counseling Alison Garber, Lisa Weingarten, Nicolas Abreu, Houda Elloumi, Tobias Haack, Clara Hildebrant, Nuria Martinez-Gil, Jennifer Mathews, Amelia Miller, Irene Valenzuela Palafoll, Connolly Steigerwald (GC ...

  13. Characterization of genetic counselor practices in inpatient care

    It establishes a foundation for future research on inpatient genetic counseling and genetic counseling outcomes in inpatient services. As demand for genetics expertise in inpatient care grows, genetic counselors can be hired to serve inpatient populations alongside genetics and non-genetics providers.

  14. Advancing the genetic counseling profession through research

    The task force developed a research agenda outlining high-priority research questions for the next 5 years. The questions are organized into four domains: (a) Genetic Counseling Clients; (b) Genetic Counseling Process and Outcomes; (c) Value of Genetic Counseling Services; and (d) Access to Genetic Counseling Services.

  15. Genetic Counseling

    Genetic counseling gives you information about how genetic conditions might affect you or your family. The genetic counselor or other healthcare professional will collect your personal and family health history. They can use this information to determine how likely it is that you or your family member has a genetic condition.

  16. Research

    At UCSF, Genetic Counseling students undertake a capstone research project consistent with their own interest. Along the way, they acquire skills to formulate a research question, design and conduct research studies, critically evaluate scientific literature, and learn to disseminate their findings. Research questions may be clinically or ...

  17. Research

    Columbia University has a long history of robust scholarly activity and our Program is no exception. Our faculty conduct cutting-edge research in genetic counseling, mentoring our students and strengthening their skills and passion for the research process. Learn more about our student resesarch. Explore our recent faculty and student publications.

  18. JHU/NIH Genetic Counseling Training Program (GCTP)

    Mission Statement. The JHU/NIH Genetic Counseling Training Program shapes genetic counseling services through student and faculty research and develops outstanding genetic counselors who are innovators and leaders in: psychotherapeutic genetic counseling, genetic counseling research and scholarship, applications of genomics and precision health.

  19. Master of Science (ScM) in Genetic Counseling

    The Johns Hopkins/National Institutes of Health Genetic Counseling Training Program is a joint effort of the Department of Health, Behavior and Society and the National Institutes of Health through the National Human Genome Research Institute (NHGRI) and the National Cancer Institute (NCI). The ScM in Genetic Counseling prepares graduates for a career in genetic counseling with an emphasis on ...

  20. The role of the genetic counsellor: a systematic review of research

    Introduction. Although the term 'genetic counselling' was coined by Sheldon Reed in 1947, 1 the genetic counselling profession is relatively young in comparison with medicine and nursing. Although genetic counselling can be undertaken by trained professionals from a range of disciplines, those describing themselves as genetic counsellors are specifically trained for the work.

  21. The role of the genetic counsellor: a systematic review of research

    A definition of genetic counselling as an activity was produced by the NSGC Taskforce in 2006 and states that 'genetic counselling is the process of helping people understand and adapt to the ...

  22. Research priorities in psychiatric genetic counselling: how to talk to

    Genetic counselling is a process through which a trained professional helps an individual to better understand and adapt to the medical, psychological, and familial implications of genetic ...

  23. Genetic Counseling (Inherited Diseases)

    Genetic counseling seeks to help you understand complex genetic information and the options available for genetic testing. During the counseling appointment, a review of your personal medical history and family health history will be conducted. A family tree, called a pedigree, will be created to document your family history.

  24. Student Research

    MS in Genetic Counseling students complete coursework to improve familiarity with the research process, understand the roles of genetic counselors in the research process, and identify and describe the growing body of research about the practice and profession of genetic counseling.

  25. Ethics in genetic counselling

    Similarly, any experimentation or research involving humans must be conducted with the consent of the research participants, who must be fully aware of any potential risks. ... While a reluctance to engage with the genetic counselling process will often need to be discussed and to become a topic within the consultation, there may be other ...

  26. Genetic Counselor in Miami, FL for Mount Sinai Medical Center Miami

    Coordinate and research all insurance coverage for genetic testing for patients. Researching the best tests based on patient's insurance coverage. Prepare letters of medical necessity. ... Certification or eligible through the ABGC (American Board of Genetic Counseling) 1-2 years of genetic counseling experience preferred Comprehensive ...

  27. Class of 2025

    Overview of Class of 2025 research projects. Join our MD program to become a physician with compassion, a sense of self, and true grit in all medical pursuits.

  28. Alumni

    Inclusive Genetic Counseling for LGBTQ+ Clients: Experiences and Perspectives of Reproductive Genetic Counselors - JEMF Student Research Grant Award; NSGC ART/Infertility Research Grant Award: Amanda Chan: Piloting an Educational Modelon Consenting for Exome Sequencing Among Non-Genetics Physicians at a Medical Institution: Carlos Dominguez ...