You are on page 1of 429

Principles of Psychiatric

Genetics

Edited by
John I. Nurnberger Jr. MD, PhD
Joyce and Iver Small Professor of Psychiatry
Professor of Medical and Molecular Genetics and Medical Neuroscience
Director of the Institute of Psychiatric Research
Department of Psychiatry
Indiana University School of Medicine
Indianapolis, IN, USA

Wade H. Berrettini MD, PhD


Karl E. Rickels Professor of Psychiatry
Director, Center for Neurobiology and Behavior
University of Pennsylvania School of Medicine
Philadelphia, PA, USA
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK

Published in the United States of America


by Cambridge University Press, New York

www.cambridge.org
Information on this title: www.cambridge.org/9780521896498

# Cambridge University Press 2012

This publication is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without
the written permission of Cambridge University Press.

First published 2012

Printed and bound in the United Kingdom by the MPG Books Group

A catalogue record for this publication is available from the British Library

Library of Congress Cataloging-in-Publication Data


Principles of psychiatric genetics / edited by John I. Nurnberger,
Wade H. Berrettini.
p. ; cm.
ISBN 978-0-521-89649-8 (Hardback)
I. Nurnberger, John I., 1946– II. Berrettini, Wade.
[DNLM: 1. Mental Disorders–genetics. 2. Mental Disorders–
epidemiology. WM 140]
616.890 042–dc23
2012000081

ISBN 978-0-521-89649-8 Hardback

Cambridge University Press has no responsibility for the persistence or


accuracy of URLs for external or third-party internet websites referred to
in this publication, and does not guarantee that any content on such
websites is, or will remain, accurate or appropriate.

Every effort has been made in preparing this book to provide


accurate and up-to-date information which is in accord with
accepted standards and practice at the time of publication.
Although case histories are drawn from actual cases, every effort has
been made to disguise the identities of the individuals involved.
Nevertheless, the authors, editors and publishers can make no
warranties that the information contained herein is totally free from
error, not least because clinical standards are constantly changing
through research and regulation. The authors, editors and
publishers therefore disclaim all liability for direct or consequential
damages resulting from the use of material contained in this book.
Readers are strongly advised to pay careful attention to information
provided by the manufacturer of any drugs or equipment that they
plan to use.
Contents
List of contributors vii
Preface xi

1 Contribution of genetic epidemiology to 11 Genetic contributions to obsessive–


our understanding of psychiatric compulsive disorder (OCD) and
disorders 1 OCD-related disorders 121
Kathleen Ries Merikangas and Anibal Cravchik Dennis L. Murphy, Pablo R. Moya,
Jens R. Wendland, and Kiara Timpano
2 A basic overview of contemporary human
12 Post-traumatic stress disorder 134
genetic analysis strategies 13
Michael J. Lyons, Tyler Zink, and Karestan
Ondrej Libiger and Nicholas J. Schork
C. Koenen
3 DNA methods 23
13 Antisocial behavior: gene–environment
David W. Craig interplay 145
4 In silico analysis strategies and resources for Laura A. Baker, Catherine Tuvblad, Serena
psychiatric genetics research 34 Bezdjian, and Adrian Raine
Ali Torkamani, Trygve Bakken, and Nicholas 14 Learning disabilities 160
J. Schork Shelley D. Smith
5 Gene expression studies in psychiatric 15 Attention-deficit hyperactivity disorder 168
disorders 49
Josephine Elia, Francesca Lantieri, Toshinobu
Alexander B. Niculescu, III Takeda, Xiaowu Gai, Peter S. White, Marcella
6 Pharmacogenetics in psychiatry 53 Devoto, and Hakon Hakonarson
Falk W. Lohoff 16 Autism and autism spectrum disorders 183
7 Functional validation of candidate genetic Daniel H. Geschwind and Maricela Alarcón
susceptibility factors for major mental 17 The genetics of bipolar disorder 196
illnesses 69
John R. Kelsoe
Akira Sawa, Wanli W. Smith, Saurav Seshadri,
Akiko Hayashi-Takagi, Hanna Jaaro-Peled, 18 The genetics of major depression 212
and Atsushi Kamiya James B. Potash
8 Epigenetic mechanisms in drug addiction 19 The genetics of schizophrenia 230
and depression 79 Hugh M. D. Gurling and Andrew McQuillin
William Renthal and Eric J. Nestler
20 The genetics of anorexia and bulimia
9 Panic disorder 90 nervosa 262
Ardesheer Talati and Myrna M. Weissman Andrew W. Bergen, Jennifer Wessel, and Walter
H. Kaye
10 The genetics of the phobic disorders
and generalized anxiety disorder 112 21 Genetics and common human obesity 272
Raymond R. Crowe R. Arlen Price

v
Contents

22 Alcoholism 279 28 Genetics of Tourette syndrome


Howard J. Edenberg and related disorders 336
Maria G. Motlagh, Thomas V. Fernandez,
23 Nicotine dependence 287
and James F. Leckman
Sarah M. Hartz and Laura J. Bierut
29 Endophenotypes in psychiatric genetics 347
24 Human molecular genetics of opioid
Andrew C. Chen, Madhavi Rangaswamy,
addiction 297
and Bernice Porjesz
Mary Jeanne Kreek, Dmitri Proudnikov,
David A. Nielsen, and Vadim Yuferov 30 Developmental disorders 363
Craig A. Erickson, Khendra I. Peay,
25 Genetics of stimulant dependence 306
and Christopher J. McDougle
Joseph F. Cubells and Yi-Lang Tang
31 Alzheimer’s disease 371
26 Genetics of personality disorders 316
Carlos Cruchaga, John S. K. Kauwe,
C. Robert Cloninger and Alison M. Goate
27 Ethical issues in behavioral
genetics 324
Stephen H. Dinwiddie, Jinger Hoop,
and Elliot Gershon Index 382
Color plate section is between pp. 212 and 213.

vi
Contributors

Maricela Alarcón, PhD David W. Craig, PhD


Center for Autism Research and Treatment, The Translational Genomics Research Institute,
Semel Institute of Neuroscience, Program in Phoenix, AZ, USA
Neurogenetics, Department of Neurology, David
Geffen School of Medicine, University of California Anibal Cravchik, MD, PhD
at Los Angeles, Los Angeles, CA, USA Genetic Epidemiology Research Branch,
National Institute of Mental Health Intramural
Laura A. Baker, PhD Research Program, National Institutes of Health,
Department of Psychology, University of Southern Bethesda, MD, USA
California, Los Angeles, CA, USA
Raymond R. Crowe, MD
Trygve Bakken Department of Psychiatry, University of Iowa Carver
The Scripps Translational Science Institute College of Medicine, Iowa City, IA, USA
and The Department of Molecular and Experimental
Carlos Cruchaga, PhD
Medicine, The Scripps Research Institute;
Graduate Program in Neurosciences and Medical Department of Psychiatry, Washington University
Scientist Training Program, University of California, School of Medicine, St. Louis, MO, USA
San Diego, La Jolla, CA, USA Joseph F. Cubells, MD, PhD
Serena Bezdjian Departments of Human Genetics and Psychiatry and
Behavioral Sciences, Emory University School of
Department of Psychology, University of Southern
Medicine, Atlanta, GA, USA
California, Los Angeles, CA, USA
Marcella Devoto, PhD
Andrew W. Bergen, PhD
Division of Genetics, Department of Pediatrics, The
Molecular Genetics Program, Center for Health
Children’s Hospital of Philadelphia and Department
Sciences, SRI International, Menlo Park,
of Biostatistics and Epidemiology, University of
CA, USA
Pennsylvania School of Medicine, Philadelphia, PA,
Laura J. Bierut, MD USA; Dipartimento di Medicina Sperimentale,
Department of Psychiatry, Washington University University La Sapienza, Rome, Italy
School of Medicine, St. Louis, MO, USA Stephen H. Dinwiddie, MD
Andrew C. Chen, MD Department of Psychiatry and Behavioral Science,
The University of Chicago Medical Center,
Department of Psychiatry, Columbia University
Chicago, IL, USA
Medical Center College of Physicians and Surgeons,
New York, NY, USA Howard J. Edenberg, PhD
Center for Medical Genomics, Department
C. Robert Cloninger, MD
of Biochemistry and Molecular Biology,
Department of Psychiatry, Washington University
Indiana University School of Medicine,
School of Medicine, St. Louis, MO, USA
Indianapolis, IN, USA

vii
List of contributors

Josephine Elia, MD Akiko Hayashi-Takagi, PhD


Department of Child and Adolescent Psychiatry, Department of Psychiatry, Johns Hopkins University
The Children’s Hospital of Philadelphia and School of Medicine, Baltimore, MD, USA
The University of Pennsylvania,
Philadelphia, PA, USA Jinger Hoop, MD, MFA
Edward Hines Jr. Veteran's Administration Hospital,
Craig A. Erickson, MD Hines, IL, USA
Department of Psychiatry, Indiana University School
of Medicine; Christian Sarkine Autism Treatment Hanna Jaaro-Peled, PhD
Center, James Whitcomb Riley Hospital for Children, Department of Psychiatry, Johns Hopkins University
Indianapolis, IN, USA School of Medicine, Baltimore, MD, USA

Thomas V. Fernandez, MD Atsushi Kamiya, MD, PhD


Child Study Center, Yale University School Department of Psychiatry, Johns Hopkins University
of Medicine, New Haven, CT, USA School of Medicine, Baltimore, MD, USA

Xiaowu Gai, PhD John S. K. Kauwe, PhD


Department of Pharmacology, Loyola University, Department of Biology, Brigham Young University,
Chicago, IL, USA Provo, UT, USA

Elliot Gershon, MD Walter H. Kaye, MD


Department of Psychiatry, The University of Chicago University of California San Diego, Department
Medicine, Chicago, IL, USA of Psychiatry and Eating Disorder Research and
Treatment Program, La Jolla, CA, USA
Daniel H. Geschwind, MD, PhD
Center for Autism Research and Treatment, Semel John R. Kelsoe, MD
Institute of Neuroscience, Program in Neurogenetics, Department of Psychiatry and Institute for Genomic
Department of Neurology, David Geffen School Medicine, University of California San Diego, VA
of Medicine, and Department of Human Genetics, San Diego Healthcare System, La Jolla, CA, USA
University of California at Los Angeles, Los Angeles,
CA, USA Karestan C. Koenen, PhD
Harvard School of Public Health, Boston,
Alison M. Goate, D.Phil MA, USA
Departments of Psychiatry, Neurology, Alzheimer’s
Disease Research Center, Genetics, and Hope Center Mary Jeanne Kreek, MD
for Neurological Disorders, Washington University The Laboratory of the Biology of Addictive Diseases,
School of Medicine, St. Louis, MO, USA Rockefeller University, New York, NY, USA

Hugh M. D. Gurling, MD Francesca Lantieri, PhD


Molecular Psychiatry Laboratory, Mental Health Department of Child and Adolescent Psychiatry,
Sciences Unit, University College London, The Children’s Hospital of Philadelphia,
London, UK Philadelphia, PA, USA

Hakon Hakonarson, MD, PhD James F. Leckman, MD


Division of Pulmonary Medicine, Department of Child Study Center and Departments of Psychiatry,
Pediatrics, The Center for Applied Genomics, Psychology, and Pediatrics, Yale University School of
The Children’s Hospital of Philadelphia and The Medicine, New Haven, CT, USA
University of Pennsylvania, Philadelphia, PA, USA
Ondrej Libiger, MA, PhD
Sarah M. Hartz, MD, PhD The Scripps Translational Science Institute and
Department of Psychiatry, Washington University Department of Molecular and Experimental
School of Medicine, St. Louis, MO, USA Medicine, The Scripps Research Institute,

viii
List of contributors

La Jolla, CA, USA; Lekarska Fakulta v Hradci Khendra I. Peay, MD


Kralove, Charles University, Czech Republic Department of Psychiatry, Indiana University School
of Medicine; Christian Sarkine Autism Treatment
Falk W. Lohoff, MD
Center, James Whitcomb Riley Hospital for Children,
Department of Psychiatry, University of Pennsylvania Indianapolis, IN, USA
School of Medicine, Philadelphia, PA, USA
Bernice Porjesz, PhD
Michael J. Lyons, PhD
Henri Begleiter Neurodynamics Laboratory,
Department of Psychology, Boston University, SUNY Downstate Medical Center, Brooklyn,
Boston, MA, USA New York, NY, USA
Christopher J. McDougle, MD
James B. Potash, MD, MPH
Professor of Psychiatry and Pediatrics, Director of the
Department of Psychiatry, Carver College of
Lurie Center for Autism, Massachusetts General
Medicine, University of Iowa, Iowa City, IA, USA
Hospital and MassGeneral Hospital for Children,
Harvard Medical School, Boston, MA, USA R. Arlen Price, PhD
University of Pennsylvania, Center for
Andrew McQuillin, PhD
Neurobiology and Behavior, Translational
Molecular Psychiatry Laboratory, Research
Research Laboratories, Philadelphia,
Department of Mental Health Sciences, University
PA, USA
College London, London, UK
Dmitri Proudnikov, PhD
Kathleen Ries Merikangas, PhD
The Laboratory of the Biology of Addictive Diseases,
Genetic Epidemiology Research Branch, National
Rockefeller University, New York,
Institute of Mental Health Intramural Research
NY, USA
Program, National Institutes of Health,
Bethesda, MD, USA Adrian Raine, DPhil
Maria G. Motlagh, MD Departments of Criminology, Psychiatry, and
Psychology, Jerry Lee Center of Criminology,
Child Study Center, Yale University School of
University of Pennsylvania, Philadelphia, PA, USA
Medicine, New Haven, CT, USA
Pablo R. Moya, PhD Madhavi Rangaswamy, PhD
Laboratory of Clinical Science, National Institute of Henri Begleiter Neurodynamics Laboratory,
Mental Health Intramural Research Program, SUNY Downstate Medical Center, Brooklyn,
National Institutes of Health, Bethesda, MD, USA New York, USA

Dennis L. Murphy, MD William Renthal, MD, PhD


Laboratory of Clinical Science, National Institute of Medical Scientist Training Program, The University
Mental Health Intramural Research Program, of Texas Southwestern Medical Center,
National Institutes of Health, Bethesda, MD, USA Dallas, TX, USA

Eric J. Nestler, MD, PhD Akira Sawa, MD, PhD


Fishberg Department of Neuroscience, Mount Sinai Departments of Psychiatry, Neuroscience, Cellular
School of Medicine, New York, NY, USA and Molecular Medicine, Human Genetics, Johns
Hopkins University School of Medicine,
Alexander B. Niculescu, III, MD, PhD Baltimore, MD, USA
Department of Psychiatry, Indiana University School
of Medicine; Indianapolis, IN, USA Nicholas J. Schork, PhD
The Scripps Translational Science Institute
David A. Nielsen, PhD and The Department of Molecular and Experimental
The Laboratory of the Biology of Addictive Diseases, Medicine, The Scripps Research Institute,
Rockefeller University, New York, NY, USA La Jolla, CA, USA

ix
List of contributors

Saurav Seshadri Catherine Tuvblad, PhD


Department of Psychiatry, Johns Hopkins University Department of Psychology, University of
School of Medicine, Baltimore, MD, USA Southern California, Los Angeles,
CA, USA
Shelley D. Smith, PhD, FACMG
University of Nebraska Medical Center, Munroe Myrna M. Weissman, PhD
Meyer Institute for Genetics and Rehabilitation, Columbia University and New York State Psychiatric
Omaha, NE, USA Institute, New York, NY, USA

Wanli W. Smith, MD, PhD Jens R. Wendland, MD


Department of Psychiatry, Johns Hopkins University Laboratory of Clinical Science, and Genetic
School of Medicine, Baltimore, MD, USA Basis of Mood and Anxiety Disorders Section,
National Institute of Mental Health Intramural
Toshinobu Takeda, MD, PhD Research Program, National Institutes of Health,
Department of Child and Adolescent Psychiatry, Bethesda, MD, USA
The Children’s Hospital of Philadelphia,
Philadelphia, PA, USA; Department of Psychiatry, Jennifer Wessel, PhD
Ryukoku University, Kyoto, Japan Molecular Genetics Program, Center for Health
Sciences, SRI International, Menlo Park, CA;
Ardesheer Talati, PhD Department of Public Health, Indiana
Columbia University and New York State Psychiatric University School of Medicine, Indianapolis,
Institute, New York, NY, USA IN, USA
Yi-Lang Tang, MD, PhD Peter S. White, PhD
Department of Psychiatry and Behavioral Sciences, Department of Pediatrics, The Children’s Hospital
Emory University School of Medicine, Atlanta, GA, USA of Philadelphia and The University of Pennsylvania,
Philadelphia, PA, USA
Kiara Timpano, PhD
University of Miami, Coral Gables, FL, USA Vadim Yuferov, PhD
The Laboratory of the Biology of Addictive Diseases,
Ali Torkamani, PhD Rockefeller University, New York, NY, USA
The Scripps Translational Science Institute and The
Department of Molecular and Experimental Tyler Zink
Medicine, The Scripps Research Institute, Department of Psychology, Boston University,
La Jolla, CA, USA Boston, MA, USA

x
Preface

The major psychiatric disorders are common illnesses While this book is, by necessity, a picture in time
with complex origins in gene–environment inter- of current knowledge, the rapidly advancing technol-
actions. As is the case with most medical diseases, ogy of DNA sequencing is likely to produce a multi-
such as diabetes and hypertension, genetic factors tude of new discoveries in the near future. During the
for psychiatric disorders are composed of many past decade, the cost of sequencing a human genome
common alleles, each with a small effect on risk. has fallen from ~ $1 000 000 000 USD to ~ $5000
Additionally, there are many rare alleles, including USD, and the “thousand dollar genome” is widely
copy number variants (CNVs), with larger effects predicted. This will allow for the sequencing of thou-
on risk. sands of affected individuals within each category of
Despite this complex picture, substantial progress psychiatric disorder. Analysis of this sequence infor-
has been made in identification of individual risk mation will permit the development of a catalogue of
alleles, through various molecular approaches, includ- common and rare genetic variants that increase risk
ing linkage and association, studies of epigenetic for these diseases. Studies of gene expression in post-
factors, and recently direct sequencing of DNA from mortem brain samples and in living blood and skin
affected persons. It is the intent of the contributors tissue from affected persons will enable catalogues of
and editors to describe these recent molecular epigenetic events to be developed as well. These
advances, in the context of the genetic epidemiology advances in genetics and epigenetics should permit
(population, family and twin studies) and our know- an explosion of knowledge concerning the genetic and
ledge of the phenomenology and course of illness of environmental risks for psychiatric disorders.
psychiatric disorders. It is now possible to develop and culture neurons
This book is organized in chapters according to in the laboratory, from easily obtained skin or blood
the current nosology of psychiatric disorders, but the cells of persons with a psychiatric disorder. This
reader should not conclude that this nomenclature is should allow for an unprecedented correlation of
based on genetic, epigenetic, neurobiologic, or envir- neuronal phenotype to genotype on a scale we can
onmental influences. The current nosology is based only imagine. There will be the potential to character-
primarily on acute signs and symptoms, few of ize in beautiful detail the electrophysiologic, morpho-
which are “pathognomonic”. Individuals with pat- logic, and neurochemical characteristics of these
terns of these signs and symptoms are categorized neurons from genetically defined origins. This should
as “having” schizophrenia, for example, but the term permit discovery of specific neuronal abnormalities.
“schizophrenia” undoubtedly refers to a very hetero- That would enable the targeting of therapeutic agents
geneous group of brain disorders which share some to pathophysiology, even if that pathophysiology is
acute signs and symptoms and course of illness vari- unique to an individual person or family
ables. In the future, it is expected that the current The book you have before you should be seen as a
nosology will be transformed into one which reflects window to that day in the future when such targeted,
knowledge of the neurobiological and experiential individualized therapies are used by doctors and
origins of these groups of heterogeneous disorders. patients around the world. We have designed Prin-
It is anticipated that new diagnostic tools and thera- ciples of Psychiatric Genetics to be useful to investi-
peutic agents will result from this expanded molecu- gators in related areas, including Psychiatry, Human
lar knowledge. Genetics, and Neurobiology. However we also expect
that it will be of value to practicing clinicians who

xi
Preface

wish to understand the sometimes confusing and and human genetics are somewhat more constant.
contradictory reports of discoveries in the media. Is You will notice that each chapter on a disease refers
there a blood test for bipolar disorder? Has “a gene” to epidemiology, twin and family studies, and linkage
been identified for alcohol addiction? Is there a gen- and association. Some disease-centered chapters also
etic test for the proper treatment for a patient with refer to epigenetic studies, bioinformatic studies, and
schizophrenia? What new drugs may we expect for drug development. There are special chapters in the
Alzheimer’s disease? first section of the book on each of these methodo-
Implicit in the foregoing discussion is a critical logical areas. These have been prepared by subspeci-
message for every reader of this book: genetic abnor- alty experts in psychiatric genetics and will be useful
malities are not immutable; they are treatable. Each in interpreting the disease-centered chapters and also
genome represents a program for the body and the the journal literature in these areas.
brain, but it is not destiny. In fact genetic programs Each disease chapter utilizes the methods
are altered in their expression by the food we eat, the described and then provides an up-to-date summary
medicine we take, and by everyday experience. The of where we stand now in identifying specific genetic
finished product of a human life is a massively com- influences on that trait or traits. Disease chapters also
plex combination of the genetic program (which in generally provide an overview of the symptoms, signs,
itself does not change, except for mutation) and the and life course of the condition described. We have
effects of our experience, beginning with the intra- concentrated on those conditions usually treated by
uterine environment. Given this complexity, it is psychiatrists and other mental health professionals,
remarkable that genetic signals are even detectable but we are aware of the similarities in origin and
in behavioral disorders, and yet they are. But one course of other neuropsychiatric conditions that
should never expect them to be constant, or may be more frequently seen by pediatricians or
unchangeable. neurologists (i.e. tuberous sclerosis, seizure disorders,
Principles of Psychiatric Genetics should be of or vascular dementia). For a more general reference
interest to every mental health professional in on medical genetics in clinical practice, we would
training. All those who wish to become psychiatrists, refer the reader to the online Mendelian Inheritance
psychologists, and counselors during the twenty-first in Man by McKusick and collaborators.
century should know about the field that we describe, As editors of this volume, we are humbled by the
because it will affect your practice profoundly. Do we contributions of those who have gone before us in
have a blood test for bipolar disorder? No, not today defining the field of Psychiatric Genetics. We would
(despite what you may read on the internet). But such like to thank our mentor and longtime collaborator
tests are not far in the future. When they arrive, they Elliot Gershon, who decades ago taught us the
will likely be based on arrays that examine multiple Principles in this volume. Others who influenced our
gene variants and biological pathways in a single test. approach to the issues in this book include Theodore
Those of you beginning practice today will likely send Reich, Seymour Kety, George Winokur, Irving
your patients for such tests. They will be designed Gottesmann, Ming Tsuang, and Robert Cloninger,
based on some of the Principles in this book. The among others. We thank our contributing authors
other questions above may be answered similarly. for their insight, their industry, their patience, and
There are clues to the answers in the appropriate their trust that our shared effort would result in a
chapters in this book. book that they and others would admire. We appreci-
Because the genetic and epigenetic variants ate the indulgence of our colleagues and families with
responsible for these disorders are not fully available the time this task took away from other activities.
today, we are titling this book with the term Prin- In this regard we thank Patricia Nurnberger and
ciples. The details of this expanding field will change Christine Berrettini most of all. Finally we thank the
daily over the coming years. The pace of discovery in patients and families who continue to teach us in our
the laboratory is daunting, and to remain current clinics every day. By this book, may we provide a
requires monitoring several hundred journals in print stepping stone to better and more productive lives
and/or online (consult the bibliographies in this book for you . . .
for examples). However, the core areas of psychiatry

xii
Contribution of genetic epidemiology to
Chapter

1 our understanding of psychiatric disorders


Kathleen Ries Merikangas and Anibal Cravchik

Introduction diseases and genetic and environmental causes of


familial resemblance. Genetic epidemiology focuses
This chapter will: (1) provide a summary of the back- on how genetic factors and their interactions with
ground disciplines and approaches to understanding other risk factors increase vulnerability to, or protec-
the role of genetic factors in mental disorders; tion against, disease [7]. Genetic epidemiology
(2) review the current knowledge in the genetic epi- employs traditional epidemiological study designs to
demiology of mental disorders; and (3) summarize explain aggregation in groups as closely related as
the role of epidemiology in the current generation of twins or as loosely related as migrant cohorts. Epi-
genome-wide association studies of mental disorders. demiology has developed sophisticated designs and
analytic methods for identifying disease risk factors.
With increasing progress in gene identification, these
Genetic epidemiology methods have been extended to include both genetic
The pioneering work of Böök, Sjögren, Angst, Perris, and environmental factors [8, 9]. In general, study
and others in Europe and Kallman, Heston, designs in genetic epidemiology either control for
Rosenthal, Wender, and Kety in the United States genetic background while letting the environment
firmly established the important role of genetic sus- vary (e.g. migrant studies, half siblings, separated
ceptibility factors in psychiatric disorders. Heston’s twins) or control for the environment while allowing
[1] original finding that adult offspring of hospital- variance in the genetic background (e.g. siblings,
ized schizophrenic mothers had significantly higher twins, adoptees nonbiological siblings). Investigations
rates of schizophrenia than offspring of parents with in genetic epidemiology are typically based on a com-
no mental illness was confirmed and extended by Kety, bination of study designs including family, twin, and
Rosenthal, and Wender’s [2, 3] studies in a much larger adoption studies.
sample of adopted away offspring of schizophrenics in
Denmark. These studies demonstrated clearly that the
presence of schizophrenia in birth parents, independ- Family studies
ent of the rearing environment, significantly increases Familial aggregation is generally the first source
offspring’s risk for the development of the disease of evidence that genetic factors may play a role in a
[4, 5]. During the last decades of the twentieth century, disorder. The most common indicator of familial
the study of genes in psychiatric disorders expanded aggregation is the relative risk ratio, computed as
beyond hospital settings to outpatient treatment set- the rate of a disorder in families of affected persons
tings, particularly in the United States [6]. With the divided by the corresponding rate in families of con-
introduction of epidemiology to the study of psychiatry, trols. The patterns of genetic factors underlying a
systematic control groups were included in family disorder can be inferred from the extent to which
studies and methods for incorporating population patterns of familial resemblance adhere to the expect-
base rates and risk assessment were developed. ations of Mendelian laws of inheritance. The degree of
The field of genetic epidemiology is defined as the genetic relatedness among relatives is based on the
study of the distribution of, and risk factors for proportion of shared genes between a particular

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

1
Chapter 1: Contribution of genetic epidemiology

relative and an index family member or proband. of 50% of their genes) provide estimates of the degree
First-degree relatives share 50% of their genes in to which genetic factors contribute to the etiology
common; second-degree relatives share 25% of their of a disease phenotype. A crude estimate of the gen-
genes in common, and third-degree relatives share etic contribution to risk for a disorder is calculated
12.5% of their genes in common. If familial resem- by doubling the difference between the concordance
blance is wholly attributable to genes, there should be rates for monozygous and dizygous twin pairs.
a 50% decrease in disease risk with each successive Modern genetic studies employ path analytic models
increase in degree of relatedness, from first to second to estimate the proportion of variance attributable to
to third, and so forth. This information can be used to additive genes, common environment, and unique
derive estimates of familial recurrence risk within and environment. There are several other applications of
across generations as a function of population preva- the twin study design that may inform our under-
lence (l) [10]. Whereas l tends to exceed 20 for most standing of the roles of genetic and environmental
autosomal dominant diseases, values of l derived risk factors for disease. First, twin studies provide
from family studies of many complex disorders tend information on the genetic and environmental
to range from 2 to 5. Diseases with strong genetic sources of sex differences in a disease. Second, envir-
contributions tend to be characterized by 50% onmental exposures may be identified through com-
decrease in risk across successive generations. parison of discordant monozygotic twins. Third, twin
Decrease in risk according to the degree of genetic studies can be used to identify the genetic mode of
relatedness can also be examined to detect inter- transmission of a disease by inspection of the degree of
actions between several loci. If the risk to second- adherence of the difference in risk between monozygotic
and third-degree relatives decreases by more than and dizygotic twins to the Mendelian ratio of 50%.
50% this implies that more than a single locus must Fourth, twin studies may contribute to enhancing the
contribute to disease risk and that no single locus can validity of a disease through inspection of the com-
largely predominate. ponents of the phenotypes that are most heritable.
The major advantage of studying diseases within The twin family design is one of the most powerful
families is that disease manifestations are more study designs in genetic epidemiology because it
likely to result within families than they are across yields estimates of heritability but also permits evalu-
families from the same underlying etiological ation of multigenerational patterns of expression of
factors. Family studies are therefore more effective genetic and environmental risk factors. Several recent
than between family designs in examining the val- updates of findings of twin studies of psychiatric
idity of diagnostic categories because they more disorders are available [12, 13].
accurately assess the specificity of transmission of
symptom patterns and disorders. Data from family
studies can also provide evidence regarding etio- Adoption studies
logical or phenotypic heterogeneity. Phenotypic het- Adoption studies have been the major source of evi-
erogeneity is suggested by variable expressivity of dence regarding the joint contribution of genetic and
symptoms of the same underlying risk factors, environmental factors to disease etiology. Adoption
whereas etiological heterogeneity is demonstrated studies either compare the similarity between an
by common manifestations of expression of differ- adoptee and his or her biological versus adoptive
ent etiological factors between families. Moreover, relatives, or the similarity between biological relatives
the family study method permits assessment of of affected adoptees with those of unaffected, or con-
associations between disorders by evaluating spe- trol adoptees. The latter approach is more powerful
cific patterns of co-segregation of two or more because it eliminates the potentially confounding
disorders within families [11]. effect of environmental factors. Similar to the familial
recurrence risk, the genetic contribution in adoption
studies is estimated by comparing the risk of disease
Twin studies to biological versus adoptive relatives, or the risk of
Twin studies that compare concordance rates for disease in biological relatives of affected versus con-
monozygotic twins (who share the same genotype) trol adoptees. These estimates of risk are often
with those of dizygotic twins (who share an average adjusted for sex, age, ethnicity, and other factors that

2
Chapter 1: Contribution of genetic epidemiology

may confound the links between adoption status and Table 1.1 Risk ratios and heritability estimates for major
mental disorders.
an index disease.
With the recent trends towards selective adoption Disorder Risk Heritability
and the diminishing frequency of adoptions in the ratios estimates
United States, adoption studies are becoming less
Mood disorders
feasible methods for identifying genetic and environ- Bipolar disorder 7–10 60–70
mental sources of disease etiology [14]. However, the Major depression 2–3 28–40
increased rate of reconstituted families (families com-
prised of both siblings and half siblings) may offer a Anxiety
All 4–6 30–40
new way to evaluate the role of genetic factors in the
Panic disorder 3–8 50–60
transmission of complex disorders. Genetic models
predict that half siblings should have a 50% reduction Schizophrenia 8–10 80–84
in disease risk compared to that of full siblings. Devi- Substance 4–8 30–50
ations from this risk provide evidence for either poly- dependence
genic transmission, gene–environment interaction,
or other complex modes of transmission.
recognition of the role of biological and genetic vul-
nerability factors for psychiatric disorders has led
Migration studies to research with increasing methodological sophisti-
Migrant studies are perhaps the most powerful study cation over the course of the second half of the
design to identify environmental and cultural risk twentieth century. There are numerous comprehen-
factors. When used to study Asian immigrants to sive reviews of genetic research on specific disorders
the United States, this study design demonstrated the of interest as well as on psychiatric genetics in general
significant contribution of the environment to the [12, 17–20].
development of many forms of cancer and heart dis- Table 1.1 presents a summary of the relative risks
ease [15]. One of the earliest controlled migrant (i.e. proportion affected among first-degree relatives
studies evaluated rates of psychosis among Norwe- of affected probands versus those of relatives of
gian immigrants to Minnesota compared to native controls) derived from controlled family studies of
Minnesotans and native Norwegians [16]. A higher selected psychiatric disorders. The risk ratios compar-
rate of psychosis was found among the immigrants ing the proportion of affected relatives of cases versus
compared to both the native Minnesotans and Nor- controls are greatest for autism, bipolar disorder, and
wegians and was attributed to increased susceptibility schizophrenia; intermediate for substance depend-
to psychosis among the migrants who left Norway. ence and subtypes of anxiety, particularly panic; and
It was found that migration selection bias was the lowest for major depression. The estimates of herit-
major explanatory factor, rather than environmental ability (i.e. the proportion of variance attributable to
exposure in the new culture. The application of genetic factors) are derived from twin studies, which
migration studies to the identification of environmen- compare rates of disorders in monozygotic and dizy-
tal factors is only valid if potential bias attributed to gotic twins. These findings reinforce the notion that
selection is considered. Selection bias has been tested genes play a major role in the extent to which mental
through comparisons of factors that may influence a disorders run in families. The heritability estimates for
particular disease of interest in a migrant sample and specific disorders shown in Table 1.1 are parallel to the
a similar sample that did not migrate. risk ratios derived from family studies. Furthermore,
adoption and half-sibling studies also support a genetic
basis for the observed familial aggregation.
Genetic epidemiology of psychiatric
disorders Schizophrenia
The wealth of data from family, twin, and adoption More is known about the genetic basis of schizophre-
studies of the major psychiatric disorders exceeds that nia than perhaps any other psychiatric disorder,
of all other chronic human diseases. The increased with genetically informative studies stemming from

3
Chapter 1: Contribution of genetic epidemiology

early in the last century. There are numerous reviews genomic copy number variants (CNVs) potentially
of this extensive body of research [21–24]. Despite affecting the expression or function of genes that are
wide differences in methods, samples, and geographic relevant to brain development [38]. Of particular
locations, controlled family studies yield a remarkably interest is the velo-cardio-facial syndrome caused by
similar average relative risk of 8.9 to first-degree rela- a deletion CNV in chromosome 22q, which confers a
tives. The four-fold greater proband-wise concordance 25% risk for schizophrenia [39]. Some of the specific
rate of schizophrenia in monozygotic compared to environmental risk factors currently under investiga-
dizygotic twins, found in 12 studies to date, demon- tion include obstetric complications [40], childhood
strates the role of genetic factors in the familial aggre- trauma [41], prenatal factors such as nutritional defi-
gation of schizophrenia. The average heritability in ciencies [42], increased paternal age [43], family inter-
liability to schizophrenia across 12 studies is 0.81 [25]. actions [28], maternal infections [44], maternal
Similarly, adoption studies using traditional paradigms cytokines [45], gluten sensitivity [46], and cannabis use
and modern diagnostic criteria (if available) demon- [47, 48]. In summary, schizophrenia is now widely
strates that the average risk to first-degree relatives was viewed as a neurodevelopmental disorder comprised of
15.5% compared to 3.6% for controls, yielding a relative a confluence of vulnerability genes and environmental
risk of 4.3. exposures [49].
Despite evidence regarding the importance of gen-
etic risk factors for schizophrenia, the lack of expected
Mendelian risk ratios in the difference in risk of schizo- Mood disorders
phrenia as a function of genetic similarity suggests that A heterogeneous group of syndromes, of which major
schizophrenia is a genetically complex disorder [10]. depression and bipolar disorder (manic depression)
Recent reviews of the genetic epidemiology of schizo- are major subtypes, comprise mood disorders. Bipo-
phrenia also converge in demonstrating the multifac- lar disorder is one of the psychiatric disorders most
torial etiology of this condition [25–29]. The largest widely studied from a genetic perspective [50, 51].
and most recent cross-fostering study of schizophrenia Both major depression and bipolar disorder have
showed that adoptive family environment was associ- important genetic components. Controlled family
ated with schizophrenia spectrum disorders among studies show a five-fold risk to relatives of major
genetically vulnerable individuals [30], implying the depression, and greater than a ten-fold risk to first-
contributions of nonspecific environmental factors degree relatives of bipolar patients for developing
(i.e. multiple factors that may affect brain develop- these disorders. The concordance rate for bipolar
ment) to schizophrenia’s etiology. monozygotic twins is over five times that of dizygotic
Another important clue about potential environ- twins, and twin concordance for depression shows
mental risk factors is the increased risk for the devel- less dramatic but still notable differences. A summary
opment of schizophrenia among immigrants in of five methodologically comparable twin studies of
several different countries including East African major depression yielded an average estimate of the
immigrants to Sweden [31], Surinamese immigrants heritability of major depression of 0.37, with the
to the Netherlands [32], Afro-Caribbean immigrants remainder (0.63) nearly totally attributable to envir-
to the UK [33], Finnish immigrants to Sweden [34], onmental factors unique to the individual [50]. The
and European immigrants to Canada [35]. Although relative risks based on the few existing adoption stud-
selective migration may be one explanation, there is ies also confirm that the familial recurrence cannot
converging evidence that socially disrupted environ- be attributed solely to environmental factors [51].
ments may trigger the onset of schizophrenia in sus- The aggregate adoption study data on mood dis-
ceptible individuals. orders reveal a moderate increase in rates of mood
Children at high risk for schizophrenia (children disorders among the biological compared to adoptive
with an affected parent) show an increased incidence relatives of adoptees with mood disorders [52]. With
of numerous neurodevelopmental abnormalities as respect to bipolar disorder, there is little evidence for
compared to offspring of parents without schizophre- differential risk among biological compared to adop-
nia [36, 37]. This discrepancy has led to a focus on tive relatives of adoptees with bipolar disorder. How-
early developmental factors in the etiology of schizo- ever, the small numbers of bipolar adoptees who have
phrenia. Several recent studies have focused on been studied (i.e. less than 50) do not provide an

4
Chapter 1: Contribution of genetic epidemiology

adequate test of genetic and environmental influences. Obsessive–compulsive disorder


The most compelling finding from adoption studies, Controlled family studies of obsessive–compulsive
however, is the dramatic increase in completed suicide disorder reveal an elevated familial risk in probands
among biologic, as opposed to adoptive, relatives of with obsessive–compulsive disorder compared to con-
mood disorder probands [2, 53]. trols, with greater familial aggregation associated with
early age of onset and obsessions rather than compul-
Anxiety disorders sions [65–67]. Twin studies of obsessive–compulsive
disorder, however, have yielded only weak evidence for
At present, relatively few studies have examined heritability [54, 68, 69].
anxiety disorders from the perspective of genetic epi-
demiology, and there is virtually no data from certain
paradigms, such as adoption studies [54, 55]. How- Substance use disorders
ever, the existing research indicates that most anxiety A positive family history of a substance use disorder
disorders aggregate in families and several investiga- is a consistent and robust risk factor for substance
tions have offered specific support for genetic etiology. use in first-degree relatives (for comprehensive
reviews of alcoholism see [70–73]). Controlled family
Panic disorder studies of alcohol use disorder reveal a three-fold
Panic disorder has the strongest degree of familial increased risk of alcoholism and two-fold increased
aggregation of any of the anxiety disorder subtypes. risk of drug abuse among the relatives of probands
A review of 13 family studies of panic disorder by with alcoholism as compared to those of controls
Gorwood [56] shows a seven-fold relative risk of [74]. Both alcohol abuse and dependence appear to
panic among relatives of panic probands compared be familial among females, whereas only dependence
to controls. In addition, early-onset panic, panic asso- aggregates among the relatives of males with alcohol
ciated with childhood separation anxiety, and panic dependence [75].
associated with respiratory symptoms have each been Twin studies have demonstrated the contribution
shown to have a higher familial loading than other of both genetic and environmental risk factors to both
varieties of panic disorder [57]. Although there has alcoholism and drug abuse [76]. Heritability of alco-
been some inconsistency reported among twin studies holism (narrowly defined) has been estimated at
of panic disorder, recent studies using contemporary 59% by some researchers [77], while the heritability
diagnostic criteria show that panic disorder has the of problem drinking (using broad definitions) has
highest heritability of all anxiety disorders (44%) [58]. been estimated at 8–44% in females and 10–50% in
males [70]. Several adoption studies conducted in
Scandinavia demonstrated the importance of genetic
Phobic states
factors underlying alcoholism [78–80]. Adoption
Although there are far fewer controlled family and study paradigms have shown not only that a disturbed
twin studies of the anxiety subtypes other than panic adoptive family environment interacts with a genetic
disorder, all of the phobic states (i.e. specific phobia, predisposition for alcoholism to increase the risk for
agoraphobia) have also been shown to be familial [59, the disorder [81], but also that the adoptive family
60]. The average relative risk of phobic disorders in environment can predict alcohol abuse or dependence
the relatives of phobics is 3.1, with greater familial independent of genetic vulnerability [82]. A recent
aggregation for the generalized subtype of social “quasi-adoption” study that investigated the associ-
phobia. The heritability of phobias according to twin ation between the biological family background (gene-
studies is about 0.35% [61]. tic factors), and a history of exposure to alcoholism
during childhood (environmental factors) revealed
Generalized anxiety disorder greater effect of genetic risk factors among men than
A limited number of studies also provide evidence for among women. The study also showed common gen-
both the familial aggregation and heritability of gener- etic and environmental risk factors contributing to
alized anxiety disorder [62]. The average familial odds alcohol dependence in both men and women [83].
ratio for the disorder is approximately 5 [60, 63] and The importance of the environment as a mediating
the heritability is 0.32 among female twin pairs [64]. factor in the transmission of substance use disorders

5
Chapter 1: Contribution of genetic epidemiology

was demonstrated in a recent study of adoptive and dimensional classification of disorders further demon-
step families [84]. strates the difficulties in creating a valid classification
Although there has been less systematic research system for psychiatric disorders because of the preva-
on the familial aggregation of drug use disorders, lence of subthreshold diagnostic categories and diag-
numerous family history studies and uncontrolled nostic spectra, and the pervasive comorbidity between
and controlled family studies have demonstrated that purportedly distinct diagnostic entities; there is wide-
rates of substance use disorders are elevated among spread agreement that the categorical classification
relatives of drug abusers compared to those of con- system in psychiatry lack validity [27, 94, 95].
trols and compared to population expectations [85, The greater complexity of psychiatric disease, as
86]. One controlled family study of drug use disorders compared to other types of disease explains the con-
using contemporary family study data [75] showed an tinued reliance on the descriptive approach as the sole
eight-fold increased risk of substance use disorders basis for diagnosis in psychiatry. The difficulty in
(opioids, cocaine, cannabis, and alcohol) among rela- classifying human cognition, behavior, and emotion
tives of probands with drug disorders compared with is understandable in light of the complex psychological
relatives of people with psychiatric disorders and and physiological states underlying mental function,
normal controls. Family, twin, and adoption studies which is the culmination of human adaptation to the
have also demonstrated common genetic and envir- environment up to the current point in time. Progress
onmental factors that contribute to cannabis use dis- in neuroscience that reveals information about the
orders and other drug use disorders [87]. The results biological pathways underlying psychiatric disorders
of numerous twin studies of substance use disorders should also advance our understanding of the classifi-
in general as well as those of specific drugs have cation of psychiatric phenotypes.
shown that there are genetic factors underlying drug
abuse in general [88], as well as unique genetic factors
associated with specific drugs of abuse including nico- Complex patterns of transmission
tine and cannabis [72, 77, 89, 90]. The application of advances in genomics to mental
disorders is still limited by the complexity of the
process through which genes influence the develop-
Sources of complexity in mental ment and progression of mental disorders. There is
disorders substantial evidence that a lack of one-to-one corres-
Two factors which contribute to the complexity of the pondence between the genotype and phenotype
patterns of inheritance of psychiatric disorders are exists for most of the major psychiatric disorders.
the lack of validity of the classification of psychiatric Psychiatric disorders, like numerous other complex
disorders (e.g. phenotypes, or observable aspects of disorders for which susceptibility alleles have been
diseases) and the complexity of the pathways from geno- identified, are characterized by phenomena such as
types to psychiatric phenotypes (i.e. heterogeneity). incomplete penetrance (i.e. probability of phenotypic
expression among individuals with susceptibility
gene), variable expressivity (i.e. variation in clinical
Lack of validity of the classification system expression associated with a particular gene), gene–
The development of structured interviews has enhanced environment interaction (i.e. expression of genotype
comparability of diagnostic methods within the United only in the presence of particular environmental
States and worldwide. There is now an exciting venture exposures), pleiotropy (i.e. capacity of genes to mani-
designed to collect information on the prevalence of fest several different phenotypes simultaneously),
mental disorders using comparable diagnostic tools in genetic heterogeneity (i.e. different genes leading to
more than 30 countries under the auspices of the World indistinguishable phenotypes), gene–environment
Health Organization and Harvard University [91]. The correlation [96] and polygenic and oligogenic modes
lack of conclusive evidence for the validity of classifica- of inheritance (i.e. simultaneous contributions of mul-
tions of psychiatric disorder phenotypes, because they tiple genes rather than Mendelian single gene models)
are based solely on clinical manifestations without path- [10, 97]. Other proposed mechanisms of transmission
ognomonic markers, continues to impede advances in include mitochondrial inheritance, imprinting, and
psychiatry [92, 93]. Growing research on the epigenetic phenomena [98].

6
Chapter 1: Contribution of genetic epidemiology

Comorbidity have been identified for cancer, heart disease, dia-


betes, and other common disorders is unknown in
The high magnitude of comorbidity and co-aggrega-
the population at large. In order to provide accurate
tion of index disorders with other major psychiatric
risk estimates, the next stage of research needs to
disorders (i.e. bipolar disorder and alcoholism, major
move beyond samples identified through affected
depression and anxiety disorders, schizophrenia and
individuals to the population as a whole. Third, iden-
drug dependence), in part induced by the classification
tification of risk profiles will require large samples to
system, has been demonstrated in both clinical and
assess the significance of vulnerability genes with
community studies [11, 86, 99, 100]. For example,
relatively low expected population frequencies. Fourth,
alcoholism, a well-established complication of bipolar
similar to the role of epidemiology in quantifying risk
illness, may mask the underlying features of bipolarity,
associated with traditional disease risk factors, applica-
leading to phenotypic misclassification in genetic
tions of human genome epidemiology can provide
studies [101]. Nonrandom mating is also a common
information on the specificity, sensitivity, and impact
phenomenon in mental disorders that impedes evalu-
of genetic tests to inform science and individuals [105].
ation of patterns of familial transmission [102].
Below we review the role of the tools of epidemiology
Assortative mating is particularly pronounced for
in ongoing and future studies designed to identify
substance use disorders for which substance depend-
genes underlying mental disorders.
ence among spouses of substance dependent probands
may be as high as 90% [103].
These phenomena serve to decrease the signal to Samples
noise ratio in defining psychiatric disorders for gen- The shift from systematic large-scale family studies
etic studies. Studies that attempt to identify the to linkage studies in psychiatry has led to the collec-
impact of these phenomena on phenotypic expression tion of families according to very specific sampling
in individuals and families will bring us closer to strategies (e.g. many affected relatives, affected sib-
understanding the role of the underlying genes on ling pairs, affected relatives on one side of the family
the components of psychiatric disorders. only, availability of parents for study, etc.) in order
to maximize the power of detecting genes according
Applications of genetic epidemiology to the assumed model of familial transmission. Des-
pite the increase in power for detecting genes, these
to gene identification sampling approaches have diminished the general-
There is a widespread consensus among geneticists izability of the study findings, and contribute little
and epidemiologists on the importance of epidemi- else to the knowledge base if genes are not dis-
ology to the future of genetics and on the conclusion covered. Recent genome-wide association studies of
that the best strategy for susceptibility risk factor psychiatric disorders have included probands from
identification for common and complex disorders will families previously collected for linkage studies and
ultimately involve large epidemiological studies from single cases collected more recently from hospital
diverse populations [73, 104–108]. It is likely that admissions, almost all of self-reported European
population-based association studies will assume descent. Future studies will attempt to collect both
increasing importance in translating the products of families and controls from representative samples of
genomics to public health. There are several reasons the population so that results can be used to esti-
that population-based studies are critical to current mate population risk parameters, to examine the
studies seeking to identify genes underlying psychi- specificity of endophenotypic transmission and so
atric disorders. First, the frequency of newly identified results can be generalized to whole populations.
polymorphisms, whether single nucleotide poly-
morphisms (SNPs) or other variants such as copy
number variation (CNVs), especially in particular Selection of controls
population subgroups, is not known. Second, current The most serious problem in the design of association
knowledge of genes as risk factors is based nearly studies is the difficulty of selecting controls that are
exclusively on clinical and nonsystematic samples. comparable to the cases on all factors except the
Hence, the significance of the susceptibility alleles that disease of interest [109, 110]. Controls should be

7
Chapter 1: Contribution of genetic epidemiology

drawn from the same population as cases, and must population attributable risk relates to the risk of a
have the same probability of exposure (i.e. genes) disease in a total population (exposed and unex-
as cases. Controls should be selected to ensure the posed) and indicates the amount the disease can
validity rather than the representativeness of a study. be reduced in a population if an exposure is elim-
Failure to equate cases and controls may lead to inated. The population attributable risk depends on
confounding (i.e. a spurious association due to an the prevalence of the exposure, or in the case of risk
unmeasured factor that is associated with both the alleles, the allele frequency. Genetic attributable risk
candidate gene and the disease). In genetic case- would indicate the proportion of a particular dis-
control studies, the most likely source of confounding ease that would be eliminated if a particular gene or
is ethnicity because of differential gene and disease genes were not involved in the disease. For example,
frequencies in different ethnic subgroups. Recent the two vulnerability alleles for Alzheimer’s disease
genome-wide association studies of psychiatric dis- include the very rare, but deterministic alleles in the
orders have included control samples recruited from b-amyloid precursor, presenilin-1, and –2 genes,
the general population using self-administered psy- which signal a very high probability of the develop-
chiatric screens and from blood bank samples that ment of Alzheimer’s disease, particularly at a young
exclude donors reporting major psychiatric diagnoses age, and the susceptibility allele ε4 in the apolipo-
or taking psychiatric medications. The matching of protein-E gene (APOE ε4) [112]. The apolipopro-
controls to cases on ethnic background is largely tein-Ε ε4 (APOE ε4) allele has been shown to
based on self-report; several methods are used to increase the risk of Alzheimer’s disease in a dose-
screen for and exclude subjects with substantial dependent fashion. Using data from a large multi-
differences in ancestry. ethnic sample collected by more than 40 research
teams, Farrer [113] reported a 2.6–3.2 greater odds
of Alzheimer’s disease among those with one copy,
Risk estimation and 14.9 odds of Alzheimer’s disease among those
Because genetic polymorphisms involved in complex with two copies of the APOE ε4 allele. Moreover,
diseases are likely to be nondeterministic (i.e. the there was a significant protective effect among those
marker neither predicts disease nor nondisease with with the ε2/ε3 genotype. As opposed to the deter-
certainty), traditional epidemiological risk factor ministic mutations, the APOE ε4 allele has a very
designs can be used to estimate the impact of these high population attributable risk because of its high
genetic polymorphisms. Increased attention to alleles frequency in the population. The APOE ε4 allele is
as a part of risk equations in epidemiology will likely likely to interact with environmental risk
resolve the contradictory findings from studies that and protective factors [114, 115]. The population
have generally employed solely environmental risk risk attributable to these mutations is quite low
factors, such as diet, smoking, alcohol use, etc. Like- because of the very low population prevalence of
wise, the studies that seek solely to identify small risk disease associated with these alleles. This model of
alleles will continue to be inconsistent because they combination of several rare deterministic alleles in a
do not consider the effects of nongenetic biological small subset of families and common alleles with
parameters or environmental factors that contribute lower relative risk to individuals but high popula-
to the diseases of interest. tion attributable risk is likely to apply to many of
There are several types of risk estimates that are the psychiatric disorders as well, and may in part
used in public health. The most common is relative explain some of the discrepancies in findings across
risk, defined as the magnitude of the association studies to date.
between an exposure and disease. It is independent Recent genome-wide association studies have
of the prevalence of the exposure. The absolute risk uncovered risk alleles associated with coronary artery
is the overall probability of developing a disease in disease, Crohn’s disease, rheumatoid arthritis, type 1
an individual or in a particular population [111]. and type 2 diabetes [116], and schizophrenia. Those
The attributable risk is the difference in the risk of genetic variants appear to confer only modest increases
the disease in those exposed to a particular risk in disease risk (odd ratios [ORs] between 1.2 and 1.5)
factor compared to the background risk of a disease compared with other established risk factors for
in a population (i.e. in the unexposed). The common chronic diseases.

8
Chapter 1: Contribution of genetic epidemiology

Use of endophenotypes for classification [129], substance use disorders [71, 130] and the pro-
tective influence of a deletion in the CCR5 gene
Numerous studies have begun to deconstruct psychi-
on exposure to the human immunodeficiency virus
atric phenotypes by their component features or sub-
(HIV) [131].
types including bipolar disorder [117, 118], general
In prospective studies, however, few environmen-
anxiety disorder [119], obsessive–compulsive dis-
tal exposures have been shown to have an etiological
order [120], schizophrenia [121], and panic disorder
role in psychiatric disorders [132]. Over the next
[122]. Identification of phenotypic traits or markers,
decades, it will be important to identify and evaluate
which are themselves heritable, and which may repre-
the effects of specific environmental factors on disease
sent intermediate forms of expression between the
outcomes and to refine measurement of environ-
output of underlying genes and the broader disease
mental exposures to evaluate specificity of effects.
phenotype, have been termed “endophenotypes”
Study designs and statistical methods should focus
[123]. Studies of the role of genetic factors involved
increasingly on gene–environment interaction [106,
in these systems may be more informative than stud-
133, 134]. Although numerous recent studies have
ies of the aggregate psychiatric phenotypes since they
reported gene–environment interaction between sev-
may more closely represent the expression of under-
eral genes that interact with nonspecific environmen-
lying biological systems. To the extent that particular
tal exposures such as life stress or childhood adversity
endophenotypes more clearly represent expression of
and a range of outcomes including depression, can-
genotypes, they may help to unravel the complexity of
nabis dependence, and conduct disorder [135], repli-
transmission of the mental disorders. For example,
cation of these findings has not been consistent [136].
some of the endophenotypes that may underlie mood
Increased knowledge of the developmental pathways
disorders include circadian rhythm, stress reactivity, and
of emotion, cognition, and behavior will expand our
mood, sleep and appetite regulation [95]. Cognitive,
ability to identify specific environmental factors such
neurophysiologic, and structural measures continue
as infection, poor diet, prenatal environment, and
to be tested as potential schizophrenia endopheno-
early life experiences that interact with the genetic
types [124, 125]. However, before applying endophe-
architecture of mood regulation and cognition [137].
notypes in gene identification studies, there should
be evidence that the endophenotype has a stronger
genetic signal than the broader phenotype. A recent
meta-analysis of psychiatric endophenotypes [126] Impact of genomics on psychiatric
and a review of the genetic architecture of traits in science and practice
model organisms do not provide evidence that endo-
Progress in genomics has far outstripped advances
phenotypes are superior to current phenotypic disease
in our understanding of psychiatric disorders and
definitions [127].
their etiologies. Technical advances and availability
of rapidly expanding genetic databases provide extra-
ordinary opportunities for understanding disease
Identification of environmental factors pathogenesis. However, the application of psychiatric
In parallel with the identification of susceptibility genetic research to study diagnostic heterogeneity,
alleles, it is important to identify environmental course and/or treatment outcome is still limited due
factors that operate either specifically or nonspecifi- to the lack of consistent genetic findings to date. Over
cally on those with susceptibility to psychiatric dis- the next decade increasing understanding of the com-
orders in order to develop effective prevention and plex mechanisms through which genetic risk factors
intervention efforts. Langholz et al. [128] describes influence disease should enhance the clinical utility of
some of the world’s prospective cohort studies that psychiatric genetics.
may serve as a basis for studies of gene–disease asso- The goal of genomics research is ultimately
ciations or gene–environment interactions. There prevention, the cornerstone of public health. An
is increasing evidence that gene–environment inter- understanding of the significance of genetic risk
action will underlie many of the complex human factors and proper interpretations of their meaning
diseases. Some examples include inborn errors of for patients and their families will ultimately become
metabolism, individual variation in response to drugs part of clinical practice. Clinicians will become

9
Chapter 1: Contribution of genetic epidemiology

increasingly involved in helping patients to compre- upon clinicians to become familiar with knowledge
hend the meaning and potential impact of genetic risk gleaned from genetic epidemiological and genomics
for both psychiatric and nonpsychiatric disorders. research. In the meanwhile, use of recurrence risk
As our knowledge of the role of genetic risk factors estimates from family studies best predicts the risk
in psychiatric disorders advances, it will be incumbent of the development of mental disorders.

References 16. Ödegaard Ö. Emigration and


Insanity: A Study of Mental
34. Leao TS, et al. J Nerv Ment Dis
2006;194(1):27–33.
1. Heston LL. Br J Psychiatry 1966; Disorders among the Norwegian
112(489):819–825. 35. Smith GN, et al. Schizophr Res
Born Population of Minnesota. 2006;87(1–3):205–211.
2. Wender PH, et al. Arch Gen Copenhagen: Levin and
Psychiatry 1986;43(10):923–929. Munksgaards; 1932. 36. Owens DG, et al. Psychol Med
2006;36(11):1501–1514.
3. Wender PH. Annu Rev Med 17. Merikangas K, et al. Epidemiol Rev
1972;23:355–371. 1996;19:1–12. 37. Brewer WJ, et al. Schizophr Bull
2006;32(3):538–555.
4. Kety SS, et al. Schizophr Bull 18. Kendler KS. Arch Gen Psychiatry
1995;52(11):895–899. 38. Cook EH, Jr., et al. Nature
1976;2(3):413–428.
2008;455(7215):919–923.
5. Kety SS, et al. Genetic Relationships 19. Tandon K, et al. Eur J Neurosci
2002;16(3):403–407. 39. Bassett AS, et al. Hum Mol Genet
within the Schizophrenia Spectrum:
2008;17(24):4045–4053.
Evidence from Adoption Studies. 20. Tsuang MT, et al. Methods Mol
New York: Raven Press; 1978. Med 2003;77:251–265. 40. Clarke MC, et al. Schizophr Bull
2006;32(1):3–8.
6. Gershon ES, et al. (eds.). Genetic 21. Zerbin-Rudin E. Nervenarzt
Approaches to Mental Disorders. 1980;51(7):379–391. 41. Morgan C, et al. Schizophr Bull
New York: American 2007;33(1):3–10.
22. Gottesman, II. Acta Psychiatr
Psychopathological Association Scand Suppl 1994;384:26–33. 42. Ludvigsson JF, et al. Scand
Press; 1994. J Gastroenterol 2007;49:179.
23. Kendler KS, et al. Schizophr Bull
7. Beaty TH, et al. Epidemiol Rev 1993;19(2):261–285. 43. Malaspina D, et al. CNS Spectr
2000;22(1):120–125. 2002;7(1):26–29.
24. Sullivan PF. PLoS Med 2005;
8. MacMahon B, et al. Epidemiology: 2(7):e212. 44. Buka SL, et al. Arch Gen Psychiatry
Principles and Methods. Boston: 2001;58(11):1032–1037.
Little Brown and Company; 1996. 25. Sullivan PF, et al. Arch Gen
Psychiatry 2003;60(12): 45. Buka SL, et al. Brain Behav
9. Kuller LH. Prog Clin Biol Res 1187–1192. Immun 2001;15(4):411–420.
1979;32:489–495. 46. Kalaydjian AE, et al. Acta Psychiatr
26. Hallmayer J. Aust N Z J Psychiatry
10. Risch N. Am J Hum Genet 1990; 2000;34 Suppl:S47–55; discussion Scand 2006;113(2):82–90.
46(2):222–228. S56–47. 47. Arseneault L, et al. BMJ 2002;325
11. Merikangas KR. Comorbity for 27. Tsuang MT. Am J Med Genet (7374):1212–1213.
Anxiety and Depression: Review of 2001;105(1):8–10. 48. Dean K, et al. Dialogues Clin
Family and Genetic Studies. Neurosci 2005;7(1):69–80.
Washington, DC: American 28. McGuffin P. Psychiatr
Psychiatric Press; 1990. Prax 2004;31 Suppl 2: 49. Dealberto MJ. Med Hypotheses
S189–193. 2007;68(2):259–267.
12. Shih RA, et al. Int Rev Psychiatry
2004;16(4):260–283. 29. Portin P, et al. Acta Psychiatr 50. Sullivan PF, et al. Am J Psychiatry
Scand 1997;95(1):1–5. 2000;157(10):1552–1562.
13. Kendler KS, et al. Psychol Med
2007;37(5):615–626. 30. Tienari P, et al. Br J Psychiatry 51. Merikangas KR, et al. Clin
2004;184:216–222. Neurosci Res 2002;2:127.
14. National Adoption Information
Clearinghouse. Families 2007 31. Selten JP, et al. Am J Psychiatry 52. Faraone SV, et al. Psychol Bull
Edition. Washington, DC: US 2002;159(4):669–671. 1990;108(1):109–127.
Department of Health and Social 32. Hanoeman M, et al. Schizophr Res 53. Mendlewicz J, et al. Nature
Services; 2007. 2002;54(3):219–221. 1977;268(5618):327–329.
15. Kolonel LN, et al. Nat Rev Cancer 33. Cooper B. Epidemiol Psichiatr Soc 54. Carey G, et al. In Rabkin J (ed.).
2004;4(7):519–527. 2005;14(3):137–144. Anxiety: New Research and

10
Chapter 1: Contribution of genetic epidemiology

Changing Concepts. Vol. 117. 76. Kendler KS, et al. Am J Psychiatry 100. Maier W, et al. Eur Arch
New York: Raven Press; 1981. 1994;151(5):707–715. Psychiatry Clin Neurosci 1993;
55. Hettema JM, et al. Am J Psychiatry 77. Kendler KS, et al. Arch Gen 243(3–4):205–211.
2001;158(10):1568–1578. Psychiatry 2000;57(3):261–269. 101. Merikangas KR, et al.
56. Gorwood P, et al. Encephale 78. Cloninger CR, et al. Recent Dev Arch Gen Psychiatry
1999;25(1):21–29. Alcohol 1985;3:37–51. 2007;64(5):543–552.
57. Goldstein RB, et al. Arch 79. Sigvardsson S, et al. Arch Gen 102. Merikangas KR. Arch Gen
Gen Psychiatry 1997;54(3): Psychiatry 1996;53(8):681–687. Psychiatry 1982;39(10):1173–1180.
271–278. 80. Goodwin DW. Arch Gen 103. Galbaud du Fort G, et al. Psychol
58. Kendler KS, et al. Psychol Med Psychiatry 1985;42(2): Med 1998;28(4):789–802.
2001;31(6):989–1000. 171–174. 104. Peltonen L, et al. Science 2001;
59. Fyer AJ, et al. Arch Gen Psychiatry 81. Cutrona CE, et al. Compr 291(5507):1224–1229.
1995;52(7):564–573. Psychiatry 1994;35(3):171–179.
105. Khoury MJ, et al. Am J Epidemiol
60. Noyes R, Jr., et al. Am J Psychiatry 82. Cadoret RJ, et al. Arch Gen 2000;151(1):2–3.
1987;144(8):1019–1024. Psychiatry 1995;52(1):42–52.
106. Yang Q, et al. Epidemiol Rev
61. Kendler KS, et al. Arch Gen 83. Light JM, et al. J Stud Alcohol 1997;19(1):33–43.
Psychiatry 1992;49(4):273–281. 1996;57(5):507–520.
107. Merikangas KR, et al. Science
62. Newman SC, et al. Psychol Med 84. Newlin DB, et al. Alcohol Clin Exp 2003;302(5645):599–601.
2006;36(9):1275–1281. Res 2000;24(12):1785–1794.
108. Risch N. Genet Epidemiol
63. Mendlewicz J, et al. Psych Genet 85. Bierut LJ, et al. Arch Gen 1990;7(1):3–16; discussion
1993;3(2):73–78. Psychiatry 1998;55(11):982–988. 17–45.
86. Merikangas KR, et al. Psychol Med
64. Kendler KS, et al. Arch Gen 109. Wacholder S, et al. J Natl Cancer
1998;28(4):773–788.
Psychiatry 1992;49(4):267–272. Inst 2000;92(14):1151–1158.
87. Agrawal A, et al. Addiction
65. Nestadt G, et al. Arch Gen 110. Ott J. Neurology 2004;63(6):
2006;101(6):801–812.
Psychiatry 2000;57(4):358–363. 955–958.
88. Kendler KS, et al. Acta Psychiatr
66. Grabe HJ, et al. Am J Psychiatry 111. Gordis L. Epidemiology.
Scand 1999;99(5):368–376.
2006;163(11):1986–1992. Philadelphia: W B Saunders;
89. Kendler KS, et al. Am J Psychiatry 2000.
67. Pauls DL, et al. Am J Psychiatry 1998;155(8):1016–1022.
1995;152(1):76–84. 112. Tol J, et al. Rev Neurol (Paris)
90. Kendler KS, et al. Br J Psychiatry 1999;155(Suppl 4):S10–S16.
68. Bellodi L, et al. Psychiatry Res 1998;173:345–350.
1992;42(2):111–120. 113. Farrer LA, et al. JAMA 1997;
91. Kessler R, et al. Hosp Manage 278(16):1349–1356.
69. Lenane MC, et al. J Am Acad Child Internat 2000:195–196.
Adolesc Psychiatry 1990;29 114. Kivipelto M, et al. Neurology
(3):407–412. 92. Kendell RE. Psychol Med 1989; 2001;56(12):1683–1689.
19(1):45–55.
70. Heath AC, et al. Psychol Med 115. Kivipelto M, et al. Ann Intern
1997;27(6):1381–1396. 93. Angst J. World Psychiatry 2007; Med. 2002;137(3):149–155.
6(2):94–95.
71. Heath AC, et al. Br J Psychiatry 116. Wellcome Trust Case Control
Suppl 2001;40:S33–S40. 94. Angst J, et al. J Affect Disord Consortium. Nature 2007;
72. Tsuang MT, et al. Harv Rev
2003;73(1–2):133–146. 447(7145):661–678.
Psychiatry 2001;9(6):267–279. 95. Lenox RH, et al. Am J Med Genet 117. Benazzi F. Lancet 2007;
73. Merikangas KR. In D’haenen
2002;114(4):391–406. 369(9565):935–945.
HAH, et al. Biological Psychiatry. 96. Dick DM, et al. Ann Clin 118. Angst J. Br J Psychiatry
Vol. 2. Chichester: John Wiley and Psychiatry 2006;18(4):223–231. 1997;190:189.
Sons; 2002. 97. Gottesman II, et al. Hum Hered 119. Angst J, et al. Psychol Med 2006;
74. Nurnberger JI, Jr., et al. Arch Gen 1971;21(6):517–522. 36(9):1283–1292.
Psychiatry 2004;61(12):1246–1256. 98. Guttmacher AE, et al. N Engl 120. Eapen V, et al. J Psychosom Res
75. Merikangas KR, et al. J Med 2002;347(19):1512–1520. 2006;61(3):359–364.
Arch Gen Psychiatry 99. Maier W, et al. Schizophr Res 121. Braff DL, et al. Schizophr Bull
1998;55(11):973–979. 2002;57(2–3):259–266. 2007;33(1):21–32.

11
Chapter 1: Contribution of genetic epidemiology

122. Smoller JW, et al. Am J Psychiatry 128. Langholz B, et al. J Natl Cancer 133. Ottman R. Genet Epidemiol 1990;
1998;155(9):1152–1162. Inst Monogr 1999(26):39–42. 7(3):177–185.
123. Gottesman II, et al. Am 129. Nebert DW. Clin Genet 1999; 134. Beaty TH. Epidemiol Rev 1997;
J Psychiatry 2003;160(4):636–645. 56(4):247–258. 19(1):14–23.
124. Horan WP, et al. Schizophr Res 130. Rose RJ, et al. Alcohol Clin Exp Res 135. Caspi A, et al. Science 2003;
2008;103(1–3):218–228. 2000;25:637. 301(5631):386–389.
125. Radant AD, et al. Schizophr Res 131. Michael NL. Curr Opin Immunol 136. Zammit S, et al. Br J Psychiatry
2007;89(1–3):320–329. 1999;11(4):466–474. 2006;188:199–201.
126. Flint J, et al. Psychol Med 2007; 132. Eaton WW. Risk Factors for 137. Meaney MJ. Annu Rev Neurosci
37(2):163–180. Mental Health Disorders. 2001;24:1161–1192.
127. Valdar W, et al. Nat Genet Bethesda, MD: National Institute
2006;38(8):879–887. of Mental Health; 2004.

12
A basic overview of contemporary human
Chapter

2 genetic analysis strategies


Ondrej Libiger and Nicholas J. Schork

Abstract only a few people or even a single individual. The types


of variation that populate the human genome and
Human genetics research has received considerable differentiate disease susceptible from nondisease sus-
recent attention as a result of major developments in ceptible individuals range from simple single nucleot-
molecular genetic technologies, large-scale coordin- ide polymorphisms (SNPs) that involve a difference in
ated research initiatives such as the Human Genome a single nucleotide, to large stretches of sequence that
Project and the International HapMap Project, and are deleted or even repeated some number of times [1].
major discoveries concerning the genetic basis of dis- If these variants influence gene function in important
ease. The strategies that human geneticists exploit ways they could perturb normal physiological function
to identify DNA sequence variations that influence and lead to disease.
disease susceptibility leverage an understanding of the Identifying the specific variants that influence
behavior of sequence variations when transmitted particular diseases is not trivial given the total
from parents to offspring across generations either number of variations that might be responsible for
within a specific family, among a set of families, or those diseases. However, as daunting as the task of
in the population at large. In this brief review we sifting through the tens of millions of nucleotides that
describe the basic principles behind the most widely a group of susceptible individuals possess may seem,
used human genetic strategies to identify inherited this task is made possible through the exploitation of
disease susceptibility factors. We also point out the a few fundamental genetic phenomena: segregation,
limitations and issues that plague these strategies as linkage, linkage disequilibrium (LD), and (causal)
well as areas for further study. gene perturbation. In this brief review we describe
these basic phenomena and how they can be leveraged
Introduction to “map” the locations of disease susceptibility vari-
Human genetic research has a long and illustrious ants in family and population-based genetic studies.
history, but has received considerable recent attention We also describe the problems that plague genetic
as a result of major developments in DNA sequencing approaches to susceptibility variant identification as
and genotyping technologies as well as the application well as human genetic research areas that will likely
of those technologies to the identification of inherited receive greater attention in the future.
DNA sequence variations (or ‘variants’) that contribute
to disease susceptibility. A human haploid genome
is over 3 billion nucleotides in length and because Establishing heritability
humans are diploid (i.e. inherit one of each of the Prior to embarking on a study to identify specific
22 pairs of autosomal chromosomes as well as the sex genetic variants that influence susceptibility to a
chromosomes comprising the complete 3 billion nucle- particular disease or trait, many researchers try to
otide genome from a mother and a father) they actually establish the degree to which the disease or trait is
possess over 6 billion nucleotides. A large number of “heritable” or attributable to inherited factors.
these nucleotides (10–20 million) vary frequently Although there are many strategies for estimating
from person to person and other nucleotides differ in the overall genetic contribution to a disease or trait,

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

13
Chapter 2: Overview of human genetic analysis strategies

virtually all of them rest on the principle that individ- identifying which subset of those variants is respon-
uals that are more closely related – and hence share sible for a particular phenotype is nontrivial. The fact
more genes – should exhibit greater phenotypic simi- that these variants differ in frequency both within
larity with respect to a trait of interest than more and across different populations further complicates
distantly related individuals, if genetic factors do, in things. However, the fundamental phenomenon
fact, influence that trait. For example, if one were to of recombination during meiosis can facilitate the
collect data on pairs of individuals who had varying search for variants influencing causative variants
degrees of biological relatedness (e.g. monozygotic for a disease. During meiosis homologous chromo-
[MZ] twins, dizygotic [DZ] twins, cousin pairs, second somes (i.e. the chromosomes pairs, one inherited
cousin pairs, etc.), the concordance rates for the dis- from the mother and one from the father) pair up
ease or trait of interest should be higher for the pairs and exchange chromosomal segments (i.e. ‘‘recom-
that were more closely related, if the trait had a bine’’) before a single recombined chromosome
genetic component. From an analysis of the correl- from every chromosome pair arises, contributing to
ation between the degree of relatedness and degree a haploid ‘‘gamete” which, when combined with a
of phenotypic similarity, one can estimate the degree gamete resulting from the same process from the
to which a trait is heritable [2, 3]. other parent, leads to fertilization and a resulting
Heritability is often estimated by comparing MZ diploid embryo. Recombination results in sequence
twins, whose genomes are identical, to DZ twins, variants neighboring each other on a single parental
whose genomes, like all sibling pairs, share on average chromosome to be inherited together except at the
half of their content. Depending on how divergent sites of recombination (i.e. the “breakpoints”). The
the concordance rates among MZ and DZ twins are, closer two positions in the DNA sequence (referred
one can infer that the trait has a heritability that is low to as “loci”) harboring variant alleles are to each
or high [2–4]. However, there are many assumptions other, the more likely they will be transmitted
behind heritability estimates, including those based together, since it is unlikely that a recombination
on twins [4] that are nontrivial to deal with. One of breakpoint will occur between them (on average there
the most vexing has to do with distinguishing the are only 1–2 recombination events that occur on any
effects of shared environment from shared genes, chromosome although there are likely to be more
since individuals that are more closely related tend recombinations on larger chromosomes).
to live in similar environments. Two loci are said to be “linked” if they are close
Many ingenious strategies have been devised to enough in proximity such that any (parental) chromo-
help disentangle shared environment from shared gen- some harboring specific nucleotides or variants at
etic influences on a trait. For example, migration studies those loci will be transmitted on a single gamete
consider the differences in rates of a particular disease more often than expected if they were not near each
among individuals living in one environment to rates other; i.e. if two loci are on different chromosomes or
achieved after those individuals have moved to a differ- very far from each other such that recombination
ent environment [5, 6]. If the rates change substantially, breakpoints are likely to occur between them, then
then a major environmental component to the disease the nucleotides or variants at those loci will be trans-
can be inferred. More compelling strategies involve the mitted or “segregate” independently of one another.
study of twin pairs in which one of the twins was Consider the members of the family depicted
brought up in a different environment (say after adop- in Figure 2.1. The mother, denoted as person X-1,
tion) than the other twin. The use of combinations of has a stretch of DNA sequence with the nucleotides
twins, some reared apart and some together, can lead to A-C-G-G-G on one of her chromosomes and on the
more precise estimates of the genetic contribution to homologous chromosome the nucleotides A-A-G-G-C.
a trait, but will never be free from concerns about the The father, Y-1, has the nucleotides A-C-G-C-G at the
context-dependency of the findings [7]. same positions as the mother on one of his chromo-
somes and on his homologous chromosome has the
nucleotides G-A-T-C-C. The nucleotide T at the third
Genetic variants: linkage and LD position is a dominant acting (i.e. only one copy of
Recombination and linkage. As noted, there are so many the variant is needed in order to get the disease, as
genetic variants that populate the human genome that opposed to “recessive” situations in which two copies

14
Chapter 2: Overview of human genetic analysis strategies

A A A G Figure 2.1 A nuclear family with five


C A C A offspring. The father, individual Y-1,
G G G T+ possesses a “T” allele on a haplotype,
G G C C G-A-T-C-C, which causes a disease,
G C G C denoted by the plus sign. Three of the
offspring inherited the “T” variant (i.e.
X-1 Y-1 individuals C-1, C-3, and C-4) and hence
have the disease but only one, C-4,
inherited the entire “G-A-T-C-C”
haplotype due to recombination
(instances of which are denoted by
the light shading).

A A A A A G A G A A
C C A C A A C A A C
G T+ G G G T+ G T+ G G
G C G C G C G C G C
G C C C C C G C C G
C-1 C-2 C-3 C-4 C-5

are needed, and “additive” or “co-dominant” situ- recombination event not occurred in the formation
ations in which having two copies is more likely to of the gamete contributing to individual C-1.
result in disease than having only one copy) and LD and variant segregation in populations. Figure
highly penetrant variant (i.e. the probability of getting 2.1 provided a simple example of the impact of
the disease given that one has the variant is close recombination on the transmission of variant nucle-
to 1.0, as opposed to situations involving “reduced otides possessed by a mother and father to their
penetrance” in which this probability is much less offspring over a single generation. Figure 2.2 depicts
than 1.0) that causes a disease, and hence the father a hypothetical situation in which a large number
has the disease (denoted by the “+” by the T allele or of generations are considered over which particular
variant). Note that the mother has a homozygous AA variants have been transmitted. Individual A-1
genotype at the first locus, the heterozygous CA geno- possesses a haplotype with nucleotides G-A-T+-C-C.
type at the second locus, the homozygous GG genotype The “+” sign next to the T variant again denotes a
at the third locus, the homozygous GG genotype at the disease causing variant. Note that unlike Figure 2.1,
fourth locus, and the heterozygous GC genotype at both homologous chromosomal haplotypes are not
the fifth locus. depicted since focus is on the transmission of the
In Figure 2.1, the parents have five children and “ancestral” G-A-T+-C-C nucleotides from the
three have been transmitted the T allele and hence “founder” individual A-1. Each broken line denotes
have the disease (i.e. the children denoted C-1, C-3, a line of descent between individuals in the latest
and C-4). Note also that affected children C-3 and generation (e.g. individuals starting with the letters
C-4 have inherited the entire paternal haplotype G-A-T- E, F, G, H, and I) and ancestors in previous gener-
C-C harboring the disease causing allele T. Individual ations. The actual number of generations separating
C-1 inherited the “T-C-C” component of this paternal the individuals is arbitrary. Individual C-1 is a des-
haplotype, but because of a recombination event in the cendent of individual A-1 but, because of a recombin-
meiosis leading up to the formation of the gamete ation event in one of his parents (who was also
contributing to the fertilized embryo of individual a descendent of individual A-1), was transmitted
C-1, the “G-A” variants were recombined with the only variants T+-C-C from the original G-A-T+-C-
“A-G” nucleotides on the paternal homolog. Thus, C haplotype possessed by individual A-1. Thus, all the
the affected offspring not only share the T allele descendants of individual C-1 who receive the T+ dis-
inherited from their father, but also the “C-C” alleles ease causing variant (i.e. individuals E-1 to E-N) will
at loci linked to the T allele and could have also be transmitted, barring further recombination or
all inherited the neighboring “G-A” variants had a mutation events, the A-C-T+-C-C haplotype rather

15
Chapter 2: Overview of human genetic analysis strategies

G
A
T+
C
C
A
A-1 C
+T
C
G

A B-1
C
T+ G
C A
C T+
G
C-1 C

D-1

A A A A G G G G G G G G G G A A G
C C C C A A A A A A A A A A C C C
T+ T T+ T+ T+ T+ T+ T T T+ T+ T T+ T+ +T +T G+
C C C C C C C C C C G G G G C C C
C C C C C C C C C C C C C C G G C
E-1 E-2 E-3 E-N F-1 F-2 F-3 F-4 F-5 F-N G-1 G-2 G-3 G-N H-1 H-N I-1
Figure 2.2 Chromosomal haplotypes and their histories back some number of generations. Each dashed line represents a line of descent
from a “founder” individual who introduced the chromosomal segment into the population. Only one homologous chromosome is drawn for
each person for simplicity and mate chromosomes are also not depicted. A “+” denotes the presence of a disease phenotype for the individual
possessing the haplotye. Generations and individuals are labeled as groups A-I, and individuals within those groups numbered 1-N.

than the G-A-T+-C-C haplotype possessed by the Individual B-1 developed a T nucleotide mutation
original founder A-1. at position 3 in the sequence de novo (i.e. independ-
Individual D-1 is also a descendant of individual ently of the event introducing the mutation to indi-
A-1. Because of a mutation event changing the fourth vidual A-1). Individual B-1’s disease causing mutation
nucleotide, C, to a G nucleotide (denoted by the is denoted +T to distinguish it from individual A-1’s
underline) this individual possesses a G-A-T+-G-C T+ mutation. Individual B-1 had haplotype A-C-+T-
haplotype. Descendants of individual D-1 who receive C-G. All descendents of individual B-1, who inherit
the T+ disease causing variant (i.e. individuals G-1 to the +T variant would, again barring recombination
G-N) will be transmitted, barring further recombin- and mutation events, also inherit the A-C-+T-C-G
ation or mutation events, the G-A-T+-G-C haplotype haplotype (individuals H-1 to H-N). Finally, individ-
rather than the G-A-T+-C-C haplotype. Note that ual I-1 does manifest the disease but does not carry
other descendants of individual A-1 received the the T variant at the third position. This could reflect
G-A-T+-C-C haplotype in tact (i.e. individuals F-1 “allelic” or “locus” heterogeneity in which different
to F-N). Also note that because of reduced or “incom- variants can cause a disease. Individual I-1 could be
plete” penetrance of the T disease-causing nucleotide, referred to as a “phenocopy” relative to individuals
some individuals with the T nucleotide do not mani- with the disease who possess the T variant.
fest the disease (e.g. individuals E-2, F-4, F-5, and Ultimately, Figure 2.2 suggests that there are
G-2). Environmental factors or additional “protective” groups of individuals in the latest generation who
variants may explain why the individuals carrying the have the disease for different reasons, some individ-
T variant do not manifest the disease. uals harboring a disease causing variant do not

16
Chapter 2: Overview of human genetic analysis strategies

manifest the disease due to incomplete penetrance of phenocopies. These statistical analysis methodologies
the T variant, and among the individuals who have are beyond the scope of this chapter; however, the
the disease because they possess the T variant, there reader is referred to excellent books on the statistical
are clusters of individuals who are descendents of methods behind them as well as the results of many
particular individuals who have different haplotypes family-based linkage studies [9, 10].
or nucleotides surrounding the T variant. LD mapping. A drawback of the linkage strategy
described above is that one needs families to embark
on the analysis. Population-based LD mapping does
Linkage and LD mapping not require families, but does have a few drawbacks of
Linkage analysis. Linkage analysis is a statistical its own. Consider Figure 2.2, Although individuals
method for identifying chromosomal regions in families E, F, and G who have the disease because
harboring variants that influence a particular pheno- they possess the T+ variant do not share a common
type. It works by examining the consistency with haplotype of which the T+ variant is a component,
which certain variants are transmitted from parents they do all share the C variant at the fifth position due
to offspring within different families. Consider Figure to the historical haplotype transmitted to them over
2.2 and families defined by individuals E-1 to E-N the generations from individual A-1. Thus, if a
(the “E” family), individuals F-1 to F-N (the “F” researcher was to genotype all E, F, and G affected
family), individuals G-1 to G-N (the “G” family), individuals at the fifth position he or she would see
and individuals H-1 to H-N (the “H” family). The that they share a C variant. If unaffected individuals
individuals within each family with the disease share a did not possess this variant (or at least not with the
certain set of variants. For the E family these variants same frequency as affected individuals) then the
are A-C-T+-C-C and for the G family these variants researcher might infer that the C variant was in LD
are G-A-T+-G-C. Note that many of the variants with a disease causing variant. He or she could then
shared within each family are different from family genotype the affected individuals at other variant sites
to family. Despite this, there is consistency of the around the site harboring the C variant and hopefully
affected family members to share variants at particu- identify the T+ causative variant.
lar positions in the genome within those families. The definition of a family or population and the
Thus, if a researcher genotyped individuals in the requisite strength of the LD providing evidence of a
families at the first position, he or she would see that region harboring a putative disease causing variant in
affected individuals in family E consistently share the context of LD mapping is somewhat arbitrary. For
an A variant, affected family members in family F example, if one confined attention to individuals in
consistently share a G variant, affected family families F and G, then the G, A, and C nucleotides at
members in family G consistently share a G variant, the first, second, and fifth positions would be in LD
and affected family members in family H consistently with the disease causing variant. In addition, if all the
share an A variant. Knowing the position of these affected individuals in families E-I were genotyped at
variants in the genome, the researcher would then the fifth locus and there were only a few affected
examine other variations in the vicinity of the pos- individuals in the sample from the H and I families,
ition harboring the shared A and G variants and then despite the fact the individuals in the H and
hopefully identify the T+ variant that causes the dis- I families do not possess the C variant, there still
eases. This strategy of identifying variants that are might be a very strong association between the
linked to a genomic position harboring a disease C variant and the disease. Suffice it to say that the
causing variant by studying families and then refining word family or pedigree is typically used to refer to
such linkages until a causative mutation is found has individuals related to each other over a small number
often been referred to as “positional cloning” [8]. of generations, whereas a population (or subpopula-
Many clever statistical analysis methodologies have tion) refers to individuals related to each other pos-
been devised to assess evidence for within-family sibly over many generations. In addition, in terms of
consistency of shared variants in the linkage analysis LD strength, it is obvious that the more affected
paradigm that account for, e.g. periodic recombin- individuals possessing a particular variant and the
ation events within a family, incomplete penetrance, fewer unaffected individuals not possessing the vari-
the frequency of variants in the population, and ant, the greater the statistical or probabilistic evidence

17
Chapter 2: Overview of human genetic analysis strategies

that the variant is likely either itself causing the dis- LD with a variant that does (e.g. the C variant in
ease or in LD with something that is. relation to the T+ variant in Figure 2.2). As straight-
In order to facilitate population-based LD map- forward as linkage and LD mapping may seem, how-
ping that does not require families but only a broad ever, there are many complicating factors that are
definition of individuals sampled from a particular discussed in the sections below.
population, the International HapMap Project was
initiated, with the goal of identifying as many genetic
variations as possible that could be interrogated in Special designs
studies seeking to determine if affected individuals There are many extensions and modifications to the
share a particular variant more often than not. In basis linkage and LD mapping approaches described
addition, the International HapMap Project also above. We describe the intuitions and basic strategies
sought to characterize the strength of the LD that behind a few of the more widely used extensions in
neighboring variations exhibit so that researchers the following.
could realistically estimate the likelihood that any Affected sibling pair tests. Instead of collecting
one variant may mark a position in the genome that entire families or large pedigrees with genealogical
is harboring a disease causing variant when geno- information – which can be costly and problematic
typed in a study [11]. To date hundreds of LD map- if many individuals in the earlier generations have
ping studies have been pursued [12]. The results of died – linkage tests that only require affected relative
these studies are listed in many publicly accessible pairs such as siblings have been proposed. The basic
databases (http://www.genome.gov/26525384). intuition behind these tests is to identify genetic
Genome-wide versus candidate gene studies. Many regions where the affected pairs share genetic variants
linkage and LD mapping studies focus on particular likely to have been transmitted from one of their
genomic regions for which some biological evidence parents to a greater degree than expected by chance.
exists that the region may harbor a gene or variants The variants in these regions that are shared by
that influence disease susceptibility. Such “candidate the relative pairs are thus likely to be “identical-by
gene” or “candidate region” studies are really only as descent” from their parents and hence suggest that
reliable as the a priori evidence linking the region to variants in those regions are responsible for the
the disease. Alternatively, researchers can merely test phenotype that all the relative pairs have in common
variants located throughout the genome for linkage or [13, 14]. The affected relative pairs strategy is intui-
LD with a disease and thereby identify regions tive, involves a sampling unit, affected relative pairs,
harboring putative susceptibility variants. For linkage that might be easy to collect and easy to perform
analyses, this would involve looking for patterns of relevant calculations with, and has been applied to a
variant-phenotype “co-segregation” within families number of diseases. However, the power to detect
[9]. A traditional measure, known as the “logarithm sharing of variants within the relative pair units that
of odds” or “LOD” score, is typically used to quantify is consistent with a common disease causing variant
evidence for a co-segregation pattern. For genomic across the pairs is notoriously low, and thus the
regions in which this LOD score has a value with a strategy requires an enormous number of pairs to
very low probability of occurring purely by chance, a identify a true effect in most realistic complex,
researcher can infer the existence of a susceptibility multifactorial disease settings [15, 16].
variant or variants. For LD mapping studies, variants The transmission-disequilibrium test (TDT). A
are simply tested in turn for association with a disease clever test for both linkage and LD involves the
(for example, by contrasting the frequencies of the assessment of the consistent transmission of a par-
variant among individuals with the disease [cases] ticular variant from a parent who is heterozygous
and without the disease [controls] using contingency at the locus harboring that variant to affected off-
table methods). For variants that exhibit association spring [17]. Figure 2.3 provides a graphical depiction
strength (e.g. based on the odds ratio or p-value of the necessary setting: Each of five parent-affected
obtained from the contingency table analysis) that is offspring trios is portrayed. Affection status is
unlikely to occur purely by chance, a researcher can denoted by the “+” sign. For each trio one of the
infer that either that variant in question directly influ- parents is heterozygous for the T variant in the second
ences the disease (e.g. T+ in Figure 2.2) or is simply in position of a haplotype (i.e. they all possess the GT

18
Chapter 2: Overview of human genetic analysis strategies

CA CA CA CA CA CA CA CA CA CA
GG GT GG G T+ GG GT GG G T+ GG G T+
GG CC GG CC GG CC GG CC GG CC

AA CA CA CC AA
G T+ G T+ G T+ G G+ G T+
GC GC GC GC GC

Figure 2.3 Depiction of the transmission of a particular variant, T, and surrounding variants, A and C, to offspring affected (denoted by
the “+” sign) across five different matings involving a parent heterozygous for the T allele. The assessment of the transmission of a particular
allele from heterozygous parents to affected offspring forms the basis for the “transmission disequilibrium test” (TDT; see text for details).

genotype at the second position). The other parent is diabetes has a higher frequency among Native Ameri-
homozygous at this position with the GG genotype). cans, and African-Americans as admixed between
Four of the five affected offspring have been transmit- European and African populations where it is known
ted the T variant. Under Mendel’s laws, a variant from that certain cancers have a higher frequency among
a heterozygous parent should, on average, be trans- African individuals. If particular variants that cause a
mitted to offspring half of the time. The TDT statistic disease are likely to be more frequent in the parental
would thus be consistent with linkage if the frequency populations for which the frequency of that disease
of the transmission of a particular variant from is greater, then one can search for variants possessed
heterozygous parents to affected offspring departed by affected admixed individuals that are known to
significantly from the expected value of half. Note be of higher frequency in that parental population.
that the TDT does not involve controls or individuals Although a gross simplification, admixture mapping
without the disease; rather, the “control” in the TDT along these lines has been used to great advantage for
setting is the variant associated with the heterozygous a few diseases for which an appropriate admixed
genotype that is not transmitted. This is important population has been identified [18].
in diseases for which the definition of a control is
problematic (e.g. neuropsychiatric and behavioral dis-
orders; diseases that only manifest when other condi- Issues in genetic analysis
tions are present such as an environmental factor; There are a number of issues that plague linkage
or age-related diseases). Also, the TDT test does not and LD mapping strategies for identifying genetic
make assumptions about the consistency of the fre- variants that influence disease susceptibility. We
quency of the test variant across different populations briefly describe a few of these issues below and note
and thus avoids issues associated with the problem that all of them are well-recognized by the genetics
of stratification described below [13, 17]. However, community and that concerted efforts to deal with
the TDT strategy does require parental genotype them in various ways have been made.
information and heterozygous parents which may Mutation and nonuniform recombination rates.
limit its applicability. In addition, because parental It is known that mutation and recombination rates
information is exploited in the TDT setting, one can differ across the genome [11]. This creates interpret-
test specific hypotheses about parent-of-origin effects ive issues for linkage and LD mapping studies since it
due to, e.g. imprinting. might be the case that, despite the proximity of a
Admixture mapping. Another clever approach to genetic variant to an actual disease causing variant it
linkage and LD mapping leverages populations may not actually exhibit linkage or LD with a causal
known to have arisen from the admixture of two variant. To overcome this problem, in-depth studies
parental populations with different variant frequency of the mutation and recombination rates in the
profiles and different rates of particular diseases. The genome have been pursued as well as detailed studies
individuals in the admixed population can be con- of the actual LD relationships of genetic variants in
sidered “hybrid” individuals of the two parental different populations [11, 12].
populations since they will possess variants of greater Multiple comparisons. In searching for genetic
frequency in parental populations. Consider, e.g. His- variants throughout the genome, small subsets of
panic populations as admixed between European and which may exhibit linkage or LD with a disease,
Native American populations where it is known that researchers must be sensitive to the potential for false

19
Chapter 2: Overview of human genetic analysis strategies

positive findings due to the massive number of statis- populations rather than European populations. If a
tical tests performed. For example, in a typical researcher sampled individuals with a disease (cases)
genome-wide association study (GWAS) leveraging from one population (e.g. Africa) and individuals
LD mapping, as many as 500 000 to 1 000 000 variants without the disease (controls) from another (e.g.
may be tested for association with a disease. To avoid Europe) then any variant with different frequencies
false positive associations conservative type I error in these two populations could exhibit an association
rates, on the order of 10–7 or 10–8 are required when with the disease in the sample of cases and controls,
declaring significance for a statistical test involving resulting in a noncausal, false positive association
any one variant [16]. [13]. To avoid this genetic background associated
Genetic heterogeneity. As noted previously, it is population “stratification” problem, researchers typ-
likely that most diseases of contemporary public ically either sample cases and controls from a single
health concern have many genetic and nongenetic population, use TDT analysis strategies, or use appro-
determinants. Thus, there may be many different sets priate statistical methods to adjust for population
of factors that can cause a disease. Such heterogeneity genetic background differences between the cases
can reduce the marginal effect of any one factor on and controls.
disease susceptibility and hence create power issues Biological significance and functional assessments.
to detect the effect of that factor in any one study. Identifying variants that are linked or exhibiting LD
To overcome this problem, researchers typically try with a particular disease can create interpretive issues
to identify individuals with more homogenous as to the biological significance of the linkage or LD.
phenotypic profiles or “endophenotypic” profiles on This is especially the case if many variants in a par-
the assumption that these individuals likely have the ticular region exhibit linkage or LD with a disease,
disease due to the same genetic and nongenetic since it will not often be obvious which variant or
factors [19, 20]. subset of variants are causally influencing the disease
Interactions. Genetic factors do not typically work and which are merely in LD with the causal variant(s).
independently of nongenetic factors to mediate dis- To identify the causal variants, sophisticated labora-
ease susceptibility. Rather, genetic factors often tory-based functional assays must be exploited,
“interact” with environmental factors such that the although large-scale functional variant characteriza-
effect of any one genetic factor could be exacerbated tion projects such as the ENCODE (ENCyclopedia
or reduced in the presence of a particular environ- Of DNA Elements) project have been initiated to
mental factor or vice versa. Such interactions have facilitate biological understanding of the role of gen-
been documented in the literature [21] and statistical etic variants in mediating disease susceptibility [23].
methods for identifying and accounting for such
interactions have been devised [22].
Phenotypic heterogeneity. In addition to genetic DNA sequencing and genetic studies
heterogeneity, phenotypic heterogeneity may plague The frequency of genetic variants strongly influences
linkage and LD mapping studies. Phenotypic hetero- the power studies have to detect their effect on a
geneity typically arises when genetic and nongenetic disease in a linkage or LD mapping study. As a result,
factors influence diseases with similar manifestations, most linkage and LD mapping studies performed to
such as different forms of autism spectrum disorder date have focused on the detection of common disease
or Alzheimer’s disease. Better clinical characterization predisposing variants. However, it is now well-recog-
and an understanding of the potential pleiotropic nized that common variants explain only a small to
effects of genes (i.e. an understanding of how perturb- moderate fraction of the heritability of most diseases
ations in a single gene may influence different bio- [1, 24, 25]. Strategies that can identify the effects of
logical and physiological processes) can help mitigate collections of rare variants – any one of which if
the untoward effects of phenotypic heterogeneity. possessed by an individual could cause disease – are
Stratification. One of the most vexing problems currently receiving a great deal of attention [1, 24, 25].
plaguing LD mapping studies is rooted in the fact that Advances in DNA sequencing technologies have
genetic variants typically exhibit different frequencies facilitated studies of rare variants, as these technolo-
in different populations. Thus, for example, a typical gies can be used to exhaustively identify all forms of
variant may be more or less frequent in African variation, common and rare, in a genomic region for

20
Chapter 2: Overview of human genetic analysis strategies

ACGG
T AGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC meaningful groups and test their collective frequency
ACGG
T AGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC differences between individuals with and without
ACGG
T AGTAGAGTAGTGTCCTAGATCGAAATCGATAGCTAGATAGCAC disease [26, 27].

Cases
ACGG
T AGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGTAC
ACCTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC
ACGTGAGTAGAGTAGACTCCTAGATCGAAATCGATAGCTAGATAGCAC
Conclusions
ACGTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC Research investigating the genetic determinants of
ACGTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGTAC disease susceptibility will continue to receive a great
deal of attention, especially since many contemporary
ACCTGAGTAGAGTAAAGTCCTAGATCGAAATCGATAGCTAGATAGCAC studies have been successful in identifying suscepti-
bility variants but have clearly not discovered the vast
ACCTGAGTAGAGTAGAGTCCTAGATCTAAATCGATAGCTAGATAGCAC
majority of such variants. Future studies will clearly
ACCTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC take advantage of more sophisticated sequencing and
ACCTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC genotyping technologies, but will also need to leverage

Controls
ACCTGAGTAGAGTAGAGTCCTAGTTCGAAATCGATAGCTAGATAGCAC better statistical and computational methods as well
ACGTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC as better ways of integrating functional annotations of
ACCTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC the genome in relevant analyses. In addition, although
ACGTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC identifying susceptibility variants is an important first
ACCTGAGTAGAGTAGAGTCCTAGATCGAAATCGATAGCTAGATAGCAC
step in understanding disease pathogenesis, further
ACCTGAGTAGAGTAGAGTCCTAGATCGAACTCGATAGCTAGATAGCAC
“translational” genomic studies that either consider
the identified susceptibility variants and the genomic
Figure 2.4 DNA sequences among a group of cases and controls. regions they reside in as drug targets or as potential
The darkened nucleotides denote variant nucleotides. The dashed prognostic or risk factors are needed to fully capitalize
boxes denote “functional” sites or elements in the sequence. Note on the information that variant discovery studies
that the left-most functional site harbors a variant “T” that is more
frequent in cases than controls and the three right-most functional can provide health researchers, clinicians and public
regions harbor variants that only a few cases and no controls health workers.
possess. Also note that there are few variants that only controls
possess but they are not in functional elements.
Acknowledgements
a group of individuals. Statistical analysis methods for This work was supported in part by the following
such studies are not trivial since the relevant tests research grants: The National Institute on Aging
must focus on the collective effect of the variants, Longevity Consortium (grant number U19
rather than any single variant. Figure 2.4 depicts a AG023122–05); The National Institute of Mental
setting in which sequences for a genomic region have Health (NIMH) -funded Genetic Association Infor-
been obtained on a number of individuals with and mation Network Study of Bipolar Disorder National
without a disease. The darkened nucleotides are vari- (grant number R01 MH078151–03); National Institutes
ant nucleotides. The dashed boxes indicate regions of Health grants: N01 MH22005, U01 DA024417–01,
in the sequence that contain functional elements. P50 MH081755–01, R01 AG030474–02, N01
The first box includes a common variant. The next MH022005, R01 HL089655–02, R01 MH080134–03,
three boxes contain rare variants possessed by only a U54 CA143906–01; Scripps Translational Sciences
few individuals. However, these variants are collect- Institute Clinical Translational Science Award (grant
ively more frequent among the cases rather than the number UL1 RR025774–03), the Price Foundation
controls. By leveraging functional annotations associ- and Scripps Genomic Medicine. Ondrej Libiger is also
ated with regions of the genome, researchers can supported by a grant from Charles University (GAUK
“collapse” rare variants in intuitive and biologically number 134609).

References San Francisco: Benjamin


Cummings; 1996.
4. Martin N, et al. Nat Genet 1997;
17(4):387–392.
1. Frazer KA, et al. Nat Rev Genet
2009;10(4):241–251. 3. Lynch M, et al. Genetics and Analysis 5. Elford J, Ben-Shlomo Y. In
of Quantitative Traits. Boston: Kuh D, et al. (eds.). A Life Course
2. Falconer DS, et al. Introduction
Sinauer and Associates; 1998. Approach to Chronic Disease
to Quantitative Genetics.

21
Chapter 2: Overview of human genetic analysis strategies

Epidemiology. Oxford: Oxford 13. Lander ES, et al. Science 1994;265 20. Morris AP, et al. Genet Epidemiol
University Press; 2004. (5181):2037–2048. 2010;34(2):188–193.
6. Schooling M, et al. Int J Epidemiol 14. Terwilliger JT, et al. Handbook of 21. Caspi A, et al. Science 2003;
2004;33(6):1219–1226. Human Genetic Linkage. Baltimore: 301(5631):386–389.
7. Lewontin RC. Am J Hum Genet Johns Hopkins Press; 1994.
22. Thomas D. Nat Rev Genet 2010;
1974;26(3):400–411. 15. Kruglyak L, et al. Am J Hum Genet 11(4):259–272.
8. Collins FS. Nat Genet 1995; 1996;58(6):1347–1363.
23. Birney E, et al. Nature 2007;
9(4):347–350. 16. Risch N, et al. Science 1996;273 447(7146):799–816.
9. Ott J. Nat Genet 1991;38(8): (5281):1516–1517.
24. Bodmer W, et al. Nat Genet
904–909. 17. Spielman RS, et al. Am J Hum 2008;40(6):695–701.
10. Sham P. Statistics in Human Genet 1994;54(3):559–560; author
Genetics. New York: John Wiley reply 560–553. 25. Schork NJ, et al. Curr Opin Genet
and Sons; 1997. Dev 2009;19(3):212–219.
18. Freedman ML, et al. Proc Natl
11. Frazer KA, et al. Nature 2007; Acad Sci U S A 2006;103 26. Li BS, et al. Am J Human Genetics
449(7164):851–861. (38):14068–14073. 2008;83(3):311–321.
12. Manolio TA, et al. Nature 19. Burmeister M, et al. Nat Rev Genet 27. Madsen BE, et al. PLoS Genet
2009;461(7265):747–753. 2008;9(7):527–540. 2009;5(2):e1000384.

22
DNA methods
Chapter

3 David W. Craig

Introduction Taken as a whole, this chapter is thus focused on tools


and study designs for identifying genetic variants to
In this chapter, we outline methods for studying the explain the heritable component of psychiatric disease.
genetic basis of psychiatric disorders. We focus on The two types of methods we describe, SNP geno-
methods and study designs emerging within the past typing and sequencing, are focused on characterizing
decade, and those likely to be influential in the near- two different types of variants predisposing to disease:
term. We caveat that numerous pivotal findings and common and rare. Common genetic variants predis-
methods emerged throughout the 1990s and in the posing to disease are those frequently found within a
early part of the 2000s that are still relevant today, and population and give a person an increased “relative
excellent review articles are available. risk” for developing a disease. As will be discussed
Before beginning our formal overview of DNA later, SNP genotyping arrays provide a mechanism
methods, it is important to highlight what we will for testing association of common SNPs within a
and won’t be covering in this chapter. This chapter region for association to a heritable disorder. Rare
will specifically be limited to two types of technologies, variants increasing one’s relative risk for developing
high-density single nucleotide polymorphism (SNP) a disorder are primarily accessible through sequencing-
genotyping arrays and so-called “next-generation based studies (noting the exception that SNP arrays
sequencing” (also referred to as second-generation can identify large rare chromosomal abnormalities).
sequencing). We focus on the types of study designs Over the past several years, there have been multiple
these technologies enable and how they have been studies investigating the majority of common variation
used to discover the genetic basis of psychiatric dis- within various psychiatric disorders at a genome-wide
orders. We do not review extensively the large level. These are termed genome-wide association stud-
breadth of genetics findings made within the last ies (GWAS). In the next several years, we can expect
decade. Since we are focusing on DNA in this chapter, that genome-wide sequencing will become realistic,
we limit our discussion to identifying germ-line gen- enabling a genome-wide search for both common
etic variants predisposing to disease. This means that and rare-variants contributing to psychiatric disorders.
we do not give a full discussion of study designs that We now expand on these topics in greater detail.
are particularly effective at identifying somatic vari-
ation. Mainly, in the case of psychiatric disorders the
ability to identify somatic DNA variants involved in
disease pathogenesis is limited by either knowing SNP genotyping
what tissue to investigate or gaining access to that SNPs as a tool for mapping diseases
tissue. Simply, most DNA samples for psychiatric
disorders come from a whole-blood sample – only a across the genome
small number of studies focus on analyzing DNA While 99.9% of the human genome is identical
within brain tissue for example. Related, we do not between individuals, it is estimated that there is on
provide an extensive discussion of epigenetic DNA the order of 10 million SNPs that differentiate indi-
changes and how they may impact disease progression. viduals [1]. SNPs are nucleotide variants at specific

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

23
Chapter 3: DNA methods

positions in the human genome, such as from an within hundreds of base pairs. Unfortunately, multi-
adenosine to thymine (A!T) or a guanine to a cyto- point logarithm of odds (LOD) score calculations
sine (C!T). By definition, they are commonly found assume independence between SNP markers, which
when sampling a population, usually when the minor is not the case for neighboring SNPs in significant
allele is found at a frequency above 1% when LD. In this case, LOD scores will be biased, likely
sampling a population. inflated. This finding is concerning since the biasing
There are three general classes of SNPs when it of a LOD score may be uneven throughout the
comes to function: (1) those with strong functional genome due to variable meiotic recombination rates
significance that dramatically alter a gene’s behavior at different positions within the genome and the
(classic single nucleotide mutations); (2) those with associated uneven physical distribution of the
more subtle functional effects that predispose an indi- markers (on most genotyping panels) throughout
vidual to disease in concert with genetic background the genome. Therefore, without correct attention to
or environment (functional SNPs); (3) those that are LD, one may falsely identify a genetic locus predis-
completely silent with respect to function (nonfunc- posing to a disease. As an alternative, one can con-
tional SNPs). While SNPs can contribute directly to struct virtual multiallelic markers in the form of
disease predisposition by modifying a gene’s function, haplotypes of markers that always travel together
it is really their ability to be used as genetic markers through a pedigree [6]. These virtual markers can
to detect nearby disease-causing mutations through be used for a two-point LOD score calculation with-
association or family-based linkage studies. As out introducing a bias to the LOD score. While
markers, they specifically allow the ability to track analysis of high-density SNP data for linkage studies
how certain regions of the genome co-segregate with is still being optimized, it is encouraging to see the
disease [2]. development of software that can handle the large
The immediate importance of genotyping a high amount of SNP data that is already being utilized to
density of SNPs does not necessarily come from their identify disease-causing genes. In the GWAS section
functional relevance, but rather the proximity of the of this chapter we further describe LD in mathemat-
SNP to the disease-causing genetic mutation or func- ical terms.
tional SNP [3]. By mapping thousands of SNPs across
the genome, it is possible to determine common Technologies for high-density SNP
regions between individuals, either ancestral or famil-
ial. In familial linkage studies, SNPs can be used to genotyping
track sections of chromosomes that co-segregate SNPs have been used since the mid 1990s as markers
through a pedigree with the trait. In case-control for identifying the genetic basis of disease. The use
association studies, SNPs can be identified that statis- was often limited to studying candidate gene studies,
tically associate to complex, multigenic diseases due since sequencing large numbers of SNPs is costly. In
to their close proximity to other, possibly unmapped the early part of the 2000s researchers began working
polymorphisms. In this review, several predominant on the International HapMap Project that had the
technologies that allow for rapid, low-cost genotyping goal of assessing common variation across multiple
of tens to hundreds of thousands of SNPs across the ethnicities by genotyping millions of SNPs. One part
genome will first be described. The basis of how, with of the project was identifying where common SNPs
increasing SNP density, SNP mapping can be used to were in the human population and the other part was
determine the genetic basis of more complex diseases genotyping those SNPs across 240 individuals. In part
will also be addressed. driven by this project, methodologies emerged for
Of particular importance is that many SNPs are inexpensively and rapidly genotyping tens of thou-
in significant linkage disequilibrium (LD) [4, 5]. LD sands of SNPs across multiple individuals. Some
refers to when two polymorphisms are inherited methods were based on sequencing, some based on
together more often than they would be expected to mass spectrometry (Sequenom), whereas others were
by random chance alone due to their physical based on single nucleotide extension of oligionucleo-
proximity to one another on a chromosome. In tide microarrays to allow differentiating alleles. These
high-density SNP data, some SNPs are expected to arrays were already widely used for quantifying RNA
be in LD due to their close proximity, in some cases expression, and a clear path was already in place to

24
Chapter 3: DNA methods

adapt them to classifying whether a person was AA, oligomer probes interrogate each of two possible SNP
AB, or BB (“B” being the reference or ancestral allele alleles. There have been several different designs of
and “A” being the alternative allele). By October 2005, Affymetrix SNP genotyping arrays, and a variety of
the HapMap project’s first phase was published with SNP genotyping analysis strategies. The number
the genotypes of approximately 2.4 million SNPs on 4 of SNPs genotyped in the Affymetrix 6.0 exceeds over
populations and 240 individuals. A planned outcome 1 million SNPs and includes additional markers
of the HapMap project was an assessment of the specific for copy number.
independence between two adjacent SNPs [7]. The The BeadArrayTM platform, manufactured by
fact that SNPs in close proximity are not independent, Illumina, also allows for high-density genotyping.
but rather in LD, meant that only a subset of SNPs In this assay, hybridization is only used for detec-
need to be genotyped in order to assess common tion, and SNP allele discrimination is enzymatic
variation. By 2005, two companies had developed [13–15]. In the enzymatic allele discrimination part
products capable of genotyping a half million SNPs, of the assay, genomic DNA is attached to a solid
and population-based study designs became possible support to better facilitate PCR amplification. Three
and termed GWAS [8]. Using these genome-wide oligomers, specific for each SNP, are used for exten-
high-density SNP genotyping arrays, it was possible sion of the sequence containing the SNP. The first
to complete genome-wide association tests [9]. two oligomers are specific at the 3´ end for the two
Simply, by genotyping hundreds to thousands of different SNP allele types. Only one oligomer will
cases and controls, one tested to see if a particular anneal depending on the particular SNP allele. The
SNP allele was associated with an increased risk of third oligonucleotide anneals in the reverse direction
a disorder. downstream and also contains an addressing
We provide a technical overview of two technolo- sequence. All three oligomers contain a universal
gies that have been used in the majority of linkage primer sequence at the 5´ end for subsequent PCR
studies and population-based GWAS: Illumina amplification. Oligomers are annealed and an allele-
BeadArray Mapping arrays and Affymertrix Gene- specific primer extension is carried out, joining the
Chip Arrays. Of course, other genotyping platforms two annealed oligomers. Depending on the SNP
exist. Sequenom based on mass-spectroscopy differ- allele, one of two possible nucleotides will be incorp-
entiation is a highly flexible platform for custom orated. In a new reaction with the extended product,
genotyping of a few thousand SNPs across a large one of two allele-specific primers containing differ-
number of people. Likewise, Applied Biosystem’s ent fluorescent dyes will anneal. A PCR reaction is
TaqMan and SNPplex assays allow SNP genotyping carried out and amplicons are hybridized to micro-
in hundreds of SNPs across thousands of samples. scopic beads at the end of a fiberoptic bundle.
Services by Perlegen using Affymetrix technology Unique to the BeadArray, the address sequence,
provided millions of genotypes for hundreds of indi- and not the SNP, hybridizes to the array. These
viduals. We focus on the Illumina and Affymetrix address sequences are not public but they do provide
technology because they are typically used within an important advantage, since each bead is not spe-
many laboratories, and are the front-end discovery cific to a particular SNP. Consequentially, probing a
engine. custom set of SNPs does not require a custom array.
®
The Affymetrix GeneChip Mapping Arrays are a
direct extension of the more commonly used expres-
Genotype calls are made from transforming
trimmed mean intensities of Cy3 and Cy5, calculated
sion profiling GeneChip Assays [10, 11]. In the SNP for each bead type, into a polar coordinate system
genotyping assay, fluorescently labeled fragments of where outliers are rejected. For large datasets, quality
DNA containing specific SNPs are genotyped by scores for each SNP can be calculated using a separ-
whether or not they hybridize to a sequence-specific ate software package that compares individual bead
oligomer probe set tiled on a silicon wafer [12]. The intensity distributions with their standard distribu-
Gene-Chip Mapping Array assay uses a whole tions. Like Affymetrix, current arrays by Illumina
genome polymerase chain reaction (PCR) amplifica- can assay over a million SNPs. While SNP genotyp-
tion of digested genomic DNA ligated to universal ing algorithms are largely static with Illumina, the
adapters. DNA is purified, biotin labeled, fragmented, technology has evolved to assay multiple samples on
and hybridized to a microarray. For each SNP, tiled a single slide.

25
Chapter 3: DNA methods

constrained when it comes to their analysis. There-


Copy-number analysis fore, for these rare CNVs it is critical to demonstrate
Recently the impact of human genomic variation in their true apparent frequency in large numbers of
copy number and the link to disease risk has been control individuals. Additionally, the discovery of a
demonstrated in studies of autism, schizophrenia, and rare CNV with a putative association with disease must
amyotrophic lateral sclerosis (ALS) among others. result from an equivalent identification strategy in the
Interestingly, copy number variant (CNV) findings cases and controls – in other words, the rare CNVs will
were often made with high-density SNP genotyping not be identified simply by examining the array data
arrays, which originally were not designed for CNV from the cases (to avoid “ascertainment bias”).
detection. As with SNPs, CNVs can be both patho-
genic or be part of normal genetic variation. Indeed
identifying CNVs in individuals with no overt symp- Linkage or family-based study designs
tomology of disease, laid the groundwork for the In linkage studies, surrogate markers are used to track
study of CNVs in disease that we are seeing emerge genomic regions as they co-segregate with a disease
in the literature today. Importantly these initial through a family structure. While linkage studies have
manuscripts showed that these so-called structural historically relied on multiallelic microsatellite markers,
variations can exist as both common and rare vari- it is now possible to use thousands of SNP markers
ants, and due to this finding it is now commonplace for this purpose. High-density SNP genotyping offers
to perform both SNP and CNV association analysis an alternative to microsatellites in that while they are
in concert for every whole genome association study. biallelic markers, the shear number of markers typed
To facilitate this, both of the major SNP chip manu- provides higher overall information that is more evenly
facturers have more recently built in probes specific- distributed. Indeed, in one of the first high-density
ally designed to facilitate CNV measurement. whole genome studies, Matise and coworkers demon-
Much of the work in the CNV field has primarily strated that a carefully selected set of 2988 SNPs is
revolved around the accurate identification of CNVs more informative than the Marshfield Clinic screening
within each assayed sample. The testing of CNVs for microsatellite version 10 [16]. Perhaps the largest
association with a heritable trait or disease is then family-based linkage study using SNPs was performed
typically performed using the standard statistical tools in autism, with nearly 7500 Affymetrix 10 000 arrays
for SNP association analysis and, as with SNPs, these across 1500 families. New linkage peaks were found
tools may vary based upon the frequency of the with evidence for the neurexin pathway – perhaps most
detected CNV (e.g. common versus rare). interesting were observations of large copy number
A CNV that occurs in greater than 1% of the changes (discussed in the previous section) [17].
general population is termed a CNP – copy number Analysis of linkage studies is well established in
polymorphism. The association analysis of CNPs util- the literature and excellent texts are available on this
izes the typical standard statistical tests (e.g. allele topic [18]. For the purposes of this review, the pri-
frequency comparisons) used for SNP associations. mary statistical output at each marker is a LOD score,
Typically CNPs are di-allelic and can therefore be which can be simplified as the likelihood that the
incorporated as if they were simply SNP genotypes underlying genetic mutation for a disease is associated
(e.g. A/A = no copies, A/B = one copy, and so on). with a particular marker. Generally, a LOD score
These will then be assessed using standard quality above 3 (1024 to 1 odds of a disease gene residing at
control (Mendelian inheritance, duplicate samples, that location in the genome) is considered significant.
and Hardy–Weinberg equilibrium [HWE]) and asso- While analysis of microsatellite genotype data are well
ciation (allelic Chi-square) tools. This is advantageous established, there are a number of practical difficulties
as these methodologies for statistical analysis already in analyzing high-density SNP data in linkage studies.
exist due to work in the SNP genotyping field. In Both Merlin [19] and GeneHunter [20] programs
some cases CNPs exhibit three of more classes. Rare are based on the Lander–Green algorithm [21] and
CNVs, which we may be powered to identify even in have been used in whole genome, high-density link-
the typically “common variant” powered GWA study age studies with SNP data. Programs relying on the
design as we are able to use multiple SNP probe Lander–Green algorithm are generally able to handle
fluorescent values to identify rare CNVs, are less large numbers of SNP markers since computational

26
Chapter 3: DNA methods

time with this algorithm scales linearly with the understand how LD is analytically characterized. LD
number of markers, and exponentially with people. is predominantly measured by D´ and R2. D is the
Conversely, other algorithms (e.g. the Elston–Stewart difference in probability of observing two marker
algorithm [22]) used in some linkage software pack- alleles on the same haplotype versus observing them
ages are not as appropriate for high-density SNP data in an independent population: D = pAB-pA*pB, where
since they scale exponentially with marker number. pAB is the frequency of observing the two calls, A and
B, on the same haplotype versus the expected prob-
ability, pA*pB, if the two calls were completely inde-
GWA studies pendent. A normalized version of D, D´, is commonly
Most diseases fall under the category of complex used so that markers with different allele frequencies
diseases – those that are influenced by multiple envir- can be compared. A measurement of D´ ¼ 0 implies
onmental and genetic factors [23]. As complex dis- complete independence, and a value of one implies
eases are not typically inherited in simple Mendelian that the lower frequency allele is consistently found
fashion, traditional linkage studies are not always with the higher frequency allele. Unfortunately,
appropriate or even possible. In 1996 Kathleen intermediate values of D´ are difficult to interpret.
Merikangas and Neil Risch predicted that if research- The second value, R2, is emerging as a more often
ers could assemble 1000 well-characterized cases and employed metric, since it allows for comparison of R2
1000 properly matched controls, these studies could measurements from two distinct SNP pairs. An R2
find disease variants with moderate risks [3]. Recent of 0 implies independence and a value of one implies
advances in genotyping technology and analytical that the marker alleles at one marker are completely
tools have now made it possible to perform GWAS predictive of what allelic genotype exists at the other
using thousands of samples from well-characterized marker. Conceptually, R2 indicates the amount of
case and control populations. These studies assay information one SNP provides towards the other
hundreds of thousands of markers across the genome SNP marker. The R2 value, including its intermediate
and have elucidated the genetic underpinnings for values, can be used to approximate sample size.
many diseases and disorders, including type 2 dia- The primary measure of effect in case-control stud-
betes, breast cancer [24], prostate cancer, rheumatoid ies is calculation of an odds ratio (OR). An OR is the
arthritis, Crohn’s disease, autism [25], bipolar dis- odds of a person having a disease given that they have a
order [26], and resulted in hundreds of publications particular SNP [27]. An OR for a SNP of greater than
in the last four years. Association studies based on one suggests an increased risk, while an OR of less than
high-density SNP panels are termed indirect associ- one implies a protective effect. An OR of one indicates
ation studies, as the interrogated SNPs are not neces- the SNP has no impact on whether a person will have a
sarily expected to contribute to disease susceptibility. disease. OR is dependent on four parameters: OR of
Rather, genotyped SNPs are expected to be in close the true disease-causing genetic variation, extent of LD
proximity to the true disease-causing polymorphism between the two markers, marker frequency, and the
such that the genotyped SNP statistically associates disease frequency. Statistical significance is typically
with the disease in a case-control manner. determined by an allelic test of association, though
Case-control GWAS are fundamentally based on various disease model tests can be implemented. One
the concept of LD. As previously discussed, LD is of the most common software packages for analyzing
when two polymorphisms are inherited together and quality control of GWA data is PLINK developed
more often than they would be expected to by random by Shawn Purcell and colleagues [28].
chance alone. This will most likely be the case with In combination between linkage and case-control
two SNPs in very close physical proximity. In an association studies are a number of family-based
association study, the SNP being genotyped is approaches largely falling under the category of trans-
expected to be in LD with the disease-causing genetic mission/disequilibrium tests (TDTs) [29]. In-depth
variant. The amount of LD between any two markers reviews of TDT and their variants are available else-
is a function not only of meiotic recombination but where, thus the major points of TDT as they relate
also natural selection, mutation, genetic drift, ances- to high-density SNP data will only be mentioned.
tral population, demographics, and mating patterns The rationale behind TDT is that in the absence of
[1]. From a practical perspective, it is useful to linkage and association between markers, alleles will

27
Chapter 3: DNA methods

be transmitted randomly from parents to children in 1999. Recently developed massively parallel “next
[30]. TDTs have a number of advantages over case- generation” sequencing technologies have begun to
control association studies. First, TDT eliminates replace the previously dominant Sanger sequencing
affects of population stratification. Second, having technology [34] for large-scale sequencing projects
parental genotype information allows for more accur- [35]. Technologies like Illumina’s Genome AnalyzerTM
ate construction of haplotypes. For example, if a child (GA) [36], Roche’s 454TM Sequencing System (454)
has the heterozygous genotype of C/T at a SNP, [37], and Applied Biosystems” SOLiDTM [37] are able
the father is T/T and the mother is C/C, it is possible to generate billions of bases of raw sequence in a
to deduce that the C is from the maternal and the matter of days. These technologies generate relatively
T is from the paternal chromosome. Construction short reads, typically from a few tens to a few hun-
of virtual multiallelic markers crossing several neigh- dreds of bases in length, with a general inverse relation
boring SNPs becomes more reliable with parental between the total number of reads and the read length.
information. Practically, TDTs are often difficult to In the context of whole human genome resequencing,
conduct since obtaining genotype information on on the order of a billion short reads are required to
parents is clinically overwhelming. This problem accurately resequence an individual genome (10–20 
is significant for high-density SNP platforms as it depth coverage) using these technologies. In the con-
has been shown that (compared with multiallelic text of sequencing a specific region across a large
markers) biases can emerge if both parents are not number of individuals, it is possible to sequence
genotyped for biallelic SNPs [31]. multiple individuals within the same sequencing run
When designing an association study, there are using bar coding methods leveraging the pseudo
many possible sources of error. These include obvious single-molecule capabilities of next-generation tech-
technical errors such as an underpowered study or nologies. While there are many potential applications
statistically expected false positives, but also includes of these technologies, we focus on those involving
unseen sampling errors resulting from population sequencing DNA for the purpose of identifying the
stratification and population admixture [32]. The dif- genetic basis of disease, specifically dividing this
ficulty in identifying truly associated SNPs is already section into targeted methods and genome-wide
evident in candidate gene association studies that approaches.
genotype a significantly fewer number of densely A key feature of all these platforms is the concept of
populated SNPs. Indeed, some reviews have found “pseudo single-molecule sequencing”, where each read
that as few as 5–30% of findings from association is derived from one molecule rather than an aggregate
scan are reproduced [33]. False associations from measure of all reads. Consequently, if an individual is
population stratification occur when a regionally or heterozygote at a base having both A and T nucleotides,
ethnically defined group with different risks of disease reads at that base will have discrete values of either A or
is unknowingly sampled. Beyond a high rate of statis- T rather than a mixture of signals from both alleles.
tical false positives and population stratification, The term “pseudo” is used only to indicate that some
population admixture can also lead to false associ- sort of clonal amplification has been used for a single
ations. Lastly, in assessing the significance of a test, molecule. Resequencing a base multiple times using
it is preferable to determine p-values empirically by pseudo single-molecule sequencing opens the window
comparison with large sets of random permutations; for distribution sampling algorithms for polymorphism
that is, sets of randomized data without association. discovery and genotype calling algorithms.
Permutational analysis is more robust to false
assumptions about the distribution, such as from
stratified populations or admixed individuals, than Next-generation sequencing technologies
standard distributions. Until recently, widespread sequencing of DNA has
largely been built around modifications to the Sanger
method developed in the 1970s [34]. Key technological
Sequencing advances permitted scaling of Sanger sequencing
We are at an inflection point in the history of geno- including capillary electrophoresis, automation, robot-
mic sequencing; a moment every bit as disruptive as ics, and fluorescent-labeled nucleotides. Over the last
the introduction of commercial capillary sequencers five years, most sequencing has been completed using

28
Chapter 3: DNA methods

capillary Sanger sequencing through machines such accomplished by bridged PCR-based amplification
as the ABI 3730. More recent efforts to develop to fragmented genomic DNA tethered to a surface
sequencing platforms with fundamentally different by a common oligo sequence ligated during sample
chemistries have matured. The result is a series of preparation. Sequencing consists of single-base
next-generation sequencing technologies with the extension using fluorescently labeled nucleotides that
implied goal of making whole genome resequencing have reversible terminators. Following imaging, the
rapid and inexpensive. Some of the technological terminators are cleaved and the process is repeated.
advances behind the current wave of commercially Important features include phasing due to incomplete
available next-generation sequencers (often referred cleaving of the terminator and pre-phasing due to
to as second generation) are pyrosequencing, revers- incorporation of unterminated nucleotides, both of
ible terminators, and advanced optics. Integrating which result in an error profile that increases at later
these advances with oligo arrays and beads has bases; and use of two lasers with four filters to detect
allowed massively parallel sequencing. Next-gener- four fluorophores leading to a bias towards over-
ation sequencing largely is also about sequencing calling C/T bases and undercalling A/G.
one molecule at a time, whether by direct measure- Roche 454. The Roche GS FLX next-generation
ment or indirectly through amplification. sequencer is based on technology acquired by Roche
Additional features include mate-pair sequencing, as part of their 2007 purchase of 454 Life Sciences
where bases from both ends of a larger DNA frag- and consequently these sequencers are commonly
ment are sequenced such that the two resultant referred to as “454 sequencers”. This 454 sequencing
reads are paired. Since both the SOLiD and GA have is based on massively parallel pyrosequencing of
relatively short reads (< 150 bp), paired data is highly DNA fragments adapter-ligated to small beads which
useful for mapping reads to the genome. We provide are then subjected to PCR and washed into a small
an overview of features for all three commercially well on a PicoTiterPlate where the well contains the
available platforms (SOLiD, GA and 454) below. enzymes required for the sequencing chemistry.
SOLiD. The Applied Biosystems SOLiD platform The 454 platform is somewhat different from the
enables massively parallel sequencing of clonally amp- other 2 platforms in that it is less parallel (approxi-
lified DNA fragments linked to beads. Clonal ampli- mately 1 million reads per run), the reads are longer
fication is accomplished through emulsion-based (400–600 bases), and runs are shorter (hours rather
PCR. The sequencing chemistry is based on sequen- than days). Sequencing is based on quantifying light
tial ligation of dye-labeled oligonucleotide octamer released as bases complementary to the bound tem-
probes that are hybridized, ligated, and imaged in plate strand are ligated. Unterminated bases are
consecutive reactions for generation of paired bases, washed over the slide one at a time so the primary
skipping five bases. Skipped bases are sequenced in failure mode of the 454 is that homopolymer runs
five additional rounds of ligation-based sequencing, result in multiple ligation events that can only be
utilizing universal sequencing primers with different recognized by quantifying the amount of light released.
start sites [37]. Each probe assays two-base positions Emerging platforms. A number of emerging
at a time and each base is assayed in two independent platforms are becoming more widespread in their
extension reactions with overlapping probes. The use. Complete Genomics now offers whole genome
system uses four fluorescent dyes to encode for the sequencing through a service center utilizing unchained
sixteen possible dibase (adjacent pairs of bases) com- reads on self-assembled DNA nanoarrays as a series
binations using a degenerate coding scheme that sat- of short ( 10–20 bp) reads connected by a defined
isfies a number of rules (H. Breu, Applied Biosystems, distance [38]. Ion Torrent (Life Technologies) utilizes
pers. comm., 2008). A single color in the read can a variation of pyrosequencing that differs in their
represent any of four dibase combinations, but the detection method utilizing direct measurement of
overlapping properties of the dibase probes and the hydrogen incorporation on a semiconductor array.
nature of the color code (so-called “color space”) Over the next few years, undoubtedly many new
allow for error-correcting properties. methods will emerge focused on the task of sequen-
Illumina sequencing by synthesis. Illumina’s Genome cing DNA (and RNA) molecules. The major challenge
Analyzer and HiSeq are built around single-base exten- will likely not be generation of data, but rather analy-
sion and reversible terminators. Clonal production is sis of data and that is the focus of the next section.

29
Chapter 3: DNA methods

Data analysis and standards for is correct. Alignments can be further refined through
a micro or local realignment step, where evidence
next-generation sequencing data from multiple reads is used to assess for more com-
Analysis of next-generation sequencing data falls into plex variants. For example, three neighboring SNPs
three categories: (1) alignment/assembly; (2) quality might be better explained using an 1 bp insertion.
control; and (3) variant calling. Finally, a process of marking or removing duplicates
Alignment of short-reads is not surprisingly one is typically completed prior to base-calling. Essen-
of the first areas to see various analysis tools emerge. tially, repeated measures of the same fragment do
Excellent reviews exist and we highlight a few aligners not provide additional information about a variant
[39]. Jeck et al. [40] have described an aligner that and can disrupt some variant calling software.
uses an extension of an aligned shorter read to com- Calling of genetic variants begins first with those
pensate for the fact that accuracy decreases as the read variants smaller than the read, before proceeding to
gets longer. Cokus et al. [41] have presented a prob- structural variants. These variants are typically SNPs
abilistic aligner, along with a pipeline of analysis tools and short insertion/deletions (indels). While almost
for Illumina’s Genome Analyzer and HiSeq. Schatz all variant calling software has some heuristic com-
et al. [42] describe an excellent use of graphic cards to ponent, at some level variant calling relies on: (1) the
accelerate the assembly and alignment of sequence quality of bases predicting a variant; (2) the quality
reads. Warren et al. [43] describe a program, SSAKE, of mapping of the read to the reference genome; and
for stringent assembly of nearly identical sequences (3) the proportion of reads predicting a variant at
that uses a prefix tree under specific search strategies. a given position. Additionally, some variant calling
Addressing data management and analysis, Trombetti approaches utilize additional information such as the
et al. [44] describe a computational pipeline with directionality of read or quality of bases immediately
integrated data storage for analyzing data from neighboring the predicted variant. In practice, most
454 sequencers. Receiving rather quick adoption for variant calling software can yield fewer than 10% false
assembly is Velvet developed by Daniel Zerbino at positives when read depth is greater than 10  cover-
EMBL-EBI [45]. A few mature nonvendor analysis age per ploidy. The most difficult to call genetic
suites to emerge are: BWA developed by Heng Li and variants are generally heterozygotes since they depend
colleagues at the Wellcome Trust Sanger Institute [46] not only on error rate, but also on binomial sampling
(WTSI) and Mosaik/GigaBayes [47]/EagleView [48] by of the variant. In other words, if three of three reads
Gabor Marth and colleagues at Boston College (BC). contain the variant, it is still reasonable that the vari-
Quality control steps are an obvious part of any ant position is an undersampled heterozygote.
analysis pipeline. For next-generation sequencing, Structural variants are typically ascertained by
quality control is less about filtering (as was the case three separate approaches involving analysis of read-
for SNP genotyping) and more about quantifying pairs or fragments: abnormal read pairs (RP) that
quality. The first step this occurs is post-alignment differed in orientation or separation from expected
recalibration of sequence quality data. Sequence data values; discrepant read depth (RD) in comparative or
is typically quantified using a PHRED scale for each absolute terms; and alignment/assembly issues in
base, where a value typically from 0 to 60 is a loga- terms of split reads, directed de novo assembly
rithmic description of the probability of the assigned (microassembly) or global de novo assembly (macro-
base. While the sequencing software typically pro- assembly). Types of structural variants assessable
vides an estimate of this value, this value is often by paired data include:
systematically off. By inspecting alignments at pos-
itions believed to be nonpolymorphic one can assess  Large deletions – Evident by a shorter than
the mismatch rate under various parameters, such as expected distance between read pairs, indicating
base-position. Recalibration, thus, corrects the quality that a deletion has occurred between read pairs.
scores to be more consistent with a logarithmic prob- The region containing the deletion will be missing
abilistic framework that feeds directly into variant alignments.
calling. Additional quality control values can include  Inversions – Evident by inconsistent direction
mapping quality that also provides a logarithmic between read pairs, indicating that one of the read
approximation of the probability that the alignment pairs is a section that has been inverted. The region

30
Chapter 3: DNA methods

at the point of the inversion will also be missing records that can include information about the
alignments reference sequence, the platform, and the run.
 Copy number variants – Evident by a large SAM also has alignment records that include
increase or decrease in alignments across a region the query sequence, quality scores and alignment
of the genome, outside of the expected Poisson information along with the ability to add
distribution and inconsistent with repeats from user-defined attributes to each alignment record.
other regions of the genome.  BAM. Binary equivalent to SAM that can
 Large insertions – Evident by systematic mapping optionally be indexed for fast nonsequential access.
of only one of the two read pairs, along both  Genotype likelihood format (GLF). A format for
read direction. Nonmapping read pairs may also describing evidence for variants at a position
be assembled into a contig. within a reference genome.
 Translocations – mapping of read pairs to  Variant consensus format (VCF). A format for
separate chromosomes in a manner that is describing genetic variants and their respective
inconsistent with mapping to repeat regions. alleles emerging out of the 1000 Genomes Project.

Data standards Sequencing


Next-generation sequencing produces a massive Current next-generation sequencing technologies can
amount of data. Sequencing a genome 30  coverage generally provide gigabases of sequence per day at
requires nearly 100 billion bases, each containing relatively low cost. At this point, it is still quite expen-
value describing quality. Typically, researchers retain sive to sequence the genomes of thousands of individ-
both the raw-sequence data and the aligned data. If uals. Moreover, often evidence from GWAS or other
not careful, it is quite easy to end up with terabytes of family-based linkage studies may point to a specific
data describing the genome of one individual. region of the genome for sequencing.
The complexity and changeability of the data The pseudo-single molecule nature of next-
flows is a powerful argument in favor of a modular generation sequencing allows an elegant solution
design based on specified data input and output for- whereby short 3–10 bp DNA barcodes are ligated to
mats. The data formats will not only ease integration each genomic fragment for multiple individuals and
of modular tools developed, but it will also help pooled prior to sequencing. Essentially, the first
establish the necessary standards for catalyzing devel- several bases or an indexing read allows the researcher
opment of third party software that is interchangeable to determine from which individual each read is
and platform independent. The key data formats derived. Multiplex sequencing by DNA barcoding can
currently in use are short read format (SRF), FASTA significantly reduce the risk of false positives and nega-
with quality (FASTQ), sequence alignment/map (SAM), tives when resequencing large genomic regions because
binary SAM (BAM), and genotype likelihood format coverage can be fine-tuned to ensure that each base
(GLF), described briefly below. is sequenced many times, making next-generation
 Sequence read format (SRF). A platform agnostic sequencing a practical method for mutation/poly-
generic format for DNA sequence data and morphism discovery. Beyond adding the obvious
capable of including platform specific markup advantage of multiplexing large numbers of samples
language. within a run, DNA barcoding offers two additional key
 Fastq. A FASTA-like format that includes two advantages: direct measure of base-by-base error rate
records for each read, one with sequence and one and reduction of day-to-day variability.
with quality scores. Quality scores are typically The next hurdle is to amplify or enrich for only
transformed so each score can be represented the genomic regions of interest. Tiling across a region
as a single printable ASCII character [33]. using PCR and then pooling is one option. If long-
 Sequence alignment/map (SAM). A simple, range PCR is used it can be practical for sequencing
compact, plain-text format for describing the tens of kilobases. More recently, multiple groups have
alignment of a query sequence to a reference demonstrated the use of various targeting strategies
sequence or assembly developed through the using oligo microarrays for enrichment or for multi-
1000 Genomes Project [49]. SAM defines header plexed amplification [50–52]. These efforts build on

31
Chapter 3: DNA methods

existing targeting strategies including molecular of writing whole genome sequencing is somewhat
inversion probes and multiplexed PCR. These recent cost-prohibitive for sequencing hundreds to thousands
papers are exciting although they have also described of individuals ($6000–30 000 for a 30  genome
a number of unresolved problems and challenges with depending on assumptions). One appeal of exon
capturing a specific portion of the genome. sequencing is that it is more readily easy to interpret
Hybridization-based target enrichment. Targets for certain types of changes, such as a one base deletion
resequencing must be isolated or enriched. However, causing a premature truncation or nonsynonymous
long-range PCR is less practical for targeting hun- amino acid change.
dreds of exons disparately distributed throughout To a certain extent, the ability to meaningfully dis-
the genome. Methods have emerged usually utilizing tinguish between benign and functional mutations will
custom oligonucleotide microarrays for hybridizing be a major challenge for making practical discoveries –
and capture fragmented genomic DNA, while nontar- both at the exome level and the genome-wide level.
geted DNA is washed away. Essentially, these methods Data collection efforts such as the 1000 Genomes Pro-
utilize biotinylated RNA “baits” to fish out targeted ject [38], which aims to characterize common human
regions from a “pond” of fragmented DNA. Magnetic variation, or efforts such as the ENCODE (ENCyclo-
streptavidin beads are used to bind and pull down pedia Of DNA Elements) project [55], which aims to
baits, as nontargeted genomic DNA is washed away. guide our understanding of how genetic variation can
Targeted sequencing of coding sequencing or the impact function will be critical, allowing researchers
exome (exome-sequencing) is one of the more obvi- to hopefully move beyond genetic associations to
ous regions of the genome to specifically capture for understanding the functional mechanism of specific
sequencing. Indeed, multiple groups now report variants causing or contributing to psychiatric disease.
exome sequencing using various hybridization-based One of the major challenges for any study is being
capture methods, and one of the first applied able to sift through the large number of germline
examples of identifying a genetic variant of large genetic variations in any one individual. Based on
effect using exome-sequencing has been shown for the 1000 Genomes Project [38], a typical person will
Freeman–Sheldon syndrome [53, 54]. Practically, have 3–4 million genetic variants mostly composed
targeted sequencing on an exome-wide scale is not of SNPs. For SNPs, typically 10 000–12 000 will be
100%, though it remains remarkably efficient. While nonsynonymous. Of these a person will carry 250 to
numbers vary, typically about 90% of the targeted 300 potential loss-of-function genetic variants and
bases are captured at sufficient depth to allow for 50–100 variants linked to disease. Clearly differentiat-
heterozygote calling. About 80–90% of exonic bases ing carrier mutations that don’t lead to disease in a
are targeted. While subjective, it is frequently viewed person with a psychiatric illness from the true disease-
that despite the inefficiencies of exon-sequencing causing mutation(s) in that person will rely heavily
from capture, it is possible to get 80% of the bases on the sample set and the simple observation of a loss-
sequenced at about 1/10th the cost of whole genome of-function does not imply a person’s psychiatric
sequencing. The primary sources of variability in illness is due to that variant. In analysis of families,
these early assays are differences in hybridization co-segregation can be used to filter to limited regions
efficiencies between probes and off-target capture of the genome. Identifying enrichment of one variant
(e.g. immediately neighboring the region or com- class in a group of cases versus controls is another
pletely off-target). Despite these inefficiencies, it is possibility. Likely, use of any of these methods may only
clear at this point that many exon-sequencing studies identify a candidate list of potentially causative variants
will be attempted in the next several years. The alter- and further functional characterization of the mutation
native is whole genome sequencing. At the time by other experimental means will be warranted.

References 3. Risch N, et al. Science


1996;273(5281):1516–1517.
6. Schaid DJ, et al. Am J Hum Genet
2004;75(6):948–965.
1. Carlson CS, et al. Am J Hum
Genet 2004;74(1):106–120. 4. Pritchard JK, et al. Am J Hum 7. Altshuler D, et al. Nature
Genet 2001;69(1):1–14. 2005;437(7063):1299–1320.
2. Kruglyak L, et al. Nat Genet 2001;
5. Reich DE, et al. Nature 2001; 8. Craig DW, et al. Expert Rev Mol
27(3):234–236.
411(6834):199–204. Diagn 2005;5(2):159–170.

32
Chapter 3: DNA methods

9. Hirschhorn JN, et al. Nat Rev 25. Weiss LA, et al. Nature 2009; 41. Cokus SJ, et al. Nature 2008;
Genet 2005;6(2):95–108. 461(7265):802–808. 452(7184):215–219.
10. Chee M, et al. Science 1996; 26. Purcell SM, et al. Nature 2009; 42. Schatz MC, et al. BMC
274(5287):610–614. 460(7256):748–752. Bioinformatics 2007;8(1):474.
11. Fodor SP, et al. Nature 1993; 27. Teng J, et al. Genome Res 1999; 43. Warren RL, et al. Bioinformatics
364(6437):555–556. 9(3):234–241. 2007;23(4):500–501.
12. Kennedy GC, et al. Nat Biotechnol 28. Purcell S, et al. Am J Hum Genet 44. Trombetti GA, et al. BMC
2003;21(10):1233–1237. 2007;81(3):559–575. Bioinformatics 2007;8(Suppl 1):
13. Fan JB, et al. Genome Res 2004; 29. Spielman RS, et al. Am J S22.
14(5):878–885. Hum Genet 1993;52(3): 45. Zerbino DR, et al. Genome Res
14. Gunderson KL, et al. Genome Res 506–516. 2008;18(5):821–829.
2004;14(5):870–877. 30. Schulze TG, et al. Am J Med Genet 46. Bentley DR, et al. Nature 2008;
15. Oliphant A, et al. Biotechniques 2002;114(1):1–11. 456(7218):53–59.
Suppl 2002;56–58:60–51. 31. Sun F, et al. Am J Epidemiol 47. Marth GT, et al. Nat Genet
16. Matise TC, et al. Am J Hum Genet 1999:150(1):97–104. 1999;23(4):452–456.
2003;73(2):271–284. 32. Marchini J, et al. Nat Genet 48. Huang W, et al. Genome Res
17. Hu-Lince D, et al. Am 2004;36(5):512–517. 2008;18(9):1538–1543.
J Pharmacogenomics 2005; 33. Neale BM, et al. Am J Hum Genet 49. Li H, et al. Bioinformatics 2009;25
5(4):233–246. 2004;75(3):353–362. (16):2078–2079.
18. Ott J, et al. Eur Child Adolesc 34. Sanger F, et al. Nature 1977; 50. Albert TJ, et al. Nat Methods
Psychiatry 1999;8 Suppl 3:43–46. 265(5596):687–695. 2007;4(11):903–905.
19. Abecasis GR, et al. Nat Genet 35. Ley TJ, et al. Nature 2008: 51. Okou DT, et al. Nat Methods
2002;30(1):97–101. 456(7218):66–72. 2007;4(11):907–909.
20. Kruglyak L, et al. Am J Hum Genet 36. Bentley DR. Curr Opin Genet Dev 52. Porreca GJ, et al. Nat Methods
1996;58(6):1347–1363. 2006;16(6):545–552. 2007;4(11):931–936.
21. Lander ES, et al. Proc Natl Acad 37. Margulies M, et al. Nature 53. Ng SB, et al. Nat Genet 42(9):
Sci U S A 1987;84(8):2363–2367. 2005;437(7057):376–380. 790–793.
22. Elston RC, et al. Hum Hered 38. Durbin RM, et al. Nature 54. Ng SB, et al. Nature 2009;461
1987;21(6):523–542. 2010;467(7319):1061–1073. (7261):272–276.
23. Carlson CS, et al. Nature 2004; 39. Li H, et al. Brief Bioinform 2010; 55. The ENCODE (ENCyclopedia
429(6990):446–452. 11(5):473–483. Of DNA Elements) Project.
24. Gudmundsson J, et al. Nat Genet 40. Jeck WR, et al. Bioinformatics Science 2004;306(5696):
2007;39(5):631–637. 2007;23(21):2942–2944. 636–640.

33
In silico analysis strategies and resources
Chapter

4 for psychiatric genetics research


Ali Torkamani, Trygve Bakken, and Nicholas J. Schork

Abstract were then genotyped at “marker” loci throughout the


Although a number of strategies for identifying genome and statistical evidence for co-segregation, or
genetic variations that influence common complex “linkage”, between a genetic marker (whose position
neuropsychiatric diseases have been proposed, imple- in the genome was known) and unobserved variants
mented, and pursued, many of these strategies have that were likely responsible for the disease was then
not been able to yield compelling insights into disease assessed. The number of markers that could be inter-
pathogenesis. The reasons for this are themselves com- rogated in early linkage studies was small due to
plex, but it is arguable that extending and integrating technical limitations with available genotyping assays.
available strategies to include detailed biological infor- In addition, the results of linkage studies often sug-
mation can only improve their yield. In this review gested that a large genomic region was likely to
we consider computational methods, databases, and harbor predisposing variations, raising questions not
related resources that can help put into perspective only about how one could identify the causative vari-
the biological and functional significance of genes ations in the region, but also about the biological
and genetic variations interrogated in contemporary meaningfulness of the result. Extensions and adapta-
gene mapping strategies for neuropsychiatric diseases. tions to the basic family-based approaches involving
Computational methods for evaluating the biological analyses limited to sibling pairs and mother–father–
significance of genetic variations, as opposed to labora- affected offspring trios were introduced in order to
tory assays, are quick and, for the most part, easy to avoid the need for large families [2, 4]. Although
use and are growing in sophistication. We provide a linkage analysis strategies did produce notable suc-
discussion of the limitations of available resources, but cesses in the identification of disease-predisposing
ultimately argue that more integrated approaches to DNA sequence variations, most of these successes
the genetic dissection of complex neuropsychiatric were confined to monogenic and overtly Mendelian
conditions are necessary and likely to be the rule rather diseases [2]. When early linkage-based methods were
than the exception in future investigations. applied to major neuropsychiatric conditions, the
results were largely equivocal, raising concerns about
the amenability of such diseases to linkage-based
Introduction strategies for gene discovery (see, e.g. [5]).
Geneticists often consider searching the entire human In the wake of the lack of overt, consistent success
genome for variations that influence susceptibility to of the application of traditional linkage-based methods
disease [1–3]. Strategies for searching the human to common complex diseases, approaches that con-
genome for disease-predisposing variations are sidered the interrogation of a larger set of markers
varied, each having certain advantages and dis- in population-based samples were proposed [6]. The
advantages [2]. The earliest methods focused on the strategies involve testing the association between
analysis of families and pedigrees that included a hundreds-of-thousands to a few million genetic vari-
number of members who had the disease of interest ations characterized on individuals, e.g. with and
and a number of members who did not have the without a specific phenotype. If evidence for associ-
disease of interest. The members of these families ation between the variation and the phenotype was

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

34
Chapter 4: In silico analysis strategies and resources

strong enough to overcome the inevitable multiple potential functional significance of a gene or variation
comparisons that must be made, then one could infer implicated in a linkage or association is now relatively
that either the genetic variation is likely to be causally easy and is becoming more commonplace. In the
related to the phenotype or in linkage disequilibrium following we describe databases and available compu-
with a variation that is causally related to the pheno- tational-based functional analysis tools that can facili-
type. Such genome-wide association studies (GWAS) tate understanding of the biological meaningfulness of
have been pursued for over 300 different disease an association involving a specific genomic region,
phenotypes [3]. Queriable databases cataloging the gene, or specific variation. We note that there are some
results of these studies have been devised to facilitate excellent reviews of available resources that provide an
the dissemination of information to the genetics extensive list of references not discussed here [14–16].
community (https://gwas.lifesciencedb.jp/cgi-bin/gwasdb/ We also note that Tables 4.1, 4.2, and 4.3 list websites
gwas_top.cgi and http://www.genome.gov/gwastudies/). describing the resources we consider.
Although GWAS have produced many unequivocal
findings, the number of variations found to be com-
pellingly associated with disease is small and collect- Sequence analysis
ively the variations do not explain a large fraction of the DNA sequence data must be dealt with at some level
risk of most diseases. In addition, GWAS for neurop- in any genetic investigation. It is therefore important
sychiatric conditions have been largely negative and not to know what tools might be available for manipulat-
replicable (though meta-analyses have been somewhat ing such data. For example, translating DNA or RNA
more positive and consistent). sequences in protein sequences is a basic task in
Despite the unmitigated successes of the GWAS sequence analysis. Depending upon the aims of the
approach to many diseases, there is now general investigator, the resultant protein sequence may be
acceptance that the primary focus of GWAS on used to search for homologous proteins, be aligned
common genetic variations that contribute to disease with other proteins of interest, or the protein sequence
is itself limited and that other approaches are needed may be queried for functional domains to gain insights
[7–9]. In fact, there are many researchers who have into the function of the protein. All these tasks will be
concluded that in order to exhaustively identify gen- described in the following sections. Since nucleotides
etic variations that contribute to disease predispos- code for amino acids in sets of three (codons), there are
ition, one must exhaustively study the genomes of three distinct translations for any nucleotide sequence
individuals, either through focused DNA sequencing depending upon the choice of a starting nucleotide.
strategies, assays that could reveal rare structural vari- Automated tools for translation, for example, the
ations or polygenic factors that contribute to disease, ExPasy Translate tool (http://www.expasy.ch/tools/
or complete genome sequencing [10–13]. dna.html) will automatically generate all three protein
No matter what genetic strategy researchers have sequences in the forward direction, as well as three
ultimately chosen in the identification of genetic vari- protein sequences reading the nucleotides in the
ations that predispose to disease (e.g. family-based reverse direction. However, the correct translation
linkage analysis or association analyses involving can- initiation site must usually be known beforehand.
didate genes, GWA study paradigms, and/or exhaust- In other cases, an investigator may have a DNA
ive DNA sequencing protocols) they have largely been sequence as well as a corresponding amino acid
pursued as a statistical exercise, with evidence impli- sequence, but be interested in determining which
cating a genomic region, gene, or specific genetic vari- specific codons correspond to the amino acids of the
ations being more or less based on the probability protein sequence. For example, an investigator may
that a linkage or association signal could have occurred want to determine where or how a disease-causing
purely by chance. It is now generally accepted that the mutation in the DNA sequence affects the resultant
more biological information one can leverage in deter- protein sequence. Specific web-based tools have been
mining the significance of a linkage or association the developed to align a nucleotide sequence with a pro-
better. However, what biological information should tein sequence. The EBI tool, Genewise (http://www.
be leveraged – or is even feasible both logistically and ebi.ac.uk/Tools/Wise2/index.html) [17], can be used
financially to leverage – is an open question. The use of to automatically generate an alignment for a given
databases and computational methods to sort out the input protein and nucleotide sequence.

35
Chapter 4: In silico analysis strategies and resources

Table 4.1 DNA sequence manipulation tools and resources.

Analysis type Function Website (name of resource)


General Genome browser http://genome.ucsc.edu/ (UCSC Genome Browser)
General Aggregate information http://www.genecards.org/ (GeneCards)
Sequence analysis Translation http://www.expasy.ch/tools/dna.html (Expasy Translate Tool)
Sequence analysis Alignment http://www.ebi.ac.uk/Tools/Wise2/index.html (Genewise)
Sequence motifs Motifs http://meme.sdsc.edu/meme/intro.html (MEME/MAST)
Sequence motifs Motifs http://genie.dartmouth.edu/scope/ (SCOPE)
Sequence motifs Motifs http://zlab.bu.edu/cluster-buster/ (Cluster-Buster)
Sequence motifs Motifs http://159.149.109.9/modtools/ (MoD Tools)
Sequence motifs Logos http://weblogo.berkeley.edu/ (WebLogo)
Alignments General alignment http://jaligner.sourceforge.net/ (JAligner)
Alignments Similarity search http://blast.ncbi.nlm.nih.gov/Blast.cgi (Blast)
Alignments General alignment http://www.ebi.ac.uk/Tools/fasta33/index.html (FASTA)
Alignments General alignment http://www.ebi.ac.uk/Tools/clustalw2/ (ClustalW)
Alignments General alignment http://www.ebi.ac.uk/Tools/t-coffee/ (T-Coffee)
Alignments General alignment http://www.ebi.ac.uk/Tools/muscle/ (MUSCLE)
Alignments General alignment http://www.ebi.ac.uk/Tools/sequence.html (EBI)
Alignments Local alignment http://www.ebi.ac.uk/Tools/emboss/align/index.html (EMBOSS)
Alignments Structural alignment http://www2.ebi.ac.uk/dali/ (Dali)
Alignments Structural alignment http://www.ncbi.nlm.nih.gov/Structure/VAST/vastsearch.html (VAST)

compilation of disease-causing mutations. A major


Sequence variation repository of information is the Online Mendelian
Obviously, the goal of genetic studies is to identify Inheritance in Man (http://www.ncbi.nlm.nih.gov/sites/
DNA sequence variations that contribute to vari- entrez?db=omim) database (OMIM) [18]. The OMIM
ations in phenotypic expression. Therefore, under- database has been maintained electronically by the
standing the biological contexts in which sequence late Victor McKusick and colleagues for over 40 years.
variations might influence a phenotype is crucial to The database has two basic types of entries, gene-
genetic studies. There are a number of databases and centric entries and disease-centric entries. Disease-
tools that can be queried or studied to obtain bio- centric entries generally begin with a basic phenotypic
logical context. Some of these databases and tools description of the disease, describe clinical features,
focus on the general properties of types of genetic inheritance patterns, biochemical features, initial
variations, such as single nucleotide polymorphisms mapping of risk loci, molecular genetics, and animal
(SNPs), such as PupaSuite (http://pupasuite.bioinfo. model information. In some cases, information about
cipf.es/), TAMAL (http://neoref.ils.unc.edu/tamal/), clinical management or other clinical or molecular
and BrainArray (http://brainarray.mbni.med.umich. features is also available. Gene-centric entries also
edu/Brainarray/Database/SearchSNP/snpfunc.aspx), or begin with a general description, describe initial
structural variations, such as DGV (http://projects.tcag. cloning, gene structure, mapping, biological func-
ca/variation/) and dbVAR (http://www.ncbi.nlm.nih. tions, early studies and history, population genetics,
gov/projects/dbvar/). However, there are a large number and eventually end in specific mutational information
of databases dedicated to the mere collection and regarding the gene or disease. Note that each entry is

36
Chapter 4: In silico analysis strategies and resources

Table 4.2 DNA sequence variation, mutation, and polymorphism analysis tools.

Analysis type Function Website (name of resource)


Sequence variation Disease mutations http://www.ncbi.nlm.nih.gov/sites/entrez?db=omim (OMIM)
Sequence variation Disease mutations http://www.hgmd.cf.ac.uk/ac/index.php (HGMD)
Sequence variation Disease mutations http://www.hgvbaseg2p.org/index (HGVBase)
Sequence variation Disease mutations http://www.mutationdiscovery.com/md/MD.com/home_page.jsp
(Mutation)
Sequence variation Alzheimer’s mutations http://www.alzforum.org/res/com/gen/alzgene/default.asp
(AlzGene)
Sequence variation Parkinson’s mutations http://www.pdgene.org/ (PDGene)
Sequence variation SNPs http://www.ncbi.nlm.nih.gov/projects/SNP/ (dbSNP)
Sequence variation SNPs http://www.hapmap.org/ (Hapmap)
Sequence variation Disease associations http://www.ncbi.nlm.nih.gov/sites/entrez?db=gap (dbGAP)
Coding SNPs Prediction http://sift.bii.a-star.edu.sg/ (SIFT)
Coding SNPs Prediction http://www.pantherdb.org/tools/csnpScoreForm.jsp (subPSEC)
Coding SNPs Prediction http://genetics.bwh.harvard.edu/pph/ (Polyphen)
Coding SNPs Prediction http://mmb2.pcb.ub.es:8080/PMut/ (PMut)
Coding SNPs Prediction http://www.snps3d.org/ (SNPs3D)
Coding SNPs Machine learning http://www.cs.waikato.ac.nz/ml/weka/ (Weka)
Noncoding SNPs Regulatory SNPs http://burgundy.cmmt.ubc.ca/cgi-bin/RAVEN/a?rm=home (RAVEN)
Noncoding SNPs Transcription factor http://www.biobase-international.com/gene-regulation/ (BIOBASE)
binding sites
Noncoding SNPs Regulatory regions http://wwwmgs.bionet.nsc.ru/mgs/gnw/trrd/ (TRRD)
Noncoding SNPs Regulatory regions http://www.oreganno.org/oregano/Index.jsp (ORegAnno)
Noncoding SNPs Enhancers http://www.oreganno.org/oregano/Index.jsp (VISTA Enhancer)
Structural variations General information http://projects.tcag.ca/variation/ (DGV)
Structural variations General information http://www.ncbi.nlm.nih.gov/dbvar/ (DBVar)

not limited to the above information, nor is the entir- hgmd.cf.ac.uk/ac/index.php) (HGMD) [19]. This
ety of this information available for all diseases/genes. database is less descriptive but represents a more
The database is easily searchable by keywords and comprehensive catalog of disease-causing mutations.
is thoroughly crosslinked to other related disease or Mutations are broken down into categories, such as
gene entries as well as references to primary literature. missense, splicing, or regulatory variations, and are
OMIM entries are not restricted to genes with vari- searchable by a number of specific identifiers such as
ation information. Suggestive disease links without gene symbol, OMIM number, or disease name. The
mutational evidence are also included. The OMIM database contains both publicly available data and
database is an excellent resource for a disease overview proprietary information, which can be accessed by
and history, however, the mutational information is paying a subscription fee. Note that both HGMD
not comprehensive, and many of the entries contain a and OMIM are hand curated, using a different set
select set of representative or classical variations. of evidence requirements for entry into the database.
Another, more comprehensive, mutational resource Some mutations reported in the literature may not
is the Human Gene Mutation Database (http://www. pass these quality requirements and thus, may not be

37
Chapter 4: In silico analysis strategies and resources

Table 4.3 Functional assessment of genetic variations, genes, and sets of genes.

Analysis type Function Website (name of resource)


SNP function General http://pupasuite.bioinfo.cipf.es/ (PupaSuite)
SNP function General http://neoref.ils.unc.edu/tamal/ (TAMAL)
SNP function General http://brainarray.mbni.med.umich.edu/Brainarray/ (BrainArray)
Protein features Structure http://www.wwpdb.org/ (PDB)
Protein features Structure http://www.rcsb.org/pdb/home/home.do (PDB)
Protein features Structure http://www.ebi.ac.uk/pdbe/ (PDBE)
Protein features Motifs http://prosite.expasy.org/ (PROSITE)
Protein features Domains http://pfam.sanger.ac.uk/ (Pfam)
Protein features Domains http://www.ebi.ac.uk/Tools/pfa/iprscan/ (InterProScan)
Protein features Signal sequences http://www.cbs.dtu.dk/services/ (Various)
Gene expression Expression http://www.ncbi.nlm.nih.gov/geo/ (GEO)
Gene expression Expression http://www.ebi.ac.uk/arrayexpress (ArrayExpress)
Gene expression Expression http://ecoliwiki.net/colipedia/index.php/Center_for_Information_
Biology_Gene_Expression_Database_(CIBEX) (CIBEX)
Gene expression eQTL analysis http://compute1.lsrc.duke.edu/softwares/SNPExpress/ (SNP Express)
Gene expression Expression patterns http://serpanalytics.com/site/tstag.molgen.mpg.de (TStag)
Gene expression Expression patterns http://biogps.gnf.org/#goto=welcome (BioGPS)
Gene expression Brain expression http://www.brain-map.org/ (Allen Brain Atlas)
Pathway analysis Analysis http://www.pathguide.org/ (Pathguide)
Pathway analysis Analysis http://www.geneontology.org/ (Gene Ontology)
Pathway analysis Analysis http://bond.unleashedinformatics.com/ (BOND)
Pathway analysis Analysis http://www.reactome.org/ (Reactome)
Pathway analysis Analysis http://www.kegg.jp/ (KEGG)
Pathway analysis Visualization http://www.cytoscape.org/ (Cytoscape)
Pathway analysis Visualization http://www.biologicalnetworks.net/ (BiologicalNetworks)
Pathway analysis Visualization http://visant.bu.edu/ (VisANT)
Pathway analysis Analysis http://www.genmapp.org/ (GenMAPP)
Pathway analysis Visualization http://www.cs.bilkent.edu.tr/patikaweb/ (PATIKAweb)
Pathway analysis Analysis http://www.ariadnegenomics.com/ (Genomics Pathway Studio)
Pathway analysis Analysis http://www.ingenuity.com/ (Ingenuity)
Pathway analysis Analysis http://www.genego.com/ (GeneGO)

included in one, or both, of the databases. Other as comprehensive as OMIM or HGMD. Disease spe-
mutational databases exist, such as Mutation Discov- cific databases exist as well, for example AlzGene
ery (http://www.mutationdiscovery.com/md/MD.com/ (http://www.alzforum.org/res/com/gen/alzgene/default.
home_page.jsp) and HGVBase (http://www.hgvbaseg2p. asp) and PDGene (http://www.pdgene.org/) dedicated
org/index), however, they do not appear to be nearly to Alzheimer’s and Parkinson’s disease respectively.

38
Chapter 4: In silico analysis strategies and resources

General mutation databases containing commonly aberration which contributes to disease risk. The data
occurring, and not necessarily disease-associated, collected from such large studies is available in the
mutations also exist. The most useful and compre- primary literature, or it may be deposited in the
hensive of these is the Single Nucleotide Polymor- dbGAP database (http://www.ncbi.nlm.nih.gov/sites/
phism database (dbSNP), maintained at the NCBI entrez?db=gap) [22]. Some information within the
website (http://www.ncbi.nlm.nih.gov/projects/SNP/) database is available publicly; however, in general, prior
[20]. Investigators may submit any sequence vari- permission or authorization is required to gain access to
ations they discover to this database, which aims to the bulk of the information contained within the data-
catalogue all known DNA variations, including those base. Access can be requested by applying to the National
falling into regions of the genome not mapping to any Institutes of Health (NIH) Data Access committee.
particular gene. The interface of this database is simi-
lar to that of OMIM, and is searchable by simple
keyword entries. Additionally, each SNP is designated Nonsynonymous polymorphisms
an identifier, which can be used to query a particular When faced with a series of potential nonsynonymous
SNP in other databases or search for phenotypic infor- polymorphisms that may underlie your phenotype of
mation in primary literature. SNPs in dbSNP are also interest, it can pay to prioritize mutations before
accompanied by flanking sequence information, gene testing them further. A number of computational
information, population frequencies if they are known, techniques aimed at differentiating neutral from func-
and validation status, among other pieces of useful tional nonsynonymous polymorphisms have been
information. dbSNP is particularly useful for identifying developed [23]. These methods may be applied to
the known genetic variations in your gene of interest. narrow down a series of common polymorphisms in
Another general SNP database is the HapMap data- candidate genes before an association study in order
base (http://www.hapmap.org/) [21]. All HapMap SNPs to increase the power of the study. Alternatively these
are contained within dbSNP; HapMap contains in- methods may be helpful in narrowing down the
depth information regarding the population distribu- number of mutations to be tested for functional sig-
tion of SNPs, as well as the linkage structure around nificance in in vitro or in vivo settings, sparing the
each SNP. Linkage structure reflects the co-occurrence investigator a lot of time, energy, and money.
of SNPs with one another, that is, how well possession Most computational methods will take advantage
of one SNP predicts the possession of another neigh- of sequence conservation in some form. SIFT, argu-
boring SNP. Many SNPs are tightly linked with one ably the most popular of these methods, takes as its
another because DNA is inherited in large blocks, input your protein of interest, searches for similar
called haplotype blocks, and thus the genetic vari- proteins, generates multiple alignments with these
ations within a block travel together from parent to proteins, and uses these alignments to determine the
child. The HapMap project essentially provides a map degree of conservation at any residue, which in turn
of genetic information so that investigators may is used to output a probability that an amino acid
pursue studies linking particular diseases or other substitution is deleterious [24]. SIFT is available at:
phenotypes to inherited genetic information. http://blocks.fhcrc.org/sift/SIFT.html. Other conser-
This haplotype block information may be lever- vation-based methods, such as subPSEC, are also
aged to design association studies. In an association available online (http://www.pantherdb.org/tools/csnp
study, SNPs which can be accurately used as surro- ScoreForm.jsp) [25]. Other methods, such as Polyphen
gates for the majority of SNPs within a single haplo- (http://genetics.bwh.harvard.edu/pph/) [26] or PMut
type block, are used to represent, or tag, a particular (http://mmb2.pcb.ub.es:8080/PMut/) [27], take advan-
haplotype block, and are thus named tag SNPs. In tage of other sequence-based or physiochemical prop-
carrying out an association study, investigators geno- erties in addition to conservation. Other methods,
type large groups of individuals with or without a such as SNPs3D, (http://www.snps3d.org/) take full
phenotype of interest, and analyze the resultant geno- advantage of available crystallographic data when
type information for frequency differences among the performing predictions [28]. It should be noted that
disease versus control group. Tag SNPs with a statis- PMut and Polyphen also use crystallographic data in
tically elevated frequency within the disease group some instances. In some cases, predictors specific
are likely to be in close proximity to a linked genetic to a protein family, or even a specific protein have

39
Chapter 4: In silico analysis strategies and resources

been generated. For example, predictors have been SNPs. However, some computational strategies
developed for DNA repair genes [29], protein kinases focused on the identification and prediction of the
[30], and G-protein coupled receptors [31]. functional effects of nucleic acid substitutions within
If training data is plentiful for your protein, or transcription factor binding sites or other regulatory
protein family, of interest, the framework for develop- sites have been developed. A recent example is RAVEN
ing custom machine learning prediction methods is (Regulatory Analysis of Variation in Enhancers), which
available as a Java-based platform, Weka (http://www. uses a combination of features, including known tran-
cs.waikato.ac.nz/ml/weka/) [32]. Weka implements a scription factor binding sites, to predict which SNPs
series of popular data-mining methods such as neural may fall into functionally important regulatory regions
networks, support vector machines, random forests, (http://burgundy.cmmt.ubc.ca/cgi-bin/RAVEN/a?rm=
decision tables, and a whole host of other predictive home) [35]. A series of tools are available from BIOBASE
algorithms. Attribute selection algorithms are also (http://www.gene-regulation.com/index.html), such as
implemented to help an investigator determine what MATCH [36], P-MATCH [37], and AliBaba2 [38], all
features are informative when developing a prediction of which identify transcription factor binding sites
model. Custom predictive methods may not be feas- in a variety of ways. These methods are based upon
ible in many cases because of the requirement for a information in the TRANSFAC database [39], a
large training set to form predictions with confidence. proprietary database of regulatory regions. Another
In such cases, a general predictor is sufficient. How- database containing information about transcription
ever, caution should be taken in interpreting the regulatory regions, rather than individual transcrip-
results of any of the above prediction methods. It is tion factor binding sites, is the TRRD database
important to understand how the predictions are (http://wwwmgs.bionet.nsc.ru/mgs/gnw/trrd/) [40].
being formed, and how the predictive attributes may Many of the previous tools rely upon identifying
apply to your gene of interest. For example, mutations known regulatory factor binding sites or regions.
occurring at relatively unconserved amino acids on However, in some cases the investigator may be inter-
the surface of a protein may be considered neutral by ested in identifying novel regulatory motifs which
structural or evolutionary predictors, even though as in may control a set of genes of interest. In this case,
the case of sickle cell anemia, these mutations can result motif search tools have been made available to iden-
in disease [33]. The training data used to train any tify de novo motifs within regions of interest. Two
predictor can greatly influence its predictive accuracy tools, MEME [41] and MAST [42] (http://meme.sdsc.
on a subset of proteins. For example, structure-based edu/meme/intro.html), can either be used to search
predictors trained solely upon DNA repair genes would for a known motif in sequence databases, or discover
not be of great predictive value for other types of pro- motifs within a specific set of sequences. Similar web-
teins. Though this is an extreme example, more subtle based tools are available, such as SCOPE (http://genie.
biases in training data could result in inaccuracies when dartmouth.edu/scope/) [43], Cluster-Buster (http://
the predictive algorithm is applied to an underrepre- zlab.bu.edu/cluster-buster/) [44], and MoD Tools
sented protein family [34]. When used appropriately, (http://159.149.109.9/modtools/) [45]. This list does
these predictive algorithms can be powerful tools in not represent a comprehensive set of the available
the hunt for disease-causing polymorphisms. tools by any means.
Much, if not all, of the publicly available infor-
mation on transcription factor binding sites, and
Regulatory polymorphisms other regulatory regions, is available on the University
Predictive algorithms have been developed less suc- of California, Santa Cruz (UCSC) Genome Browser
cessfully for noncoding polymorphisms. The reasons (http://genome.ucsc.edu/; [46]) (Figure 4.1). A number
for the lack of predictive algorithms for identifying of different tracks relevant to gene regulation are avail-
regulatory noncoding polymorphisms are the relative able. Information from the ENCODE regions, regions
scarcity of training data for disease-associated SNPs of the genome which have been heavily characterized
falling outside protein coding regions, and the relative in terms of transcription factor binding sites, DNA
ease of assigning predictive attributes, such as amino hypersensitivity, and epigenetic modifications, among
acid conservation and structural features of proteins, other characteristics, is available at the UCSC Genome
to protein coding SNPs as compared to noncoding Browser [47]. Some of the elements within the

40
Chapter 4: In silico analysis strategies and resources

Export Navigate the genome


sequences
Switch
genomes

Search
sequences View
genomic
Export features
data

Rearrange
features Select
features to
view

Figure 4.1 University of California, Santa Cruz (UCSC) genome browser layout with key features highlighted.

ENCODE regions have been expanded beyond those contains a number of tracks which describe sequence
regions to the whole genome. For example, transcrip- conservation at a variety of levels. Sequence conservation
tion start sites as determined by luciferase promoting can be queried across 28 vertebrate species or across
activity have been identified for nearly the whole 17 placental mammal species. Regions are denoted by a
genome [48]. Other tracks on the UCSC Genome conservation percentage and ultra-conserved elem-
Browser include but are not limited to the identification ents are highlighted in a separate track.
of CpG islands, conserved transcription factor Motifs, conservation, and regulatory polymorph-
binding sites, microRNA sites, sites with regulatory isms are all interrelated. When a regulatory motif,
potential, and regulatory factor binding sites as deter- or simply a motif of interest, is identified, an investi-
mined by chromatin immunoprecipitation. A track of gator may be interested in determining what nucleo-
note, which combines many of the previous tracks tides of the motif are conserved across all instances
and provides the type of evidence used for each assign- of that motif in a set of sequences of interest.
ment is the ORegAnno track (http://www.oreganno. Nucleotides conserved across motifs are likely to be
org/oregano/Index.jsp) [49]. ORegAnno is an open important in mediating the function of that motif.
access community driven resource for the annotation A useful visualization tool for sequence motifs of
of regulatory elements which is consistently updated any sort, nucleotide or protein, is WebLogo (http://
with new findings by the scientific community. weblogo.berkeley.edu/) [51]). WebLogo takes as its
Another track of note, which is available at the UCSC input a series of alignments and generates a graphic
Genome Browser as well as at a standalone site, is the where the heights of each character at each position
VISTA Enhancer Browser (http://enhancer.lbl.gov/) represent its preponderance at that position. This
[50]. This resource contains experimentally validated provides a very simple and powerful tool for identi-
human noncoding fragments with gene enhancer fying functionally important nucleotides or amino
activity as assessed in transgenic mice. These enhancers acids within a motif.
correspond to regions of extreme conservation across
vertebrates.
Conservation can be a useful tool in identifying Alignments
many different types of functional elements in the Sequence alignments may be generated by many of
human genome. The UCSC Genome Browser the above tools; however they are not generated in all

41
Chapter 4: In silico analysis strategies and resources

cases. Sequence alignments are a robust tool used for Blast (http://blast.ncbi.nlm.nih.gov/Blast.cgi) [52], and
many purposes, the most common of which is the the EBI tool, FASTA (http://www.ebi.ac.uk/Tools/
identification of similar stretches of sequence across fasta33/index.html) [53]. For most purposes, word
DNA or protein regions. Similar stretches of sequence methods are sufficiently accurate. NCBI Blast is also
are identified by arranging sequences in such a way useful for finding similar genes, or stretches of DNA,
that equivalent characters, single letter representa- to a sequence of interest. The investigators’ sequence of
tions of nucleotides or amino acids, are aligned. interest can be used to query databases for similar
Gaps, or missing chunks of sequence, are generally sequences contained within large databases.
represented by a “-” character. Alignment allows the Multiple alignments present a much more chal-
investigator to compare equivalent positions in differ- lenging problem, in which one attempts to find the
ent sequences, to identify patterns occurring across optimal alignment for numerous sequences. Multiple
the sequences. alignments are useful for discovering highly con-
Alignments fall into two general categories: global served elements, for example catalytic site residues
alignments and local alignments. Global alignments or other types of conserved sequence motifs. Dynamic
are used to determine the best overall alignment for programming algorithms can be used to tackle this
the entirety of a set of sequences. This approach is problem; however, because of the exponentially
generally applicable when an investigator knows of an increasing complexity as the number of sequences is
a priori similarity between a set of sequences, and increased, such approaches are generally limited to a
would like to identify corresponding residues across very small number of sequences. NCBI has a multiple
sequences. For example, global alignments would be sequence alignment package available [54]. The major
useful when aligning sequences from different indi- approaches to multiple sequence alignments include
viduals but from the same gene in order to identify the progressive approach or motif finding approach.
sequence variations between the individuals. Local In the progressive approach, the two most related
alignments are used to find regional similarities sequences are aligned, and the resultant alignment is
between two or more sets of sequences. For example, used as a sequence for the next pairwise alignment.
a set of sequences that share a protein domain but This process is iterated until all sequences are
contain different accessory domains would be subject aligned. A number of different implementations
to local alignment. using slightly different approaches exist, including
Pairwise alignments are generally local alignments ClustalW (http://www.ebi.ac.uk/Tools/clustalw2/) [55],
that span only two sequences. These alignments are T-Coffee (http://www.ebi.ac.uk/Tools/t-coffee/) [56],
efficient and quick to calculate, given that the two and MUSCLE (http://www.ebi.ac.uk/Tools/muscle/)
input sequences are of sufficient similarity. They are [57]. Motif-based alignments attempt to discover
calculated in two major ways, dynamic programming short sequences of high similarity that are common
algorithms or word methods. Dynamic programming elements of all the sequences being aligned. Pairwise
algorithms, the Needleman–Wunsch algorithm or alignments guided by these common motifs are exe-
Smith–Waterman algorithm, identify the optimal cuted. Then the pairwise alignments to the common
alignment by calculating scores for all possible align- motifs are combined to provide the resultant overall
ments and retracing a path which selects the highest alignment. One example uses Gibbs sampling to gen-
scoring alignment. Though dynamic programming erate motifs for pairwise alignments [58]. A large
algorithms are more exact, they are also very compu- number of tools exist for generating pairwise or
tationally intensive and excessive for general pur- multiple alignments, some of which are provided in
poses. Implementations of these algorithms are a web interface by EBI (http://www.ebi.ac.uk/Tools/
available as software packages, such as JAligner sequence.html).
(http://jaligner.sourceforge.net/). Word methods, on Local alignments, in contrast, are used to identify
the other hand, are extremely efficient alignment tools, short spans of similarity shared by two or more
but are not guaranteed to find the optimal solution. sequences. This approach is useful when a shared
Word methods search the input sequences for short short element is suspected to be contained within
subsequences of high similarity, and evaluate only a set of sequences. This shorter element may be
matched pairs of sequences with high similarity. Two surrounded by stretches of sequence which differ
major implementations of word methods are NCBI greatly across all the sequences being compared.

42
Chapter 4: In silico analysis strategies and resources

For example, an investigator may use a local align- ebi.ac.uk/pdbe/) [61]. A tutorial for E-MSD is avail-
ment search to attempt to find a short transcription able on the web (http://www.ebi.ac.uk/pdbe/docs/
factor binding site motif within the long upstream roadshow_tutorial/).
promoter sequence of a gene of interest. The
EMBOSS pairwise alignment tool at EBI can be set
to generate local alignments (http://www.ebi.ac.uk/ Protein functional sites and domains
Tools/psa). Protein structures are not necessary for the identifica-
Though sequence similarity is indicative of func- tion of functional sites, such as catalytic sites, or
tional similarity, it is still possible that disparate pro- functional domains. Proteins contain independently
tein sequences may result in similar protein folds and folding functional regions, called functional domains,
similar structure and function. The so-called “twilight which perform specific biochemical functions as inde-
zone” involves comparison of sequences with less pendent subunits of a whole protein. Domains of
than 25% sequence identity. At this point, sequence- importance can be identified by patterns of physio-
based alignments are extremely difficult to generate chemical properties; for example, transmembrane
with high confidence. There are numerous examples domains tend to contain long stretches of hydropho-
of proteins with similar, or the same, function, with bic residues. Important functional sites can also be
low sequence identity. In this case, structural align- identified by particular amino acid patterns; for
ments become useful tools. Though the sequences of example, post-translational modification sites tend
two related proteins may differ to the point where to be defined by specific amino acid motifs. These
their similarities are almost unrecognizable, proteins functional domains or sites can be identified by
of similar function tend to fold into a very similar searching the amino acid sequence for particular pat-
three-dimensional shape. Structural alignment algo- terns indicative of the presence of a functional
rithms take advantage of solved three dimensional domain. These patterns are defined in a number of
structures to find the closest structural superposition different ways including basic amino acid patterns in
of the two proteins. That is, the two structures are PROSITE (http://prosite.expasy.org/) [62], groups of
superimposed so that the distance between “equiva- motifs, or fingerprints, in PRINTS [63], or hidden
lent” amino acids in the two structures is minimized. markov models in Pfam (http://pfam.sanger.ac.uk/) [64],
Corresponding amino acids can be assigned based among others. EBI provides an integrated tool, Inter-
upon the closest amino acid pairs across the two ProScan (http://www.ebi.ac.uk/Tools/pfa/iprscan/)
structures to define a sequence alignment. Numerous [65], which provides the above tools as well as numer-
structural alignment tools are available, including ous other functional annotation tools in a single
EBI Dali (http://www2.ebi.ac.uk/dali/) [59] and NCBI web-based query system.
VAST (http://www.ncbi.nlm.nih.gov/Structure/VAST/ In additional to functional sites, proteins may
vastsearch.html) [60]. contain targeting signals which determine the cellular
distribution of some proteins. These signal peptides
tend to be the most N-terminal amino acids and
Protein structures contain specific amino acid patterns, patterns of
Structural alignments require as their input, protein physiochemical properties, or a combination of the
structure files. The WorldWide Protein Data Bank two, which determine what cellular compartment the
(http://www.wwpdb.org/) is a federation of organ- proteins should be sorted to. The Center for
izations which act as deposition, data processing, Biological Sequence Analysis, at the Technical Uni-
and distribution centers for protein structural infor- versity of Denmark, contains a suite of prediction
mation. The RCSB PDB (http://www.rcsb.org/pdb/ servers for bacterial, plant and animal signal
home/home.do) contains a simple web interface for sequences (http://www.cbs.dtu.dk/services/). In add-
performing searches for an investigator’s protein of ition, the DTU prediction servers contain a series of
interest. More advanced searches by sequence, post-translational modification site prediction
motifs, structural domains, and a whole host of other algorithms that can perform specific post-transla-
filtering parameters are also possible. PDB informa- tional modification predictions (e.g. phosphorylation,
tion can be accessed by numerous web portals. O-glycosylation and N-glycosylation) similar to some
Another example is the PDBe by EBI (http://www. of the predictions provided by the PROSITE tool.

43
Chapter 4: In silico analysis strategies and resources

Expression which the expression assays have been performed


as well as sample size. One of the best, however, is
If a gene or genetic variation within a gene is identified
the SNPExpress database (http://people.genome.duke.
via a linkage or association study, a logical question
edu/dg48/SNPExpress).
might be whether that gene is expressed in a tissue of
relevance to the disease under study (e.g. the brain, in
neuropsychiatric disease). Gene expression datasets are Pathway analysis
scattered widely throughout a large number of gene Genes and proteins do not act in isolation within a
expression data repositories. There are databases and cell, and one can use information about their inter-
resources that provide information on the expression actions to choose candidate SNPs for a study and to
patterns of genes, but these are limited by the tissues help interpret the significance of a study’s results. For
and assays studied. Some of the most relevant data- example, one can apply gene set expression analysis
bases for neuropsychiatric disease include TStag (GSEA) [70, 71] to microarray data to determine the
(http://tstag.molgen.mpg.de/), BioGPS (http://biogps. significance of perturbations to an entire pathway
gnf.org/#goto=welcome), and the evolving Allen Brain of interest, which can be more powerful than com-
Atlas (http://www.brain-map.org/). paring expression levels of individual genes. Pathway
Other gene expression resources harbor actual analysis may also help explain the connections between
gene expression datasets that can be mined and used significant results in GWAS. Torkamani et al. showed
in meta-analyses. The major gene expression reposi- that SNPs in the Wellcome Trust study that did not
tories are the Gene Expression Omnibus (http://www. meet genome-wide significance for disease association
ncbi.nlm.nih.gov/geo/) (GEO) [66], maintained by were disproportionately found within particular path-
NCBI, ArrayExpress (http://www.ebi.ac.uk/arrayex- ways [72]. Some of these results were expected, such as
press) [67], and the Center for Information Biology that SNPs associated with type I diabetes were in
Gene Expression Database (http://ecoliwiki.net/ immune response pathways, but other pathways may
colipedia/index.php/Center_for_Information_Biology_ be novel targets for experimental validation.
Gene_Expression_Database_(CIBEX)) [68]. Data There are three major sources of information
adheres to the “Minimum Information about a Micro- about biological pathways in the cell – single experi-
array Experiment” guidelines, which should guarantee ments, high-throughput assays, and bioinformatics –
access to adequate information to conduct additional and each technique has its trade-offs. First, cell and
analyses using third party generated data. The infor- molecular biological experiments over the past few
mation generally presented is the experimental design, decades have provided detailed data about many meta-
array design, samples, hybridization protocol, measure- bolic pathways. These studies provide a gold standard
ments, and normalization controls [69]. The GEO data- for accuracy, but they are time-consuming and hence
base allows you to either browse through the available are not comprehensive. More recently, high through-
information, or search it using keywords, experiment put experiments have captured a more global picture
identifiers (known as the GEO accession, which is of gene and protein interactions within the cell at the
generally published within the original manuscript cost of some false positives. These data include, for
describing the experiment), or platform identifiers cor- example, binding affinities of thousands of pairs of
responding to the type of microarray used. proteins, enhancer and transcription factor binding
An emerging set of resources and databases sites, and the effects of noncoding RNAs such as
for gene expression analysis of relevance to genetic microRNAs. Bioinformatic tools can identify patterns
studies of neuropsychiatric diseases are “expression in this high-throughput data in order to predict add-
quantitative trait locus” or “eQTL” resources. These itional interactions. For example, a whole genome
databases harbor information on the results of associ- search for microRNA nucleotide binding motifs can
ation studies that identify genetic variations that help predict which gene transcripts will be degraded by
influence the expression levels of particular genes; this these microRNAs. Bioinformatic predictions are the
information may be useful in arguing that a particular least reliable sources of biological interactions, and
variation influences the regulation of a particular they must eventually be verified in living cells, but they
protein rather than the protein’s structure. Many of can help guide experiments to elucidate metabolic
these databases are severely limited by the tissues for pathways about which we know surprisingly little [73].

44
Chapter 4: In silico analysis strategies and resources

Pathway data repositories in their fields, cross-referenced to sequence databases


such as NCBI and Ensembl, and finally checked by
While there are many online repositories of specific
the editorial board. Contributing authors and review-
types of interaction data (e.g. PIPs [74], UniProbe
ers are listed with each pathway, which may increase
[75, 76], VISTA), there are relatively few sites that
the accuracy of the entries. Pathways also include a
attempt to combine this information together into
short description of their biological role as well as the
pathways. This merging has been facilitated by the
interacting components.
adoption of standard data formats. Three of the most
Kyoto Encyclopedia of Genes and Genomes
popular formats are BioPAX, SBML and PSI-MI,
(KEGG) (http://www.kegg.jp/) is perhaps the most
which are appropriate for different types of pathway
comprehensive free resource of pathway information
data [77]. Pathguide (http://www.pathguide.org/)
available [81, 82]. KEGG was started in 1995 and is
lists many pathway data sites and the formats that
now composed of 19 interlinked databases in three
they support, and it is a good place to look for
categories: genomic, chemical, and systems. The
specific datasets, such as kinase pathways or HIV/
GENE database contains genomic information about
host interactions [78]. However, it may make sense to
the building blocks of the cell – genes and proteins.
start with a more comprehensive database of pathway
The COMPOUND and DRUG databases contain
information.
chemical information about endogenous and exogen-
The Gene Ontology (GO) (http://www.geneontology.
ous substances, such as enzymes and pharmaceuti-
org/) database describes the cellular components,
cals approved in the United States and Japan.
molecular functions, and biological processes of gene
PATHWAY has interactions between these genetic
products with a species-independent vocabulary (The
objects that form the molecular wiring diagram
Gene Ontology Project, 2006). For example, a search
of a cell. Finally, BRITE ties together the other data-
for the neuron receptor GluR2 would locate it in
bases and gives higher level information about
the plasma membrane, and associate it with AMPA/
interactions between genes, proteins and the environ-
kainate receptor activity, signal transduction, and syn-
ment. Much of this information is manually curated
aptic transmission. These terms exist in a hierarchical
or annotated data from other public databases such
tree, and so, for example, synaptic transmission is a
as NCBI’s RefSeq.
type of cell-cell signaling, which is a type of cell
communication, which is a cellular process, and so
on. One can easily search for all genes associated with Pathway visualization and analysis
GO terms across species or vice versa, which is a Pathways in a cell are better visualized as a network
powerful way to look for common functional themes rather than a list of parts. While the aforementioned
in a set of SNPs and their associated genes. sites have some visualization features, there are a
The Biomolecular Object Network Databank (BOND) plethora of more powerful tools available. Suderman
(http://bond.unleashedinformatics.com/) includes path- and Hallett recently compared the features of 35 of
way information from the Biomolecular Interaction these tools [83], and we will highlight a couple that
Network Databank (BIND) [79] and a comprehensive, balance power with flexibility and ease of use.
annotated list of patented DNA sequences from Cytoscape (http://www.cytoscape.org/) is a free tool
GENESEQ. BIND has information about more than that can display and analyze large (> 100 000 nodes)
200 000 interactions from over 1500 organisms, many molecular interaction networks derived from many of
of which have been manually curated from the litera- the above resources, including KEGG and Reactome
ture. GENESEQ allows one to screen, for example, drug [84, 85]. Researchers have written almost 100 plugins
targets for potential patent infringements world-wide. for it that extend its functionality, and many of the
Thomson Reuters also offers a pay version called network display options are user customizable. Cytos-
BONDplus that is updated more frequently than BIND cape allows one to connect gene nodes into a network
(http://bond.unleashedinformatics.com/). based on multiple interaction types, including pro-
Reactome (http://www.reactome.org/) makes it tein–protein and protein–DNA. Gene expression data
easy to browse for specific pathways, although it has can be overlaid on this network by coloring genes
a less comprehensive database of interactions than based on their fold change in expression. Visual
BOND [80]. Each pathway is contributed by experts inspection of the network can reveal clusters of

45
Chapter 4: In silico analysis strategies and resources

interconnected genes that show similar expression proprietary pathway databases based on manual
changes, and the jActiveModules plugin automates this curation of full text journal articles. This keeps the
process. A search within these sub-networks may reveal information current and automatically provides ref-
hub genes that regulate the function of neighboring erences for all the interactions. Like PATIKAweb,
nodes. The BiNGO plugin lists biological processes MetaCore is a web application, and so it uses the most
from the Gene Ontology database that are significantly current biological knowledge. MetaCore has a higher
over-represented within the network [86]. Another quality database than free tools such as Cytoscape,
useful plugin, Cerebral, will alter the network layout and it provides many of the same visualization and
so that each gene node is in the appropriate sub-cellular analysis features. In addition, it facilitates import of
location [87]. This makes the co-regulation of sub- many experimental data formats, has flexible options
networks more plausible by showing that the gene for network generation, and allows functional annota-
products could physically interact within the cell. tion based on a customized version of the free GO
Many of these network analyses can be accom- database.
plished in other tools. BiologicalNetworks (http:// Pathway-based analysis is already a useful biological
www.biologicalnetworks.net/) is the visualization inter- tool, and it will become much more powerful as more
face for the PathSys database, which contains much of interaction data is collected between a broad range of
the same information as KEGG. PathSys has a more objects within a cell, including proteins, DNA, RNA,
flexible architecture than Cytoscape and can incorpor- and small molecules. These interactions are dynamic
ate a broader range of molecular data and interactions processes that should be classified by direction, stabil-
[88]. It also has a powerful query language called ity, sub-cellular compartment, developmental time
BioNetSQL that can be used for advanced analysis of point, and many other attributes [95]. Visualization
a network. Another tool worth trying is VisANT tools must be developed that can display this context
(http://visant.bu.edu/), which supports large networks dependence of pathways in a way that highlights the
and plugins like Cytoscape, but it makes it easier biological significance of experimental data.
to integrate data from different sources [89, 90]. Gen-
MAPP (http://www.genmapp.org/) includes thousands
of hand curated pathway MAPPs and will soon use Conclusions
Cytoscape for network visualization [91]. PATIKAweb As the costs and labor associated with genotyping
(http://www.cs.bilkent.edu.tr/patikaweb/) allows one and sequencing are reduced through technological
to quickly visualize a pathway using a pre-built data- developments, more and more research investigating
base of interactions derived from numerous free the role of inherited DNA sequence variations in
sources such as Reactome, NCBI, and GO [92]. disease pathogenesis will be pursued. It is likely that
Commercial pathway analysis tools benefit from these investigations will rely on family-based linkage
access to high-quality databases covering a wide range strategies, genotyping-based GWA study strategies,
of molecular interactions, but their lack of plugin and sequencing-based strategies, but will leverage a
support makes them less flexible. Ariadne Genomics more complete understanding of the full array of
Pathway Studio (http://www.ariadnegenomics.com/ variations that populate the genome. Given that more
products/pathway-studio/) combines data from public variations will be interrogated in future investiga-
resources such as KEGG with data mined from current tions, the need to leverage biological and functional
biomedical literature [93]. Their MedScan technology annotations of genes and genetic variations will
applies natural language processing to the full text of become more pronounced in order to make sense of
journal articles retrieved from PubMed in order to potential linkages and associations. There is growing
extract interactions between biological objects, such precedent with such approaches. For example, The
as genes, proteins, or drugs [94]. MedScan allows International Schizophrenia Consortium recently
researchers to easily build targeted data content for a pursued a study in which they found a greater
specialized field that may not be available in conven- number of deletions in the genomes of individuals
tional pathway databases. with schizophrenia than in the genomes of controls
Ingenuity Pathway Analysis (http://www.ingenuity. [13]. However, since the same deletions were not seen
com/) and GeneGo MetaCore (http://www.genego. across all schizophrenic individuals, the researchers
com/) are two commercial tools that have generated needed to determine whether the various deletions

46
Chapter 4: In silico analysis strategies and resources

possessed by the schizophrenic subjects were likely functional elements in the genome, such as the
to disrupt genes with similar biological effect. With- ENCODE project [96] and the Human Epigenome
out this link between the deletions, the association Project [97], along with attempts to develop uniform
involving a greater number of deletions among standards in referring to and assessing gene function
schizophrenics would not have been compelling from and nomenclature [98], will lead to resources that will
a biological standpoint, and might have been dis- facilitate more integrated approaches to the genetic
missed as a false positive. In addition, Torkamani analysis of human neuropsychiatric diseases.
et al. [72] examined the biological pathways influ-
enced by variations exhibiting weak associations with
a number of traits and showed that the genes and Acknowledgements
pathways harboring weakly associated variations The authors are supported in part by the following
were biologically meaningful and not likely to have research grants: the National Institute on Aging
occurred by chance. Longevity Consortium (grant number U19 AG023122–
Although the pursuit of genetic studies exploiting 01); the National Institute of Mental Health (NIMH)
as much biological information about genes and -funded Genetic Association Information Network
genetic variations as possible are logical, they are Study of Bipolar Disorder National (grant number
obviously limited by the community’s incomplete 1 R01 MH078151–01A1); National Institutes of
knowledge of gene function and genome biology. Health grants (grant numbers N01 MH22005, U01
However, this limitation will be overcome over time DA024417–01, P50 MH081755–01); the Scripps
in much the way inefficiencies in genotyping and Translational Sciences Institute Clinical Translational
sequencing technologies were overcome via techno- Science Award (grant number U54 RR0252204–01),
logical developments. Large-scale efforts to disclose the Price Foundation and Scripps Genomic Medicine.

References 13. International Schizophrenia


Consortium. Nature 2009;
23. Ng PC, et al. Annu Rev Genomics
Hum Genet 2006;7:61–80.
1. Carlson CS, et al. Nature 2004; 460(7256):748–752.
429(6990):446–452. 24. Ng PC, et al. Genome Res 2002;
14. Baxevanis AD. Curr Protoc 12(3):436–446.
2. Lander ES, et al. Science 1994;
Bioinformatics; 2006:Chapter 1:
265(5181):2037–2048. 25. Thomas PD, et al. Proc Natl
Unit 1.1. Acad Sci U S A 2004;101(43):
3. Manolio TA, et al. J Clin Invest
2008;118(5):1590–1605. 15. Plumpton M, et al. In Barnes 15398–15403.
MR (ed.). A Bioinformatics Primer 26. Sunyaev S, et al. Hum Mol Genet
4. Elston RC. Stat Methods Med Res for the Analysis of Genetic Data,
2000;9(6):527–541. 2001;10(6):591–597.
2nd edition. New York: John
5. Risch N, et al. Nat Genet 1996; Wiley and Sons; 2007. 27. Ferrer-Costa C, et al.
12(4):351–353. Bioinformatics 2005;21(14):
16. Zweig AS. Genomics 2008; 3176–3178.
6. Risch N, et al. Science 1996; 92(2):75–84.
273(5281):1516–1517. 28. Yue P, et al. BMC Bioinformatics
17. Birney E, et al. Genome Res 2006;7:166.
7. Frazer KA, et al. Nat Rev Genet 2004;14:988–995.
2009;10(4):241–251. 29. Nakken S, et al. Neuroscience
18. McKusick VA. Am J Hum Genet 2007;145(4):1273–1279.
8. Manolio TA, et al. Nature 2009; 2007;80:588–604.
461(7265):747–753. 30. Torkamani A, et al. Bioinformatics
19. Stenson PD, et al. Hum Mutat 2007;23(21):2918–2925.
9. Schork NJ, et al. Curr Opin Genet
2003;21:577–581.
Dev 2009;19(3):212–219. 31. Xue D, et al. J Hum Genet 2008;
10. Ng SB, et al. Nature 2009; 20. Sherry ST, et al. Nucleic Acids Res 53(5):379–389.
461(7261):272–276. 2001;29:308–311.
32. Witten IH, et al. Data Mining:
11. Ng SB, et al. Nat Genet 2010; 21. The International HapMap Practical Machine Learning Tools
42(1):30–35. Consortium. Nature 2003; and Techniques, 2nd edition. San
426(6968):789–796. Francisco: Morgan Kaufmann; 2005.
12. International Schizophrenia
Consortium. Nature 2008; 22. Mailman MD, et al. Nat Genet 33. Tchernitchko D, et al. Clin Chem
455(7210):237–241. 2007;39:1181–1186. 2004;50(11):1974–1978.

47
Chapter 4: In silico analysis strategies and resources

34. Care MA, et al. Bioinformatics 56. Notredame C, et al. J Mol Biol 79. Bader GD, et al. Nucleic Acids Res
2007;23(6):664–672. 2000;302(1):205–217. 2001;29(1):242–245.
35. Andersen MC, et al. PLoS Comput 57. Higgins D, et al. Nucleic Acids 80. Joshi-Tope G, et al. Nucleic Acids
Biol 2008;4(1):e5. Res 1994;22:4673–4680. Res 2005;33(Database issue):
36. Kel AE, et al. Nucleic Acids Res 58. Neuwald AF, et al. Protein Sci D428–D432.
2003;31(13):3576–3579. 1995;4:1618–1632. 81. Aoki-Kinoshita KF, et al. Methods
59. Holm L, et al. Science Mol Biol 2007;396:71–91.
37. Chekmenev DS, et al. Nucleic
Acids Res 2005;33(Web Server 1996;273:595–603 82. Kanehisa M, et al. Nucl Acids
issue):W432–W437. 60. Madej T, et al. Proteins Res 2008;36(Database issue):
1995;23:356–369. D480–D484.
38. Grabe N. In Silico Biol 2002;
2(1):S1–S15. 61. Golovin A, et al. Nucleic Acids 83. Suderman M, et al. Bioinformatics
Res 2004;32(Database issue): 2007;23(20):2651–2659.
39. Wingender E, et al. Nucleic Acids
Res 1996;24(1):238–241. D211–D216. 84. Cline MS, et al. Nat Protoc 2007;
62. Sigrist CJA, et al. Brief Bioinform 2(10):2366–2382.
40. Wingender E, et al. Nucleic Acids
Res 1997;25(1):265–268. 2002;3:265–274. 85. Shannon P, et al. Genome Res
63. Attwood TK. Brief Bioinform 2003;13(11):2498–2504.
41. Timothy L. Bailey, et al. Nucleic
Acids Res 2006;34:W369–W373. 2002;3:252–263. 86. Maere S, et al. Bioinformatics
64. Finn RD, et al. Nucleic Acids Res 2005;21(16):3448–3449.
42. Timothy L. Bailey, et al.
Bioinformatics 1998;14:48–54. 2006;34:D247–D251. 87. Barsky A, et al. Bioinformatics
65. Zdobnov EM, et al. Bioinformatics 2007;23(8):1040–1042.
43. Carlson JM, et al. Nucleic Acids
Res 2007;35(Web Server issue): 2001;17:847–848. 88. Baitaluk M, et al. BMC Bioinform
W259–W264. 66. Edgar R, et al. Nucleic Acids Res 2006;7:55.
44. Frith MC, et al. Nucleic Acids Res 2002;30:207–210. 89. Hu Z, et al. Brief Bioinform 2008;9
2003;31(13):3666–3668. 67. Parkinson H, et al. Nucleic Acids (4):317–325.

45. Pavesi G, et al. Nucleic Acids Res Res 2007;35:D747–D750. 90. Hu Z, et al. Nucleic Acids Res
2006;34(Web Server issue): 68. Ikeo K, et al. C R Biol 2005;33(Web Server issue):
W566–W570. 2003;326:1079–1082. W352–W357.

46. Kent WJ, et al. Genome Res 69. Brazma A, et al. Nat Genet 2001; 91. Dahlquist KD, et al. Nat Genet
2002;12(6):996–1006. 29:365–371. 2002;31(1):19–20.

47. Thomas DJ, et al. Nucleic Acids 70. Subramanian A, et al. Proc Natl 92. Demir E, et al. Bioinformatics
Res 2007;35(Database issue): Acad Sci U S A 2005;102(43): 2002;18(7):996–1003.
D663–D667. 15545–15550. 93. Nikitin A, et al. Bioinformatics
48. Kim TH, et al. Genome Res 71. Mootha VK, et al. Nat Genet 2003;19(16):2155–2157.
2005;15(6):830–839. 2003;34(3):267–273. 94. Novichkova S, et al.
49. Montgomery SB, et al. 72. Torkamani A, et al. Genomics Bioinformatics 2003;19(13):
Bioinformatics 2006;22(5): 2008;92(5):265–272. 1699–1706.
637–640. 73. Duarte NC, et al. Proc Natl Acad 95. Beyer A, et al. Nat Rev Genet
50. Visel A, et al. Nucleic Acids Res Sci U S A 2007;104(6):1777–1782. 2007;8(9):699–710.
2007;35(Database issue):D88–D92. 74. McDowall MD, et al. Nucleic Acids 96. ENCODE Project Consortium,
51. Crooks GE, et al. Genome Res Res 2008;6:6. et al. Nature 2007;447(7146):
2004;14(6):1188–1190. 799–816.
75. Philippakis AA, et al. J Comput
52. Altschul SF, et al. J Mol Biol Biol 2008;15(7):655–665. 97. American Association for Cancer
1990;215:403–410. Research Human Epigenome
76. Berger MF, et al. Nat Biotechnol Task Force, et al. Nature 2008;
53. Pearson WR. Methods Mol Biol 2006;24(11):1429–1435. 454(7205):711–715.
1994;25:365–389. 77. Stromback L, et al. Bioinformatics 98. Genome Group of the Gene
54. Lipman DJ, et al. Proc Natl Acad 2005;21(24):4401–4407. Ontology Consortium. PLoS
Sci U S A 1989;86:4412–4415. 78. Bader GD, et al. Nucleic Acids Comput Biol 2009;5(7):
55. Edgar RC. BMC Bioinform Res 2006;34(Database issue): e1000431.
2004;5:113. D504–D506.

48
Gene expression studies in psychiatric
Chapter

5 disorders
Alexander B. Niculescu, III

Identifying genes for psychiatric disorders through disorders have produced interesting leads [20, 21].
classic genetic approaches has proven arduous, des- However, this important line of work, if pursued by
pite some recent successes mentioned in this book. itself, suffers from multiple caveats [22] – genetic
This is due to the likely complex, polygenic and het- variability, difficulty of building large enough cohorts,
erogeneous nature of these disorders – multiple genes, uncertainty about exact pre-mortem diagnosis, agonal
multiple polymorphisms in those genes, in different artifacts [23], impact of comorbid medical conditions,
combinations of variable penetrance – involved in and the potential effects of environmental variables
different subtypes of the illnesses. As a consequence, (medications, drugs of abuse, stress, nutrition) on
most genome-wide association studies to date have brain gene expression changes. Animal model gene
been underpowered to detect the full complement of expression studies avoid these caveats, but suffer from
genes and polymorphisms, or even a small portion of the potential limited relevance of the animal model
it [1–3]. The imprecise nature of broad psychiatric used to the human condition [24].
phenotypes has also been a major rate limiting step A combined approach, termed convergent func-
[4–6], and is the subject of study of a new field, tional genomics (CFG) (Figure 5.1), which integrates
psychiatric phenomics [7, 8]. In addition to complex- genetic and gene expression data, in humans and
ity and heterogeneity, there is also a growing appreci- animal models, has been developed as a way of
ation of the genetic, neurobiologic, and phenotypic avoiding the limitations of the individual approaches
overlap and interdependence of various major neuro- mentioned, and reinforcing their strengths in a Baye-
psychiatric disorders [6, 9–12]. The use of case–case sian fashion [25]. This approach has been applied
designs, subphenotypes, and boot-strapping with other with some success to bipolar disorder [16, 17], alco-
lines of work – neurophysiology [13], imaging [14], holism [26], and schizophrenia [11]. Candidate genes
and animal models [15–17], may provide for an accel- identified by such an approach can be pursued in
erated pace of gene identification in the years to come. a prioritized fashion to obtain additional unambigu-
The completion of the sequencing of the human ous evidence for involvement in the illness, through
genome and that of other model organisms, coupled human candidate gene association studies and human
with the advent of microarray technology over the last transgenic mouse studies. The list of prioritized genes
decade, have made large-scale gene expression studies identified by approaches such as CFG also provides
scientifically and economically feasible. After some testable hypotheses for epistatic interactions among
initial debate about different microarray platforms the co-expressed genes [16, 25]. Last, but not least, a
[18], there is an emerging consensus that different combination of CFG with the emerging and growing
platforms perform with similar accuracy and reliabil- large genome-wide association datasets may be
ity if employed well [19]. The main room for particularly powerful in breaking the genetic code of
improvement has been and is at the level of designing psychiatric disorders [27, 28].
appropriate biological experiments, and integrating More recently, there has been renewed interest in
multiple independent lines of evidence in a Bayesian identifying peripheral correlates of neuropsychiatric
fashion. Human postmortem brain gene expression disorders, termed biomarkers. There are to date no
profiling studies from subjects with neuropsychiatric well established, specific clinical laboratory blood tests

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

49
Chapter 5: Gene expression studies

Figure 5.1 Convergent functional


Expanded convergent functional genomics: genomics. QTL, quantitative trait loci.
Multiple independent lines of evidence for Bayesian cross-validation

Animal model studies Human studies


pharmacogenomic, transgenic, inbred strains
Animal model genetic evidence Human
QTL or transgenics genetic
linkage or
association

Animal model brain Candidate Human


evidence gene/ postmortem brain
biomarker evidence

Human blood
Animal model blood evidence
evidence

Sensitivity Specificity

for psychiatric disorders. Given the complex nature of illness state and environmental history, including
psychiatric disorders, the current reliance on patient medications and street drugs, on gene expression, it
self-report of symptoms and the clinician’s impres- is questionable if they have sufficient power to extract
sion on interview of patient is a rate limiting step in bona fide findings about trait and diagnosis, despite
delivering the best possible care with existing treat- the variety of sophisticated statistical methodologies
ment modalities, as well as developing new and used. Combined approaches, such as CFG, may be
improved treatment approaches, including new medi- useful in terms of overcoming current limitations
cations. Identifying molecules in the blood that reflect [34]. (2) Use of lymphoblastoid cell lines. Fresh blood,
illness in the brain would be a major advance. These with phenotypic state information gathered at time
molecules could be used to develop clinical laboratory of harvesting, may be more informative than immor-
tests to aid: (1) diagnosis of illness; (2) early interven- talized lymphocytes, and avoid some of the caveats of
tion and prevention efforts, as well as; (3) prognosis Epstein–Barr virus (EBV) immortalization and cell
of course of illness; and (4) monitoring response to culture passaging [35].
various treatments, including medications. In con- The key litmus test for any biomarker (or genetic)
junction with other clinical information, such tests study should be whether the panels of markers show
will play an important part of personalizing treatment predictive ability in independent cohorts. A series of
to increase effectiveness and avoid adverse reactions – recent studies looking in whole blood at state gene
personalized medicine in psychiatry. Moreover, they expression biomarkers for discrete phenes (mood,
will be of immediate use to pharmaceutical com- hallucinations, delusions) [34, 36], using a CFG meth-
panies engaged in new neuropsychiatric drug devel- odology, have provided reasons for optimism. Panels
opment efforts, at both a pre-clinical and clinical of DNA markers in genes prioritized by CFG have
(Phase I, II, and III) stages of the process. also shown predictive ability in independent cohorts
Lymphocyte protein studies [16] and gene expres- in terms of distinguishing between bipolar disorder
sion profiling [29–33] have emerged as a particularly and controls, and between more severe and less severe
interesting area of research in the search for periph- forms of bipolar disorder [28].
eral biomarkers. Most early studies suffer from one or P-values at best permit the precise ranking of find-
more of the following limitations: (1) The sample size ings within a study. At worst, they are over-interpreted
used in most reports so far is small. Given the genetic and a source of frustration in terms of reproducibility
heterogeneity in human samples and the effects of of findings across studies. This “p-value illusion” in

50
Chapter 5: Gene expression studies

genetics [37], that a significant p-value in one study bringing to bear other large datasets and databases
should necessarily reproduce in another independent relevant to that disease, resulting in essence in a field-
study, is based on an under-appreciation of two wide collaboration. Cross-validating signals from
factors. The first one is the fit-to-cohort effect of other sources, with different noise factors, can
classic statistical analyses of genetic studies, the increase the signal–noise ratio and decrease the
second one is the above discussed complexity and required sample size.
heterogeneity of most disorders. In essence, genes A relatively recent development has been the focus
are identified from the data of a single cohort. This on regulatory RNAs [39], and on epigenetic modifi-
guarantees that best findings will not be at the top of cations [40]. It is clear that gene expression studies
the list in a study conducted in a different cohort, will play a major role in understanding how these
since complexity and heterogeneity will ensure that mechanisms come into play, and what impact the
no two cohorts are alike. This phenomenon is some- environment has. After all, Gene  Environment =
times described as the “winner’s curse”, a strong ini- Gene Expression!
tial finding not being as strong in subsequent In conclusion, gene expression studies have
independent cohorts. proven to be a useful partner to classic genetics
A solution is to use a fit-to-disease approach, like approaches, and combined approaches may provide
CFG [11, 17, 26, 27, 34, 36, 38]. Such an approach, in shortcuts to discovery of genes and overall under-
addition to p-values, uses multiple independent lines standing of the neurobiology involved. For a complete
of evidence related to illness, including gene expres- understanding of the illness, the analyses then need to
sion studies, as a way of prioritizing findings within a be pursued at a biological pathway and mechanistic
cohort, similar to a Google PageRank algorithm. level, integrating environmental effects as key modu-
Genes prioritized in such a way may not have the lators of gene expression and phenotype manifest-
highest p-values, but will generalize and reproduce ation. More progress in quantitative profiling of
well in independent cohorts. Since this approach is psychiatric phenotypes, and borrowing of concepts
based on a gene level integration of data rather than a and paradigms from other medical fields that are
single nucleotide polymorphism (SNP) level analysis, farther along, such as cancer genetics and genomics,
it reduces heterogeneity. It can and has been used are exciting areas of advance for the near future.
profitably to mine genome-wide association studies A (r)evolution in medical nosology in general, and
[27, 28] and biomarker [34, 36] datasets, and to psychiatric nosology in particular, will occur as a
extract panels of top genes or biomarkers that repro- result of such studies. It is hoped that together, all
duce well in independent cohorts. Studies that by these approaches will provide in the long term a
themselves are relatively under-powered can be sound scientific basis for the development of person-
mined and made to yield results using CFG, by alized medicine in psychiatry [41–43].

References 7. Freimer N, et al. Nat Genet


2003;34:15–21.
14. Meyer-Lindenberg A, et al.
Nat Rev Neurosci 2006;7:
1. Baum AE, et al. Mol Psychiatry 818–827.
2008;13:197–207. 8. Kelsoe J, et al. CNS Spectrums
2002;7:215–226. 15. Le-Niculescu H, et al. Am J Med
2. Smith EN, et al. Mol Psychiatry Genet B 2008;147B:134–166.
2009;14:755–763. 9. Berrettini W. Am J Med Genet C
Semin Med Genet 2003;123:59–64. 16. Niculescu AB, III, et al. Physiol
3. Wellcome Trust Case Control
10. Craddock N, et al. Schizophr Bull Genomics 2000;4:83–91.
Consortium. Nature 2007;447:
661–678. 2006;32:9–16. 17. Ogden CA, et al. Mol Psychiatry
11. Le-Niculescu H, et al. Am J Med 2004;9:1007–1029.
4. Schulze TG, et al. Hum Hered
2004;58:131–138. Genet B Neuropsychiatr Genet 18. Tan PK, et al. Nucleic Acids Res
2007;144:129–158. 2003;31:5676–5684.
5. Bearden CE, et al. Trends Genet
2006;22:306–313. 12. Purcell SM, et al. Nature 19. Shi L, et al. Nat Biotechnol
6. Niculescu AB, III, et al. Am J Med 2009;460:748–752. 2006;24:1151–1161.
Genet B Neuropsychiatr Genet 13. Dick DM, et al. Behav Genet 20. Mirnics K, et al. Neuron 2000;
2006;141:653–662. 2006;36:112–126. 28:53–67.

51
Chapter 5: Gene expression studies

21. Choudary PV, et al. Proc Natl 29. Glatt SJ, et al. Proc Natl 36. Kurian SM, et al. Mol Psychiatry
Acad Sci U S A 2005;102: Acad Sci U S A 2005;102: 2011;16(1):37–58.
15653–15658. 15533–15538. 37. Niculescu AB, III, et al. Am J Med
22. Niculescu AB, III. Genome Biol 30. Middleton FA, et al. Am J Med Genet B Neuropsychiatr Genet
2005;6:215. Genet B Neuropsychiatr Genet 2010;153B:847–849.
23. Vawter MP, et al. Mol Psychiatry 2005;136:12–25. 38. Niculescu AB, III, et al. Physiol
2006;11:615. 31. Segman RH, et al. Mol Psychiatry Genomics 2000;4:83–91.
24. Gould TD, et al. Neurosci 2005;10:500–513. 39. St Laurent G, III, et al. Neurosci
Biobehav Rev 2007;31:825–831. 32. Tsuang MT, et al. Am J Med Lett 2009;466:81–88.
25. Bertsch B, et al. Methods 2005; Genet B Neuropsychiatr Genet 40. Murgatroyd C, et al. Nat Neurosci
37:274–279. 2005;133:1–5. 2009;12:1559–1566.
26. Rodd ZA, et al. Pharmacogenomics 33. Vawter MP, et al. Schizophr Res 41. Niculescu AB III. Psychiatr Genet
J 2007;7:222–256. 2004;67:41–52. 2006;16:241–244.
27. Le-Niculescu H, et al. Am J Med 34. Le-Niculescu H, et al. 42. Niculescu AB, III, et al.
Genet B Neuropsychiatr Genet Mol Psychiatry 2009;14: J Affect Disord 2010;123(1–3):
2009;150B:155–181. 156–174. 1–8.
28. Patel SD, et al. Am J Med Genet B 35. Rollins B, et al. Am J Med Genet 43. Niculescu AB, III, et al.
Neuropsychiatr Genet 2010; B Neuropsychiatr Genet Neuropsychopharmacology
153B:850–877. 2010;153B:919–936. 2010;35:355–356.

52
Pharmacogenetics in psychiatry
Chapter

6 Falk W. Lohoff

Introduction in the context of environmental influences comprises


essentially the action of a drug in an individual. By
The introduction of medication for the treatment of identifying genetic components implicated in drug
psychiatric disorders in the 1950s marked the beginning response and side effects, the hope is to precisely
of a new era in psychiatry. For the first time psychiatric match a medication to the patients’ genetic makeup
symptoms could be treated and alleviated with medica- and thus maximize treatment response while minimiz-
tion. This paradigm shift in psychiatry brought with it a ing potential adverse events. The environment of
new focus on research, in particular investigation of a patient will always remain unique and “personal”;
biological and neurochemical aspects in the pathophy- however, advances in genomic medicine promise a
siology of psychiatric disorders. The field of neuropsy- more comprehensive “personalized” pharmacotherapy.
chopharmacology was born and subsequently many
new compounds were developed for the treatment of Pharmacogenetics of antidepressants
psychiatric symptoms. However, despite the continued Major depressive disorder (MDD) is a common psy-
effort to optimize chemical compounds and the effort chiatric illness with high levels of morbidity and mor-
to selectively target specific neurotransmitter systems, tality. It is estimated that 10–15% of the general
response to and tolerability of medication remains population will experience clinical depression during
highly variable with some patients responding to one their lifetime and 5% of men and 9% of women will
treatment but not another. There are several potential experience a depressive disorder in a given year [1].
explanations for the variability in drug response rates, The introduction of tricyclic antidepressants (TCAs) in
including clinical heterogeneity and diagnostic uncer- the 1950s represented a great advance in the treatment
tainty, environmental and social factors, and genetic of depression; however, it soon became clear that ser-
factors. This chapter will describe the recent develop- ious side effects and toxicity limited their use and also
ments in the field of psychiatric pharmacogenetics. varied substantially between patients. Although there
are multiple new pharmacological treatments available
Definition of pharmacogenetics with better side effect profiles, response and tolerability
Pharmacogenetics is the term used to describe the to medication continue to be highly variable and in
phenomenon that genes influence drug response and many cases poor. In general, only one-third of treated
side effects. Drug response and tolerability are influ- MDD patients respond to pharmacological treatment
enced by two main processes referred to as pharma- and achieve remission of symptoms [2, 3]. Based on
cokinetics and pharmacodynamics. Pharmacokinetics these circumstances, the search for biomarkers that
involves the absorption, distribution, metabolism, predict treatment response and efficacy in depression
and excretion of a drug while pharmacodynamics is of high interest to the field.
refers to the effects of the compound on receptors, Genetic influences on antidepressant
transporters, and other downstream targets. Both of
these systems are influenced by heritable variation in drug pharmacokinetics
genes. The complex interaction of multiple genes Pharmacokinetic factors in antidepressant response
involved in pharmacokinetic and dynamic processes have been subject to extensive research over the last

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

53
Chapter 6: Pharmacogenetics in psychiatry

several decades [4, 5]. Most attention has been given still unclear and need to be established. The idealistic
to the cytochrome P450 systems, involved in phase 1 vision of predicting side effects based on genotypes
oxidation. Cyptochrome P450 (CYP450) enzymes are becomes complicated by the reality of environmental
drug-metabolizing hemoproteins present in multiple factors and gene–environment interactions. For
tissues with predominance in liver. Over 50 enzymes example dietary factors (grapefruit juice, caffeine,
have been described which are encoded by over nicotine) have known influences on drug metabolism
63 CYP450 genes. The main P450 enzymes involved [16]; however, perhaps the most common influencing
in antidepressant drug metabolism are CYP2D6, factor is taking another medication. A drug that
CYP2C19, CYP3A4, and CYP1A2 [6]. The CYP2D6 strongly inhibits CYP2D6, for example cimetidine
system has been studied extensively and is so far the or amiodarone, may make an individual who is other-
best characterized [7]. Early experiments with debri- wise a normal metabolizer, appear like a poor meta-
soquin and nortriptyline documented that patients bolizer with increased side effects. Future studies will
fall into different categories, including poor (PM), be necessary to dissect the complicated interaction
intermediate (IM), extensive (EM), and ultra-rapid between clinical presentation, environmental factors,
metabolizers (UM). Cloning and characterization and CYP450 genotype.
of the CYP2D6 gene led to the identification of over
75 CYP2D6 alleles which determine metabolizer Genetic influences on antidepressant
status. For example some individuals carry alleles that
encode for an inactive enzyme or no enzyme at all, drug pharmacodynamics
while others have gene duplications resulting in UM The term pharmacodynamics is used to describe the
of CYP2D6 substrates. Metabolizer status is also effects of a drug on the body. Pharmacodynamic
influenced by ethnicity, with 7% of Caucasians being aspects thus include interactions of a drug with
PM compared to 1% in the Asian population. Certain neurotransmitter receptors, transporters and other
African populations have higher proportions of UM downstream targets. Although the primary mechan-
resulting from multiple copies of the CYP2D6 gene. ism of action is known for most antidepressant drugs,
Such individuals can have an inadequate therapeutic and thought to involve predominantly monoaminer-
response to standard doses of the drugs metabolized gic neurotransmitter systems, the exact mechanisms
by CYP2D6. In fact, plasma levels of the TCA nor- by which antidepressant medications work remain
triptyline and imipramine were closely related to the elusive. Most pharmacogenetic studies in MDD so
number of functional CYP2D6 gene copies [8, 9]. far have focused on candidate genes involved in
Another study showed that patients who lack either monoaminergic neurotransmission. Some of the
one or both functional copies of the gene reach thera- most obvious targets that have been studied include
peutic plasma levels with starting doses of nor- genes encoding the serotonin transporter and recep-
triptyline and reach potentially toxic plasma levels tor, norepinephrine transporter, dopamine receptors,
with high-normal doses. Patients with two to four monoamine oxidase A, tryptophan hydroxylase, and
copies on the other hand require high-normal doses cathechol-O-methyltransferase [17]. It should be noted
just to reach therapeutic plasma levels [10]. Similar that candidate gene association studies in pharmaco-
CYP2D6-allele/plasma level concentration correl- genetics face similar conceptual issues, as do genetic
ations have been reported for selective serotonin association studies in general. Many studies have been
reuptake inhibitors (SSRIs), such as fluoxetine and limited by small sample sizes, clinical and diagnostic
paroxetine [11], and serotonin norepinephrine reup- heterogeneity, incomplete clinical data, differences in
take inhibitors (SNRIs), such as venlafaxine [12, 13]. definition of treatment response, genetic complexity,
While a clear correlation between drug plasma levels unknown functional relevance of single nucleotide
and therapeutic antidepressant efficacy has been polymorphisms (SNPs), and limited biological evi-
described for TCA [14], it is less apparent for SSRIs dence for candidate gene selection. Despite these
and SNRIs [15]. Knowledge of the genetic metabolizer obstacles, several candidate genes deserve mention as
status of a patient is thus helpful to the clinician in they have been suggested repeatedly to be implicated
order to potentially avoid side effects, and might in treatment response to antidepressant drugs.
further help to reach therapeutic levels faster; never- One of the most widely studied genes in pharma-
theless, the overall effects on improved efficacy are cogenetic studies of depression is the serotonin

54
Chapter 6: Pharmacogenetics in psychiatry

transporter gene (SLC6A4). This gene is located on As part of the trial, 1953 patients gave blood for
chromosome 17q and several polymorphisms have pharmacogenetic testing.
been reported in association studies. One polymorph- Several groups have investigated whether genetic
ism in the promoter region of the gene (5HT trans- factors might predict clinical outcome in the STAR*D
porter gene linked polymorphic region: 5HTTLPR) sample. McMahon et al. [27] investigated 68 candi-
consists of an insertion or deletion of a repetitive date genes and genotyped 768 SNPs in this sample.
sequence, producing a short (S) allele or a long (L) They detected an association between the serotonin
allele [18]. Although the 5HTTLPR was originally receptor gene HTR2A and treatment outcome [27].
described as bi-allelic, rare very-long and extra-long Although citalopram does not bind directly to
alleles have been described in Japanese and African- 5HT2A receptors, citalopram downregulates 5HT2A
Americans [19]. The 5HTTLPR L-allele has been receptors in animal models [28, 29], which could
shown to affect transporter function, resulting in play a role in differences in treatment response. Other
higher rates of serotonin reuptake by the transporter. candidate genes implicated in antidepressant drug
Based on this observation, several pharmacogenetic efficacy in the STAR*D trial are FKBP5 and GRIK4
studies have investigated this polymorphism with [30, 31]; however, all of these findings require inde-
regards to antidepressant drug response [20–22]. pendent replication.
Results have been mixed, with some studies showing Given the fact that response and remission are
an effect of the L-version while others failed to dem- complex and influenced by multiple factors, perhaps
onstrate an effect. A recent meta-analysis of 15 pub- focusing on the side effect profile of antidepressants
lished studies however indicated that there was a might reveal a closer link to genetics. For example,
significant association between the L-allele and better while an effect of the 5HTTLPR on primary efficacy
treatment response to SSRIs [23]. Interestingly, the could not be established in the STAR*D sample [32],
association between the L-allele and antidepressant this polymorphism has been reported to be associated
treatment response within four weeks was the most with side effects [33]. In addition, several recent stud-
robust finding in this meta-analysis, suggesting that ies have suggested an association between genetic
the 5HTTLPR might predict not only treatment markers and treatment emergent suicidal ideation
response but also the time-course of response and using the STAR*D sample [25, 34].
remission. Despite these interesting findings, it is still All of the abovementioned studies using the
premature to recommend 5HTTLPR genetic testing STAR*D sample were candidate gene studies. Genes
on a widespread level for patient with MDD. Future and polymorphisms were selected based on the
large-scale prospective clinical studies are needed, to current understanding of the neurobiology of mood
confirm the effect of the 5HTTLPR on antidepressant disorders. With the advance of technology, it is now
treatment response and to identify potential other gen- possible to investigate over a million SNPs in an
etic variants that in combination with the 5HTTLPR individual simultaneously without a priori SNP selec-
will predict treatment outcome. tion based on biological plausibility. Such studies,
While several other genes have been investigated called genome-wide association studies (GWAS),
as candidates for antidepressant drug response, the are currently underway in the STAR*D and other
results are ambiguous and a clear established relation- samples. Recently, results from the first GWAS of
ship between genetic variants and antidepressant drug antidepressant response became available using a
treatment outcome has not yet been demonstrated sample from Germany [35]. Despite the inclusion of
[4, 5, 24, 25]. Recently, results from the Sequenced more than 1500 patients with depression, 700 of them
Treatment Alternatives to Relieve Depression (STAR*D) with genome-wide genotyping, the study failed to
trial became available for review. This multicenter identify single SNP signals that satisfied the criteria
trial included 3671 patients with MDD who were for genome-wide statistical significance, suggesting
treated with the SSRI citalopram as a first-line agent that multiple loci are involved with only modest effect
for 12 weeks [26]. Main outcome measures were cat- sizes. Even though these promising new approaches
egorical response, remission, tolerance, and adverse to investigate genetic variants at a large scale are
effect burden. Analyses of the first phase of the trial intriguing, caution should be used in the interpret-
showed that only 30% of treated patients reached ation of findings. While the STAR*D sample has the
remission after adequate citalopram treatment [3]. advantages of large sample size and a consistent

55
Chapter 6: Pharmacogenetics in psychiatry

design in which all patients received citalopram, it is A recent study investigated the serotonin transporter
important to note that the STAR*D trial was not gene promoter polymorphism in relation to mirtaza-
designed to answer pharmacogenetic questions. There pine and paroxetine efficacy and adverse events in
are several concerns regarding this sample from a geriatric major depression. Results showed that the
pharmacogenetic perspective, including heterogeneity short allele of the 5HTTLPR polymorphism had a
in the severity of depression, multicenter design, significant effect on adverse events among paroxe-
inclusion of different ethnic populations, and lack of tine-treated subjects [40]. Similarly, another study
data on drug levels. In order for pharmacogenetic evaluated the association between adverse events
studies in MDD to succeed, there is a clear need for during SSRI treatment in 214 MDD patients and 2
well-designed, large prospective trials with a focus on polymorphisms in the serotonin transporter gene and
the genetic component [36]. One such attempt is cur- showed that patients with the 5HTTLPR S/S or S/L
rently being undertaken by the European Consortium genotype appeared to have an increased risk of
project, Genome-Based Therapeutic Drugs for Depres- adverse events, especially general adverse events such
sion (GENDEP). This is the first large-scale multicen- as dermatological reactions, weight change, and
ter human pharmacogenomics study focused on the fatigue [41]. Hu et al. [33] investigated the 5HTTLPR
prediction of therapeutic response to antidepressants polymorphism in the STAR*D sample and showed
and adverse effects [37]. This open-label, flexible dose, that the S-allele is associated with citalopram adverse
multicenter trial included 760 patients with MDD that effects, in particular diarrhea. These data suggest that
were treated with either citalopram or nortriptyline for the short allele of the 5HTTLPR polymorphism pre-
12 weeks. Initial analysis of 10 candidate genes dicts adverse effects of SSRI therapy.
involved in serotonin, norepinephrine, neurotrophic, Another serious adverse event of antidepressant
and glucocorticoid signaling revealed an association therapy has recently gained much attention. Treat-
between treatment response to escitalopram and sev- ment emergent suicidal ideations (TESI) are serious
eral variants in the serotonin receptor gene (HTR2A) adverse events in the management of psychiatric
with one marker (rs9316233) explaining 1.1% of the disorders with antidepressants. It is estimated that
response variance. SNPs in the norepinephrine trans- approximately 4% of patients treated with anti-
porter gene (SLC6A2) predicted response to nortripty- depressants develop TESI compared to 2% of
line, and variants in the glucocorticoid receptor gene patients treated with placebo [42–44]. Since the US
(NR3C1) predicted response to both antidepressants Food and Drug Administration (FDA) -mandated
[38]. These data further support a role for the influence black box warning on antidepressant-induced sui-
of genetic variants on treatment response to anti- cidal ideations, the number of antidepressant pre-
depressant drugs. Since single marker analysis only scriptions has significantly decreased. The recent rise
explains a small fraction of the variance, future studies in the number of suicides and drop in the number of
will have to use a multiple variant approach in order to antidepressant prescriptions [45], possibly related to
find clinically meaningful genetic prediction the TESI black box warning, mandates thorough
algorithms. investigation of TESI both clinically and preclinically.
Two genetic association studies have investigated
Pharmacogenetic studies of antidepressant- whether genetic variants contribute to TESI in the
STAR*D sample [34, 46]. Laje et al. [34] investigated
induced adverse events 768 SNPs in 68 candidate genes in the STAR*D sample
Prediction and prevention of adverse events are with respect to TESI. They reported two markers within
important goals of pharmacogenetics and personal- GRIK2 and GRIA3 that were associated with TESI
ized medicine. Most pharmacogenetic studies of anti- during citalopram therapy. Perlis et al. [46] showed that
depressants have focused predominantly on treatment polymorphisms in the CREB1 gene were associated with
response, perhaps also due to the fact that most cur- TESI among men with depression. These studies await
rent antidepressants are well tolerated and fairly safe. replication and in both cases interpretation of results is
Nevertheless, SSRIs and SNRIs can also cause signifi- complicated by several factors including the a priori
cant side effects including sexual dysfunction, gastro- selection of candidate genes (Laje et al. [34] 68
intestinal disturbances, and weight gain, all of which candidate genes; Perlis et al. [46] one candidate gene),
might influence compliance with treatment [39]. the heterogeneous clinical characteristics of the

56
Chapter 6: Pharmacogenetics in psychiatry

STAR*D population (see above) and the lack of bio- Genetic influences on antipsychotic drug
logically functional assessment of genetic variants on
the cellular level. Results of the first GWAS for TESI pharmacokinetics
have been recently reported for the STAR*D sample, As discussed above for antidepressant drugs, recent
suggesting a potential role of variants in the PAPLN pharmacokinetic genetic studies of antipsychotics
and IL28RA genes. PAPLN encodes papilin, a proto- have mostly focused on the cytochrome P450 system;
glycan-like sulfated glycoprotein. IL28RA encodes an however, other areas in which genetic variation may
interleukin receptor [47]. Another study showed that impact pharmacokinetics include blood–brain and
polymorphisms in the BDNF gene were significantly blood–intestine barrier systems [55, 56]. The cyto-
associated with an increase in suicidal ideation in the chrome P450 (CYP450) enzymes mediate phase 1
GENDEP cohort [48]. Interestingly, a significant oxidation of many antipsychotic drugs. Here again,
interaction between variants in BDNF and NTRK2, patients can be grouped into different phenotypic
the gene encoding the BNDF receptor, could be dem- metabolizer categories, including PM, IM, EM, and
onstrated, suggesting the importance of genetic vari- UM. Several genetic polymorphisms in genes encod-
ants in biological pathways and risk to TESI. ing CYP450 enzymes have been associated with these
phenotypes and are predictive of metabolizer status.
Pharmacogenetics of antipsychotic Although antipsychotic drugs are usually metabolized
by a variety of different CYP450 enzymes, major
drugs metabolic pathways that are clinically relevant have
Pharmacotherapy is the treatment of choice for been identified for most commonly prescribed drugs
psychotic symptoms of mental conditions such as [57, 58]. Detailed description of individual antipsy-
schizophrenia, bipolar disorder, and psychotic chotic compounds and their metabolism can be found
depression. Antipsychotic drugs are traditionally elsewhere [59].
divided into two groups: typical (first-generation) CYP2D6 is the major metabolic enzyme for classic
antipsychotics, with strong affinity for the dopamine antipsychotic medications such as chlorpromazine
receptor, and atypical (second-generation) antipsy- and haloperidol but also for the atypical risperidone
chotics, with multiple receptor targets. Antipsycho- [59]. Several polymorphisms in the CYP2D6 gene are
tic drugs have significantly improved the clinical responsible for PM status and gene duplication can
outcome for schizophrenia patients, although several lead to UM status. As mentioned earlier, the frequen-
serious side effects remain important limitations, cies of these phenotypes vary between ethnic groups,
including metabolic abnormalities [49, 50], cardio- with 7–10% of Caucasians being PM in contrast to
vascular events [51], and movement disorders [52]. only 1–2% of Asians. Recent data show that the meta-
Despite their widespread use, about 70% of patients bolic ratio of antipsychotic substrates for CYP2D6
with chronic schizophrenia discontinued their anti- can be affected by genetic variants. For example, the
psychotics drugs in a recent multicenter study due to metabolism of haloperidol is severely reduced in PM
poor effectiveness or tolerability [53]. Antipsychotic and dose adjustment is recommended [60]. On the
pharmacotherapy achieves clinical improvement in other hand, for patients that are UM, higher thera-
the treatment of psychosis in about 50% of patients peutic doses are necessary to compensate for their
[53, 54]. This high degree of response variability and rapid elimination of drugs.
subsequent treatment failure can be attributed to Another important CYP450 enzyme involved in
multiple factors including clinical heterogeneity, antipsychotic pharmacokinetics is CYP1A2, for which
environmental factors, and genetic variations. While clozapine and olanzapine are substrates [61, 62]. Sev-
epidemiological studies of schizophrenia suggest a eral variations in the CYP1A2 gene result in decreased
strong genetic component, no similar studies exist enzyme activity [63, 64] although the clinical rele-
in relation to response to antipsychotics. Neverthe- vance remains to be determined. Some reports indi-
less, the search for genetic variation that predicts cate UM status for CYP1A2 is associated with delayed
treatment response and occurrence of adverse events response to clozapine [65, 66]. While CYP1A2
is the subject of current research and might, in the enzyme activity is in part genetically determined, out-
near future, alter the way antipsychotic medications side factors are also involved. This is an important
are prescribed. consideration in particular for the treatment of

57
Chapter 6: Pharmacogenetics in psychiatry

patients with schizophrenia, since 80–90% are nico- treatment response but no clear association could be
tine dependent and smoking significantly increases documented [57].
the activity of CYP1A2 resulting in higher metabolic The serotonergic system has been also the focus of
ratios, reduced drug levels, and subsequent lack of pharmacogenetic studies of antipsychotics, in part
treatment response [67]. Genotyping and identifica- driven by the high occupancy of 5HT2 receptors by
tion of CYP1A2 UM status, in particular in patients atypical antipsychotics [82]. Several groups have
who smoke, is recommended for accurate dose investigated the relationship between the 102-T/C
adjustment. Numerous polymorphisms have been substitution polymorphism in the 5HT2A gene and
discovered for CYP3A4, CYP2C9, CYP2C19, and response to clozapine and risperidone with both posi-
CYP2A5; however; no clear connection between vari- tive and negative results reported [83–88]. Another
ants and level of response to antipsychotic medication variant in this gene, the amino-acid substitution poly-
has yet been established. morphism His452Tyr, was also found to be associated
with clozapine treatment response in several datasets
[89, 90]. A recent meta-analysis further supports an
Genetic influences on antipsychotic drug association of the 102-T/C and His452Tyr variants with
clozapine response, reinforcing the role of the 5HT2A
pharmacodynamics receptor as an important therapeutic target [84].
Antipsychotic drugs display a wide range of affinities A variety of other serotonin receptor subunit genes
for different neurotransmitter receptors, including have been investigated as well with respect to antipsy-
dopaminergic, serotoninergic, histaminergic, muscar- chotic treatment response, including 5HT1A, 5HT2C,
inergic, glutaminergic, and adrenergic receptors [68, 5HT3A, 5HT3B, 5HT5A, and 5HT6 [57]. Results of
69]. The dopaminergic system is thought to play a these studies are mixed and follow-up studies will have
major role in the pathophysiology of psychosis and to use larger samples in randomized prospective treat-
dopamine receptor blockade has been associated with ment trials. Future studies using GWAS designs are
amelioration of psychotic symptoms [70]. Positron also likely to be used for investigations of pharmaco-
emission tomography (PET) studies show that 60% genetic aspects of antipsychotic treatment response;
minimum occupancy of D2-like receptors is necessary however, methodological issues regarding sample size
to obtain therapeutic response and occupancy above and multiple testing are significant obstacles that
80% may lead to extrapyramidal side effects [71, 72]. impact the feasibility of such endeavors.
Variants of the genes encoding receptors targeted by
antipsychotic medications are logical candidates for Pharmacogenetic studies of antipsychotic
pharmacogenetic investigations. Several studies have
investigated polymorphisms in the DRD2 gene and drug-induced adverse events
response to antipsychotic treatment. The promoter The side effects associated with antipsychotic treatment
insertion/deletion polymorphism -141-C Ins/Del was often contribute to poor compliance and treatment
associated with lower treatment response [73–75]. failure. Drug-induced weight gain and tardive dis-
Although the functional relevance of this polymorph- kinesia (TD) have been the primary focus of pharma-
ism remains unclear, some evidence suggests an asso- cogenetic studies in the recent past, although
ciation of the Del allele with higher striatal D2 receptor accumulating data are also available for less common
density [76]. side effects like drug-induced agranulocytosis and
The DRD3 gene has also been investigated with neuroleptic malignant syndrome.
respect to antipsychotic treatment response. The Antipsychotic-induced weight gain is a serious
Ser9Gly polymorphism in the DRD3 gene has been side effect that is particularly high among patients
associated with effects on dopamine binding affinity receiving multiple antipsychotic drugs simultaneously
[77]. Several studies suggest that the Gly allele is associ- [91]. Evidence from twin studies suggests a genetic
ated with good treatment response [78–80]; however, component to this phenomenon [92, 93]. Since sero-
one study in patients of Chinese descent indicated pre- tonin and histamine receptors play important roles
diction of poor treatment response [81]. Several other in eating behavior, genes encoding these receptors
genes involved in dopaminergic neurotransmission are logical candidates for pharmacogenetic inquiry.
have been investigated with regards to antipsychotic Numerous studies have investigated the 5HT2C

58
Chapter 6: Pharmacogenetics in psychiatry

receptor gene, and although some studies did not find


an association, accumulating evidence suggests a role
Pharmacogenetics of anticonvulsants
for the functional -759-T/C polymorphism in medi- and mood stabilizers
ating weight gain [94–98]. Interestingly, the T-allele Anticonvulsant drugs are widely used in the manage-
has strong protective properties against antipsychotic- ment of behavioral disorders, including bipolar dis-
induced weight gain (odds ratio [OR] = 6) [99]. order, mood disorders, and impulse control disorders
Another candidate that has been investigated with [113–115]. Similar to other drugs, inter-individual
regards to weight gain is the leptin gene. Results from differences exist in drug response and tolerability,
a nine-month antipsychotic treatment trial suggest a at least partially related to genetic factors [116].
significant association between the 2548-A/G poly- Analogous to the pharamacokinetic pharmacoge-
morphism and weight gain [98]; however, no biological netics of antidepressants and antipsychotics, variation
functional data on this SNP has been reported. Several in metabolizing enzymes play an important role in
other genes have also investigated with regards to anti- determining successful treatment with certain anti-
psychotic weight gain; however, results await con- convulsants, including phenytoin, carbamazepine,
firmation and are reviewed in detail elsewhere [100]. and benzodiazepines. In particular genetic variants
TD and abnormal extra pyramidal movements are in the CYP2C19 and CYP3A4/5 pathway have been
serious side effects observed in some patients treated documented to influence benzodiazepine plasma
chronically with antipsychotic drugs. Although the levels and related adverse events [117]. While the extent
mechanism of TD remains unknown, symptom of pharmacogenetic factors for other anticonvulsants,
occurrence has been directly associated with drug such as valproate, lamotrigine, gabapentin, pregabalin,
dosage and plasma levels [52]. Several genes have and topiramate remains to be determined, accumu-
been investigated as potential candidates for antipsy- lating evidence suggests genetic factors contribute
chotic-induced TD [57, 101]. One of the most prom- to lithium response and carbamazepine-induced
ising findings has been the association between the Stevens–Johnson syndrome.
Ser9Gly variant in the dopamine D3 receptor gene
(DRD3) and antipsychotic-induced movement dis-
orders [102–104]. A combined analysis of 780 patients
from different ethnic groups showed that the DRD3 Pharmacogenetic studies of lithium
Gly allele increased susceptibility to TD with a pooled Lithium is an alkali metal that is used in salt form,
OR of 1.33 [105]. A recent meta-analysis further such as lithium carbonate and lithium citrate, for
confirmed a significant contribution of the DRD3 the treatment of bipolar disorder. The therapeutic
Gly allele with an increased risk of developing TD with efficacy of lithium for bipolar illness was first dis-
an OR of 1.17 [106]. While these data support the covered by John Cade in 1949 [118] and further
DRD3 Ser9Gly polymorphism as a pharmacogenetic systematically evaluated by Mogens Schou in a series
marker for TD, the modest ORs indicate a weak effect of clinical trials [119]. Lithium is currently considered
and it is likely that other genetic variants are involved first-line treatment for bipolar disorder, in particular
in the complex pathophysiology of TD. for patients with euphoric mania. Epidemiological
Antipsychotic-induced agranulocytosis is a rare studies in bipolar disorder have consistently docu-
(0.7–3.0%) but severe adverse event, most often mented a strong genetic component to its etiology,
observed in patients treated with clozapine [107]. with heritability estimates ranging from 65 to 80%
While the etiology of this severe side effect remains [120]. Despite these strong genetic predispositions,
elusive, some reports have indicated an involvement identification of susceptibility alleles for bipolar dis-
of immune mechanisms. In particular, genes coding order has been difficult due to the complex mode
for human leukocyte antigens (HLA) have been asso- of inheritance, moderate effect sizes and genetic and
ciated with clozapine-induced agranulocytosis [108– clinical heterogeneity. Clinical heterogeneity is par-
112]. However, as with other antipsychotic-induced ticularly apparent with regards to treatment response.
adverse events, the genetic architecture of drug- Some patients respond remarkably well on lithium
induced agranulocytosis is complex, likely involving monotherapy, while for others the medication has
multiple genes, gene–gene, and gene–environment– no benefit [121]. The “endophenotype” of lithium
drug interactions. responsiveness has been used diagnostically in clinical

59
Chapter 6: Pharmacogenetics in psychiatry

practice and several family studies have suggested that response to lithium augmentation in acutely depressed
lithium response is a heritable trait. Mendlewicz antidepressant nonresponders, expanding the potential
carried out the first family study of lithium response role of lithium responsiveness across diagnostic
and showed a higher risk for bipolar disorder in first- categories [134]. Several other candidate genes have
and second-degree relatives of responders compared been investigated with regards to pharmacogenetics
to nonresponders [122]. Since then, several other of lithium response [135, 136], most of them involved
studies replicated this finding [123–125]. Based on in monoamine neurotransmission, including COMT,
the high degree of heritability of bipolar disorder, DRD2, DRD3, MAOA, serotonin transporter, and
and more so on evidence for a role of genetics on receptor genes [137–139]. So far no clear correlation
lithium responsiveness, pharmacogenetic studies are between variants in these genes and lithium response
logical steps in the dissection of bipolar disorder has been established and caution should be used in
pathophysiology. Similar to the early genetic studies interpretation of the results given the significant clin-
of bipolar disorder investigating disease susceptibility ical heterogeneity and differences in treatment
loci via molecular linkage methods [120], the earliest response criteria between studies [17, 137, 140–158].
pharmacogenetic studies of lithium response were While the field of psychiatric genetics has advanced
pedigree linkage studies. A genome-wide linkage to GWAS, no data exist for lithium response. However,
study of bipolar families from Quebec with positive a recent GWAS on bipolar disorder [159] genotyped
lithium response found a major locus on chromo- over 550 000 SNPs in two independent case-control
some 12q23-q24 [126]. Another linkage study of samples of European origin. Several genes were iden-
31 families with lithium responsiveness reported sig- tified, with rather modest effects. Nevertheless, the
nificant results for chromosome 15q14 and suggestive strongest result of this study was related to variation
results for 7q11 [127]. Despite these interesting data, in the diacylglycerol kinase, eta (DGKH) gene,
none of the linkage regions has been confirmed involved in the lithium-sensitive phosphatidyl inositol
independently. pathway. Although the field of lithium response
The search for genetic factors involved in lithium genetics is mostly unexplored, results obtained so far
response has also used a candidate gene approach. are encouraging. Future studies will have to include
This methodology depends upon biological collateral careful characterization of the bipolar phenotype,
data to inform potential genetic targets of investiga- improved methodological quality and standardized
tion. Several biological pathways have been suggested criteria for lithium response.
to play an important role in the mechanism of action
of lithium [128]. The phosphoinositide pathway has Pharmacogenetic studies of carbamazepine-
been shown to be affected by lithium, which causes
inhibition of inositol monophosphatase and inositol induced Stevens–Johnson syndrome
polyphosphate 1-polyphosphatase [129]. Inhibition of Carbamazepine (CBZ) is an important treatment for
these enzymes by lithium causes a reduction in the seizure disorders, bipolar disorder and chronic pain.
amount of free inositol available for regeneration of Tolerability varies among patients and common side
phosphatidylinositol biphosphate, ultimately leading effects include drowsiness, headaches, motor coordin-
to diminishing cellular levels of inositol triphosphate ation impairment and/or upset stomach. Beside these
and diacylglycerol. In addition, lithium’s inhibitory common side effects, several serious adverse events
effects on glycogen synthase kinase 3 [130], with have been reported under CBZ therapy, such as aplas-
downstream effects on transcription, are thought to tic anemia, fatal arrhythmias, and life-threatening
be involved in its clinical effectiveness. Based on these cutaneous disorders. Stevens–Johnson syndrome
data, recent studies have investigated an association (SJS) and toxic epidermal necrolysis (TEN) have both
between the promoter -50T/C polymorphism in the been associated with CBZ use. The incidence of CBZ-
GSK3-b gene and lithium response. While two associ- induced SJS/TEN is approximately 1–6 per 10 000
ation studies documented a relationship between the of new users in mainly Caucasian countries and is
promoter variant and lithium prophylaxis [131, 132], 10-fold higher in some Asian countries [160]. The
another group failed to replicate the finding [133]. skin lesions of SJS and TEN consist of blisters that
Interestingly, a more recent study documented an arise on erythematous or purpuric macules and
association between the GSK3-b -50T/C SNP and involve two or more mucosal surfaces. Both disorders

60
Chapter 6: Pharmacogenetics in psychiatry

carry a mortality of almost 30% [161, 162]. While Koreans. In Caucasians, African-Americans, Hispanics,
these two disorders are considered distinct entities, and Native Americans, the HLA-B*1502 allele is rare
they lie on a continuous clinical spectrum, with TEN and below 2%; however, limited data exist to estimate
being considered the most severe form [163]. The allele frequencies accurately in these populations
pathophysiological mechanisms for the development [170, 171].
of SJS/TEN remain elusive, although current evidence Based on these compelling data, the FDA released
suggests an immune-complex-mediated hypersensi- a warning in 2007 stating that serious and potentially
tivity with subsequent keratinocyte cell apoptosis. fatal skin reactions may occur with the administration
Recently, a strong association between the major of CBZ in patients positive for the HLA-B*1502 allele.
histocompatibility complex HLA-B*1502 allele and In addition the FDA recommended genotyping
CBZ-induced SJS among Han Chinese has been iden- patients of Asian descent before initiation of CBZ
tified [164]. The initial study included 44 patients that treatment [172]. While the strong link between the
developed SJS under CBZ and 101 controls that toler- HLA-B*1502 marker and CBZ-induced SJS is the first
ated CBZ well. All subjects were Han Chinese patients robust and clinically relevant example of the use of
living in Taiwan. Genotyping of 157 SNPs in candi- pharmacogenetics, future studies not only have to
date genes was carried out and results showed that replicate above findings in large cohorts of different
all 44 patients with SJS were positive for the HLA- ethnic background, they also have to carefully exam-
B*1502 but only 3% of CBZ-treated patients without ine potential class effects of compounds with similar
SJS had the same allele. The allele was present in 8.6% chemical structures. It will be important for future
of normal controls from the general population. studies to investigate the extent of the risk for a severe
The OR for developing CBZ-induced SJS if positive skin reaction in HLA-B*1502 carrier of non-Asian
for HLA-B*1502 was  2500 with a positive predictive descent.
value of 93.6% [164]. A follow-up study by the same
research group included 16 new CBZ-induced SJS
cases of Chinese descent, again demonstrating a
Future directions and implications
strong association of the HLA-B*1502 allele [165]. for clinical practice
Another smaller study from Hong Kong included The field of psychiatric pharmacogenetics is develop-
eight Han Chinese patients with CBZ-induced SJS ing rapidly and there is hope for identifying genetic
and reported that six of them had the HLA-B*1502 variants that are clinically meaningful. This expect-
allele [166]. Furthermore, Locharernkul et al. [167] ation is in particular supported by the strong associ-
showed for the first time a strong association between ation between the HLA-B*1502 allele and CBZ-
the HLA-B*1502 allele and CBZ-induced SJS in the induced SJS, which not only showed a large effect in
Thai population, demonstrating an effect of the allele a subset of patients with a very specific adverse event,
in a non-Chinese population. In contrast to these but also demonstrated that it is possible to identify
robust positive reports, analysis of European patients clinically relevant pharmacogenetic associations in small
with CBZ-induced SJS failed so far to identify a clear to moderate-sized groups of patients. “Pharmaco-
relationship between HLA-B*1502 and development genetics” by definition has the conceptual problems
of this severe adverse event [168, 169]. One possible of both disciplines. Psychopharmacology has been
reason that no clear association was found between plagued traditionally by high placebo response rates,
HLA status and phenotype might be due to the low poor treatment response rates, and low remission
HLA-B*1502 allele frequency in patients of European rates. For example, pharmaceutical trials in depres-
descent. In fact, the frequency of HLA-B*1502 varies sion show response rates of about 50%, remission
greatly across ethnic groups. Individuals of Han rates of about 30% [26], and a placebo response rate
Chinese ancestry show allele frequencies of greater of 30–60% [173–175]. Furthermore, the influence of
than 15% in individuals from Hong Kong, Thailand, environmental factors on drug response is poorly
Malaysia, and parts of the Philippines. Other Asian understood and thus difficult to control. The field
ethnic groups show less frequent occurrence of the of genetics has been challenged by lack of power in
HLA-B*1502 allele with estimates of 9–10% in Taiwan- samples studied, clinical and genetic heterogeneity,
ese individuals, 4% in North Chinese, 2–4% in South and lack of biological functional data on genetic vari-
Asians and Indians, and less than 1% in Japanese and ants. Genetic studies can only suggest an association

61
Table 6.1 Pharmacogenetic tests in clinical practice.

Genetic test Gene Drug Useful for Commercially Notes


available
Cytochrome CYP450 2D6 Antidepressants  Identifying patients who are poor, Roche AmpliChip; FDA approved
P450 Drug intermediate, extensive, or ultrarapid available in special
Mainly metabolized:
metabolizer metabolizers of psychotropic drugs laboratories
desipramine
status metabolized by CYP 2D6 and 2C19
fluoxetine
 Adjusting dosages for psychotropic
nortriptyline
drugs that are metabolized by
paroxetine CYP 2D6, 2C19
Partly metabolized:
amitriptyline
bupropion
citalopram
duloxetine HCL
escitalopram
fluvoxamine
maprotiline
mirtazepine
venlafaxine

Antipsychotics
Mainly metabolized:
aripiprazole
fluphenazine
perphenazine
risperidone
thioridazine

Partly metabolized:
chlorpromazine
haloperidol
olanzapine
Stimulants
Mainly metabolized:
atomoxetine

Partly metabolized:
amphetamine/
dextroamphetamine
dextroamphetamine
Cytochrome CYP 450 amitriptyline  Identifying patients who are poor, Roche AmpliChip; FDA approved
P450 Drug 2C19 citalopram intermediate, extensive, or ultrarapid Available in special
metabolizer clomipramine metabolizers of psychotropic drugs laboratories
status escitalopram metabolized by CYP 2D6 and 2C19
imipramine  Adjusting dosages for psychotropic
diazepam drugs that are metabolized by CYP
2D6, 2C19
CBZ-induced HLA-B*1502 carbamazepine  Identifying patients who are at Available in special FDA recommended for
SJS increased risk for CBZ-induced SJS laboratories patients of Asian
 Patients who test positive for descent started on CBZ
HLA-B*1502 may be at increased risk
of SJS/TEN from other antiepileptic
drugs that have been associated with
SJS/TEN. Therefore, in HLA-B*1502-
positive patients, consider avoiding use
of other antiepileptic drugs associated
with SJS/TEN when alternative therapies
are equally acceptable
Serotonin SLC6A4 fluoxetine  Predicting response time to Available in special
transporter fluvoxamine improvement with SSRIs (L-allele) laboratories
escitalopram  Identifying patients who have increased
sertraline risk of side effects under SSRI treatment
citalopram (S/S, S/L genotype)
paroxetine  Identifying patients who have reduced
amounts of the serotonin transporter
and thus might have an altered
response to SSRIs
 Evaluating patients who have failed
therapy with SSRIs and
who might respond favorably to
nonselective antidepressants (S-allele)
Table 6.1 (cont.)

Genetic test Gene Drug Useful for Commercially Notes


available
Serotonin HTR2A fluoxetine  Guiding treatment choice of an SSRI or Available in special
receptor HTR2C fluvoxamine non SSRI antidepressant laboratories
escitalopram  Guiding treatment choice in individuals
sertraline who have drug-metabolizer
citalopram phenotypes discordant with CYP450
paroxetine genotypes
 Identifying patients who may benefit
from treatment with the atypical
antipsychotic clozapine
Clozapine- HLA-DQB1 Clozapine  Identifying patients who are at PGxPredict:Clozapine
induced increased risk for clozapine-induced
Agranulocytosis agranulocytosis
 guiding decisions about the frequency
of hematological monitoring, and about
treatment decisions in the face of falling
white blood cell counts
 a negative test does not eliminate the
need for blood monitoring
Abbreviations: CBZ, carbamazepine; FDA, US Food and Drug Administration; SJS, Stevens–Johnson syndrome; SSRI, selective serotonin reuptake inhibitor; TEN, toxic epidermal necrolysis.
Chapter 6: Pharmacogenetics in psychiatry

between a polymorphism and treatment. Currently, poor or ultra-rapid metabolizers, since some individ-
very little is known about the biological relevance uals might have rare, previously unknown, gene
of most genetic polymorphisms and how genetic vari- variants that drive their metabolizer status. In addi-
ation influences treatment response on the cellular tion, cost effectiveness and long-term benefits for
level. In the future, SNPs associated with drug treat- patients taking antidepressant drugs have yet to be
ment response might lead to promising new drug established. Several laboratories now also offer the
development, but first these findings must undergo HLA-B*1502 allele test for CBZ-induced SJS, and
extensive validation and neurobiological investiga- additional tests are anticipated to become available
tion. To address these issues, there is a strong need in the near future.
to develop appropriate and standardized methodolo- Although pharmacogenetics in clinical practice is
gies for pharmacogenetic studies, as proposed by currently limited to “side effects” and “metabolism”,
some groups [36]. It is necessary to develop prospect- comprehensive pharmacogenetic profiling is already a
ive large pharmacogenetic clinical trials to evaluate reality in many pharmaceutical companies. Many
the effects of genetic variants on treatment outcome phase II and phase III trials now have genetic com-
comprehensively. In addition, other questions should ponents and the FDA has recently issued the first
be addressed, including investigation of the genetics genotype-based indications for warfarin [178]. This
of placebo response, genetics of psychotherapy response, paradigm shift offers great opportunities in identify-
or the genetics of treatment response in general. ing particular subgroups of patients for which a
Despite many obstacles, it is likely that genetic compound works especially well or causes severe side
patient information will influence clinical practice effects. On the other hand, pharmaceutical companies
in the very near future. Currently, there are only a might limit their phase II and III trials to certain
few commercial pharmacogenetic tests available “genetic” populations, in order to document safety
(Table 6.1) which can be ordered through a few com- and efficacy, leaving patients with “complicated risk
mercial and academic laboratories. The Roche Diag- genetics” out of trials and drug development. It is
nostic AmpliChip CYP450 test was FDA approved in thus important to develop comprehensive pharmaco-
2005 and provides genotypes for the two cytochrome genetic policies and regulations in order to avoid
P450 genes CYP2D6 and CYP2C19. Theoretically, misuse of genetic information [179]. While psychiatry
by genotyping patients for variation in these genes, has entered the new area of pharmacogenetics, it is
the metabolizer status of a patient may be predicted important to remember that this new technology will
and this can be used to help guide medication choice only provide additional information on one aspect
and dosing. This might be useful for example in a of the complex and personal history of psychiatric
depressed patient who has a history of being very patients. It is the sum of inside and outside factors
sensitive to antidepressant medication preferentially that contribute and influence mental pathology and
metabolized through CYP2D6, or in a depressed well-being.
patient who appears treatment resistant despite
adequate dosing of antidepressant drugs. Currently
there are no universal recommendations regarding Acknowledgements
which patients should get tested, but several sugges- This work was supported by the Center for Neuro-
tions have been published [176, 177]. Limitations of biology and Behavior, Department of Psychiatry,
the AmpliChip test include false positive and false University of Pennsylvania. The author would like
negative results, an issue with every laboratory test, to thank Thomas Ferraro for very helpful comments,
but more importantly the test does not identify all suggestions, improvements, and corrections.

References 4. Binder EB, et al. Ann Med


2006;38:82–94.
7. Weinshilboum RM, et al. Annu
Rev Genomics Hum Genet
1. Kessler RC, et al. Arch Gen 2006;7:223–245.
Psychiatry 2005;62:617–627. 5. Serretti A, et al.
Psychopharmacology (Berl) 8. Dalen P, et al. Clin Pharmacol
2. Rush AJ, et al. Am J Psychiatry 2004;174:490–503. Ther 1998;63:444–452.
2006;163:1905–1917.
6. Staddon S, et al. 9. Schenk PW, et al. Mol Psychiatry
3. Trivedi MH, et al. Am J Psychiatry Psychopharmacology (Berl) 2008;13:597–605.
2006;163:28–40. 2002;162:18–23.

65
Chapter 6: Pharmacogenetics in psychiatry

10. Bertilsson L, et al. Br J Clin 33. Hu XZ, et al. Arch Gen Psychiatry 54. Miyamoto S, et al. Mol Psychiatry
Pharmacol 2002;53:111–122. 2007;64:783–792. 2005;10:79–104.
11. Charlier C, et al. Ther Drug Monit 34. Laje G, et al. Am J Psychiatry 55. Cascorbi I. Pharmacol Ther
2003;25:738–742. 2007;164:1530–1538. 2006;112:457–473.
12. Shams ME, et al. J Clin Pharm 35. Ising M, et al. Arch Gen Psychiatry 56. Linnet K, et al. Eur
Ther 2006;31:493–502. 2009;66:966–975. Neuropsychopharmacol 2008;
13. Veefkind AH, et al. Ther Drug 18:157–169.
36. Serretti A, et al. Pharmacogenomics
Monit 2000;22:202–208. J 2008;8:90–100. 57. Arranz MJ, et al. Mol Psychiatry
14. Perry PJ, et al. J Clin 2007;12:707–747.
37. Uher R, et al. Psychol Med
Psychopharmacol 1994;14: 2008;38:289–300. 58. Spina E, et al. Basic Clin
230–240. Pharmacol Toxicol 2007;
38. Uher R, et al. Pharmacogenomics J 100:4–22.
15. Gex-Fabry M, et al. Ther Drug 2009;9:225–233.
Monit 2008;30:474–482. 59. Urichuk L, et al. Curr Drug Metab
39. Murphy GM, Jr., et al. Am 2008;9:410–418.
16. Alvares AP, et al. Drug Metab Rev
J Psychiatry 2003;160:1830–1835.
1979;9:185–205. 60. Kirchheiner J, et al. J Clin
40. Murphy GM, Jr., et al. Arch Gen Psychopharmacol 2006;26:
17. Serretti A, et al. Pharmacogenomics
Psychiatry 2004;61:1163–1169. 440–442.
J 2004;4:267–273.
41. Smits K, et al. Int Clin 61. Eiermann B, et al. Br J Clin
18. Lesch KP, et al. Science
Psychopharmacol 2007;22: Pharmacol 1997;44:439–446.
1996;274:1527–1531.
137–143.
19. Gelernter J, et al. Am J Med Genet 62. Ring BJ, et al. J Pharmacol Exp
42. FDA. http://www.fda.gov/Drugs/ Ther 1996;276:658–666.
1999;88:61–66.
DrugSafety/PostmarketDrugSafety
20. Baune BT, et al. Depress Anxiety 63. Murayama N, et al. J Pharmacol
InformationforParients and
2008;25:920–925. Exp Ther 2004;308:300–306.
Providers/DrugSafetyInformation
21. Mrazek DA, et al. Am J Med Genet forHealthcareProfessionals/Public 64. Sachse C, et al. Br J Clin
B Neuropsychiatr Genet HealthAdvisories/ucm161679.htm; Pharmacol 1999;47:445–449.
2008;147B:1337–1344. 2004. 65. Eap CB, et al. J Clin
22. Serretti A, et al. Prog 43. Jick H, et al. JAMA 2004;292: Psychopharmacol 2004;24:214–219.
Neuropsychopharmacol Biol 338–343. 66. Ozdemir V, et al. J Clin
Psychiatry 2005;29:1074–1084. 44. Licinio J, et al. Nat Rev Drug Psychopharmacol 2001;21:603–607.
23. Serretti A, et al. Mol Psychiatry Discov 2005;4:165–171. 67. Bondolfi G, et al. Ther Drug Monit
2007;12:247–257. 45. Gibbons RD, et al. Am J Psychiatry 2005;27:539–543.
24. Lin E, et al. Pharmacogenomics 2007;164:1356–1363. 68. Buckley PF. J Clin Psychiatry
2008;9:935–946. 46. Perlis RH, et al. Arch Gen 2007;68(Suppl 6):5–9.
25. Perlis RH, et al. Arch Gen Psychiatry 2008;65:882–892. 69. Gardner DM, et al. CMAJ
Psychiatry 2007;64:689–697. 47. Laje G, et al. Pharmacogenet 2005;172:1703–1711.
26. Rush AJ, et al. Control Clin Trials Genomics 2009;19:666–674. 70. Tamminga CA, et al. Mol
2004;25:119–142. 48. Perroud N, et al. Psychiatry 2005;10:27–39.
27. McMahon FJ, et al. Am J Hum Neuropsychopharmacology 71. Kapur S, et al. Am J Psychiatry
Genet 2006;78:804–814. 2009;34:2517–2528. 2000;157:514–520.
28. Manji HK, et al. Nat Med 49. Henderson DC. J Clin Psychiatry 72. Mamo D, et al. Am J Psychiatry
2001;7:541–547. 2008;69:e04. 2004;161:818–825.
29. Strome EM, et al. Biol Psychiatry 50. Scheen AJ, et al. Diabetes Metab 73. Himei A, et al. Psychiatry Clin
2005;57:1004–1010. 2007;33:169–175. Neurosci 2002;56:97–102.
30. Lekman M, et al. Biol Psychiatry 51. Meltzer HY, et al. J Clin Psychiatry 74. Lencz T, et al. Am J Psychiatry
2008;63:1103–1110. 2002;63(Suppl 9):25–29. 2006;163:529–531.
31. Paddock S, et al. Am J Psychiatry 52. Kane JM. Am J Psychiatry 75. Wu S, et al. Neurosci Lett
2007;164:1181–1188. 2006;163:1316–1318. 2005;376:1–4.
32. Kraft JB, et al. Biol Psychiatry 53. Lieberman JA, et al. N Engl J Med 76. Jonsson EG, et al. Mol Psychiatry
2007;61:734–742. 2005;353:1209–1223. 1999;4:290–296.

66
Chapter 6: Pharmacogenetics in psychiatry

77. Lundstrom K, et al. Biochem 98. Templeman LA, et al. 120. Lohoff FW, et al. In Charney D,
Biophys Res Commun Pharmacogenet Genomics et al. (eds.). Neurobiology of
1996;225:1068–1072. 2005;15:195–200. Mental Illness. Oxford: Oxford
78. Joober R, et al. J Psychiatr Res 99. Reynolds GP, et al. Lancet University Press; 2008.
2000;34:285–291. 2002;359:2086–2087. 121. Fountoulakis KN, et al. Int
79. Scharfetter J, et al. Eur 100. Muller DJ, et al. Pharmacogenomics J Neuropsychopharmacol
Neuropsychopharmacol 2006;7:863–887. 2008;11:999–1029.
1999;10:17–20. 101. Muller DJ, et al. Pharmacogenomics 122. Mendlewicz J, et al. Am
80. Szekeres G, et al. Am J Med Genet J 2004;4:77–87. J Psychiatry 1973;130:1011–1013.
B Neuropsychiatr Genet 123. Grof P, et al. J Affect Disord
102. Liao DL, et al. Neuropsychobiology
2004;124B:1–5. 1994;32:85–95.
2001;44:95–98.
81. Reynolds GP, et al. Eur 124. Prien RF, et al. Arch Gen
103. Steen VM, et al. Mol Psychiatry
Neuropsychopharmacol Psychiatry 1974;31:189–192.
1997;2:139–145.
2005;15:143–151.
104. Segman R, et al. Mol Psychiatry 125. Smeraldi E, et al. J Affect Disord
82. Meltzer HY, et al. 1984;6:139–151.
Arzneimittelforschung 1999;4:247–253.
126. Morissette J, et al. Am J Med Genet
1992;42:268–272. 105. Lerer B, et al.
1999;88:567–587.
83. Arranz M, et al. Lancet Neuropsychopharmacology
1995;346:281–282. 2002;27:105–119. 127. Turecki G, et al. Mol Psychiatry
2001;6:570–578.
84. Arranz MJ, et al. Schizophr Res 106. Bakker PR, et al. Schizophr Res
1998;32:93–99. 2006;83:185–192. 128. Coyle JT, et al. Nat Med
2002;8:557–558.
85. Joober R, et al. J Psychiatry 107. Munro J, et al. Br J Psychiatry
Neurosci 1999;24:141–146. 1999;175:576–580. 129. Berridge MJ, et al. Cell
1989;59:411–419.
86. Lane HY, et al. Am J Psychiatry 108. Amar A, et al. Int
2002;159:1593–1595. J Neuropsychopharmacol 130. Klein PS, et al. Proc Natl Acad Sci
1998;1:41–44. U S A 1996;93:8455–8459.
87. Lin CH, et al. Neuroreport
1999;10:57–60. 109. Dettling M, et al. Pharmacogenomics 131. Benedetti F, et al. Neurosci Lett
J 2007;7:325–332. 2004;355:37–40.
88. Yu YW, et al. Neuropsychobiology
2001;43:79–82. 110. Dettling M, et al. Pharmacogenetics 132. Benedetti F, et al. Neurosci Lett
2001;11:135–141. 2005;376:51–55.
89. Arranz MJ, et al. Neurosci Lett
1996;217:177–178. 111. Lieberman JA, et al. 133. Szczepankiewicz A, et al. World
Arch Gen Psychiatry J Biol Psychiatry 2006;7:158–161.
90. Masellis M, et al.
Neuropsychopharmacology 1990;47:945–948. 134. Adli M, et al. Biol Psychiatry
1998;19:123–132. 112. Valevski A, et al. Eur 2007;62:1295–1302.
91. Correll CU, et al. Schizophr Res J Immunogenet 1998; 135. Mamdani F, et al.
2007;89:91–100. 25:11–13. Pharmacogenomics J 2004;
113. Bowden CL. Expert Rev Neurother 4:161–170.
92. Theisen FM, et al. Psychiatr Genet
2005;15:285–289. 2007;7:9–16. 136. Serretti A, et al. Am
114. Ettinger AB. Neurology J Pharmacogenomics 2003;3:
93. Wehmeier PM, et al. Psychiatry
2006;67:1916–1925. 17–30.
Res 2005;133:273–276.
115. Post RM, et al. Neuropsychobiology 137. Serretti A, et al. Psychiatry Res
94. Ellingrod VL, et al. Am J Med
1998;38:152–166. 1999;87:7–19.
Genet B Neuropsychiatr Genet
2005;134B:76–78. 116. Ferraro TN, et al. Epilepsy Behav 138. Serretti A, et al. Am J Med Genet
2005;7:18–36. 2002;114:370–379.
95. Miller DD, et al. Am J Med Genet
B Neuropsychiatr Genet 117. Fukasawa T, et al. J Clin Pharm 139. Serretti A, et al. J Psychiatr Res
2005;133B:97–100. Ther 2007;32:333–341. 2000;34:89–98.
96. Reynolds GP, et al. 118. Cade JF. Med J Aust 1949; 140. Benmessaoud D, et al. BMC
J Psychopharmacol 2006;20:15–18. 2:349–352. Psychiatry 2008;8:40.
97. Reynolds GP, et al. Am 119. Schou M. Nervenarzt 141. Bremer T, et al. Mol Diagn Ther
J Psychiatry 2003;160:677–679. 1983;54:331–339. 2007;11:161–170.

67
Chapter 6: Pharmacogenetics in psychiatry

142. Cavazzoni P, et al. Psychiatry Res 154. Serretti A, et al. J Psychiatr Res 168. Alfirevic A, et al. Pharmacogenomics
1996;64:91–96. 1999;33:371–377. 2006;7:813–818.
143. Dmitrzak-Weglarz M, et al. 155. Serretti A, et al. 169. Lonjou C, et al. Pharmacogenet
Pharmacogenomics 2008;9: Pharmacogenomics J 2001; Genomics 2008;18:99–107.
1595–1603. 1:71–77. 170. Ferrell PB, Jr., et al.
144. Dmitrzak-Weglarz M, et al. 156. Squassina A, et al. Pharmacol Res Pharmacogenomics 2008;9:
Pharmacol Rep 2005;57:761–765. 2008;57:369–373. 1543–1546.
145. Duffy A, et al. J Psychiatry 157. Turecki G, et al. Psychiatry Res 171. Lonjou C, et al. Pharmacogenomics
Neurosci 2000;25:353–358. 1996;63:17–23. J 2006;6:265–268.
146. Mamdani F, et al. Am J Med Genet 158. Zill P, et al. Psychiatr Genet 2003; 172. FDA. Safety Alerts for Drugs,
B Neuropsychiatr Genet 13:65–69. Biologics, Medical Devices, and
2008;147B:500–504. Dietary Supplements. Silver
159. Baum AE, et al. Mol Psychiatry
147. Masoliver E, et al. Spring, MD: US Food and Drug
2008;13:197–207.
Psychiatr Genet 2006; Administration; 2008.
160. Mockenhaupt M, et al. Neurology
16:25–29. 173. Charney DS, et al. Arch Gen
2005;64:1134–1138.
148. Masui T, et al. Int Psychiatry 2002;59:262–270.
J Neuropsychopharmacol 161. Rzany B, et al. Lancet 174. Khan A, et al. J Psychiatr Res
2006;9:83–88. 1999;353:2190–2194. 2008;42:791–796.
149. Masui T, et al. Psychiatr Genet 162. Tennis P, et al. Neurology 175. Montgomery SA. Eur
2006;16:49–50. 1997;49:542–546. Neuropsychopharmacol
150. Masui T, et al. Prog 163. Bastuji-Garin S, et al. Arch 1999;9:265–269.
Neuropsychopharmacol Biol Dermatol 1993;129:92–96. 176. de Leon J, et al. Psychosomatics
Psychiatry 2008;32:204–208. 164. Chung WH, et al. Nature 2006;47:75–85.
151. Michelon L, et al. Neurosci Lett 2004;428:486. 177. Mrazek DA. CNS Spectr
2006;403:288–293. 165. Hung SI, et al. Pharmacogenet 2006;11:3–4.
152. Rybakowski JK, et al. Pharmacol Genomics 2006;16:297–306. 178. Tan GM, et al. Pharmacogenomics
Rep 2005;57:124–127. 166. Man CB, et al. Epilepsia 2010;11:439–448.
153. Rybakowski JK, et al. 2007;48:1015–1018. 179. Katsanis SH, et al. Science
Pharmacopsychiatry 2005; 167. Locharernkul C, et al. Epilepsia 2008;320:53–54.
38:166–170. 2008;49:2087–2091.

68
Functional validation of candidate
Chapter

7 genetic susceptibility factors for major


mental illnesses
From protein chemistry, cell biology,
and animal studies, to human brain imaging
Akira Sawa, Wanli W. Smith, Saurav Seshadri, Akiko Hayashi-Takagi,
Hanna Jaaro-Peled, and Atsushi Kamiya

Introduction studies on disrupted in schizophrenia 1 (DISC1) over


the past decade as examples [7, 9]. The gene coding
Genetic susceptibility factors for major mental illnesses, for DISC1 was originally identified at the breakpoint
such as schizophrenia (SZ) and bipolar disorder, have of a balanced chromosomal translocation in a Scottish
now become available due to the long-term efforts of pedigree in which many family members suffer from
psychiatric genetics [1, 2]. Recent progress in whole mental illnesses, such as major depression, SZ, and
genome association studies, and work on copy number bipolar disorder [10]. Genetic studies in the general
variations have further enhanced our knowledge of population have indicated that DISC1 affects endophe-
such genetic factors [3–5]. In contrast to Huntington’s notypes that underlie several major mental illnesses
disease (HD) in which a specific mutation in a single [7, 9]. Consequently, we can find genetic associations of
gene, huntingtin (htt), causes the disease [6], none of the DISC1 with a wide range of psychiatric disorders [7, 9].
susceptibility genes for major mental illnesses available Functional studies examining how such genetic
thus far is such a direct causal factor. Instead, com- factors influence the pathology of major mental illnesses
binations of genetic factors together with environ- are conducted at many levels. Most fundamentally, bio-
mental stressors underlie the pathogenesis of major chemical studies on protein interaction among these
mental illnesses [7]. Genetic variants in some of these factors are performed. Such molecular mechanisms are
factors occur commonly and result in mild risk in the first validated and characterized in cell models. Roles for
functional context; whereas other genetic variations such molecular mechanisms in the neuronal circuitry
occur very rarely or de novo, and result in greater are then tested in animal models. Finally, brain imaging
impact on biological processes underlying disease studies allow us to examine how genetic variations are
pathology, though they are not causal per se [8]. correlated with human brain functions in vivo.
How can we utilize such genetic information for a
better understanding of the biology underlying disease
pathology? The goal of this chapter is to address this Protein chemistry
important question. Before such genes became available, Once genetic studies identify candidate susceptibility
molecular studies for major mental illnesses had been genes, the first step of functional studies is to charac-
limited to pharmacological approaches, in which cellular terize the proteins that are encoded by these genes.
responses to drugs used in clinical settings were studied. As described above, no one factor by itself can cause
We will review functional studies based on genetic major mental illnesses. Instead, combinations of more
information for better understanding of major mental than one factor result in the diseases. Thus, a key
illnesses, such as SZ, by using the many biological question in psychiatric genetics is how such genetic

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

69
Chapter 7: Validation of candidate genetic susceptibility factors

Neurotrophic factors
BDNF EGFR
NT-3
NT-4
IGF-1
TRK EGFR AMPAR LICAM NMDAR EPHB2 DOPAMINE RECEPTORS GPCR

AKAP9
ADENYLATE CYALASE
KALRN
INTEGRIN CAMKK2 SNAP91 CDK5RAP3
RAC
MAP3K71P2

AB12 FLJ13386 CDKS cAMP


SPTBN1
NEF3 NCK
SOS1

SPTAN1 SNX6 PDE4B

TNIK
NDEL1 GNB1
TRAF3IP1
RAS SH3BP5 CDCSL
PAFAH1B1
ANK2
CLU

JNK2 JNK3 JNK1 DST

SEC3L1 DISC1 SRGAP3

TRIM9
MACF1
RHO GTPASES
TRI0
SNAP25

DMD
RHO GTPASES

Figure 7.1 Protein interactome simulated by pathway analyses. Iterative yeast-two hybrid screens, combined with detailed pathway, and
functional analyses suggest DISC1 interacts with proteins involved in key processes involved in neurodevelopment. Deficits in these processes
may underlie decreased dendritic branching, arborisations and neuropil size observed pathologically. (Adapted from [16], with permission.)
See plate section for color version.

susceptibility factors may have molecular and/or co-localization by using confocal microscopy and
functional interactions, which may result in the dual-label electron microscopy.
pathology. To explore possible protein interactions, DISC1 is an intracellular protein that contains
a yeast two-hybrid assay is frequently used [11]. This several coiled-coil domains. This structure suggested
assay is especially powerful in detecting intracellular that DISC1 might be an anchoring molecule that
protein interactions. Protein interactions suggested by would interact with many proteins, but at that time
yeast two-hybrid need to be further confirmed by no concrete indication of its function was available.
biochemical approaches, such as in vitro protein Therefore, several groups conducted yeast two-hybrid
binding assays with purified recombinant proteins screening, and have identified many interesting inter-
and co-immunoprecipitation. The former informs us actors, such as nuclear distribution element-like
whether such protein interactions are direct or indirect. (NDEL1), phosphodiesterase 4 (PDE4), and pericen-
By utilizing site-directed mutagenesis as well as gener- triolar material 1 (PCM1) [12–15]. One of the most
ating several deletion mutants, we can further deter- extensive studies using this approach is the work by
mine which small domains of the proteins are required Camargo et al. in which the possible interactome of
for their interaction. This information is very useful in DISC1 is suggested [16] (Figure 7.1). Importantly,
addressing the functional significance of the protein many of these protein interactors are also genetic
interactions, which are validated in cell and in vivo susceptibility factors for major mental illnesses pro-
animal models, as described in the following subsections. posed by independent genetic studies [17–20].
In addition to biochemical approaches, protein inter- At present, many investigators agree on a general
actions in the cellular context are assessed by their concept that genetic susceptibility factors for mental

70
Chapter 7: Validation of candidate genetic susceptibility factors

illnesses frequently interact with each other, forming Because these cells are selected from homogenous
meaningful “pathways” [7, 21]. colonies with ectopic plasmids (either for overexpres-
sion or knocking down the targets), the potential to
Cell biology expand the manipulated cells in a large scale and
homogeneous manner is an advantage of this system.
Once genetic studies identify candidate susceptibility
Thus, stable expression cell models can be effectively
factors for the diseases, functions of such proteins can
used for compound screening.
be tested in cells by modulating expression of the target
molecules or by expressing their genetic variants.
Types of cells used for assays
Molecular manipulations For molecular manipulations, cell lines are frequently
There are three types of molecular manipulations used. In general, we can introduce plasmids in cell
in cells. First, we can simply overexpress molecules lines by transfection with a high efficiency. For
of interest exogenously. Overexpression of wild-type example, PC12 cells, which can be differentiated from
proteins may be useful in addressing the cellular a proliferating to postmitotic status by reducing
effects of genomic duplication. If nonsynonymous serum concentration and adding nerve growth factor
genetic variations are available, overexpression of in media, are frequently used to address questions of
such mutants may address the effects of the genetic how target molecules may play roles in neurite out-
variations. Especially if biological significance of such growth. However, while cell lines are easy to use, they
genetic variants is the result of a dominant-negative may not truly reflect endogenous conditions, as the
mechanism, we frequently see meaningful phenotypic artificial processes used in their production, such as
changes following overexpression. Second, we can immortalization, may affect cellular signaling.
knockdown expression of target molecules by RNA To address this concern with cell lines, primary
interference (RNAi) or other approaches. This knock- cell cultures are useful. Established protocols for
down approach is regarded to be the best way to neuron, astrocyte, microglia, and oligodendrocyte
address the physiological roles of target molecules. cultures are available. If necessary, mixed culture with
As described below, if knockout mice are available, different types of cells is also possible. In recent years,
cells primarily prepared from such animals can be investigators have paid more attention to the use of
used for this aim. A knock-in approach, in which stem cells. A technical difficulty in using these cells
endogenous wild-type molecules are replaced by gen- (primary and stem cells) in molecular manipulations is
etic variants, is the third strategy. To address subtle the relatively lower transfection efficiency compared
functional differences between wild-type mice and with that for cell lines. To overcome this issue, virus-
mutants, this approach is more sensitive and more mediated gene transfer with vectors of lentivirus,
closely reflects endogenous conditions compared with retrovirus, and adeno-associated virus (AAV) is used.
mere overexpression of genetic variants. A knock-in In addition, in molecular psychiatry, it is espe-
approach can be used to address the significance of cially important to consider using cells derived from
protein interactions. Phenotypes elicited by knock- patients and controls. Many studies have used periph-
down expression of a target protein (protein X) can eral cells from human subjects, such as lymphocytes,
be rescued (normalized) by overexpressing wild-type lymphoblasts, and fibroblasts. To obtain neuronal-
protein X. If interaction of proteins X and Y is crucial, like cells in a minimally invasive manner, olfactory
mutant protein X that lacks the domain required for neurons obtained via nasal biopsy can be useful [22].
binding with protein Y may not rescue the phenotypes Furthermore, recent progress in stem cell biology
elicited by knockdown of endogenous protein X. allows us to establish induced pluripotent stem cells
Such molecular manipulations are usually per- (iPS cells) from peripheral cells, such as fibroblasts
formed by transfection of plasmids or infection of [23]. Although there are still many technical barriers,
viral vectors, and their effects are assayed within this iPS cell approach is regarded as the most promis-
several days. In contrast to this type of “transient” ing strategy to obtain any type of cells in the central
manipulation, some studies utilize “stable” expression nervous system in the near future. A derivative of this
cell models. Although lower throughput and labor- approach is the direct transformation of fibroblasts
intensive, these stable models have several advantages. into neurons [24].

71
Chapter 7: Validation of candidate genetic susceptibility factors

Table 7.1 In vitro cellular assays relevant to mental illnesses. Examples of possibly altered cellular events in individuals with mental
disorders, and assays used for detection of those deficits in cellular model system in vitro are shown.

Cellular event Assay


Neurotransmitters Production Concentration measurement in cell lysate
Release Concentration measurement in culture media
Uptake Fluorescence-based assay, Synaptosomal assay
Neurodevelopment Proliferation BrdU cell proliferation assay
Migration Wound healing assay, Boyden chamber assay
Neurite outgrowth Measurement of neurite length (neuroblastoma such as PC12,
SHSY-5Y cells)
Dendritic arborization Scholl assay (primary neuron culture)
Synaptogenesis/reorganisation Morphometry analysis (primary neuron culture)
Metabolism Oxidative stress Measurement for DNA/RNA damage, oxidative protein, ROS,
and lipid peroxidation

Readouts relevant to mental disorders many proteins. Indeed, yeast two-hybrid screening
and biochemical validations have identified many
(Table 7.1) proteins associated with DISC1. Although no func-
What are the relevant cellular assays associated with tion of DISC1 was originally available, cellular roles
mental illnesses? Mechanisms associated with neuro- for these protein interactors were known. Thus, to
transmitter production, release, and uptake may be address the possible function of DISC1, investigators
the first category of candidates. Recent studies of used assays that were relevant to the interactors. For
brain imaging and neuropathology in patients example, neurite outgrowth was used based on the
suffering from mental illnesses have suggested that observation that DISC1 binds with cytoskeletal regula-
there are anatomical and histological abnormalities tors, such as fasciculation and elongation protein zeta-1
originating from neurodevelopment. Thus, the key (FEZ1), lissencephaly 1 (LIS1), and NDEL1 [12, 27–29].
components underlying these changes, such as cell Interaction of PDE4 with DISC1 was the rationale
proliferation, cell migration, neurite outgrowth, den- that investigators used in choosing assays of cellular
dritic arborization, and synaptic spine formation, are response to cyclic adenosine monophosphate (cAMP)
the targets to be examined. Accumulating evidence in signaling [20, 30]. Likewise, assays of cell proliferation
clinical studies has also pointed out that mental ill- and Wnt signaling were based on interaction of GSK3b
nesses, especially schizophrenia, are associated with with DISC1 [31]. Assays of cAMP and GSK3b/Wnt
intrinsic abnormalities in metabolic signaling [25]. In signaling are also relevant in the light of clinical
accordance with this notion, some studies have pro- pharmacology. Aberrant connectivity of neuron net-
posed a role for oxidative stress in mental illnesses works is believed to underlie the pathology of major
[26]. Thus, effects of target molecules on cellular mental illnesses; especially in schizophrenia, decrease in
responses to metabolic and oxidative stress may be synaptic spine density is reproducibly reported [32].
candidate readouts relevant to mental disorders. A role for DISC1 in spine reorganization has been
established in primary neuron cultures [33]. In addi-
tion, cell models can be used to address the functional
Cell biology in addressing roles relationships of genetic susceptibility factors, even if
of genetic factors for mental illnesses: they do not directly bind at molecular levels: one
DISC1 studies as examples example is a study that demonstrates the effects of
Neuregulin-1 on expression of a DISC1 isoform [34].
When the gene coding for DISC1 was first reported
from a Scottish pedigree [10], nobody knew its cellu-
lar function. As described above, structural prediction Animal models
from the deduced amino acids suggested that it is an Although cell models are useful to identify the signal
intracellular anchoring molecule that interacts with transduction pathways involving genetic susceptibility

72
Chapter 7: Validation of candidate genetic susceptibility factors

factors that result in cellular manifestations, such cycle, optical clarity, rapid external embryonic devel-
as neurite outgrowth, cell proliferation, and synaptic opment, and multiple molecular genetic techniques.
spine reorganization, they cannot address the most A recent paper by Wood [42] studied the function
important question of psychiatry directly; that is, how of DISC1 and Neuregulin-1 in zebrafish, suggesting
do genetic susceptibility factors affect formation and that these factors affect similar neurodevelopmental
functions of neuronal circuitry in the brain in vivo? processes. With similar reasons, fruit fly Drosophila
Therefore, we need to generate animal models based melanogaster may be useful in testing functions of
on information from psychiatric genetics, in which genetic susceptibility factors for mental illnesses.
knowledge from cell biology should be fully utilized. In particular, many outstanding studies in memory,
Rodents, especially mice, are the species in which sleep, and circadian cycle have been done using the
genetic engineering technology is the most advanced. fruit fly. For example, Swamura et al. [43] reported
Thus, in testing brain functions of genetic suscepti- that transgenic files overexpressing DISCI display
bility factors for major mental illnesses, mice are the disturbance homeostasis.
first choice. In this chapter, we will describe rodent models
The onset of major mental illnesses (addictions, with manipulations for genetic susceptibility factors
SZ, mood disorders, and related conditions) occurs of mental illnesses in greater detail.
in adolescence or young adulthood. However, accu-
mulating evidence has suggested neurodevelopmen-
tal etiology of these disorders, especially in SZ [35, Molecular manipulations in rodents
36]. Recent genetic studies indicate that some of the As in cell models, there are three types of molecular
most promising susceptibility factors of these dis- manipulations in animal models. The first strategy is
eases, such as Neuregulin-1, ErbB4, and DISC1, play to establish gain-of-function models by exogenously
important roles in development [37, 38]. Thus, gen- overexpressing wild-type or mutant proteins under
etic model mice in which such development-related the control of appropriate promoters. The second
disease-associated factors are altered may be useful in one is to establish loss-of-function models in which
characterizing how the disease etiologies develop over target genes are knocked out or their expressions are
time until development of full-blown disease [7]. It is knocked down. The third strategy is knock-in, in
likely that nonlethal but significant insults in early which endogenous wild-type molecules are replaced
development affect postnatal brain maturation, which, by genetic variants.
in turn, results in manifestation of psychiatric symp- Mouse genetic engineering techniques allow us to
toms after puberty. Early intervention in these diseases, generate hereditary mouse models, such as transgenic,
including preventive medication to prodromal subjects knockout, and knock-in mice in which these three
is a hot topic in translational research [39]. To address types of molecular manipulations are performed,
these questions, although we could design longitudinal respectively. In transgenic mice, exogenous protein
studies with human subjects, research with genetic expression can be spatially and temporally controlled
mouse models has significant advantages. by choosing an appropriate promoter. Such pro-
There is still debate about whether it is possible to moters include the a-CaMKII promoter mainly for
use rodents to model psychiatric disorders in which the pyramidal neurons and the GAD67 promoter for
high brain functions, that are probably in part unique interneurons. Furthermore, by combining this with a
to humans, are impaired [40]. Thus, genetic manipu- tetracycline-inducible system, an additional temporal
lations in primates by using AAV technology com- control in expression is available [44]. In knockout
bined with stereotaxic injection into targeted brain mice, spatial and temporal control of knockout
regions are attempted [41]. Although the primate becomes possible by utilizing the Cre-loxP system
model has high potential, we should also acknowledge [45]. Such spatial and temporal expression control is
its current limitations, including high costs, shortage very important in considering neural circuitry-
of experimental tools, and ethical concerns. dependent functional validation.
On the other hand, investigators have also paid In addition, by injecting plasmids or virus-mediated
attention to zebrafish (Danio rerio), a vertebrate lower constructs, expression of target molecules can be modu-
than rodents, as a model for functional studies due to lated (Figure 7.2) [46, 47]. First, in utero gene transfer
its advantages of low cost maintenance, rapid life is a technique to modulate expression of target genes

73
Chapter 7: Validation of candidate genetic susceptibility factors

(a) Embryo Figure 7.2 Brain region specific


gene manipulations. (a) Schematic
representation of bilateral in utero
injection of constructs followed by their
E14 incorporation by electroporation into
Injection progenitor cells in the ventricular zone at
– + embryonic day 14 (E14). Migrating cells
+ + with green fluorescent protein (GFP) are
visualized at E18 after injection of a GFP
expression construct. (Adapted from [46],
with permission.) (b) Stereotaxic
coordinates and actual injection of
lentivirus-based enhanced GFP (EGFP).
– – Stereotaxic coordinates were determined
from the rat brain in stereotaxic
coordinates. The coordination in
E18 Sprague–Dawley rat at postnatal day 21
was stereotaxically injected with lentivirus
+ – containing EGFP at the coordinates of
AP = +2.2; ML = +0.9; DV = +2.0, +1.5,
+1.0 from the bregma. The crosses
indicate injected sites. Cx, cerebral
cortex; Hip, hippocampus; Th, thalamus;
Str, striatum; Amy, amygdala; Hypo,
hypothalamus; VTA, ventral tegmental
area; P, pontine tegmentum. (Adapted
from [47], with permission.) See plate
section for color version.

(b) Adult brain


medial lateral
0 1 2 3 4 mm

Cx
Cx Hip

Str Th
Amy

AP +2.2 mm AP –1.34 mm
anterior bregma posterior
3 2 1 0 –1 –2 –3 –4 –5 –6 –7 –8 mm
0 dorsal
Cx 1
Hip 2
Str 3
Th
VTA 4
Hypo 5
6 ventral

by introducing expression or RNAi plasmid(s) in the procedures. Second, sterotaxic injection of virus-
developing brain [48]. Expression of more than one mediated expression and/or RNAi constructs in the
gene can be simultaneously modulated with this tech- brain is another approach to modulate expression of
nique. This feature is particularly advantageous in target molecules in a temporally and spatially specific
studying genetic roles for mental disorders, in which manner [47, 50].
a combination of multiple genes plays a role in the
etiology. Temporal expression control again becomes
possible by using an inducible system [49]. Regional Readouts relevant to mental disorders
specific expression can be controlled by changing the Functional outcomes in the brain elicited by genetic
orientation and position of the electrodes used in the susceptibility factors for major mental illnesses are

74
Chapter 7: Validation of candidate genetic susceptibility factors

Table 7.2 Behavioral assays relevant for different assayed in several ways. As previously published,
neuropsychological domains, which are used for testing
of mouse models of psychiatric diseases.
both histological/anatomical and behavioral assays
are utilized (Table 7.2). When genetic factor(s)
Putative Characteristic Typical test are modulated, it is important to address their
domain causal links to the observed phenotypes by both
Cognitive Working memory Delayed nonmatch cell autonomous and noncell autonomous/circuitry
to place mechanisms.
Behavioral Reversal learning Some of the most promising genetic factors for
flexibility major mental illnesses most likely play roles during
Learning and Water maze neurodevelopment. To address how the effects of
memory genetic factors in early development result in adult
Attention Latent inhibition phenotypes, we may need to consider postnatal
Cognitive/ Sensorimotor Prepulse inhibition brain maturation [7]. Especially in adolescence, sev-
positive gating eral key characteristics in the brain are dynamically
Positive Novelty induced Open field changed (Figure 7.3). Thus, when we consider read-
hyperactivity outs, their age-dependent progression should be
Hypersensitivity to Methamphetamine examined.
psychostimulants induced hyperactivity Molecular profiling studies with human cells
in the open field and tissues have identified genes and proteins that
Negative Social interaction Three chamber social are differentially expressed between normal controls
interaction test and patients. It is important to examine whether
Anhedonia Sucrose consumption such molecules are also differentially expressed
when genetic factors are modulated in animals. Such
Depressive Behavioral despair Forced swim test
study design opens a window for translational
Anxiety Anxiety Elevated plus maze research.

Figure 7.3 Dynamic changes in the adolescent brain. Disturbances generated by susceptibility genes and environmental insults during early
development (three stars on the left-hand side) may impair some of the crucial processes in early development, including progenitor cell
proliferation, neuronal migration and dendritic arborization and outgrowth. Independent of such initial risks/insults, intrinsic disease-associated
factors might also directly affect postnatal brain maturation (two central stars) contributing to the emergence of schizophrenia (SZ) in
young adulthood. (Adapted from [7], with permission.) See plate section for color version.

75
Chapter 7: Validation of candidate genetic susceptibility factors

(a) Human Mouse

Control SZ 5

Ventricle volume (mm3)


4

0
WT Tg

100
WT
(b)
WT Tg Tg
80

Cells/section
Control SZ
nCI/g
2 080
60
1 200 PV
860
620
390 40
210
90
0 1 mm 1 mm
0 PV mRNA 20
CB
0
PV CB
Figure 7.4 Anatomical and histological abnormalities in schizophrenia patients and in mouse models. (a) Enlarged lateral ventricles as
detected by in vivo magnetic resonance imaging (MRI) in schizophrenia patients (left) and in a DISC1 mouse model (right). (b) Decreased
parvalbumin (PV) expression in the prefrontal cortex of schizophrenia patients (mRNA, left) and a DISC1 mouse model (right).
(Adapted from [60], with permission.) See plate section for color version.

Animal models in addressing roles of DISC1 seem to result in behavioral deficits, which
include endophenotypes associated with both schizo-
of genetic factors for mental illnesses: phrenia and mood disorders [55]. In SZ, several histo-
DISC1 studies as examples logical and anatomical changes are reported [56].
Mice with eight distinct genetic manipulations on These include enlargement of lateral ventricles [57]
DISC1 have been published. Gogos and colleagues and interneuron deficits, which is easily detected by a
[51] published a model in which a spontaneous muta- decrease in the immunoreactivity of parvalbumin [58].
tion in the DISC1 gene is utilized in combination with By using in vivo or ex vivo MRI scans, enlarged lateral
genetic manipulation, achieving knockdown of some ventricles or shrinkage of some brain regions, consist-
types of DISC1 isoform. Clapcote and colleagues [52] ent with reports from patients with SZ, were observed
selected two lines with point mutations in the DISC1 in several DISC1 mice [52, 53, 59, 60] (Figure 7.4a).
gene from ENU mutagen-treated animals. Four Moreover, immunohistochemistry revealed that there
groups have generated transgenic mice expressing was a selective decrease in the immunoreactivity of par-
dominant-negative mutants of DISC1, two of which valbumin in different types of DISC1 mice [52, 59, 61]
are under an inducible expression system [53, 54]. (Figure 7.4b).
Most recently, a model generated by in utero gene The DISC1 model generated by in utero gene
transfer in which transient knockdown of DISC1 in transfer is a good example to present the importance
the pre- and perinatal stages, specifically in a lineage of assays in different developmental ages [54]. In this
of pyramidal neurons mainly in the prefrontal cortex, model, although dendritic pathology exists at post-
was reported [46]. natal day 14 (P14) due to transient knockdown of
In these animals, many behavioral assays have been DISC1 in early development, no robust changes in
conducted. As far as reported, genetic manipulations neurochemistry and behavior are observed before

76
Chapter 7: Validation of candidate genetic susceptibility factors

puberty (at P28). Interestingly, selective disturbance relationships. To identify mechanistic links from
of dopaminergic neurotransmission and associated genetic factors to the phenotypes, especially those
behavioral alterations become prominent after puberty observed during brain development and maturation,
(at P56 and later). a combination of human studies with animal
experiments is expected.
Human studies with brain imaging
A series of studies by Weinberger and associates has Concluding remarks
pioneered the possible correlation of brain dysfunc- In this chapter, we describe how functions of genetic
tion with genetic variations in susceptibility factors susceptibility factors can be validated, specifically
associated with mental illnesses [62, 63]. This approach using DISC1 as an example. Studies at multiple levels,
is very important in addressing genetic effects on from protein chemistry, cell biology, animal study, to
human functions directly. Nonetheless, these studies clinical work will provide comprehensive understand-
are limited to descriptively presenting correlative ing of the functions of susceptibility factors.

References 16. Camargo LM, et al. Mol Psychiatry


2007;12(1):74–86.
33. Hayashi-Takagi A, et al. Nat
Neurosci 2010;13(3):327–332.
1. Allen NC, et al. Nat Genet 2008;
40(7):827–834. 17. Gurling HM, et al. Arch Gen 34. Seshadri S, et al. Proc Natl
Psychiatry 2006;63(8):844–854. Acad Sci U S A 2010;107(2):
2. Sun J, et al. Am J Med Genet 5622–5627.
B Neuropsychiatr Genet 2008; 18. Burdick KE, et al. Hum Mol Genet
2008;17(16):2462–2473. 35. Lewis DA, et al. Ann Rev Neurosci
147B(7):1173–1181.
2002;25:409–432.
3. Cook EH, Jr., et al. Nature 2008; 19. Tomppo L, et al. Biol Psychiatry
2009;65(12):1055–1062. 36. Rapoport JL, et al. Mol Psychiatry
455(7215):919–923.
2005;10(5):434–449.
4. Sebat J, et al. Trends Genet 2009; 20. Millar JK, et al. Science 2005;
310(5751):1187–1191. 37. Mei L, et al. Nat Rev Neurosci
25(12):528–535. 2008;9(6):437–452.
5. Cichon S, et al. Am J Psychiatry 21. Harrison PJ, et al. Mol Psychiatry
38. Ishizuka K, et al. Biol Psychiatry
2009;166(5):540–556. 2005;10(1):40–68.
2006;59(12):1189–1197.
6. The Huntington Disease 22. Tajinda K, et al. Mol Psychiatry 39. White T, et al. Am J Psychiatry
Collaborative Research Group. 2010;15(3):231–232. 2006;163(3):376–380.
Cell 1993;72(6):971–983. 23. Takahashi K, et al. Cell 2006; 40. Arguello PA, et al. Neuron
7. Jaaro-Peled H, et al. Trends 126(4):663–676. 2006;52(1):179–196.
Neurosci 2009;32(9):485–495. 24. Vierbuchen T, et al. Nature 41. Sanftner LM, et al. Exp Neurol
8. Manolio TA, et al. Nature 2010;463(7284):1035–1041. 2005;194(2):476–483.
2009;461(7265):747–753. 25. Holmes E, et al. PLoS Med 2006; 42. Wood JD, et al. Hum Mol Genet
9. Chubb JE, et al. Mol Psychiatry 3(8):e327. 2009;18(3):391–404.
2008;13(1):36–64. 26. Do KQ, et al. Curr Opin Neurobiol 43. Sawamura N, et al. Mol Psychiatry
10. Millar JK, et al. Hum Mol Genet 2009;19(2):220–230. 2008;13(12):1138–1148, 1069.
2000;9(9):1415–1423. 27. Kamiya A, et al. Nat Cell Biol 44. Bockamp E, et al. Physiol
11. Fields S, et al. Nature 1989; 2005;7(12):1167–1178. Genomics 2002;11(3):115–132.
340(6230):245–246. 28. Kamiya A, et al. Hum Mol Genet 45. Kos CH. Nutr Rev 2004;62
12. Ozeki Y, et al. Proc Nat Acad Sci 2006;15(22):3313–3323. (6 Pt 1):243–246.
U S A 2003;100(1):289–294. 29. Miyoshi K, et al. Mol Psychiatry 46. Niwa M, et al. Neuron 2010;
13. Morris JA, et al. Hum Mol Genet 2003;8(7):685–694. 65(4):480–489.
2003;12(13):1591–1608. 30. Murdoch H, et al. J Neurosci 47. Seshadri AJ, et al. Prog Brain Res
14. Miyoshi K, et al. Biochem Biophys 2007;27(35):9513–9524. 2009;179:17–27
Res Commun 2004;317(4): 31. Mao Y, et al. Cell 2009;136(6): 48. LoTurco J, et al. Cereb Cortex
1195–1199. 1017–1031. 2009;19 Suppl 1:i120–125.
15. Millar JK, et al. Biochem Biophys 32. Glantz LA, et al. Arch Gen 49. Matsuda T, et al. Proc Natl
Res Commun 2003;311(4): Psychiatry 2000;57(1): Acad Sci U S A 2007;104(3):
1019–1025. 65–73. 1027–1032.

77
Chapter 7: Validation of candidate genetic susceptibility factors

50. Cetin A, et al. Nat Protoc 2006; 55. Jaaro-Peled H. Prog Brain Res 60. Jaaro-Peled, et al. Schizophr Bull
1(6):3166–3173. 2009;179:75–86. 2009;35:865–873.
51. Koike H, et al. Proc Natl Acad Sci 56. Harrison PJ. Brain 1999;122 61. Ayhan Y, et al. Mol Psychiatry
U S A 2006;103(10):3693–3697. (Pt 4):593–624. 2011;16(3):293–306.
52. Clapcote SJ, et al. Neuron 2007; 57. Shenton ME, et al. Schizophr Res 62. Meyer-Lindenberg A, et al.
54(3):387–402. 2001;49(1–2):1–52. Nat Rev Neurosci 2006;7(10):
53. Hikida T, et al. Proc Natl Acad Sci 58. Lewis DA, et al. Am J Psychiatry 818–827.
U S A 2007;104(36):14501–14506. 2001;158(9):1411–1422. 63. McGuire P, et al. Trends
54. Pletnikov MV, et al. Mol Psychiatry 59. Shen S, et al. J Neurosci 2008; Pharmacol Sci 2008;29(2):
2008;13(2):173–186, 115. 28(43):10893–10904. 91–98.

78
Epigenetic mechanisms in drug
Chapter

8 addiction and depression


William Renthal and Eric J. Nestler

Introduction “depressed” transcriptional program in neurons,


much in the same way environmental cues differen-
Epigenetics is classically defined as the interaction tiate a stem cell into specific lineages. While this field
between genes and the environment that gives rise is still in its infancy, great progress is being made in
to a specific phenotype. An example of this process identifying epigenetic alterations in drug addiction
is observed in cellular differentiation where chemical and depression, as well as in several other neuropsy-
signals induce totipotent stem cells to differentiate chiatric syndromes such as schizophrenia, Alzheimer’s
into genetically identical cell types with vastly differ- disease, and Rett syndrome. Focusing on drug addic-
ent functions. This is easily apparent when comparing tion and depression, this chapter will briefly review
the brain, for example, which is extremely sensitive the molecular machinery underlying epigenetic mech-
to tissue damage and unable to regenerate, to other anisms and discuss how their dysregulation may con-
organs (e.g. liver) that can regenerate rapidly. This tribute to these chronic psychiatric disorders.
is due in part to the vastly different sets of genes
expressed between neurons and hepatocytes, despite
their identical DNA templates. Mechanistic insight Epigenetic mechanisms
into this process has been gained over the past Chromatin is the complex of DNA, histones, and
20 years and involves the transduction of unique associated nonhistone proteins in the cell nucleus.
environmental signals into precise and highly stable DNA wraps around histone octamers made up of
alterations in chromatin structure that ultimately gate two copies of histone H2A, H2B, H3, and H4 [2],
access of transcriptional machinery to specific gene which then undergo a complex supercoiling process
programs, thereby providing unique gene expression to form a highly condensed structure (Figure 8.1) [3].
profiles in response to specific environmental cues Initially, it was thought that this elaborate chromatin
[1]. Importantly, many of these chromatin remodel- structure only functioned to condense meters of DNA
ing mechanisms are highly stable, which contributes into the microscopic cell nucleus, but it is now known
to the maintenance of specific gene expression pro- to participate in the transcriptional status of nearly
grams in the correct tissues throughout the life of an every eukaryotic gene. Because DNA is tightly associ-
individual. ated with histones and often embedded deep within
The strong control epigenetic mechanisms exert chromatin supercoils that are structurally inaccessible
on gene expression and their potential stability over to transcriptional activators [4, 5], cellular mechan-
time suggests a potential role in mediating aspects of isms exist to modify and remodel chromatin structure
the long-lasting neural plasticity that ultimately result to allow for the coordinated expression of specific
in psychiatric syndromes such as drug addiction or transcriptional programs and the silencing of others
depression, as well as their reversal during effective [6]. Such modifications typically occur on N-terminal
treatment. Thus, epigenetic research in psychiatry histone tails and include acetylation, phosphorylation,
attempts to identify whether environmental stimuli and methylation of histones, methylation of DNA, and
(e.g. drugs of abuse, stress) induce certain changes many others, with each modification either directly
in chromatin structure that mediate an “addicted” or altering histone/DNA interactions or serving as a mark

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

79
Chapter 8: Epigenetic mechanisms in drug addiction and depression

(a) Histone tail Figure 8.1 Chromatin remodeling. (a) Picture of a


nucleosome showing a DNA strand wrapped around
a histone octamer composed of two copies each of
the histones H2A, H2B, H3, and H4. The amino (N) termini
of the histones face outward from the nucleosome
DNA
complex. (b) Chromatin can be conceptualized as
H2B existing in two primary structural states: as active, or
H4
open, euchromatin (top left) in which histone acetylation
(A) is associated with opening the nucleosome to allow
H3
binding of the basal transcriptional complex and other
H2A activators of transcription; or as inactive, or condensed,
heterochromatin where all gene activity is permanently
silenced (bottom left). In reality, chromatin exists in a
A Acetylation continuum of several functional states (active; permissive
Histone [top right]; repressed [bottom right]; and inactive).
M Methylation Enrichment of histone modifications such as acetylation
(b) Active P Phosphorylation and methylation (M) at histone N-terminal tails and
Histones Histone Transcription factor
related binding of transcription factors and co-activators
tail +
Permissive
(Co-Act) or repressors (Rep) to chromatin modulates the
A A A M M transcriptional state of the nucleosome. Recent evidence
P A A suggests that inactivated chromatin may in some cases
be subject to reactivation in adult nerve cells, although
this remains uncertain. (c) Summary of common
covalent modifications of H3, which include acetylation,
A Co-Act A
methylation and phosphorylation (P) at several amino
A M A
M acid residues. H3 phosphoacetylation commonly
DNA Basal transcription
complex
involves phosphorylation of S10 and acetylation of K14.
Acetylation is catalysed by histone acetyltransferases
Inactive (HATs) and reversed by histone deacetylases (HDACs);
M M
P P lysine methylation (which can be either activating or
repressing) is catalysed by histone methyltransferases
Repressed
Rep Rep (HMTs) and reversed by histone demethylases (HDMs);
Rep
M M M
M and phosphorylation is catalysed by protein kinases (PK)
Rep M
A M M and reversed by protein phosphatases (PP), which have
Rep
A M not yet been identified with certainty. K, lysine residue;
M
M
? S, serine residue. (From [3].) See plate section for
color version.
Rep M
M M
M
M M

(c)
Demethylation Demethylation

HDM HMT HDM HMT

Methylation Methylation
(activating) (repressing)

M M
M M
H3 K4 K27 S28 M
K9 K23 K36
S10 K14 K18 K79
P
A A
P A A
Histone
tail
Acetylation Phosphorylation
(activating) (activating)

HAT HDAC PK PP

Deacetylation Dephosphorylation

80
Chapter 8: Epigenetic mechanisms in drug addiction and depression

that recruits specific transcriptional co-regulators to intrinsic HAT activity, suggesting that part of their
positively or negatively regulate the underlying gene’s function in regulating emotional behaviors could be
activity. Ultimately, dozens of potential modifications a result of histone acetylation [13, 14].
that occur at many distinct histone residues summate Histone deacetylases (HDACs), which remove
to determine the final transcriptional output of a acetyl groups from histones, are divided into four
given gene [7]. An especially important aspect of classes. Class I HDACs (e.g. HDAC1, 2, 3, and 8)
certain chromatin modifications is their apparent are ubiquitously expressed and likely mediate the
stability, as is seen with maternal imprinting in majority of deacetylase activity within cells. Recent
Prader–Willi syndrome or X-inactivation, where evidence points to HDAC2 in the hippocampus as a
DNA methylation contributes to life-long gene silen- key regulator of learning and memory [15]. Class II
cing [3]. However, despite the apparent stability of HDACs (e.g. HDAC4, 5, 9) are only expressed in
epigenetic mechanisms in vivo, all types of chromatin specific tissues such as heart and brain and are much
modifications identified to date are potentially revers- larger enzymes that also contain an N-terminal regu-
ible and have specific enzymes or processes which latory domain that enables them to be shuttled in
mediate the addition or removal of each mark [6]. and out of the nucleus in a neural activity-dependent
The reasons for this apparent biochemical and manner [16]. While Class II HDACs can deacetylate
physiological disconnect are currently not clear. histones, they are far less efficient enzymes than Class
I HDACs [17]. There is currently one Class IV
HDAC, HDAC11, and it has characteristics of both
Histone acetylation Class I and Class II enzymes [18]. Class III HDACs
Acetylation of histone lysine residues reduces the (Sirtuins) are mechanistically distinct from the other
electrostatic interaction between histone proteins and HDACs and have been implicated in the regulation of
DNA, which is thought to relax chromatin structure life span and metabolism [19]. While Classes I–III are
and make DNA more accessible to transcriptional evolutionarily conserved in yeast, there are far more
regulators (Figure 8.1) [6]. Genome-wide studies indi- members in each class in mammals, with each indi-
cate that the presence of high levels of histone acetyla- vidual HDAC performing a unique function. Perhaps
tion in gene promoter regions is almost invariably the best studied example of this is HDAC6, which is a
associated with transcriptional activation while low Class II HDAC that deacetylates tubulin and is not
levels of acetylation correlate with less gene activity even present in the nucleus. The individual functions
[8, 9]. Most genome-wide studies of histone acetyla- of other HDACs remain an active topic of investigation.
tion have focused on acetylation of histones H3
and H4, but histone acetylation can occur on other
histone proteins as well. Histone phosphorylation
Histone acetylation is a dynamic process, con- Histone phosphorylation is generally associated with
trolled by specific enzymes which either add or transcriptional activation; it can be observed on the
remove the acetyl mark. Histone acetyltransferases promoters of immediate early genes such as c-fos
(HATs) catalyze the addition of acetyl groups onto when they are induced after cyclic adenosine mono-
lysine residues of histone proteins. There are over a phosphate (cAMP) stimulation or glutamate treat-
dozen known HATs that have varying degrees of ment in cultured striatal neurons [20, 21]. One of
specificity for different lysine residues and histone the best characterized histone phosphorylation sites
proteins and, importantly, many of them can also is serine 10 on histone H3 (H3S10). This modification
acetylate nonhistone proteins such as transcription stabilizes the HAT, GCN5, on gene promoters while
factors (e.g. p53). Certain HATs, such as CREB-binding antagonizing the repressive modification – methylation
protein (CBP) and p300/CBP-associated factor of lysine 9 on histone H3 (H3K9) and its subsequent
(PCAF), have been implicated in behavioral responses recruitment of HP1 (heterochromatin protein 1) (see
to drugs of abuse, stress, and learning and memory below) [6]. Since phosphorylation at H3S10 recruits a
[10–12]. Recently, several transcription factors HAT, the neighboring lysine residue at H3K9 is often
already implicated in behavioral responses to emo- acetylated in concert with phosphorylation, a process
tional stimuli (e.g. ATF2 [activating transcription called phospho-acetylation that further potentiates
factor 2], CLOCK) have also been shown to possess gene activation.

81
Chapter 8: Epigenetic mechanisms in drug addiction and depression

There are several nuclear protein kinases and Similarly, the HDM, KDM3A (JHDM2a), can
protein phosphatases known to regulate histone demethylate one or two methyl groups on H3K9,
phosphorylation [6]. The mitogen-activated protein requiring a distinct demethylase (e.g. KDM4D
kinase, MSK1, and the dopamine and cAMP regu- [JMJD2D]) to fully demethylate the trimethylated state.
lated protein phosphatase inhibitor, DARRP-32, are Thus, many enzymes, which are often found in large
elegant examples shown to regulate H3S10 phosphor- complexes with other transcriptional co-regulators, are
ylation in the adult brain in response to cocaine expos- required to move between the unmethylated and fully
ure [22, 23]. Furthermore, genetic disruption of the trimethylated states. Proper balance of histone methyl-
histone modifying ability of MSK1 or DARRP-32 ation has already been strongly implicated in normal
in vivo has dramatic effects on behavioral responses brain function, as the HDM, KMT5C (SMCX), con-
to cocaine. Thus, histone phosphorylation likely plays trols dendritic spine density and is mutated in patients
an important role in the regulation of brain function. with intellectual disability [25, 26].

Histone methylation DNA methylation


Histone methylation is thought to function primarily DNA methylation refers to the enzymatic methylation
by generating docking sites that recruit transcrip- of cytosine bases; this is, a fundamental cellular pro-
tional co-regulators to specific gene loci. Histone cess required for development, tissue-specific gene
methylation occurs on lysine residues and can exist expression, X-inactivation, and genetic imprinting,
in mono-, di-, or tri-methylated states, enabling each to name a few examples [27]. DNA methylation is
state to recruit unique co-regulators and exert distinct thought to repress gene expression by interfering with
effects on transcriptional activity [6]. Additionally, transcription factors binding to their target sequences
methylation of different histone lysine residues can or by initiating the recruitment of co-repressors.
result in distinct or even opposite effects on transcrip- For example, the cAMP-response element (CRE) con-
tion. For example, tri-methylation of H3K4 is highly tains a cytosine-guanine dinucleotide in the middle
associated with gene activation, while tri-methylation of its consensus sequence, which, when methylated,
of H3K9 or H3K27 are usually associated with gene prevents the transcription factor CRE-binding protein
repression [5]. The repression caused by tri-methyla- (CREB) from binding [28]. Thus, for genes at which
tion of H3K9 is mediated in part via the recruitment CREB is necessary to initiate transcription, methyla-
of co-repressors, such as HP1, as stated earlier. How- tion at this site is repressive. Methylated DNA can
ever, even this is an oversimplification, as methylated also recruit methyl-binding domain-containing pro-
H3K9 is often found in the coding region down- teins, such as MeCP2, which can then recruit and
stream of a gene promoter and may be involved in stabilize transcriptional co-repressors such as HDACs
transcriptional elongation [6, 24]. Thus, through a on specific gene promoters. Mutations in MeCP2
vast array of combinatorial options, histone methyla- cause the autistic spectrum disorder, Rett syndrome,
tion provides each cell with exquisite control over an illustrating the importance of DNA methylation in
individual gene’s activity. normal brain development [29]. While there is a
Like histone acetylation and phosphorylation, the strong correlation between methylated DNA and
enzymes which regulate histone methylation can repressed gene activity, recent studies of MeCP2 indi-
be divided into two main groups: those which add cate it may also serve to activate gene activity under
the mark, histone methyltransferases (HMTs), and some circumstances [30], suggesting that the context
enzymes which remove it, histone demethylases in which DNA methylation occurs is an important
(HDMs) (Figure 8.1). HMTs and HDMs not only factor in its ultimate effect on transcription.
discriminate between specific histone lysine residues, There are three known enzymes which catalyze
but each enzyme is also unique in its ability to cata- DNA cytosine methylation: DNMT1, DNMT3a, and
lyze mono-, di-, or tri-methylation or demethylation DNMT3b. DMNT2 was recently shown to methylate
at that site [6]. For example, the HMT, KMT1C RNA rather than DNA [31]. Together, these enzymes
(G9a), is specific for histone H3K9 but only adds establish and maintain the unique methylation
1 or 2 methyl groups, with the distinct HMT, KMT1A patterns that exist within each cell type. While the
(SUV39H1), catalyzing trimethylation of this site. regulation of these enzymes in brain remains unclear,

82
Chapter 8: Epigenetic mechanisms in drug addiction and depression

pharmacological inhibition of DNA methylation in past decade, as the knowledge of molecular/epigenetic


the brain in vivo results in rapid demethylation of mechanisms by which individual genes are activated
specific gene targets and severe deficits in learning or silenced in cell culture has exploded, the question
and memory [32]. The mechanism by which this of how drugs of abuse can mediate long-lasting
occurs, however, remains unclear because, unlike changes in the activity of specific genes is now a major
other chromatin modifications, the existence of DNA focus in addiction research. While these epigenetic
demethylases remains highly controversial [33]. Never- studies are ongoing for a variety of substance abuse
theless, regulation of DNA methylation by environ- models, we focus here on the psychostimulants
mental stimuli remains an attractive mediator of cocaine and amphetamine because these studies are
long-lasting changes in transcription in adult neurons. more mature.
Both cocaine and amphetamine are frequently
Epigenetic mechanisms in drug studied in rodents using forced investigator adminis-
tration of the drug. The molecular changes in brain
addiction can then be studied after a single injection or repeated
Drug addiction is a chronic relapsing disorder injections to provide insight into changes that occur
whereby behavior related to seeking and taking drugs initially in response to the drug versus changes which
of abuse becomes compulsive and pathological [34]. require chronic exposure. While this model of forced
The process by which repeated drug experimentation drug administration conveniently analyzes the acute
transitions into a chronically addicted state is the and chronic pharmacological effects of drugs, it only
focus of intense research, as clues into these mechan- permits an elementary assessment of addictive-like
isms may help better manage or perhaps fully treat behaviors. Thus, rodent self-administration para-
addicted patients. Currently, treatment of drug addic- digms, in which animals are trained to push a lever
tion is a major clinical challenge because of high rates to obtain drug volitionally, while much lower in
of relapse, which are observed even after long periods throughput, better model human addiction. Both
of drug abstinence. Thus, another major research goal forced administration and self-administration models
is to identify the mechanisms underlying drug relapse have been used to study epigenetic changes in
and target them with novel therapeutics that improve response to psychostimulants.
treatment outcomes. The first studies to identify changes in chromatin
One such mechanism involves long-lasting alter- structure in response to drugs of abuse focused on
ations in gene expression in brain reward regions that histone acetylation on the promoters of genes previ-
contribute to both the pathogenesis and persistence of ously implicated in cocaine action (e.g. c-fos, fosB,
drug addiction. For example, the transcription factor, cdk5) [37]. These studies found that acute adminis-
ΔFosB, a splice product of the immediate early gene tration of cocaine significantly increased histone H4
fosB, accumulates several-fold uniquely in brain acetylation on the immediate early genes c-fos and
reward regions after repeated drug exposure [35]. fosB in striatum with a rapid time course consistent
This establishes a positive-feedback loop whereby with the induction time of their mRNAs (Figure 8.2).
higher levels of ΔFosB facilitate greater drug-seeking The histone acetyltransferase CBP appears to be
behavior, which in turn induce higher levels of required for the drug-induced acetylation of the fosB
ΔFosB. Thus, ΔFosB is one molecule that may con- promoter [10]. Interestingly, an acute cocaine dose
tribute to the transition to an addicted state. Activator increases total levels of histone H4 acetylation, and
of G-protein signaling 3 (AGS3) is an example of a histone H3 phospho-acetylation in striatum, while
protein that may contribute to drug relapse, since it cocaine does not induce histone acetylation at several
remains elevated after weeks of drug withdrawal and, control gene promoters [22, 37]. These global increases
when genetically manipulated in brain, alters relapse in histone acetylation, which are also observed in
behavior [36]. ΔFosB and AGS3 not only exemplify response to environmental enrichment and tests of
how drug-induced gene expression changes in brain learning and memory [38, 39], may be accounted for
can contribute to clinically important aspects of the by high levels of acetylation on specific subsets of genes.
addicted state, they also illustrate the potential for The promoters of certain genes induced by chronic
new therapeutic targets that could prevent or reverse cocaine exposure are hyperacetylated for days to weeks
the stable effects of chronic drug exposure. Over the after the last drug exposure (Figure 8.2). For example,

83
Chapter 8: Epigenetic mechanisms in drug addiction and depression

Figure 8.2 Regulation of chromatin structure by drugs of abuse. Drug-induced signaling events are depicted for cocaine and amphetamine.
Cocaine and amphetamine can increase cyclic adenosine monophosphate (cAMP) levels in striatum, which activates protein kinase A
(PKA) and leads to phosphorylation of its targets. This includes the cAMP response element binding protein (CREB), the phosphorylation of
which induces its association with the histone acetyltransferase, CREB binding protein (CBP) to acetylate histones and facilitate gene activation.
This is known to occur on many genes including fosB and c-fos in response to psychostimulant exposure. ΔFosB is also upregulated by chronic
psychostimulant treatments, and is known to activate certain genes (e.g. cdk5) and represses others (e.g. c-fos) where it recruits HDAC1.
This repression of c-fos also involves increased repressive histone methylation, which is thought to occur via the induction of specific histone
methyltransferases. It is not yet known how cocaine regulates histone demethylases (HDM) or DNA methyltransferases (DNMTs). Cocaine
also activates the mitogen activated protein kinase (MAPK) cascade, which through MSK1 can phosphorylate CREB and histone H3 at serine 10.
Cocaine promotes H3 phosphorylation via a distinct pathway, whereby PKA activates protein phosphatase 2A, leading to the
dephosphorylation of serine 97 of DARPP32. This causes DARPP32 to accumulate in the nucleus and inhibit protein phosphatase 1 (PP1) which
normally dephosphorylates H3. Chronic exposure to psychostimulants is also known to increase glutamatergic signaling from the prefrontal
cortex to the NAc. Glutamatergic signaling elevates Ca2+ levels in NAc synapses and activates CaMK (calcium/calmodulin protein kinases)
signaling, which, in addition to phosphorylating CREB, also phosphorylates HDAC5. This results in nuclear export of HDAC5 and increased
histone acetylation on its target genes (e.g. NK1R [NK1 or substance P receptor). (From [3].) See plate section for color version.

bdnf (brain-derived neurotrophic factor) and npy potentiates the locomotor-activating and rewarding
(neuropeptide Y) expression were found to be upre- responses to cocaine and amphetamine [37, 41, 42].
gulated after cocaine self-administration and their Conversely, reducing histone acetylation by overex-
gene promoters were hyperacetylated, while egr-1 pressing certain HDACs, or knockdown of the HAT,
(early growth response 1) was found to be downregu- CBP, results in less sensitivity to cocaine [10, 36, 41].
lated and hypoacetylated after cocaine withdrawal Cocaine alters histone acetylation through many
[40]. These findings suggest a role of histone acetyl- enzymes in the NAc, but one particular HDAC,
ation in the maintenance of gene expression involved HDAC5, responds uniquely to chronic cocaine
in drug withdrawal and relapse. administration, raising the interesting possibility that
Cocaine-induced alterations in chromatin struc- this HDAC is involved in the behavioral transitions
ture in the nucleus accumbens (NAc), the ventral which occur between acute and chronic cocaine
portion of striatum heavily implicated as a brain exposure (e.g. drug experimentation to compulsive
reward region, have been shown to regulate behav- drug use). Chronic cocaine administration increases
ioral responses to drugs of abuse. Pharmacological the phosphorylation of HDAC5 and shuttles it out of
inhibition of HDACs in the NAc, which increases the nucleus, permitting hyperacetylation of HDAC5
histone acetylation in this brain region, significantly target genes (Figure 8.2) [41]. This phosphorylation

84
Chapter 8: Epigenetic mechanisms in drug addiction and depression

reaction may be mediated via Ca2+/calmodulin- (SNRIs) are the most common. Unfortunately, less
dependent protein kinase II (CaMKII). Consistent than 50% of patients exhibit a complete response to
with its regulation by cocaine, mice deficient for SSRIs, SNRIs, and related antidepressants, thus leav-
HDAC5 display normal rewarding responses to initial ing a substantial portion of depressed patients with a
cocaine exposure, but become hypersensitive when treatment-resistant depression, which may become a
treated with a chronic course of cocaine [41]. Thus, chronic condition [see 44]. Psychiatric research is
pharmacological and genetic manipulations that thus focused on identifying new mechanisms that
increase histone acetylation appear to potentiate behav- are involved in the pathogenesis and maintenance of
ioral responses to cocaine and suggest that altered depression, which may serve as new targets for more
histone acetylation may contribute to establishment effective therapeutics.
of an addicted state. One of the most challenging obstacles for depres-
Histone H3 phosphorylation and phospho-acety- sion research has been the development of an animal
lation also appear to play key roles in drug-regulated model that accurately recapitulates the symptoms of
behaviors. Global levels of histone H3 phosphoryl- human depression. While no model can effectively
ation at serine 10 are induced by acute cocaine in model all aspects of human depression (e.g. suicide),
striatum, a process which requires MSK1 [22, 37]. some of the major symptoms such as anhedonia and
The function of MSK1 is behaviorally important, as sleep and weight disturbances, and their reversal by
mice lacking this kinase have attenuated locomotor antidepressant treatment, can be studied in rodents.
responses to cocaine. The genes at which histone The pathogenesis of depressed-like states is typically
phosphorylation is occurring in response to cocaine modeled in rodents by chronic exposure to stress [44].
remain mostly undefined, with the exception of c-fos, One such model, chronic social defeat stress, involves
where histone phosphorylation occurs in conjunction the repeated exposure of an experimental mouse to a
with acetylation (phospho-acetylation). series of aggressive mice over 10 days. Each day the
Preliminary findings suggest that histone methyl- stress begins as a brief physical encounter (typically
ation is also regulated by cocaine and, in turn, alters 5–10 minutes) followed by a full day of sensory con-
behavioral responses to the drug. For example, inhib- tact (e.g. smell, sight) as the mice are separated by a
ition of a particular H3K9 histone methyltransferase, screen. After 10 days of social defeat, the experimental
KMT1C (G9a), whose expression is regulated in the mice develop a chronic syndrome (lasting more than
NAc by chronic cocaine administration, potentiates a month) that is characterized by anhedonia, anxiety-
behavioral responses to the drug [43]. These findings like symptoms, weight loss, and loss of interest in
are consistent with histone acetylation findings, social interaction. Importantly, SSRIs or SNRIs
since inhibition of H3K9 methylation would also be reverse most of these behavioral end points, making
expected to enhance gene activity. Together, these data chronic social defeat stress an attractive model in
suggest that, in general, increases in gene expression which to study the molecular adaptations associated
potentiate behavioral sensitivity to drugs of abuse. with a depressed-like state and those involved with
Overall, these findings implicate changes in his- antidepressant action [45, 46].
tone acetylation, phosphorylation, and methylation in BDNF plays a critical role in the development of
mediating expression changes in specific sets of genes the social defeat phenotype and its reversal by anti-
that are crucial for controlling behavioral responses depressant treatment. It was observed that BDNF in
to drugs of abuse. the hippocampus is downregulated for at least one
month after chronic social defeat stress, and that
chronic antidepressant treatment reversed this down-
Epigenetic mechanisms in depression regulation [46]. A mechanism for this long-lasting
Depression is a chronic disorder characterized by regulation of gene expression was identified as meth-
many debilitating symptoms including dysphoria, ylation of H3K27, a repressive histone modification,
anhedonia, sleep disturbances, and weight changes. that remains hypermethylated on the bdnf promoter
Most people diagnosed with depression are prescribed within hippocampus for at least a month after defeat
some type of antidepressant treatment, of which stress (Figure 8.3 [3]). While chronic antidepressant
selective serotonin reuptake inhibitors (SSRIs) or treatment of mice exposed to chronic social defeat
mixed serotonin–norepinephrine reuptake inhibitors ameliorates many of the behavioral deficits and

85
Chapter 8: Epigenetic mechanisms in drug addiction and depression

(a) Figure 8.3 Regulation of the bdnf gene


HDAC5 HDAC5 by social defeat. (a) In the absence of
stress, the chromatin state of brain-
A A A derived neurotrophic factor (Bdnf) is at
a basal level, characterized by moderate
+ levels of histone H3 acetylation and
Bdnf expression
virtually no H3K27 dimethylation. In this
state, histone deacetylase 5 (HDAC5)
A A might repress unnecessary activation of
M
M BDNF and maintain a chromatin balance.
(b) Chronic defeat stress induces the
specific and prolonged dimethylation
(b) HDAC5 of histone H3K27. This induces a more
HDAC5
M M M M M “closed” chromatin state at bdnf promoters
A M M A M M A M P3 and P4, and a corresponding repression
of Bdnf transcripts III and IV expression.
H3 acetylation and HDAC5 regulation are

Bdnf expression not affected after chronic defeat stress
alone, corroborating the idea that the
main repressive marker after chronic stress
A A
M
is histone methylation. (c) Chronic
M M M M
M M M M M imipramine (antidepressant) treatment
after defeat stress downregulates Hdac5
(c) expression and increases H3 acetylation,
M M M M
M M M M with little if any change in H3K27
A A A A A dimethylation. Imipramine-dependent H3
hyperacetylation at the bdnf promoters
+ P3 and P4 allows partial “reopening” of
Bdnf expression
the repressed chromatin state caused by
defeat stress, and results in transcriptional
M
A A A A A M H3–K27 dimethylation reactivation of the bdnf gene. K, lysine
M M M M residue. (From [3].) See plate section
M M M M A H3 acetylation for color version.

restores Bdnf mRNA to normal levels, H3K27 remains tranylcypromine derive from its blockade of KMT1A
hypermethylated. The maintenance of H3K27 methyl- and the subsequent facilitation of H3K4 methylation.
ation even after chronic antidepressant treatment sug- Arguing against this interpretation is the knowledge
gests that BDNF expression might revert to a repressed that several structurally unrelated monoamine oxidase
state if drug administration were stopped. This novel inhibitors, which have not been shown to inhibit his-
epigenetic mechanism, which was proposed as a form tone demethylases, are still effective antidepressants.
of “molecular scar”, may describe a potential mech- The increase in H3 acetylation by antidepressant
anism by which the symptoms of depressed patients treatment suggested that HDAC inhibitors may also
reappear after cessation of antidepressant treatment; have antidepressant-like effects. Indeed, in both the
however, this remains speculative and further research chronic social defeat model and in the forced swim
is needed. test, HDAC inhibitors demonstrated antidepressant-
The recovery of Bdnf expression after antidepres- like prosperities [46, 48]. This was especially apparent
sant treatment is likely mediated by the antidepressant- when an HDAC inhibitor was administered in addi-
induced increase in histone H3K4 methylation and tion to the SSRI fluoxetine. While these inhibitors target
H3 poly-acetylation in hippocampus, which are associ- numerous HDACs, one specific isoform, HDAC5,
ated with gene activation [46]. Interestingly, tranylcy- stood out because it was oppositely regulated by stress
promine, which inhibits monoamine oxidases and is and antidepressant treatment [46]. Indeed, overexpres-
used as an antidepressant, is actually a much stronger sion of HDAC5 in the hippocampus blocks the behav-
inhibitor of the histone H3K4 demethylase KMT1A ioral effects of chronic antidepressant treatment,
(formerly, LSD1) than it is of either monamine oxidase suggesting that increased histone acetylation on the
A or B [47]. Thus, it would be interesting to determine bdnf promoter is a key mechanism to overcome the
whether any of the antidepressant properties of repressive effects of H3K27 methylation.

86
Chapter 8: Epigenetic mechanisms in drug addiction and depression

Another intriguing aspect of chronic social defeat remodeling pathways could be a novel approach to
stress is that the severity of the depressed-like pheno- new antidepressant development.
type varies within a cohort of inbred (i.e. nearly
genetically identical) mice. It was observed that more
susceptible mice in order to defeat stress show signifi-
Challenges of epigenetics research
cantly higher firing rates of dopaminergic neurons in in psychiatry
the ventral tegmental area (VTA). However, less sus- The study of gene expression and chromatin remodel-
ceptible mice had normal VTA firing rates because of ing in adult brain is made difficult by two major
an upregulation of potassium channels in this brain technical challenges. The first challenge is that brain
region [44]. Why do certain mice upregulate protect- tissue is highly heterogeneous and contains many
ive potassium channels in the VTA while others subtypes of neurons and glia, each of which is further
fail to do this and become “depressed?” Perhaps distinguished based on connectivity and function.
an epigenetic mechanism is involved in altering the Thus, the cellular and molecular responses to cocaine
promoters of certain potassium channels to ultimately can vary tremendously even within a small subregion
determine if the gene will be induced in response to of the NAc. In response to cocaine, for example,
chronic stress. If so, which life experiences trigger neurons which express Gs-coupled dopamine D1
these chromatin remodeling events? These are receptors in the NAc show higher levels of cAMP
important questions that may shed new light into an and also show its downstream consequences (such
extraordinarily complex syndrome and provide new as phosphorylation of CREB), while neurons express-
avenues for the development of more effective ing Gi-coupled D2 receptors respond oppositely.
antidepressants. There is even a small population of striatal neurons
Another important epigenetic mechanism that which express both D1 and D2 receptors, which seem
may contribute to long-lasting changes in neural to form hetero-oligomers that couple to Gq signaling
function and behavior is DNA methylation. Early [50], leading to more complex cellular responses to
insight into the role of DNA methylation in behavior cocaine. The end result of studying tissue lysates from
followed from studies of maternal care. Rats that such a heterogeneous structure is that most of the
receive poor maternal care as pups grow up to deliver observed effects are very small since they are averaged
poor maternal care to their pups (as defined by licking together with several other, differently responding,
and grooming). In addition to delivering poor mater- cell types. This is particularly problematic for micro-
nal care, these rats also develop long-lasting heightened array analyses, which depend on strong effects for
anxiety and stress responses. Meaney and colleagues statistical significance. One solution that is being cur-
identified a region of the glucocorticoid receptor (GR) rently developed is fluorescence-activated cell sorting
gene, which was hypermethylated throughout adult- (FACS), which may be used in conjunction with
hood in rats that received poor maternal care. Treat- BAC-transgenic mice that express green fluorescent
ment with an HDAC inhibitor not only reduced DNA protein (GFP) in specific cell types (e.g. dopamine D1
methylation on the GR receptor gene but also improved versus D2 neurons) [51]. This new technology will
anxiety and stress responses in these rats [49]. While permit the study of the mechanism by which cocaine
these data are correlative, they suggest an important alters gene expression and histone modifications in
role for epigenetic mechanisms in anxiety and stress isolated cell populations.
and suggest that DNA methylation at the GR gene The second major challenge is distinguishing
promoter (and probably other genes) may contribute causality from correlation. How does one determine
to this phenomenon. whether histone acetylation at the bdnf promoter
Taken together, these studies demonstrate that causes a transcriptional or behavioral response
chromatin structure is an important substrate for in vivo? To address this question, one must induce
long-lasting changes in behavioral responses to stress or prevent histone acetylation on the bdnf gene spe-
and antidepressant action. While the precise signaling cifically. Simply overexpressing a HAT or inhibiting
mechanisms by which environmental stresses con- HDACs would likely hyperacetylate numerous other
verge on chromatin remain unclear (e.g. Figure 8.2), target genes in addition to bndf, thus confounding the
these early studies suggest the exciting possibility transcriptional and behavioral interpretation of the
that pharmacological manipulation of chromatin experiment. Recently, an exciting breakthrough using

87
Chapter 8: Epigenetic mechanisms in drug addiction and depression

zinc finger proteins may enable investigators to dir- transcription or the transcriptional potential of genes
ectly address this very challenging question. Zinc which eventually affect neural function [37]. Thus,
finger peptides can be designed or screened for highly any study of chromatin regulation is inexorably
sequence-specific DNA binding properties. These linked with the study of the underlying gene activity.
zinc finger peptides can then be fused to a chromatin This is an especially important point for the purposes
remodeling enzyme, which effectively targets that of treating chronic psychiatric diseases, because a
enzyme to the promoter of a specific gene. This was drug that can permanently reverse the underlying
accomplished for the first time using a DNA methyl- transcriptional defect may circumvent the side effects
transferase in cell culture [52, 53], and may now and compliance issues that plague chronic therapies
permit behavioral neuroscientists, using viral-medi- based on antagonizing the dysregulated gene targets.
ated gene transfer or genetic mutant mice, to ask While extremely exciting, epigenetic research in
whether histone modifications at specific genes are psychiatry is still in its infancy, and far more research
indeed causally linked to the transcriptional and is needed to identify both the dysregulated genes and
behavioral phenotypes observed. chromatin modifications responsible for individual
psychiatric diseases and their improvement during
Concluding remarks effective therapy. Work is needed to determine
whether meaningful epigenetic modifications can be
There is now growing evidence that epigenetic mech-
detected in human peripheral tissues (blood, epithelial
anisms, such as histone acetylation and methylation,
are involved in regulating the salience of environmen- cells) that are useful in establishing a diagnosis or
tal stimuli [3, 10, 37–39, 41, 54]. In most studies to tracking treatment. Consideration of epigenetic
date, including learning and memory, chronic pain, mechanisms should inform and eventually facilitate
the identification of genetic factors that contribute to
addiction, and depression models, manipulations that
psychiatric syndromes. Future advances might
elevate histone acetylation potentiate the behavioral
include the development of novel positron emission
effects of the associated environmental stimulus. This
tomography (PET) or other brain imaging ligands
has important implications for the pathogenesis of
that would make it possible to assess global levels of
drug addiction and depression, since novel therapeut-
ics could target such mechanisms to ultimately block chromatin modifications in a patient’s brain. Finally,
or even reverse a chronically addicted or depressed it will be important to determine whether small mol-
state. As illustrated by the persistent hypermethyla- ecules that target chromatin modifying mechanisms
might be useful in the treatment of mental illness.
tion in mice exposed to chronic social defeat stress
(H3K27 methylation at the bdnf gene) and rats sub-
jected to poor maternal care (DNA methylation of the
GR gene), epigenetic mechanisms are also attractive Acknowledgements
candidates for novel molecular substrates that medi- Parts of this chapter were based on Tsankova et al. [3],
ate long-lived changes in brain. Ultimately, the key no permission needed, and Renthal et al. [55], with
function of chromatin remodeling is to alter the permission.

References 7. Strahl BD, et al. Nature 2000;


403(6765):41–45.
12. Oliveira AM, et al. Learn Mem
2007;14(9):564–572.
1. Surani MA, et al. Cell 2007;
128(4):747–762. 8. Kurdistani SK, et al. Cell 2004; 13. Doi M, et al. Cell 2006;125(3):
117(6):721–733. 497–508.
2. Luger K, et al. Curr Opin Genet Dev
1998;8(2):140–146. 9. Pokholok DK, et al. Cell 2005; 14. Kawasaki H, et al. Nature 2000;
122(4):517–527. 405(6783):195–200.
3. Tsankova N, et al. Nat Rev Neurosci
2007;8(5):355–367. 10. Levine AA, et al. Proc 15. Guan JS, et al. Nature 2009;
4. Felsenfeld G, et al. Nature 2003; Natl Acad Sci U S A 2005; 459(7243):55–60.
421(6921):448–453. 102(52):19186–19191. 16. Chawla S, et al. J Neurochem 2003;
5. Li B, et al. Cell 2007;128(4):707–719. 11. Maurice T, et al. 85(1):151–159.
6. Kouzarides T. Cell 2007;128(4): Neuropsychopharmacology 17. Fischl W, et al. Mol Cell 2002;
693–705. 2007;33(7):1584–1602. 9(1):45–57.

88
Chapter 8: Epigenetic mechanisms in drug addiction and depression

18. Yang XJ, et al. Nat Rev Mol Cell 31. Goll MG, et al. Science 2006; 43. Maze I, et al. Science 2011;
Biol 2008;9(3):206–218. 311(5759):395–398. 327:213–216.
19. Haigis MC, et al. Genes Dev 32. Miller CA, et al. Neuron 2007; 44. Krishnan V, et al. Nature
2006;20(21):2913–2921. 53(6):857–869. 2008;455(7215):894–902.
20. Brami-Cherrier K, et al. J 33. Ooi SK, et al. Cell 2008;133(7): 45. Berton O, et al. Science 2006;
Neurochem 2007;101(3): 1145–1148. 311(5762):864–868.
697–708.
34. Hyman S, et al. Annu Rev Neurosci 46. Tsankova N, et al. Nat Neurosci
21. Li J, et al. J Neurochem 2004; 2006;29:565–598. 2006;9(4):519–525.
90(5):1117–1131.
35. Nestler EJ. Philos Trans R Soc 47. Lee MG, et al. Chem Biol 2006;
22. Brami-Cherrier K, et al. J Neurosci Lond B Biol Sci 2008;363 13(6):563–567.
2005;25(49):11444–11454. (1507):3245–3255.
48. Schroeder FA, et al. Biol
23. Stipanovich A, et al. Nature 36. Bowers MS, et al. Neuron 2004; Psychiatry 2007;62(1):55–64.
2008;453(7197):879–884. 42(2):269–281.
49. Weaver IC, et al. Nat Neurosci
24. Vakoc CR, et al. Mol Cell 2005; 37. Kumar A, et al. Neuron 2005; 2004;7(8):847–854.
19(3):381–391. 48(2):303–314.
50. Rashid AJ, et al. Proc Natl Acad Sci
25. Iwase S, et al. Cell 2007;128(6): 38. Fischer A, et al. Nature 2007; U S A 2007;104(2):654–659.
1077–1088. 447(7141):178–182.
26. Jensen LR, et al. Am J Hum Genet 51. Lobo MK, et al. Nat Neurosci
39. Levenson JM, et al. J Biol Chem 2006;9(3):443–452.
2005;76(2):227–236. 2004;279(39):40545–40559.
27. Suzuki MM, et al. Nat Rev Genet 52. Li F, et al. Nucleic Acids Res
40. Freeman WM, et al. 2007;35(1):100–112.
2008;9(6):465–476. Neuropsychopharmacology
28. Zhang X, et al. Proc Natl Acad Sci 2008;33(8):1807–1817. 53. Smith AE, et al. J Biol Chem
U S A 2005;102(12):4459–4464. 2008;283(15):9878–9885.
41. Renthal W, et al. Neuron 2007;
29. Ramocki MB, et al. Nature 56(3):517–529. 54. Zhang C, et al. Cell 2002;110(4):
2008;455(7215):912–918. 479–488.
42. Schroeder FA, et al.
30. Chahrour M, et al. Science Neuropsychopharmacology 55. Renthal W, et al. Trends Mol Med
2008;320(5880):1224–1229. 2008;33(12):2981–2992. 2008;14:341–350.

89
Panic disorder
Chapter

9 Ardesheer Talati and Myrna M. Weissman

Panic disorder (PD) is a complex genetic anxiety medical services – notably emergency rooms [7–9] –
disorder. Although data from family and twin studies and is therefore costly not only to the individual
suggest familial heritability, genetic studies have been patient but to society at large. PD is frequently comor-
to date only partially successful in identifying putative bid with depression, other anxiety disorders, and sub-
genetic regions underlying the disorder. This chapter stance use and abuse [6, 10, 11]. Agoraphobia (AG)
provides a broad overview of the clinical and epidemi- (the fear of having attacks in public or unfamiliar
ological profile of PD, and a review of genetic studies, settings, in extreme cases resulting in complete con-
including promising areas of research and ways of finement to the home) is also very common [4]. It
subtyping the disorder for greater possible genetic should be noted however that even though AG most
tractability. We conclude with a review of neuroima- frequently presents as a complication to the primary
ging studies of PD to show how neurobiological data symptoms, not all agoraphobics meet criteria for PD
may be coupled with genetics to more comprehen- [5]; furthermore, of those that do, a third experience
sively probe the underlying mechanisms. AG symptoms prior to panic onset [12], and of those
that do not, some may fail to express overt panic
symptoms due to successful avoidance rather than a
Diagnosis and epidemiology true lack of vulnerability to the disorder itself. The
PD is a complex anxiety disorder characterized by epidemiology of PD and AG relative to each other is
recurrent attacks of sudden and uncontrollable fear. therefore not clear cut, and consensus on whether AG
The attacks typically come out of the blue, peak should be classified independently of, or as part of, PD,
rapidly, and are accompanied by a wide range of has been the subject of an enduring debate.
autonomic, cardiovascular and gastrointestinal symp- There is a confluence of evidence from family,
toms, as illustrated in Figure 9.1. DSM-IV criteria [1] twin, and molecular genetic studies, suggesting a gen-
require that the attacks be both unexpected and recur- etic contribution to PD. If genes influence the suscep-
rent, and that they be followed by at least one month tibility to PD, biological relatives of individuals with
of either the fear of having additional attacks, or the the disorder should have a higher prevalence of PD
development of phobic avoidance or other significant than family members of controls without the dis-
changes in behavior. The attacks must also not be order. Twin studies further examine the relative con-
explainable by neurological or general medical condi- tribution of genetic and environmental influences on
tions, medication, or substance use. the variance in PD risk. Finally, molecular genetic
The lifetime prevalence of PD is approximately studies seek to identify specific genes that increase
1–5% [2–4], although isolated panic attacks occur risk of developing the disorder. These studies are
more frequently [5]. The epidemiology is similar across reviewed respectively in the following sections.
cultures, with first onset typically in early adulthood
(though it can occur in children), and higher rates Family and twin studies
among women [4, 6]. Because the symptoms of a panic A number of family studies have reported familial
attack can mimic serious medical conditions (such as a co-aggregation of the disorder, with first-degree
heart attack), the disorder is associated with high use of relatives (FDRs) of persons with PD, as compared to

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

90
Chapter 9: Panic disorder

Figure 9.1 Panic attacks and panic


Increased disorder. The panic attack is the central
Sweating feature of the disorder, but is typically
heart rate
accompanied by a range of
cardiovascular, gastrointestinal, and
Derealization or Chest pain neurological symptoms. Not all panic
depersonalization attacks lead to, or are associated with
the full panic disorder syndrome, which
require repeating panic attacks with
Fear of going crazy
Panic Shortness of breath significant and lasting impairment.
attack
or dying
Requires ≥ 4 of the
surrounding symptoms
Stomach/abdominal Dizziness
distress

Numbing or tingling
Hot /cold flashes

Attacks are recurrent and unexpected

And accompanied by:


(at least one or more, for one or more month)

Worry about Concern about Modification in


consequences additional attacks behavior

And not better accounted for by:

Other anxiety General medical


Substance use
disorder condition

Panic disorder

those of persons without, at several-fold greater risk familial transmission does not appear to particularly
for PD themselves [13–17] (A recent weighted meta- vary by gender or parent-of-origin [20, 21], earlier
analysis reported an adjusted mean seven-fold onset has been associated with greater familial risk.
increase [18].) Despite the frequent comorbidity with Goldstein et al. for example reported that whereas
depression and other anxiety disorders such as gener- FDRs of PD probands with onset after age 20 were
alized anxiety disorder (GAD), at least some of the at 6-fold higher risk than those of controls, those of
transmission appears specific to PD [14, 19]. Though probands with onset under age 20 were at 17-fold risk

91
Chapter 9: Panic disorder

100 of induction of panic attacks using 35% carbon dioxide


Percent of variance due to heritable factors

among healthy MZ (56% concordant) than DZ (12%)


twins [29].
Even though twin studies clearly document herit-
ability for PD, they also suggest that much of this
heritability may be shared with other mental dis-
50 orders. In a large study of more than 5600 twin pairs
drawn from the population-based Virginia Twin
Registry, Kendler et al. [30] examined the underlying
structure of risk factors for five internalizing dis-
orders, including PD, major depressive disorder
(MDD), GAD, animal phobia, and situational phobia.
0 The authors identified a genetic factor shared with
sm
d
ize r rd
er bi
a
ism ni
a n’s MDD and GAD that explained  10% of the total
tici ral rde iso p ho h ol h re g to se
a
ro e o d l co p n
ti se variance in liability to PD; a separate genetic factor
u en is c cia Al zo un di
Ne G ty d ani o c hi H
xie
P S S with strong loadings on animal and situational
an phobias that explained 2% of the liability, and about
Figure 9.2 Comparison of heritability estimates for panic disorder 4% of variance that was specific to PD. In a subse-
with those of other major psychiatric disorders. Comparison of the quent examination of anxiety disorders within the
occurrence of mental illness in monozygotic and dizygotic twins
reveals the influence of genetic factors for major psychiatric same sample, Hettema et al. [31] identified another
disorders. Heritability for panic disorder is higher than for genetic factor that explained 25% of the total variance
generalized and social anxiety disorders, but lower than for for PD and GAD. Furthermore, the underlying liabil-
alcoholism and schizophrenia. (From [9], with permission from
Macmillan Publishers Ltd.) ity structure appeared similar across gender, despite
the higher prevalence of anxiety disorders among
[22], suggesting that early-onset PD may represent a women. The authors hypothesized that rather than
phenotypic subtype with greater genetic loading, and individual vulnerability for each disorder, a set of
thereby be of particular value in genetic studies. genes may instead increase vulnerability to some
Family studies alone however cannot tell us intermediate phenotype (such as an anxious person-
whether a disorder runs in families for genetic or ality trait like neuroticism), which in interaction with
environmental reasons, as family members are not other genes and/or stressful life events may increase
only more genetically similar to each other but are the risk of developing one or more anxiety disorders.
also more likely to share the same environmental
exposures. Twin studies have the advantage that they
allow us to further disentangle the relative contribu- Molecular genetic studies
tions of genetic and environmental influences. For Whereas family and twin studies can identify the
PD, twin studies suggest a significant genetic contri- presence of genetic risk, molecular genetic studies
bution, with heritability estimates typically ranging allow us to further probe the actual genetic markers
between 0.3 and 0.4, and as high as 0.48 [23–28]. conferring that risk. The most common types of
(See Figure 9.2 for a comparison with estimates of molecular genetic approaches are linkage and associ-
other major disorders.) Precise estimates vary by ation studies. Linkage studies investigate the entire
ascertainment criteria, a point we will revisit later in genomes of individuals, using DNA markers to locate
this chapter. Inclusion of AG with panic attacks [28] chromosomal areas that are passed on together with a
or phobic avoidance [24] for instance increases the particular disorder. Linkage studies have implicated
heritability estimate; inclusion of sporadic panic a number of chromosomal regions in PD, including
attacks on the other hand diminishes it, as illustrated 1q [32, 33], 2p,q [34], 4q [35], 7p [32, 36, 37], 9p,q
by a study by Perna et al. [25] showing that whereas [34, 38], 11p [33], 12q [39], 13q [40, 41], 15q [34], and
73% of monozygotic (MZ) and no dizygotic (DZ) 20p [36]. However, these studies have failed to iden-
twin pairs were concordant for PD, 57 and 43% were tify major gene loci, underscoring the genetic com-
concordant respectively for panic attacks. Finally, plexity and likely multifactorial inheritance patterns
twin studies report significantly greater concordance of the disorder.

92
Chapter 9: Panic disorder

Association studies instead test whether specific variation within the transporter predict who would
genetic variants (polymorphisms) of interest are asso- respond to treatment and who would not? To test this
ciated with a given disorder, typically by comparing question, Perna et al. [49] employed a 12-week study
affected cases with unrelated controls. Genome-wide of 92 PD patients being treated with the SSRI parox-
association studies (GWAS), which allow testing of etine and measured whether the polymorphism was
entire genomes without requiring a priori hypotheses, associated with treatment response or remission.
can play an important first-step to identify candidates They found that female patients with one or more
which can then be more narrowly targeted in associ- long variants of the transporter promoter (i.e. “ll” or
ation studies. There are however no published (or “sl”) had a significantly greater reduction in panic
to our knowledge, ongoing) GWAS of PD; rather, symptoms than those with only the short variant
the genetic research here has largely focused on (“ss”), suggesting that bioavailability of serotonin
hypotheses within specific neurotransmitter systems, may be a primary determinant of the panicolytic
suggested by animal models or clinical data. effect. The specificity to females may reflect inter-
action with sex hormones, a notion supported by
evidence that progesterone can protect against panic
Genes involved in the serotonin system attacks [50], and estrogen exacerbate them [50, 51].
Given that drugs modulating serotonin levels are the A further study compared serotonin transporter bind-
most efficacious and widely used line of treatment ing within current and remitted PD patients to study
[42, 43], the greatest convergence of genetic studies the effect of treatment on receptor binding [52].
of PD has also not surprisingly been on genes involv- Patients with PD exhibited a significant reduction in
ing serotonergic transmission. Of these, the most binding within midbrain raphe nucleus, temporal and
targeted is the gene encoding the serotonin trans- thalamic regions of the brain. Binding within the
porter protein (SLC6A4), the binding site of the raphe and temporal regions was dose dependent, with
selective serotonin reuptake inhibitors (SSRIs) class symptom severity inversely related to binding levels.
of drugs (e.g. fluoxetine, paroxetine, sertraline, citalo- Interestingly, remitted patients still had lower binding
pram, escitalopram), and particularly a functional in the thalamus, but in all other regions they were
polymorphism within the promoter known as indistinguishable from controls. The authors posit
5HTTLPR (serotonin transporter linked polymorph- that the reduced binding in PD patients may reflect
ism region). This polymorphism was originally iden- either a deficit of the transmitter or some compen-
tified as a 14 (short) or 16 (long) base pair repeat satory mechanism to increase transmitter levels, and
sequence (additional functional variants have since that treatment may serve to re-normalize transporter
been identified) that determines the amount of sero- binding within the temporal and thalamic regions.
tonin and is related to depression and anxiety related Because thalamic binding does not change based on
traits [44–46]. But numerous investigative groups remission, the authors advance that this may reflect a
have failed to find any association [44–57] as elegantly trait marker that is independent of symptom severity.
summarized in a recent meta-analysis [47] revealing SLC6A4 is only one of a number of genes that can
a null overall odds ratio of 0.9. We also found no modulate serotonin levels. The serotonin receptor
evidence of association within the promoter in a family includes a heterogeneous class of G-protein
recent case-control study of PD [48]; a single nucleo- coupled receptors (except for 5-HT3, which activates
tide polymorphism (SNP) within an intron however ion channels) that serve multiple pre- and post-synaptic
was associated with the panic phenotype (but not roles in serotonin neurotransmission [53]. A number of
social anxiety), and we are currently investigating its polymorphisms within the receptor family have been
functional significance. Though our observations of tested, but with inconclusive results. The 5-HT1A
course require replication, they nevertheless under- receptor is the binding site for serotonergic agonists
score the importance of incorporating the entire gen- such as buspirone, and animal studies have suggested
etic region of the transporter rather than zeroing in a role for this subtype in anxiety and depression.
on the promoter region only. Klemenhagan et al. [54] found that 5-HT1A knockout
If selective serotonin reuptake inhibitors (SSRIs) (KO) mice freeze to a larger array of familiar and
are the first line of treatment for PD, and if they work unfamiliar cues following contextual fear condition-
by targeting the serotonin transporter, can genetic ing than their wild type counterparts, who only freeze

93
Chapter 9: Panic disorder

when presented with specific preconditioned threat- association studies [68, 71, 72], each of which showed
associated stimuli. The behavior by the KO mice an association among females only. Although a link-
parallels that of PD and post-traumatic stress disorder age study did not replicate these findings [73] that
(PTSD) patients, who tend to overly generalize threat may have been because power was limited, and only
and therefore may respond with fear even to neutral half of the families in that study stemmed from female
situations. In human studies, PD patients have also probands. Regardless, the sex specificity is interesting,
been shown to have reduced 1A binding in the raphe especially given the location of the MAOA gene is
nucleus (the primary site of serotonergic neurons), located on the X-chromosome, and may reflect inter-
cingulate cortex, orbitofrontal cortex, and amygdale actions with sex hormones or other X-linked genes.
[55, 56]. Interestingly, remitted patients showed A separate polymorphism (T941G) within the MAOA
reduction in presynaptic binding but unimpaired gene itself has been associated with generalized anx-
post-synaptic binding, suggesting that post-synaptic iety but not PD [74], and a study testing synergy
receptor availability may modulate vulnerability to between MAOA, VNTR, and the serotonin trans-
PD [55]. There is also some genetic evidence that a porter promoter polymorphism (5-HTTLPR) did
cytosine-guanine substitution at position 1019 of the not find any interactions.
5-HT1A receptor may predispose to anxiety, but spe- Finally, whereas MAOA degrades serotonin, tryp-
cificity to PD is questionable, as in one study only PD tophan hydroxylase (TPH) serves as the rate-limiting
cases with comorbid AG showed any association [57], enzymatic step in it synthesis. Unlike MAOA, how-
while in another, there was an association with panic ever, reports of TPH have been mostly negative [68,
attacks but not PD per se [58]. 75–78] including tests of a recently recognized variant
For the class 2 receptors, findings are inconsistent. preferentially expressed in neurons [79].
Selective ablation of 5-HTR2A in mice reduces con-
flict anxiety but not depressive or fear-conditioning
related behaviors [59]. But PD patients typically do Genes involved in the dopamine system
not respond to 5-HT2A agonists [60]. A T102C poly- Of the genes related to dopaminergic function,
morphism within 5-HT2A was associated with the catechol-O-methyltransferase (COMT), a methylation
pure panic phenotype in one sample [61], but in enzyme that degrades catecholamines (such as dopa-
another, was only associated among the AG subtype mine, epinephrine, and NE) and is the primary deter-
[62], and in a third there was no association even if minant of cortical dopamine levels, has been the most
AG was present [63]. A fourth case-control study also encouraging. A conversion of a valine to a methionine
did not identify any genetic differences between cases at position 158 has been much focused on due its
and controls, but within the group of PD cases, a functional significance: the valine variant results in
number of SNPs predicted symptom severity [64]. three to four times as much enzymatic activity as the
Animal models have also implicated 5-HT2C in methionine. This polymorphism, commonly known
panic-like behaviors [65], but the human receptor, as val158met (dbSNP rs4680) has been associated
though associated with anxiety and depressive pheno- with PD in a number of studies, particularly among
types [66] is not associated with panic [60, 62, 67]. women [80–85], but these are tempered by some
Finally, polymorphisms within 5-HT1B [60, 61, 68] failures to replicate [72, 86, 87] or reports of nonspe-
and 5-HT3A [61, 69] have yielded exclusively negative cificity [23]. Even within the positive studies, some
findings. report the valine encoding allele (i.e. an excess of
Monoamine oxidase A (MAOA) is another COMT) associated with presence of PD, but the stud-
important candidate as it plays a central role in the ies by Woo et al. [84, 85] finding the reverse.
degradation of serotonin as well as norepinephrine This apparent discrepancy might perhaps be
(NE). A functional polymorphism upstream of the explained due to ethnic heterogeneity, given the find-
gene that consists of a 30 base pair variable number ings of a recent meta-analysis showing no overall
tandem repeat (VNTR) sequence [70] has been association between COMT val158met and PD, but a
targeted as it alters transcription levels of the gene, val encoding allele effect on PD among Caucasians,
resulting in the longer version having greater enzym- and a met effect in Asian populations [80]. In both
atic activity. The VNTR polymorphism has been asso- cases, though, the association was predominantly
ciated with the panic phenotype in at least three among females, paralleling gender-specific reports of

94
Chapter 9: Panic disorder

COMT in obsessive–compulsive disorder [88] and (CCK) and its receptors have been of particular inter-
schizophrenia [89]. Though the mechanisms of this est as administration of CCK can induce panic attacks
gender specificity are not entirely clear, estrogen may in healthy individuals, PD patients are more sensitive
play a modulating role, as the COMT gene contains to CCK than controls, and anti-panic medications
an estrogen response element within its promoter can alter the CCK gene [102, 103]. SNPs within the
region [90]. Interestingly, the COMT enzyme also CCK gene itself have been linked to PD [104–106];
plays a role in the conjugation of estrogens, so the polymorphisms within the receptor however have
two products may cross-regulate each other [91]. yielded both positive and negative findings for CCK-A
Freitag et al. [92] further reported an interaction [107–109] and largely negative findings for CCK-B
between the aforementioned COMT val158met and [106, 108, 110–112].
5-HT1A 1019C/G polymorphisms among PD cases. Keck et al. [113] examined the relationship
Interestingly, they did not find any additive effects: between the vasopressin (AVP) neuropeptide system
instead, the risk for PD associated with each variant and the corticotropin releasing hormone (CRH), gen-
was highest in the presence of the low risk allele of otyping 71 SNPs within the CRH, CRHR1, CRHR2,
the other polymorphism, suggesting “a ceiling effect AVP, AVPR1A, and AVPR1B genes. A number of
at the molecular level . . . that if one element of mono- polymorphisms within CHRH1 and AVPR1B were
aminergic transmission is dysfunctional, dysfunction nominally associated with PD, and 15 of them showed
of other elements does not further increase the risk significant gene-by-gene (epistatic) interactions.
for PD” [92]. Even though this observation should be These findings are provoking given that the expres-
treated cautiously given the small sample, it serves to sion of CRHR1 and AVPR1B has been shown to most
remind us of the complexity and interdependency of strongly co-localize in the hippocampus, amygdala,
genetic risk factors. A separate study similarly tested and pituitary gland, all brain regions relevant to anx-
for interactions between COMT and the MAOA, iety. Furthermore, unlike so many other polymorph-
but did not find any [81]. Finally, dopamine receptor isms identified, the most significant SNPs lie within
1 (DRD1) has been associated with PD [61], but genetically consequential exon (AVPR1B) and 3’ UTR
studies of DRD2 and 4, and the dopamine transporter (CRHR1) regions, allowing for the genetic variation
(DAT) have yielded mostly negative results [93, 94]. to have a direct functional impact. Finally, the associ-
ations are concordant both with murine models asso-
Genes involved in the adenosine system ciating these hormone receptors with anxiety related
behaviors [114], and with human linkage studies
Though not a neurotransmitter in the classical sense
of PD implicating regions of chromosome 1q that
(it is a purine nucleoside), adenosine can alter levels
contains the human AVPR1B gene [32].
of several other neurotransmitters and neuromodula-
tors [95]. There are four receptor subtypes, but the
adenosine receptor 2A (ADORA2A) on chromosome Genes involved in GABA-ergic system
22, is the most provoking candidate as it is preferen-
The efficacy of benzodiazepines in treating panic
tially expressed in the brain, and excess caffeine – a
symptoms spotlighted the role of the g-aminobutyric
potent 2A antagonist – can trigger panic attacks [96].
acid (GABA) system. No significant associations to
The subtype has been associated with a number of
our knowledge have been found thus far for any of
psychiatric phenotypes in mice [97]. There is now
the GABA receptor subtypes [93], although the b3
evidence that exons within the ADOR2A but not
subunit has been implicated in PTSD [115]. However,
other receptor subtypes may be involved in human
lower levels of glutamic acid decarboxylase (GAD), a
PDs as well, with two Caucasian populations showing
key enzyme in the synthesis of GABA, has been asso-
a significant association [98, 99].
ciated with depression and bipolar disorder, as well as
in neurotic patients [116]. In humans, GAD exists as
Genes involved in neuropeptides systems one of two isoforms, encoded by the GAD1 and GAD2
A number of neuropeptides have also been associated genes. Though there have been no concrete findings
with PD including neuropeptide Y Y5 receptor [100], for PD, variation within the GAD2 gene has been
tachykinin receptor 1 (TACR1) [101], and gastrin related to behavioral inhibition, and is anxiety related
releasing peptide (GRP) [101]. Cholecystokinin [117], whereas that within GAD1 has been linked to a

95
Chapter 9: Panic disorder

genetic risk commonly shared across multiple anxiety in sensitivity to hypercapnia may also serve as useful
and depressive phenotypes [118]. markers to study the etiology of human PD, especially
given that this phenotype appears selective to panic.
Genes involved in the regulators Serotonin is thought to play an important role in
hypersensitivity: serotonergic neurons in the medulla
of G-protein signaling family are central respiratory chemoreceptors, and seroto-
Finally, associations with PD have been found for a nergic neurons in the midbrain of rats are also highly
family of proteins known as the regulator of G-pro- chemosensitive to small changes in blood CO2, indu-
tein signaling (RGS). The RGS proteins can decrease cing changes in arousal, anxiety, and cerebrovascular
G-protein function via GTPase activity, and have tone [127, 128]. In a genetic study of the serotonin
been shown to decouple autoreceptors from noradre- transporter gene, Schmidt et al. found that healthy
nergic neurons in the locus coeruleus [119]. Two subjects homozygous for the long allele of the 5-
members of the large RGS family have been involved HTTLPR promoter polymorphism showed higher
in the etiology of PD; these two genes lie on chromo- sensitivity and a stronger fear response to 35% CO2
some 1q, one of the risk loci suggested for PD [33]. [129] than subjects with the heterozygous or homo-
The RGS2 variant has been linked to PD in males zygous for the short allele. But a subsequent extension
[120] and RGS7 [121] in females. of this hypothesis to study a population of PD
patients [130] failed to find any differences. Philbert
Genes related to carbon dioxide et al. [131] examined lactate dehydrogenase (LDH),
the enzyme that converts pyruvate (the final product
hypersensitivity of glycolysis) to lactic acid under anaerobic condi-
PD is unique within the anxiety spectrum in that its tions. LDH has two separately encoded variants,
central manifestation, the panic attack (PA), can be LDH-A and LDH-B. The authors found that a haplo-
incited in laboratory settings by specific chemical type within the fifth exon of the LDH-A gene predicted
challenges such carbon dioxide (CO2) inhalation. ventilation response to CO2 inhalation that could
The recognition that an increase in brain CO2 levels be differentiated between subjects at high and low risk
is an indicator of potential suffocation, coupled with for PD. Finally, the cholinergic system has also been
the ability of CO2 inhalation to generate panic attacks, postulated to play an evolutionary role in CO2 sensi-
has promoted the false suffocation alarm theory of tivity [132], although there is no molecular evidence
PD [122], in which a hypersensitive suffocation alarm implicating specific gene variants to our knowledge.
system triggers the suffocation alarm erroneously (i.e.
in the absence of any real danger) producing respira-
tory distress, hyperventilation, and panic. In that Genetic studies: conclusions
sense, initial attacks may arise from an unconditioned If there is a single conclusion from the above studies,
biological origin, but among those who go on to it is one of inconsistency. For many of the reported
develop the full disorder, successive associative learn- associations, replication in independent samples – so
ing or classical conditioning may continue to generate vital to psychiatric genetics [133] – has either not
learned alarm responses to internal or external cues. been attempted, or has failed. Some failures to repli-
There is a body of converging evidence for a cate might be method-related, such as small sample
genetic basis to CO2 sensitivity. Unaffected FDRs of sizes or variation in how the panic phenotype is
PD patients have increased hypersensitivity to CO2 operationalized. But other failures could represent
[123, 124], and MZ twins are substantially more con- true negatives, with the originally reported associ-
cordant on CO2 induced panic induction than DZ ations having being the false-positive victims of popu-
twins [29]. Genetic modeling suggests that common lation stratification or use of overly liberal statistical
genetic factors may account for trait sensitivity to tests, etc. Regardless, given that negative findings have
CO2 challenge and naturally occurring PAs [125], a lower overall likelihood of being published, true
and that transmission of PD is higher among families replication rates are likely even lower than what a
in which the proband is responsive to the CO2 chal- literature search would bear forth.
lenge [126]. Given these observations, the discovery of Given these concerns, it is imperative that investi-
genetic variants associated with individual differences gators rigorously document their study characteristics

96
Chapter 9: Panic disorder

so that when discrepant findings emerge, they can be variation and then study the mechanisms via which
probed analytically rather than simply descriptively. genes can impact behavior.
Some characteristics such as gender, age, and ethni-
city, are obvious, yet critical, as they can substantially
impact the findings. (Consider the earlier discussion
of COMT: not only were associations gender-specific,
The “expanded spectrum” approach
but they veered in opposite directions across different and a possible PD syndrome
ethnic groups.) Others, such as onset age or family In this section, we describe a series of studies on
history, however are less frequently considered even a potential PD syndrome conducted by our group,
though these may impact genetic loading. Interview- which grew out of a single initial observation: specific
ing relatives directly may be prohibitive, but there are medical conditions tended to cluster nonrandomly
a number of standardized instruments such as the within some PD families but not others. The objective
Family History Screen (FHS) [134] and the Family of this section is not only to provide an overview of
Interview for Genetic Studies (FIGS) [135] through this syndrome, but also to serve as an example of an
which family history can be rapidly and reliably ascer- investigative approach that can be used to refine the
tained. Finally, the importance of selecting appropri- clinical phenotype in ways that may be useful to
ate controls cannot be understated, as illustrated in a genetic investigation. The overall arc of these studies
recently published analysis comparing effect sizes in is conceptualized in Figure 9.3; details of individual
PD studies using different control populations [136]. studies are summarized in Table 9.1 [48, 136–139].
Controls should be recruited and assessed using the A distinguishing feature of PD is that is frequently
same methods and instruments as cases. Ideally, they accompanied by a range of medical symptoms,
should be free of all anxiety disorders, depression, and including cardiovascular (e.g. shortness of breath,
substance use, since these are frequently comorbid palpitations), gastrointestinal (nausea, abdominal
with PD and could fuzzy the diagnosis. They should distress), urinary, and neurological (dizziness, trem-
not only be matched on age, gender, and ethnicity, bling) [140–143] (see also Figure 9.1). It is in fact
but also be selected to ensure that they have passed often these clinical manifestations that drive individ-
the age of greatest risk for the disorder of interest. PD uals to their doctor or emergency room, where their
for instance onsets in early adulthood; inclusion of PD is first identified [7–9]. In the course of a previous
teenagers as controls might contaminate the sample genetic study of multiplex families with PD, we
with some subjects who though not symptomatic, observed that some medical problems – particularly
may still be at the same genetic risk as persons with those related to bladder, thyroid, and mitral valve
the disorder. Having said that, the more restrictive the function – were not randomly distributed. Instead,
selection criteria, the smaller will be the available pool they tended to cluster within some families with PD
of subjects to recruit from, and the trade-offs between but not others. We hypothesized that families who
sample size and sample homogeneity will need to be had this clustering of conditions might perhaps be
carefully evaluated for each context. genetically different from those that did not, and
Genetic studies, while instrumental, cannot alone examined the linkage patterns after reclassifying the
address the etiological complexities of most psychiatric families based on the presence or absence of these
disorders. In the remaining sections of this chapter, we medical conditions. Indeed, reclassification revealed a
turn to two integrative approaches that combine gen- strong linkage peak (logarithm of odds [LOD] score:
etics with other clinical or biological methods to target 3.6) on marker D13S779 of chromosome 13q, raising
the underlying mechanisms. In the first, we discuss the hypothesis of a possible syndrome wherein these
exploiting the relationship between psychiatric and diverse symptoms could be sub-served by a common
nonpsychiatric medical manifestations (the “expanded genetic mechanism related to this chromosome
spectrum” approach). This approach is particularly (pleiotropy) [41, 136]. These original findings were
relevant to PD, where the panic attacks are accompan- subsequently replicated in a larger sample of 587
ied by a range of physiological responses that may be individuals from 60 multiplex pedigrees (including
central to the etiology. Then, we focus on neurobio- the original 19 families), that reported a linkage peak
logical phenotypes, and in particular, on using meas- on marker D13S793, only 7 cMs from the original
ures of brain structure and function to identify genetic peak [40].

97
(a)
PD probands First-degree relatives
80

70

60
Frequency (%)

50

40

30 +

20

10

0
IC

MVP

MIG

TH

ANY

IC

MVP

MIG

TH

ANY
Medical conditions

(b)
IC probands First-degree relatives
80
+
70

60
Frequency (%)

50

40

30

20

10

0
PD

MVP

MIG

TH

ANY

PD

MVP

MIG

TH

ANY

Medical conditions
Figure 9.3 Case-control studies of panic disorder (PD) and interstitial cystitis (IC). (a) Rates of IC and other hypothesized medical
conditions among probands with and without PD (left) and their first-degree relatives (FDRs) (right). PD probands were at significantly
increased risk for IC, mitral valve prolapse (MVP), and migraines than controls; FDRs of PD probands were also at increased risk for IC, MVP,
migraines, and thyroid problems than FDRs of controls. (b) Rates of PD and other hypothesized medical conditions among probands with and
without IC (left) and their FDRs (right). IC probands were at significantly increased risk for PD and thyroid problems; FDRs of IC probands were
at increased risk for PD and migraines than FDRs of controls. ANY, any syndrome condition (that is, PD, IC, MVP, MIG, or TH); IC, interstitial
cystitis; MIG, migraine headaches; MVP, mitral valve prolapse; PD, panic disorder; TH, thyroid problems.
**** p < 0.001; *** p < 0.005; ** p < 0.01; * p < 0.05; + p < 0.1
P-values reflect odds ratio, adjusted for age and gender. For data on FDRs, the odds ratios are further adjusted for whether or not the
proband reported the same condition. For example, the odds ratio for whether a FDR of a panic proband was more likely to have mitral valve
prolapse than a FDR of a control proband, were adjusted for the FDRs age, gender, and whether or not the corresponding proband had
mitral valve prolapse. Darker bars indicate probands (i.e. PD cases in part (a); IC cases in part (b)); lighter gray bars indicate controls. In each
figure, the left panel shows data for probands versus controls; the right panel shows data for FDRs of probands versus FDRs of controls.
PD was diagnosed based on DSM-IV criteria using the Schedule for Affective Disorders-Lifetime version (SADS-LA-IV). Medical conditions in the
proband were ascertained by a medical checklist at the time of the interview; IC was also assessed using screening criteria developed by
the National Institute of Diabetes and Digestive and Kidney Diseases (NIDDK) for use in genetic studies of IC. Medical conditions in the FDRs
were assessed using the Family History Screen, with the proband as informant.

98
Chapter 9: Panic disorder

Table 9.1 Potential PD syndrome: study characteristics.

Study Sample Assessments Ref.


1. Family study of PD 34 families Schedule for Affective Disorders and Fyer et al., 1999;
Schizophrenia – Lifetime Version IV Weissman et al., 2000
2. Replication of (1) 60 families Schedule for Affective Disorders and Hamilton et al., 2003
Schizophrenia – Lifetime Version IV
3. Case-control study 67 IC cases; 79 non-IC Assessment of IC: Urodynamics and/or Weissman et al., 2004
of IC urological controls cystoscopy with bladder distention
Assessment of PD and other psychiatric
disorders: Schedule for Affective Disorders
and Schizophrenia – Lifetime Version IV
4. Clinical case- Cases: 219 (PD); 199 Assessment of IC: Medical history; NIDDK Talati et al., 2008
control (SAD); 173 (PD + SAD); screen for IC
study of PD Controls: 102a. Assessment of psychiatric disorders: Schedule
for Affective Disorders and Schizophrenia –
Lifetime Version IV
Required onset by age 30 and family history
of anxiety in  1 FDR
5. Genetic case- Casesb: 179 (PD); 161 Schedule for Affective Disorders and Strug et al., 2008
control (SAD); 140 (PD + SAD); Schizophrenia – Lifetime Version IV
study of PD Controls: 470c.
6. Replication of IC Casesb: 179 (PD); 161 Assessment of IC: Medical history; NIDDK Underway
linkage findings (SAD); 140 (PD + SAD); screen for IC
from (1) and Controls: 574d. Assessment of psychiatric disorders: Schedule
(2) using case- for Affective Disorders and Schizophrenia –
control sample (4) Lifetime Version IV
a
These controls were collected as part of our study, and were required to have no lifetime personal or family history of any psychiatric disorder.
b
These are the same cases as those in the clinical study (no. 4). The sample is smaller here due to subject attrition (some subjects
completed the interview but did not give blood), or genotyping failure.
c
The controls for this analysis were not originally collected as part of our sample but rather obtained from a national repository (the Rutgers
University Cell and DNA repository [RUCDR]) that had been set up by the National Institute of Mental Health’s (NIMH’s) Human Genetics
Initiative to aid genetic studies. Subjects were assessed using an on-line self-report based on the Composite International Diagnostic
Interview–Short, assessing eight syndromes: major depressive episodes, panic attacks, agoraphobia, social phobia, specific phobia, generalized
anxiety disorder, obsessive–compulsive disorder, and alcohol and drug dependence (see Talati et al. 2008 [136] for more information on
the NIMH controls). From the available sample of more than 3000 subjects, we selected the 480 subjects who reported no history of any
anxiety, depressive or substance use disorders, and were in the same age and gender range as our cases.
Abbreviations: FDR, first-degree relative; IC, interstitial cystitis; NIDDK, National Institute of Diabetes and Digestive and Kidney Disease;
PD, panic disorder; SAD, severe anxiety disorder.

Of the various conditions clustering with PD, the women than men, although with later onset at around
strongest linkage evidence was for kidney and bladder 40 years [145], and is accompanied by neurological,
problems (LOD scores > 3). It was unclear at the time gastrointestinal, and musculoskeletal distress [145–147].
what these symptoms represented, so we presented Data from twin studies allude to possible genetic
the findings to expert urologists, who suggested these underpinnings [148, 149], and a large genetic study
problems may be related to interstitial cystitis (IC). IC of IC (Maryland Genetics of Interstitial Cystitis
is a low-prevalence ( 0.5%, lifetime) but chronic and [MAGIC]) is currently underway.
debilitating syndrome characterized by frequent or To follow up on the linkage patterns from our
urgent need to urinate, and pain and discomfort on family studies, we designed a case-control study of
bladder filling, not accounted for by a urinary tract IC, followed by a larger case-control study of PD.
infection [144]. Like PD, IC is more common among For the former, we recruited 67 IC cases who were

99
Chapter 9: Panic disorder

(a) Figure 9.4 Conceptualization of the


studies to test the panic disorder
Initial observation during a genetic linkage study of PD that
syndrome. FDR, first-degree relatives; IC,
specific medical problems appear to cluster nonrandomly among
interstitial cystitis; LOD, logarithm of odds;
some families but not others
NIDDK, National Institute of Diabetes and
Linkage study of PD

Digestive and Kidney Diseases; PD, panic


Reclassification of families based on presence/absence disorder; SAD, social anxiety disorder.
of medical conditions and reanalyze data

Linkage peak on chromosome13q, marker D13S779

Replication of above finding in a larger sample of 60 families, with


linkage peak on marker D13S793, only 7 cM from first peak

(b)
Association study of IC

Largest LOD score for bladder problems; urologist consultation


suggests symptoms represent interstitial cystitis (IC)

Design new case-control study of IC versus non-IC urological


patients

Findings: IC patients are at increased risk for PD; so are their first
degree relatives, regardless of whether or not the proband had PD

(c)
Case-control study of PD and SAD using revised criteria to screen
Association study of PD

for IC developed by the NIDDK for use in genetic studies

Panic probands at five-fold higher risk for IC, and two-fold or


greater risk for mitral valve, migraines. FDRs also at higher risk for
these disorders regardless of proband having these conditions

Fine mapping to chromosome 13q in the area originally


identified by the linkage analysis to hone in on the functional
genetic region that may be involved in this pleiotropic syndrome

well-documented by urologists using urodynamics PD status, consistent with a pleiotropic model of


and/or cystoscopy with bladder distention [138]; con- transmission that extends beyond either PD or IC
trols (N = 79) were drawn from patients with other alone. The results are summarized in Figure 9.4b.
urological disorders but not IC. A total of 815 FDRs The second case-control study was designed to
were also assessed using the family history screen examine the same question from the opposite per-
[134]. We found that 27% of IC patients in the spective: i.e. patterns of IC within PD cases (Figure
sample, as compared to only 8% of controls, had been 9.4c). In order to test for specificity, we also included
diagnosed with PD, an adjusted 4-fold increase. Very an additional group of persons with social anxiety
similar patterns have since been independently disorder (SAD). IC was now assessed using revised
observed in the Events Preceding Interstitial Cystitis criteria developed by the National Institute of Dia-
(EPIC) study, where 27% of IC cases but 3% of con- betes and Digestive and Kidney Diseases (NIDDK)
trols had a PD diagnosis [Warren et al., unpublished]. for use in genetic studies [150]. To increase the likeli-
Strikingly, in our study, the FDRs of probands with hood of capturing a genetic phenotype (see earlier
IC were also three-fold more likely to have PD than discussion) cases were required to have onset prior
the FDRs of controls, regardless of the proband’s own to age 30, and a documented history of anxiety in

100
Chapter 9: Panic disorder

at least one FDR. Controls, by contrast, were required innervate smooth muscles throughout the body. The
to have no evidence of any psychiatric disorder over dysregulation leading to the increased fear response
their lifetime, have no family history of anxiety, and could therefore couple with the excess sympathetic
be at least 30 years old at the time of assessment projections to muscles in pulmonary, cardiovascular,
(the last criteria serving to minimize capturing those genitourinary, and gastrointestinal organs, accounting
still at risk). for the coexistence of PD and the syndrome’s other
The findings in this latest sample were even more medical conditions.
robust than our previous studies. Here, PD probands, Though not the only explanation, a number of the
as compared to controls, were more than nine-fold as observed conditions fit nicely within the autonomic
likely as controls to report IC symptoms (but not any dysregulation model. For example, autonomic nuclei
other genitourinary problems), and more than 4% of are the major source of innervation to the bladder
them had been formally diagnosed by a physician or and urethra [155, 156], and IC patients demonstrate
urologist with IC [150]. Probands with PD were also abnormal vasomotor tone, increased density of sym-
at two-fold or greater risk for mitral valve prolapse, pathetic neurons in their bladder, and increased
migraines, hypercholesterolemia, colitis and ulcers. excretion of NE, all suggestive of autonomic dysfunc-
Paralleling the observations of the case-control study tion [145, 157, 158]. Animal models also show a
of IC, the FDRs of PD probands were also at increased marked decrease in cortisol response to stress in cats
risk for these disorders vis-á-vis those of controls, with feline IC [159]. Patients with mitral valve pro-
regardless of the proband’s status. lapse (MVP) similarly manifest elevated NE levels and
increased vagal tone [160]. Strikingly, MVP in panic
patients tends to be disproportionately of the noncal-
Potential mechanisms cified type, whose etiology has been linked to auto-
Why is it that some medical conditions, but not nomic function [161], and MVP symptoms can be
others, cluster with PD, and what mechanisms could ameliorated by treating with PD medication only
explain the clustering patterns? One possibility is that [162]. Autonomic abnormalities have also been
the medical conditions are themselves stress-provoking, reported in individuals with migraine [163] and
and lead to panic symptoms. Though this may be true migraine pathophysiology involves inappropriate dila-
in some cases, the relative onset ages of the medical tion of cerebral blood vessels, an action which is under
conditions vis-á-vis PD make it an unlikely primary autonomic control [164]. Similar links have been
explanation: PD onsets in early adulthood whereas IC reported for gastrointestinal disorders [165, 166] and
does not typically until well into the forties. An alter- hypercholesterolemia [167, 168].
nate explanation may stem from the observation that Collectively, our observations are consistent with
a number of the clustering medical conditions involve a pleiotropic model in which vulnerability conferred
smooth muscle functions, suggesting possible dysre- by a genetic change (possibly on chromosome 13q)
gulation of the autonomic nervous system (ANS). PD gives rise to dysfunction in multiple autonomically
results in part from a failure to regulate the ANS, regulated sites, resulting in the host of seemingly
which results in increased reactivity and an excess of disparate clinical phenotypes. The previously identi-
NE release from the brainstem, and particularly the fied region of chromosome 13q encompasses, among
locus coeruleus (LC) (which comprises more than others, genes encoding the endothelin B receptor
70% of all NE projections to the forebrain) [151– (a promising candidate, related to vasoconstriction),
153]. In asymptomatic individuals, the NE activity the serotonin 2A (HTR2A) receptor, and multiple
in the LC is cross-regulated by serotonin projections transcription factors [40, 41], and further genotyping
from the raphe nuclei and by neuromodulators such is currently underway to more specifically identify
as CRF, maintaining a homeostatic balance [154]. the susceptibility locus for this interesting clinical
Breakdown of this regulation results in the increased syndrome.
reactivity and stress associated with many of the anx- The coupling of psychiatric and nonpsychiatric
iety phenotypes. But the same locus coeruleus–nore- symptoms we found is hardly novel, having been
pinepherine (LC-NE) system also activates the previously reported for disorders such as Huntington’s
hypothalamus–pituitary–adrenal axis (HPA) and the disease [169], Wolfram’s syndrome [170], and rheuma-
sympathetic branches of the ANS , which subsequently toid fever [171, 172]. But few psychiatric genetic studies

101
Chapter 9: Panic disorder

have exploited the links between psychiatric and within the amygdala, hippocampus, thalamus, and
nonpsychiatric manifestations to explore underlying midbrain of PD patients [184].
mechanisms, instead typically relegating the latter Although an exhaustive neuroanatomical review is
to comorbidities or confounders [173]. Contrast beyond the scope of this chapter, it bears note these
this with diabetes research, for example, where the regional findings are consistent with the proposed
study of glucose metabolism has informed our under- cortical circuitry of fear and anxiety. The anterior
standing of the associated neuropathy, retinopathy, cingulate is centrally involved in the processing of
and nephropathy [174]. Our observations support emotional (especially negatively valenced) stimuli,
using “expanded spectrum” approaches wherein psy- and volume reduction may diminish the ability to
chiatric phenotypes can be refined or morphed with ascribe the appropriate emotional weight to incoming
other associated clinical symptoms in ways that may information, leading to even neutral stimuli being
provide more tractable substrates at the biological level. perceived as threatening. The anterior cingulate
cortex also downregulates sub-cortical structures such
Neuroimaging studies of PD as the amygdala that generate the fear response;
reduced volume may therefore reflect a decoupling
In the course of the last decade, neuroimaging has
of this regulation. The insula is additionally involved
emerged as a potent tool via which to address the
in integration of afferent and efferent cortical connec-
machinery of the brain and the mind. Understanding
tions to the brain, and PD patients have reduced
the neural circuitry of normal brain functioning
GABA binding here [185], suggesting that inhibition
allows us to predict and then test how breakdown
by the insula may protect against PD. On the other
of that circuitry might lead to psychopathological
hand, the increase in midbrain nuclei is consistent
phenotypes of interest. In this section we discuss some
with their role in unconditioned fear responses [186].
of the recent advancements in neuroimaging of PD.
Furthermore, the reduced serotonin transporter bind-
Although we review both structural and imaging stud-
ing found in the midbrain of PD patients [52], and the
ies, we especially focus on imaging-genetics approaches,
observation that treated PD patients show response-
as integration of genetics with neuroimaging data
related changes in glucose binding [187] allude to a
may ultimately allow us to most comprehensively
functional role for a lack of serotonin inhibition
probe the underlying relationship between genes and
within this region.
psychopathology.

Structural studies Functional studies


Structural imaging studies have identified changes in There are few functional imaging studies of PD.
a number of cortical and sub-cortical brain regions of Most studies have relied on experimental proxies
PD patients, including volume reductions in the right for the panic phenotype, as directly imaging panic
anterior cingulate cortex [175, 176], bilateral amyg- attacks is, for a variety of ethical and logistic reasons,
dala [177, 178], putamen [179], left hippocampus unfeasible. There is a case report of a 26-year-old
[178] and parahippocampal gyrus [180], and volu- woman who experienced a spontaneous attack while
metric increases within the rostral pons [181] and left being scanned. Her scan shows increased amygdala
insula [176]. Both decreases [178, 182, 183] and firing during the time period corresponding to the
increases [176] of gray-matter volume have been attack [188] but because this was a single scan, it is
reported within the temporal cortex, and temporal difficult to dissect causal from compensatory or medi-
lobe abnormalities have been associated with earlier cation-related (she was on an SSRI) elements. In an
age of onset and higher rates of panic attacks [183]. unrelated PET study, another healthy female subject
Oddly, in the hippocampus, later onset has been asso- experienced an unexpected panic attack in the scanner
ciated with greater volumetric reduction [178], while being subject to electric shocks as part of a fear-
though this has not been replicated. These findings conditioning paradigm. Here, the panic attack was
are summarized in Table 9.2 [175–183]. Consistent with associated with reduced blood flow in right orbito-
the structural studies, resting state positron emission frontal, prelimbic, and anterior cingulate regions
tomography (PET) studies have also reported higher but no changes in the amygdala [189]. Induction of
glucose uptake (indicative of greater regional activity) panic attacks among healthy volunteers using sodium

102
Chapter 9: Panic disorder

Table 9.2 Structural brain changes associated with PD.

Anatomical Source Sample Hemisphere Effect (PD versus controls)


region (Cases/
Controls)
Amygdala Massana et al., 2003 12/12 L and R Decrease in volume
Uchida et al., 2003 11/11 L and R Decrease in volume
Anterior cingulate Asami et al., 2008 26/26 R Decrease in volume
Uchida et al., 2008 19/20 R Decrease in volume
Hippocampus Uchida et al., 2003 11/11 L Decrease in volume
Parahippocampal Massana et al., 2003 18/18 L Decreased gray matter intensity
gyrus
Insula Uchida et al., 2008 19/20 L Increase in volume
Midbrain/pons Protopopescu et al., 10/23 L Increase in volume
2006
Uchida et al., 2008 19/20 L Increase in volume
Putamen Yoo et al., 2005 18/18 L and R Decrease in volume
Temporal lobe Fontaine et al., 1990 31/20 R Decrease in volume
Ontiveros et al., 1989 30/20 R Greater MRI abnormalities, correlated with
younger age of onset
Uchida et al., 2003 11/11 L and R Decrease in volume; preferentially in the
left hemisphere
Uchida et al., 2008 19/20 L Increase in volume
Abbreviations: L, left; MRI, magnetic resonance imaging; PD, panic disorder; R, right.

lactate has been reported to increase activation in the of the scan will also vary across patients, and the
insula, temporal lobe and the periacqueductal gray scanner can itself be a panicogenic agent. Current
region (PAG) [190], whereas induction by CCK-4 symptoms can greatly influence task performance
reported increased activity within hypothalamus, and outcome; therefore, even if the target is under-
insula, and PAG [191]. The activation of the PAG is lying trait differences, it is imperative that state
important as this region plays a critical role in the differences be accounted for. One approach is via
processing of unconditioned fear, a hallmark of panic quantitative symptom measures such as the Spielber-
attacks [192]. But because these are single studies ger State Trait Anxiety Inventory (STAI), Acute Panic
or case reports, the findings should be interpreted Inventory, or the Hamilton Rating Scale for Anxiety
cautiously. Additionally, data from panic induction (HRSA), which can be administered at the time of
studies should not be overly generalized, as experi- the scan (ideally at multiple points), and can then be
mentally induced panic attacks do not evoke all of the subsequently modeled into the statistical analysis.
symptoms of the disorder with equal reliability and The most widely used functional paradigms to
may represent a more restrictive or etiologically study anxiety disorders involve testing brain
biased phenotype [193]. responses to fear, based on the idea that anxiety states
Most functional imaging studies of PD therefore may arise from a failure of the brain to regulate fear
do not target “active” panic symptoms, but rather, processing. The model of anxiety disorders as an
underlying trait differences that distinguish those exaggerated extension of the natural fear system is
with and without the disorder. This of course is an attractive one to study, as the neural circuitry of
contingent upon the differences of interest being fear has been well worked out in both animal models
detectable in the scanner by appropriately targeted and humans [194]. Most of the proposed pathways
experimental paradigms, even if the subject is currently have implicated a top-down control mechanism,
asymptomatic. But current symptom levels at the time wherein various frontal structures downregulate

103
Chapter 9: Panic disorder

activity sub-cortical structures (notably the amyg-


dala), resulting in the generation and extinguishing Run
cycle of the fear response [195–197]. Anxiety disorders
may result from a breakdown of this regulation,
with the precise mechanisms varying by phenotype.
15s “on” 15s “off”
A recent meta-analysis [197] reported that the specific
and social phobias, where exaggerated fear was the
Epoch
primary feature, invoked greater activation in the
amygdala and insula in response to viewing negatively
valenced emotional stimuli; PTSD, in contrast, wherein
fear was part of a larger emotional disturbance, showed Event
reduced activation within cingulate cortex. In PD 200 ms 400 ms 200 ms 1200 ms
patients, symptom-related reductions in cingulate activ- Nonmasked:
ity have been reported while viewing fearful faces [198] Fearful (F)
or
(as well as a converse increases while viewing happy Neutral (N)
faces [199]); however, the above meta-analysis did
or
not consider PD owing to an overall paucity of studies. 200 ms 400 ms 33 ms 167 ms 1200 ms
Masked:
We currently have an ongoing study to test where Masked
PD falls along this spectrum. Subjects with PD, along Fearful (FN)
or
with a comparison group of those with social anxiety Masked
disorder (SAD) and a healthy control group, are Neutral (NN)

scanned while performing experimental paradigms Figure 9.5 Experimental paradigm to test brain responses to
that target facets of emotional processing. In the first masked versus unmasked fearful stimuli. Stimuli were faces with
either fearful (F) or neutral (N) expressions, which were pseudo-
paradigm, illustrated in Figure 9.5 [200], subjects are colored in red, yellow, or blue. The faces were presented in either
imaged while passively viewing masked (uncon- masked or unmasked conditions. For the unmasked condition, the
sciously perceived) or unmasked (consciously per- F or N face was shown for the full stimulus duration (200 ms).
For the masked conditions (FN and NN), a fearful or neutral face
ceived) fearful faces. (The masked conditions are was presented for the first 33 ms, followed by a neutral face of
generated by presenting a fearful face first briefly the same gender and color for the remaining 167 ms. Activity
[in this paradigm, 33 ms], followed by a matched attributable to unconscious fear processing was calculated as
[activity masked fearful] – [activity masked neutral]); conscious fear
neutral face for the remaining duration of the stimu- processing was [activity unmasked fearful] – [activity unmasked neutral]).
lus [167 ms], such that the fearful face is not con- (From [200], with permission from Elsevier.)
sciously perceived by the subject.) Brain scans on
healthy volunteers have found that the conscious compare the processing of consciously and uncon-
and unconscious fear processing activate different sciously perceived fear and address this question.
subregions of the amygdala [200] (Figure 9.6): uncon-
scious fear (experimentally speaking, [activity masked
fearful] – [activity masked neutral]) activates the basolateral
Imaging genetics
complex, whereas conscious fear [activity unmasked fear- We conclude with an overview of imaging genetic
ful] – [activity unmasked neutral]) activates the dorsal studies of PD, and particularly, of how data from
region. Importantly, unconscious, but not conscious, imaging studies can be used to enhance the tractabil-
processing in the amygdala correlates with trait anx- ity of genetic targets. Brain imaging, and particularly
iety. This makes sense behaviorally: when faced with functional imaging, can provide invaluable insights to
unknown or unrecognized threats, subjects are more genetic studies, as it allows identification of specific
likely to respond based on their pre-existing individual neural mechanisms that may mediate the effects of
proclivities to anxiety (trait). But when the context genes on psychiatric disorders. Additionally, because
of the threat is clarified (unmasked), these differences variation in brain structure or function may be more
converge as subjects respond more uniformly. The above objectively indexed to the underlying genetic architec-
data raise the question of whether (and if so, which?) ture than behavioral outcomes, neural phenotypes
anxiety phenotypes are disorders of extreme uncon- may provide more penetrant targets through which
scious emotional vigilance. Our study allows us to to identify genes related to clinical psychopathology.

104
Chapter 9: Panic disorder

connectivity resulting from the low functioning vari-


ant of the MAOA gene has been found to predict
propensity to violent behaviors, harm avoidance,
and reward dependence [204–206], demonstrating
that genetic predispositions can indeed impact behav-
ioral outcomes via changes in brain functioning.
But can such models predict psychopathology?
Nonmasked Masked fear
fear (F-N) (FN-NN) vs. STAI-T The first study [81] to apply such an approach to
PD focused on the COMT gene, and in particular
the aforementioned “val158met” polymorphism. This
polymorphism is relevant not only because it func-
tionally reduces COMT levels [206] but it also reduces
cortical activity during emotional processing tasks
among healthy volunteers [207, 208]. The investiga-
tors tested whether the association between COMT
and PD was mediated via these changes in emotional
T processing. Twenty PD patients were scanned while
value being presented with a series of standardized faces
6
5
5
[209] bearing fearful, angry, happy, or neutral expres-
4
4 sions, and their DNA was genotyped on the val158met
3 3 polymorphism. The investigators reported that
2 2 presentation of fearful faces resulted in an increased
1 1
0 0
activation in the right amygdala as well as the left
orbitofrontal cortex – but this was only observed
among patients carrying at least one val encoding
Figure 9.6 Amygdala activity related to masked and unmasked
fear processing in healthy subjects. Enlarged views of the right allele; patients with both copies of the met encoding
amygdala illustrating: (1) the dorsal amygdalar cluster from the allele did not show these changes (Figure 9.7). Angry
nonmasked fear (F-N) comparison (coronal view at Y ¼ 8 [A] and or happy faces activated the ventromedial cortex but
axial view at Z ¼ 16 [B]); and (2) the basolateral amygdalar cluster
from the correlation of masked fear-induced activity (FN-NN) with not either the amygdala or the orbitofrontal cortex.
trait anxiety (coronal view at Y ¼ 8 [C] and axial view at Z ¼ 28 The authors speculate that the excessive orbitofrontal
[D]). The color bar indicates the significance, with lighter colors activation in their study might reflect a failure to
indicating a greater difference between the respective fearful and
neutral conditions. (From [200], with permission from Elsevier.) appropriately inhibit the processing of anxiety related
See plate section for color version. emotional cues by the amygdala, and that this may be
an outcome of reduced dopamine levels conferred by
In other words, imaging can help both in identifying the val allele (val allele ! higher COMT levels !
genetic risk and in exploring the mechanisms of that more breakdown of dopamine ! lower cortical dopa-
genetic risk [201]. Such approaches have already been mine). This explanation parallels data from animal
effectively applied to healthy subjects. In an elegant models showing dopamine loss in the prefrontal
and pioneering study, Hariri et al. [202] found that cortex to result in delayed fear extinction [210].
individuals with one or more short alleles of the The investigators then extended this strategy to the
serotonin transporter had greater amygdala activation serotonergic system, focusing on both the promoter
in response to viewing of fearful faces. Following polymorphism within the transporter (5-HTTLPR) as
up on these results, the same group [203] found that well as a polymorphism involving a cytosine to guanine
the regulatory circuit between the amygdala and the switch at position 1019 of the serotonin receptor 1A,
cingulate cortex was uncoupled among individuals that had been previously shown to result in lower
with the short allele of the serotonin transporter. serotonin transmission [211] and had been associated
Furthermore, this uncoupling accounted for a third with anxiety-related traits in healthy volunteers [212].
of variation in temperamental anxiety, suggesting that They found that patients homozygotes for the g alleles
anxiety disorders may be an extreme form of this in the 1A receptor had reduced activation in orbito-
decoupling. Subsequently, variation in cortical frontal, ventromedial prefrontal, and cingulate cortices

105
Chapter 9: Panic disorder

Figure 9.7 Amygdala activation in


response to fearful faces is modulated by
catechol-O-methyltransferase (COMT)
genotype. Whole brain voxel-wise t-map
(uncorrected, P > 0.001) of a single panic
disorder patient with a genotype 472G/A
(i.e. val-met heterozygous) comparing
activation for the fearful versus the
no-face condition. Note that there is
significantly greater amygdala activation
(center of cross hair) in the fearful
condition. This increase was not observed
among patients homozygous for the met
encoding allele. The transverse plane is
the original one; the coronal and sagittal
planes are planar reconstructions
orthogonal to the original image.
Reader’s right is subject’s right. Emotional
face stimuli and no-face control stimuli
were controlled for dynamics and
luminance. (From [81], with permission
from Cambridge University Press.)
See plate section for color version.

when viewing unmasked fearful faces, as compared to ability to zero in on the most salient candidates.
patients with at least one c allele (Figure 9.8) [213]. Another limitation is the lack of GWAS, which
These observed reductions were also largely within are particularly valuable for candidate gene identifi-
the right hemisphere, consistent with the role of that cation as they allow probing of entire genomes at
hemisphere in processing emotional faces [214]. The relatively high resolution without requiring a priori
authors postulate that the genetic variation within the hypotheses [215]. GWAS are opening doors in genetic
receptor may increase the odds of psychopathology by research on major depressive disorder, autism,
impairing processing within specific cortical sites schizophrenia, and bipolar disorder; yet, they have
related to emotion and anxiety, including the orbito- been largely overlooked for most anxiety phenotypes,
frontal and cingulate cortices. Even though the samples including PD.
in these two studies were relatively small and bereft of a In the course of this chapter, we touched upon a
true control group of persons without PD, the findings number of approaches aimed at refining the panic
are important as they are the first to submit that that phenotype to increase its tractability. We saw that
genetic variation may alter vulnerability to PD via early onset substantially magnified familial loading
specific, behaviorally relevant, brain pathways. of PD, and may therefore represent a particularly
valuable subtype. Requiring cases with family history
of anxiety may similarly increase the genetic loading
Conclusions while minimizing sporadic cases (by corollary, con-
There is ample evidence that PD runs in families and trols should be selected to be free not just of personal
is heritable; yet the search for specific genes under- but of family history as well). Other fine-tuning could
lying the disorder has only been partially informative. include restricting to cases with AG, those showing
A number of factors may contribute to this: Many of CO2 hypersensitivity, or those lying within extreme
the individual studies have been jeopardized by rather ranges of quantitative anxiety measures such as neur-
small sample sizes, and the substantial phenotypic oticism. Finally, operationalization need not be
heterogeneity across studies has further impeded the restricted to the psychiatric domain, as illustrated by

106
Chapter 9: Panic disorder

further as gene products regularly cross-talk. For


example, successful treatment of PD patients with
the SSRI fluoxetine, increases plasma levels not only
of serotonin metabolites, but of NE too [154]. Finally,
the detection of certain genetic effects may
be conditional on other genes or environmental
factors, as so elegantly exemplified by the work
of Caspi and colleagues [44] who found a dose-
dependent association between the serotonin trans-
porter polymorphism 5-HTTLPR and depression,
but only among offspring subjected to stressful life
events. There were no main effects, and had the
authors not tested for interaction, they would have
obtained null findings. Yet few of the PD studies
have searched for gene-by-gene, and even fewer,
for gene-by-environment, effects.
Given these layers of complexity, studies relying
on 1 : 1 gene-to-disorder approaches will likely be
of increasingly limited utility. On the one hand, as
Figure 9.8 Serotonin 1A receptor genotype moderates fear is already clear, genetic findings need to be coupled
processing in panic disorder patients. Random effects statistical with epistatic, epigenetic, gene expression, and
parametric map for the fearful versus neutral faces contrast overlaid translational studies in order to probe the func-
on a three-dimensional canonical Montreal Neurological Institute
brain showing right-lateralized activity differences in the prefrontal tional consequences of the genetic variation. (Many
cortex between the two patient groups (5HT1A1019GG versus of the variants identified thus far lie within introns or
CC/CG; p < 0.001, uncorrected). Patients homozygotes for the do not alter expression, so their significance remains
g alleles had reduced activation in orbitofrontal, ventromedial
prefrontal, and cingulate cortices when viewing unmasked fearful unclear.) But the amassing genetic data will also need
faces, as compared to patients with the CG or CC genotype. to be coupled with other downstream functional
(From [213], with permission from Cambridge University Press.) phenotypes (such as the brain structure and function
See plate section for color version.
measures provided by neuroimaging techniques)
our observations that transmission was higher among that can help shed light on how the genes and their
families in which the panic clustered with other blad- products translate into the psychiatric outcomes of
der and cardiovascular problems. interest.
But these considerations aside, some of the diffi- As the field moves toward DSM5, the classifica-
culty in uncovering genetic associations may also lie tion of psychiatric disorders also appears to be evolv-
in the very complexity of the disorder and in the ing toward a more etiological approach integrating
expectation implicit in many study designs that there knowledge from genes, environments, and basic
is a direct or linear relationship between gene and neuroscience [217]. Functional imaging can play an
disease. Complete penetrance is rare, and anxiety important guiding role in this evolution: by identify-
disorders, at least, more likely arise from complex ing specific cortical pathways relevant to fear and
interplay between multiple genetic and environmen- anxiety, we can generate more targeted neurobio-
tal variables, each with only a marginal role [201, logical markers which can then be used both as a tool
216]. Any given genetic variation is unlikely to be to identify novel genetic variation, and to study the
either specific or sufficient, and in the case of neuro- neurobiological pathways through which genes might
transmitter-encoding genes, specificity is muddied ultimately influence psychopathology.

References IV). Washington, DC: American


Psychiatric Association; 1994.
3. Kessler RC, et al. Arch Gen
Psychiatry 2005;62(6):617–627.
1. American Psychiatric Association.
Diagnostic and Statistical Manual of 2. Kessler RC, et al. Arch Gen 4. Weissman MM, et al. Arch Gen
Mental Disorders, Version IV (DSM- Psychiatry 2005;62(6):593–602. Psychiatry 1997;54(4):305–309.

107
Chapter 9: Panic disorder

5. Kessler RC, et al. Arch Gen 28. Torgersen S. Arch Gen Psychiatry 50. Perna G, et al. Biol Psychiatry
Psychiatry 2006;63(4):415–424. 1983;40(10):1085–1089. 1995;37(8):528–532.
6. Kessler RC, et al. Arch Gen 29. Bellodi L, et al. Am J Psychiatry 51. Price WA, et al. Psychosomatics
Psychiatry 1998;55(9):801–808. 1998;155(9):1184–1188. 1988;29(4):433–435.
7. Demiryoguran NS, et al. Emerg 30. Kendler KS, et al. Arch Gen 52. Maron E, et al. Psychiatry Res
Med J 2006;23(2):99–102. Psychiatry 2003;60(9):929–937. 2004;132(2):173–181.
8. Marchesi C, et al. Emerg Med J 31. Hettema JM, et al. Arch 53. Olivier B, et al. Eur J Pharmacol
2004;21(2):175–179. Gen Psychiatry 2005;62(2): 2005;526(1–3):207–217.
9. Marchesi C, et al. Int J Psychiatry 182–189. 54. Klemenhagen KC, et al.
Med 2001;31(3):265–275. 32. Crowe RR, et al. Am J Med Genet Neuropsychopharmacology
2001;105(1):105–109. 2006;31(1):101–111.
10. Goisman RM, et al. Depress
Anxiety 1998;7(3):105–112. 33. Gelernter J, et al. Am J Med Genet 55. Nash JR, et al. Br J Psychiatry
2001;105(6):548–557. 2008;193(3):229–234.
11. Zimmermann P, et al. Psychol
Med 2003;33(7):1211–1222. 34. Fyer AJ, et al. Biol Psychiatry 56. Neumeister A, et al. J Neurosci
2006;60(4):388–401. 2004;24(3):589–591.
12. Bienvenu OJ, et al. Br J Psychiatry
2006;188:432–438. 35. Kaabi B, et al. Am J Hum Genet 57. Rothe C, et al. Int J
2006;78(4):543–553. Neuropsychopharmacol 2004;
13. Fyer AJ, et al. Anxiety 1996; 7(2):189–192.
2(4):173–178. 36. Knowles JA, et al. Am J Med Genet
1998;81(2):139–147. 58. Huang YY, et al. Int J
14. Goldstein RB, et al. Arch Gen Neuropsychopharmacol 2004;
Psychiatry 1994;51(5):383–394. 37. Logue MW, et al. Am J Med 7(4):441–451.
15. Horwath E, et al. Arch Gen Genet B Neuropsychiatr Genet
2003;121B(1):95–99. 59. Weisstaub NV, et al. Science
Psychiatry 1995;52(7):574–582. 2006;313(5786):536–540.
16. Maier W, et al. J Psychiatr Res 38. Thorgeirsson TE, et al. Am J Hum
Genet 2003;72(5):1221–1230. 60. Fehr C, et al. Psychiatry Res
1995;29(5):375–388. 2000;97(1):1–10.
17. Noyes R, Jr., et al. Arch Gen 39. Smoller JW, et al. Am J Med Genet
2001;105(2):195–206. 61. Maron E, et al. Psychiatr Genet
Psychiatry 1986;43(3):227–232. 2005;15(1):17–24.
18. Smoller JW, et al. Am J Med 40. Hamilton SP, et al. Proc Natl
Acad Sci U S A 2003;100(5): 62. Inada Y, et al. Psychiatry Res
Genet C Semin Med Genet 2003;118(1):25–31.
2008;148(2):118–126. 2550–2555.
41. Weissman MM, et al. Am J Med 63. Rothe C, et al. Neurosci Lett
19. Weissman MM, et al. Arch Gen 2004;363(3):276–279.
Psychiatry 1993;50(10):767–780. Genet 2000;96(1):24–35.
42. Kent JM, et al. Biol Psychiatry 64. Yoon HK, et al. J Anxiety
20. Battaglia M, et al. J Psychiatr Res Disord 2008;22(8):
1999;33(1):37–39. 1998;44(9):812–824.
1529–1534.
21. Haghighi F, et al. Am J Med Genet 43. Nutt DJ. J Clin Psychiatry 1998;
65. Jenck F, et al. Eur Neuro-
1999;88(2):131–135. 59(S8):24–28; discussion 29.
psychopharmacol 1998;8(3):
22. Goldstein RB, et al. Arch 44. Caspi A, et al. Science 2003; 161–168.
Gen Psychiatry 1997;54(3): 301(5631):386–389.
66. Ribases M, et al. J Psychiatr Res
271–278. 45. Heils A, et al. J Neural Transm 2008;42(1):50–57.
23. Hettema JM, et al. Am J Psychiatry 1997;104(10):1005–1014.
67. Deckert J, et al. Int J
2001;158(10):1568–1578. 46. Lesch KP, et al. Science 1996; Neuropsychopharmacol 2000;
24. Kendler KS, et al. Psychol Med 274(5292):1527–1531. 3(4):321–325.
1993;23(2):397–406. 47. Blaya C, et al. Behav Brain Funct 68. Maron E, et al. Int J
25. Perna G, et al. Psychiatry Res 2007;3:41. Neuropsychopharmacol 2005;
1997;66(1):69–71. 48. Strug LJ, et al. Mol Psychiatry 8(2):261–266.
26. Scherrer JF, et al. J Affect Disord 2010;15(2):166–176. 69. Mizuta N, et al. Psychiatr Genet
2000;57(1–3):25–35. 49. Perna G, et al. Neuropsycho- 2008;18(1):44.
27. Skre I, et al. Acta Psychiatr Scand pharmacology 2005;30(12): 70. Sabol SZ, et al. Hum Genet
1993;88(2):85–92. 2230–2235. 1998;103(3):273–279.

108
Chapter 9: Panic disorder

71. Deckert J, et al. Hum Mol Genet 92. Freitag CM, et al. Psychiatr Genet 113. Keck ME, et al. Am J Med Genet B
1999;8(4):621–624. 2006;16(2):59–65. Neuropsychiatr Genet 2008;
72. Samochowiec J, et al. Psychiatry 93. Gratacos M, et al. Genes Brain 147B(7):1196–1204.
Res 2004;128(1):21–26. Behav 2007;6(S1):2–23. 114. Griebel G, et al. Curr Drug
73. Hamilton SP, et al. Mol Psychiatry Targets CNS Neurol Disord 2003;
94. Hamilton SP, et al. Am J Med
2000;5(5):465–466. 2(3):191–200.
Genet 2000;96(3):324–330.
74. Tadic A, et al. Am J Med 115. Feusner J, et al. Psychiatry Res
95. Sebastiao AM, et al. Trends
Genet B Neuropsychiatr Genet 2001;104(2):109–117.
Pharmacol Sci 2000;21(9):341–346.
2003;117B(1):1–6. 116. Kaiya H, et al. Psychiatry Res
96. Charney DS, et al. Arch Gen
75. Fehr C, et al. Prog Neuro- 1982;6(3):335–343.
Psychiatry 1985;42(3):233–243.
psychopharmacol Biol Psychiatry 117. Smoller JW, et al. Am J Med Genet
2001;25(5):965–982. 97. Moreau JL, et al. Brain Res Brain 2001;105(3):226–235.
Res Rev 1999;31(1):65–82.
76. Kim W, et al. Prog Neuro- 118. Hettema JM, et al. Mol Psychiatry
psychopharmacol Biol Psychiatry 98. Deckert J, et al. Mol Psychiatry 2006;11(8):752–762.
2006;30(8):1413–1418. 1998;3(1):81–85.
119. Jedema HP, et al. Eur J Neurosci
77. Maron E, et al. Neurosci Lett 99. Hamilton SP, et al. Neuro- 2008;27(9):2433–2443.
2007;411(3):180–184. psychopharmacology 2004;
29(3):558–565. 120. Leygraf A, et al. J Neural Transm
78. Unschuld PG, et al. Am J Med 2006;113(12):1921–1925.
Genet B Neuropsychiatr Genet 100. Domschke K, et al. Am J Med
Genet B Neuropsychiatr Genet 121. Hohoff C, et al. J Neural Transm
2007;144B(4):424–429. 2009;116(111):1523–1528.
2008;147B(4):510–516.
79. Zhang X, et al. Science 2004; 122. Klein DF. Arch Gen Psychiatry
305(5681):217. 101. Hodges LM, et al. Am J Med
Genet B Neuropsychiatr Genet 1993;50(4):306–317.
80. Domschke K, et al. Am J Med 2008;147B(8):1476–1480. 123. Coryell W, et al. J Affect Disord
Genet B Neuropsychiatr Genet 2006;92(1):63–70.
2007;144B(5):667–673. 102. Charney DS. Acta Psychiatr Scand
Suppl 2003(417):38–50. 124. van Beek N, et al. Biol Psychiatry
81. Domschke K, et al. Int J 2000;47(9):830–835.
Neuropsychopharmacol 2004; 103. Hamilton SP, et al. Am J Med
7(2):183–188. Genet B Neuropsychiatr Genet 125. Battaglia M, et al. Am J Med
2004;126B(1):111–115. Genet B Neuropsychiatr Genet
82. Hamilton SP, et al. Biol Psychiatry 2008;147B(5):586–593.
2002;51(7):591–601. 104. Ebihara M, et al. Am J Med
Genet B Neuropsychiatr Genet 126. Cavallini MC, et al. Biol Psychiatry
83. Rothe C, et al. Neuro- 1999;46(6):815–820.
2003;118B(1):32–35.
psychopharmacology 2006;
31(10):2237–2242. 105. Hattori E, et al. Mol Psychiatry 127. Bradley SR, et al. Nat Neurosci
2001;6(4):465–470. 2002;5(5):401–402.
84. Woo JM, et al. J Psychiatr Res
2004;38(4):365–370. 106. Hosing VG, et al. J Neural Transm 128. Severson CA, et al. Nat Neurosci
Suppl 2004(68):147–156. 2003;6(11):1139–1140.
85. Woo JM, et al. Am J Psychiatry
107. Ise K, et al. Am J Med 129. Schmidt NB, et al. J Abnorm
2002;159(10):1785–1787.
Genet B Neuropsychiatr Genet Psychol 2000;109(2):308–320.
86. Ohara K, et al. Psychiatry Res 2003;118B(1):29–31. 130. Perna G, et al. Am J Med Genet B
1998;80(2):145–148.
108. Kennedy JL, et al. Mol Psychiatry Neuropsychiatr Genet 2004;
87. Zintzaras E, et al. Psychiatr Genet 1999;4(3):284–285. 129B(1):41–43.
2007;17(5):267–273.
109. Miyasaka K, et al. Am J Med 131. Philibert RA, et al. Am J Med
88. Alsobrook JP, et al. Am J Med Genet B Neuropsychiatr Genet Genet B Neuropsychiatr Genet
Genet 2002;114(1):116–120. 2004;127B(1):78–80. 2003;117B(1):11–17.
89. Lee SG, et al. Hum Genet 2005; 110. Hamilton SP, et al. Mol Psychiatry 132. Battaglia M. Mol Psychiatry 2002;
116(4):319–328. 2001;6(1):59–65. 7(3):239–246.
90. Bourdeau V, et al. Mol Endocrinol 111. Hattori E, et al. Am J Med Genet 133. Chanock SJ, et al. Nature 2007;
2004;18(6):1411–1427. 2001;105(8):779–782. 447(7145):655–660.
91. Liehr JG, et al. Free Radic Biol Med 112. Kato T, et al. Am J Med Genet 134. Weissman MM, et al. Arch Gen
1990;8(4):415–423. 1996;67(4):401–405. Psychiatry 2000;57(7):675–682.

109
Chapter 9: Panic disorder

135. Maxwell ME. Manual for the 158. Stein PC, et al. Urology 1999; 182. Fontaine R, et al. Biol Psychiatry
FIGS. Clinical Neurogenetics 53(6):1140–1143. 1990;27(3):304–310.
Branch, Intramural Research 159. Westropp JL, et al. J Urol 2003; 183. Ontiveros A, et al. J
Program. National Institute of 170(6 Pt 1):2494–2497. Neuropsychiatry Clin Neurosci
Mental Health 1992. 1989;1(4):404–408.
160. Pasternac A, et al. Am J Med
136. Talati A, et al. Mol Psychiatry 1982;73(6):783–790. 184. Sakai Y, et al. Neuroreport 2005;
2008;13(2):122–130. 16(9):927–931.
161. Gorman RC, et al. J Thorac
137. Fyer AJ, et al. Am J Med Genet Cardiovasc Surg 1995;109(4): 185. Cameron OG, et al. Arch
1999;88(2):173–181. 684–693. Gen Psychiatry 2007;64(7):
138. Weissman MM, et al. Arch Gen 162. Coplan JD, et al. Am J Psychiatry 793–800.
Psychiatry 2004;61(3):273–279. 1992;149(11):1587–1588. 186. Coplan JD, et al. Biol Psychiatry
139. Hamilton SP, et al. PNAS 2003; 163. Zigelman M, et al. Headache 1998;44(12):1264–1276.
5(100):2550–2555. 1994;34(10):569–577. 187. Sakai Y, et al. Neuroimage 2006;
140. Gorman JM, et al. Psychosom Med 164. Welch KM. Semin Neurol 1997; 33(1):218–226.
1988;50(2):114–122. 17(4):335–341. 188. Pfleiderer B, et al. World J Biol
141. Lydiard RB, et al. Am J Psychiatry 165. Goddard E, et al. Primary Psychiatry 2007;8(4):269–272.
1994;151(1):64–70. Psychiatry 2007;14(4):69–73. 189. Fischer H, et al. Neurosci Lett
142. Placidi GP, et al. Neuropsychobiology 166. Drossman DA, et al. Gut 1999; 1998;251(2):137–140.
1998;38(4):222–225. 45 S2:1125–1130. 190. Reiman EM. J Clin Psychiatry
143. Stewart W, et al. Neurology 1994; 167. Agargun MY, et al. Can J 1997;58(S16):4–12.
44(10/S7):S23–S27. Psychiatry 2004;49(11):776–778. 191. Javanmard M, et al. Biol
144. Hanno PM. Urol Clin North Am 168. Bajwa WK, et al. Am J Psychiatry Psychiatry 1999;45(7):872–882.
1994;21(1):63–66. 1992;149(3):376–378. 192. Graeff FG. Neurosci Biobehav Rev
145. Curhan GC, et al. J Urol 1999; 169. Wexler NS, et al. Nature 1987; 2004;28(3):239–259.
161(2):549–552. 326(6109):194–197. 193. Perna G, et al. Psychiatry Res
146. Buffington CA. J Urol 2004; 170. Swift RG, et al. Mol Psychiatry 1994;52(2):159–171.
172(4 Pt 1):1242–1248. 1998;3(1):86–91. 194. LeDoux JE. Annu Rev Neurosci
147. Alagiri M, et al. Urology 1997; 171. Herdy GV, et al. Braz J Med Biol 2000;23:155–184.
49(5A Suppl):52–57. Res 1992;25(8):789–794. 195. Bush G, et al. Trends Cogn Sci
148. Warren JW, et al. Urology 2001; 172. Swedo SE, et al. Am J Psychiatry 2000;4(6):215–222.
57(6 Suppl 1):22–25. 1997;154(1):110–112. 196. Etkin A, et al. Neuron 2006;51(6):
149. Zondervan KT, et al. Behav Genet 173. Frank ECG, et al. CNS Spectrums 871–882.
2005;35(2):177–188. 1998;3(4):23–34. 197. Etkin A, et al. Am J Psychiatry
150. Talati A, et al. Biol Psychiatry 174. Mahtani MM, et al. Nat Genet 2007;164(10):1476–1488.
2008;63(6):594–601. 1996;14(1):90–94. 198. Pillay SS, et al. J Affect Disord
151. Chrousos GP, et al. JAMA 175. Asami T, et al. Psychiatry Clin 2006;94(1–3):173–181.
1992;267(9):1244–1252. Neurosci 2008;62(3):322–330. 199. Pillay SS, et al. J Anxiety Disord
152. Gorman JM, et al. Am J Psychiatry 176. Uchida RR, et al. Psychiatry Res 2007;21(3):381–393.
2000;157(4):493–505. 2008;163(1):21–29. 200. Etkin A, et al. Neuron 2004;
153. Sullivan GM, et al. Biol Psychiatry 177. Massana G, et al. Neuroimage 44(6):1043–1055.
1999;46(9):1205–1218. 2003;19(1):80–90. 201. Meyer-Lindenberg A, et al.
154. Coplan JD, et al. Arch Gen 178. Uchida RR, et al. Braz J Med Biol Nat Rev Neurosci 2006;7(10):
Psychiatry 1997;54(7):643–648. Res 2003;36(7):925–929. 818–827.
155. Blok BF, et al. Brain 1998;121 179. Yoo HK, et al. Eur J Neurosci 202. Hariri AR, et al. Science 2002;
(Pt 11):2033–2042. 2005;22(8):2089–2094. 297(5580):400–403.
156. Gosling JA, et al. J Urol 1977; 180. Massana G, et al. Am J Psychiatry 203. Pezawas L, et al. Nat Neurosci
118(2):302–305. 2003;160(3):566–568. 2005;8(6):828–834.
157. Hohenfellner M, et al. J Urol 181. Protopopescu X, et al. Neuroreport 204. Buckholtz JW, et al. Mol
1992;147(3):587–591. 2006;17(4):361–363. Psychiatry 2008;13(3):313–324.

110
Chapter 9: Panic disorder

205. Meyer-Lindenberg A, et al. Proc CA, Consulting Psychologists, 213. Domschke K, et al. Int J
Natl Acad Sci U S A 2006; 1976. Neuropsychopharmacol 2006;
103(16):6269–6274. 9(3):349–355.
210. Morrow BA, et al.
206. Chen J, et al. Am J Hum Genet Neuroscience 1999;92(2): 214. Schmitt JJ, et al. Cortex 1997;
2004;75(5):807–821. 553–564. 33(1):65–81.
207. Egan MF, et al. Proc Natl Acad Sci 211. Lemonde S, et al. 215. Pearson TA, et al. Jama 2008;
U S A 2001;98(12):6917–6922. J Neurosci 2003;23(25): 299(11):1335–1344.
208. Smolka MN, et al. J Neurosci 8788–8799. 216. Risch N. Genet Epidemiol 1990;
2005;25(4):836–842. 212. Strobel A, et al. J Neural 7(1):3–16; discussion 17–45.
209. Ekman PFW. Pictures of Transm 2003;110(12): 217. Watson D. J Abnorm Psychol
Facial Affect. Palo Alto, 1445–1453. 2005;114(4):522–536.

111
The genetics of the phobic disorders
Chapter

10 and generalized anxiety disorder


Raymond R. Crowe

With the phobic disorders genetics returns to its or endured with marked distress; in addition, they
origin because Mendel suffered from a phobia of must predictably provoke anxiety, the phobic individ-
disease [1]. His symptoms would satisfy modern diag- ual must recognize that the fear is excessive and
nostic criteria for a phobic disorder because their unreasonable (in adults), and the avoidance or attend-
severity made it impossible for him to attend to ill ant anxiety must interfere with some realm of life.
or dying parishioners. It was for that reason that he Social phobia is further subclassified as generalized
could not be assigned to a parish church so he was social phobia when the fear and anxiety is not limited
assigned instead to the monastery at Altbrun where to a few situations but is caused by many social
he did his remarkable work on heredity. With his encounters. Specific phobia, previously termed “simple
broad-ranging interests and mathematical mind, no phobia”, is characterized by persistent fear of discrete
doubt he would be fascinated to learn of modern objects and situations. Animal phobias are examples
genetic analyses of phobias: that the magnitude of of feared objects, and heights, storms, and medical
genetic and environmental effects can be separated encounters are examples of feared situations. The diag-
and measured; that environmental influences can be nosis requires that, in addition to the phobic fear,
separated into family environment and nonfamilial the object or situation predictably elicit the feared
perturbations; that some of the genetic liability is response, the person recognize that the fear is excessive
shared with other anxiety disorders; and most of all and unreasonable, that the stimulus be avoided or
that attempts are being made to find the genes endured with marked distress, and that the avoidance
responsible for phobias. or phobic anxiety interfere with some realm of the
person’s life. Specific phobia is further subclassified,
depending on the feared stimulus, as animal type,
Diagnosis and classification natural environment type, blood-injection-injury type,
The phobias fall into three major groups with sub- situational type, and other type (not fitting any of the
types within two of these groups. Agoraphobia is previous subtypes).
characterized by phobic anxiety over being in situ-
ations from which escape might be difficult or embar-
rassing or in which help might not be available; for Epidemiology
example, sitting in large audiences. In order to be Unreasonable fears and phobias are quite prevalent in
diagnosed as agoraphobia the feared situations must the general population. A representative population
be either avoided or endured with marked distress. survey using DSM-III-R criteria (Table 10.1) found
Agoraphobia is further classified as occurring with or the lifetime prevalence of social phobia to be 13.3%,
without panic disorder. Social phobia is characterized specific phobia 11.3%, and agoraphobia 5.3%, with
by a fear of embarrassing oneself in social or perform- females affected more frequently in each case [2].
ance situations. For example, persons with social These findings have now been updated with DSM-IV
phobia might be unable to make presentations before criteria and Table 10.1 shows that the results for
groups for fear of fainting. The diagnosis of social the combined-sex totals are essentially unchanged
phobia requires that the feared situations be avoided for social and specific phobia but the prevalence of

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

112
Chapter 10: Genetics of the phobic disorders and GAD

Table 10.1 Lifetime prevalence and age at onset of phobic Table 10.2 Prevalence of fears and phobias in male and
disorders. female twins.

Phobia DSM-III-R DSM-IV Prevalence Male twins Female twins

Male Female Total Fear Phobia Fear Phobia


Agoraphobia 3.5 7.0 5.3 1.4 Agoraphobia 8.6 4.0 12.7 9.1
w/o panic
Animal 28.3 5.2 34.5 10.5
Specific 6.7 15.7 11.3 12.5
Blood-injury 23.3 5.7 20.2 6.6
phobia
Situational 31.2 9.4 30.4 11.3
Social phobia 11.1 15.5 13.3 12.1
Prevalences are expressed in percentages.
Social 32.7 6.3 27.5 14.1
Data from Kessler et al. [2, 3]. Prevalences are expressed as percentages.
Data from Kendler et al. [6, 7].

agoraphobia was only 1.4% [2, 3]. Phobic disorders


develop early in life; the same survey found the The magnitude of the additive genetic variance to the
median age of onset to be 7 years for specific phobia, total variance measures the heritability of a trait.
13 years for social phobia, and 20 years for agorapho- A related question is whether the current classification
bia. Among specific phobias, animal phobia develops of phobias has genetic validity. If all phobias are caused
at an earlier age than situational phobias [4, 5]. Table by the same genetic predisposition but the subclasses are
10.2 shows the relationship of irrational fears to phobic shaped by different environmental influences, then the
disorders in a population-based sample of twins [6, 7]. phobias as a group have genetic validity but the sub-
In both sexes irrational fears are substantially more classes do not. This has obvious implications for phe-
prevalent than the corresponding phobia, the ratio notyping subjects for gene searches. Understanding the
being more pronounced in social phobia and the spe- genetic basis for sex differences in prevalence rates of the
cific phobias. Table 10.2 also presents the prevalence phobias is important in modeling how genes cause the
of the major subclasses of specific phobia, which were phenotype. Since phobias and unreasonable fears are
not included in the survey presented in Table 10.1. both highly prevalent, it will be important to know
The prevalence of phobias is greater in females (with whether phobias represent the severe end of a con-
the possible exception of illness-injury phobia), but tinuum of irrational fears or whether they are separate
the prevalence of irrational fears is similar in both disorders. If they are genetically distinct, then phobic
sexes (with the possible exception of agoraphobia). disorders must be diagnosed carefully in order to maxi-
Comorbidity with multiple phobias is common; the mize the homogeneity of samples used in genetic studies.
tetrachoreic correlation among phobias within indi- The familial transmission of several phobia sub-
viduals ranges from 0.29 to 0.54 [4, 5]. The reliability types demonstrates that these phobias are familial and
and long-term stability of the diagnoses of phobias implies that the transmission is specific for each of the
was examined with repeated interviews after one subclasses examined [8]. The findings in Table 10.3
month and again after eight years [6]. Kappas ranged reflect rates of phobias in interviewed relatives of
from 0.48 to 0.52 for all phobias but blood-injury, for probands with specific phobia, social phobia, and
which the kappa was 0.29. Long-term stability was agoraphobia. The rates of all three phobias are
even lower, with kappas ranging from 0.24 to 0.38. increased in the relatives, but only in relatives of
These findings speak for the importance of follow-up probands with each respective phobia. The magnitude
interviews in phenotyping the phobic disorders. of the increase is moderate since the relative risks
were 3.4 for specific phobia, 2.5 for social phobia,
and 3.3 for agoraphobia. It might be assumed that
Genetic epidemiology agoraphobia with panic would increase the familial
Genetic epidemiology can address a number of ques- risk for agoraphobia without panic but no trend in that
tions of practical importance for a search for genes direction was found. Familial transmission patterns,
for phobic disorders. The fundamental question is therefore, are consistent with genetic transmission of
whether additive genes predispose to phobic disorders. these phobias in a subtype-specific manner.

113
Chapter 10: Genetics of the phobic disorders and GAD

Table 10.3 Familial specificity of phobias.

Specific Social Agoraphobia Unaffected


phobiaa phobia with panic controls
No. of probands 15 39 49 77
No. of interviewed relatives 49 105 131 231
Phobia types (%)
Specifica 31 13 10 9
Social 10 15 8 6
Agora with panic 0 2 10 3
a
Specific phobia was referred to as “simple phobia” in the reference.
Rates in bold are statistically significant at p < 0.05 or less.
From Fyer et al. [8], with permission.

Table 10.4 Genetic and environmental variance in twin by both twins such as family environment, and specific
concordance. environment refers to environmental influences that
Phobia a2 c2 e2 affect one twin and not the other. In extrapolating these
effects from twins to the general population it is easily
Agoraphobia w/o panic seen that the same concepts apply equally to families.
Females 0.39 – 0.61
The principal finding is that additive genes make a
Males 0.37 – 0.63
relatively small, though not an inconsiderable contri-
Social phobia bution to all four phobias, the heritability ranging
Females 0.30 – 0.70 from 20 to 39%. Differences in heritability between
Males 0.20 – 0.80 the sexes are not large, arguing against sex-specific
Animal phobia genetic effects that might be expected, considering
Females 0.32 – 0.68 that all phobia subtypes except the blood-injury group
Males 0.35 – 0.65 are more common among women. Although the
Situational phobia preponderance of the variance was environmental,
Females – 0.27 0.73 no common environmental influence was found, the
Males 0.25 – 0.75 only exception being in women with situational
a2 ¼ additive genetic variance; c2 ¼ common environmental phobia where all of the variance was environmental
variance; e2 ¼ specific environmental variance. The variance with a minor proportion of it belonging to common
estimates are from the best fitting model from each analysis. environment. The authors were skeptical of the find-
The data on males are for irrational fears plus phobias; figures
on females are for phobias. From Kendler et al. [4, 7], with ing, and indeed, the profile of male twins resembles
permission. that of the other three phobias. As noted above, the
phobias have rather modest diagnostic reliability,
and this has the effect of underestimating the genetic
variance and overestimating the environmental vari-
Twin studies are able to separate genetic and envir- ance. In this regard, it is reassuring that when test–retest
onmental components of transmission and quantitate reliability was maximized with repeated interviews,
the magnitude of each liability. Table 10.4 presents the the genetic variances were found to be substantially
results of a population-based study of male and female greater: agoraphobia 0.67, animal phobia 0.47, situ-
twins [4, 5, 7]. The variance of each subclass of phobic ational phobia 0.46, social phobia 0.51 [6].
disorder is partitioned into additive genes, common When the phobic themes of blood, needles, hos-
environment and individual-specific environment in pitals, and disease were examined separately, irrational
male and female twins separately. Common environ- fear or phobia of the first three themes had additive
ment refers to environmental influences that are shared genetic influences accounting for 34–56% of the

114
Chapter 10: Genetics of the phobic disorders and GAD

variance in concordance, with unshared environment With regard to genes shared among the phobias,
accounting for the remainder [9]. Fear and phobia of twins can be used to partition the variance into com-
disease, however, had no genetic component, the vari- ponents that are shared by two or more disorders and
ance being fully accounted for by specific (70%) and unshared (i.e. specific to each disorder). Likewise, the
common environment (30%). specific environmental variance can be partitioned
Do irrational fears lie on a genetic continuum into shared and unshared components. (N.B. the
with phobias or are they genetically distinct dis- terms shared and unshared are used in place of the
orders? This question was examined by applying the terms “common” and “specific”, used in the publica-
multiple threshold model to twin data; the broad tions cited to avoid confusion with common and
threshold included fears and phobias and the narrow specific environmental variance as those terms are
one only phobias [6]. The phobic disorders examined used elsewhere in the chapter.) The results of such
in the analysis included agoraphobia, social, situ- an analysis are presented in Table 10.5 [7]. Shared
ational, animal, and blood-injury phobias. Multiple additive genes contributed 5–35% of the total variance
threshold models fit each of the individual phobias, for four phobias; animal phobias had the greatest
consistent with the conclusion that irrational fears proportion of shared genes (21–35%); and social
and phobias do indeed lie on a genetic continuum. phobia the least (5–10%). Unshared genetic influences
Similar to the relationship of fears to phobias, ranged from 2 to 29%, with social, animal and situ-
the greater prevalence of phobias in women might ational phobias having the largest proportion at
be explained by two genetic hypotheses. Women 13–18%. Specific environment accounted for most
might have a lower threshold for manifestation of of the residual variance, which was also composed of
the phenotype; or alternately, the same liability genes shared and unshared components. The question that
could differ between the sexes. Here again, the two is likely of most interest to psychiatric geneticists
hypotheses can be tested with the multiple threshold is what proportion of the total genetic variance is
model, assuming that women have a broader thresh- accounted for by shared versus unshared genes. Aver-
old than men to account for their greater prevalence. aging the sexes in Table 10.5 and expressing shared
For agoraphobia, situational, and blood-injury genetic variance as a percent gives the following
phobias the findings indicate similar heritabilities proportions: agoraphobia 52%, social phobia 30%,
between the sexes but some differences in the liability animal phobia 79% (with a large difference between
genes [10]. Animal phobias have no differences in the sexes), and situational phobia 36%.
either parameter, implying the same heritability and The anxiety disorders could also share common
the same genes for men and women. In this study, the genes as an explanation for the observed comorbidity
best-fitting model estimated a heritability in women among these conditions, though there are other
for social phobia of zero. The investigators had previ- explanations. For six anxiety disorders (generalized
ously found the heritabilities of men and women to be anxiety disorder, panic disorder, agoraphobia, social,
similar (for example, see Table 10.4) and suggested animal, and situational phobias) the odds ratios for
that the discrepant outcome may have resulted from comorbidity between all possible pairs of disorders
stochastic variation in the data. ranged from 1.5 to 6.3 [11]. To what extent could this
Summarizing, the phobias have an additive gen- comorbidity be due to shared genes? The best fitting
etic liability of small to moderate magnitude and a genetic model in twins, shown in Table 10.6, found
moderate-to-large specific environmental compon- two additive genetic factors contributing to the six
ent but no common environmental influences. Do disorders: one factor loaded on generalized anxiety
all of the phobias share the same genetic liability disorder, panic disorder, and agoraphobia; the other
with the type of phobia being determined by the loaded on the animal and situational phobias, leaving
environment; alternatively, is the genetic liability social phobia intermediate between the two [11].
for each phobia specific, or does the truth lie some- Although each shared genetic factor accounted for
where in between? This question can be broadened only 20–27% of the total variance (genetic plus envir-
to ask whether the phobic disorders as a group have onmental) of the disorders on which they loaded,
specific “phobia” genes or whether some of the together they accounted for all of the genetic variance
liability genes are shared with other anxiety dis- for each phobia, with the exception of agoraphobia
orders, and even with nonanxiety disorders. where they accounted for 58% (with unshared genetic

115
Chapter 10: Genetics of the phobic disorders and GAD

Table 10.5 Proportion of variance.

Phobia Genetic variance Environmental variance

Specific environment Common


environment
Shared Unshared Shared Unshared
Agoraphobia
Female 0.07 0.29 0.64 0.00 0.00
Male 0.11 0.02 0.25 0.39 0.23
Social
Female 0.10 0.21 0.32 0.36 0.00
Male 0.05 0.13 0.42 0.35 0.05
Animal
Female 0.35 0.00 0.05 0.59 0.00
Male 0.21 0.15 0.15 0.49 0.00
Situational
Female 0.09 0.20 0.17 0.53 0.00
Male 0.13 0.18 0.12 0.57 0.01
In this table the total variance is partitioned into genetic, common environmental, and specific environmental components as in the text
and in Table 10.4. The genetic and specific environmental variance are further partitioned into shared and unshared components. Shared
variance consists of genetic and environmental liabilities that influence two or more disorders. Specific variance is comprised of genetic
and environmental liabilities that influence only that disorder. The terms “shared” and “unshared” in the table replace the terms “common”
and “specific” in the publication cited in order to avoid confusion with the way the latter two terms are used elsewhere in the chapter and
in Table 10.4. From Kendler et al. [7], with permission.

Table 10.6 Components of the genetic variance for six anxiety disorders.

Disorder Shared genetic Unshared genetic Total genetic

First Second
General anxiety disorder 20 3 0 23
Panic disorder 27 1 0 28
Agoraphobia 20 1 15 36
Social phobia 8 2 0 10
Animal phobia 1 23 0 24
Situational phobia 1 23 0 24
The total genetic variance can be partitioned into shared and unshared components. Shared genetic variance is common to all disorders
and was further partitioned into two components labeled as the first and second common genetic component. Unshared genetic variance
refers to additive genes that contribute only to that disorder. The terms “shared” and “unshared” in the table replace the terms “common”
and “specific” in the publication cited in order to avoid confusion with the way the latter two terms are used elsewhere in the chapter
and in Table 10.4. From Hettema et al. [11], with permission.

factors explaining the remaining 42%). Environmental and externalizing disorders in twins [3]. Internalizing
variance was predominantly specific and composed of disorders included major depressive disorder, gener-
shared and unshared sources. alized anxiety disorder, and phobic disorder; the
Further evidence for shared liability with the externalizing disorders were substance dependence,
phobic disorders comes from a study of internalizing antisocial behavior, and conduct disorder. The results

116
Chapter 10: Genetics of the phobic disorders and GAD

supported separate additive genetic liabilities for hypersensitivity whereas the key feature of social
internalizing and externalizing disorders, and each phobia is an unreasonable fear of embarrassment.
genetic factor was shared by all disorders within its Are these two conditions unrelated or are they different
respective group. The shared genetic liability to the manifestations of the same disorder? A study of female
internalizing disorders was further subdivided into twins from the Norwegian population found that
two liabilities: one shared by major depression and the same additive genetic component contributed to
generalized anxiety disorder, the other by the phobic both avoidant personality disorder and social phobia,
disorders (panic disorder correlated weakly with the accounting for 37 and 39% of the variance, respectively
first genetic liability). The genetic liability of the [14]. The remainder of the variance for social phobia
internalizing disorders was virtually entirely shared, was accounted for by specific environment. These find-
panic disorder being the only exception with ings suggest that social phobia and avoidant personality
unshared genetic liability. The results were summar- disorder share their genetic liability with a phobic
ized by suggesting that the internalizing disorders outcome determined by the environment.
could be grouped into those characterized by “anx- With the genetics of the phobic disorders estab-
ious misery” (depression and generalized anxiety) and lished, studies are beginning to ask more specifically
those characterized “fear” (situational and animal what constitutes the additive genetic component.
phobias), with panic disorder not falling cleanly into Extraversion and neuroticism are recognized person-
either group. Agoraphobia, social, specific, and ality traits that are reliably measurable and known
blood-injury phobias were not examined and whether to be genetic. Could high neuroticism and low extra-
they would also cluster into the “fear” group remains version be genetic liabilities to phobic disorders,
unresolved. particularly social phobia? This hypothesis was tested
Social phobia can be further specified as general- in twins by examining correlations between the two
ized or not, although generalized social phobia is not personality traits and individual phobic disorders,
considered as a separate subclass in DSM-IV. First- with differences between monozygotic and dizygotic
degree relatives of probands with generalized social twin pairs estimating the genetic influences under-
phobia have an increased risk for that phobia com- lying the correlations [15]. Each personality trait
pared with families of unaffected probands at an and three phobias (agoraphobia, social, and animal
odds ratio of 9.7 [12]. The transmission was limited phobia) had genetic predispositions, as evidenced by
to generalized social phobia as the relatives did not greater monozygotic than dizygotic twin concord-
have an increased risk of nongeneralized social ance. High neuroticism and low extraversion correl-
phobia. The rate of avoidant personality disorder ated moderately with agoraphobia and social phobia;
was also increased in the families of generalized social but animal phobia was uncorrelated with extraversion
phobia probands, being diagnosed in 19.8% com- and only weakly correlated with neuroticism. When
pared with 0% of controls. correlations between the two personality traits and the
The observed comorbidity between generalized three phobias were included in genetic models the
social phobia and generalized anxiety disorder was personality traits accounted entirely for the genetic
examined by comparing the rates of both disorders liability to agoraphobia and social phobia but not to
in relatives of probands with each diagnosis, probands animal phobia. These findings would imply that the
comorbid for both disorders, and unaffected controls genes responsible for extraversion and neuroticism
[13]. The rate of generalized social phobia was fully account for the genetic liability to agoraphobia
increased in family members of probands with gener- and social phobia.
alized social phobia (with or without comorbid gen- Further traits that could predispose to social phobia
eralized anxiety disorder) at an odds ratios greater are social anxiety, trait anxiety, and personality traits.
than 3. These results suggest that generalized social These traits were examined in a family study by measur-
phobia is familial and its transmission is limited to ing social anxiety with assessments of fear: negative
generalized social phobia. evaluation, social performance and social interaction;
Social phobia may overlap to a substantial extent trait anxiety: the State-Trait Anxiety Inventory; and per-
with avoidant personality disorder. The cardinal fea- sonality: the Tridimensional Personality Questionnaire
ture of that disorder according to DSM-IV is social [16]. First-degree relatives of probands with general-
inhibition due to pervasive feelings of inadequacy and ized social phobia from an earlier study [12] had

117
Chapter 10: Genetics of the phobic disorders and GAD

higher social anxiety, higher trait anxiety, and greater analyzed for logarithm of odds (LOD) and heterogen-
harm avoidance, compared with relatives of control eity LOD (HLOD) scores, as well as nonparametric
probands who were free of social phobia. Since these linkage (NPL) scores. The parametric analyses were
measures were strongly intercorrelated, a principal conducted with dominant and recessive genetic
components analysis was performed and resulted in models; both a broad (definite and probable diag-
a single component that accounted for 82% of the noses) and a narrow (definite diagnoses only) pheno-
variance that loaded on all of the measures enumer- type were analyzed. The strongest evidence for linkage
ated above. These findings are correlative and cannot to agoraphobia appeared on chromosome 3 at pos-
address whether the traits predisposed to or were ition 167.5, where an NPL score of 2.75 was associated
influenced by the phobic disorders, but the analyses with a p-level of 0.005. None of the parametric ana-
do identify a number of potential endophenotypes lyses in this region supported linkage. On inspection
for further genetic studies of social phobia. of the individual pedigrees one family generated an
Apart from personality traits, another possible NPL score of 10 and was responsible for the NPL peak
endophenotype is susceptibility to fear conditioning, in the full pedigree set.
which has been shown to be genetic in humans [17]. The same pedigree set was analyzed for linkage
Fear-relevant images (snakes and spiders) and fear- to specific phobia [19]. In these analyses 14 pedigrees
irrelevant images (geometric shapes) were paired with were informative for linkage. Fifty-seven family
a mild electric shock, and galvinic skin response was members had simple phobia, 66% of which were
measured as the conditioned response in twins. One phobias of the natural environment, with the remain-
image of each pair (fear-relevant and fear-irrelevant) der phobic for animals, blood-injury, situations, and
was presented without shock, providing a measure other conditions. The linkage scan was strongly indi-
of generalization of the conditioned response to an cative of linkage to a locus on chromosome 14 at
unconditioned stimulus. Three phases of condition- 37 cM. A marker at that locus generated a LOD score
ing were examined: habituation (presentation of of 3.17 with the narrow-phenotype, dominant model,
images without shock), acquisition (presentation of and a Zlr score of 3.93, associated with a p-value of
images with shock), and extinction (re-presentation 4  10–5. This result is particularly interesting because
of images without shock). The conditioning process the provisionally linked region is homologous to a
was found to be moderately heritable: the heritability region of the mouse genome that has been linked to
of habituation was 35%, acquisition 35–43%, and an anxiety model. Furthermore, a candidate gene for
extinction 36%. The heritability of acquisition is similar anxiety, human somatostatin receptor-1, is located in
to the heritability found in animal phobias, 32 and 35% that region as well.
respectively (Table 10.4). Genetic modeling implied Genes for social phobia were also sought in these
different additive genes for the conditioned response pedigrees [20]. In 17 informative pedigrees, the 163
to fear-relevant and fear-irrelevant stimuli. This finding relatives included 56 with social phobia in addition
is consistent with the theory that phobias evolved to to 4 probable cases. Parametric analyses of a broad
protect against environmental threats. phenotype (definite and probable social phobia) and
a recessive genetic model resulted in a LOD score
of 2.22 on chromosome 16 at 71.1 cM. A nonpara-
Genomic studies of phobic disorders metric linkage analysis at the locus resulted in a
Genome searches have been reported for agorapho- p-value of 0.002. Thus, searches for genes for three
bia, simple phobia, social phobia, and for a broad of the phobic disorders in these pedigrees found
phenotype that includes these three phobias in addi- separate regions of interest for each disorder with
tion to other anxiety disorders [18–21]. All four stud- no overlapping regions among them. This outcome
ies are based on the same sample of families selected is consistent with evidence reviewed earlier that
through probands with panic disorder and genotyped agoraphobia, social phobia, and situational phobias
with a 10 cM marker map. have different genetic predispositions.
A genomic search of agoraphobia was conducted A final analysis of this pedigree set included the
in 153 family members of panic disorder probands in above three phobias (agoraphobia, social, and simple
20 American families. Approximately two-thirds of the phobia) and panic disorder as the affected phenotype
61 agoraphobics had panic disorder. The families [21]. Cluster analysis created two overlaping groups

118
Chapter 10: Genetics of the phobic disorders and GAD

by scoring each diagnostic assessment on a three- [24, 25] and found the odds ratio for that disorder
point scale (absent, probable, or definite diagnosis). among first-degree relatives to be 6.1, compared with
The phenotype used in the linkage analysis was a relatives of unaffected probands. The study by Noyes
“group of membership” score that was analyzed by et al. [24] found a rate of 19.5% in family members of
the Haseman–Elston linkage algorithm. The analysis probands with GAD but the rate was not increased in
identified a locus at chromosome 4q31.21–32.3 families of probands with panic disorder or agora-
(D4S413) with nominal, 2-point and multipoint p- phobia, supporting a hypothesis of genetic specificity
values less than 10–5, reaching the accepted signifi- for GAD. An additional family study published since
cance level for linkage. The genome-wide, empirical the meta-analysis is important because, in contrast
multipoint p-value for linkage was 5.6  10–4. The to the above studies, the families were ascertained
region is of interest because it contains the gene for from a population-based sample [26]. The analyses
neuropeptide Y receptor, which has been implicated controlled for potential confounding by comorbid
in anxiety in rodent models. This analysis was par- depression and other anxiety disorders, and found
ticularly interesting because of the novel way it com- that the odds ratios for first-degree relatives ranged
bined multiple disorders into an affected phenotype, from 1.4 to 1.8. These odds ratios are lower than the
thus capitalizing on findings that many of the anxiety estimate of 6.1 from the meta-analysis, as might be
disorders share additive liability genes. expected since the studies analyzed were from clinic-
based samples.
Twin studies indicate that the familial transmis-
Conclusion sion of generalized anxiety disorder is due to genes.
Genetic epidemiology indicates that finding genes for A population-based sample of 20 monozygotic and
the phobic disorders will require samples with con- 29 dyzygotic twins with GAD from Norway found
siderable statistical power since the heritability’s of 40% of monozygotic twins concordant compared
these disorders are modest. Thus, it is encouraging with 10% of dizygotes [27]. Kendler et al. [4, 5]
that heritabilities can be increased substantially by studied DSM-III-R diagnoses in a population-based
repeated assessments. It will be critical to choose sample of female twins. Cases with a minimum of
phenotypes wisely to avoid degrading statistical one-month and six-months duration were analyzed
power with genetic heterogeneity. A number of the together with a multiple threshold model, the former
phobias share additive genes and understanding this being the broad and the latter the narrow threshold.
will lead to more informed phenotype definitions. This analysis found 34% concordance in monozygo-
Well-validated endophenotypes could provide an tic compared with 13% in dyzgotic twins. Genetic
index of gene expression that is closer to the genome factors accounted for 28% of the variance with spe-
than clinical diagnosis. All of these considerations will cific environment accounting for the remainder.
require careful balancing since the large samples Like the phobias, GAD appears to have a definite
needed for genome-wide association studies will place but modest genetic predisposition. When cases of
limits on the amount of time and resources that can GAD that only occurred during episodes of depres-
be applied to phenotype specification. sion or panic disorder were excluded, the results
were similar to those found without a diagnostic
hierarchy, indicating that the concordance was
Generalized anxiety disorder indeed due to GAD and not comorbid conditions.
The cardinal feature of generalized anxiety disorder In a similar analysis of these data, additive genes
(GAD) is at least six months of apprehensive expect- contributed 14% of the variance in both males and
ation, characterized by excessive anxiety and worry; females, with specific environment contributing the
the apprehension is manifest through such symptoms remaining 86% [28]. If depression excluded GAD in
of anxiety as restlessness, fatigue, poor concentration, hierarchical diagnoses, the best-fitting model esti-
irritability, muscle tension, and insomnia [22]. This is mated the additive genetic variance at 21%, some-
a frequent condition, affecting 5.7% of the population what closer to the 28% found in the Kendler et al.’s
at some time during their lives [3]. analyses [4, 5].
Hettema et al. [23] performed a meta-analysis A twin panel of Vietnam-era veterans provides
of two family studies of generalized anxiety disorder additional evidence on the genetics of GAD [29, 30].

119
Chapter 10: Genetics of the phobic disorders and GAD

Diagnoses were based on DSM-III-R criteria except section [11]. With regard to GAD, additive genes
that the illness duration was one month (as in DSM- shared with panic disorder and agoraphobia
III) rather than six months (as in DSM-III-R and accounted for 20% of the total variance of GAD;
DSM-IV). The data were analyzed for GAD and panic indeed, this shared variance component made up
disorder jointly. Additive genes were found to have a 87% of its additive genetic variance. The genetic pre-
small but appreciable influence, accounting for 38% disposition to generalized anxiety disorder is also
of the variance of GAD with unshared environment shared with major depressive disorder as well [4, 5].
contributing the remaining 62%. The additive genetic GAD in female twins was defined by three durations
component of GAD also accounted for 22% of the total of symptoms between one and six months. With each of
variance of panic disorder and was the sole additive the three durations of GAD symptoms the best-fitting
genetic factor for panic. genetic model predicted that the genetic predisposition
The extent to which the anxiety disorders share was completely shared between the two disorders, and
some of their genetic predisposition has been con- the remaining variance was predominantly specific
sidered from the perspective of the phobias in that environment.

References 11. Hettema JM, et al. Arch Gen


Psychiat 2005;62:182–189.
22. American Psychiatric Association.
Diagnostic and Statistical Manual
1. Orel V. Gregor Mendel – The First of Mental Disorders, Fourth
Geneticist. Translator: Finn S. 12. Stein MB, et al. Am J Psychiat
1998;155:90–97. Edition, Text Revision (DSM-IV-
Oxford: Oxford University Press; TR). Washington, DC: American
1996. 13. Coelho HF, et al. J Affect Dis Psychiatric Association; 1994;
2. Kessler RC, et al. Arch Gen 2007;100:103–113. pp. 441–456.
Psychiat 1994;51:8–19. 14. Reichborn-Kjennerud T, et al. 23. Hettema JM, et al. Am J Psychiatry
3. Kessler RC, et al. Arch Gen Am J Psychiat 2007;164: 2001;158:1568–1578.
Psychiat 2005;62(6):593–602. 1722–1728.
24. Noyes R. Jr., et al. Am J Psychiatry
4. Kendler KS, et al. Arch Gen 15. Bienvenu OJ, et al. Am J Psychiat 1987;144:1019–1024.
Psychiatr 1992;49:273–281. 2007;164:1714–1721.
25. Mendlewicz J, et al. Psych Genet
5. Kendler KS, et al. Arch Gen 16. Stein MB, et al. Am J Med Genet 1993;3:73–78.
Psychiatr 1992;49:716–722. 2001;105:79–83.
26. Newman SC, et al. Psychol Med
6. Kendler KS, et al. Psychol Med 17. Hettema JM, et al. Arch Gen 2006;36:1275–1281.
1999;29:539–553. Psychiat 2003;60:702–708.
27. Skrye OI, et al. Acta Psychiatr
7. Kendler KS, et al. Arch Gen 18. Gelernter J, et al. Am J Med Genet Scand 1993;88:85–92.
Psychiat 2001;58:257–265. 2001;105:548–557.
28. Hettema JM, et al. J Nerv Ment Dis
8. Fyer AJ, et al. Arch Gen Psychiat 19. Gelernter J, et al. Molecular Genet 2001;189:413–420.
1995;52:564–573. 2003;8:71–82. 29. Scherrer JF, et al. J Affect Dis
9. Neale MC, et al. Am J Med Genet 20. Gelernter J, et al. Am J Psychiat 2000;57:25–35.
1994;54:326–334. 2004;161:59–66. 30. Chantarujikapong SI, et al.
10. Kendler KS, et al. Psychol Med 21. Kaabi B, et al. Am J Hum Genet Psychiatry Res 2001;103:
2002;32:209–217. 2006;78:543–553. 133–145.

120
Genetic contributions to obsessive–
Chapter

11 compulsive disorder (OCD) and


OCD-related disorders
Dennis L. Murphy, Pablo R. Moya, Jens R. Wendland, and Kiara Timpano

Introduction symptoms in first-degree relatives of OCD probands


when compared to first-degree relatives or
Obsessive–compulsive disorder (OCD) is a major, psychiatrically healthy controls. In one of the first of
severe neuropsychiatric disorder, with an estimated these, the age-corrected risk of “broadly defined OCD”
lifetime prevalence based on population surveys con- was higher in first-degree relatives of OCD subjects
ducted in many communities nationally and inter- compared to relatives of psychiatrically normal con-
nationally of 2–3% with a > 2-fold prevalence than trols (10% versus 1.9%) [6]. More recently, an increased
schizophrenia and an approximately similar prevalence risk of definite OCD in first-degree relatives of OCD
as bipolar disorder [1]. probands was substantiated when compared to first-
Onset of OCD can be in the pre-teen years as degree relatives of matched community controls, an
young as age three, although most commonly, the excess in the range of four- to eight-fold [4, 7, 8].
period of greatest risk is from adolescence to early Some segregation analyses of OCD families have
adulthood; less commonly the disorder may first suggested a dominant major gene [9–12]. Validation
occur in later decades. Most patients experience a with larger samples which focused solely on OCD
chronic course, while some have an episodic cyclic diagnosis rather than OCD spectrum symptoms
disorder, with exacerbations and remissions. In either found results consistent with a complex genetic model
case, OCD symptoms usually are only partially including a possible single major locus, which was
responsive to various pharmacological and behavior- neither dominant nor recessive, in combination with
ally based treatments, being persistent and lifelong, multiple other contributing loci – suggestive of a mixed
substantially impairing function. model of inheritance [13, 14].
There is now compelling evidence for biological con- In a still-limited number of studies, monozygotic
tributions to OCD, primarily based on brain imaging twins have been reported as concordant for obsessive–
studies that have defined likely brain circuitry involved in compulsive symptoms [15–18]. Although reports con-
OCD symptoms plus increasing contributions from gen- trasting the rates of the disorder in monozygotic
etic studies. Early studies had suggested a role for hered- versus dizygotic twins with OCD are few in number,
ity in OCD. More recently, investigations have indicated a Japanese series found concordance for obsessive–
likely contributions of chromosomal risk regions, spe- compulsive symptoms in 80% of monozygotic twins,
cific candidate genes and gene pathway networks to compared to 50% of dizygotic twin pairs [19]. In the
OCD etiology, as will be reviewed in this chapter. Maudsley Twin Register, the concordance rates in
monozygotic and dizygotic twin pairs were 87% and
Family and twin studies of OCD 47%, respectively, giving a heritability estimate of 80%
[20]. No adoption or separation studies, comparing
and OCD-related disorders the rates of OCD in twins raised together or apart,
Early family studies of OCD were conducted with vary- have been reported. The most recent reviews of OCD
ing methodological rigor [2–5]. More recent family or OCD symptoms in twins found intermediate esti-
studies, using quantitative diagnostic methods, have mates of heritability which were higher in pediatric
found substantially higher risks for OCD and OC than adult OCD probands [21, 22].

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

121
Chapter 11: Genetic contributions to OCD and related disorders

Formal genetics of OCD probands, a relatively strong association peak was


detected nearby at allele D3S3600 (p ¼ 0.002).
Linkage approaches to the genetics of OCD
The first genome linkage scan for well-diagnosed Age of OCD onset as a secondary
OCD in seven extended pedigrees identified one can-
didate region on 9p24 that met criteria for suggestive phenotype in linkage studies
significance (logarithm of odds [LOD] 1.97) [23]. In Age at onset, used as a secondary phenotypic approach
an attempt to replicate this finding, a second study of to investigate linkage regions in this second genome
50 OCD families focused on microsatellite markers linkage study, led to additional results of interest [26].
spanning the 9p24 candidate region. In support of the The median age of onset in this sample of OCD pro-
original report, a nonparametric analysis identified bands and families was 7 years, reflecting in part, that a
a linkage signal at marker D9S1813 with an nonpara- criterion for entry into this study was onset before age
metric linkage (NPL) of 2.52 (p ¼ 0.006); this peak 18. In the analyses using MERLIN, suggestive linkage
lies within 0.5 cM of the original 9p24 linkage signal for the younger age of onset group (LOD ¼ 3.21; p ¼
[23, 24]. A follow-up linkage analysis by authors of 0.0001) was found at 1q23–24; a smaller shift in LOD
the first scan included 121 subjects from 26 families score peaks was also observed in one of the chromo-
and reported a maximum NPL score of 2.43 at some 3 peaks when the sample was divided into groups
chromosome 10p15, considered “suggestive” only [25]. with age of onset before or after 7 years [26]. This was
In a second genome-wide linkage study, OCD pro- confirmed when families with two or more young age of
bands and their sibling pairs and some family members onset OCD relatives were evaluated as an additional
(N ¼ 1008 total) were evaluated using a NPL method. subgroup. When the sample was divided based on the
Both Kong and Cox LOD all and Kong and Cox LOD proband’s gender (78 male proband families; 141 female
pair statistics were computed and empirical p-values proband families), a greater increase in the linkage
for all “significant signals” were evaluated with Merlin, signal at 11p15 was observed [27]. After genotyping
using 10 000 replicates. Evidence for susceptibility loci additional microsatellite markers, the gender-stratified
was found on chromosomes 1q, 3q, 6q, 7p, and 15q analysis using MERLIN revealed a LOD ¼ 3.02 (p ¼
[26] and were genotyped at the Center for Inherited 0.0001) in the male group. This region contains plaus-
Disease Research (CIDR) using 386 microsatellite ible candidate genes, such as tyrosine hydroxylase (TH)
markers with an average spacing of 9 cM. Of interest, and the dopamine D4 receptor (DRD4).
no peak in chromosome 9 exceeded a LOD p-value of
1.0, thus not replicating the two prior studies that had
pointed to a 9p24 region of interest [26]. Rather, the
Hoarding among probands with OCD as
multipoint nonparametric analyses found several sug- a secondary phenotype in linkage studies
gestive linkage regions that included chromosome 1 (p- When this sample was stratified based on the presence of
value ¼ 0.003) and chromosome 3 (p-value ¼ 0.0002). two or more relatives with OCD plus hoarding symp-
The highest Kong and Cox LOD score (2.67, p ¼ 0.0002) toms (74 hoarding families; 145 nonhoarding families), a
was obtained at the marker D3S2398 located at 209 cM on suggestive linkage signal was found at 14q31–32 (LOD ¼
chromosome 3. None of these linkage peaks exceeded 2.99; p ¼ 0.0001). Candidate genes in this region include
values needed for more than “suggestive” linkage with three serotonin receptor genes (3C, 3D, and 3E),
OCD. although only 3C is known to be expressed in brain [28].

Fine mapping of chromosome 3: 3q27–28 Genome-wide association analysis


The strongest linkage result with the second linkage As with other genetically complex medical disorders,
study near marker D3S2398 was further assessed by methodological approaches in psychiatric genetics are
fine mapping using a combined set of microsatellite shifting to genome-wide association studies (GWAS),
and SNP markers to obtain an average marker density which make more comprehensive single nucleotide
of 360 kb [26]. This follow-up study narrowed the polymorphism (SNP) -based assessments possible
linkage region from 37 to 14 cM. When family-based (e.g. [29, 30]). One GWAS study investigating over
association analyses were performed using the same 1000 OCD probands and over 1000 controls has been

122
Chapter 11: Genetic contributions to OCD and related disorders

Figure 11.1 An extended


multiply-affected family. Obsessive–
compulsive disorder (OCD) -affected
males (shaded squares) and females
(shaded circles); nonaffected family
members shown by non-shaded squares
and circles. (Modified from [23].)

partially analyzed and preliminary reports have been There are five major studies of SLC1A1 in OCD,
reported in abstract form only [31]. The highest all of which reported significant association of the
genome-wide p-values point towards glutamate gene with OCD. Four of these were family-based
system gene involvement [31]. studies, primarily using trio samples [32–35]; these
were supported by a fifth large case-control study
(Table 11.1) [36].
Candidate chromosomal regions The glutamate transporter gene SLC1A1 (also
known as EAAC1), is located 350 kb centromeric to
and candidate genes for OCD the linkage peak, 9p24, that was described in the first
Potential candidate genes for OCD based on genome linkage scan of OCD. Its protein product, the
neuronal transporter gene, seems critically important
functional studies and prior investigations based on the central role of glutamate in circuits con-
Numerous gene products seem highly relevant to sistently implicated in OCD. Although SLC1A1 was
neurotransmitter system pathways and developmen- not specifically identified in the first genome scans, nor
tal sequences important in OCD, but only relatively in a follow-up study of the 9p24 region where SLC1A1
few have been investigated. These include glutamate, is located [23, 24], and no signal was detected in the
dopamine, serotonin, and other systems, neuro- largest genome scan, the series of direct SNP-based
trophic factor genes and their affiliated receptors, studies finding significant associations with this neur-
and genes indirectly implicated via comorbid dis- onal glutamate transporter gene, and other studies
orders or suggested from animal models of OCD suggesting additional glutamate system candidates,
and OCD-related behaviors (such as perseverative such as GRIN-2B and GRIK-2, sustain continued inter-
behaviors, over-grooming, and hoarding). est in the brain glutamate system in OCD [36–38].
Biological and neurodevelopmental studies also
Candidate genes for OCD support the glutamate system as being involved in
the neurocircuitry of OCD, including brain imaging
SLC1A1 and other glutamate system genes studies and direct magnetic resonance study (MRS)
SLC1A1 encodes the only neuronal glutamate trans- -based evaluations of brain glutamate concentrations
porter and is an attractive OCD candidate gene for in OCD treatment studies [39]. In addition, riluzole,
multiple reasons. As a positional candidate, it is the a glutamate antagonist has been evaluated in several
most evident brain gene of interest located in the treatment trials of OCD patients, with positive
chromosomal region 9p24, the region identified in results, although replicated, placebo-controlled trials
the first genome-wide linkage study of mixed large of this agent have not yet been reported [40–42].
and small families with OCD [23]; an adapted version
of the four-generation pedigree of one large family
from this study is reproduced in Figure 11.1. As noted
SLC6A4, the serotonin transporter (SERT, 5-
above, this linkage result was supported by a second, HTT) gene and other serotonin system genes
more directed study of this region [24], but was not SLC6A4 was among the very first brain genes associ-
replicated as a candidate chromosomal region in the ated with an OCD diagnosis. In particular it is the
subsequent largest whole genome linkage study [26]. insertion/deletion variant (the long L 5-HTTLPR) in

123
Table 11.1 Compilation of five studies investigating SLC1A1 and obsessive–compulsive disorder (OCD) [36].

chr9: 4400000 4450000 4500000 4550000


13–1518,19
Markers 7 9 11 21
Markers 1 2 3 4 5 6
RefSeq Genes 8 10 12 16,’7 20
SLC1A1
C9orf68
A’nold et al, 2006 Dickel et al, 2006 Stewart et al, 2007 Wendland et al, 2009 Shugart et al, 2009
Polymorphism HG18 position Single locus Haplotype Single locus Haplotype Single locus Haplotype Single locus Haplotype Single locus Haplotype
1 rs3933331 (LCL eQTL) 4,379,941 not tested not tested not tested ns not tested
2 rs10814987 4,473,303 not tested ns not tested not tested 0
3 rs10739062 4,492,848 not tested ns not tested not tested 0 P = .046
4 rs1980943 4,504,246 ns ns not tested not tested 0 P = .046
5 rs10115600 4,509,369 not tested ns not tested not tested 0
6 rs3780415 4,545,573 ns not tested not tested not tested 0
7 rs7858819 (LCL & DLPFC eQTL) 4,549,892 not tested not tested not tested ns P =.0002 not tested
8 rs10974625 4,551,596 not tested not tested ns not tested not tested
9 rs7856209 (syn cSNP) 4,554,432 ns not tested ns not tested PBAT dominant: P = .04
10 rs3780413 4,557,353 not tested not tested ns not tested not tested
11 rs3780412 4,562,480 ns P =.04 ns ns 0
12 rs12682807 4,564,022 not tested not tested ns P =.004 not tested not tested
13 rs2072657 4,566,451 not tested not tested ns P =.004 not tested not tested
14 rs301430 (syn cSNP, LCL & DLPFC eQTL, luciferase) 4,566,680 ns P =.03 P =.03 ns P =.004 ns P =.0002 0
15 rs301979 4,566,851 ns ns P =.03 ns not tested 0
16 rs301434 4,572,082 P = .0007 P = . 003 ns ns ns 0
17 rs301435 4,572,843 P = .0009 not tested ns not tested 0
18 rs3087879 (3*.-UTR) 4,576,808 P = .006 P = . 003 not tested ns P =.02 P =.0002 0 P = .02
19 rs301437 4,577,560 not tested not tested ns not tested not tested
20 rs301440 4,580,305 not tested not tested ns not tested not tested
21 rs301443 4,584,919 not tested rs not tested not tested PBAT dominant: P= .000007 P = .02
Chapter 11: Genetic contributions to OCD and related disorders

STin2 IIe425Val
SNPs VNTR
rs25531,
67 Alternate
(a) Gene rs25532
2 34 5 8 9
1011
1A 1C 1B 12 14 polyadenylation sites
13
5’

3’

5HTTLPR
(LA,LG,SA,SG)
G56A

Alternative
splicing
34

(b) SERT SNPs: 8 Extracellular


11 12
6 7
1. T4A1 9. P339L TMS
9 1013
2. G56A2 10. L362M 2 14 15 Intracellular
16
3. S214S 11. L383L
COOH
NH2
4. E215K1 12. A419A 1
5. H235H 13. 2X1425V(/L)
6. L255M 14. T439T
7. S293F 15. K605N1 or 5HT transport in transfected cells
1
No response to PKG/p38 MAPK activation
8. G308G 16. P621S1 2
No response to 8BrcGMP
Figure 11.2 Human SERT gene organization, with multiple functional variants. See plate section for color version.

the promoter region which is associated with greater case-control study did not find significant associ-
transporter expression that has been associated with ations of the triallelic 5-HTTLPR (5-HTTLPR plus
OCD in case-control and family-based studies [43– rs25531) with OCD, instead discovering only a nom-
45]. While a series of nonreplications were reported inal association with rs25531 alone that did not sur-
in different countries and ethnic groups, recent vive correction for multiple testing [49]. The most
reviews and a recent meta-analysis found that the recent study reported another new variant, rs25532,
L allele was significantly associated with OCD in also in the 5-HTTLPR region (Figure 11.2) and also
family-based studies [46, 47]. Further interest in affecting luciferase reporter gene expression by 15–
SLC6A4 arose when SNPs in the 5-HTTLPR region 80%, (depending on combinations of variants and
were discovered and found to differentially affect cells chosen to evaluate expression). Association with
expression, with the LG variant found to convert the OCD of a novel haplotype that included this variant
L allele into the equivalent of an S allele [48]. Prior together with the triallelic 5-HTTLPR plus another
studies that did not separate the LG and LA alleles may SNP, rs16965628 (which is located in intron1) was
have introduced a genotyping error of 1–24%, found in a sample of 295 OCD probands and 657
depending upon ethnicity [48]. controls [50]. Of great interest, it was the higher
In this study by Hu and colleagues [48], associ- expressing allele at each locus that was associated with
ations between the LA allele and LALA genotype and OCD diagnosis. Functional variation in the 3’
OCD diagnosis were found in a family-based study of untranslated region of SLC6A4 was recently reported,
175 trios and also in a replicate case-control study of but has not yet been evaluated in OCD [51]. Thus,
169 OCD probands and 253 controls. A following from the very earliest studies, which genotyped only

125
Chapter 11: Genetic contributions to OCD and related disorders

the 5-HTTLPR alone, to the most recent studies that Dopamine system candidate genes in OCD
investigated rs25531 and rs25532 alone or together,
Dopamine D2, D3, and most frequently, the D4
all supported the nation of increased serotonin trans-
receptor gene plus the dopamine transporter (DAT)
porter functioning with decreased extracellular fluid
gene, as well as genes in the catabolic pathways for
serotonin but increased serotonin in vesicles available
dopamine such as catechol-O-methyltransferase
to be released as possibly contributing to OCD.
(COMT) and other catecholamines, have yielded
Additionally, re-interest in SLC6A4 was also
some results that have been replicated, although con-
stimulated by observation of a rare coding region
flicting gender-related differences were also found
functional variant, SERT I425V, found associated
[70–76]. Of note, a low activity allele of the COMT
with OCD as a complex phenotype [52], a finding
gene was found to be associated with OCD in females
subsequently replicated in a larger study [53], as dis-
and subsequent studies also observed positive associ-
cussed in greater detail below. Functional studies of
ations, although only in specific subgroups or in
this variant found altered regulation of the SERT
males only [77–81]. Likewise, the monoamine oxidase
protein in cell expression systems. An enhanced basal
isoform A (MAOA) gene which is involved in both
serotonin transporter function could not be further
catecholamine and serotonin metabolism, was found
stimulated by nitric oxide precursors, unlike the
to be associated with OCD in two studies, although
common 425I allele; however a nearly two-fold
gender differences were again found [82, 83].
greater basal expression was found with I425V [54].
A follow-up evaluation of the I425V literature in
OCD cases and controls as well as autism subjects Other candidate genes in OCD
led to the conclusion that OCD was most consistently Multiple other candidate genes for OCD and OCD-
associated with this mutation (p ¼ 0.004, odds ratio related disorders have been reported but most thus far
[OR] ¼ 6.54; Fisher’s Exact Test, corrected by family have not been replicated. Several have been fairly
coefficient of identity), with SLC6A4 I425V found in comprehensively evaluated and perhaps the strongest
1.5% of 530 individuals with OCD and 0.23% of data suggest that a variant in the neurotrophin gene,
1300 controls [50]. SLC6A4 I425V has now been BDNF (e.g. V66M), is strongly associated with OCD
designated “OCD 1” in OMIM (Online Mutations [84]. Other candidates include transcription factor
in Man in Pubmed, http://www.ncbi.nlm.nih.gov/ genes such as Hoxb-8 (grooming [85, 86]; MOG [87,
pubmed), in recognition of its high penetrance and 88]), neuropeptide genes and their receptors (e.g.
replicated associations with OCD. BDKRB27 [89]), modulators of genes otherwise
SLC6A4 was initially considered a prime candidate implicated in OCD, and, of course, genes already
for investigation in OCD because the only well-valid- associated with neuropsychiatric disorders with
ated drug treatments for OCD are the selective sero- common comorbidity with OCD.
tonin reuptake inhibitors [43, 45, 55–59]. While
depressive disorders seem to be equally well treated
with tricyclic and MAO-inhibiting antidepressants as
Chromosomal anomalies and rare
well as selective serotonin reuptake inhibitors (SSRIs), gene involvement in OCD
these other antidepressants do not seem of benefit in Additional uncommon chromosomal anomalies or
OCD in placebo-controlled trials [43, 45, 55–59]. gene variants have been investigated in OCD and
SERT and SLC6A4 are also associated with other neu- OCD-related or OCD-comorbid disorders, especially
ropsychiatric disorders and with diverse medical dis- Tourette syndrome (TS) families. As OCD is often
orders in humans [58], and are likewise associated with found as a triad together with tic disorders plus atten-
multiple behavioral and neurochemical alterations in tion-deficit hyperactivity disorder (ADHD) in TS, for
mouse and rat models of serotonin transporter defi- the purposes of this review we will consider chromo-
ciency and over-expression [60–66]. Endophenotypes some regions and genes associated with TS as possible
or changes related to SLC6A4 genotypes include alter- OCD-related. The two findings with the most
ations in amygdala and cortico-amygdala function in common TS–OCD comorbidity are associated with
brain imaging studies of healthy humans, human and the 22q11 microdeletion syndrome [90–93] and OCD
mouse anxiety and stress responses, and serotonin associated with the dystonia myoclonus syndrome
receptor density and function [44, 61, 67–69]. [94–97]. The latter is of note because its 7q21-q31

126
Chapter 11: Genetic contributions to OCD and related disorders

region is near that of the chromosomal anomalies patients with focal dystonias (relative to the 2.5%
described in cases at 7q31 and 7q35–q36 (reported OCD prevalence in the general population) [106,
to be associated with OCD and TS) [98–100]. Add- 107]. All of these disorders, including TS, appear to
itionally, a family-based association study using share alterations in cortico-striato-thalamocortical
markers in the 7q31 region demonstrated biased (CSTC) pathways, although some differences in
transmission of marker alleles in individuals with pathway involvement require further elucidation via
comorbid TS, OCD, and ADHD [101]. ongoing brain imaging, electrophysiological, and
postmortem brain studies [106, 108–110].

OCD and the dystonia myoclonus syndrome OCD and the chromosome 22q deletion
In one study of three extended dystonia myoclonus
syndrome (DMS) families, OCD was present in 25% of syndrome.
symptomatic DMS carriers with the 7q21 haplotype, The 22q11.2 contiguous microdeletion syndrome
but in only 9% (1/11) of nonsymptomatic carriers and occurs in approximately 0.3% of live births, generally
0% (0/28) of nonhaploytpe carriers [96]. OCD comor- in sporadic cases, although a minority of cases
bid with generalized anxiety disorder and major ( 10%) may follow an autosomal dominant inherit-
depressive disorder was also significantly increased in ance pattern. The size of the deletion is 1.5–3.0 mb,
these three families (but neither generalized anxiety involving approximately 25 genes. Initially recognized
disorder nor major depression was increased with because of facial and cardiac malformations associ-
DMS alone) [96]. In another study of three extended ated with learning disabilities, and therefore named
families with DMS, OCD was diagnosed in three the velo-cardio-facial syndrome (VCFS), more exten-
members of one of the families; all three were symp- sive evaluation of children and adults with this dele-
tomatic DMS carriers; OCD was not present in any of tion uncovered diverse psychopathology, ranging
the other 10 members of this family nor in 14 members from ADHD, pervasive developmental disorder
of another 2 families from which psychiatric profiles (PDD), and anxiety disorders – including OCD – to
were obtained [94]. Individuals in the family with schizophrenia and bipolar disorder [111].
DMS and OCD had a 7q21 deletion mutation shown In an evaluation of children with VCFS, obses-
to truncate the DYT1 (e-Sarcoglycan) locus. sive–compulsive behaviors were reported with the
In another single family with 14 members, 3 indi- 22q11 deletion syndrome although none in this early
viduals with DMS plus OCD (together with diagnoses study received an OCD diagnosis [112]. Ascertain-
of depression) among 6 total had DMS attributed to a ment strategies and diagnostic evaluations have been
truncating mutation within the SGCE gene [95, 97]. quite varied in these VCFS studies, and in one of the
DMS has also been found linked to two other loci few investigations which compared children with
besides DYT1 at 7q21: a 16.9 cM region between VCFS to children with similar cognitive impairments,
DISS1132 and D183843 on 18p11 [102]. However, a a high, but equal number of behavior problems and
direct examination of the SGCE gene in 32 TS patients psychiatric disorders were found in the two groups
with OCD using WAVE DHPLC analysis (plus direct [90]. For OCD specifically, 11% of the children with
sequencing of 5 of the patients) detected no abnor- VCFS received an OCD diagnosis, as did 14% of the
malities in this gene in comparison to 60 CEPH cognitive ability-matched control [90].
controls [103]; nor was OCD increased in mutation In more detailed studies, groups of adult and
carriers of the DYT1 gene among family members of mixed-age clinic subjects meeting diagnostic criteria
dystonia probands [104]. for OCD have been noted in four studies of individ-
Some prior evidence of a dystonia–OCD connection uals with 22q11 deletions, although all but one of
had been suggested on the basis of an elevated frequency of these reports were primarily focused on the evalu-
OCD symptoms measured by the Maudsley Obsessive– ation of schizophrenia or affective disorders [91–93,
Compulsive Disorder Scale and the Yale–Brown 113]. Psychiatric evaluation of 1 cohort of 14 VCSF
Obsessive–Compulsive Scale (YBOCS) in individuals patients over age 15 revealed 4 with schizophrenia or
with two dystonic syndromes, spasmodic torticollis schizo-affective disorder, 2 of whom also had OCD
and blepharospasm [105]. Also, a 14–20% incidence [93]. A similar cohort of patients with VCFS who
of OCD has been found in 2 studies of over 100 were psychiatrically evaluated found that most of the

127
Chapter 11: Genetic contributions to OCD and related disorders

patients shared common mood, anxiety, and obses- Given the current state of the evidence that OCD
sive–compulsive symptoms, but the majority received is a genetically and phenotypically complex and het-
bipolar, schizoaffective, and ADHD diagnoses (64%), erogeneous disorder, it is not surprising that some of
while only 8% met OCD criteria [92]. the most interesting and persuasive animal models
In the only more comprehensive study that used the have focused on specific OCD sub-phenotypes. Sev-
YBOCS scale to evaluate OCD symptomatology in psy- eral recent reviews have been directed to animal
chiatric interviews of a VCSF clinic sample, 33% models of OCD features [27, 117–121] and this
received primary OCD diagnoses [91]. The most review will primarily be directed to the most recent
common comorbid diagnoses in those with OCD in findings reported.
this study were ADHD, simple phobia, and social
phobia. The highly varied forms of psychopathology
found in VCFS probands and the large range of 8–
Over-grooming behavior in mice with
33% occurrence of OCD in the 4 studies with a total SAPAP3 deficiences, and two initial studies
of 120 VCFS probands clearly requires further evalu-
ation. This will be of interest, as the COMT locus
of SAPAP3 in humans with OCD,
previously nominated as an OCD candidate gene, lies trichotillomania, and grooming disorders
within the 22q deletion and several studies found that Knockout of the excitatory glutamate-system-related
the V158M COMT variant was associated with OCD post-synaptic density gene, Sapap3 was recently
and other disorders, as noted above and as recently reported to be associated with excessive facial self-
reviewed [78]. Changes in the forebrain metabolism grooming [122]. Facial hair and some skin were
of dopamine have been shown in individuals with this essentially denuded in mice lacking Sapap3, a
functional COMT V158M variant [114]. Individuals DLGAP3-related gene located on human chromo-
with VCSF and the 22q microdeletion have not yet some 1p35 that encodes a protein localized to den-
been evaluated in similar magnetic resonance imaging drites and cell bodies and is enriched in synaptosomal
(MRI) studies, or positron emission tomography membrane fractions. Rescue of this prominent
(PET) studies used in these studies of COMT variants. phenotype was essentially complete following chronic
Study of such variants may provide clues about more treatment with the serotonin reuptake inhibitor (SRI),
common variants nearby. The study of rare variants fluoxetine, as well as by replacement expression of
represents a complementary approach to the identifi- SAPAP3 in the striatum using a lentiviral vector.
cation of disease genes and ultimately drug targets In follow-up studies in humans, resequencing of
for disease interventions, as increasingly discussed SAPAP3 was carried out in individuals with OCD
[115, 116]. alone, with OCD plus trichotillomania (TTM, pri-
marily characterized by self-removal of body hair),
TTM together with a control group [123]. The major
Animal models of OCD-like features findings were an excess of rare SAPAP3 variants
that suggestively implicate specific (Table 11.2), some with estimated likelihood of dam-
aging effects, in the OCD/TTM groups from the two
genes in behavior sites. The incidence of these rare variants in cases was
Many models of behavioral changes resembling 2.1% (7/330) versus 0.56% (2/356) in controls. As
OCD-related features such as perseveration, “compul- noted in Figure 11.3b, SAPAP3 SNPs are highly con-
sive” grooming, food-restriction-induced compulsive served across mammalian species including the rare
wheel running, or drinking have been reported [27, variant SNPS found in greater proportions in patients
55, 117–121]. However, despite “face validity”, and in than controls. It is noteworthy that conservation
some instances, apparent validity based upon bio- of these SNPs is  97% identical among specific
logical neurocircuitry or drug response similarities amino acids.
to human OCD, relatively few of these models have A second study of SAPAP3 in sibling pairs with
implicated specific genes. Most of the models which “grooming disorders” found that 32% of 1638 study
have done so have utilized knockout or transgenic participants (65% of whom met lifetime criteria for
mouse models or evaluations of behavioral features OCD) met defined criteria for “grooming disorders”
in mice of different background strains. [124]. Six SNPs, constituting three haplotype blocks,

128
Chapter 11: Genetic contributions to OCD and related disorders

Table 11.2 Identified rare variants in SAPAP3 and predicted functional relevance.

Detected nonsynonymous Analyzed samples Prediction algorithms


variants
121 Duke TTM 44 NIMH 48 130 PMuta PolyPhena
and NIMH OCD wlo NIMH Duke
OCD w. TTM TTM controls controls
R13C; c.38C>T 1 0 0 0 Pathological Possibly
damaging
A148insGPAGA 0 1 0 0 NAb NAb
c.441_442insGGGCCAGCAGGGGCA
T156M; c.467C>T 0 0 0 1 Neutral Benign
A189V; c.566C>T 1 1 0 1 Neutral Benign
T523K; c.1569-70CC>AA 1 0 0 0 Pathological Possibly
damaging
P606T; c.1816C>A 0 1 0 0 Neutral Possibly
damaging
K910R; c.2728A>G 1 0 0 0 Neutral Benign
Combined allele frequencies 7/330TTM and OCD (2.1%) 2/356 controls
(0.56%)
a
Pmut: http://mmb2.pcb.ub.es:8080/Pmut; PolyPhen: http://genetics.bwh.harvard.edu/pph.
b
Prediction of the effects of in/dels is not possible with PMut and PolyPhen, but a functional effects is likely for a five amino acid insertion.
Reproduced from [141] with permission from Nature Publishing group.
Abbreviations: NA, not applicable; NIMH, National Institute of Mental Health; OCD, obsessive–compulsive disorder; TTM, trichotillomania.

(a)
A189V

Variations in OCD/ R13C A148insGPAGA T523K P606T K910R


TTm patients

Guanylate-Kinase-associated protein (GKAP) domain

Variations in
T156M A189V
controls subjects

(b) Human MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPEPRSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP


Rhesus MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPEPRSE SHSLBAPGKRDY OLPLLATPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Cow MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPEARSE SHSLBAPGKRDY OLPLLAAPASVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Horse MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPEARSE SHSLBAPGKRDY OLPLLAAPASVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Mouse MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGFGPGP--GSGAAPEARSE SHSLBAPGKRDY OLPLLAAPASVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Rat MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGFGPGP--GSGAAPEARSE SHSLBAPGKRDY OLPLLAAPASVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP

R13C A148insGPAGA T156M A189V T523K P606T K910R

Figure 11.3 Identified rare nonsynonymous polymorphisms in synapse-associated protein 90/postsynaptic density-95-associated protein 3
(SAPAP3). (a) Schematic of SAPAP3, which consists of 10 coding exons (boxes). Seven rare changes were identified in trichotillomania and
obsessive–compulsive disorder (OCD) patients (a), but only two in controls (b). Most mutations fell into exon 1; however, three changes
affected the conserved guanylate-kinase-associated protein (GKAP) domain. (b) SAPAP3 is highly conserved between species ( 97% identical
amino acids between human and mouse). Accordingly the identified rare changes affected conserved residues. See plate section for color
version.

129
Chapter 11: Genetic contributions to OCD and related disorders

were examined. OCD diagnosis itself was not signifi- highest importance in the first genome-wide scan
cantly associated with any of the six SNPs or three [23]) was supported by a study using more dense
haplotypes, but the “grooming disorder” subgroup targeted markers [24]. This region includes SLC1A1,
showed a nominal association of at least one thus far the most-replicated candidate gene in OCD,
“grooming disorder” (p < 0.05) with four of the six with positive findings in five studies. A single major
SNPs genotyped and with all three haplotypes [124]. functional gene variant has not yet been identified,
These results were not corrected for multiple testing although some of the SNPs associated with OCD
and must be considered preliminary. However, rela- show evidence of functional changes in HapMap gene
tively few groups studying OCD have been examining expression data, in human brain expression data, and
the “grooming disorder” subgroup. in studies of luciferase-measured gene expression in
cultured cell systems [50].
Environmental contributions to OCD, Several other uncommon, less well quantitated

including traumatic life events and


genetic variations occur with an OCD phenotype,
including chromosomal anomalies and some other
possible contributions from immune rare gene variants, such as SLC6A4 I425V, SGCE,
system challenges such as streptococcal GCH1, and SAPAP3. A tentative conclusion is that
OCD resembles other complex disorders in being
infection as postulated in the PANDAS etiologically heterogeneous and in having both highly
syndrome penetrant familial subtypes associated with rare alleles
or chromosomal anomalies [115, 116, 132], as well as
Environmental and gene  environment having relatively more common, polygenic contribu-
interactional contributions to OCD tions that may involve polymorphisms in such genes
as SLC1A1, SLC6A4, BDNF, and perhaps COMT,
There is some evidence for contributions to OCD
DRD4, and HTR2A.
onset, OCD severity, and other features of OCD from
OCD seems likely to be genotypically and pheno-
diverse environmental factors, from psychological
typically heterogeneous. Phenotypic variables such as
trauma, head trauma, and autoimmune reactions
compulsive hoarding behavior, age of onset, associ-
[125–130]. Interpretation on these studies is limited
ated comorbid disorders, and gender, have shown
by study design differences, small subject numbers,
influences in genome-wide linkage analyses and other
and the paucity of studies examining interactions
studies. This has primarily been due to the large-scale
between any of these factors and gene variants.
quantitative phenotypical variable assessment built
into the second GWAS of OCD and other studies
Comments and conclusions from the OCD Collaborative Genetics Study (OCGS)
Except for candidate gene studies, OCD has been group [28, 131, 133–137] and also in the pending
neglected as a disorder of interest to geneticists until GWAS study from another consortium [31].
recently. Although it is a chronic, only partially treat- Compulsive hoarding behavior as a subtype of
able disorder of higher frequency than schizophrenia OCD or an entity of its own has emerged as a major
and of similar frequency as bipolar disorder (both of target in genetic studies [28, 135, 138, 139]. An earlier
which have been the object of 20+ genome-wide study also found important genome-wide linkage
scans, recent GWAS reports, and many other studies regions associated with hoarding in a TS cohort
of twins, families, and gene-associated neurobio- [140]. Compulsive hoarding and related perseverative
logical studies), OCD has only very recently been behaviors have also been found in other human dis-
the subject of three small genome-wide linkage stud- orders with repetitive behaviors or obsessions
ies. One of these was based on only seven families, the (autism, schizophrenia, “punding” in Parkinson’s dis-
second was primarily a sibling-pair based study, and ease, “obsessive flashbacks” in post-traumatic stress
the third was an extension of the first genome-wide disorder) as well as in animal models of compulsive
linkage study [23, 26, 131]. A first GWAS of OCD behaviors such as over-grooming [122, 124, 141],
probands has been accomplished, but as of 2009, only canine acral lick and other behaviors [142], feline
preliminary analyses and results have been reported wool chewing, and additional behaviors in other
in abstract form [31]. However, the 9p24 region (of species [143–145]. There are many potentially fruitful

130
Chapter 11: Genetic contributions to OCD and related disorders

network and pathway hypotheses, including those traumatic events such as drug exposure in utero,
involving serotonergic, dopaminergic, glutamatergic streptococcal infections, and personal traumatic events.
neuropeptide, neurodevelopmental, as well as genes
related to frequent OCD–comorbid disorders. We
may most likely anticipate a combination of rare Acknowledgements
genes ( 0.5%), together with a more extensive poly- This research was supported by the Intramural
genic group contributing to the OCD subphenotypes. Research Program of the National Institute of Mental
Other phenotypic subgroups such as those associated Health (NIMH), National Institutes of Health (NIH).
with comorbid disorders may be partially genetically The authors are grateful to Theresa B. DeGuzman for
based. Among these, gene variants may interact with her editorial and artwork assistance.

References 17. Mcguffin P, et al. Br J Psychiatry


1980;137:285–287.
34. Shugart YY, et al. Am J Med Genet
B Neuropsychiatr Genet
1. Kessler RC, et al. Arch Gen 2009;150B(6):886–892.
Psychiatry 2005;62:593–602. 18. Woodruff R, et al. Am J Psychiatry
1964;120:1075–1080. 35. Stewart SE, et al. Am J Med Genet
2. Bellodi L, et al. Psychiatry Res B Neuropsychiatr Genet 2007;
1992;42:111–120. 19. Inouye E. Am J Psychiatry
1965;121:1171–1175. 144B:1027–1033.
3. Lenane MC, et al. J Am Acad Child 36. Wendland JR, et al. Arch Gen
Adolesc Psychiatry 1990; 20. Carey G, et al. In Klein DF et al.
(eds.). Anxiety: New Research and Psychiatry 2009;66:408–416.
29:407–412.
Changing Concepts. New York: 37. Arnold PD, et al. Psychiatry Res
4. Pauls DL, et al. Am J Psychiatry
Raven Press; 1981. 2009;172:136–139.
1995;152:76–84.
21. Jonnal AH, et al. Am J Med Genet 38. Sampaio AS, et al. CNS Neurosci
5. Riddle MA, et al. J Am Acad
2000;96:791–796. Ther 2011;17(3):141–147.
Child Adolesc Psychiatry
1990;29:45–48. 22. Van Grootheest DS, et al. Twin 39. Moore GJ, et al. J Am Acad Child
Res Hum Genet 2005;8:450–458. Adolesc Psychiatry 1998;
6. Black DW, et al. Arch Gen
23. Hanna GL, et al. Am J Med Genet 37:663–667.
Psychiatry 1992;49:362–368.
2002;114:541–552. 40. Coric V, et al. Biol Psychiatry
7. Hettema JM, et al. Am J
24. Willour VL, et al. Am J Hum 2005;58:424–428.
Psychiatry 2001;158:
1568–1578. Genet 2004;75:508–513. 41. Grant P, et al. J Child Adolesc
25. Hanna GL, et al. Biol Psychiatry Psychopharmacol 2007;
8. Nestadt G, et al. Arch Gen
2007;62:856–862. 17:761–767.
Psychiatry 2000;57:358–363.
26. Shugart YY, et al. Mol Psychiatry 42. Schwenkreis P, et al. Exp Brain Res
9. Cavallini MC, et al. Am J Med
2006;11:763–770. 2000;135:293–299.
Genet 1999;88:38–43.
27. Wang Y, et al. Am J Hum Genet 43. Bengel D, et al. Mol Psychiatry
10. Hanna GL, et al. Am J Med Genet
2009;84:52–59. 1999;4:463–466.
B Neuropsychiatr Genet 2005;
134B:13–19. 28. Samuels J, et al. Am J Psychiatry 44. Lesch K-P, et al. Science 1996;
2007;164:493–499. 274:1527–1531.
11. Nestadt G, et al. Am J Hum Genet
2000;67:1611–1616. 29. Sklar P, et al. Mol Psychiatry 45. Mcdougle CJ, et al. Arch Gen
12. Nicolini H, et al. Ursus Medicus 2008;13:558–569. Psychiatry 1994;51:302–308.
1991;1:25–28. 30. The Psychiatric GWAS 46. Bloch MH, et al. Am J Med Genet
13. Alsobrook JP, et al. Am J Med Consortium Steering (TPGCS) B Neuropsychiatr Genet
Genet B Neuropsychiatr Genet Committee. Mol Psychiatry 2008;147B:850–858.
1999;88:669–675. 2009;14:10–17. 47. Dickel DE, et al. Biol Psychiatry
14. Eapen V, et al. J Psychosom Res 31. Stewart E, et al. ACNP Meeting 2007;61:322–329.
2006;61:359–364. Miami, FL; 2010. 48. Hu XZ, et al. Am J Hum Genet
15. Cryan EM, et al. Br J Psychiatry 32. Arnold PD, et al. Arch Gen 2006;78:815–826.
1992;161:694–698. Psychiatry 2006;63:769–776. 49. Wendland JR, et al.
16. Marks IM, et al. Br J Psychiatry 33. Dickel DE, et al. Arch Gen Neuropsychopharmacology
1969;115:991–998. Psychiatry 2006;63:778–785. 2007;32(12):2543–2551.

131
Chapter 11: Genetic contributions to OCD and related disorders

50. Wendland JR, et al. Psychiatr 73. Hemmings SM, et al. Eur Neuro- 94. Doheny D, et al. Neurology 2002;
Genet 2008;18:31–39. psychopharmacol 2003;13: 59:1244–1246.
51. Vallender EJ, et al. Genes Brain 93–98. 95. Marechal L, et al. Am J Med Genet
Behav 2008;7:690–697. 74. Millet B, et al. Am J Med Genet B Neuropsychiatr Genet 2003;
B Neuropsychiatr Genet 119:114–117.
52. Ozaki N, et al. Mol Psychiatry
2003;116:55–59. 96. Saunders-Pullman R, et al.
2003;8(895):933–936.
75. Nicolini H, et al. Mol Psychiatry Neurology 2002;58:242–245.
53. Delorme R, et al. Mol Psychiatry
1996;1:461–465. 97. Zimprich A, et al. Nat Genet
2005;10(12):1059–1061.
76. Walitza S, et al. J Neural Transm 2001;29:66–69.
54. Kilic F, et al. Mol Pharmacol
2008;115:1071–1078. 98. Boghosian-Sell L, et al. Am J Hum
2003;64:440–446.
77. Alsobrook JP, 2nd, et al. Am J Med Genet 1996;59:999–1005.
55. Altemus M, et al. Am J Med Genet Genet 2002;114:116–120.
1996;67:409–411. 99. Petek E, et al. Am J Hum Genet
78. Azzam A, et al. Am J Med Genet 2001;68:848–858.
56. Insel TR, et al. Biol Psychiatry B Neuropsychiatr Genet 100. Verkerk AJ, et al. Genomics 2003;
1985;20:1174–1188. 2003;123:64–69. 82:1–9.
57. Mcdougle CJ, et al. Mol Psychiatry 79. Erdal ME, et al. Depress Anxiety 101. Diaz-Anzaldua A, et al. Am J Med
1998;3:270–273. 2003;18:41–45. Genet A 2004;127:17–20.
58. Murphy DL, et al. Mol Interv 80. Niehaus DJ, et al. J Affect Disord 102. Grimes DA, et al. Neurology
2004;4:109–123. 2001;65:61–65. 2002;59:1183–1186.
59. Torres GE, et al. Mol Pharmacol 81. Schindler KM, et al. Am J Med 103. De Carvalho Aguiar P, et al. Mov
2003;64:196–198. Genet 2000;96:721–724. Disord 2004;19:1237–1238.
60. Fox MA, et al. Psycho- 82. Camarena B, et al. Int J 104. Heiman GA, et al. Am J Med
pharmacology (Berl) Neuropsychopharmacol Genet Part B Neuropsychiatr
2008;201:203–218. 2001;4:269–272. Genet 2007;144B(3):361–364.
61. Holmes A. Neurosci Biobehav Rev 83. Karayiorgou M, et al. Biol 105. Bihari K, et al. Psychiatry Res
2008;32:1293–1314. Psychiatry 1999;45: 1992;42:267–272.
1178–1189.
62. Jennings KA, et al. J Neurosci 106. Cavallaro R, et al. Biol Psychiatry
2006;26:8955–8964. 84. Hall D, et al. Am J Hum Genet 2002;52:356–361.
2003;73:370–376.
63. Murphy DL, et al. Neuro- 107. Voon V, et al. Mov Disord
pharmacology 2008;55:932–960. 85. Chen SK, et al. Cell 2010;
2010;25:2249–2252.
141:775–785.
64. Murphy DL, et al. Nat 108. Greenberg BD, et al. Am J Med
Rev Neurosci 2008;9: 86. Greer JM, et al. Neuron 2002;
Genet 2000;96:202–216.
85–96. 33:23–34.
109. Rauch SL, et al. Adv Neurol
65. Homberg J, et al. BMC Genet 87. Atmaca M, et al. J Affect Disord
2001;85:207–224.
2010;11:37. 2010;123:258–263.
110. Lerner A, et al. Cereb Cortex
88. Zai G, et al. Am J Med Genet
66. Homberg JR, et al. Neuroscience 2009;19:218–223.
B Neuropsychiatr Genet
2007;146:1662–1676. 111. Antshel KM, et al. Neuropsychol
2004;129B:64–68.
67. Greenberg BD, et al. Am J Med Dev Cogn C Child Neuropsychol
89. Gratacos M, et al. Am J Med Genet
Genet 1999;88:83–87. 2005;11:5–19.
B Neuropsychiatr Genet 2009;
68. Hariri AR, et al. Arch Gen 150B:808–816. 112. Eichstedt JA, et al. Clin Psychol
Psychiatry 2005;62:146–152. Rev 2001;21:137–157.
90. Feinstein C, et al. Biol Psychiatry
69. Lothe A, et al. Neuroimage 2002;51:312–318. 113. Fisher SE, et al. Am J Hum Genet
2009;47:482–492. 2002;70:1183–1196.
91. Gothelf D, et al. Am J Med Genet
70. Billett EA, et al. Psychiatr Genet B Neuropsychiatr Genet 2004; 114. Meyer-Lindenberg A, et al. Nat
1998;8:163–169. 126:99–105. Neurosci 2005;8:594–596.
71. Catalano M, et al. Am J Med Genet 92. Papolos DF, et al. Am J Psychiatry 115. Asimit J, et al. Annu Rev Genet
1994;54:253–255. 1996;153:1541–1547. 2010;44:293–308.
72. Cruz C, et al. Neurosci Lett 93. Pulver AE, et al. J Nerv Ment Dis 116. Cirulli ET, et al. Nat Rev Genet
1997;231:1–4. 1994;182:476–478. 2010;11:415–425.

132
Chapter 11: Genetic contributions to OCD and related disorders

117. Joel D. Prog Neuro- 126. Snider LA, et al. Mol Psychiatry 136. Samuels JF, et al. Behav Res Ther
psychopharmacol Biol Psychiatry 2004;9:900–907. 2008;46:836–844.
2006;30:374–388. 127. Gershuny BS, et al. Depress 137. Samuels JF, et al. Behav Res Ther
118. Joel D, et al. Behav Brain Res Anxiety 2008;25:69–71. 2008;46:1040–1046.
2005;157:253–263.
128. Gershuny BS, et al. Behav Res Ther 138. Lasalle VH, et al. Depress Anxiety
119. Korff S, et al. Psychiatr Clin North 2003;41:1029–1041. 2004;19:163–173.
Am 2006;29:371–390.
129. Cromer KR, et al. Behav Res Ther 139. Wheaton M, et al. J Anxiety Disord
120. Leonard HL, et al. Int J 2007;45:1683–1691. 2008;22:243–252.
Neuropsychopharmacol
2001;4:191–198. 130. Sasson Y, et al. Psychiatry Res 140. Zhang H, et al. Am J Hum Genet
2005;135:145–152. 2002;70:896–904.
121. Wang L, et al. Behav Pharmacol
2009;20:119–133. 131. Samuels JF, et al. Am J Med Genet 141. Zuchner S, et al. Mol Psychiatry
B Neuropsychiatr Genet 2006; 2009;14:6–9.
122. Welch JM, et al. Nature
141B:201–207.
2007;448:894–900. 142. Dodman NH, et al. Mol Psychiatry
132. Gorlov IP, et al. Am J Hum Genet 2010;15:8–10.
123. Zuchner S, et al. Ann Hum Genet
2008;82:100–112.
2008;72:725–731. 143. Luescher UA, et al. Vet Clin North
133. Hasler G, et al. Biol Psychiatry Am Small Anim Pract 1991;
124. Bienvenu OJ, et al. Am J
Med Genet B Neuropsychiatr
2007;61:617–625. 21:401–413.
Genet 2008;150B:710–720. 134. Samuels J, et al. Behav Res Ther 144. Rapoport JL, et al. Arch Gen
125. Grisham JR, et al. Eur Arch 2002;40:517–528. Psychiatry 1992;49:517–521.
Psychiatry Clin Neurosci 135. Samuels JF, et al. Behav Res Ther 145. Voon V. Mov Disord 2004;
2008;258:107–116. 2007;45:673–686. 19:367–370.

133
Post-traumatic stress disorder
Chapter

12 Michael J. Lyons, Tyler Zink, and Karestan C. Koenen

The “diathesis–stress” model [1] has been an influen-


tial approach for conceptualizing mental disorders.
History of the construct
It reconciles the roles of nature and nurture in the The definition of PTSD as a mental disorder, how-
etiology of a disorder. “Diathesis” refers to a predis- ever, has evolved over time. While ideas about the
position, tendency, or pre-existing vulnerability to negative psychological impact of trauma have been
develop a disorder when the individual is exposed to accepted for centuries [3], PTSD did not have a
an environmental stressor. Post-traumatic stress dis- formal psychiatric definition until 1980 in DSM-III
order (PTSD) is a quintessential example of a diathesis– [4]. Inclusion of PTSD in DSM-III marked formal
stress phenomenon because it describes a process in acknowledgement from the psychiatric community
which an individual reacts to an environmental stres- that the effects of trauma should be considered from
sor (the “trauma”) by developing a mental disorder a mental health perspective, that inherent personal
(PTSD). This chapter will provide some background weakness does not drive traumatic sequelae, and that
about PTSD and review the evidence for the role of the negative experiences of those having suffered
genetic factors as a diathesis for PTSD. trauma are legitimate [5]. However, before this shift,
recognition of the psychological importance of trauma
was not universal. In fact, psychological sequelae of
Overview of diagnosis traumatic events only became an important focus of
According to the DSM-IV-TR, PTSD is an anxiety medicine during the American Civil War. Pizarro et al.
disorder characterized by a series of symptoms dir- [6] reported that 44% of soldiers reported signs of
ectly related to the experience of a traumatic event [2]. mental or “nervous” disease after the Civil War, which
In fact, PTSD is only one of two mental disorders in was often called “irritable heart” by nineteenth-century
DSM-IV-TR for which a precipitating event is neces- physicians. Also, individuals suffering from mental and
sary, although not sufficient, for diagnosis. In order physical reactions to train accidents spurred the notion
for an event to be considered a potential antecedent to of “railway spine” or “postconcussion syndrome” ([79],
PTSD, the individual must: (1) experience, witness, or as cited in [5]). Considering that the symptoms of
be confronted by actual or threatened death or serious these ailments included sleep disruptions, nightmares
injury or a threat to the physical integrity of the self or about train accidents, avoidance of train travel, and
others; and (2) experience intense fear, helplessness, chronic pain, Oppenheim renamed the phenomenon
or horror as a result of the traumatic event. Once such “traumatic neurosis” ([78] as cited in [5]).
a traumatic event has been established as a possible Reference to psychological reaction to trauma
antecedent, an individual must re-experience the event, appeared in DSM-I under the label of traumatic neur-
avoid stimuli related to it, and experience increased osis [7]. However, this disorder was not present in
arousal as a result of it in order to qualify for a DSM-II [8]. The understanding of psychological reac-
diagnosis of PTSD. Additionally, such symptoms must tion to traumatic events underwent a metamorphosis
persist for at least one month and significantly impair throughout this time period, leading up to the inclu-
daily functioning [2]. sion of PTSD as a mental disorder in DSM-III [4].

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

134
Chapter 12: Post-traumatic stress disorder

Before World War I, traumatic neurosis was explored estimates than later studies because the DSM-III and
through the uncovering of dissociated experiences via DSM-III-R employed a more narrow definition of
hypnosis [9]. Ideas developed from Charles Samuel trauma exposure than that of the DSM-IV. Later
Myers’s concept of World War I trench warfare vet- surveys also queried participants about a wider
erans experiencing “shell shock” to Abram Kardiner’s range of traumatic events using interview probes
notion of “war neurosis”, which was lumped together designed to increase the likelihood that respondents
with hysteria under the general umbrella of “psychic would provide sensitive and personal information.
trauma” from nonspecific damage to the nervous For example, in the National Comorbidity Survey-
system [10]. The return of Vietnam War veterans to Replication, participants were specifically asked
the United States and the feminist movement of the whether they were physically abused as children. This
1960s, served as an impetus to combine the under- question was not asked of participants in the ECA
standing of psychological reaction to traumatic events survey [12, 28].
across several categories (combat exposure, sexual Because individuals exposed to trauma do not
abuse, natural disaster survival, and so forth) for always develop PTSD, it is important to identify
inclusion as one disorder in DSM-III [11]. factors associated with the probability of developing
PTSD. Many researchers have reported that more
Epidemiology of trauma exposure women receive a diagnosis of PTSD than do men
[12, 29, 30]. A recent meta-analysis [31] suggested
While PTSD earned formal psychiatric recognition
that women are at greater risk for developing PTSD
due to the large number of suffering war veterans
across all types of trauma except for assaultive vio-
[5, 11], many different forms of trauma affecting
lence. However, this difference may reflect gender
different groups of people have been demonstrated
differences in type of trauma exposure. Specifically,
to be antecedents to the development of PTSD [2].
women have a higher conditional risk for developing
Given that researchers differ in their methods of
PTSD because of their increased exposure to assault-
assessing traumatic events and the definition of what
ive violence [29, 30]. Some of the other factors con-
constitutes a traumatic event has changed with each
tributing to the development of PTSD among those
revision of the DSM [12], there is a wide range of
exposed to trauma include being physically wounded
estimates for the prevalence of trauma exposure. For
during the event [15], having lower cognitive ability
example, some researchers combine all combat
[32–34], having behavioral problems in adolescence
related traumatic events [13, 14], while others specify
[15, 33, 35], being of lower socioeconomic status [14],
two types [15] or three types [16] of combat-related
being sexually assaulted by an acquaintance [36, 37],
PTSD antecedents. Keeping potential definitional and
being high in trait neuroticism or introversion [38],
methodological differences in mind, estimates of the
living in a country in the midst of political or ethnic
prevalence of exposure to traumatic events typically
violence [28], and prior trauma exposure [39, 40].
range from 40% (using DSM-III/III-R criteria for
trauma exposure) to 85% (using DSM-IV criteria for
trauma exposure) in community samples [17–22]. More
recent studies using DSM-IV criteria show much higher Genetic studies of PTSD
estimates of trauma exposure [23, 24]. Quantitative genetic studies
There are three approaches that have traditionally
Epidemiology of diagnosis been used to investigate the presence of genetic
Many individuals experience traumatic events, how- influences on psychopathology: family studies, twin
ever, not all develop PTSD. In fact, as compared to studies, and adoption studies. In general, the logic of
estimates of trauma exposure, PTSD develops in only a family study is that if a disorder is influenced by
about 2–15% of the general population [15, 17, 20, 22, genetic factors, relatives who share their genes with an
25–27]. This wide range of estimates reflects differ- individual with the disorder of interest (a “proband”)
ences in methodology and changes in diagnostic should have a higher risk for the disorder than indi-
criteria. Earlier studies, including the Epidemiologic viduals who are not related to a proband. Davidson
Catchment Area (ECA) study and those using similar et al. [41] investigated female rape victims and found
methodology, typically provided lower prevalence that subjects with a family history of major depression

135
Chapter 12: Post-traumatic stress disorder

were more likely to develop PTSD following rape than susceptibility to PTSD symptoms. In another twin
individuals without such a family history. Yehuda and study, Stein et al. [45] reported heritabilities ranging
her colleagues [42, 43] have investigated the relation- from 28 to 38% for the individual PTSD symptom
ship of PTSD in Holocaust survivors and PTSD clusters and no influence from environmental factors
in their offspring. In one study [42] they found that shared by twins. Jang et al. [46] conducted a twin study
offspring who had experienced traumatic events investigating genetic and environmental contributions
themselves were more likely to develop PTSD if their to susceptibility for developing post-traumatic stress
Holocaust-surviving parents had chronic PTSD. In a symptoms (PTSS) following exposure to traumatic
second study that used some of the subjects from their events. They studied over 400 twin pairs and deter-
1998 study, but added additional subjects, Yehuda mined lifetime frequency of exposure to assaultive and
et al. [43] found that the offspring of Holocaust sur- nonassaultive trauma. They found that in the case
vivors with PTSD had a higher risk of PTSD than of nonassaultive trauma, PTSS were directly affected
offspring of Holocaust survivors without PTSD. They by environmental factors that also influence exposure
also observed that maternal PTSD was a stronger to nonassaultive trauma. Both genetic and nonshared
predictor of offspring PTSD that paternal PTSD. environmental influences jointly affected PTSS for
These observations are consistent with a familially assaultive trauma, and the number of traumatic events
transmitted genetic vulnerability to the development moderated the severity of PTSS. The influence of gen-
of PTSD following exposure to trauma, but are etic factors became less important beyond some
equally compatible with a socially transmitted vulner- threshold number of traumatic events.
ability. A weakness of family studies is the inability Adoption studies are another method for disen-
to distinguish between genetic and environmental tangling genetic factors from family environmental
factors as the mechanism of transmission. factors, but we know of no adoption studies of PTSD.
Twin studies can avoid a shortcoming of family One of the factors that contributes to the complexity
studies by distinguishing genetic influences from of conducting research on genetic factors in PTSD is
influences due to the family environment. This is the role that trauma exposure plays in the disorder.
accomplished by capitalizing on a naturally occurring Exposure to a qualifying trauma is necessary, but not
experiment, that is, the difference in the genetic simi- sufficient for the diagnosis of PTSD. Therefore, it is
larity between monozygotic (MZ) and dizygotic (DZ) not possible to assess an individual’s vulnerability to
twins. MZ twins share 100% of their genes, while DZ PTSD in the absence of exposure to trauma. These
twins share, on average, 50%. To the extent that MZ considerations make the logistics of conducting an
twins are more similar to one another on some trait, it adoption study of PTSD extremely difficult.
can be inferred that genetic factors have some influ-
ence on the trait. Twin studies can decompose the
sources of individual differences into those that reflect Genetic influence in exposure to trauma
genetic differences among individuals, those that Stein and colleagues [45], using the same sample
reflect the aspects of the environment that are shared utilized in the report by Jang et al. [46] described
by twins, such as neighborhood and parental educa- above, found that exposure to assaultive trauma, such
tion, and environmental features that are not shared as robbery and sexual assault, was influenced signifi-
by both members of a twin pair. cantly by genetic factors, but not by shared environ-
True and colleagues [44] studied over 4000 pairs mental factors. Interestingly, exposure to nonassaultive
of male twins who had served in the US military trauma, such as natural disasters and motor vehicle
between 1965 and 1975 comprising the Vietnam Era accidents, was significantly influenced by the shared
Twin Registry. Data on PTSD symptomatology were environment, but not genetic factors. They also found
collected by questionnaire. Genetic factors contrib- that to some extent genetic factors that influenced
uted to susceptibility for nearly all symptoms of PTSD exposure to assaultive traumatic events also influ-
with heritabilities ranging from 13 to 34%. Combat enced vulnerability to PTSD symptoms. Certainly
exposure was a strong predictor only of the many traumatic events are fateful and befall an indi-
“re-experiencing” symptom cluster and the “avoided vidual regardless of his or her behavior. However, it
activities” symptom. There was no detectable influ- may be that personality traits, such as high sensation
ence from environmental factors shared by twins on seeking and/or low harm avoidance, make it more

136
Chapter 12: Post-traumatic stress disorder

likely that an individual will seek out environments in reasonable candidates for influencing the risk of
which he or she will be exposed to trauma. This developing PTSD following exposure to trauma.
reflects the phenomenon of “gene–environment cor- The most common form of association study is
relation” in which to some extent an individual’s very similar to the traditional case-control study.
genetically influenced characteristics affect the type A sample of unrelated individuals with the disorder
of environment that the individual seeks out. of interest is compared to a matched control sample of
Lyons and colleagues [47] studied members of unrelated individuals without the disorder in terms of
the Vietnam Era Twin Registry to examine genetic the presence of a risk factor. For example, a sample
and nongenetic factors that influence wartime expos- of individuals with emphysema (cases) might be com-
ure to traumatic events. Specific events examined pared to a matched sample of individuals without
were volunteering for service in Vietnam, actual ser- emphysema (controls) for the putative risk factor
vice in Southeast Asia, a composite index of 18 of smoking. If a significantly greater proportion of
combat experiences, and information from military individuals with emphysema were found to be
records about being awarded combat decorations. smokers than individuals without emphysema, then
There was a significant genetic influence on volun- an association between smoking and emphysema
teering for service in Vietnam. Among twin pairs in would be inferred. In the genetic case-control associ-
which both siblings served in Southeast Asia there ation study, samples of cases and controls are estab-
was a significant genetic influence on self-reported lished and the risk factor that is examined is a genetic
combat experiences. The family environment did not polymorphism. Genetic polymorphisms refer to two
have a significant effect on any of the variables. or more genetic markers or alleles at the same genetic
Analyses of data from military records regarding locus, each of which occurs more often than could
being awarded a combat decoration provided very be explained by mutation. The basic logic of the case-
similar results to those found for self-reported control candidate gene study is that if it is determined
combat experiences. that one (or more) of the alleles is found more fre-
quently among cases than it is among controls, that
allele is associated with the phenotype. There are
Molecular genetic studies a number of factors that go into determining the
There have been a number of case-control association statistical power (the probability of detecting a true
studies conducted to try to identify genetic variants genetic effect) of association studies. As with other
that are related to a vulnerability to develop PTSD types of studies, the size of the sample and the thresh-
following exposure to trauma. The strategy that has old set for significance influence the probability of
been used thus far in this research has been the detecting a significant effect. The power is also influ-
“candidate gene” approach because of the very large enced by the prevalence of the “risk” genotype and
number of genes in the human genome and the tech- its effect size (i.e. how much is risk of the disorder
nical limitations that existed until recently in terms of increased in the presence of the specified marker).
genotyping very large numbers of markers simultan- Table 12.1 includes information about the 18 asso-
eously. In this approach genes are selected for interro- ciation studies of PTSD of which we are aware
gation based on the plausibility that they play a role in (although several of these studies investigated clinical
the phenotype of interest. For behavioral and psychi- features of subjects with PTSD, rather than the risk of
atric phenotypes, such as PTSD, the genes with the PTSD, per se). Eight of the studies investigated genes
greatest biological plausibility are those that operate in the dopaminergic system; six looked at the D2
in the central nervous system. However, there are dopamine receptor gene, one at the dopamine trans-
thousands of genes that influence the brain, so it has porter gene, and one at the dopamine beta-hydroxylase
been helpful to narrow the field of candidates further. (DBH) gene. Comings et al. [48] reported a greater
Information that we know about the neurobiology of frequency of the A1 allele of the D2 dopamine recep-
PTSD can inform decisions about where to look for tor among 35 patients with PTSD. However, after
relevant genes. Genes that influence the hypothalamic- correcting for multiple comparisons, this was not
pituitary-adrenal (HPA) axis, genes that influence the significant. In 1996, Comings et al. [49] reported the
locus cerulius/noradrenergic system, and genes that results from a study of Vietnam veterans who were
influence limbic-frontal brain systems are all very patients on an addiction treatment unit. There was

137
Chapter 12: Post-traumatic stress disorder

Table 12.1 Candidate gene association studies of post-traumatic stress disorder (PTSD).

First Year Trauma Trauma Gene name Finding


author exposed type
controls?
Comings 1991 No Combat Dopamine receptor Excess D2A1 allele in PTSD cases p ¼ 0.007
D2 (DRD2)
Comings 1996 Yes Combat Dopamine receptor (Two samples) Excess D2A1 allele in PTSD
D2 (DRD2) cases p ¼ 0.041
Excess D2A1 allele in PTSD cases p ¼ 0.002
Gelernter 1999 No Combat Dopamine receptor No significant association between D2A1
D2 (DRD2) allele/DRD2 haplotypes and PTSD
Young 2002 No Combat Dopamine receptor Excess D2A1 allele only in PTSD cases with
D2 (DRD2) harmful drinking p < 0.001
Mustapić 2007 Yes Combat Dopamine No main effect for DBH gene on risk of PTSD
β-hydroxylase (DBH)
Segman 2002 Yes Various Dopamine Excess 9-repeat allele in PTSD cases p ¼ 0.012
transporter (DAT1)
Bachman 2005 Yes Combat Glucocorticoid No significant association between GCCR
receptor (GCCR) polymorphisms and PTSD
Lappalainen 2002 No Combat Neuropeptide Y No significant association between Leu7Pro
(NPY) polymorphism and PTSD
Lu 2008 No Various Cannabinoid significant association between a CNR1
receptor (CNR1) haplotype and PTSD, p < 0.04
Freeman 2005 No Combat Apolipoprotein E Association between APOE ε2 and poorer
allele (APOE) memory scores as well as more severe trauma
re-experiencing
Koenen 2005 No Acute Glucocorticoid Two FKBP5 SNPs are significantly associated
injury receptor-regulating with dissociation during (p < 0.05) and after
cochaperone (FKBP5) ( p < 0.03) traumatic accidents in children
Binder 2008 No Various Glucocorticoid Genetic variation in the FKBP5 gene may
receptor-regulating place individuals with significant past child
cochaperone (FKBP5) abuse at significant risk for PTSD, p < 0.0004
Lee 2005 No Various Serotonin transporter Excess s allele in PTSD cases p ¼ 0.04
(SLC6A4)
Kilpatrick 2007 Yes Hurricane Serotonin transporter Significant association between s/s genotype
(SLC6A4) and PTSD in adults with high hurricane
exposure and low social support, p < 0.03
Lee 2006 No Not Brain-derived No association between BDNF Val66Met
specified neurotrophic factor and PTSD
(BDNF)
Zhang 2006 Not Not Brain-derived No significant association between three
specified specified neurotrophic factor BDNF variants and PTSD
(BDNF)

138
Chapter 12: Post-traumatic stress disorder

Table 12.1 (cont.)

First Year Trauma Trauma Gene name Finding


author exposed type
controls?
Lawford 2003 No Combat Dopamine receptor Possible association between D2A1 allele
D2 (DRD2) and response to SSRI paroxetine, p ¼ 0.03
Lawford 2006 No Combat Dopamine receptor DRD2 gene is associated with comorbid
D2 (DRD2) depression, anxiety, and social dysfunction
in untreated veterans with PTSD, p < 0.05
SNPs, single nucleotide polymorphisms.

a significantly greater frequency of the A1 allele of the and Bcll) in 118 Vietnam veterans with PTSD and
D2 dopamine receptor among the 37 subjects with 42 controls who were Vietnam veterans with combat
PTSD compared to the 19 controls who had been exposure, but without PTSD. The frequency of the
exposed to high levels of combat but did not develop N363S and Bcl1 GR polymorphisms were not greater
PTSD. Gelernter et al. [50] attempted to replicate in PTSD patients than controls and did not differ
Comings’ findings using a sample of 52 Vietnam from reported population frequencies.
veterans with PTSD and 87 controls without PTSD. In a study that was focused on alcohol depend-
They found no association between the A1 allele of ence, Lappalainen, Kranzler, and Malison [55]
the D2 dopamine receptor and PTSD. Young et al. included a group (n = 77) of Vietnam-era combat
[51] examined the frequency of the A1 allele among veterans with PTSD who were compared to 202 com-
91 patients with PTSD versus 51 controls and found munity controls. They found that the groups did
that A1 allelic frequency was significantly higher not differ in the frequency of the neuropeptide
among the PTSD patients. They also found that PTSD Y Leu7Pro polymorphism. Lu and colleagues [56]
patients with the A1 allele were more likely to have conducted an association study focused on atten-
alcohol problems than PTSD patients without the tion-deficit hyperactivity disorder (ADHD) that also
A1 allele. Mustapić et al. [52] investigated the DBH assessed PTSD and several other disorders that are
gene which converts dopamine to norepinepharine. potentially comorbid with PTSD. The candidate gene
Among a sample of individuals exposed to combat that they investigated was the cannabinoid receptor
trauma, they did not find a relationship between gene (CNR1). They found a significant association
PTSD and the polymorphism in the DBH gene that between a single nucleotide polymorphism (SNP)
they investigated. There was, however, a trend for an haplotype and PTSD. While suggesting that their
interaction between PTSD and DBH genotype; PTSD finding was preliminary, they concluded that the
subjects with one version of the DBH polymorphism CNR1 gene warrants further investigation for a role
(the CC genotype) had lower levels of plasma DBH in psychiatric disorders, including PTSD.
activity than subjects without PTSD, but in subjects Some studies have looked at the association of
with other genotypes, there was no relationship candidate genes with characteristics among individ-
between PTSD and plasma DBH activity. uals with PTSD, rather than looking at the genotype
In another study of the dopaminergic system, as a risk factor for PTSD. Freeman et al. [57] looked
Segman et al. [53] examined the dopamine trans- at the relationship of APOE genotype (which has been
porter gene (DAT). Specifically, they investigated the implicated in the risk of Alzheimer’s disease) among
association of the nine repeat allele of the variable 54 veterans with PTSD. They found that the APOE
number tandem repeat in the dopamine transporter ε2 allele was associated with significantly worse
gene. There was a significantly higher frequency of re-experiencing symptoms and impaired memory
the 9 repeat allele among the 102 chronic PTSD function in this population. Koenen and colleagues
patients compared to 104 controls who had been [58] examined the FKBP5 gene, which is a gluco-
exposed to trauma, but did not develop PTSD. corticoid receptor-regulating cochaperone of stress
Bachmann et al. [54] investigated two common proteins. They studied 46 children following their
glucocorticoid receptor (GR) polymorphisms (N363S admission to a hospital for an acute medical injury.

139
Chapter 12: Post-traumatic stress disorder

They found that two SNPs in the FKBP5 gene were D2 dopamine receptor genotype on features related to
significantly associated with dissociation during and PTSD. They found that among their subjects with
since the injury. PTSD, those with the A1 allele of the D2 dopamine
Binder et al. [59] reported a cross-sectional study receptor gene had higher scores for anxiety/insomnia,
examining genetic and psychological risk factors social dysfunction, and depression [65]. PTSD sub-
in 900 nonpsychiatric clinic patients (762 included jects with the A1 allele had more severe comorbid
for all genotype studies) with significant levels of psychopathology. In an earlier study, Lawford et al.
childhood abuse as well as nonchild abuse trauma. [64] found that PTSD subjects with the A1 allele of
Eight SNPs spanning the FKBP5 locus were utilized. the D2 dopamine receptor gene showed significant
Although FKBP5 SNPs did not directly predict PTSD improvement in social functioning when treated with
symptom outcome or interact with level of nonchild paroxetine compared to PTSD subjects without the
abuse trauma to predict PTSD symptom severity, four A1 allele.
SNPs in the FKBP5 locus significantly interacted with
the severity of child abuse to predict level of adult
PTSD symptoms.
Genetic influences on comorbidity
Several studies have investigated the serotonin There have been a number of twin studies that have
transporter gene (SLC6A4; SERT). Lee and colleagues examined possible genetic influences that contribute
[60] examined the serotonin-transporter-linked poly- to the comorbidity of PTSD with other mental dis-
morphic region (SERTPR or 5HTTLPR). They stud- orders. Koenen et al. [66] found that a genetic liability
ied 100 PTSD patients and 197 unrelated healthy that influenced both PTSD and major depression was
controls using a case-control design. PTSD patients the primary reason that the two disorders co-occur. In
had a higher frequency of the low expression (s/s) 2003, Koenen and colleagues reported that comorbid-
genotype than normal controls. Kilpatrick et al. [61] ity between PTSD and depression and PTSD and
also investigated whether the same polymorphism dysthymia reflect, in part, genetic factors that impart
influenced the risk of PTSD following exposure to a risk for all three disorders. Xian et al. [67] demon-
a hurricane. Subjects were 589 adults from Florida. strated that genetic factors contributed to the comor-
The low-expression variant of the 5HTTLPR poly- bidity observed among PTSD, depression, and alcohol
morphism was associated with risk for PTSD following dependence. McLeod et al. [68] found that genetic
high exposure to hurricane, but only if there was low factors contribute significantly to the frequently
social support. High risk individuals (high hurricane observed comorbidity between PTSD and alcohol
exposure, the low-expression 5HTTLPR variant, low dependence. Chantarujikapong et al. [69] found that
social support) had 4.5 times the risk of developing some of the genetic vulnerability to PTSD also confers
PTSD compared to low risk individuals. vulnerability to generalized anxiety disorder and
Lee et al. [62] studied the genotype and allele panic disorder. Koenen et al. [58] found that most
frequencies of the brain-derived neurotrophic factor of the association between PTSD and nicotine
(BDNF) gene Val66Met polymorphism in 106 PTSD dependence reflects shared genetic influences, but
patients and 161 unrelated healthy controls using the relationship between late-onset smoking and
a case-control design. The genotype and allele fre- PTSD is not influenced by genetic factors. Overall,
quencies for the BDNF gene polymorphism did not these various studies indicate that at least some of the
differ between the two groups. Zhang et al. [63] also genetic factors that impart a risk for PTSD also
investigated the relationship of the BDNF Val66Met impart some risk for a number of other disorders.
polymorphism to PTSD along with two other BDNF
polymorphisms. They studied 96 patients with PTSD Future directions
(as well as a number of other psychiatric disorders)
and 250 normal controls. No association was found Epigenetics
between any of the three included BDNF gene variants An area that is likely to play an increasingly important
and PTSD. role in understanding how genes influence psycho-
Two studies by Lawford and colleagues [64, 65] pathology in general and PTSD in particular is epige-
did not look at the association between genotype and netics. Epigenetics refers to changes in gene expression
risk for PTSD, but investigated the possible effects of that are not caused by the DNA sequences, but are

140
Chapter 12: Post-traumatic stress disorder

instead caused by other mechanisms. Perhaps more available acetylcholine and depressed cholinergic neu-
than any other psychiatric disorder, PTSD is amenable rotransmission following exposure to stress. They
to this type of approach because trauma might plaus- suggested that while this mechanism might have the
ibly be associated with changing gene expression. beneficial short-term effect of quieting brain activity
At least two studies of PTSD have utilized this following trauma, the long-term effects could be
approach. Segman and colleagues [70] used oligonu- damaging.
cleotide microarrays to measure peripheral blood Several animal model studies have investigated the
mononuclear cell (PBMC) gene expression among promoter polymorphism of the serotonin transporter
survivors of trauma at an emergency room and four (5-HTTLPR) gene that is associated with increased or
months later. The profile of gene expression signa- decreased transcription (several human case-control
tures at both times distinguished survivors with PTSD candidate gene studies of the 5-HTTLPR are
at one and four months, from those who met no described above). Barr et al. [73] examined macaques
criteria for PTSD. Results demonstrated a general who had either been raised with their mothers or in
reduction in PBMC ’s expression of transcription acti- peer-only groups. They examined adrenocortico-
vators among trauma survivors affected by PTSD. tropic hormone (ACTH) and cortisol levels during
Several genes that differentiated the groups had pre- separation stress. Macaques with the l/s allele of
viously been found to have a role in stress response. 5-HTTLPR had higher ACTH levels during stress
Zieker et al. [71] studied eight individuals who and there was a significant interaction of separation
had been exposed to the Ramstein air show catas- stress  rearing  5-HTTLPR genotype, which
trophe in 1989 and eight controls using cDNA micro- indicated that the effect of 5-HTTLPR genotype on
arrays. The genes that were selected for inclusion on hormonal response during stress is modulated by
the custom chip used for the molecular analyses were early experience. Bennett et al. [74] examined the
those that played a role in stress and immune effect of the 5HTTLPR promoter polymorphism in
response (inflammation, apoptosis, stress response rhesus monkeys. They found that the 5HTTLPR poly-
and related pathways). Their statistical analysis morphism interacted with deleterious early rearing
identified 4 upregulated genes and 14 downregulated experiences. Specifically, monkeys with the genotype
genes in the PTSD group versus controls. Most of the associated with less transcription had lower 5-hydro-
genes that were downregulated were associated with xyindoleactic acid (a 5HT metabolite) if they had been
immune functions or with reactive oxygen species. reared in deleterious conditions, but not if they had
The authors concluded that transcript differences that been raised in more favorable conditions. Rasmusson
they observed suggest a temporary perturbance of the et al. [75] studied the effects of psychological stress
oxidative stress system and specific immune parameters (cues conditioned to foot-shock) on the expression of
in patients with PTSD. BDNF in male rats. Tones that had previously been
paired with foot-shock downregulated BDNF; BDNF
mRNA was significantly decreased in the hippocam-
Animal models pus following exposure to the aversive conditioned
There have been a number of studies that have used stimulus.
animal models to investigate the role of genetic Animal models are very valuable because more
factors in PTSD. Clearly one cannot examine most invasive techniques can be utilized (e.g. measuring
of the psychological and behavioral features of human mRNA in the hippocampus) and controlled exposure
PTSD in animals. However, plausible animal models to stressors is possible. It is not clear how precise an
can be proposed because traumatic stress can be analogue to PTSD one can create in an animal model,
delivered very reliably and various species display but it seems likely that many of the same genetic and
behavioral and physiological characteristics that rep- physiological phenomena are operating.
resent a reasonable analogue to PTSD. Kaufer et al.
[72] investigated long-lasting changes in cholinergic
gene expression in mice as a possible mechanism by Genome-wide association studies
which exposure to trauma is transduced into a rela- The cutting-edge research design for gene discovery
tively stable phenotype such as PTSD. They found is currently the genome-wide association studies
that modulated cholinergic gene expression reduced (GWAS). In the most typical case, a large number of

141
Chapter 12: Post-traumatic stress disorder

unrelated individuals with the disorder are identified effects that are present in current cases of PTSD
and compared to a large number of well-matched might not be present in individuals who have
individuals without the disorder. The density of gen- recovered.) A simple and reasonable answer to defin-
etic markers and the extent of linkage disequilibrium ing a case is to apply the DSM criteria for the dis-
must be adequate to capture a large proportion of order. However, evidence suggests that PTSD, as
the common variation in the human genome. In defined by the DSM criteria, is a heterogeneous dis-
order to identify variants of modest effect, which order composed of at least two subtypes, internalizing
will probably include most if not all relevant to psy- and externalizing [76, 77]. Future genetic studies may
chiatric disorders, a large number of subjects for benefit from a more homogenous definition of cases
genotyping is necessary. At this writing, no GWAS using these subtypes. An important consideration in
studies of PTSD have been conducted. selecting controls is whether or not they have been
exposed to a traumatic event. If an individual has
not been exposed to a traumatic event, he or she
Methodological challenges cannot manifest PTSD regardless of the genetic sus-
As with most psychiatric disorders, it seems likely ceptibility that he or she might have. For the purpose
that more than one gene, and quite possibly a large of studies such as association studies, it is probably
number of genes, play a role in imparting vulnerabil- best to classify individuals without trauma exposure
ity to PTSD. There may also be genetic heterogeneity, as “phenotype unknown” and exclude them from the
that is, there may be a number of different genetic study. It might be that putative PTSD endopheno-
pathways to PTSD, with none being necessary or types, such as hypothalamic-pituitary-adrenal axis
sufficient. There may be incomplete penetrance (i.e. functioning, acoustic startle response, and physio-
the phenotype may not appear in all individuals with logical markers of increased arousal might improve
the relevant genotype) and variable expressivity the ability to detect genes beyond sole reliance on
(the extent to which the genotype is expressed in an the clinical phenotype.
individual’s phenotype may vary). The situation may Population stratification is another issue that
be further complicated by interactions among genes. must be considered in designing association studies
The interaction of PTSD relevant genes with the of PTSD. This refers to a situation in which groups
environment (G  E) is assured by the criteria for may differ in the frequency of an allele and may differ
PTSD which require that an individual be exposed to in some other, unrelated characteristic that is relevant
a traumatic event. By this definition, genetic factors to the phenotype of interest. For example if Northern
cannot produce PTSD in the absence of a stressor. Europeans and South Asians differ in the frequency of
The definition of a G  E interaction is that genetic some genetic marker and they also differ in social
factors determine sensitivity to the environment. support, which has a demonstrated influence on the
In the case of PTSD we are searching for genes that risk of PTSD, one might draw the spurious conclu-
determine whether an individual develops the charac- sion that the marker was associated with PTSD if the
teristic signs and symptoms of PTSD after he or she is two ethnic groups were differentially represented
exposed to an environmental stimulus (the stressor). among cases and controls.
An important methodological issue in the design
of studies to identify genes that influence PTSD is the
selection of cases and controls. The question is “who Conclusions
should be considered to be a PTSD case?” The selec- Evidence from twin and family studies clearly
tion of cases for this type of study is considered demonstrates that genetic factors exert a significant
“phenotype definition.” Researchers must decide influence on the risk of PTSD, but at the level of
whether to select on the basis of current PTSD or identifying specific genes that contribute to risk, the
PTSD at any time during the individual’s lifetime. evidence is less clear. There have been a number of
Since the primary sequence of DNA does not change, positive findings with regard to genes that play a role
it seems feasible to utilize individuals who have ever in the dopaminergic system, but none have a compel-
had PTSD, whether it is current or not. (This ling record of replication. Similarly, some findings
reasoning does not apply to epigenetic studies because point to the relevance of genes related to serotonin
phenomena-like gene expression do change and and glucocorticoid functioning, but replication will be

142
Chapter 12: Post-traumatic stress disorder

required. It seems that we are only at the beginning of matched by the commitment of clinical and epidemi-
explicating what is likely to be a very complex rela- ological researchers, it is not unreasonable to look
tionship between genes and PTSD. However, if the forward to exciting discoveries that eventually will
dramatic advances that are taking place in technology help to reduce or eliminate the suffering associated
at the molecular level and at the statistical level are with PTSD.

References 13. Gallers J, et al. J Trauma Stress


1988;1:181–192.
34. Kremen WS, et al. Arch Gen
Psychiatry 2007;64:361–368.
1. Zubin J, et al. J Abnorm Psychol
1977;86:103–126. 14. Sutker PB, et al. Psychol Rep 35. Storr CL, et al. Am J Psychiatry
1990;66:912–914. 2007;164:119–125.
2. American Psychiatric Association.
Diagnostic and Statistical 15. Helzer JE, et al. N Engl J Med 36. McLeer SV, et al. J Am Acad
Manual of Mental Disorders, 1987;317:1630–1634. Child Adolesc Psychiatry 1988;
Fourth Edition, Text Revision 16. Delimar D, et al. Mil Med 27:650–654.
(DSM-IV-TR). Washington, 1995;160:635–639. 37. Arata CM, et al. J Psychol Hum Sex
DC: American Psychiatric 17. Breslau N, et al. Arch Gen 1995;8:79–92.
Association; 2000. Psychiatry 1991;48:216–222. 38. McFarlane AC. Br J Psychiatry
3. Tomb DA. Psychiatr Clin North 18. Breslau N, et al. Arch Gen 1988;152:116–121.
Am 1994;17:237–250. Psychiatry 1997;54:81–87. 39. Brewin CR, et al. J Consult Clin
4. American Psychiatric Association. 19. Creamer M, et al. Psychol Med Psychol 2000;68:317–36.
Diagnostic and Statistical Manual 2001;31;1237–1247. 40. Ozer EJ, et al. Psych Bull
of Mental Disorders, Third Edition 2003;129:52–73.
(DSM-III). Washington, DC: 20. Kessler RC, et al. Arch Gen
American Psychiatric Association; Psychiatry 1995;52:1048–1060. 41. Davidson JRT, et al. J Psychiatr
1980. 21. Norris FH. J Consult Clin Psychol Res 1998;32:301–309.
5. Lasiuk GC, et al. Perspect Psychiatr 1992;60:409–418. 42. Yehuda R, et al. Am J Psychiatry
Care 2006;42:13–20. 22. Resnick HS, et al. J Consult Clin 1998;155:841–843.
Psychol 1993;61:984–991. 43. Yehuda R, et al. J Psychiatr Res
6. Pizarro J, et al. Arch Gen
Psychiatry 2006;63:193–200. 23. Breslau N, et al. Psychol Med 2001;35:261–270.
2004;31:889–898. 44. True WR, et al. Arch Gen
7. American Psychiatric Association.
Diagnostic and Statistical Manual 24. Breslau N, et al. J Urban Health Psychiatry 1993;50:257–264.
of Mental Disorders, First Edition 2004;8:530–544. 45. Stein MB, et al. Am J Psychiatry
(DSM-I). Washington, DC: 25. Shore JH, et al. J Nerv Ment Dis 2002;159:1675–1681.
American Psychiatric Association; 1989;177:681–685. 46. Jang KL, et al. Twin Res Hum
1952. Genet 2007;10:564–572.
26. Davidson JRT, et al. Psychol Med
8. American Psychiatric Association. 1991;21:1–19. 47. Lyons MJ, et al. Am J Med
Diagnostic and Statistical Manual Genet B Neuropsych Genet 1993;
27. Kessler RC, et al. Arch Gen
of Mental Disorders, Second 48:22–27.
Psychiatry 2005;62:593–602.
Edition (DSM-II). Washington,
DC: American Psychiatric 28. Kessler RC. J Clin Psychiatry 48. Comings DE, et al. JAMA
Association; 1968. 2000;61:4–14. 1991;266:1793–1800.
9. Herman JL. Trauma and Recovery: 29. Breslau N, et al. Psychol Med 49. Comings DE, et al. Biol Psychiatry
The Aftermath of Violence – 1999;29:813–821. 1996;40:368–372.
From Domestic Abuse to Political 30. Breslau N, et al. Arch Gen 50. Gelernter J, et al. Biol Psychiatry
Terror. New York: Basic Books; Psychiatry 1998;55:626–632. 1999;45:620–625.
1997. 51. Young BR, et al. Alcohol
31. Tolin DF, et al. Psychol Bull
10. Lamprecht F, et al. Psychosom 2006;132:959–992. Alcoholism 2002;37:451–456.
Med 2002;64:222–237. 52. Mustapić M, et al. Am J Med
32. Breslau N, et al. Arch Gen
11. Paige SR. Integr Physiol Behav Sci Psychiatry 2006;63(11): Genet B Neuropsych Genet
1997;32:75–83. 1238–1245. 2007;144:1087–1089.
12. Breslau N. Can J Psychol 2002;47: 33. Koenen KC, et al. Psychol Med 53. Segman RH, et al. Mol Psychiatry
923–929. 2007;37:181–192. 2002;7:903–907.

143
Chapter 12: Post-traumatic stress disorder

54. Bachmann AW, et al. Psycho- 63. Zhang H, et al. Am J Med Genet B 71. Zieker J, et al. Mol Psychiatry
neuroendocrinology 2005;30: Neuropsych Genet 2006;141: 2007;12:116–119.
297–306. 387–393. 72. Kaufer D, et al. Nature 1998;
55. Lappalainen J, et al. Arch Gen 64. Lawford BR, et al. Eur 393:373–377.
Psychiatry 2002;59:825–831. Neuropsychopharmacol 2003;
73. Barr CS, et al. Biol Psychiatry
56. Lu AT, et al. Am J Med Genet B 13:313–320.
2004;55:733–738.
Neuropsych Genet 2008;147: 65. Lawford BR, et al. Eur
1488–94. 74. Bennett AJ, et al. Mol Psychiatry
Psychiatry 2006;21:
2002;7:118–122.
57. Freeman T, et al. J Neuropsychiatry 180–185.
Clin Neurosci 2005;17:541–543. 66. Koenen KC, et al. J Affect Disord 75. Rasmusson AM, et al. Neuro-
2008;105:109–115. psychopharmacology 2002;
58. Koenen KC, et al. Arch Gen
27:133–142.
Psychiatry 2005;62:1258–1265. 67. Xian H, et al. Drug Alcohol
59. Binder EB, et al. JAMA Depend 2000;61:95–102. 76. Miller MW, et al. Psychol Assess
2008;299:1291–1305. 2003;15:205–215.
68. McLeod DS, et al. J Trauma Stress
60. Lee HJ, et al. Depress Anxiety 2001;14:259–275. 77. Miller MW, et al. Behavior Ther
2005;21:135–139. 2007;38:58–71.
69. Chantarujikapong SI, et al.
61. Kilpatrick DG, et al. Am Psychiatry Res 2001; 78. Weisaeth L, et al. PTSD Res Q
J Psychiatry 2007;164:1693–1699. 103:133–146. 1991;2(2):1– 7.
62. Lee HJ, et al. Stress Health 70. Segman RH, et al. Mol Psychiatry 79. Cohen ML, et al. Pain Rev 1996;
2006;22:115–119. 2005;10:500–513. 3:181–202.

144
Antisocial behavior: gene–environment
Chapter

13 interplay
Laura A. Baker, Catherine Tuvblad, Serena Bezdjian, and Adrian Raine

The term “antisocial behavior” is a broad one that net around antisocial behavior and these related
encompasses many facets of deviance, most of which behaviors under the term “externalizing disorders”.
bring harm to another person or involve the violation Externalizing disorders include different disruptive
of the rights of others in some way. Violence and and problem behaviors within the disinhibitory spec-
aggression bring physical or psychological harm to trum such as substance use, antisocial behavior,
other individuals, while property destruction and ADHD, conduct disorder, as well as personality traits
theft show disregard and possible damage to another such as impulsivity and sensation seeking. Several
person by way of their possessions. Antisocial behav- studies have shown that these traits and behaviors
ior often involves breaking the law, although less are related on a phenotypic level [2, 3]. Twin studies
serious or more subtle forms of rule violations such have shown that they also share a common genetic
as skipping classes, disruptive behavior, lying, and component [4].
deception are also considered to be antisocial in the Genetic influences are well evident for antisocial
broadest definition. behavior, both in the broadest sense of externalizing
As such, antisocial behavior in various forms is of behaviors as well as in various narrower forms such as
key importance in several psychiatric disorders, both aggression, law-breaking, and psychopathic personal-
in children and adults. The American Psychiatric ity traits. This research stems largely from over
Association [1] definitions for oppositional defiant three decades of twin, family, and adoption studies
disorder and conduct disorder in children, as well as in which concordance for disorders or trait covaria-
antisocial personality disorder in adults, all include tion among relatives has been amply demonstrated to
some form of antisocial behavior in their symptoms, vary according to their genetic relatedness. That is,
while aggression features significantly in conduct dis- monozygotic (MZ) or identical twins show greater
order and antisocial personality disorder. Further- similarity than dizygotic (DZ) or fraternal twins and
more, these primary disorders for which antisocial nontwin siblings, who in turn are more similar than
behavior is symptomatic are often comorbid with unrelated siblings for individual and composite meas-
other disorders, including attention-deficit hyper- ures of antisocial behavior. Although genetic influ-
activity disorder (ADHD) and substance dependence, ences on antisocial behavior appear to hold across
while aggression is either a comorbid condition of both sexes, their effects to some extent depend on
(or is central to) many psychiatric disorders, includ- age, and the ways in which antisocial behavior
ing autism, schizophrenia, depression, dementia, and is defined and measured. Most importantly, genetic
intermittent explosive disorder. Alcohol and other predispositions, though important, often appear more
substance abuse in particular are often considered deleterious in the presence of adverse environments.
forms of antisocial behavior, because of the harm they Indeed, antisocial behavior is one domain of human
bring to family members, friends and neighbors, or behavior in which gene  environment interactions
society as a whole. Although not currently defined in have been found repeatedly.
the DSM-IV-TR, psychopathy is another disorder for In spite of the overwhelming evidence for genetic
which antisocial behavior plays a major role. Indeed, contributions to the risk for engaging in antisocial
some clinicians and researchers cast an even broader behavior, the specific nature of the genetic risk is only

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

145
Chapter 13: Antisocial behavior: gene–environment interplay

beginning to be understood. For the most part, the compared and combined. One recent meta-analysis
genetic effects in antisocial behavior remain unspeci- [5] included 51 distinct twin and adoption studies
fied, and heritable influences are considered largely as that focused primarily on some dimension of anti-
a black box. This lack of specificity is also true for social behavior in children or adults. These included
environmental influences in behavioral genetic stud- studies of trait aggression, criminal offending, and
ies of antisocial behavior, in that these effects are symptoms for the major psychiatric disorders involv-
“anonymous” and we have little certainty about which ing antisocial behavior (i.e. conduct disorder, opposi-
experiences or circumstances exert true causal effects tional defiant disorder, antisocial personality disorder),
on deviant outcomes. Advances in molecular genetics and involved a variety of methods of assessment
and multivariate genetic models, however, have (e.g. self-report, parent ratings, and official records).
provided methods for unpacking both the genetic Heritable effects, including additive genetic influ-
and environmental black boxes for antisocial behav- ences (0.32), nonadditive genetic influences (0.09),
ior. Based on molecular genetic techniques that are explained in total 41% of the variance in antisocial
now widely accessible to social scientists, we now have behavior when combining across studies, and these
opportunities for identifying specific genes, and for effects did not differ for males and females. Environ-
identifying environmental factors that limit or mental effects shared by relatives accounted for less
enhance genetic expression in antisocial behavior. than half as much variance (16%), with the remaining
Even without specific DNA measures, multivariate variance (43%) being explained by nonshared envir-
genetic models help unpack the genetic and environ- onment. These effects were comparable for males
mental effects by examining covariation between and females, although they differed significantly
antisocial behavior and its various social (e.g. socioeco- according to definition, method of assessment of
nomic status, peer characteristics, parental monitoring, antisocial behavior, and age of the subjects. Criminal
and discipline) and biological (e.g. neurotransmitters, convictions appear more influenced by nonadditive
physiological arousal, frontal lobe function, and hor- genetic effects (i.e. due to genetic dominance or epis-
mones) risk factors and how they might be genetically tasis), while parental ratings and younger samples are
and environmentally mediated. more influenced by shared environmental factors.
This chapter reviews the evidence for genetic and Given the confound between age and method of
environmental influences – both specified and assessment across studies – studies of children tend
unspecified – in antisocial behavior. We discuss her- to rely on parental ratings – it is difficult to know
itability of both adult and child mental disorders in whether the greater effect of shared environment in
DSM-IV-TR, for which antisocial behavior is central children may simply be an artifact of rater bias.
to their diagnosis, but also review heritability of the It should also be noted that error variance and the
related externalizing disorders. Genetic studies of variance due to gene  environmental interactions
crime in particular also illustrate the importance of are included in the component attributed to non-
genotype  environment interactions, which are now shared environmental influences, thus tending to
beginning to be understood at a more molecular level. underestimate genetic contributions and overestimate
We highlight some of the most exciting new direc- environmental contributions.
tions in this field, which aim to unpack the genetic and Individual studies have also examined the genetic
environmental black boxes in antisocial behavior, and influence in the broad externalizing factor by com-
understand the complexities of the gene–environment bining various antisocial behavior measures from the
interplay in antisocial development. same subjects into one model (Figure 13.1). A recent
study of antisocial behavior in 9–10-year-old twins
Heritability of the broader construct attempted to evaluate the effects of raters’ identity
on estimates of genetic and environmental influences
of externalizing problems during childhood [6]. The study was based on a socio-
One approach for understanding the role of genetic economically and ethnically diverse sample of 605
influences in the wider realm of externalizing behav- pairs of twins (MZ, DZ same sex, and DZ opposite
ior problems is through meta-analysis of behavior sex) and their caregivers who participated in a com-
genetic studies, in which heritability estimates for prehensive assessment of the twins’ antisocial behav-
the varying forms of antisocial behavior can be ior and related risk factors. Both the child and his/her

146
Chapter 13: Antisocial behavior: gene–environment interplay

Figure 13.1 Standardized parameter


AC CC EC estimates for common pathways model
of antisocial behavior in multiple
informants. (From [111], with permission.
Copyright © 2007 by the American
0.00 Psychological Association.)
0.98* 0.19

Shared
view

0.67* 0.55*
0.42*

Caregiver Child Teacher


report report report

0.24 0.38* 0.58* 0.55* 0.15 0.71* 0.33 0.53* 0.45* 0.33*

AM RM EM AK CK EK AT RT CT ET

caregiver provided reports of the child’s antisocial Table 13.1 Proportions of variance in antisocial behavior
behavior, in addition to teacher ratings of each child. explained by genetic, shared and nonshared environment in
different informants.
Composite measures of antisocial behavior were com-
puted for each rater – parent, teacher, and child – Source of influence Caregiver Child Teacher
based on several standardized instruments measuring
Genetic
rule-breaking behaviors, including theft and violence,
as well as reactive and proactive aggression and con- Common factor 0.436 0.168 0.289
duct disorder symptoms, as rated by each informant. Informant specific 0.059 0.304 0.108
The pattern of genetic and environmental influences
Total 0.495 0.472 0.397
in the composite antisocial behavior measures for
these pre-adolescent children, as summarized in Shared environment/
Figure 13.1 and Table 13.1, was similar to that found rater effects
in Rhee and Waldman’s meta-analytic review, in that: Common factor 0.000 0.000 0.000
(1) genetic effects were significant for antisocial a
Informant specific 0.146 0.021 0.203
behavior as assessed by each of the three informants;
(2) shared environmental effects were larger for Rater effects 0.281
parent and especially for teacher ratings of antisocial (correlated errors)b
behavior as compared to the children’s self-report; Total 0.146 0.021 0.484
and (3) the respective magnitudes of genetic and Nonshared
environmental effects were comparable for males environment
and females. A larger effect of shared environment
for childhood antisocial behavior was exhibited, as Common factor 0.016 0.006 0.011
rated by parents and teachers, although genetic influ- Informant specific 0.344 0.500 0.108
ences are still significant at this early age. The larger Total 0.360 0.506 0.119
shared environmental effect estimated in twin studies a
Informant-specific shared environment and rater effects cannot
relying on parent or teacher ratings of antisocial be differentiated in caregiver reports.
b
behavior may thus be due in part to a form of rater Rater effects are not applicable in child reports.
bias, rather than to a true, shared environmental From Baker et al. [111]. Copyright © 2007 by the American
Psychological Association. Reproduced with permission.
effect. Perhaps most compelling is the finding that

147
Chapter 13: Antisocial behavior: gene–environment interplay

the latent common factor underlying the three differ- broader or secondary ones (ADHD and substance
ent raters of the child’s antisocial behavior was almost use disorders) are summarized below.
completely explained by genetic factors (96%). This
very high heritability of the latent common factor Antisocial personality disorder
is due to less error of measurement in a construct
Antisocial personality disorder is probably one of the
derived from three raters, and as such could be con-
most extensively researched personality disorders. It
sidered as the relative importance of genetic factors on
is characterized as a pervasive pattern of disregard for
the true score variance underlying antisocial behavior
and violation of the rights of others occurring since
measures across informants. When such true score
childhood or early adolescence. Individuals diagnosed
models are used to assess the underlying externalizing
with antisocial personality disorder must be at least 18
problem factor in a more reliable manner, genetic
years old, and have had a history of conduct disorder
influences generally become most evident.
before age 15. To be diagnosed, symptoms or behav-
iors of the disorder must have been occurring since
Heritability of DSM-IV disorders childhood or early adolescence. These behaviors
related to antisocial behavior include fighting, setting fires, running away from
home, cruelty to animals, and conflicts with authority
As noted earlier, antisocial and aggressive behaviors
figures [1].
are fundamental in the diagnoses of several psychi-
Estimates of antisocial personality disorder in the
atric disorders. At the symptom level, antisocial per-
community are thought to be 3% [1]. In forensic
sonality disorder in adulthood includes behaviors
settings, the prevalence ranged from about 3 to 30%.
such as aggression, impulsivity, irresponsibility, and
The disorder is more prevalent in males compared to
recklessness. Consequently, it becomes difficult for
females [1]. Having a relative that is affected increases
affected individuals to conform to social or cultural
the risk of developing the disorder [7]. In addition,
norms, that is, to keep a job, complete an education,
despite the prevalence of antisocial personality disorder
and have a long-term romantic relationship. During
being lower in females, affected female probands
childhood and adolescence conduct disorder involves
normally have more affected relatives than male
externalizing behaviors including aggression towards
probands. This indicates that females may require
people and animals, destruction of property, dis-
stronger environmental and genetic influences than
honesty, theft, and other serious violations of age-
males to develop antisocial personality disorder [7].
appropriate rules such as truancy. Conduct disorder
The early adoption studies from the 1970s and
tends to be preceded by oppositional defiant disorder
1980s probably provide the most important findings
which is characterized by overly aggressive and defi-
that heritable influences account partly for the devel-
ant behaviors, and a negative and hostile behavior
opment of antisocial and criminal behavior. These
towards authority figures. These disorders are typic-
studies demonstrated that having a criminal birth
ally measured as diagnostic criteria categories; either
parent increased one’s own risk of having a criminal
the individual meets the criteria for the disorder, or
conviction as an adult, regardless of if the person was
not. In addition to these disorders for which antisocial
reared by pro-social and law-abiding but genetically
behavior is symptomatic, there are several other DSM
unrelated foster parents. This finding was replicated
disorders which may be considered as antisocial in
in adoptive cohorts across different populations, includ-
the broader sense (e.g. drug use), or which are often
ing Scandinavia [8, 9] and the United States [10]. Based
comorbid with conduct disorder and oppositional
on these adoption studies, the genetic effect on crim-
defiant disorder in children, and antisocial personality
inal outcomes appears important for both sexes,
disorder in adults. These additional disorders include
although individual genetic risk is typically more
substance use disorders and ADHD. The influence of
extreme for female than male offenders [11].
genetic and environmental factors in these different
disorders has been examined in both twin and adop-
tion studies. Behavior genetic studies of the three Conduct disorder
primary disorders related to antisocial behavior (anti- The characteristic feature of conduct disorder in chil-
social personality disorder, conduct disorder, and dren and adolescents is a repetitive and persistent
oppositional defiant disorder), as well as the other pattern of behaviors including aggressive behaviors

148
Chapter 13: Antisocial behavior: gene–environment interplay

causing physical harm to people and/or animals, non- authority figures. Oppositional defiant disorder includes
aggressive behaviors causing property loss or damage, behaviors such as lashing out at adults, losing one’s
deceitfulness, theft, and violation of rules. Conduct temper, and intentionally annoying others [1]. The
disorder is mainly diagnosed in individuals 18 years prevalence of oppositional defiant disorder varies
or younger. Conduct disorder can be further divided depending on age and tend to decrease across age: in
into two subtypes: childhood-onset type and adoles- children 6–8 years of age the prevalence ranges from
cent-onset type. The childhood-onset type requires 1.1 to 13.3%, in children 8–14 years of age the preva-
an onset prior to the age of 10, and these individuals lence ranges from 1.7 to 6.9%, and in children 14–17
are generally characterized as frequently being aggres- years of age the prevalence ranges from 1.0 to 5.8%
sive towards others and having disturbed peer [16]. Results from twin studies have shown that
relationships. Compared to the childhood-onset type, genetic, as well as shared environmental influences
adolescent-onset individuals tend to be less aggressive are important in the development of oppositional
and to have more normative peer relationships [1]. defiant disorder symptoms, explaining approximately
The prevalence of conduct disorder for children one-third each of the total variance [25, 26, 33, 34].
and adolescents ranges between 1 and 15% in commu-
nity samples [12, 13] and makes it one of the most
common child psychiatric disorders. A general finding Attention-deficit hyperactivity disorder
is that conduct disorder is more common among boys ADHD is one of the most common childhood psychi-
[14, 15]. There is also an increasing prevalence with atric disorders. The worldwide prevalence of ADHD
age, in both sexes, especially in the mid-teens [16]. was recently estimated in a meta-analysis to be around
Conduct disorder is also associated with drug 5.29% in children and adolescents [35]. It is a disorder
abuse and dependence [17, 18] and children with with two separate clusters of symptoms: hyperactivity–
conduct disorder are more likely to later develop impulsivity, including excessive activity and impulsiv-
other psychiatric disorders, including depression and ity; and inattention including difficulties in sustaining
anxiety [19]. Most importantly, numerous studies attention, distractibility, lack of task persistence, and
have shown that conduct disorder is one of the disorganization [1]. Although DSM-IV recognizes
strongest predictors of antisocial personality disorder either high levels of hyperactivity–impulsivity or high
in adulthood [20, 21]. This association is particularly levels of inattention, most affected individuals exhibit
strong in combination with having a biological parent high levels of both types of symptoms. ADHD tends
with antisocial personality disorder [22]. to persist across development, but whereas symptoms
Results from twin studies suggest genetic factors of hyperactivity and impulsivity decrease with age,
contribute to the development of conduct problems symptoms of inattention tend to persist. The preva-
in children [23]. A recent study showed that conduct lence of ADHD in adulthood has been estimated to be
disorder symptoms are fairly heritable, with estimates 4.4% [35]. The disorder is associated with academic
ranging between 27 and 78% [24]. Further, shared underachievement, substance use and dependence,
environmental factors also seem to be of importance social maladjustment, and antisocial behavior [36].
in explaining the development of conduct disorder Childhood ADHD is related to antisocial behavior
[25, 26]. Shared environmental influences refer to in two different ways. First, several studies have
nongenetic influences that contribute to similarity shown childhood ADHD is associated with a later
within pairs of twins. Shared environmental risk onset of antisocial behavior in adolescence and adult-
factors for conduct disorder may for example include hood [37, 38]. Second, individuals having a combin-
family-related factors (e.g. poor child rearing practices, ation of ADHD and conduct disorder have been
maltreatment [27–29]) and contextual factors (e.g. found to have a worse outcome, compared with
neighborhood disadvantage and poverty [30–32]). individuals having either ADHD or conduct disorder
[39]. It has also been suggested that children with
both types of disorders manifest more severe forms
Oppositional defiant disorder of antisocial and aggressive behaviors [39].
Oppositional defiant disorder is also a childhood- Results from twin and adoption studies provide
onset psychiatric disorder, and is characterized by a strong evidence suggesting a genetic basis for ADHD
pattern of disobedient and hostile behavior towards symptoms [40]. The heritability of ADHD symptoms

149
Chapter 13: Antisocial behavior: gene–environment interplay

have been found to be approximately in the range It should be mentioned that although genetic
of 55–80% [41–43]. A general finding is that shared influences partly explain the risk for disorders involv-
environmental factors are of minor importance for ing antisocial behavior as well as those involving
the underlying symptoms of ADHD, whereas the substance dependence; this does not necessarily mean,
nonshared environment explains some part of the however, that the same genetic and/or environmental
variance, around 15–25% [42]. The relative lack of factors influence both types of disorders. To investi-
shared environmental influences is consistent with gate the genetic and environmental overlap between
findings for antisocial behavior in general [5]. Non- substance abuse and other antisocial behavior related
shared environmental factors may include unique disorders, a genetic informative design is required.
individual experiences, and possibly biological factors Several twin studies have in fact investigated this.
that may not be genetic in origin. Such nonshared Different disruptive and problem behaviors within
environmental factors refer to experiences that may the disinhibitory spectrum such as substance use,
affect one twin in a pair but not the other, for example antisocial behavior, ADHD, conduct disorder, and
head injury or differences in parental treatment. impulsive and sensation-seeking personality traits
In sum a vast majority of twin and adoption can on a phenotypic level be united by a common
studies provide consistent evidence for a strong gen- higher-order externalizing factor [2, 3]. The genetic
etic contribution to ADHD; environmental factors influences on a common externalizing factor describ-
seem to be largely nonshared. In addition, several ing conduct disorder, substance use, ADHD, and
studies have reported comorbidity among ADHD, novelty seeking was found to account for more than
conduct disorder and oppositional defiant disorder 80% of the variation in an adolescent sample [4].
in both epidemiological and clinical samples [16, Similar findings were reported in a study linking
44–46]. Studies examining the co-occurrence among antisocial personality disorder, conduct disorder,
these disorders generally show that these different alcohol and drug dependence and unconstrained
disruptive and problem behaviors can be united by personality style [47]. Strong heritable influences on
an externalizing higher-order factor [2, 3]. This an externalizing factor of antisocial behavior, sub-
higher-order externalizing factor has been found to stance abuse and conduct disorder were also found
be highly heritable [47], explaining about 60% of the in an adult sample [51]. Consequently, the broader
total variance [34]. construct of externalizing behaviors based on mul-
tiple measures shows approximately the same high
level of heritability that has been demonstrated for
Substance use and abuse a multiple-measure, multiple-informant construct of
The relationship of antisocial behavior (including broad antisocial behavior [6].
violence and aggression) with alcohol and other drug Importantly, it does appear that the same genetic
use is well established [18]. Family and twin studies risk factors leading to substance dependence may in fact
can be used to understand further the relationship provide risk for a wider range of antisocial disorders,
between antisocial behavior and substance use. The both in adolescent and adult samples. Given the
co-occurrence of substance use and aggression may common genetic link across these various DSM-IV
be due to shared common risk factors, such as genetic disorders pertaining to antisocial behavior and sub-
or temperamental traits, antisocial personality dis- stance use, it is also important to examine the genetic
order, or parental substance use and violence. As with underpinnings of some of the personality traits that may
criminal and antisocial behavior [48], substance underlie antisocial behavior. We thus turn attention to
dependence tends to run in families. Most of this the heritability psychopathic personality traits, which
family resemblance is explained by genetic influences have been studied extensively in twins in recent years.
[49]. Additionally, the relative influence of genetic
and environmental influences on substance depend-
ence tends to be the same in both men and women. Heritability of psychopathic traits
However, as with antisocial and criminal behavior [5], Psychopathy is, in its adult and full manifestation,
it has also been indicated that women require greater considered to be a serious personality disorder that
familial loading, that is, greater genetic propensity is associated with severe and persistent antisocial and
or liability, to express these disorders [50]. criminal behavior [52]. The principal view of

150
Chapter 13: Antisocial behavior: gene–environment interplay

psychopathy originates from the personality-based and adolescents. These studies together suggest that
clinical conception of the syndrome in which core psychopathic personality shows significant heritability
interpersonal affective deficits are the focus, e.g. [62–64]. For example, a study conducted by Larsson
superficial charm, lack of remorse, unemotionality and colleagues [62] using a set of 16–17-year-old
[53]. Its base rate in the general population is not twins, showed that psychopathic personality dimen-
known, but has been estimated at between 0.5 and sions, callous–unemotional, grandiose–manipulative,
1.0% [54]. Psychopathy is today generally conceptual- and impulsive–irresponsible, could be explained by a
ized as a disorder defined by callous and unemotional common higher-order “psychopathic personality”
affects, a grandiose and manipulative interpersonal factor. Genetic or heritable influences explained as
style, pervasive impulsive and irresponsible behavior, much as 63% of the total variance in a latent “psycho-
and criminal offending [54]. The most common pathic personality” factor, while nonshared environ-
measure used to assess adults is the Psychopathy mental (i.e. refers to experiences unique to the
Checklist – Revised (PCL-R) [54, 55]. The PCL-R is individual) factors explained the remaining 37% of
a semi-structured interview that consists of 20 items the variance. Shared environmental influences (i.e.
based on Cleckley’s criteria [53]. The measure was nongenetic influences that contribute to similarity
initially designed to assess the traits and behaviors within pairs of twins) did not contribute to the
of psychopathy in incarcerated settings [54, 56]. explanation of “psychopathic personality”.
Psychopathy is distinct, but related to and overlaps Additional twin studies have shown that a
with the more behavioral-based diagnosis of anti- common genetic factor largely accounts for the pat-
social personality disorder [1]. It should be noted that tern of covariation between psychopathic personality
individuals identified as psychopathic very often traits and antisocial and delinquent behavior [65].
receive a diagnosis of antisocial personality disorder, Blonigen and colleagues conducted a study in 2005
but the reverse is not true. on a set of 17-year-old twins. They found that the
In recent years there has been a growing interest overlap between self-reported psychopathic personality
to study the developmental aspects of this severe traits (Fearless Dominance and Impulsive Antisociality
personality disorder [57]. Researchers have attempted measured using the Multidimensional Personality
to extend the construct of psychopathy downward Questionnaire) and externalizing psychopathology was
to children and adolescents [58, 59]. So far, the vast due to common genetic influences [66].
majority of research on psychopathy has been There are also a few longitudinal twin studies
conducted on male incarcerated samples. Though investigating stability of and change in genetic and
important, such research does not provide any infor- environmental influences over time [67]. A recent
mation regarding the development of psychopathic study conducted by Forsman et al. [68] used a longi-
traits. To understand the developmental origins of tudinal sample including 16–17 and 19–20-year-old
adult psychopathy it is necessary to use samples twins. Results showed that three psychopathic person-
including younger individuals. Further, to understand ality dimensions, callous–unemotional, grandiose–
genetic and environmental contributions to the devel- manipulative, and impulsive–irresponsible, were highly
opment of psychopathy genetically informative stable across time and linked to a higher-order latent
samples are required. Male incarcerated samples are factor (i.e. psychopathic personality factor). Genetic
probably also biased in that they include comorbid factors contributed substantially to the stability of this
cases and only identify the “unsuccessful” (caught) general higher-order factor, whereas environmental
psychopath. Recently several studies have shown that factors were of little importance [68].
psychopathic personality traits are present and can be
meaningfully assessed during adolescence [60, 61]. It is
however important to stress that the purpose of assess- Gene × environment interactions
ing psychopathic personality traits in youths is not to Gene–environment (G  E) interaction is defined as
assign a formal diagnosis of psychopathy, but rather a genetic influence on a given phenotype that depends
for prevention purposes to identify at-risk adolescents. upon certain environmental factors, or vice versa.
There are a few twin studies that have examined A genetic influence on a specific phenotype can be
the relative influence of genetic and environmental moderated by an individual’s experiences or exposure
effects on psychopathic personality traits in children to certain environments. Likewise, various individuals

151
Chapter 13: Antisocial behavior: gene–environment interplay

may respond differently to the same environmental environmental factors interact in the development of
exposure because they have different genotypes. Sev- antisocial and criminal behavior.
eral studies, using different research designs, including Several twin studies have also reported G  E
molecular genetic, twin and adoption designs have interaction [75]. Tuvblad et al. [76] showed that socio-
shown that genetic influences and environmental risk economic status moderated the influence of genetic and
factors interact in the development of antisocial and environmental factors on antisocial and delinquent
criminal behaviors. behavior. Using a sample of Swedish 16–17-year-old
Evidence of G  E interaction for antisocial and twins, heritability for antisocial and delinquent
delinquent behavior has recently been demonstrated behavior was found to be higher in more advantaged
in molecular genetic studies [27]. Caspi et al. [27] neighborhoods (37%) compared to the less advan-
reported that a functional polymorphism in the gene taged neighborhoods (1%). In contrast, the shared
encoding the neurotransmitter-metabolizing enzyme environment was higher in the less advantaged neigh-
monoamine oxidase A (MAOA) moderated the impact borhoods (69%) compared to more advantaged
of early childhood maltreatment on the development neighborhoods (13%). A similar, but less evident
of antisocial behavior in males. Maltreated boys pattern was found for girls. Raine [77] has argued
with a genotype conferring high levels of MAOA for a “social push hypothesis” in which the influence
expression were found to be less likely to develop of biological risk factors are more likely to become
antisocial problems than maltreated boys who had a expressed in an environment where an antisocial
genotype conferring low levels of MAOA expression. individual lacks the environmental risk factors that
The authors concluded that their findings may partly push or predispose him/her to behave antisocially.
explain why not all victims of maltreatment grow In contrast, the relationship between antisocial behavior
up to victimize others. There has so far been a few and biological risk factors will be weaker in adolescents
replications of this interesting finding [69–71] and from disadvantaged backgrounds because the envir-
one published failure to replicate [72]. onmental determinants of antisocial behavior will
An indirect way to test for G  E interaction is to camouflage the genetic contribution [78]. An inter-
use data from adoption studies. The coherence of pretation of these results in view of the “social push
research based on adoption samples showed that the perspective” would suggest that the influences of
combination of a genetic predisposition (i.e. psycho- genes on antisocial behavior are more expressed in a
pathology in biological parents) with a high risk socioeconomically advantaged environment where
environment (i.e. adverse adoptive home environ- the environmental risk factors are absent. On the
ment) lead to greater pathology than what would be contrary, genetic factors for antisocial behavior will
expected from either factor acting alone or both in an be weaker and the shared environment will be more
additive combination [9, 73, 74]. This was demon- important in a socioeconomically less advantaged
strated in a study by Cloninger et al. [8]. They studied environment because the environmental risk factors
the interaction of a genetic predisposition (i.e. will conceal the genetic contribution [76].
whether the biological parents were criminal) and a
high risk environment (i.e. adverse rearing experi-
ences and adoptive placement) in 862 Swedish men
Unpacking the genetic and
adopted at an early age. When both hereditary and environmental “black boxes”
adverse environmental factors were present, 40% of In spite of the extensive evidence for genetic influ-
the males were found to be criminal; if only genetic ences in antisocial behavior from twin and adoption
factors were present 12.1% became criminal; if only studies, the exact nature of either genetic or environ-
environmental factors were present 6.7% were crim- mental effects remains largely unknown. As in other
inal; and when neither hereditary nor adverse envir- domains of human behavior, the behavior genetic
onmental factors were present 2.7% were criminal. studies that focus on the heritability of antisocial
The fact that the combination of both hereditary behavior are often criticized for their “black box”
and environmental factors (40%) was higher than approach, since both genes and environment remain
either hereditary (12.1%) or environmental (6.7%) anonymous. Moreover, early behavior genetic studies
factors, or hereditary and environmental factors failed to specify the exact biological or social mechanisms
together (18.8%), indicated that hereditary and that underlie these global genetic and environmental

152
Chapter 13: Antisocial behavior: gene–environment interplay

influences. There is a growing body of research, One important and widely cited study demon-
however, which attempts to unpack the genetic and strating a link between a specific genotype and anti-
environmental black boxes, and understand the social behavior is the large Dutch pedigree study by
physiological pathways and the specific nature of Han Brunner and colleagues [79]. Several family
individual circumstances or experiences that may dir- members (especially males) who showed a history of
ectly influence antisocial outcomes, or moderate the violence and impulsive aggression were found to share
genetic influences themselves. We describe here two a mutant form of the gene on the X-chromosome that
primary approaches being used to unpack these black codes for MAOA. Specifically, MAOA is an enzyme
boxes and understand the biosocial pathways for that metabolizes several neurotransmitters involved
antisocial behavior development: These include: (1) in impulse control, attention, and other cognitive
a measured gene/measured environment approach, functions, including dopamine, norephinephrine,
in which researchers are attempting to identify spe- and serotonin [79]. Mutations in the normal MAOA
cific genes that increase risk for antisocial outcomes, gene lead to deficient production of the MAOA
along with narrowly defined measures of individual enzyme, which in turn disrupts the normal function
circumstances and experiences that attempt to flesh of these neurotransmitters, resulting in a wide range of
out the nature of environmental influences; and disorders, including ADHD, alcoholism, drug abuse,
(2) a measured risk factor approach, in which various impulsivity, and other risky behaviors [80–83]. Other
biological and social risk factors for antisocial behav- studies have found significant associations between
ior are studied in a genetically informative design. X-linked MAOA alleles with greater numbers of
The measured gene/measured environment approach dinucleotide repeats, and various behavior disorders,
may be used to investigate main effects of both genes including ADHD, conduct disorder, major depressive
and environment, or to understand more complex disorder, drug abuse, alcoholism, reward dependence,
interactions whereby environments may moderate and learning disabilities (see [84]). One hypothesis is
genetic effects or vice versa. The measured risk factor that longer repeat allele variations result in greater
approach includes the identification of endopheno- disruptions in gene regulation, leading to variation
types for antisocial behavior, and relies extensively in MAO function and hence individual differences
on multivariate genetic models that aim to under- in these related phenotypic disorders [85].
stand the genetic and environmental mediation of This finding of increased aggression being associ-
relationships between measured risk factors and ated with MAO deficiency produced by a genetic
antisocial behavior. mutation in the MAOA allele in the Brunner et al.
[79] pedigree coincides with animal research using
knockout strains of mice [86], where the same associ-
Gene identification ation between MAOA and violent aggression has been
Rapid advances in molecular genetic techniques have repeatedly demonstrated. Although the main effects
provided increased interest and practical opportun- of the MAOA mutation on aggression and violence
ities to search for specific genes that contribute to have not been replicated yet in any other large human
the risk for antisocial behavior and its correlates, pedigrees, this genetic defect remains the first such
such as impulsivity. This research, nonetheless, is still link to antisocial behavior in humans. One paradox is
in its infancy and there are only a handful of studies that MAOA deficiency is associated with particularly
to date that have evidenced associations between high high levels of the neurotransmitter serotonin, and yet
risk alleles and antisocial outcomes, particularly using low serotonin is a very well-replicated neurochemical
narrower definitions such as criminal offending or correlate of antisocial and aggressive behavior [87].
psychiatric disorders for which antisocial behavior Furthermore, MAOA-deficient mice show significantly
is primary. There have been no genome-wide linkage better fear conditioning [88], yet poor fear condition-
studies of antisocial behavior, in particular, although ing is a well-replicated correlate of criminal and psy-
if one widens the scope of antisocial behavior to chopathic behavior [87]. These conflicting findings
include substance abuse and other correlates such clearly require further investigation.
as impulsivity, molecular genetic studies are much One reason for the failure to find main effects
more extensive and include both association and of the MAOA mutation may be due to G  E inter-
linkage studies. actions, whereby the deleterious effects of the mutant

153
Chapter 13: Antisocial behavior: gene–environment interplay

gene are only observed under certain environmental A gene–gene interaction between DRD2 and DRD4
circumstances. The idea that environmental factors in predicting conduct disorder in childhood and
may moderate the effects of the MAOA mutations criminal offending in adults has also been shown
on aggression could in general be predicted from the [91]. Still, the finding of enhanced risk for later onset
well-replicated finding of G  E interactions in crim- criminal offending as a function of high genetic risk
inal behavior in adults and conduct problems and combined with low environmental risk is contrary
antisocial behavior in youth. Caspi et al. [27] did in to predictions from other developmental models of
fact demonstrate such a G  E interaction, whereby antisocial behavior. Moffitt [94] has suggested that
the functional polymorphism in the MAOA gene was persistent and early-onset antisocial behavior would
found to increase the risk for antisocial behavior in be more influenced by genetic factors than late onset,
males, but only for those who experienced early child- transient forms of antisocial behavior, yet the converse
hood maltreatment. The fact that the main effect of pattern was found by DeLisi et al. [90]. Nonetheless,
the MAOA mutation as found in Brunner et al. [79] the DeLisi et al. findings are particularly interesting
was not found in the Caspi et al. [27] study under- in that they also demonstrate a G  E interaction in
scores the importance of investigating specific criminal offending using specific genetic markers and
genetic effects under a variety of environmental cir- well-defined measures of the environment [90].
cumstances in order to fully understand the risk for Gene identification studies showing direct associ-
criminal offending and other antisocial behavior. The ations with more narrowly defined antisocial behavior
findings of Caspi et al. are particularly intriguing are still rare. Dick et al. [95] reported a linkage to a
since it was one of the first studies to illustrate the region of chromosome 7, which appears to contain
well-replicated G  E interaction in criminal behavior genes conferring risk to the externalizing spectrum
using a measured gene/measured environment (including alcohol, drug dependence, conduct disorder,
approach [27]. At the same time, one study has shown antisocial personality disorder, novelty, and sensation
that while the MAOA  abuse interaction holds for seeking). Specifically, the CHRM2 gene could be asso-
whites, it is not observed in African-Americans [89] ciated with the spectrum of externalizing psycho-
and consequently caution should be exercised as pathology [95].
interaction effects may not apply to all subgroups. There is however a more extensive literature
Another study using a measured gene/measured examining specific genetic effects in behavioral and
environment approach to understand antisocial out- neurochemical, and other physiological correlates
comes focused more specifically on the age of onset of antisocial behavior, and for psychiatric disorders
for criminal offending. Among those adolescents with for which antisocial behavior is symptomatic. For
a history of criminal offending, polymorphisms in example, numerous studies in both mice and humans
genes related to the neurotransmitter dopamine (also have identified specific genes involved in the produc-
discussed further below) were shown to be associated tion and function of the neurotransmitter serotonin
with age of first police contact and arrests, but only (5-hydroxytryptamine, or 5-HT), which has been
for youth from low risk family environments [90]. shown to play a role in impulsive and other risky
That is, the individuals at greatest risk for later onset behaviors such as drug abuse, gambling, and suicide
criminal offending were those with the A1 allelic form [96–99]. A recent study reported an association
of the DRD2 gene, in combination with favorable between genes related to 5-HT function and impulsivity
home environments as defined by maternal attach- in children with ADHD [100]. Evidence of an associ-
ment, involvement, and engagement. Unlike the ation between catechol-O-methyltransferase (COMT)
Brunner et al. [79] and Caspi et al. [27] studies which valine/methionine polymorphism and the development
involved the overall risk for engaging in antisocial of antisocial behavior among children with ADHD
behavior, the DeLisi et al. [90] finding involves the has also been found [101] and replicated [102].
age of onset of first police contact, and not the overall
risk for offending versus not offending. Nonetheless,
other studies have also shown that different forms Measured risk factors (endophenotypes)
of the DRD2 allele show associations with criminal In the absence of measured genes, it is still possible to
victimization [91] and age of first sexual intercourse unpack the genetic and environmental black boxes
[92], as well as normal personality variation [93]. using multivariate studies in a genetically informative

154
Chapter 13: Antisocial behavior: gene–environment interplay

106]. Additional genetic influences that are unique to


A C E
either antisocial behavior or the risk factor are also
indicated in this model (A: genetic effects; C: shared
environmental effects; E: nonshared environmental
effects for each phenotype). Thus, the total heritability
Antisocial
for a given trait such as antisocial behavior may be
Risk
behavior factor parsed into components that are shared and not
shared with the risk factor, in order to specify more
exactly the nature of the total genetic influences in
antisocial behavior.
a c e a c e Many correlates of antisocial behavior may be
considered as possible endophenotypes, in particular
Figure 13.2 Multivariate genetic model for antisocial behavior personality traits such as impulsivity and attention
and a measured risk factor. deficits. Other correlates include both biological and
social risk factors. Key biological risk factors include
design such as a classical twin or adoption study. hormones, physiological (autonomic) underarousal,
More specifically, it is possible to examine the genetic frontal lobe function (and dysfunction), and neuro-
and environmental influences between different traits transmitters. Any of these traits may be examined for
and behaviors that are known to correlate with the common genetic relationships with antisocial behav-
risk of antisocial behavior. In this “measured risk ior, although little work has been done using this
factor” approach multivariate genetic models are used approach to date.
to explain sources of genetic and/or environmental One example of a biological risk factor for antisocial
covariance that underlie the associations between risk behavior is autonomic underarousal in antisocial –
factors and antisocial behavior [103]. This is similar including violent – individuals. Antisocial individuals
to the “endophenotype approach”, in which research- and criminal offenders have shown a lower resting
ers identify highly heritable traits that show associ- heart rate and lower electrodermal (skin conductance)
ations with antisocial behavior [104]. In addition to responsivity both in orienting and fear conditioning
estimating the components of genetic and or environ- paradigms [107, 108]. The connections of these find-
mental variance common to both antisocial behavior ings to antisocial behavior have been interpreted in
and a specific risk factor, multivariate genetic twin several ways. For example, low heart rate may reflect
models can also be used to assess the correlation fearlessness, or reduced anxiety [109], lack of social-
between genes influencing antisocial behavior and ization, poor learning abilities due to cognitive def-
genes influencing the risk factor or endophenotype. icits or emotional withdrawal or both, or reduced
A large genetic correlation would result if the same brain functioning in areas involved in mediating psy-
gene(s) influences the risk factor and antisocial behav- chophysiological responding [77]. Another explan-
ior; this is referred to as pleiotropy. A multivariate ation is that low arousal may lead to stimulation-
genetic model involving antisocial behavior and one seeking behaviors, including violence, in an attempt
risk factor is presented in Figure 13.2. Common gen- to raise autonomic arousal to optimal levels.
etic and environmental effects (A, C, and E) influence Although low resting heart rate is the best-replicated
both the risk factor and antisocial behavior. biological correlate of antisocial behavior in child and
To the extent that genes have pleiotropic effects, adolescent populations [110], and while heart rate
that is, to the extent that they influence more than has been shown to be at least partially heritable, it
one phenotype, common genetic influences on the is unclear to what extent cardiovascular underarousal
risk factor and antisocial behavior should result. Close may be related to antisocial behavior due to a
proximity – “genetic linkage” – of the genes influ- common genetic link. This question can be addressed
encing antisocial behavior to the risk factor, or certain using multivariate genetic analyses as indicated in
patterns of assortative mating, such as antisocial indi- Figure 13.2. In our own longitudinal twin study of
viduals paired systematically to mates with extreme antisocial behavior, we found using multivariate ana-
values of the risk factor, or both, can also contribute lyses a significant genetic correlation did in fact exist
to the common genetic factors in Figure 13.2 [105, between heart rate and antisocial behavior. Analyses

155
Chapter 13: Antisocial behavior: gene–environment interplay

showed that the relationship between low resting of antisocial behavior [112]. The model proposes that
heart rate and antisocial behavior was significantly gene abnormalities lead to structural brain abnormal-
and entirely explained by common genetic influences, ities which result in emotional/cognitive/behavioral
although the heritable component of heart rate abnormalities, which in turn predispose to antisocial
explained only a small portion (1–4%) of the substan- behavior. There is increasing evidence for brain
tial genetic variance in antisocial behavior. Despite the impairments in antisocial individuals, with particu-
effect size being small, children with low resting heart larly strong evidence for the prefrontal cortex [113].
rate appear to be genetically predisposed towards exter- For example, neurological patients suffering damage
nalizing behavior problems as early as age nine [111]. to the ventral prefrontal cortex exhibit psychopathic-
Environmental mechanisms may also be investi- like, disinhibited behavior, autonomic and emotional
gated using the same “measured risk factor” approach blunting, and bad decision-making [114]. This indi-
in multivariate genetic models. As shown in Figure cates that there is a significant brain basis to antisocial
13.2, the common and specific environmental influ- behavior, and that these neurobehavioral processes
ences for the risk factor and antisocial behavior may are relevant to understanding the development of anti-
also be estimated. This approach may be used to help social behavior. The next step is to understand how
elucidate the nature of environmental influences brain structural/functional impairments translate into
important to antisocial behavior by determining, for the cognitive, emotional, and behavior risk factors
example, the extent to which certain measured social predisposing to antisocial behavior. For example, the
risk factors and antisocial behavior may have correl- amygdala is centrally involved in fear conditioning.
ated etiologies. Individual social risk factors that have Poor fear conditioning may result in a failure to fully
been identified as being important to antisocial develop a conscience – a set of conditioned emotional
behavior include various aspects of parenting, such responses that motivate individuals to desist from
as harsh discipline and monitoring or awareness of previously punished behavior. Poor conscience devel-
children’s activities and behaviors. These and other opment is in turn viewed as a predisposition to anti-
environmental factors need to be investigated in gen- social behavior. Similarly, ventral prefrontal damage
etically informative designs to determine the extent to results in disinhibited behavior that predispose to
which their effects may be moderated by individual lawless behavior.
genetic predispositions. Even though the “genes to brain to antisocial
behavior” model argues for a direct causal pathway
from genes to brain to antisocial behavior, the
Conclusions importance of environmental influences must be
Compelling evidence from twin and adoption stressed. This is well in line with results from twin
research show that at least half of the total variance and adoption studies, which consistently show that
in antisocial behavior is due to heritable factors environmental influences explain a substantial part of
[5, 94], and the heritability may be as high as 90% the total variance in antisocial behavior, explaining
for more refined, true score estimates of antisocial approximately more than 50% [5]. Environmental
behavior (e.g. Baker et al. [111]). Nonetheless, influences early in development could directly change
molecular genetic research has identified only a few gene expression, in turn altering brain functioning
specific genes associated with antisocial behavior or and resulting in antisocial behavior. Social environ-
its correlates. Clearly more research is needed to ment can interact with genetics and biological risk
identify which genes are involved in these heritable factors for antisocial behavior in different ways [77].
influences. Our understanding of the link between Results from adoption studies indicate that antisocial
genes and antisocial behavior also requires further behavior increases, perhaps even synergistically, when
investigation of the brain processes that underlie social and biological risk factors are combined. There
antisocial behavior throughout the lifespan. is also replicated evidence that an abnormality in
The “genes to brain to antisocial behavior” model the MAOA gene interacts with early child abuse in
hypothesizes that specific genes result in structural and predisposing to adult antisocial behavior [27].
functional brain alterations which in turn predispose Evidence of genetic influences on antisocial
to antisocial behavior. The environment is also con- behavior does not implicate that individuals exhibit-
sidered to play an important role in the development ing antisocial behavior are immune or resistant to

156
Chapter 13: Antisocial behavior: gene–environment interplay

interventions. The importance of genetic influences environmentally based methods which are tailored to
implies that biological processes are involved in the individuals based on their measured genetic risk.
etiology of antisocial behavior. However, it does not In summary, antisocial behavior underlies the
imply that a genetic influence on antisocial behavior diagnosis of several psychiatric disorders, for which
requires a biological intervention, but instead it may the genetic effects are clearly established. The substan-
require an environmental intervention. For example, tial genetic influence on antisocial and aggressive
a genetic liability to antisocial behavior may be best behavior itself may be at the root of the comorbidity
prevented or treated through, for example, parental among various externalizing disorders, including
and/or teacher training. Further, poor nutrition in the conduct and oppositional disorders and ADHD in
first three years of life has been associated with long- children, and antisocial personality disorder and sub-
term antisocial behavior throughout childhood and stance abuse in adults. Future research with combined
late adolescence [115]. Simple prevention programs approaches from behavior genetics and neuroscience
that manipulate nutrition early in life have resulted will lead to better understanding of specific genes that
in reduced delinquency [116] and criminality [117]. result in structural and functional brain impairments
Environmental manipulations can in theory reverse that in turn give rise to antisocial, violent, and psy-
brain risk factors for antisocial behavior. chopathic behavior. As in other domains of human
Understanding the complex interactions between behavior, a major challenge will be to understand the
specific genes and measured environments will pre- processes involved in gene expression, including ways
sumably give rise to the most effective prevention and the environment may exacerbate or protect individ-
treatment courses for these antisocial disorders. As in uals from risk. Longitudinal, genetically informative
other genetic disorders – such as phenyletonuria studies that involve sophisticated neuroscience
(PKU), which is entirely treated through an environ- methods such as brain imaging will undoubtedly be
mental (dietary) intervention – it is entirely possible the key to understanding individual pathways toward
that treatment avenues for genetically based psychi- psychiatric outcomes for which antisocial behavior is
atric disorders could potentially involve noninvasive, involved.

References 9. Mednick SA, et al. Science


1984;224:891–894.
20. Loeber R, et al. In Loeber R and
Farrington DP (eds.). Serious
1. American Psychiatric Association. and Violent Juvenile Offenders.
Diagnostic and Statistical Manual of 10. Cadoret RJ, et al. Behav Genet
1983;13:301–310. Risk Factors and Successful
Mental Disorders, Fourth Edition, Interventions. Thousands Oaks,
Text Revision (DSM-IV-TR). 11. Baker LA, et al. Behav Genet CA: Sage Publications; 1998.
Washington, DC: American 1989;19:355–370.
Psychiatric Association; 2004. 21. Rutter M, et al. Antisocial
12. Costello EJ, et al. Arch Gen Behavior by Young People.
2. Krueger RF, et al. J Abn Psychology Psychiatry 2003;60(8):837–844. Cambridge: Cambridge
2007;116(4):645–666.
13. Loeber R, et al. J Am Acad Child University Press; 1998.
3. Krueger RF, et al. J Abn Psychology Adolesc Psychiatry 2000;39(12): 22. Robins LN. Deviant Children
2005;114(4):537–550. 1468–1484. Grown Up. Baltimore: Williams
4. Young SE, et al. Am J Med Genet B 14. Costello EJ, et al. Arch Gen and Wilkins; 1966.
Neuropsych Genet 2000;96:684–695. Psychiatry 1996;53:1129–1136. 23. Scourfield J, et al. Arch Gen
5. Rhee SH, et al. Psychol Bull Psychiatry 2004;61:489–496.
15. Hipwell AE, et al. Crim Behav
2002;128:490–529.
Mental Health 2002;12(1):99–118. 24. Gelhorn HL, et al. J Child
6. Baker LA, et al. J Abn Psychology Psychol Psychiatry 2005;46(6):
16. Maughan B, et al. J Child Psychol
2007;116(2):219–235. 580–591.
Psychiatry 2004;45(3):609–621.
7. Cloninger CR, et al. In Mednick
17. Miles DR, et al. Am J Med Genet 25. Burt AS, et al. J Abn Psychology
et al. (eds.). The Causes of Crime:
2002;114:159–618. 2001;110(4):516–525.
New Biological Approaches.
Cambridge: Cambridge University 18. White HR, et al. Psychol Addict 26. Dick DM, et al. J Abn Child
Press; 1987. Behav 2001;15(3):210–218. Psychology 2005;33(2):219–229.
8. Cloninger CR, et al. Arch Gen 19. Kim-Cohen J, et al. Arch Gen 27. Caspi A, et al. Science 2002;
Psychiatry 1982;39:1242–1247. Psychiatr 2003;60(7):709–717. 297(5582):851–854.

157
Chapter 13: Antisocial behavior: gene–environment interplay

28. Cottle CC, et al. Crim Justice 48. Farrington DP, et al. Legal 67. Blonigen DM, et al. J Abn
Behav 2001;28(3):367–394. Criminol Psychol 1996;1: Psychology 2006;115(1):85–95.
29. Lipsey MW et al. In Loeber R 47–63. 68. Forsman M, et al. J Abn
and Farrington DP (eds.). Serious 49. Hicks BM, et al. Arch Gen Psychology 2008;117(3):606–617.
and Violent Juvenile Offenders. Psychiatry 2004;61:922–923. 69. Foley DL, et al. Arch Gen
Risk Factors and Successful 50. Merikangas KR, et al. Arch Gen Psychiatry 2004;61:738–744.
Interventions. Thousands Oaks, Psychiatry 1998;55(11):973–979.
CA: Sage Publications; 1998. 70. Kim-Cohen J, et al. Mol Psychiatry
51. Kendler K, et al. Arch Gen 2006;11:903–913.
30. Beyers JM, et al. J Abn
Psychiatry 2003;60:929–937. 71. Nilsson KW, et al. Mol Psychiatry
Child Psychology 2001;29(5):
379–381. 52. Hare RD. In Gray J, et al. (eds.). 2006;59:121–127.
Criminal Justice, Mental Health, 72. Haberstick BC, et al. Am J Med
31. Lynam DR. J Abn Psychology
and Politics of Risk. London: Genet B Neuropsych Genet
2000;109(4);563–574.
Cavendish Publishing; 2002. 2005;135B(1):59–64.
32. Sampson RJ, et al. Science
53. Cleckley HJ. The Mask of Sanity. 73. Bohman M, et al. Arch Gen
1997;277:918–924.
St. Louis: Mosby; 1941. Psychiatry 1982;39:1233–1241.
33. Eaves L, et al. J Child Psychol
54. Hare RD. The Hare Psychopathy 74. Cadoret RJ, et al. Arch Gen
Psychiatry 1997;38:965–980.
Checklist-Revised (PCL-R), 2nd Psychiatry 1995;52(11):916–924.
34. Tuvblad C, et al. J Abn Child edn. Toronto: Multi-Health 75. Rowe DC. In Plomin R, et al.
Psychology 2009;37(2):153–167. Systems; 2003. (eds.). Behavioral Genetics in the
35. Polanczyk G, et al. Am J Psychiatry 55. Hare RD. The Hare Psychopathy Postgenomic Era. Washington,
2007;164(6):856–858. Checklist-Revised Manual. DC: American Psychological
36. Spencer TJ, et al. J Pediatric Toronto: Multi-Health Systems; Association; 2003.
Psychology 2007;32(6): 1991. 76. Tuvblad C, et al. J Child Psychol
631–642. 56. Cooke DJ, et al. Psychol Assess Psychiatry 2006;47(7):734–743.
37. Hechtman L, et al. Am J 2001;13(2):171–188. 77. Raine A. J Child Psychol Psychiatry
Orthopsychiatry 1984;54:415–425 57. Kotler JS, et al. Clin Child Fam 2002;43(4):417–434.
38. Satterfield R, et al. Am J Psychiatry Psychol Rev 2005;8(4):291–325. 78. Raine A. J Abn Child Psychology
1982;139:795–798. 58. Andershed H, et al. In Blaauw E 2002;30(4):311–326.
39. Lahey BB, et al. J Nerv Ment Dis and Sheridan L (eds.). 79. Brunner HG, et al. Science
1979;167:734–741. Psychopaths: Current 1993;262(5133):578–580.
International Perspectives. The
40. Barkley RA. ADHD and the 80. Von Knorring A-L, et al. Alcohol
Hague: Elsevier; 2002.
Nature of Self-control. New York: Alcohol 1991;26:409–416.
Guilford Press; 1997. 59. Forth AE, et al. The Psychopathy
81. Shekim WO, et al. Psychiatry Res
Checklist: Youth Version. Manual.
41. Rhee SH, et al. J Abn Psychology 1989;27:81–88.
North Tonawanda, NY: Multi-
1999;108(1):24–41. 82. Buchsbaum MS, et al. Science
Health Systems, Inc.; 2003.
42. Rietveld MJ, et al. Am J Med 1976;194:339–341.
60. Lynam DR, et al. Ann Rev Clin
Genet B Neuropsych Genet 2003; 83. Gottfries CG, et al. J Neurochem
117(1):102–113. Psychol 2005;1:381–407.
1975;25:667–673.
43. Thapar A, et al. J Am Acad 61. Vitacco MJ, et al. Assessment
2003;10(2):143–150. 84. Comings DE. In Fishbein DH
Child Adolesc Psychiatry 2000; (ed.). The Science, Treatment, and
39:1528–1536. 62. Larsson H, et al. J Abn Psychology Prevention of Antisocial Behaviors:
44. Angold A, et al. J Child Psychol 2006;115(2):221–230. Application to the Criminal Justice
Psychiatry 1999;40:57–87. 63. Taylor A, et al. J Abn Child System Kingston, NJ: Civic
45. Biederman J, et al. Am J Psychiatry Psychology 2003;31(6):633–645. Research Institute; 2000.
1991;148(5):564–577. 64. Viding E, et al. J Child Psychol 85. Gade R, et al. Mol Psychiatry 1998;
46. Faraone SV, et al. J Am Acad Child Psychiatry 2005;46(6):592–597. 3:50–60.
Adolesc Psychiatry 1998;37(2): 65. Larsson H, et al. Psychol Med 86. Shih JC. Neurotoxicology 82;
185–193. 2007;37(1):15–26. 25(1–2):21–30.
47. Krueger RF, et al. J Abn 66. Blonigen DM, et al. Psychol Med 87. Raine A. The Psychopathology of
Psychology 2002;111:411–424. 2005;35:637–648. Crime: Criminal Behavior as a

158
Chapter 13: Antisocial behavior: gene–environment interplay

Clinical Disorder. San Diego: 98. Roy A, et al. In Emil F, et al. (eds.). 108. Raine A, et al. Dev Psychol 1996;
Academic Press; 1993. Serotonin in Major Psychiatric 32:624–630.
88. Kim JJ, et al. Proc Natl Acad Sci Disorders. Washington,
109. Farrington DP. In Raine A,
U S A 1997;94(11):5929–5933. DC: American Psychiatric et al. (eds.). Biosocial Bases of
Press; 1990.
89. Widom CS, et al. Biol Psychiatry Violence. New York: Springer-
2006;1(1):60. 99. Virkkunen M, et al. Archiv Gen Verlag; 1997.
Psychiatry 1989;46:600–601.
90. DeLisi M, et al. J Crim Justice 110. Ortiz J, et al. J Am Acad Child
2008;36:217–223. 100. Oades RD, et al. Behav Brain Adolesc Psychiatry 2004;43(2):
Funct 2008;4:48. 154–162.
91. Beaver KM, et al. Int J Offender
Ther Comp Criminol. 2007; 101. Thapar A, et al. Hum Mol Genet 111. Baker LA, et al. Dev Psychopathol
51(6):620–645. 2005;14:R275–R282. 2009;21(3):939–960.
92. Miller WB, et al. J Biosoc Sci 102. Caspi A, et al. Arch Gen Psychiatry 112. Raine A. Curr Direct Psychol Sci
1999;31(1):43–54. 2008;65(2):203–210. 2008;17(50):323–328.
93. Munafò MR, et al. Mol Psychiatry 103. Evans DM, et al. Biol Psychology 113. Raine A, et al. Soc Cogn Affect
2003;8(5):471–484. 2002;61:33–51. Neurosci 2006;1:203–213.
94. Moffitt TE. Psychol Bull 2005; 104. Gottesman II, et al. Am J 114. Damasio A. Descartes’ Error:
131(4):533–554. Psychiatry 2003;1160: Emotion, Reason, and the Human
636–645. Brain. New York: G P Putnam’s
95. Dick DM, et al. Arch Gen
Psychiatry 2008;65(3):310–318. 105. Krueger RF, et al. Behav Genet Sons; 1994.
1998;28(3):173–186. 115. Liu JH, et al. Am J Psychiatry
96. Åsberg M, et al. In Meltzer HY
(ed.). Psychopharmacology: 106. Maes HH, et al. Twin Res 2004;161:2005–2013.
The Third Generation of Progress. Hum Genet 2007;10(1): 116. Olds D. JAMA 1999;281:
New York: Raven Press; 1987. 136–150. 1377–1378.
97. Moore TM, et al. Aggressive Behav 107. Raine A, et al. Am J Psychiatry 117. Raine A, et al. Am J Psychiatry
2002;28:299–316. 1990;147:933–937. 2003;160:1627–1635.

159
Learning disabilities
Chapter

14 Shelley D. Smith

Learning disabilities: definitions of 6 year olds [5], is defined as unexplained deficits in


receptive and/or expressive language skills, without
and prevalence any evidence of deficits in nonverbal IQ, neurological
Clinically, learning disabilities are generally defined as impairment, or environmental or emotional prob-
a deficit in a given cognitive ability compared to the lems that could explain the language delays. These
“expected” ability based on intelligence, abilities in deficits can be seen in some or all of five global
other areas, and opportunity to learn. Specific reading language domains: phonology, morphology, syntax,
disability (SRD or dyslexia) is the most common semantics, and pragmatics [5–7] (see [8] for review).
learning disability with a prevalence of about 9% in There have been several theories regarding the com-
American school children [1], and has been defined as ponent skills necessary for language, and this has
a specific learning disability that is neurobiological in affected the endophenotypes used in genetic studies.
origin. It is characterized by difficulties with accurate There is growing consensus that nonword repetition
and/or fluent word recognition and by poor spelling is important, and this has been used as the diagnostic
and decoding abilities. These difficulties typically phenotype in some studies such as those by the SLI
result from a deficit in the phonological component Consortium [9, 10]. Mastering the morphology of
of language that is often unexpected in relation to grammar is also a core skill [11], and at least one
other cognitive abilities and the provision of effective study has shown that both may be needed to produce
classroom instruction. Secondary consequences may SLI [12].
include problems in reading comprehension and An additional language disability, speech sound
reduced reading experience that can impede growth disorder (SSD), concerns developmentally inappro-
of vocabulary and background knowledge [2]. priate articulation problems that interfere with intel-
Although the popular idea of dyslexia is that of visual ligibility. The prevalence is about 8% [13]. SSD is
confusion of letters, this definition emphasizes that sometimes included with SLI, and has also been
the auditory (phonologic) component is the primary divided into a phonologic disorder, considered to be
cause of reading disability. Furthermore, genetic stud- a central deficit, and developmental dyspraxia (or
ies have suggested that the auditory and visual routes childhood apraxia of speech), which is a motor dis-
to reading are not totally distinct pathways since they order confined to articulation [14]. It is not clear that
show association with the same genes [3, 4]. A number these are etiologically distinct disorders, however.
of studies have established the necessity for phonemic Overall, SSD shares a phonological component with
awareness in reading ability, and recent studies have reading disability, but also includes an articulation
also demonstrated that rapid serial naming is also an component.
important skill for reading (reviewed in [1]). These These definitions and subtypes raise a number of
endophenotypes, presumably closer to the etiology of questions. All of these disorders occur on a con-
reading problems, appear to be better measures for tinuum of severity, so the determination of an appro-
genetic studies of reading. priate cut-off for diagnosis is arbitrary and may be
Similarly, specific language impairment (SLI), the tied more to the requirements for eligibility for ser-
second most common learning disability at about 7% vices than to qualitative difference between affected

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

160
Chapter 14: Learning disabilities

and unaffected children. Similarly, the designation of around 0.56 [19–20]. Some authors have suggested
“specific” reading or language disability may not that this is due to a combination of many genes; in
reflect a qualitative difference in deficits; for example, particular, Plomin and Kovas [21] have developed a
the nature of the reading problems of children with “generalist gene” hypothesis, which proposes that
IQs in the normal range may not be different than there are many genes that influence overall cognitive
those of children whose (lower) IQs are similar to ability and that learning disabilities result when
their reading achievement. There may be etiological enough of these genes are disrupted in their function.
differences across IQ, however, especially when gen- The exact nature of the learning disability is then
etic causes are considered. Another question regards determined by environmental factors acting in com-
the use of diagnostic categories to determine under- bination with genetic susceptibility. In contrast, other
lying disability. Diagnosis of a particular learning studies have suggested that there is a finite number of
disability or its subtype may be helpful in classifying genes involved, and that different genes may have
deficits at a point in time, targeting them for reme- different phenotypic effects. Several types of studies
diation, but a diagnostic category may not stay the have supported the idea that there are a smaller
same over time, and thus may not reflect the under- number of genes influencing RD. Segregation ana-
lying etiology. Rather than being separate and dis- lyses have been consistent with the existence of a
tinct, learning disabilities are often comorbid in single major gene effect with reduced penetrance,
children, suggesting that there may be overlaps in although this would not necessarily be the same gene
deficits and etiologies. Through the identification of in different families, and the decrease in penetrance
genes influencing processes that are important to would presumably be due to interaction with add-
learning, new developments in genetic analysis may itional genes [22]. Another type of segregation analy-
help define alternate ways of conceptualizing different sis estimated 2–5 genes affecting different aspects of
types of learning disabilities based on the genes and reading [23, 24]. This distinction between a few or a
endophenotypes that are involved in each one. Since large number of genes has strong implications for
reading disability has been studied in depth, it can be studies to try to identify these genes; if there are very
used as an illustration of what is known about the many genes with individually small effects, very large
phenotypic, neurodevelopmental, and genetic aspects study populations with thousands of cases will be
of learning disability. needed to detect any one of the genes, but if there
are a smaller number of genes, smaller studies will be
sufficient.
Reading disability Gene localizations: The results of linkage and
Gender differences: Studies of clinical populations association analyses support the findings of segrega-
have consistently shown male : female sex ratios of 3– tion analysis, in that studies of relatively small popu-
5 : 1, but at least some of the increase in males appears lations of families have resulted in linkages and
to be bias of ascertainment since population-based associations implicating at least nine loci. These loci
studies showed sex ratios closer to 1–2 : 1 [15], leading are indicated by the prefix DYX and are numbered in
to the suggestion that boys tended to be more disrup- the order of their publication: DYX1, on chromosome
tive and thus more often referred for treatment. Stud- 15q21; DYX2 at 6p22; DYX3 at 2p16-p15; DYX4 at
ies of a large twin population indicated that heritability 6q13-q16.2; DYX5 at 3p12-q13; DYX6 at 18p11.2;
was different for males than for females [16], but a twin DYX7 at 11p15.5; DYX8 at 1p36; and DYX9 at
population with more detailed behavioral testing did Xq27.3. In addition, several other loci have been pro-
not support this conclusion [17]; rather, this study has posed: 2p11, 2q22.3, 12p13.3, 13q12, and 13q21. Some
found evidence that males showed greater variation in of these loci have been replicated in independent stud-
severity than females [18]. Since sex-linked genes do ies, particularly DYX1, DDYX2, DYX3, DYX5, and
not appear to play a prominent role in reading disabil- DYX8.
ity (RD), these differences could be due to some sex-
influenced effects of autosomal genes.
Genetics: There has been evidence since the turn Candidate genes for RD
of the last century that RD occurs in families, and Several candidate genes have been identified within
twin studies have consistently shown heritabilities DYX loci. While there have been conflicting results

161
Chapter 14: Learning disabilities

for nearly all of the candidates, with some studies Certain motifs in the sequence of DYX1C1 are
replicating an association and other studies finding consistent with this involvement in neuronal migra-
lack of support, several genes have consistent support tion, but another area in the DYX1C1 protein binds
which is backed up by functional analyses. to heat-shock proteins involved in cancers, and the
DYX1: 15q21, DYX1C1: The first region which gene was found to be upregulated in tumors [43, 44].
showed linkage to RD phenotypes was on chromo- These apparently unrelated functions may be recon-
some 15 [25–27]. Subsequently, a family was identi- ciled by a recent report that the gene has estrogen-
fied in which RD segregated with a translocation with receptor binding sites [45]. Since hormone receptors
breakpoints in the linkage region. One breakpoint are known to influence brain development, it is
disrupted a gene of unknown function, and single tempting to hypothesize that this gene could influence
nucleotide polymorphisms (SNPs) within the gene the gender differences that have been observed in RD.
showed association in a separate population [28]. This DYX2: 6p22, DCDC2, KIAA0319: Linkage of RD
gene has been given the name DYX1C1 (DYX1 Can- to the 6p22 region was one of the first replicated
didate 1), and has also been called EKN1. Independent linkages of a behavioral trait, and was a strong
follow-up studies initially failed to replicate the asso- impetus to the expansion of molecular genetic studies
ciation of RD with the same SNPs [29–31], but others of RD [27, 46–49]. Subsequent association analyses
found association with different SNPs in the gene [32, narrowed the region to a set of about five genes; VMP,
33]. Subsequently, additional studies replicated the DCDC2, KIAA0319, THEM2, and TTRAP [3, 4] with
association [34–36]. In the Italian population that most results centering around DCDC2 and KIAA0319
initially rejected association with RD, later studies [50–53]. The functions of these two genes are not
by Marino et al. [37] found association with DYX1C1 clearly established, but the homology of DCDC2
with short-term memory. (doublecortin 2) to the X-linked lissencephaly gene
One of the early questions in the genetic studies of DCX (doublecortin) made it an obvious candidate. In
RD was whether different genes would affect different the X-linked recessive disorder, males who are hemi-
component phenotypes. Examples of these phenotypes zygous for mutations in the DCX gene have a severe
were phonemic awareness, phonological coding, ortho- disruption of neuronal migration with an essentially
graphic coding, and word recognition. One linkage smooth cortex or sub-cortical heterotopias resulting
study of DYX1 suggested that the primary influence from a defect in microtubule stabilization [54, 55;
of this locus was on word recognition, but an editorial reviewed in 56]. This is associated with significant
accompanying that paper cautioned against conceptu- cognitive deficits and a severe seizure disorder. As
alizing these components as separate brain functions, would be expected, heterozygous females have a
particularly since the abilities were highly correlated somewhat milder phenotype, but still show cortical
[38]. Subsequent studies verified this caution, finding abnormalities and cognitive deficits. Based on this,
that the linkage distinctions between component one could postulate that less severe mutations involv-
phenotypes of reading could not be made [39]. On ing the autosomal DCDC2 gene might result in less
the other hand, the finding of linkage of short-term severe neuronal migration defects and a less severe
memory to DYX1C1 may indicate that there could be cognitive outcome. This has been supported by the
differences in effects on related cognitive abilities. finding that miRNA knockdown of DCDC2 in develop-
The first functional study of DYX1C1 examined ing rat brain results in delays in neuronal migration,
the effects of knockdown of gene expression in brain similar to that seen in DYX1C1 [52]. In addition,
development in the rat and demonstrated delays in sequencing of DCDC2 in disabled readers has not
neuronal migration [40]. These results are reminis- produced obvious causal disruptions in the coding
cent of the early anatomic studies of postmortem sequence, but a set of variants and a deletion in the
human brains which showed cortical heterotopias second intron has been associated with RD [52]. At
and microgyri indicative of migrational abnormalities least one of these includes a transcription factor binding
[41]. Interestingly, when treated rats were examined site. Support for association of the intronic deletion
postnatally, juvenile and adult rats showed cortical with RD has been conflicting, however [34, 57, 58].
microgyri. Depending on the location of the micro- Support for association of KIAA0319 with RD
gyri, they also showed deficits in auditory processing has been a little more consistent, particularly with
and spatial learning [42]. SNPs in the 5’ and first intronic regions of the gene

162
Chapter 14: Learning disabilities

[4, 50, 51, 57, 59]. Association has also been found in a large family showing an autosomal dominant pat-
with SNPs in KIAA0319 and variation in reading abil- tern of inheritance of severe reading disability. Subse-
ity in the normal range [60]. As with DCDC2, sequen- quently, an adult male was found with reading disability
cing has not found mutations in the coding region, and a translocation in that region of chromosome 3.
and the positions of the associated SNPs in the intronic Although his sister had also been diagnosed with dys-
and 5’ regions suggest that the effect is in gene regula- lexia and did not share the translocation, testing of both
tion [61]. Indeed, one study showed that individuals individuals was not possible, so the authors considered
with a “risk haplotype” of SNP alleles had decreased it likely that there could be another cause for learning
expression of KIAA0319 in lymphoblastoid cells [51], problems in the sister. The translocation disrupted the
and more recently a particular SNP in a regulatory ROBO1 gene, which is homologous to the Robo (round-
region of the gene has been associated with decreased about) gene in the fruitfly, mouse and rat and is
gene expression [59]. As with DYX1C1 and DCDC2, involved in the migration of axons across the midline
miRNA knockdown experiments in developing rat of the brain and spinal cord. The ROBO1 gene was
brain have shown that decreased expression of this sequenced in affected individuals from the large linkage
gene also produces delays in neuronal migration [51]. family, but no mutations were found in the coding
Some studies of RD find association primarily with region. A SNP haplotype was identified which marked
DCDC2 and others primarily with KIAA0319. If both the chromosome segregating with RD in the family, and
are real, the differences between studies could be due to analysis of allelic expression in lymphoblasts from sev-
population stratification or in diagnostic differences. It eral affected individuals indicated that the risk haplo-
seems most likely that the small populations in most of type was underexpressed. Allelic expression of nearby
the studies have magnified nonsignificant differences genes was normal, leading to the identification of
in association, but it is interesting that the studies that ROBO1 as a candidate gene [63]. This appears to be a
find evidence for DCDC2 and not KIAA0319 have rare form of reading disability, however, since the local-
tended to be the studies from Germany, with a focus ization has not been replicated in other families.
on spelling ability, and those with stronger evidence Overall, there is evidence that disorders of neuronal
for KIAA0319 have tended to be English and focused migration affect RD, and Galaburda et al. [64] have
on reading phenotypes. An exception is the Colorado suggested a mechanism whereby DYX1C1, DCDC2,
RD twin population, which found associations in both KIAA0319, and ROBO1 proteins interact in neuron
genes [3]. Recently, in a study of children ascertained and axon guidance. Several other candidate genes have
for attention-deficit hyperactivity disorder (ADHD), also been suggested for reading disability or related
association of ADHD was found with DCDC2 and the phenotypes. MRPL19 and C2ORF3 are in the region
adjacent VMP gene but not with KIAA0319. This 2p12, which may be the locus responsible for the linkage
population did not show association of RD phenotypes signal for DYX3. The original linkage was to markers
to either gene [62]. Interestingly, Pennington and around 2p16–15, however, so it is possible that these are
Bishop [1] have postulated that rapid serialized naming actually separate loci. Association was found to SNPs
may be the common deficit explaining comorbidity very close to these two genes in Finnish and German
between RD and ADHD, so perhaps this endopheno- families, and it was hypothesized that they are jointly
type is a more accurate measure of the phenotype regulated [65]. It is not known how these genes might
of the gene. Further studies are going to be required influence RD; MRPL19 is a mitochondrial ribosomal
to determine whether these two genes actually have protein and the function of C2ORF3 is unknown. In
different phenotypic effects. Since these differences another study, SNPs in SLC2A3 (GLUT3) at 12p13.3
may be more quantitative than qualitative and the was found to be associated with mismatch negativity in
phenotypes are correlated, it may be very difficult to dyslexic children in a genome-wide association study
establish differential influences of the two genes with- (GWAS) of 200 German families. This region had not
out larger populations. A further complicating factor been detected in previous linkage studies [66]. The gene
is that the two genes may interact and may have a product is known to be a glucose transporter, and thus
common regulatory system [57]; in that case, mutations could represent a mechanism besides neuronal migra-
affecting either gene should have similar outcomes. tion that can affect RD. The use of an electrophysiologic
DYX5: 3p12.3, ROBO1: Linkage of RD to the endophenotype is also an important development in the
centromeric region of chromosome 3 was first reported investigation of the causal mechanisms.

163
Chapter 14: Learning disabilities

Language impairment influence on reading with lesser influence on LI, so


that their association was missed in genome-wide
Although the definitions of RD and language impair-
studies for LI. Similarly, the SLI1 and SLI2 loci have
ment (LI) are based on measures that are very differ-
primary effects on LI phenotypes, but the SLI2 locus
ent, the disorders share some similarities; the
also has a lesser effect on reading which was not
heritabilities for deficits are similar [67], males are
detectable in genome scans for RD.
affected more often than females, and young children
with LI are at greater risk for RD [1, 68]. Although
limited studies of gene segregation have been done, Candidate genes for LI
analyses have not found the evidence for major genes SLI1: 16q23–24, CMIP; 16q24.1, ATP2C2: Two
[69] that have been seen for RD. Despite this, genes have been cited as independent candidates for
genome-wide linkage studies were done and resulted the SLI1 locus, CMIP and ATP2C2 [73]. Consistent
in localizations of gene effects. Those loci were not in with the linkage results, the primary phenotype that
the same regions as RD loci, suggesting that that showed association with SNPs in those genes was
language impairment may have a separate etiology. phonological short-term memory (nonword repeti-
The largest linkage studies were done by the SLI tion), and there was no evidence of association with
Consortium [9, 10], and resulted in two loci, SLI1 reading measures. Association was found in two inde-
on 16q24 and SLI2 on 19q13. Follow-up studies pendent populations selected for LI, but was not
narrowed the regions on each chromosome and char- found in an unselected population, indicating that
acterized the phenotypes [70]. The SLI1 locus showed the genes may not affect variation of nonword repeti-
association with phonological short-term memory, tion in the normal range. The mechanism of effect on
while the SLI2 locus showed association with expres- LI for these genes is unknown, although the functions
sive language, a grammatical measure of the ability to of these genes allow some speculation. Mutation of
use past tense, and literacy. Phonological short-term the CMIP gene can cause a renal disorder called
memory, often measured by nonword repetition, and minimal change nephritic syndrome which results
grammatical measures such as tense marking [11] from truncation of the gene product [74], but the
have been used as primary component phenotypes gene is also known to be expressed in the brain and
for LI, and earlier genetic studies indicated that they additional functions are being discovered. The gene
might be under separate genetic control [12]. More- product has been found to interact with filamin-A
over, based on their analysis of the published litera- [75], which suggests a role in the formation of den-
ture, Pennington and Bishop [1] postulated that dritic spines, and is involved in the control of NF-kB
deficits in both abilities are necessary for SLI. [76], which is also important in synaptic activity [73].
The association of reading measures to the The protein produced by ATP2C2 is a transporter of
chromosome 19 SLI2 locus also demonstrates that calcium and manganese ions and localizes to intracel-
LI and RD can overlap in etiology, despite the fact lular vesicles, where it is thought to affect calcium
that these overlaps were not seen in the genome homeostasis necessary for neuronal and synaptic
screens for each disorder. Evidence for overlapping function. Abnormalities in magnesium ion homeo-
genetic influences was also seen in a smaller linkage stasis have been associated with neurodegenerative
study of LI which found linkage to chromosome disorders and memory problems [73].
13q21 in 5 extended families with LI, but the linked SPCH1: 7q31, FOXP2: The influence of the
phenotype was a reading measure rather than a lan- FOXP2 gene on speech and language was discovered
guage measure [71]. Finally, linkage studies that through a unique kindred known as the KE family in
targeted RD loci in families ascertained for LI repli- which severe language impairment and oromotor
cated the linkage to the DYX8 and DYX2 regions, and apraxia segregated as a dominant trait, found to be
association analysis in these LI families replicated due to a missense mutation in the FOXP2 gene [77].
association of language and reading phenotypes to A few additional children with developmental verbal
several of the same SNPs in KIAA0319 that showed dyspraxia have been identified with disruptions (either
association with RD [72]. It appears, then, that there translocations or mutations) of FOXP2 [77, 78], but
are loci that affect both RD and SLI, but the magni- studies of larger LI populations have indicated that
tude of the effect differs; the DYX loci have greater mutations in the gene are not a common cause of LI.

164
Chapter 14: Learning disabilities

Most studies have looked within the coding region, substitutions resulted in changes in the basal ganglia
however, and some evidence has been found for link- and increased synaptic plasticity as well as some changes
age or association near the gene [72, 79], so changes in in ultrasonic vocalizations [83].
regulatory regions cannot be ruled out. More import- CNTNAP2: The CNTNAP2 gene is located distal
antly, as a gene regulator itself (forkhead-winged helix to FOXP2 on 7q35. It was first identified as candidate
transcription factor), it can be a key to a network of for autism through genome screens and expression
genes that affect language [80]. The findings with the studies [84, 85]. In the study by Alarcón et al., age
gene CNTNAP2, discussed below, are an example. at first word was used as a quantitative phenotype.
Structural MRI characterization of the KE family The gene has several possible functions in neuronal
demonstrated that the gene mutation resulted in a migration and synaptic function, and mutations in
25% decrease in volume of the caudate nuclei, and CNTNAP2 have been associated with mental retard-
the extent of decrease correlated with the degree of ation, severe epilepsy, and cortical dysplasia [86, 87].
oral dyspraxia [81]. Abnormalities were also noted in In an independent study designed to identify a genetic
cortical regions. Language and oral motor areas, network influencing LI, Vernes et al. [88] identified
including Broca’s area and the ventral portion of the CNTNAP2 as a downstream target of FOXP2. They
precentral gyrus, showed decreased gray matter, and demonstrated significant association between nine
these differences were reflected in decreased activation intronic SNPs in the gene and performance in a non-
in these areas seen with fMRI. Overall, the deficits word repetition task in families with a child with LI.
appeared to be in frontostriatal and frontocerebellar Further studies will undoubtedly expand upon the
networks subserving language and oromotor skills. function of the FOXP2–CNTNAP2 network.
Studies of the function of FOXP2 also illustrate an
important evolutionary principle: language ability is
likely to be built on existing neurodevelopmental Speech sound disorder
systems, rather than on new genes that are unique to Speech sound disorder (SSD) involves both language
humans [80]. The FOXP2 gene is found to be conserved and articulation deficits and is often grouped with LI.
across species; for example, the homologous gene in It is also often comorbid with RD. Pennington and
zebra finches is expressed in a similar pattern in the Bishop [1] have pointed out that the comorbidity
brain and is involved in the learning of song [82]. between SSD and RD appears primarily when SSD is
Furthermore, decreased expression of the gene caused included with LI. In their analysis of published results,
deficits in song learning [82]. In mice, knockouts of the relative risk for RD in children with earlier SSD þ
FoxP2 or introduction of the human KE family muta- LI ranged from 4.6 to 8.9, but the rate of RD in
tion caused severe motor problems and decreased children with SSD alone was negligible. Unfortu-
cerebellar size in homozygotes. Phenotypic effects in nately, published data for RD in children with LI
heterozygotes were less clear, possibly depending upon without SSD was inconsistent; in one study, the rate
the genetic background of the mice, indicating that of RD was 3.2–3.6, while the other LI study had a rate
interaction with other genes is important. Some reports of RD of only 0.5. Thus, SSD without LI may be
had indicated that vocalizations of heterozygous pups separate from SSD + LI. This is supported by a study
were abnormal, but in an extensive review, Fisher and which showed that the heritability of LI was largely
Scharff [80] have noted that these may be secondary to driven by the heritability of LI + SSD [89].
other developmental problems. They argue that, rather Molecular genetic studies support common gen-
than assuming that the gene is important for communi- etic influences on RD and SSD. Linkage analyses
cation across species, the evolutionary function of have demonstrated linkage of SSD phenotypes to
FOXP2 could be conceptualized as facilitating synaptic the DYX5/ROBO1 region on chromosome 3 [90],
plasticity for planning and coordination of very rapid to the DYX2 region on chromosome 6 and to the
movements. Such abilities would be fundamental to the DYX1C1 region on chromosome 15 [91] as well as
production of words, and changes affecting this type of to a more proximal region on 15 [92], and to the
gene could be the basis for the evolution of language. DYX8 region on chromosome 1 [93]. These linkages
While such changes have not been identified yet, two were found with articulation phenotypes and phono-
“human specific” amino acid substitutions have been logical phenotypes, but it is unknown whether these
noted in FOXP2, and when introduced into mice, these results were driven primarily by children who later

165
Chapter 14: Learning disabilities

developed RD. More fine-grained association studies Such studies could have a high payoff. As the genes
have not yet been done for most of these regions, so it is influencing learning disabilities are defined, they will
not certain that the localizations to RD regions mean lead to the neurodevelopmental pathways, such as
that the same genes that affect RD also affect SSD. neuronal migration, involved in reading, language, or
articulation, and may demonstrate where the pathways
Conclusions for these disorders converge and diverge. Genetic stud-
ies also suggest that endophenotypes are more valuable
Studies of RD, LI, and SSD have demonstrated that
than clinical definitions in determining the etiology of
subjects with these conditions have overlapping cog-
these disorders. As the cognitive overlaps between
nitive deficits, and the distinctions between them may
disorders are worked out, e.g. the relative contribu-
depend upon particular combinations of deficits (as
tions of phonology, rapid naming, grammar, or articu-
measured by endophenotypes) that produce greater
lation to each disorder (e.g. [1]), concurrent genetic
liability for one disorder over another [94]. Molecular
studies can confirm which endophenotypes have
genetic studies support this concept, with some genes
common genetic effects and which are distinct. By
appearing to primarily affect one endophenotype,
starting with etiology and working up to clinical
such as the linkage of nonword repetition to SLI1,
phenotype, it is likely that the current diagnostic cri-
some affecting just a few, such as the effects of SLI2
teria may be modified. For example, is LI + SSD etio-
on nonword repetition and literacy, and other genes
logically distinct from LI alone or SSD alone, and
appearing to affect several endophenotypes and sev-
should diagnostic criteria take this into account?
eral disorders, such as the association of KIAA0319 to
Should LI be diagnosed based on nonword repetition
reading and language measures in RD and LI popula-
alone, or must measures of grammar also be included?
tions. To characterize the effects of individual genes
Can RD be diagnosed solely by assessment of phon-
on the endophenotypes and their resulting clinical
emic awareness or decoding, or will the inclusion of
manifestations, studies will need to be done with the
rapid serial naming result in dignosis of a more com-
same phenotypes and genotypes in all three disorders,
prehensive disorder with a distinct etiological path-
with large enough populations to reliably distinguish
way? Better definition of the cognitive components of
any qualitative influences of the candidate genes on
the three disorders should lead to optimal diagnostic
different phenotypes. Given the differences in the ages
and treatment procedures based on deficits that are
at which RD, LI, and SSD are usually diagnosed,
closer to etiology than clinical symptoms alone.
longitudinal studies would be especially valuable.

References Genetics, 4th edn. New York:


Elsevier; 2007.
17. Hawke JL, et al. Dyslexia 2006;
12(1):21–29.
1. Pennington BF, et al. Annu Rev
Psychol 2009;60:283–306. 9. SLI Consortium. Am J Hum Genet 18. Hawke JL, et al. Dyslexia 2009;
2002;70(2):384–398. 15(3):239–242.
2. Lyon GR, et al. Ann Dyslexia
2003;53:1–14. 10. SLI Consortium. Am J Hum Genet 19. DeFries JC, et al. Mental Retard
2004;74(6):1225–1238. Devel Disab Res Rev 1996;
3. Deffenbacher KE, et al. Hum Genet 2:39–47.
2004;115(2):128–138. 11. Rice ML, et al. J Speech Hear Res
1996;39(6):1239–1257. 20. DeFries JC, et al. In Plomin R,
4. Francks C, et al. Am J Hum Genet et al. (eds.). Nature, Nature, and
2004;75(6):1046–1058. 12. Bishop DV, et al. Genes Brain Psychology. Washington, DC:
Behav 2006;5(2):158–169. APA Press; 1993.
5. Tomblin JB, et al. J Speech Lang
Hear Res 1997;40(6):1245–1260. 13. Shriberg LD, et al. J Speech Lang 21. Plomin R, et al. Psychol Bull
Hear Res 1999;42:1461–81. 2005;131(4):592–617.
6. Tallal P. In Adelman G (ed.).
Encyclopedia of Neuroscience. 14. Sanger TD, et al. Pediatrics 22. Pennington BF, et al. JAMA
Vol. 1. Boston: Birkhauser; 1987. 2006;118(5):2159–2167. 1991;266(11):1527–1534.
7. Leonard LB. Appl Psycholinguistics 15. Shaywitz SE, et al. JAMA 1990;264 23. Wijsman EM, et al. Am J Hum
1989;10:179–202. (8):998–1002. Genet 2000;67(3):631–646.
8. Smith SD, et al. In Rimoin DL, et al. 16. Harlaar N, et al. J Child Psychol 24. Chapman NH, et al. Am J Med
(eds.). Emery and Rimoin’s Psychiatry 2005; Genet B Neuropsychiatr Genet
Principles and Practice of Medical 46(4):373–384. 2003;121B(1):60–70.

166
Chapter 14: Learning disabilities

25. Smith SD, et al. Science 1983; 48. Gayán J, et al. Am J Hum Genet 71. Bartlett CW, et al. Am J Hum
219(4590):1345–1347. 1999;64(1):157–64. Genet 2002;71(1):45–55.
26. Smith SD, et al. Reading Writing: 49. Fisher SE, et al. Am J Hum Genet 72. Rice ML, et al. J Neurodev Disord
Interdis J 1991;3:285–298. 1999;64(1):146–56. [epub ahead of print].
27. Grigorenko EL, et al. Am J Hum 50. Cope N, et al. Am J Hum Genet 73. Newbury DF, et al. Am J Hum
Genet 1997;60(1):27–39. 2005;76(4):581–591. Genet 2009;85(2):264–272.
28. Taipale M, et al. Proc Natl 51. Paracchini S, et al. Hum Mol 74. Grimbert P, et al. J Exp Med
Acad Sci U S A 2003;100 Genet 2006;15(10):1659–1666. 2003;198(5):797–807.
(20):11553–11558. 52. Meng H, et al. Proc Natl Acad Sci 75. Grimbert P, et al. Mol Immunol
29. Marino C, et al. Eur J Hum Genet U S A 2005;102(47):17053–17058. 2004;40(17):1257–1261.
2005;13(4):491–499. 53. Schumacher J, et al. Am J Hum 76. Kamal M, et al. Mol Immunol
30. Meng H, et al. Hum Genet Genet 2006;78(1):52–62. 2009;46(5):991–998.
2005;118(1):87–90. 54. des Portes V, et al. Hum Mol 77. Lai CS, et al. Nature 2001;
31. Bellini G, et al. J Mol Neurosci Genet 1998;7(7):1063–1070. 413(6855):519–523.
2005;27(3):311–314. 55. Sossey-Alaoui K, et al. Hum Mol 78. MacDermot KD, et al. Am J Hum
32. Wigg KG, et al. Mol Psychiatry Genet 1998;7(8):1327–1332. Genet 2005;76(6):1074–1080.
2004;9(12):1111–1121. 56. Leventer RJ. J Child Neurol 79. O’Brien EK, et al. Am J Hum
33. Scerri TS, et al. J Med Genet 2005;20(4):307–312. Genet 2003;72(6):1536–1543.
2004;41(11):853–857. 57. Harold D, et al. Mol Psychiatry 80. Fisher SE, et al. Trends Genet
34. Brkanac Z, et al. Am J Med 2006;11(12):1085–1091. 2009;25(4):166–177.
Genet B Neuropsychiatr Genet 58. Ludwig KU, et al. Psychiatr Genet 81. Vargha-Khadem F, et al. Nat Rev
2007;144B(4):556–560. 2008;18(6):310–312. Neurosci 2005;6(2):131–138.
35. Bates TC, et al. Mol Psychiatry 59. Dennis MY, et al. PLoS Genet 82. Haesler S, et al. J Neurosci 2004;
2010;15(12):1190–1196. 2009;5(3):e1000436. 24(13):3164–3175.
36. Dahdouh F, et al. Psychiatr Genet 60. Luciano M, et al. Biol Psychiatry 83. Enard W, et al. Cell 2009;137(5):
2009;19(2):59–63. 2007;62(7):811–817. 961–971.
37. Marino C, et al. Genes Brain 61. Couto JM, et al. Am J Med Genet 84. Alarcón M, et al. Am J Hum Genet
Behav 2007;6(7):640–646. B Neuropsychiatr Genet 2008;82(1):150–159.
38. Pennington BF. Am J Hum Genet 2010;153B(2):447–462. 85. Arking DE, et al. Am J Hum Genet
1997;60(1):13–16. 62. Couto JM, et al. Biol Psychiatry 2008;82(1):160–164.
39. Grigorenko EL, et al. Am J Hum 2009;66(4):368–75. 86. Strauss KA, et al. N Engl J Med
Genet 2000;66(2):715–723. 63. Hannula-Jouppi K, et al. PLoS 2006;354(13):1370–1377.
40. Wang Y, et al. Neuroscience Genet 2005;1(4):e50. 87. Zweier C, et al. Am J Hum Genet
2006;143(2):515–522. 64. Galaburda AM, et al. Nat Neurosci 2009;85(5):655–666.
41. Galaburda AM. Neurol Clin 2006;9(10):1213–1217. 88. Vernes SC, et al. N Engl J Med
1993;11(1):161–173. 65. Anthoni H, et al. Hum Mol Genet 2008;359(22):2337–2345.
42. Threlkeld SW, et al. Brain Res Bull 2007;16(6):667–677. 89. Bishop DVM, et al. Genes Brain
2007;71(5):508–514. 66. Roeske D, et al. Mol Psychiatry Behav 2008;7:365–372.
43. Chen Y, et al. J Cancer Res Clin 2011;16(1):97–107. 90. Stein CM, et al. Am J Hum
Oncol 2009;135(9):1265–1276. 67. Spinath FM, et al. Child Dev 2004; Genet 2004;74(2):283–297.
44. Kim YJ, et al. J Cancer Res Clin 75(2):445–454. 91. Smith SD. J Child Psychol
Oncol 2009;135(2):265–270. 68. Catts HW, et al. J Speech Psychiatry 2005;46(10):1057–1066.
45. Massinen S, et al. Hum Mol Genet Lang Hear Res 2002;45(6): 92. Stein CM, et al. Behav Genet
2009;18(15):2802–2812. 1142–1157. 2006;36(6):858–868.
46. Cardon LR, et al. Am J Hum Genet 69. Lewis BA, et al. Behav Genet 93. Miscimarra L, et al. Hum Hered
1994;55(4):825–833. 1993;23(3):291–297. 2007;63(1):47–58.
47. Cardon LR, et al. Science 1995; 70. Falcaro M, et al. Genes Brain 94. Pennington BF. Cognition 2006;
268(5217):1553. Behav 2008;7(4):393–402. 101:385–413.

167
Attention-deficit hyperactivity disorder
Chapter

15 Josephine Elia, Francesca Lantieri, Toshinobu Takeda, Xiaowu Gai,


Peter S. White, Marcella Devoto, and Hakon Hakonarson

Introduction were suggested to be responsible for the increased


familial aggregation by adoption studies, reporting
Attention-deficit hyperactivity disorder (ADHD) is 18% ADHD in biological versus 6% in adoptive
one of the most common neuropsychiatric disorders parents [18].
with a best estimate prevalence of 5–10% in school Twin studies have attempted to distinguish
age children [1], 4% in college students [2] and between genetic and environmental influences with
 2.5% in adulthood [2, 3]. Relatively consistent reported heritabilities ranging between 60 and 100%
prevalence rates have been reported for North for an overall heritability of 0.76 reported in a
America and Europe [4]. review of earlier studies [19]. Additional recent
ADHD is characterized by developmentally studies, summarized in Figure 15.1, estimate herit-
inappropriate levels of hyperactivity, impulsivity and ability at 29–90% [20–27]. The lower heritability
inattention as well as executive function deficits ranging between 29 and 37% have been reported
leading to significant learning, behavioral and social primarily in studies that included older subjects
impairments throughout the lifespan [5]. Children [28–33]. Parental rating bias [34, 35] can mask
with ADHD are reported to have higher rates of other nonadditive genetic contributions, leading to
academic, emotional, conduct, and social difficulties over-estimation of the heritability in younger
[6]. Compared to non-ADHD peers, young adults cohorts [34, 36].
with ADHD have poorer academic outcomes [5, 7], The rest of the variance is explained by shared
work performance [8], and self-esteem [9]. In add- and nonshared environmental influences [20, 23,
ition they have increased rates of smoking [10], drug 37, 38] and non-additive genetic influences, such
use [11], and automobile accidents [12]. They also as a dominant effects [25] resulting from epigenetic
have more social problems including fewer close factors [39]. Environmental effects [40] may
friends, more dating partners and shorter duration also play a role during the different developmental
of dating relationships, more sex partners; they stages [41]; therefore, a cohort effect may also be
become parents earlier (38% versus 4% of controls important.
by age 25), and have higher rates of sexually
transmitted diseases (16% versus 4%) [5].
Genome-wide linkage studies
Family and twin studies Summarized in Table 15.1 are the ADHD linkage
Family, twin, and adoption studies have unequivo- studies with logarithm of odds (LOD) scores > 1.5
cally demonstrated that ADHD is influenced by gen- conducted in sibling pairs [33, 42–48], isolated popu-
etic factors. A higher risk for ADHD has been lations [49, 50], unrelated extended pedigrees [51, 52],
reported in siblings of ADHD probands (20.8% and affected members of single families [53]. Several
versus 5.6% in controls) [13], among first-degree chromosomal regions with potential linkage, some
family members of ADHD male [14] and female overlapping in two or more studies have been identi-
probands [15, 16], and in second-degree relatives fied including 2p, 5p, 7p, 16p, 17p, 5q, 9q, 11q, 12q,
[17]. Shared genes, rather than shared environment, 14q, 16q, and 18q. In particular, a pooled analysis of

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

168
Chapter 15: Attention-deficit hyperactivity disorder

Rietveld 2004 3–12y

Laarson et al. 2004 13–14y

Saudino et al. 2005 7y

Price et al. 2005 2y *1

Dick et al. 2005 14y *1

Kuntsi et al. 2005 8y *1

Knopik et al. 2005 11–23y

Ehringer et al. 2006 12–19y

Schultz et al. 2006 middle-age

Haberstick et al. 2008 22y

Ouellet-Morin et al. 2008 6–7y

Derks et al. 2008 12y


*2
Young et al. 2009 12y

Young et al. 2009 17y *2

0.00 0.20 0.40 0.60 0.80 1.00


Heritability
Figure 15.1 Heritabilities reported from 2004–09 for attention-deficit hyperactivity disorder (ADHD). The estimates reported here regard
the ADHD diagnosis, with distinguishing for subtypes. The darker bars refer to parental or teacher reports, while the lighter bars refer to
self-reports. *1 and *2 indicate that the same sample has been investigated at successive follow up (longitudinal studies).

the US and Dutch data identified a single region of Dopaminergic pathways


overlap at 5p13 [54] and a meta-analysis identified a
We will review candidate gene-specific investigations
region with genome-wide significance on chromo-
beginning in 1995 when Cook and colleagues [62]
some 16 between 64 and 83 Mb [55].
reported a significant association of ADHD with
Overall, the linkage analysis studies conducted to
SLC6A3–10R (DAT1), a variant of the dopamine
date have achieved only limited success in identifying
transporter gene, located in the perisynaptic regions
genetic determinants of common complex diseases;
of neurons [63, 64] and involved in recycling
this may be due to the generic problem that the
dopamine back into the releasing neurons. SLC6A3
linkage analysis approach has generally low power in
is expressed predominantly in basal ganglia where
identifying common genetic variants that have
it preferentially influences caudate volume [42].
modest effects [56, 57].
In humans, the 3’ untranslated region (UTR) of the
gene has a 40 base pair tandem repeat polymorphism
Selected candidate gene studies (VNTR) consisting of 3–11 copies [65]. As reviewed
The initial search for ADHD genes was hypothesis- by Barr and Misener, [66] and Gizer et al. [67] since
driven and focused on genes involved in neurotrans- Cook’s initial report, replication studies of SLC6A3–
mission, based on evidence from effective pharma- 10R have been inconsistent, and meta-analyses and
cotherapeutic agents [58], animal models [59, 60], analyses of pooled odds ratios are in the range
and neuroimaging studies [61]. Numerous studies of 1.03–1.27 suggesting a modest contribution to
have thus far been conducted and a summary of disease risk.
candidate gene polymorphisms associated with ADHD Additional polymorphisms that may confer risk
identified in the most recent meta-analysis is presented include single nucleotide polymorphisms (SNPs)
in Table 15.2. rs27072, a VNTR in intron 8 [66, 67]. Recently, a

169
Chapter 15: Attention-deficit hyperactivity disorder

Table 15.1 Genome-wide linkage studies.

Studies Sample Chromo- ADHD Closest Position LOD Results


(Year) some candidate marker (cM)
gene
Fisher et al. 126 ASP 5p12 SLC6A3 D5S418 68 1.04 None exceeded
(2000) 104 families (USA) 10q26 D10S212 193 1.66 genome-wide
12q24 D12S79– 131 1.16 significance
D12S324 165 1.099 thresholds
12p13 SLC2A1 D12S1725– 10 1.51
D12S336 22 2.60
16p13 D16S3075 14 1.51
Ogdie et al. 270 ASP 5p13 SLC6A3 16p13 (GRIN2A)
(2003, 2004) 204 families 6q14 (GRIN2A) was significant
(160 from Fisher 11q25
et al. [2000]) 16p13
17p11
20q13
Bakker et al. 238 ASP 5p13 SLC6A3 D5S2500 10 1.43
(2003) (106 families; 7p13 DDC D7S1818 3.04
the Netherlands) 9q33 DBH D9S1825 2.05
13q33 D13S796 1.91
15q15 GATA50C03 3.54
Loo et al. 283 ASP 16p D16S3046 18 3.73 Reading disability
(2004) (expanded 17q D17S787 46 2.98 and ADHD
sample from 10q D10S196 59 1.26
Ogdie et al. 2004)
Ogdie et al. 308 6q12 D6S430 83 3.30 Suggestive
(2004) (269 from Bakker 16p13 D16S3114 23 3.73 evidence for
et al. 2003) 17p11 D17S947 38 3.63 5p13
Arcos-Burgos 375 4q13.2 D4S409 81.6 2.4 Significant linkage
et al. (2004) (Columbian 5q33.3 D5S2117 138.5 1.5 for all families
genetic isolate 8q11.23 D8S1110 67 3.2 together at these
community: 16 11q.22 D11S1377 129 2.4 sites. However,
multigenerational, 17p11 D17S1876 11.9 3.4 individual families
extended D17S678 16.9 3.5 had significant
pedigrees) D17S1881 17.9 3.9 linkages at
D17S1844 22.8 3.8 different sites
D17S1791 25.1 3.7 (Family 9 at
4q13.2, 5q33.3,
8q11.23; Family 8
at 11q22; Family
14 at 17p11.)
Hebebrand 229 ASP 6q D6S1053 75 0.58 No linkage for
et al. (2006); 102 families 7p D7S2490 88 0.92 SLC6A3 VNTR;
Valera et al. (Germany) 9q D9S1851 104 0.68 Inattentive
(2007) 11q D11S4176 94 94 subtype more
12q D12S392 166 2.10 strongly
17p D17S1308 1 1.39 influenced by
gene in 5p region

170
Chapter 15: Attention-deficit hyperactivity disorder

Table 15.1 (cont.)

Studies Sample Chromo- ADHD Closest Position LOD Results


(Year) some candidate marker (cM)
gene
Faraone et al. 601 ASP 8 54.2 1.85 No significant
(2008) (217 families) regions
1000 sibs 8 93.4 0.8
(260 families)
Doyle et al. 1212 individuals 3q13 RS2062834 115.4 2.11 Suggested
(2008) (271 families) linkage with
neurocognitive
tasks and
inattentive
symptoms
Zhou et al. (2545 ASP 1p36 45 3.2 Significant
(2008) (1094 families) 9p23 27 2.2 linkage at 1p36
11q21 100 2.6
Abbreviations: ASP, affected sibling pair; cM, centiMorgan; VNTR: variable number of tandem repeats.

Table 15.2 Candidate attention-deficit hyperactivity disorder (ADHD) genes and risk for structural and functional neuroimaging
intermediate ADHD phenotypes.

Gene variant Brain area


Structural studies (Year)
Durston, S. (2005) SLC6A3-10/10 Reduced caudate volume
Monuteaux, M. (2008) DRD47R Reduced superior frontal and cerebellar cortex
Shaw, P. ( 2007) DRD47R Cortex (thickness).
Functional studies (Year)
Durston, S. (2008) SLC6A3–10R Increased striatal activity
Durston, S. (2008) SLC6A3–10R Decreased cerebellar activity
Fan, J. (2003) DRD4, MAOA Activation in anterior cingulated gyrus

rare SLC6A3 coding variant, Ala559Val has also Hetereogeneity in effect size was noted across
been identified in two male children with ADHD. studies [67]. It is of interest that even 1-year-old
This variant, which results in a normal SLC6A3 infants with this allele have been reported to show
protein and dopamine uptake, also enhances non- less sustained attention than those without it. Infants
vesicular SLC6A3-dependent DA release which is that also carried the seven repeat in addition to the
interestingly also blocked by effective ADHD homozygous form of 5HTTLPR had the shortest
medications [68]. attention span [71]. Promoter variants of DRD4 have
Gene variants in dopamine receptors (DRD1, also been explored in ADHD and a modest associ-
DRD2, DRD4, DRD5) have also been reported to ation was reported in the meta-analyses for the “T” of
be associated with ADHD. The most widely SNP 521 (rs1800955) [67].
studied has been DRD4, a gene that has a poly- Other dopaminergic genes conferring risk for
morphic region of 2–10 repeats. A positive associ- ADHD include a 148 bp allele of a microsatellite
ation of the seven-repeat, a variant considered to be repeat polymorphism on DRD5 while the 136bp
less responsive to dopamine, has been reported in allele confers a protective effect. DRD1, DRD2 and
several meta-analyses indicating moderate risk [66, DRD3 variants have not been shown to confer risk
67, 69, 70]. [66, 67].

171
Chapter 15: Attention-deficit hyperactivity disorder

Catecholaminergic pathways Cholinergic pathways


ADHD children homozygous for the the catechol- CHRNA4, the nicotinic acetylcholine receptor a subunit
O-methyltransferase (COMT) Val allele have been gene, is also considered important in ADHD; however,
reported to have reduced prefrontal cortex dopamine it has not been widely studied. A meta-analyses detected
neurotransmission [72], and ADHD adult subjects with a trend for association of the “T” allele in exon 2
the Val/Val genotype have slower reaction times [73]. (rs2273506) and in intron 2 (rs6090384) [67]. The
However this polymorphism was not found to modulate pedigree disequilibrium test in a genetic isolate found
executive function [74, 75] and several meta-analyses the T allele of rs6090384 to be undertransmitted in
have provided no evidence for association between affected individuals [81]. Individual variations in
ADHD and the Val allele polymorphism [67, 76]. spatial attention have been reported to be associated
Several polymorphisms (1021C/T, 2*rrrG/A, with variants in CHRNA4 [82].
GT microsatellite, TaqI with an undigested band
of 464 base pair (A1) and two bands of 300 bp and Glutaminergic neurotransmission
164 bp (A2)) in dopamine beta-hydroxylase (DBH), the
Altered glutamatergic function in frontal and striatal
enzyme that catalyzes the conversion of dopamine to
regions in ADHD has been reported [83–85], and this
noradrenaline, have been associated with ADHD
appears to be normalized with medications used to treat
in some but not all studies [77]. A recent meta-
ADHD such as the stimulants and atomoxetine [86].
analyses showed a trend towards association only
SLC1A3 is located on 5p13 and encodes for a high-affinity
for TaqI [67] and a recent study found an impaired rate
glutamate transporter (GLAST; EAAT1). An association
of perceptual processing for rapidly presented visual
between several markers (rs 2269272 and haplotypes
events in ADHD subjects carrying the A2 allele [78].
rs2269272/rs3776581 and rs2269272/rs2032893) and
Meta-analyses results did not detect association for
ADHD was reported by Turic et al. [87] but not replicated
variants of the norepinephrine transporter (NET) gene
for the same markers and haplotypes in 257 nuclear
(SLC6A2) or alpha-2A adrenergic receptor gene
families by Laurin and colleagues [88]. Results from the
(ADRA2A) [67]. However, NET deserves further inves-
CHOP gene-wide study of SLC1A3 [89] found a nominal
tigation given the results from a study using a dense-
association for one intron SNP (rs3776571) and two
mapping strategy that captured all genetic variations
downstream SNPs (rs1529461; rs 6863386) but did not
within the SLA6A2 that found an association between
replicate Turic’s findings. Reports for the glutamate
ADHD and two SNPs (rs3785143 and rs11568324)
receptor, ionotropic, N-methyl D-aspartate 2A (GRIN2A)
that had also been identified in the IMAGE study [79].
have also been mixed [77].

Serotonergic pathways SNAP-25 and other genes regulating


Results from candidate gene studies investigating gene
variants in serotonin receptors (HTR1B, HTR2A), and vesicular neurotransmission
in the serotonin transporter (5-HTT) in ADHD The SNAP-25 gene, on human chromosome 20p11.2,
cohorts have been inconsistent [77]. A population-based encodes for a presynaptic membrane protein (synap-
association study that investigated 19 genes encoding tosomal associated protein 25kDA), that in combin-
serotonin receptors, transporters, and enzymes involved ation with synaptobrevin and syntaxin, forms the
in serotonin synthesis and degradation reported an asso- SNARE (soluble N-ethylmalemide-sensitive factor
ciation with polymorphisms of DDC (involved in attachment protein receptors) complex, which is
serotonin synthesis), MAOB (serotonin degradation), responsible for binding and fusion of neurotransmit-
and 5-HTR2A [80]. A meta-analysis has detected a sig- ter vesicles to the plasma membrane [90]. Barr et al.
nificant association only for an HTR1B variant [67]. [91] first reported a haplotype defined by two bi-
Association studies for monoamine oxidase A allelic SNPs (T87610G, T87614C), MnlI (rs3746544)
(MAOA), which also codes for an enzyme involved and the DdeI polymorphism (rs1051312), associated
in metabolism of norepinephrine, dopamine, and with ADHD in the 3’UTR of SNAP25. A modest
serotonin, show mixed results [77]. Results from significant association has also been reported for
meta-analyses indicate a weak associaton for a VNTR s3746544 in a study that included pooled data [92]
in the promoter region [67]. as well as in meta-analyses [67].

172
Chapter 15: Attention-deficit hyperactivity disorder

Gene expression may also be playing a role. An a sizable constellation of genes, presumably with
example, not yet explored in ADHD is the cis-acting related functions. Given the high prevalence and her-
element, SP1, in the promoter region of SNAP-25; SP1 itability of ADHD, a corollary of the rare variant
has been reported to increase SNAP-25 expression while hypothesis would be that most risk variants arose as
inhibition of SP1 results in reduced expression [93]. relatively recent events, which would considerably
In summary, several gene variants in genes primar- lessen the ability of association studies to detect them.
ily involved in neurotransmission have been replicated CNVs in genes or sets of genes related through the
in several studies; however, their contribution to ADHD same biological or clinical concept may be more likely
risk is modest. This may be in part due to phenotypic to explain most forms of ADHD.
and genotypic heteregenity and epigenetic factors. In the first ADHD study investigating structural
variation in ADHD, 222 CNVs were identified in 335
Genome-wide association studies ADHD patients and their parents that were not detected
in 2026 unrelated healthy controls. The ADHD CNV-
Genome-wide analyses (GWA) allow investigation for
associated gene set was significantly enriched for genes
association between a disorder and a marker virtually
important for psychological and neurological functions,
anywhere in the genome. Results of several GWA stud-
including learning, behavior, synaptic transmission,
ies (GWAS) conducted on data from the International
and central nervous system development. It was also
Multicenter ADHD Genetics (IMAGE I) samples, based
significant enriched for genes reported in autism,
on ADHD diagnoses and related phenotypes, did not
schizophrenia, and Tourette syndrome [104]. Four
identify any markers with genome-wide significance
other subsequent studies also identified a number of
[94, 95]. One IMAGE I GWAS, that used quantitative
variants [105–107]. As expected, different variants were
ADHD phenotypes, identified two SNPs (rs552655 and
identified by the four different studies suggesting that
rs6565113) with nominal significance, and one of these
there could be hundreds or even thousands of variants
SNPs, rs6565113, is located in an intron of CDH13 that
disrupting genes in neuronal pathways that could
codes for cadherin 13, a member of a family of cell–cell
lead to similar phenotypic manifestations. The genes
adhesion proteins [96]. This same SNP was also
impacted by structural variants reported in one or more
reported to have nominal significance and was in the
studies are listed in Table 15.3.
top 25 SNPs reported by Neale and colleagues in the
same sample and also in a GWAS of an independent
sample of adults with ADHD [51, 97]. However, a meta- Gene–environment interactions
analysis that included samples from IMAGE I, IMAGE Heritability studies also suggest that epigenetic factors
II, The Children’s Hospital of Philadelphia, and the play a role in ADHD. As reviewed by Elia et al. [39],
Pfizer-funded study from the University of California animal studies of malnutrition, maternal stress, infec-
with a combined sample size of 2064 trios, 896 cases, tion, and toxic compounds have been shown to affect
and 2455 controls did not detect any genome-wide prenatal development of the brain circuitry relevant
significant associations [98]. to ADHD. Human studies have reported an associ-
ation between ADHD and low birth weight [108],
Structural variants maternal smoking [109], and psychosocial adversity
[110], but not prenatal alcohol exposure [109, 111].
In addition to single nucleotide variations, larger and
The dopamine transporter gene (DAT; SLC6A3) has
more complex variations have been recently identified in
also been reported to be affected by epigenetic factors
the human genome, referred to as copy number variants
that impact on chromatin structure, DNA methylation,
(CNVs) [99, 100]. These duplications and deletions can
and transcriptional regulation [112]. DAT mRNA
result in changes in gene dosage that may contribute to
expression has also been reported to decrease with age
genomic instability [101] and phenotypic variations
[113], in response to medication [114], environmental
[102, 103] in complex disorders such as ADHD.
factors [114], and pathogens [115].
The lack of success in identifying common vari-
ants contributing substantial risk for ADHD in the
European population also led to an alternative ADHD: a complex phenotype
hypothesis, namely that ADHD genetic risk is trans- ADHD is highly heteregenous. Symtoms such as
mitted largely by rare variants that collectively disrupt hyperactivity vary in individual patients over

173
Chapter 15: Attention-deficit hyperactivity disorder

Table 15.3 ADHD genes impacted by CNVS identified in two or more studies.

Elia et al. Williams et al. Lesch et al. Lionel et al. Elia et al. 2011
2009 2011 2011
Cases/ 335/2026 366/1047 UK 99/2026 248/2353 CHOP 1013/4105
Controls US 825/35243 Iceland Germany Canada Replic 2493/9222
CNV size 7.75–1775 kb > 500 kb 110 kb-3 Mb >20 kb CHOP 62.2 kb
Repl 53.6 kg
Genotyping Illumina Human aCGH Affy 6.0
Platform 550 K 660 W-Quad
Cases/controls Illumina 550 K

Chromosome Gene(s)
1p31.1 NEGR1 Dup
1p32.3 UPS24 Del
3p26.1 GRM7 Del
3p26.3 CHL1 dup dup
CNTN6 dup del
3q26.1 ZZBX Dup Del Del
Serpini 12 Dup Del
WDR49
PDC10
3q11.2 EPHA6 Dup del
4p15.2 DHX15 Dup dup
4q22.1 PPM1K Del dup
5p15.2 CTNND2 Del Dup
5q12.3 SGTB/NLN Del
5q13.3 SV2C Dup Dup
IQGAP2 Dup
5q35.2 CPLX2 Del dup
6q22.33 PTPRK Dup del
6q24.3 GRM1 Dup
7p22.2 SDK1 Del/dup del
7q31.33 GRM8 Del
7q32.3 CHCHD3 Dup Dup
7q35 CNTNAP2 Del dup
7q36.2 DPP6 dup
7q.11.22 AUTS2 Del/dup Dup
7q.32 CHCHD3 dup dup
8p23.2 CSMD1 Del del
8p22 SGCZ Dup del
8q21.3 CNBD1 Del del

174
Chapter 15: Attention-deficit hyperactivity disorder

Table 15.3 (cont.)

Elia et al. Williams et al. Lesch et al. Lionel et al. Elia et al. 2011
2009 2011 2011
11q13.4 DNAJB13 dup dup
CHCHD8
MRPL48
PAAF1
UCP2
11q14.3 GRM5 Del
15q13.3 FMN1 Dup dup
16p11.2 40 genes Del/dup Dup
16q23.3 CHD13 Del del
16q24.1 ATP2C2 Del del
20p12.2 PAK7 Dup dup
20p12.1 MACROD2 del Del/dup del
Del-deletion; Dup-duplication

time [116]. With the exception of the inattentive sub- activation [133], and decreased cerebellar activity [133].
type which has strong familial clustering in an isol- The 10 repeat (10R) homozygous variant of SLC6A3, a
ated Dutch population [117], family studies reported gene expressed primarily in the striatum [134], was
a lack of similarity for ADHD subtypes in first-degree shown to be associated with decreased caudate volume
relatives of probands with ADHD [17, 118, 119]. in subjects with ADHD, their unaffected siblings, and
Comorbid conditions occurring in approximately controls [42]. The 10R allele, which is associated with
33% add to the heterogeneity of ADHD [120, 121]. decreased dopamine transporter activity [135], may also
These include oppositional defiant disorder (35%), be related to reduced gene expression.
conduct disorder (30–50%), anxiety disorders (25%), Reduced prefrontal gray matter volumes have
mood disorders (15–75%), and learning disabilities been reported for DRD4-R4/4; however there was no
(25%) [122–127]. Epidemiological samples of chil- genotype effect detected in ADHD children versus
dren and adolescents also report significant comor- controls [42], which was not surprising given that this
bidity suggesting that this is not due to referral bias is not the variant that has been associated with
that might be expected in clinical samples [128–130]. ADHD. In contrast, the DRD4-7R variant, associated
It is not known whether these associated conditions with ADHD, has been associated with decreased cor-
modify the ADHD phenotype or whether single genes tical thickness in ADHD children [136] and ADHD
influence multiple phenotypic traits, as suggested in adults [137], where it has also been associated with
a study by Jain et al. [131] where ADHD was found decreased volume in the cerebellar cortex [137].
to co-segregate with oppositional defiant disorder The differences in brain structure in ADHD do not
(ODD) and conduct disorder (CD). appear to be static. As reported by Shaw et al. [136],
To address this phenotypic heterogeneity, neurocog- delays in prefrontal cortical maturation of  3 years
nitive tests [82] and statistical methods used to identify were detected in ADHD children compared to con-
distinct homogenous ADHD subgroups [132] are being trols. ADHD subjects with the DRD47R genotype had
used to further define ADHD phenotypes. Structural and the thinnest cortex and also had better clinical out-
functional neuroimaging are additional measures util- comes suggesting that this neuroanatomical correlate
ized as intermediate phenotypes for the investigation of of DRD4 genotype resolved by late adolescence [136].
candidate genes conferring risk for ADHD (Table 15.2). Recent neuroimaging genetic studies that also
Thus far, SLC6A3 variants have been associated with included unaffected siblings of ADHD children
reduced caudate volume [42], increased striatal brain are allowing the assessment of familial risk for

175
Chapter 15: Attention-deficit hyperactivity disorder

candidate genes. The SLC6A3 10R allele was found to normal adult volunteers with homozygous (COMT)
be associated with increased striatal activation in Val/Val genotype, improvement with amphetamine
ADHD children and their nonaffected siblings but treatment was demonstrated, whereas the homozygous
not in controls, indicating that this variant alone does Met/Met subjects performed more poorly after medica-
not result in clinical symptoms [133]. In this same tion in working memory tasks [143].
study, decreased activation of the cerebellar vermis, Thus pharmacogenetic studies, considered as com-
which was found in ADHD subjects, siblings, and plementary approaches in the search for biologically
healthy controls with the SLC6A3 10R homozygous relevant disease genes, have not yet been helpful in the
genotype, suggested a lack of familial risk [133]. How- quest for the identification of causal genes. This may be
ever, in a subsequent study that did not investigate due in part to the fact that our pharmacotherapeutic
genotype, and utilized the same basic go/no-go para- probes are nonspecific, as shown in one study indicat-
digm (but had the stimuli occurring at both expected ing changes in gene expression in over 700 genes in the
and unexpected time intervals), unaffected siblings striatum of rats treated with methylphenidate [144]. In
showed similar anomalous activation to that of addition, as Levy [145] recently hypothesized, genes
ADHD affected subjects when processing a stimulus conferring risks for ADHD may be different from
presented at unexpected intervals [138]. genes involved in medication response.
Brain imaging studies in twins are being used Genetic variants that influence the pharmacoki-
to decipher changes that could be attributed to genetic netics of ADHD medications may be more clinically
or environmental factors. A structural magnetic useful. A gene variant for caboxylesterase 1, the enzyme
resonance imaging (MRI) study in monozygotic twins used to esterify methylphenidate to d, l-ritalinic acid
concordant or discordant for ADHD symptoms and l-ethylphenidate has been reported in one subject
showed reduced function of attentional networks identified as a poor methylphenidate metabolizer [146].
of cerebellar, occipital/parietal and temporal brain Amphetamine compounds are metabolized through
regions interacting with the prefrontal cortex. In the hepatic CYP450 system, primarily through CYP3A4
ADHD of genetic origin (defined as brain differences and to a lesser extent through CYP2D6 [147], while
between concordant high risk twin pairs for ADHD atomoxetine is metabolized primarily through CYP2D6.
and a low risk twin comparison group), the medial Poor metabolizers of atomoxetine were shown to
orbitofrontal subdivisions were abnormal; while in have greater symptom improvement [148].
ADHD of environmental origin (brain differences
between the low risk and high risk twins from dis- ADHD animal models
cordant MZ twin pairs), the right inferior dorsolateral
ADHD animal models have been developed through neu-
prefrontal cortex was abnormal [139]. Functional
rotoxic brain lesions [149, 150], neonatal anoxia [151,
MRI studies indicated decreased activation of the left
152], toxin exposure such as lead [153], polychlorinated
dorsolateral prefrontal cortex and right parietal lobe
biphenyls [154], and X-irradiation [155]. More recently, as
for ADHD of genetic origin, and decreased activation
reported in several excellent reviews [156–159] and sum-
in left and right temporal lobe areas in ADHD of
marized in Figure 15.2, the development of knockout
environmental origin [140].
(KO), knockdown (KD), selected inbred strains, trans-
genic strains, and ENU mutants [160] are providing
Pharmacogenetics insight into the complex genetic components of ADHD.
As reviewed by Froehlich et al. [141], pharmacogenetic
studies have for the most part focused on ADHD can- KO animal models
didate gene variants involved in neurotransmission and KO animals, allowing the exploration of single-
methylphenidate reponse with mixed results. A GWAS ADHD candidate genes, indicate that a number of
investigating response to a methylphenidate transder- different genes, in the same as well as in different
mal system reported nominal evidence for SLC6A2 neurotransmitter pathways, may be responsible for
SNPs (rs17841329 and rs192303) as well as for GRM7, similar behaviors. As reviewed by Viggiano and col-
a metabotropic glutamate receptor gene (rs3792452) leagues [59], in the dopaminergic pathway, the homo-
[142]. Amphetamine compounds have not been as thor- zygous SLC6A3 (DAT1) KO mouse, which clears
oughly studied in ADHD subjects. In one study in synaptic dopamine at a slower rate than the hemizygous

176
Chapter 15: Attention-deficit hyperactivity disorder

MAO-B MAO-A COMT


NET
DAT1 α2c
DRD1 α2a α1b
K-O DRD2 TH
DRD3
H1R DRD4 5H1B DBH
DA
TH
DARPP-32

K-D DAT1

Hemizygous Deletions SNAP-25

SHR
WK/HA WKY Hyposexual
Selected In Bred
Strains Rat NHE
NLE C57BL/6

α2c++ TRβ1 Caly


Transgenic α1b++ TRα1R348C

ENU Models Mutations can occur anywhere in the genome and can affect different alleles

Figure 15.2 Attention-deficit hyperactivity disorder (ADHD): genetic animal models.

and the wild type (WT), is also more hyperactive than spatial learning and long-term potentiation (LTP) are
these two strains [161] as is the DRD3 KO model [162]. normal but perform less well on behavioral tasks and
Decreased locomotor activity occurs in the DA [163], have lower activity levels which are restored by
tyrosine hydroxylase (TH) [164], D2 [165], DARPP-32 increasing NE activity [177]. Interestingly, DBH KO
[166], and DRD4 KO [167]. D5-KO had normal activity mice which lack NE have normal locomotor activity.
level [168] while results for the D1-KO animal show Activity level changes are not limited to the dopa-
both normal and elevated locomotor activity [169, 170]. minergic and noradrenergic systems, as hyperactivity
Double KO strains result in increased levels for D1 + D3 has also been reported in mice lacking the 5-HT1B
[171] and decreased levels for D2 + D3 [59]. receptor [178], while reduced locomotor activity has
A review of the noradrenergic pathway by Vig- been noted in both histamine-1-receptor (H1R) -KO
giano et al. [60] indicates that locomotor activity is and WT mice [179].
decreased in SLA6A2 [172] and a2a [173] KO animals Toxic lesions in KO animals are also elucidating
and in strains where a2c [174] and a1b [175] are signaling necessary for locomotor activity changes.
over-expressed. Locomotor activity is increased in An example are rat pups injected with intracisternal
KO a1b [176] and a2c [174] models. TH is necessary 6-hydroxydopamine (6-OHDA), one of the earliest
for the synthesis of both dopamine (DA) and nor- animal models for ADHD [149]. Compared to control
epinephrine (NE) [163, 164], since 3,4-dihydroxyphe- littermates, they presented with hyperactivity and
nilalanine (l-DOPA) is converted to DA (by TH) deficits in brain dopamine that was reversed by
which is then converted to NE (by DBH). Pure DA methylphenidate and d-amphetamine [180]. 6-OHDA
KO mice can be generated by normalizing expression injection in DRD4 KO mice, however, prevented the
of TH and these mice are severely hypoactive [163]. development of hyperactivity, showing that D4 recep-
Animals with a TH mutation resulting in decreased tor signaling is essential for hyperactivity [181].

177
Chapter 15: Attention-deficit hyperactivity disorder

KD animal models may be compensatory to the excessive dopamine


present during early development [196].
DAT-KD mice have DAT expression reduced to 10%
Naples high-excitability (NHE) rats have high activ-
compared to WT animals, most likely due to reduced
ity while the low-excitability strain (NLE) has reduced
clearance [182]; these animals provide a model of a
locomotor activity; both strains have deficits of atten-
chronic hyper-dopaminergic state. These animals are
tion [191, 197]. In the NHE animals, SLC6A3 and tyro-
hyperactive and have alterations in habituation. Elec-
sine hydroxylase expression are increased in prefrontal
trophysiological recordings from medium-sized spiny
cortex while D1 receptor expression is decreased [197];
neurons in the dorsal striatum of KD mice showed
they also have an imbalance between NMDA and non-
alterations in both amplitude and frequency of spon-
NMDA sensitive [L-3H] glutamate receptors [198].
taneous glutamate-receptor-mediated synaptic cur-
The C57BL/6 is another inbred strain showing
rents [183]. Since D2 receptors normally dampen
hyperactivity [199] and has low expression of D2-R
glutamate release signaling [184, 185], it is hypothe-
autoreceptors in the ventral tegmental area (VTA) [200].
sized that in a chronic hyperdopaminergic state there
may be alterations in presynaptic D2 receptor function.
Transgenic animals
Transgenic male mice with a human mutant thyroid
Hemizygous deletions receptor (TRb1) are hyperactive, inattentive, and
The Coloboma mutant mouse, characterized by impulsive, and with the exception of a brief period
hyperactivity, head bobbing, eye dysmorphology, during postnatal development they are euthyroid [201].
and neurobehavioral delays [186] has a deletion in Rats with a knock-in TRa1R384C, a mutation of
chromosome 2 that includes the gene for synaptoso- the TRa1 gene that lowers affinity to thyroid hormone,
mal associated protein 25 (SNAP-25). These animals were noted to have extreme anxiety, reduced memory
respond to amphetamine but not to methylphenidate and locomotor activity [202]. A mouse engineered for
and this effect has been shown to be mediated the over-expression of the Calcyon gene, a vesicular
through D2 dopamine receptors [187]. Hyperactivity protein involved in endocytosis which is important in
in this mouse model is also mediated through a2C- the recycling of neurotransmitters, is hyperactive with
adrenergic receptors [188] suggesting that both nor- reduced anxiety [203].
adrenergic and dopaminergic systems may be involved.
ENU mutants
Selected inbred strains Novel mutant animals, produced with chemicals such
Inbred strains have included the spontaneously hyper- as N-ethyl-N-nitrosurea (ENU), are expected to be
tensive rats (SHR) [189], the hyposexual rat [190], and better models of the human disease. These mutants
the Naples high- and low-excitability (NHE and NLE) are not limited to loss of single gene function since
rats [191]. mutations can occur anywhere in the genome and can
One of the most frequently studied inbred models affect different alleles [160].
is the SHR, one of the few animal models known to
display several ADHD features including hyperactiv- Gene–environmental interactions:
ity, impulsivity, and difficulties with attention [158].
By selective inbreeding of this strain with a progeni- animal models
tor, the Wistar Kyoto strain (WKY), a hyperactive but Animal models will also be invaluable in deciphering
not hypertensive model was developed [192]. Sequen- gene and environmental interactions. Studies with
cing studies of candidate dopaminergic genes DRD2, the NHE rat have shown that increased maternal
DRD4, and SLC6A3 in the SHR strain compared to care and high fat diet (induced by small litter size)
the WKY strain (nonhyperactive) showed variations resulted in decreased activity and longer scanning times
only in the SLC6A3 gene and not in DRD2 or DRD4 [204]. Prepubertal handling of NHE animals also
[193], as well as reduced SLC6A3 gene expression resulted in decreased expression of excitatory amino
in the first postnatal month and increased SLC6A3 acids (L-Glu, L-Asp, and D-asp) and decreased activity
gene expression in adulthood [194, 195]. Mesocortical levels compared to nonstimulated animals [205].
SLC6A3 expression, increased in SHR strains [193, 196] Increased stress in the C57BL/6 strain has been shown

178
Chapter 15: Attention-deficit hyperactivity disorder

to result in comorbid depressive behaviors [200]. A pilot Interestingly, in SHR pups, performance in a training
study that examined brain RNA from rats exposed to task involving cognition and attention was signifi-
polychlorinated biphenyl (PCB) compared to SHR and cantly enhanced with methylphenidate and also with
control rats, found different sets of genes expressed H3 receptor antagonists [216].
[206]. A rat model exposed to nitrogen during the first
seven days of postnatal life showed increased anxiety.
While MRI and histological analysis did not detect any Summary
brain damage, expression studies detected decreased Candidate genes, linkage, and GWAS have identified
expression of group-I metabotropic glutamate receptors several gene variants involved in neurotransmission
[207]. In contrast, mice lacking the 5-HT1B receptor are which confer a modest risk for ADHD. In part, this
reported to be hyperactive throughout the lifespan but may be due to phenotypic heterogeneity as well as
show reduced anxiety [178]. environmental factors, gene–environment inter-
actions, or multiple variants within genes conferring
risk for ADHD. It is also possible that few common
Pharmacogenetics of animal models variants conferring risk for ADHD exist in the European
Animal models such as the DAT1 KO mouse, population and that very large samples will be necessary
remains responsive to methylphenidate in spite of to identify them.
the lack of a dopamine transporter [208]. SLC6A3 Structural variant studies are indicating that ADHD
KO animals also respond to fluoxetene [208]. Hyper- genetic risk is likely to be transmitted largely by rare
activity in these mice can be increased by NMDA- variants that collectively disrupt a sizable constellation
receptor blockers and suppressed by drugs that of genes, presumably with related functions. This is
increase glutamatergic transmission [209]. supported by animal genetic models which also impli-
Methylphenidate and d-amphetamine decreased cate numerous genes involved in complex interactions
symptoms in the SHR compared to WKY controls between neural pathways. Medication effects also sup-
[210, 211]. In the SHR model, reduced a2 adrenocep- port this, given that over 700 genes involved in the
tor-mediated inhibition of NE release mediates hyper- formation, maturation, and stability of neural connec-
activity [212], pointing to a noradrenergic focus that tions were reported to have increased expression in the
has been bolstered by evidence showing the efficacy of striatum of rats treated with methylphenidate [144].
the selective NE transporter (NET) inhibitor, atomox- Epigenetic factors, environmental factors, and
etine, in human ADHD studies [213, 214]. However, gene regulatory elements most likely also play a role
the dysfunction may be resulting from a defect in in ADHD genetics, and these are just beginning to be
glutamate-stimulated release of dopamine [215]. explored.

References 7. Barkley R. Attention-Defict/


Hyperactivity Disorder.
13. Biederman J, et al. Arch
Gen Psychiatry 1992;49(9):728–738.
1. Scahill L, et al. Child Adolesc A Clinical Workbook.
Psychiatr Clin N Am 2000;9 14. Lombroso PJ, et al. J Am Acad
New York: Guilford Press; Child Adolesc Psychiatry 1994;
(3):541–555. 1998. 33(7):921–938.
2. Heiligenstein E, et al.
8. Kessler RC, et al. J Occup Environ 15. Faraone SV, et al. Am J Psychiatry
J Am Coll Health 1998;46(4):
Med 2005;47(6):565–572. 1991;148(1):112–117.
185–188.
3. Kooij J, et al. Psychol Med 2005; 9. Weiss G, et al. Hyperactive 16. Faraone SV, et al. J Abnorm
35(6):817–827. Children Grown. New York: Psychol 1995;104(2):334–345.
Guilford Press; 1993.
4. Polanczyk G, et al. Curr Opin 17. Faraone SV, et al. Biol Psychiatry
Psychiatry 2007;20(4):386–392. 10. Pomerleau OF, et al. J Subst Abuse 1994;35(6):398–402.
1995;7(3):373–378.
5. Barkley R, et al. J Am Acad Child 18. Sprich S, et al. J Am Acad Child
Adolesc Psychiatry 2006;45(2): 11. Biederman J, et al. Biol Psychiatry Adolesc Psychiatry 2000;39(11):
192–202. 1998;44(4):269–273. 1432–1437.
6. Strine TW, et al. Prev Chronic Dis 12. Barkley RA, et al. Pediatrics 19. Biederman J, et al. Lancet 2005;
2006;3(2):A52. 1996;98(6 Pt 1):1089–1095. 366(9481):237–248.

179
Chapter 15: Attention-deficit hyperactivity disorder

20. Kuntsi J, et al. Biol Psychiatry 42. Fisher SE, et al. Am J Hum Genet 65. Vandenbergh DJ, et al. Genomics
2005;57(6):647–654. 2002;70(5):1183–1196. 1992;14(4):1104–1106.
21. Price TS, et al. Behav Genet 2005; 43. Smalley S, et al. Am J Hum Genet 66. Barr CL, et al. Future Neurology
35(2):121–132. 2002;71(4):959–963. 2008;3(6):705–728.
22. Saudino KJ, et al. J Abnorm Child 44. Bakker SC, et al. Am J Hum Genet 67. Gizer IR, et al. Hum Genet
Psychol 2005;33(1):113–130. 2003;72(5):1251–1260. 2009;126(1):51–90.
23. Hay DA, et al. Biol Psychiatry 45. Ogdie M, et al. Am J Hum Genet 68. Mazei-Robison MS, et al.
2007;61(5):700–705. 2003;73(5):493. J Neurosci 2008;28:7040–7046.
24. Polderman TJ, et al. J Child Psychol 46. Hebebrand J, et al. Mol Psychiatry 69. Faraone SV, et al. Am J Psychiatry
Psychiatry 2007;48(11):1080–1087. 2006;11(2):196–205. 2001;158(7):1052–1057.
25. Derks EM, et al. Behav Genet 47. Asherson P, et al. Mol Psychiatry 70. Li D, et al. Hum Mol Genet
2008;38(1):11–23. 2008;13(5):514–521. 2006;15(14):2276–2284.
26. Ouellet-Morin I, et al. Am J Med 48. Zhou K, et al. Biol Psychiatry
Genet B Neuropsychiatr Genet 71. Auerbach JG, et al. Psychiatr
2008;64(7):571–576. Genet 2001;11(1):31–35.
2008;147B(8):1442–1449.
49. Arcos-Burgos M, et al. Am J Hum 72. Lachman HM, et al.
27. Wood AC, et al. Behav Genet
Genet 2004;75(6):998–1014. Pharmacogenetics 1996;6(3):
2008;38(3):266–276.
50. Amin N, et al. Eur J Hum Genet 243–250.
28. Ehringer MA, et al. J Abnorm
2009;17(7):958–966. 73. Boonstra AM, et al. Am J Med
Child Psychol 2006;34(1):1–17.
51. Lesch KP, et al. J Neural Transm Genet B Neuropsychiatr Genet
29. Schultz MR, et al. Twin Res Hum 2008;147(3):397–402.
2008;115(11):1573–1585.
Genet 2006;9(2):220–232.
52. Romanos M, et al. Mol Psychiatry 74. Mills S, et al. BMC Psychiatry
30. van den Berg SM, et al. Am J Med 2004;4:15.
Genet B Neuropsychiatr Genet 2008;13(5):522–530.
2006;141B(1):55–60. 53. Vegt R, et al. Eur J Hum Genet 75. Taerk E, et al. BMC Med Genet
2010;18(2):206–211. 2004;5:30.
31. Haberstick BC, et al. Psychol Med
2008;38(7):1057–1066. 54. Ogdie MN, et al. Mol Psychiatry 76. Cheuk DK, et al. Behav Genet
2006;11(1):5–8. 2006;36(5):651–659.
32. Young SE, et al. J Abnorm Psychol
2009;118(1):117–130. 55. Zhou K, et al. Am J Med 77. Faraone SV, et al. Psychiatr
Genet B Neuropsychiatr Genet Clin North Am 2010;33(1):
33. Saviouk V, et al. Am J Med Genet 159–180.
2008;147B(8):1392–1398.
B Neuropsychiatr Genet
2011;156B(3):352–362. 56. Weiss KB, et al. J Allergy Clin 78. Bellgrove MA, et al. Biol
Immunol 2000;106(3):493–499. Psychiatry 2006;60
34. Simonoff E, et al. Psychol Med (10):1039–1045.
1998;28(4):825–837. 57. Mannino DM, et al. MMWR
Surveill Summ 2002;51(1):1–13. 79. Kim JW, et al. Am J Med Genet
35. Nadder TS, et al. J Child Psychol B Neuropsychiatr Genet 2007;
Psychiatry 2001;42(4):475–486. 58. Solanto MV. Behav Brain Res 144(6):781–790.
1998;94(1):127–152.
36. Sherman DK, et al. Am J 80. Ribases M, et al. Mol Psychiatry
Psychiatry 1997;154(4):532–535. 59. Viggiano D, et al. Neurosci 2009;14(1):71–85.
37. Hay D, et al. Aust J Psychology Biobehav Rev 2003;27(7):
623–637. 81. Wallis D, et al. Atten Defic
2004;56(2):99–107. Hyperact Disord 2009;1(1):
60. Viggiano D, et al. Neural Plast 19–24.
38. Larsson JO, et al. J Am Acad Child
2004;11(1–2):133–149.
Adolesc Psychiatry 2004; 82. Bellgrove MA, et al. Ann N Y Acad
43(10):1267–1275. 61. Volkow ND, et al. Neuroimage Sci 2008;1129:200–212.
2007;34(3):1182–1190.
39. Elia J, et al. Curr Top Behav 83. Jin Z, et al. Neurosci Lett 2001;
Neurosci 2011 [epub ahead of 62. Cook EH Jr., et al. Am J Hum 315(1–2):45–48.
print]. Genet 1995;56(4):993–998.
84. MacMaster FP, et al. Biol
40. Lenroot RK, et al. J Child Psychol 63. Hersch SM, et al. J Comp Neurol Psychiatry 2003;53(2):184–187.
Psychiatry 2011;52(4):429–441. 1997;388(2):211–227. 85. Courvoisie H, et al. J Neuro-
41. Roth TL, et al. J Child Psychol 64. Nirenberg MJ, et al. J Neurosci psychiatry Clin Neurosci 2004;
Psychiatry 2011;52(4):398–408. 1997;17(18):6899–6907. 16(1):63–69.

180
Chapter 15: Attention-deficit hyperactivity disorder

86. Carrey N, et al. Clin 108. Hultman CM, et al. J Am Acad 129. Angold A, et al. J Child Psychol
Neuropharmacol 2003;26(4): Child Adolesc Psychiatry 2007; Psychiatry 1999;40(1):57–87.
218–221. 46(3):370–377. 130. Costello EJ, et al. J Am Acad Child
87. Turic D, et al. Biol Psychiatry 109. Knopik VS, et al. Psychol Med Adolesc Psychiatry 2006;45(1):8–25.
2005;57(11):1461–1466. 2006;36(10):1461–1471. 131. Jain M, et al. Biol Psychiatry
88. Laurin N, et al. Biol Psychiatry 110. Laucht M, et al. Arch 2007;61(12):1329–1339.
2006;59(11S):115S. Gen Psychiatry 2007;64 132. Elia J, et al. Psychiatry Res
89. Elia J, et al. Psychiatr Genet (5):585–590. 2009;170(2–3):192–198.
2009;19(3):134–141. 111. Neuman RJ, et al. Biol Psychiatry 133. Durston S, et al. J Am Acad
90. Sorensen JB. Trends Neurosci 2007;61(12):1320–1328. Child Adolesc Psychiatry
2005;28(9):453–455. 112. Shumay E, et al. PLoS One 2010; 2008;47:61–67.
91. Barr CL, et al. Mol Psychiatry 5(6):e11067. 134. Heinz A, et al. Neuro-
2000;5(4):405–409. 113. Bannon MJ, et al. Neurology psychopharmacology 2000;22:
92. Kim JW, et al. Mol Psychiatry 1997;48(4):969–977. 133–139.
2008;13(6):624–630. 114. Volkow ND, et al. Biol Psychiatry 135. Jacobsen LK, et al. Am J Psychiatry
93. Cai F, et al. J Neurochem 2008; 2005;57(11):1410–1415. 2000;157:1700–1703.
105(2):512–523. 115. Wang GJ, et al. Brain 2004; 136. Shaw P, et al. Proc Natl Acad Sci
127(Pt 11):2452–2458. U S A 2007;104:19649–19654.
94. Franke B, et al. Hum Genet
2009;126(1):13–50. 137. Monuteaux MC, et al. J Med Genet
116. Lahey BB, et al. Arch Gen
B Neuropsychiatr Genet 2008;
95. Neale BM, et al. J Am Acad Child Psychiatry 2005;62(8):896–902.
147B:1436–1441.
Adolesc Psychiatry 2010;49(9): 117. Croes EA, et al. Eur J Epidemiol
906–920. 138. Mulder MJ, et al. J Am Acad
2005;20(9):789–794.
Child Adolesc Psychiatry 2008;
96. Patel SD, et al. Curr Opin Struct 118. Faraone SV, et al. Am J Psychiatry 47(1):68–75.
Biol 2003;13(6):690–698. 2000;157(7):1077–1083.
139. Van ‘t Ent D, et al. Neuroimage
97. Neale BM, et al. Am J Med Genet B 119. Smalley SL, et al. J Am Acad Child 2007;35(3):1004–1020.
Neuropsychiatr Genet 2008; Adolesc Psychiatry 2000;39(9):
147B(8):1337–1344. 140. Van ‘t Ent D, et al. Neuroscience
1135–1143.
2009;164:16–29.
98. Neale BM, et al. J Am Acad Child 120. Jensen PS, et al. J Am Acad Child
Adolesc Psychiatry 2010;49: 141. Froehlich TE, et al. CNS Drugs
Adolesc Psychiatry 2001;40(2): 2010;24(2):99–117.
884–897. 147–158.
142. Mick E, et al. Am J Med Genet
99. Eichler EE. Nat Genet 2006; 121. Elia J, et al. Child Adolesc B Neuropsychiatr Genet
38(1):9–11. Psychiatry Ment Health 2008; 2008;147B(8):1412–1418.
100. Sharp AJ, et al. Nat Genet 2006; 2(1):15.
143. Mattay VS, et al. Proc Natl
38(9):1038–1042. 122. Biederman J, et al. Am J Psychiatry Acad Sci U S A 2003;100(10):
101. Emanuel BS, et al. Nat Rev Genet 1991;148(5):564–577. 6186–6191.
2001;2(10):791–800. 123. Hinshaw SP. J Consult Clin 144. Adriani W, et al. Ann N Y Acad
102. Iafrate AJ, et al. Nat Genet 2004; Psychol 1992;60(6):893–903. Sci 2006;1074:52–73.
36(9):949–951. 124. Cantwell DP. J Am Acad Child 145. Levy F. Aust N Z J Psychiatry
103. Sebat J, et al. Science 2004; Adolesc Psychiatry 1996; 2007;41(1):10–16.
305(5683):525–528. 35(8):978–987.
146. Patrick KS, et al. Clin Pharmacol
104. Elia J, et al. Mol Psychiatry 125. Jensen PS, et al. J Am Acad Child Ther 2007;81(3):346–353.
2010;15(6):637–646. Adolesc Psychiatry 1997;36(8): 147. Markowitz JS, et al. Clin
105. Williams NM, et al. 1065–1079. Pharmacokinet 2001;40(10):
Lancet 2010;376 126. Brown RT, et al. Pediatrics 753–772.
(9750):1401–1408. 2001;107(3):E43. 148. Michelson D, et al. J Am Acad
106. Lesch KP, et al. Mol Psychiatry 127. Spencer TJ. J Clin Psychiatry Child Adolesc Psychiatry 2007;
2011;16(5):491–503. 2006;67(Suppl 8):27–31. 46(2):242–251.
107. Lionel AC, et al. Sci Transl Med 128. Caron C, et al. J Child Psychol 149. Shaywitz BA, et al. Science
2011;3(95):75–95. Psychiatry 1991;32(7):1063–1080. 1976;191(4224):305–308.

181
Chapter 15: Attention-deficit hyperactivity disorder

150. Cardinal RN, et al. Science 173. Lahdesmaki J, et al. Neuroscience 195. Leo D, et al. Neurosci Biobehav
2001;292(5526):2499–2501. 2002;113(2):289–299. Rev 2003;27(7):661–669.
151. Dell’Anna ME, et al. Behav Brain 174. Sallinen J, et al. Mol Psychiatry 196. Viggiano D, et al. Behav Brain Res
Res 1991;45(2):125–134. 1999;4(5):443–452. 2002;130(1–2):181–189.
152. Dell’Anna ME, et al. Brain Res 175. Zuscik MJ, et al. Nat Med 2000; 197. Viggiano D, et al. Behav Genet
Bull 1993;32(2):159–170. 6(12):1388–1394. 2002;32(5):315–333.
153. Silbergeld EK, et al. Exp Neurol 176. Drouin C, et al. J Neurosci 2002; 198. Sadile AG, et al. Behav Brain Res
1974;42(1):146–157. 22(7):2873–2884. 1996;78(2):163–174.
154. Holene E, et al. Behav Brain Res 177. Kobayashi K, et al. J Neurosci 199. Cabib S, et al. Physiol Behav
1998;94(1):213–224. 2000;20(6):2418–2426. 1990;47(4):749–753.
155. Diaz-Granados JL, et al. Behav 178. Brunner D, et al. Behav Neurosci 200. Cabib S, et al. Behav Brain Res
Neural Biol 1994;61(3):251–259. 1999;113(3):587–601. 2002;130(1–2):103–109.
156. Davids E, et al. Brain Res Brain 179. Zlomuzica A, et al. Eur J Neurosci 201. Siesser WB, et al. Genes Brain
Res Rev 2003;42(1):1–21. 2008;27(6):1461–1474. Behav 2006;5(3):282–297.
157. Russell VA, et al. Behav Brain 180. Luthman J, et al. Psycho- 202. Venero C, et al. Genes Dev
Funct 2005;1:9. pharmacology (Berl) 1989; 2005;19(18):2152–2163.
158. Sagvolden T, et al. Biol Psychiatry 99(4):550–557. 203. Trantham-Davidson H, et al.
2005;57(11):1239–1247. 181. Avale ME, et al. Mol Psychiatry Behav Brain Res 2008;189
159. Mill J. J Neurosci Methods 2004;9(7):718–726. (2):244–249.
2007;166(2):294–305. 182. Zhuang X, et al. Proc Natl 204. Fresiello A, et al. Behav Brain Res
160. Godinho SI, et al. Eur J Hum Acad Sci U S A 2001; 2002;130(1–2):111–115.
Genet 2006;14(6):651–659. 98(4):1982–1987. 205. Ruocco LA, et al. Behav Brain Res
161. Giros B, et al. Nature 1996; 183. Wu N, et al. J Neurophysiol 2009;198(1):29–36.
379(6566):606–612. 2007;98(1):423–432. 206. DasBanerjee T, et al. Am J
162. Accili D, et al. Proc Natl Acad Sci 184. Cepeda C, et al. J Neurophysiol Med Genet B Neuropsychiatr
U S A 1996;93(5):1945–1949. 2001;85(2):659–670. Genet 2008;147B(8):1554–1563.
163. Zhou QY, et al. Cell 1995; 185. Bamford NS, et al. J Neurosci 207. Casolini P, et al. J Neurochem
83(7):1197–1209. 2004;24(43):9541–9552. 2005;95(1):137–145.
164. Kobayashi K, et al. J Biol Chem 186. Heyser CJ, et al. Brain Res 208. Gainetdinov RR, et al. Science
1995;270(45):27235–27243. Dev Brain Res 1995;89(2): 1999;283(5400):397–401.
165. Vallone D, et al. Behav Brain Res 264–269. 209. Gainetdinov RR, et al. Proc Natl
2002;130(1–2):141–148. 187. Fan X, et al. Neurobiol Dis 2007; Acad Sci U S A 2001;98(20):
166. Nally RE, et al. Neuro- 26(1):201–211. 11047–11054.
psychopharmacology 2003; 188. Bruno KJ, et al. Neurobiol Dis 210. Wultz B, et al. Behav Neural Biol
28(12):2055–2063. 2006;23(3):679–688. 1990;53(1):88–102.
167. Rubinstein M, et al. Cell 1997; 189. Okamoto K, et al. Jpn Circ J 211. Sagvolden T, et al. Behav Neural
90(6):991–1001. 1963;27:282–293. Biol 1992;58(2):103–112.
168. Hollon TR, et al. J Neurosci 190. Kohlert JG, et al. Physiol Behav 212. Russell V, et al. Behav Brain Res
2002;22(24):10801–10810. 1993;53(6):1215–1218. 2000;117(1–2):69–74.
169. Drago J, et al. Proc Natl Acad Sci 191. Gonzalez-Lima F, et al. Neurosci 213. Michelson D, et al. Am J Psychiatry
U S A 1994;91(26):12564–12568. Biobehav Rev 2000;24(1): 2002;159(11):1896–1901.
170. Xu M, et al. Cell 1994;79(4): 157–160. 214. Spencer T, et al. J Clin Psychiatry
729–742. 192. Hendley ED, et al. Behav Neural 2002;63(12):1140–1147.
171. Wong JY, et al. Psycho- Biol 1986;45(1):1–16. 215. Russell VA. Neurosci
pharmacology (Berl) 2003; 193. Mill J, et al. Behav Brain Funct Biobehav Rev 2003;27(7):671–6,
167(2):167–173. 2005;1:24. et al.82.
172. Xu F, et al. Nat Neurosci 2000; 194. Watanabe Y, et al. J Nucl Med 216. Fox GB, et al. Behav Brain Res
3(5):465–471. 1997;38(3):470–474. 2002;131(1–2):151–161.

182
Autism and autism spectrum disorders
Chapter

16 Daniel H. Geschwind and Maricela Alarcón

Abstract neuropsychiatric or neurodevelopmental syndrome


(Table 16.1; Figure 16.1). The evidence for this comes
Autism is a heterogeneous and broadly defined neu- from three major areas: (1) the existence of known
rodevelopmental disorder. After a decade or so of chromosomal or genetic disorders in which a high
intense genetic investigations, diverse approaches, proportion of children have a diagnosis of an autism
from genetic linkage and whole genome association, spectrum disorder; (2) twin studies clearly show
to comparative genomic hybridization have begun to markedly increased risk for the second twin among
yield significant dividends. These genetic findings monozygotic twins or a concordance rate of > 70%
demonstrate more heterogeneity than initially antici- for monozygotic twins versus 10% for dizygotic twins;
pated, but have also begun to shed light on potential and (3) the increased risk to first-degree relatives of
neurobiological mechanisms of this vexing spectrum autistic probands; that is, the risk to a sibling of an
of conditions. autistic proband is at least 10–25-fold greater than
other members of the general population [8]. Unlike
Introduction schizophrenia, there does not seem to be a large
Like other psychiatric conditions, autism is a syn- distinction between the concordance rate among non-
drome that is defined by observed behavior and cog- twin siblings and dizogotic twins. Thus, although
nition, not etiology. As is described in the DSM-IV- there may be a role for shared in utero environment,
TR, autism is characterized at its core by deficits in it is not as significant a risk factor for autism spec-
social communication, language, and the presence of trum disorders (ASD) as it may be in schizophrenia
repetitive, restrictive behaviors, all with onset prior to [9]. All of these data clearly indicate that studies that
age three [1]. Although in the original description of focus on genetic etiology would prove fruitful in
11 boys with autism in 1943 [2], Kanner clearly high- advancing our understanding of autism.
lighted what he thought to be its biological etiology,
for many decades subsequently, autism was con-
sidered to be primarily caused by certain aspects of A brief history of genetic linkage studies
family environment, including a lack of parental Given the high genetic risk for autism [8], and a
warmth or “cold mothering”. This sentiment was paucity of multigenerational families, linkage studies
widely held in spite of the recognition that autism using a sibling pair design represented the first
had a significant association with many medical con- attempts to screen genome-wide for major autism risk
ditions having genetic or biochemical etiologies [3, 4]. loci (see Table 16.1, which provides a brief history of
Rather, many of the rare conditions identified with genetic studies in autism). Such studies began in
autism were thought to be relatively unique circum- earnest in the 1990s and have culminated in at least
stances, and it was not until the last three decades that a dozen linkage studies [10–21].
the true nature of its genetic etiology was properly The first linkage study was performed by a collab-
recognized [5–7]. orative group, the International Molecular Genetic
Subsequently it has become widely accepted that Study of Autism Consortium (IMGSAC) and pub-
autism has among the highest genetic liability of any lished in 1998. This study, based on a multinational

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

183
Chapter 16: Autism and autism spectrum disorders

Table 16.1 Selected historical highlights of genetic studies of autism.

Year Study Contribution


1943 Kanner Described 11 cases of individuals with autism
1977 Folstein and Rutter First twin study of autism
1989 Le Couteur et al. First published diagnostic instrument for autism. Currently
gold standard for research studies
1985/1992 Gillberg and Wahlsrom, 1985; Affiliation of autism with medical conditions and chromosomal
Gillberg and Coleman, 1992 disorders
1994 Bolton et al. First-degree relatives of individuals with autism have mild
cognitive and behavioral traits characteristic of disorder
1995 Bailey et al. Expanded twin study that considered broader phenotypes
of autism
1997 Cook et al. 15q maternally derived duplication causes autism
1997 Geschwind et al. Cure Autism Now established first collaborative gene bank for
the study of autism – Autism Genetic Resource Exchange (AGRE)
is a publicly available database: http://agre.autismspeaks.org
1998 International Molecular Genetic First genome-wide linkage scan for autism
Study of Autism Consortium
(IMGSAC)
1999 Risch et al. Early linkage study that interprets evidence in favor of the
existence of multiple autism genes, each with a small effect
2001 Wassink et al. Documented that up to 9% of cases with autism are due to
chromosomal abnormalities
2002 Alarcón et al. First genome-wide quantitative trait loci scan in autism
2005 Cantor et al. First replication of linkage finding in autism; confirmed 17q21
peak that was attributed to families with male-only affecteds
2006 Jacquemont et al. Microarray comparative genomic hybridization identifies
chromosomal rearrangements in significant proportion of
affected cases and thereby illustrates new methodology for
autism gene search
2007 AGP Consortium; Szatmari et al. Largest linkage study to date. International collaborative effort
of 50 centers from United States and Europe with 1168 families.
Typed for 10 000 markers
2007 Nishimura et al. First blood genomics study with independently validated
results in neural tissue
2007 Sebat et al. Evidence that copy number variants (CNVs) are associated
with autism
2008 Morrow et al. Used homozygosity mapping to detect inherited causal
mutations for autism. Provided further evidence for extreme
heterogeneity, as no loci were shared
2008 Kumar et al.; Weiss et al. 16p11.2 CNV associated with autism
2009 Wang et al. First genome-wide association study that identified significant
common variants

184
Chapter 16: Autism and autism spectrum disorders

Chr 1 Chr 2 Chr 3 Chr 4 Chr 5 Chr 6 Chr 7 Chr 8 Chr 9 Chr 10 Chr 11 Chr X
1.1 5.1 8.1
3.1 10.1 X.1
9.1
7.1

5.2 11.1
2.1
10.2
3.2 5.3
3.3 7.2 9.2 11.2 X.2
4.1
11.3
4.2 10.3
4.3 6.1 7.4 7.3
X.3
7.5
7.6 9.3 X.4
7.7
3.4 7.8 9.4
6.2 9.5
7.10 7.9 X.5
1.2 7.11 X.6
2.2 2.3
2.4
3.5
3.6 4.4 12.1

1.3 16.1 15.1


2.5 15.2
17.2 12.2
17.1 15.3 13.1
20.1 16.2
19.1 17.3
21.1 20.2 16.3 14.1
17.4
17.5
22.1 16.4
21.2 15.4
22.2

Chr 22 Chr 21 Chr 20 Chr 19 Chr 18 Chr 17 Chr 16 Chr 15 Chr 14 Chr 13 Chr 12 Chr Y

ID Feature Position Refs ID Feature Position Refs ID Feature Position Refs


1.1 Loss 1p36 3 7.4 RELN 7q22 117–121 15.3 Gain 15q11–15q13 4
1.2 Linkage 1q21–1q23 116 7.5 MET 7q31 70,123 15.4 Linkage 15q22–15q26 5
1.3 DISCI 1q42 122 7.6 Loss 7q31 5 16.1 TSC2 16p13 110
2.1 NRXNI 2p16 5,34 7.7 Linkage 7q32–7q34 15 16.2 Loss 16p11 4,20,35,44
2.2 Loss 2p24 4 7.8 CADPS2 7q31 5,45 16.3 Gain 16p11 20,35,44
2.3 Linkage 2p24–2q31 57,58,112 7.9 Linkage 7q34–7q36 14,52,68 16.4 Loss 16q21 5
2.4 SLC 25A12 2p24 124–126 7.10 CNTNAP2 7q35–7q36 37–40 17.1 Loss 17p12 5
2.5 Loss 2q37 4 7.11 EN2 7q36 129,130 17.2 Gain 17p12 107
3.1 OTXR 3p25 127,128 8.1 Gain 8p23 5 17.3 SLC6A4 17q11 131–134
3.2 Loss 3p14 4 9.1 Linkage 9p24 5 17.4 Linkage 17q11–17q21 51,54,135
3.3 Gain 3p14 4 9.2 Loss 9q12 5 17.5 ITGB3 17q21 136,137
3.4 Linkage 3q22 15 9.3 Linkage 9q33 5 19.1 Linkage 19p13 140
3.5 Linkage 3q25–3q27 138,139 9.4 Linkage 9q34 15 20.1 Loss 20p13 5
3.6 Loss 3q27–3q28 3 9.5 TSCI 9q34 110 20.2 Loss 20p13 5

4.1 Loss 4q21 3 10.1 Loss 10q14–10p15 4 21.1 Linkage 21q11 55


4.2 Loss 4q21–23 3 10.2 Gain 10q11–10q21 4 21.2 Loss 21q22 3
4.3 Linkage 4q22–4q25 15 10.3 PTEN 10q23 141 22.1 Loss 22q13 4
4.4 Loss 4q35 5 11.1 Linkage 11q12–11p13 5 22.2 SHANK3 22q13 21,22,142,143
5.1 Linkage 5p15 5 11.2 DHCR7 11q13 108 X.1 NLGN4X Xp22 28
5.2 Linkage 5p13–5q11 140 11.3 Linkage 11q13–11q14 15 X.2 NLGN3 Xq13 28

5.3 Linkage 5q12 5 12.1 CACNAIC 12p13 24 X.3 Linkage Xq21–Xq25 140
6.1 GRIK2 6q21 144–146 12.2 AVPRIA 12p14–12q15 147 X.4 Gain Xq24 3
6.2 AHII 6q23 106 13.1 Gain 13p14 5 X.5 FMRI Xq27 105,148

7.1 Loss 7p21 4 14.1 Linkage 14q23 149 X.6 MECP2 Xq28 109
7.2 Loss 7q11 3 15.1 UBE3A 15q11 102,103
7.3 Linkage 7q22–7q32 52,111–113 15.2 GABRB3 15q12 25,114,115

185
Chapter 16: Autism and autism spectrum disorders

collaborative effort, consisted of fewer than 100 fam- increase in sample size in the most recent whole
ilies, characteristic of genetic studies of neuropsychia- genome linkage study, performed by an international
tric disease at this time [10]. Despite its size, a few collaborative group of the Autism Genome Project
very interesting findings emerged from this study and (AGP), no region reached clear genome-wide signifi-
are discussed in more detail below. The second major cance [24].
linkage study came from the Paris Autism Research This is not to say that linkage has been entirely
International Sibpair Study (PARIS) group [13] and without success. In 1998, the IMGSAC published a
involved approximately 60 families, identifying no two-stage genome scan, in which they studied a total
loci even suggestive of a genome-wide significance of 99 families from multiple collaborating institutions
level [22]. The small study was based on the optimis- in Europe and the United States. They identified
tic assumption that there would be only a few major several regions reaching nominal significance and a
loci shared by most children with autism, those with broad region on distal chromosome 7q reaching an
“idiopathic autism”, who did not have another under- MLS of 3.55 in a smaller subset of families from the
lying genetic disorder. At the time, most cases were United Kingdom. This same region was identified in
thought to be idiopathic and not to be syndromic further investigations by this group [25] and a study
forms, which were considered rare and atypical, des- with an expanded sample revealed a locus on chromo-
pite evidence to the contrary [23]. These assumptions some 2q and another on 17q [15]. The original locus
were prevalent then and had they been true, a few on chromosome 7q was the second most significant,
hundred families would have been sufficient to detect reaching a LOD score of 3.2 in a region near that of
the genome-wide significant linkage and identify a the original report in a total of approximately 150
gene within a locus. sibling pairs (including the original sample). A third
Important in this regard, was one of the other region on chromosome 17 was also identified at a
early linkage studies, which involved over 100 families suggestive level with a maximum LOD score of 2.3
and was performed by a group from Stanford [14]. at about 50 cM.
These investigators found no loci at a genome-wide It was noted early on that a region on distal
suggestive significance level, which they interpreted as chromosome 7q reached nominal significance in
evidence of extreme heterogeneity and suggested the nearly every linkage study performed [26], and Bad-
interaction of multiple genes was necessary to cause ner and Gershon identified this as the only region
autism in idiopathic cases. The authors concluded meeting genome-wide significance for autism using
that “positional cloning of susceptibility loci by link- meta-analysis [27]. This region overlaps with the ori-
age analysis may be a formidable task and other ginal IMGSAC linkage region and was recently con-
approaches may be necessary”. These conclusions of firmed in an independent sample [21].
multiple interacting genes causing most autism cases Following the initial IMGSAC paper, the PARIS
and the need for alternative strategies to identify group [13]) performed a genome-wide linkage scan
them, contrasted with the prevailing wisdom, includ- at < 10 cM density in a small set of multiplex
ing that of this chapters’ authors; but they have turned families. Although none of the regions yielded even
out to be largely correct. Despite a nearly 10-fold suggestive linkage, several did overlap with other

Figure 16.1 Loci implicated in autism spectrum disorder (ASD) etiology. Entries in the ID column of the table map are entries to the
ideograms of individual chromosomes. Red and yellow bars correspond to de novo losses and gains, respectively, which are observed in cases
but not in controls. Green bars correspond to genes that are observed to modulate ASD risk (either through a rare syndrome or through
genetic association): light green and dark green bars represent promising or probable candidate genes, respectively, as defined in the
table map. Regions shaded in purple correspond to linkage peaks. Only human data were considered in the assembly of the table. AHI1,
Abelson helper integration site 1; AVPR1A, arginine vasopressin receptor 1A; CACNA1C, calcium channel voltage-dependent L type α 1C
subunit; CADPS2, Ca2+-dependent activator protein for secretion 2; CNTNAP2, contactin associated protein-like 2; DHCR7, 7-dehydrocholesterol
reductase; DISC1, disrupted in schizophrenia 1; EN2, engrailed homeobox 2; FMR1, fragile X mental retardation 1; GABRB3, g-aminobutyric
acid (GABA) A receptor β3; GRIK2, glutamate receptor ionotropic kainate 2; ITGB3, integrin β3; MECP2, methyl CpG binding protein 2; MET,
met proto-oncogene; NLGN3, neuroligin 3; NLGN4X, neuroligin 4 X-linked; NRXN1, neurexin 1; OXTR, oxytocin receptor; PTEN, phosphatase
and tensin homologue; RELN, reelin; SHANK3, SH3 and multiple ankyrin repeat domains protein 3; SLC25A12, solute carrier family 25
(mitochondrial carrier, Aralar) member 12; LC6A4, solute carrier family 6 (neurotransmitter transporter, serotonin) member 4; TSC1, tuberous
sclerosis 1; TSC2, tuberous sclerosis 2; UBE3A, ubiquitin protein ligase E3A. (From [8], with permission from Macmillan Publishing Ltd.)
See plate section for color version.

186
Chapter 16: Autism and autism spectrum disorders

suggestive loci identified by the IMGSAC, including these findings [8]. To date, a very large region on
the peaks on 2q, 7q, 16p, and 19p. Three major chromosome 7q likely represents many distinct loci
studies were published in 1999, one from the due to its size and because it has been observed in
Cooperative Linkage Study in Autism (CLSA) with multiple studies with independent samples. Another
75 families [12], the Stanford study with 139 families region on chromosome 17q was identified in the
[14], and the PARIS group with 51 families [13]. AGRE cohort by Stone and colleagues [29] with the
Subsequent linkage studies involved Finnish cohorts striking result that its effect was predominantly due to
[17] with 38 families, an American group in 2002 male affectation status. Subsequently, this signal was
with 90 families [18], and the first paper from the confirmed in an independent sample, the first formal
Autism Genetic Resource Exchange (AGRE; [28]) confirmation of an autism-susceptibility region at
which contained 110 families [16]. genome-wide significance [30]. There are several
It was not until 2003 that a substantial increase in other regions that have been identified in more than
the number of AGRE families warranted another one study of suggestive linkage [8], but again, none of
genome-wide linkage investigation [19]. The AGRE these would meet stringent statistical criteria for
project was prompted by previous studies that had genome-wide significance, including those on
been clearly underpowered and by the fact that a chromosomes 5q and 11.
three- to five-fold increase in sample size would be One of the most recent high-resolution genome-
necessary to detect loci at genome-wide significance wide linkage and association studies of autism
[22]. Despite this increase in the sample to 345 fam- included multiplex families ascertained from the
ilies, Yonan and colleagues did not detect any signifi- AGRE and from the US National Institute for Mental
cant loci. However, several loci, including those on Health (NIMH) repositories [31]. Two novel putative
chromosomes 11 and 5, did garner additional support linkage regions were identified: chromosomes 20p13
as more markers and families were added relative to and 6q27. Although there were no markers with evi-
the earlier work [16]. dence for genome-wide association in families, there
These early results provided more evidence for were many that just missed the significance threshold
heterogeneity in autism, which has been bolstered by suggesting that the study lacked power to detect
recent studies. The University of Washington group’s common variants. Using a case-control approach,
genome-wide scan in 200 families [21] revealed the the authors did find nominal evidence for association
strongest linkage signal on chromosome 7q which at several chromosomes (4q13, 5p15, 6p23, 9p24,
overlapped with the same region identified previously 9q21, 10q21, and 11p14), but replication analyses only
[10, 27]. Even more definitive support for heterogen- identified one significant marker – rs10513025 on
eity came from the largest linkage study to date per- chromosome 5p15, about 80 kb upstream of an axo-
formed by the AGP, an international collaborative nal guidance gene (SEMA5A). No copy number vari-
effort of most of the larger autism genetics groups ations (CNVs) were detected surrounding this SNP or
[24]. Here, despite over 1000 families each with 2 in the SEMA5A locus. Brain expression for SEMA5A
affected individuals and a 10-fold sample increase was reduced in individuals affected with autism com-
over the originally published linkage scans from pared with controls, although this was a marginal
almost a decade earlier, there was not an increase in effect. The authors propose that the inconsistency in
power substantial enough to detect linkage signals. their linkage and association results could be due to
Although no signal reached genome-wide signifi- the limitations of the respective methods, in that
cance, the most prominent region on chromosome linkage scans detect rare variants with strong effects
11p12–13 did reach the suggestive linkage threshold. (20p13), while association tests detect novel common
The unexpectedly modest results from this inter- variants with more modest effects (rs10513025).
national collaboration could be due to the merging All of these linkage studies relied on the qualita-
of samples from distinct geographic regions; the large tive diagnosis of autism for their analysis. Their
sample size may not have compensated for the results strongly suggest that individual common
increase in heterogeneity introduced by pooling of genetic risk factors are not likely to cause the entire
the samples. core deficits required for the broad diagnosis of
A thorough summary of all the major linkage autism or ASD. Thus, we need to define more precise
studies has been published and includes a table of intermediate phenotypes or endophenotypes that

187
Chapter 16: Autism and autism spectrum disorders

comprise components of the disorder that might be affected female [29]. Using this novel approach of
more closely related to a few single genes of small effect stratification by sex, they reanalyzed data from the
size. Endophenotypes are heritable traits characteristic AGRE cohort that had previously been subjected to a
of the disorder and are present in relatives of affected genome-wide linkage analysis [19]. Stone and col-
individuals more frequently than in the unrelated gen- leagues showed significant enrichment for loci due
eral population. In support of this view, the psychiatric to sex related to both female and male sex. Most
genome-wide association studies Consortium Steering importantly, in the male-only families, they identified
Committee recently stated that “Careful attention to a locus with genome-wide significance on chromo-
phenotypic measures (rather than reliance on diag- some 17q. As mentioned previously, Cantor et al.
noses) could prove to be important for identifying replicated this finding in an independent sample from
and replicating susceptibility genes” ([32], p 15). AGRE and performed further fine mapping to narrow
the region to chromosome 17q11 [30]. Although this
is a formal replication on an independent sample,
Endophenotypes in ASD other studies [21] have not identified the same locus
Although it has taken a long time for the autism on chromosome 17, again likely reflecting the signifi-
genetics community to widely adopt the concept of cant heterogeneity in autism.
intermediate endophenotypes, several groups have Investigators of autism genetics have also used
tried to decrease heterogeneity using such phenotypes behavioral endophenotypes (e.g. language deficits) as
to subset patients into potentially more homogeneous covariates in linkage analyses [41, 42] to increase
groups. Among the first of these studies that separ- phenotypic and genetic homogeneity of the sample.
ated probands (with their families) into groups of Shao and colleagues applied the ordered subset analy-
those with and without language delay, resulted in sis (OSA) method to a covariate representing “insist-
the identification of a putative locus reaching a sug- ence on sameness” derived from a principal
gestive level of significance on chromosome 2q [33]. components analysis, and identified a subset of
Bradford et al. used a similar strategy in the CLSA homogeneous families with autism that were respon-
cohort, identifying loci on chromosomes 7 and 13 sible for linkage to a previously reported region on
[34]. However, subsequent analyses in a larger sample 15q11–13 [37]. And in a novel application of latent
using the same method of sample stratification based class analysis to autism, Liu and colleagues [43] used
on language-related endophenotypes while taking symptom counts from the Autism Diagnostic Inter-
parental language history into account, was unable view, Revised (ADI-R) to compute latent class-
to replicate previously reported linkage [35]. derived phenotypes. Nonparametric linkage analysis
Others have successfully used endophenotypes of these phenotypes showed stronger signals in
related to rigidity or obsessive–compulsive disorder chromosomes 3q24, 6q14, 8p12, and 16p13 than com-
to identify putative linkage regions on chromosomes parable analyses using the diagnosis of autism.
1q [36], 15q [37], and 17q [38]. Regression, which is Another way to use endophenotype data is to
observed in up to one-third of autistic probands identify quantitative intermediate phenotypes or
[39], has also been used to select families for linkage endophenotypes that one can use directly as linkage
in the AGRE cohort [40]. This resulted in two seem- variables [44–46]. In other complex diseases, this
ingly strong linkage signals on chromosomes 21q strategy has been used for many metabolic parameters
and 7q in a very small subset of families. Since the (such as plasma triglyceride concentration [47],
21q peak is novel, it may be an artifact due to the fasting insulin concentration [48], and even gene
small sample size and thus it needs additional expression, so-called eQTLs [49–52]). The use of
confirmation. eQTLs for gene identification via linkage methods
Perhaps the most remarkable findings have has already begun in plant and animal models [53]
involved using sex as a phenotype for stratification. and holds much promise for human gene expression
This trait was selected based on the observation that traits. In one recent example, a dominant mutation in
on average males are four times more likely than a gene that causes retinitis pigmentosa (PRPF31) was
females to have autism. Stone et al. reasoned that subjected to eQTL analysis; it was found that the
the genetic risk for affected male-only containing expression of PRPF31, which is correlated with dis-
families might be distinct from those that contain an ease penetrance, was associated with an eQTL in an

188
Chapter 16: Autism and autism spectrum disorders

8.2 Mb region on chromosome 14q21–23 in 15 CEPH stratifying of patients on other biomedical parameters
families [54]. The authors also showed that a nearby or other medical conditions that co-occur with
penetrance factor modulated the expression of both autism, such as specific forms of dysmorphology,
PRPF31 alleles. seizures, or gastrointestinal dysfunction, could likely
With regards to autism, results of the first quali- prove fruitful in the future.
tative trait loci (QTL) analysis were published in 2002
and used language-related endophenotypes for analy-
sis. Alarcón et al. used “age at first word” to identify a Candidate gene association
locus on chromosome 7q35 [55], which they later The vast majority of association studies have involved
confirmed via additional linkage analyses [56] and the assessment of single candidate genes whose selec-
high-density association [57]. In addition, they also tion was based either on biological hypotheses or on
identified a novel locus on chromosome 3q [56], published linkage regions. With few exceptions [57],
which awaits independent replication. SNP mapping these investigations have not relied on dense thor-
of the entire 7q35 QTL region at 5.6 kb density ough screening of entire regions of linkage, thus
revealed a polymorphism in contactin associated pro- making it very difficult to rank the importance of
tein-like 2 (CNTNAP2) associated with age at first such genes relative to each other. Several detailed
word, the first successful linkage-directed association reviews of genetic association studies in autism have
study in autism [57]. Interestingly, as was the case for been published recently [8, 61]. Thus here we focus
the 17q11 locus, male-only containing families were on only a few of the more interesting genes that were
mainly responsible for the CNTNAP2 association identified based on biological hypotheses, as well as
signal. To test the hypothesis that this language endo- regional linkage signals. Please refer to Figure 16.1 for
phenotype was representative of normal population a more comprehensive representation of candidate
variation (e.g. [58]), the relationship of this poly- genes for autism.
morphism with specific language impairment was EN2 1and MET. Two genes, EN2 [62] and MET
investigated: association with the same region was [63], were interesting candidate genes based on their
shown in a sample ascertained for specific language roles in neural development and their presence in the
impairment [59]. Even though this CNTNAP2 asso- 7q2–3 linkage region. In both cases, genetic associ-
ciation with language performance appears to have ation was identified first using the AGRE sample and
been confirmed, the signal from this one gene does was later replicated in other samples; EN2 by the same
not entirely explain the strong linkage signal on group who found the initial association [64], and
chromosome 7q35 in the autism sample, indicating MET by a different group [65]. Notably, with MET,
that additional susceptibility loci for autism must the same promoter polymorphism that was initially
exist within this region. identified by Campbell and colleagues was not repli-
Other more recent studies have also used QTL cated; however, there was other evidence for associ-
approaches. But, none have identified a region of ation of a different SNP in the gene – again, this
genome-wide significance, although certain loci look may reflect the presence of genetic heterogeneity.
promising [21, 43]. The reason for this is probably In the case of the promoter variant that was originally
manifold, including the fact that the endophenotypes identified, Campbell et al. have shown that it is func-
being studied represent relatively broad cognitive tional, whereas the role of the other MET variant is
domains, such as language, and certainly need further unclear. Recently, this group reported an association
refinement. The problem is that most data that have of the MET promoter variant with autism in multi-
been collected in autism are based on diagnostic plex AGRE families with co-occurring gastrointest-
interviews and caregiver questionnaires, which may inal dysfunction [66]. In addition, they found no
not have the optimal properties for accurately meas- association with autism in simplex families without
uring the most relevant aspects of cognition and gastrointestinal conditions. Thus, the authors con-
behavior. More measurable and heritable phenotypes cluded that a disruption in MET signaling may con-
[60], such as brain structure volumes, brain structural tribute to an increased risk for autism with
and functional connectivity, and even more gross gastrointestinal dysfunction. Rather than focus exclu-
phenotypes such as head size, could be interesting sively on a single gene, this line of investigation has
endophenotypes for genetic studies. Further been extended to include other genes that may

189
Chapter 16: Autism and autism spectrum disorders

interact with MET (such as PLAUR and SERPINE1, disrupted in children with autism, with different
[67]), thereby shifting to a pathway approach. brain regions showing variable levels of this disrup-
OXTR and AVPR1a. Other genes of interest that tion [77]. Sutcliffe and colleagues [78] investigated the
have shown at least nominal association with autism serotonin transporter in great detail, performing rese-
in different studies include oxytocin (OXTR) and quencing and identifying several potential deleterious
arginine vasopressin 1a receptor (AVPR1a). Oxytocin variants in the gene. However, once again, these rare
and vasopressin are interesting because of their roles variants that potentially cause functional changes
in parent–offspring social behavior and maternal cannot account for the entire linkage signal. It is
bonding [68, 69]. Copy number variations (CNVs) remarkable that the serotonin receptor sits directly
have been detected in 20p13 [70, 71], a region that under the main linkage peak on 17q11 that is
includes both the OXTR and AVPR1a genes, in indi- genome-wide significant in families with male-only
viduals with autism. Oxytocin has been shown to affected [30]. However, given the data that have so far
decrease activation of the amygdala which affects been published and the fact that both common and
circuitry involved in fear and trust [72]; moreover, rare variants in this gene cannot account for the 17q
variants of OXTR and AVPR1a appear to be involved signal, it is highly likely that other polymorphisms
in amygdala activation and regulation. Variation in and other genes also contribute to the linkage signal.
three microsatellites in AVPR1a has not only been What is becoming clear is that when dense SNP
associated with a diagnostic measure of autism studies of several suggested linkage regions have been
(ADIR), but also with a measure of social skills (the performed, no single common variant can clearly
Vineland Adaptive Behavior Scales) suggesting that account for the linkage signal. This includes regions
this gene may contribute to the social deficits charac- on chromosome 17 [79] and on chromosome 5
teristic of the disorder [73]. Moreover, two of these [Stone, Geschwind, and Nelson, unpublished data].
variants were also associated with amygdala activation Therefore, there may be many common variants, each
and personality in unaffected, healthy individuals with small effect size, that underlie the linkage peaks
[72]. Taken together these studies suggest that the or, alternatively, rare genetic mutations may be con-
neural mechanisms that underlie the disorder may tributing more to autism linkage than was previously
be influenced by the same genes that contribute to suspected.
variability in brain function and behavior in healthy
individuals.
SLC6A4. One of the most persistently studied Whole genome association
genes in autism is the serotonin transporter Recently, the first genome-wide association study in
(SLC6A4). This gene was initially the research focus autism was performed on the AGRE cohort of 700
of Ed Cook and colleagues [74]. Since this gene is multiplex families, consisting of over 3000 subjects.
notoriously difficult to genotype, and evaluating link- Replication was performed in a second cohort of over
age and association results that were unexpectedly 1200 subjects with autism and  6500 controls, all of
positive, investigators noticed that genotypes from a European background. This study was done using the
control group were not in Hardy–Weinberg equilib- Illumina HumanHap550 BeadChip from which geno-
rium and thereby suspected genotype errors [75]. The types from 486 000 markers were used (http://www.
authors further reported that variation in conditions AGRE.org). In the combined analysis of these two
for amplifying the SLC6A4 caused genotype errors cohorts, one family based and one control, a SNP
[75]. An analysis of the corrected SLC6A4 genotypes between cadherin 9 and 10 reached genome-wide
in two autism samples provided no more evidence for significance [80]. This finding was replicated in
linkage than was originally reported [19] and no another sample of nearly 500 multiplex autism fam-
evidence for association [75]. Despite these results, ilies, genotyped on the Illumina HapMap 1M Bead-
serotonin remains a very plausible candidate given Chip, and again confirmed in another small case-
its role in some of the cognitive features found in controlled cohort of 100 cases and 500 control sub-
autism as well as in modulating sleep, which is a jects genotyped on a less-dense array platform. The
considerable problem for patients with autism [76]. combined association from pooled analysis revealed
Also, there is evidence from functional imaging stud- six SNPs in this region on chromosome 5 between
ies that serotonin synthesis during development is cadherin 9 and 10 with a p-value < 8  10–8. The

190
Chapter 16: Autism and autism spectrum disorders

same study also showed significant association when regression analysis [82]. As described previously,
neurexin and cadherin variants were tested as a stratification by sex is an approach that has been used
group. To perform preliminary functional analysis, successfully in linkage studies [29], and, not surpris-
in situ hybridization in developing human fetal brain ingly, sex also proved to be useful in detecting associ-
was performed. There was significant enrichment of ation. Two candidate genes that surpassed the
mRNA expression in the frontal lobe of the develop- genome-wide significance threshold were identified:
ing human embryo for cadherin 10, but no significant the first was a novel gene involved in calcium channel
expression for cadherin 9 at all in mid-gestation fetal detects, RyR2 (rs6683048), and the second was previ-
human brain, making it unlikely that cadherin 9 is ously identified via linkage and association studies,
responsible. However, further studies will have to be UPP2 (rs17420138).
performed to determine this and to identify the crit-
ical functional variants. Nevertheless, this work rep-
resents the first successful whole genome association
Rare versus common variation
in autism, and demonstrates that enormous samples and the role of CNV
are going to be needed to identify variants with Similar to other complex genetic diseases, identifying
smaller effect sizes. The implication of cadherin 10 significant genome-wide linkage and association sig-
adds support to the role of genes involved in various nals in autism has been challenging, as the above
forms of cell adhesion, such as CNTNAP2 and discussion demonstrates. Even when significant link-
Neurexin 1 in ASD. age signals are identified and variants underlying
To address the need for large sample sizes to these signals are found, none of the common variants
continue with genome-wide association attempts, clearly account for the signal, and their effect sizes are
the AGP Consortium reported a genome-wide associ- less than 1.5. This pattern is consistent with two
ation study of 1369 ASD families from North Amer- genetic models: one in which many common variants,
ica and Europe and 1880 controls genotyped with the each with small effect size, interact to produce the
Illumina 1M-Infinium BeadChip array [81]. For this phenotype, and the second in which the role of rare
study, 595 families from the AGRE were used for variants are predominant. Of course, these models are
independent replication although they were not gen- presented as dichotomous as possible, but in many
otyped on the same platform as the AGP sample. The families, both of these and intermediate models are
Illumina HapMap550 markers (described above) that likely involved [58].
were typed in the AGRE families were included in the Support for the rare variant hypothesis has
1-M array thereby facilitating a procedure used to developed over the last two years with the identifica-
infer the missing genotypes. Despite the effort tion of significant contributions from de novo struc-
involved in characterizing the families, genotyping tural chromosomal or CNVs [70]. Sebat and
the samples, and pooling data across continents, only colleagues identified rare chromosomal structural
one SNP (rs4141463) in MACROD2 met genome- variation in 10% of simplex autism cases (i.e. sporadic
wide significance in the primary analysis under the autism) and 3% of the multiplex cohort (i.e. families
specific condition of pooling the AGP families with including at least 2 affected members). A subsequent
the controls. Interestingly, when the exploratory study has identified a similar proportion of de novo
analysis focused on verbal individuals from the AGP rare variation in large autism cohorts [71], while a
and AGRE families, two additional SNPs (rs3784730 recent report using the Illumina Infinium 1-M
in ST8SIA2 and rs2196826 in PLD5) had strong asso- genome-wide microarray did not detect a difference
ciation signals. Results of this study strongly suggest in de novo CNV rates between simplex and multiplex
that endophenotypes should be considered an essen- families with autism [83]. Moreover, Pinto and col-
tial part of the primary, a priori, hypothesis-driven leagues found that genes that were previously impli-
investigation rather than a secondary part of the cated in ASD were disrupted by rare CNV
exploratory, post hoc analysis. significantly more frequently in cases versus controls.
In support of the use of endophenotypes in pri- In addition, rare CNV (i.e. those that occurred only
mary genome-wide association study analyses, Lu and one time) were also present at higher rates in cases
Cantor increased the power to detect association by than in controls, although this effect was marginal.
including sex as a risk factor in their logistic These CNV implicated novel ASD genes including

191
Chapter 16: Autism and autism spectrum disorders

SHANK2, SYNGAP1, and DLGAP2. Genes that over- causing intellectual disability that also lead to autism
lapped with CNV and that had a similar function or a in a small percentage of cases, such as fragile X, Joubert
common pathway were examined as gene-sets and, syndrome, or Smith–Lemli–Opitz syndrome. To date,
subsequently, many novel ASD pathways such as over 100 disease genes and 44 genomic loci involved
GTPase/Ras signaling, cellular proliferation, projec- with intellectual disability have been identified and
tion, and motility were suggested by this approach. may also contribute to the heterogeneity of ASD [92].
Further work is needed to replicate these novel path- Although a significant proportion of patients with
ways and assess their generalizability. each of these disorders have an ASD, each accounts
As studies progress, sample sizes become larger, for no more than 1% of autism cases [8].
and the platform resolution continues to increase,
more and more contribution from rare variation will
be detected [84]. Some models have been proposed to
Is there a single pathway or coherent
suggest that the majority of autism can be attributed biochemical explanation?
to Mendelian inherited factors [85]. However, irre- These new genetic findings beg the question: Despite
spective of the model, these studies continue to iden- the enormous heterogeneity of autism, are there pat-
tify dozens of new potential autism candidate genes, terns of genes or pathways that are implicated in this
regions, and pathways providing a significant advance disorder? It is probably too early to state whether
for the field. these exist or not, but certain patterns are emerging,
In this regard, the recent identification of several such as the role of specific genes involved in neural
forms of recurrent chromosomal structural variation, development (e.g. neural adhesion molecules, as dis-
including 16p11 [86] and 22q11–13 [87], in addition cussed above). The genes identified so far relate to
to the 15q11–13 maternally inherited duplications many distinct biochemical and biological functions.
discovered by Ed Cook and colleagues over a decade Recently, we performed an Ingenuity network analy-
ago [88], also represent important advances. Each of sis with a subset of the most probable candidate genes
these recurrent variations accounts for about 1% of to date. This preliminary analysis demonstrates
ASD, so genetic testing to identify such variants is remarkable connections among many different
warranted. The challenge of moving to individual groups of genes [93]; however, these Ingenuity path-
genes from these findings is that most of these dupli- ways rely on interactions in many different cell types
cations or deletions involve multiple genes and there- and much further investigation is needed to deter-
fore identifying which gene is responsible requires mine whether these genes are related to each other.
large-scale mutation screening and further study in To consider the existence of genetic pathways we
larger populations. Again, the use of high resolution must bear in mind that there may be only a few ways
arrays and the focus on deletions or duplications of in which the brain can respond to a wide variety of
specific exons or genes, in addition to resequencing, developmental insults. It is also important to note that
may help lead to the culprits in these cases. At 22q11– the functions of social cognition, language, and
13, one of the primary candidates is shank3, a synap- mental flexibility, which likely underlie the repetitive,
tic adapter protein. Rare point mutations in shank3 restrictive behavior of autism, represent a huge chunk
have been identified, but when CNV and point muta- of the early brain’s repertoire; therefore many distinct
tions involving shank3 are combined, they are pathways can lead to dysfunction in these areas. Fur-
thought to account for between 0.5 and 1.0% of ther, though many pathways may be involved, it is
ASD [89, 90]. also clear that there is a specific anatomy to social
The identification of each of the recurrent muta- cognition in the human brain, and this relies heavily
tions that underlie ASD risk is an important advance; on anterior temporal and anterior frontal lobes and
but, it is also becoming clear that no single gene their interconnections with sub-cortical circuits
accounts for a major proportion of autism cases, including the striatum. These same regions are also
let alone even 5%. Furthermore, some of these recur- important for language and have been implicated in a
rent CNV, such as those at 16p, are risk factors for a variety of imaging studies [94–96]. This, coupled with
broad range of neurodevelopmental disorders, includ- the anatomical restriction of several major autism
ing schizophrenia [84, 91]. This is consistent with the candidate genes to anterior regions, suggests that
known contribution of many single gene disorders perhaps a critical component of this disorder is

192
Chapter 16: Autism and autism spectrum disorders

what is called the “developmental disconnection” is also seen in some of the single gene disorders that
of these important frontal circuits [97]. For example, are related to autism, such as fragile X and others
not only is contactin associated protein-like 2 [103]. This may also explain why some of the same
(CNTNAP2), strongly enriched in human frontal risk factors that underlie autism, such as neurexin 1
cortex [57, 98], but so is cadherin 10 [80] and MET and perhaps others, are also observed in other dis-
[99]. Modification of the proper development of orders, such as schizophrenia [91, 104], which also
human frontal and anterior temporal lobe circuitry have neurodevelopmental etiologies and also involve
via mutations or polymorphisms in these genes pro- frontal systems. It is further likely that disorders that
vides one potential mechanism by which genetic risk phenotypically overlap with autism, such as attention-
factors could influence specific brain circuits that are deficit hyperactivity disorder (ADHD), share some
involved in the cognitive and behavioral phenotype genetic risk with autism; one can imagine that there
of autism. will be other discrete genetic components that will
It is notable that the hypothesis of developmental distinguish these disorders from each other despite a
disconnection represents a convergence of many years potentially overlapping core set of genes affecting
of imaging and electrophysiological studies based on these circuits [91]. In support of this hypothesis,
the pioneering work by Marcel Just and colleagues Ronald and colleagues reported that there are
[100–102], which has implicated functional connect- common genetic influences on ADHD behaviors
ivity in the brain as a potential cause of the autism and autism traits both in a community sample of
phenotype. Genetic evidence now seems to support twins and in a subsample of the extreme individuals
this hypothesis, as deficits in many distinct levels of that are likely comorbid for these disorders. To con-
molecular function, from synapse to neuronal migra- tinue to move forward in autism genetics, the field
tion to neuronal path finding, as well as neuronal needs very large samples, some of which should
maturation, have been implicated, and all could con- include undiagnosed family members, that should be
verge on disconnection [97]. These ideas and data phenotyped using instruments that measure concepts
suggest that in addition to molecular cellular studies and constructs related to a broad range of cognitive
of genes they must also be put into an anatomical functions that overlap with other disorders such as
context, since autism likely reflects the disruption of a ADHD, bipolar disorder, and schizophrenia. Func-
variety of specific circuits. Given its clear phenotypic tional imaging studies connecting gene to brain in
heterogeneity, it is also likely that there is a wide humans, and the use of animal models to incorporate
variety of distinct brain regions involved; for the molecular and developmental mechanisms, are both
disorder to comprise deficits in social cognition and very exciting, and given the number of genes that
language at its core there must be some convergence have been identified, there will likely be an explosion
in frontal systems. Involvement of these brain regions of such studies over the next several years.

References 5. Ritvo ER, et al. Am J Psychiatry


1990;147:1614–1621.
13. Philippe A, et al. Hum Mol Genet
1999;8:805–812.
1. American Psychiatric Association.
Diagnostic and Statistical 6. Rutter M. J Abnorm Child Psychol 14. Risch N, et al. Am J Hum Genet
Manual of Mental Disorders, 2000;28:3–14. 1999;65:493–507.
Fourth Edition, Text Revision 7. Folstein SE, et al. Nat Rev Genet 15. IMGSAC. Am J Hum Genet
(DSM-IV-TR). Washington, DC: 2001;2:943–955. 2001;69:570–581.
American Psychiatric
8. Abrahams BS, et al. Nat Rev Genet 16. Liu J, et al. Am J Hum Genet
Association; 2000.
2008;9:341–355. 2001;69:327–340.
2. Kanner L. Nervous Child
9. Patterson PH. Behav Brain Res 17. Auranen M, et al. Am J Hum
1943;10:217–250.
2009;204:313–321. Genet 2002;71:777–790.
3. Gillberg C. J Intellectual Disabilities
10. IMGSAC. Hum Mol Genet 18. Shao Y, et al. Am J Med Genet
Res 1992;36:201–214.
1998;7:571–578. 2002;114:99–105.
4. Gillberg C. et al. (eds.). The
Biology of the Autistic Syndromes. 11. Ashley-Koch A, et al. Genomics 19. Yonan AL, et al. Am J Hum Genet
Vol. 126. New York: MacKeith 1999;61:227–236. 2003;73:886–897.
and Cambridge University 12. Barrett S, et al. Am J Med Genet 20. McCauley JL, et al. BMC Med
Press; 1992. 1999;88:609–615. Genet 2005;6:1.

193
Chapter 16: Autism and autism spectrum disorders

21. Schellenberg GD, et al. Mol 43. Liu XQ, et al. Biol Psychiatry 65. Sousa I, et al. Eur J Hum Genet
Psychiatry 2006;11:1049–1060. 2008;64:561–570. 2009;17:749–758.
22. Lander E, et al. Nat Genet 44. Streeten EA, et al. 66. Campbell DB, et al. Pediatrics
1995;11:241–247. J Bone Miner Res 2009;123:1018–1024.
23. Gillberg C, et al. Dev Med Child 2006;21:1433–1442. 67. Campbell DB, et al. Autism Res
Neurol 1985;27:293–304. 45. Doyle AE, et al. Am J Med 2008;1:159–168.
24. Szatmari P, et al. Nat Genet Genet B: Neuropsych Genet 68. Nelson EE, et al. Neurosci
2007;39:319–328. 2008;147B:1399–1411. Biobehav Rev 1998;22:437–452.
25. IMGSAC. Hum Mol Genet 46. Mottl AK, et al. Kidney Int 69. Heinrichs M, et al. Prog Brain Res
2001;10:973–982. 2008;74:1185–1191. 2008;170:337–350.
26. Cook EH Jr. Child Adolesc 47. Coletta DK, et al. Diabetes 70. Sebat J, et al. Science
Psychiatr Clin N Am 2001; 2009;58:279–284. 2007;316:445–449.
10:333–350. 48. Duggirala R, et al.
71. Marshall C, et al. Am J Hum Genet
27. Badner JA, et al. Mol Psychiatry Am J Hum Genet
2008;82:477–488.
2002;7:56–66. 2001;68:1149–1164.
72. Meyer-Lindenberg A,
28. Geschwind DH, et al. Am J Hum 49. Cheung VG, et al. Nature
et al. Mol Psychiatry
Genet 2001;69:463–466. 2005;437:1365–1369.
2009;14:968–975.
29. Stone JL, et al. Am J Hum Genet 50. Evans DM, et al. Trends Genet
73. Yirmiya N, et al. Mol Psychiatry
2004;75:1117–1123. 2006;22:350–354.
2006;11:488–494.
30. Cantor RM, et al. Am J Hum 51. Nica AC, et al. Hum Mol Genet
74. Cook EH Jr., et al. Mol Psychiatry
Genet 2005;76:1050–1056. 2008;17:R129–R134.
1997;2:247–250.
31. Weiss LA, et al. Nature 52. Cookson W, et al. Nat Rev Genet
75. Yonan AL, et al. Psych Genet
2009;461:802–808. 2009;10:184–194.
2006;16:31–34.
32. Psychiatric GWAS Consortium 53. Schadt EE, et al. Nature
76. Liu X, et al. Child Psychiatry Hum
Steering Committee. Mol 2003;422:297–302.
Dev 2006;37:179–191.
Psychiatry 2009;14:10–17. 54. Rio Frio T, et al. Hum Mol Genet
77. Chugani DC. Mol Psychiatry
33. Buxbaum JD, et al. Am J Hum 2008;17:3154–3165.
2002;7:S16–S17.
Genet 2001;68:1514–1520. 55. Alarcón M, et al. Am J Hum Genet
78. Sutcliffe JS, et al. Am J Hum Genet
34. Bradford Y, et al. Am J Med Genet 2002;70:60–71.
2005;77:265–279.
2001;105:539–547. 56. Alarcón M, et al. Mol Psychiatry 79. Stone JL, et al. Hum Mol Genet
35. Spence SJ, et al. Am J Med 2005;10:747–757. 2007;16:704–715.
Genet B: Neuropsych Genet 57. Alarcón M, et al. Am J Hum Genet 80. Wang K, et al. Nature 2009;459:
2006;141B:591–598. 2008;82:150–159. 528–533.
36. Buxbaum JD, et al. Mol Psychiatry 58. Geschwind DH. Cell 81. Anney R, et al. Hum Mol Genet
2004;9:144–150. 2008;135:391–395. 2010;19:4072–4082.
37. Shao Y, et al. Am J Hum Genet 59. Vernes SC, et al. N Engl J Med 82. Lu A, et al. Mol Psychiatry
2003;72:539–548. 2008;359:2337–2345. 2012;17:215–222.
38. McCauley JL, et al. Am J Med 60. Glahn DC, et al. Hum Brain Mapp 83. Pinto D, et al. Nature 2010;466:
Genet 2004;127B:104–112. 2007;28:488–501. 368–372.
39. Richler J, et al. J Autism 61. Losh M, et al. J Neuropathol 84. Morrow EM. J Am Acad Child
Dev Disord 2006;36: Exp Neurol 2008;67: Adolesc Psychiatry 2010;
299–316. 829–837. 49:1091–1104.
40. Molloy CA, et al. Mol Psychiatry 62. Gharani N, et al. Mol Psychiatry 85. Zhao X, et al. Proc Natl
2005;10:741–746. 2004;9:540. Acad Sci U S A
41. Dementieva Y, et al. Proceedings 63. Campbell DB, et al. Proc Natl 2007;104:12831–12836.
of the IMFAR. Orlando: IMFAR; Acad Sci U S A 2006;103: 86. Weiss LA, et al. N Engl J Med
2002. 16834–16839. 2008;358:667–675.
42. Shao Y, et al. Am J Hum Genet 64. Benayed R, et al. Am J Hum Genet 87. Manning MA, et al. Pediatrics
2002;70:1058–1061. 2005;77:851–868. 2004;114:451–457.

194
Chapter 16: Autism and autism spectrum disorders

88. Cook EH Jr., et al. Am J Hum 94. Kleinhans NM, et al. Brain Res 99. Mukamel Z, et al. J Neurosci,
Genet 1997;60:928–934. 2008;1221. 2011;31:11437–11442.
89. Durand CM, et al. Nat Genet 95. Mundy P. J Child Psychol 100. Hughes JR. Epilepsy Behav 2007;
2007;39:25–27. Psychiatry 2003;44:793–809. 11:20–24.
90. Moessner R, et al. Am J Hum 96. Redcay E, et al. Biol Psychiatry 101. Just MA, et al. Cereb Cortex 2007;
Genet 2007;81:1289–1297. 2008;64:589–598. 17:951–961.
91. Walsh T, et al. Science 2008; 97. Geschwind DH, et al. Curr 102. Koshino H, et al. Cereb Cortex
320:539–543. Opin Neurobiol 2007; 2008;18:289–300.
92. Betancur C. Brain Res 2011; 17:103–111. 103. Abrahams BS, et al. Arch Neurol
1380:42–77. 98. Abrahams BS, et al. Proc Natl 2010;67:395–399.
93. Bill BR, et al. Curr Opin Genet Dev Acad Sci U S A 2007;104: 104. Cantor RM, et al. Neuron 2008;
2009;19:271–278. 17849–17854. 58:165–167.

195
The genetics of bipolar disorder
Chapter

17 John R. Kelsoe

Introduction the general population, suggesting a partially shared


genetic basis for these two mood disorders. The study
The heritable nature of mood disorders has been by Wozniak et al. further addresses risk in the families
observed since ancient times. Over the last century of probands with childhood onset [1], where an 18-
great advances have been made in understanding the fold elevation in risk is observed compared to the
genetic transmission and more recently identifying general population. This risk is 2–3 times higher than
some of the responsible genes. Study of the genetics in adult-onset probands and reflects a substantially
of bipolar disorder has great potential to improve our higher genetic loading.
understanding of etiology and mechanism and may
lead to new treatments and methods of diagnosis. In
this chapter, the family studies supporting the role of Twin studies
genetics will be reviewed and recent molecular genetic Family transmission, however, is not equivalent to
results discussed. The possible role of this informa- genetic contribution. Families share many factors
tion in clinical treatment is also considered. besides genes, such as behavioral environment, or
alternatively, diet or infectious factors. Twin studies
have been the preferred strategy for separating nature
Family epidemiology and nurture. Typically, monozygotic (MZ) and dizy-
Family studies gotic (DZ) twins are ascertained where one twin has
bipolar disorder, and the rate of illness in the co-twin,
Numerous family studies over many years have docu- or concordance rate, is determined. In the most
mented the familiality of bipolar disorder. In this common design, twins are selected who were raised
design, typically, each proband in a clinic or other together in the same household in order to control for
source is systematically queried regarding psychiatric environmental effects. Table 17.2 illustrates many of
illness in their first-degree relatives. Though such the twin studies conducted in bipolar disorder or
studies are limited by recollection of the proband affective illness over the last 80 years. Though, there
and knowledge of family members, these studies have is much methodological and diagnostic variability
provided much evidence regarding the genetic trans- between studies, overall they demonstrate an approxi-
mission of bipolar disorder. Table 17.1 summarizes a mate 70% concordance rate for MZ compared to 25%
selection of 20 of these family studies, including over for DZ twins, thereby arguing for a substantial herit-
11 000 relatives, that together show a risk of approxi- ability of bipolar disorder. That all co-twins are not
mately 7% in first-degree relatives. Compared to the affected is evidence for reduced penetrance of bipolar
general population rate of approximately 1%, this susceptibility genes, and implies that other nonheri-
indicates a 7-fold increase in risk and a substantial table factors also play a role in etiology.
level of familiality. This increase in risk is consistently
observed despite the variability in diagnostic methods
employed amongst these studies. It is also of interest Mode of transmission
that unipolar disorder (UPD) or major depression Though, in principle, one should be able to examine
(MDD) is observed at roughly twice the rate as in the patterns of distribution of illness in families and

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

196
Chapter 17: Genetics of bipolar disorder

Table 17.1 Morbid risk among first-degree relatives of probands with bipolar disorder.

Study Relatives at risk Morbid risk (%)

Bipolar Unipolar
Perris, 1966 [100] 627 10.2 0.5
Winokur et al., 1967 [101] 167 10.2 20.4
Mendlewicz et al., 1974 [102] 606 17.7 22.4
Goetzl et al., 1974 [103] 212 2.8 13.7
Helzer et al., 1974 [104] 151 4.6 10.6
Gershon et al., 1975 [105] 341 3.8 6.8
James et al., 1975 [106] 239 6.4 13.2
Johnson et al., 1977 [107] 126 15.5 19.8
Petterson, 1977 [108] 472 3.6 7.2
Smeraldi et al., 1977 [109] 172 5.8 7.1
Angst, 1980 [110] 400 2.5 7
Dunner et al., 1980 [111] 1199 4.2 8.2
Taylor et al., 1980 [112] 600 4.8 4.2
Fieve et al., 1984 [113] 2171 6.6 9
Gershon et al., 1982 [5] 598 8.0 14.9
Tsuang et al., 1985 [114] 608 3.9 9.1
Rice et al., 1987 [4] 557 5.7 23
Sadovnick et al., 1994 [115] 1102 3.5 5.7
Birmaher et al., 2009 (Offspring) [116] 388 10.6 24.3
Wozniak et al., 2009 (Child probands) [1] 487 18 49
Total 11223 6.9 12.1

Table 17.2 Twin studies of bipolar disorder.

Monozygotic Dizygotic

Concordant Total % Concordant Total %


Luxenberger, 1930 [117] 3 4 75.0 0 13 0.0
Rosanoff et al., 1935 [118] 16 23 69.6 11 67 16.4
Slater, 1953 [119] 4 7 57.1 4 17 23.5
Kallman, 1954 [120] 25 27 92.6 13 55 23.6
Harvald et al., 1965 [121] 10 15 66.7 2 40 5.0
Allen et al., 1974 [122] 5 15 33.3 0 34 0.0
Bertelsen, 1979 [123] 32 55 58.2 9 52 17.3
Kendler, 1993 [124] 107 154 69.5 114 326 35.0
Total 202 300 67.3 153 604 25.3

197
Chapter 17: Genetics of bipolar disorder

infer mode of transmission, in practice this approach multiple genes may be contributing to illness. Evi-
has yielded limited information. Though such segre- dence for a parent-of-origin effect has also been
gation analyses can be quite successful in simple reported for bipolar disorder. Maternal transmission
Mendelian traits, their relative lack of success in bipo- has been reported to be more common than paternal
lar disorder is consistent with a large amount of transmission [11]. This would be consistent with two
epidemiological and molecular data arguing for a atypical forms of genetic transmission: genomic
complex form of transmission involving many genes. imprinting or mitochondrial transmission. Genomic
Attempts to identify a Mendelian form of transmis- imprinting is seen in disorders such as Prader–Willi
sion in family data have provided some modest sup- or Angelman syndrome where different syndromes
port for a dominant single major locus effect [2–4]. result from transmission of the mutation through
Alternatively, other investigators have examined the father or the mother. However, maternal trans-
family data for evidence of multifactorial or polygenic mission has not been consistently observed to be
transmission in which numerous genes of small effect increased [12]. Despite continued study, it has been
combine to produce liability to illness. Such models difficult to obtain molecular evidence for either
are usually applied to quantitative traits, and may be imprinting or mitochondrial transmission. However,
consistent with the range of severity of illness seen in the area of epigenetics in general is a very active area
family members. A related model is the multithres- of investigation.
hold multifactorial model, in which numerous genes
of small effect contribute to a latent liability to
psychiatric illness that is normally distributed in Do subphenotypes breed true?
the population. Individuals carrying more suscepti- Bipolar disorder manifests itself in a broad range of
bility alleles are affected with more severe forms of symptoms and presentations. An important question
illness [5]. is whether any of these features or presentations of
Family studies have revealed some other interest- bipolar disorder represent distinct gene effects or
ing aspects of the genetic transmission of bipolar genetically distinct forms of illness. For example, do
disorder. The cohort effect describes the observation different genes predispose to bipolar I versus bipolar
that the age of onset of bipolar disorder was earlier for II disorders? If bipolar disorder can be broken down
cohorts born later in the twentieth century than those into different subphenotypes that are genetically dis-
born earlier [4]. The cause of this well replicated effect tinct, then each would potentially represent a simpler
is unclear. It may simply be differences in recollection problem for gene mapping studies. Furthermore,
of onset events in older versus younger people. How- knowledge of which genes are involved in different
ever, it has also been proposed that this is an indica- forms of illness might lead to an understanding of the
tion of the phenomenon of anticipation. Anticipation different biological mechanisms operating in those
is a hallmark of genetic disorders in which the patho- forms of illness. Family studies of different types or
genic mutation is an expanding trinucleotide repeat, features of bipolar disorder have been conducted in
such as fragile X syndrome or Huntington’s disease. order to identify those forms of illness that might be
McInnis et al. [6] have reported a decrease in the age genetically distinct. If such a subphenotype is prefer-
of onset and increase in disease severity in successive entially transmitted in families or breeds true, then it
generations. Though the family data seemed intri- might result from a distinct set of predisposing genes.
guing, molecular studies have not been successful in A variety of clinical features of bipolar disorder
clearly identifying a trinucleotide repeat expansion have been examined for their validity as subpheno-
sequence associated with bipolar disorder [7]. types. Studies of the bipolar II phenotype suggest that
Another complicating phenomenon in the genet- it may be partially genetically distinct. Affected family
ics of bipolar disorder is assortative mating [8–10]. members of bipolar I probands primarily exhibit an
This is the tendency for patients with affective dis- increase in bipolar I illness, while family members of
orders to marry spouses who also have affective dis- bipolar II probands are more likely to have bipolar II
orders, thereby, producing bilineal families where [13]. Psychotic features have been reported to occur
illness comes from both the mother’s and father’s side more often in affected relatives of psychotic bipolar
of the family. This presumably results in greater gene- probands than nonpsychotic probands [14, 15].
tic loading, and more complex genetics where Familiality has also been reported for bipolar disorder

198
Chapter 17: Genetics of bipolar disorder

3.5 Figure 17.1 Familiality of different


Adjusted for sex, age, Dx subphenotypes of bipolar disorder. The
3 odds ratio for different subphenotype are
Adjusted for sex, age, Dx, site/wave presented calculated as the risk to other
2.5 bipolar family members in families in
which the proband has the
Odds ratio

2
subphenotype of interest. Hatched bars
indicate the odds ratio when covaried for
1.5
diagnosis (Dx), age, and sex. The solid
bars indicate a generally lower odds ratio
1
when site and wave are also included
.67
as covariates. Stars indicate
subphenotypes with at least nominal
significance. MDE, major depressive
episode. (Adapted from [19].)
m er is ts pt d s g g E DE
lis rd os gh tem oo ode clin chin DE DE DE DE MD DE
ho di
so ch ou at
m is cy it n
M
n
M
n
M
n
M
in n
M
in
M
lco c Ps
y
lt
h
de b l e e p
pi
d s w i
a
i
a
i
e
i e y
i s
a ni da ita nic d AM ni ni tit tit rg es
bi
d pa ici ici irr a Ra oo in som som ppe pe ha sn
or id Su Su i th d m m s e n r a ap L et l es
rb i d I e t
om o w e p wo
r p ed e d s
C m ia Mix Ra Hy eas as Re
Co an d re
M oo e cr I nc
M D

Subphenotypes

with comorbid panic attacks [16], suicidality, and temperaments have been shown to have different
rapid mood switching [17]. Several studies have courses of illness [22]. Evans et al. [23] examined five
investigated a number of traits systematically. Schulze temperaments (dysthymia, hyperthymia, cyclothymia,
et al. [18] examined 1246 individuals in 172 families irritable, and anxious) as assessed by the TEMPS-A
for 32 clinical variables from the diagnostic interview. and found evidence for familiality for cyclothymia,
Five clinical features were significantly correlated hyperthymia, and anxious temperaments.
between family members after correction for multiple
comparisons. These included: social relations, sub-
stance abuse, alcoholism, psychosis, and suicide Gene mapping methods
attempts. A similar analysis was conducted by Saun- In the last two decades, genetic studies of bipolar
ders et al. [19], who examined 1416 bipolar siblings in disorder, as with other genetic disorders, have focused
589 sibships which were part of the National Institute largely on using molecular genetic mapping methods
of Mental Health (NIMH) Genetics Initiative collec- to identify the specific genes responsible for the herit-
tion. As shown in Figure 17.1, rapid cycling showed ability described above. The human genome project
the strongest evidence for familiality. Other signifi- has vastly accelerated this process by providing power-
cantly familial traits included: comorbid alcoholism, ful tools for gene identification. A detailed review of
comorbid panic disorder, psychosis, suicidal mapping methods can be found elsewhere in this
thoughts, and rapid mood switching. Similar results volume, but these methods will be briefly described
were obtained by Potash et al. in their examination of here. The real power of these mapping methods is: (1)
the Phenome Database. They in particular reported their ability to identify novel genes not previously
strong familiality for missed work, obsessive–compul- suspected; and (2) their comprehensive coverage of
sive disorder (OCD), hospitalization, and age at onset all genes in the genome. This has led in psychiatric
[20]. These studies suggest subphenotypes of bipolar disorders and many other genetic disorders to the
disorder that may be useful in mapping genes and discovery of completely new genes not previously
understanding the pathophysiology. implicated in illness. This in turn spawns new hypoth-
Affective temperament is another feature of eses regarding these genes, and new ideas about mech-
bipolar disorder that has been hypothesized to reflect anism. Prior to this approach, most hypotheses of
genetically distinct subphenotypes. In this model, pathophysiology were based on the mechanism of
affective temperaments are genetically transmitted action of medications effective for the disorders. This
and in turn predispose to episodes of mania or depres- has been useful, but has restricted the range of hypoth-
sion [21]. Bipolar patients with different underlying eses to a few systems affected by these drugs.

199
Chapter 17: Genetics of bipolar disorder

There are two major approaches to genetic map- GWAS were proposed over a decade ago [24], but
ping that are employed. Genetic linkage examines the awaited two developments to become practical. First,
co-segregation of anonymous DNA markers with ill- the Human Genome Project identified millions of
ness in families. If an allele of a marker is consistently SNP markers suitable for covering the genome at high
transmitted from affected parent to affected offspring density. Secondly, microarray-based methods were
in a family, this suggests that a gene for the disease developed that enabled the inexpensive and rapid
maps near that marker on a chromosome. If the genotyping of hundreds of thousands of markers
chromosomal location of the marker is known then simultaneously. Currently, GWAS typically genotype
the approximate location of the gene is known. Practic- approximately 1 million markers at an average inter-
ally, this has been accomplished by using 400–500 DNA val of 3 kb using 1 of 2 genotyping platforms.
markers spread evenly across the genome at about A strength of the association method is its ability to
10 million base pair (Mb) intervals. Historically, this detect small gene effects, i.e. mutations that contrib-
has been done using markers that have a highly variable ute only a small amount to disease risk. While linkage
number of tandem repeats. More recently, dense maps has been successfully used for Mendelian traits, in
of single nucleotide polymorphisms (SNPs), or single which one gene is primarily responsible for genetic
base substitutions have been used as well. Such studies transmission; association is ideally suited to detect
would approximately localize a gene. More detailed fine polygenic transmission, where many genes each con-
mapping studies were then required to identify the tribute a small amount to an overall cumulative sus-
exact gene in the region. The Human Genome Project ceptibility. This is also referred to as the common
made available dense maps of such markers, and later variant–common disease model. However, an issue
maps of the genes within the linkage peaks. with GWAS is that very large sample sizes may be
Genetic association is an alternative approach that required to detect such small gene effects. Also, very
compares the frequency of a given marker allele in low p-values are required to survive the multiple
cases versus controls. If a specific allele occurs more comparisons correction for the one million statistical
often in cases than controls, this suggests that that tests conducted on datasets of one million or more
marker or a mutation very near it contributes to markers. The standard threshold for genome-wide
susceptibility. A key difference between linkage and significance is p < 5  108, which is roughly the
association is the genetic resolution. Linkage can be Bonferroni correction.
detected if the marker is within approximately 10 Mb The future of genetic mapping is whole genome
of the causative mutation. Association is a much sequencing. This will allow for complete detection of
higher resolution technique and requires generally all genomic variation. Rapid and inexpensive “next-
that the marker is within 2–100 kilobases (kb) of the generation” sequencing methods are evolving quickly
mutation. Association can detect nearby mutations and sequencing costs are dropping rapidly. Soon, all
because of the phenomenon of linkage disequilibrium. previous methods will be replaced by the collection of
If a new mutation causing a disease occurs on a complete sequence information. This approach will
chromosome very near a marker SNP, then it will be required to test the multiple rare variant model of
tend to remain associated with the marker allele as transmission. This model posits that numerous rare
the population expands. The closer they are together, variants of strong effect cause genetic vulnerability.
the fewer opportunities for recombination or other Many different mutations in a number of different
events to disconnect them. This is statistically evalu- genes transmit susceptibility, but each mutation may
ated by observing whether the two alleles (disease and be so rare as to essentially be a “private mutation” in
marker) occur together more often than expected by that family. Genes involved in illness may be detected
chance. Comparing the two approaches, association by demonstrating a larger number of rare likely func-
generally is able to identify specific genes, while link- tional mutations in cases as compared to controls.
age identifies much broader regions. As a result, the Such studies are just now beginning.
entire genome can be tested for linkage with several
hundred markers, whereas, many more markers are
required to test the whole genome for association. Linkage studies
This has made genome-wide association studies Linkage studies in bipolar disorder began in 1987
(GWAS) impractical until recently. with the report of linkage to chromosome 11p15 in

200
Chapter 17: Genetics of bipolar disorder

1 2 3 4 5 6 7 8 9 10 11
4 8
1
1

1
13
13 12
1
2
11

18
15
14

9
1
5
10 2
16 11 6
9
3 1
6

12 13 14 15 16 17 18 19 20 21 22 X Y

Figure 17.2 Chromosomal regions implicated by linkage studies. Chromosomal regions reported to be linked to bipolar disorder are
illustrated. Circles represent individual study reports, stars represent regions implicated by a meta-analysis or combined analysis. Numbers
in the circles or stars refer to referenced papers as follows: 1 ¼ Nurnberger et al. [180], Detera-Wadleigh et al. [181], Rice et al. [182],
Edenberg et al. [183]; 2 ¼ Kelsoe et al. [184]; 3 ¼ Detera-Wadleigh et al. [185]; 4 ¼ Blackwood et al. [186]; 5 ¼ Berrettini et al. [29]; 6 ¼ Straub
et al. [187]; 6 ¼ Morissette et al. [58]; 7 ¼ Pekkarinen et al. [188]; 8 ¼ Egeland et al. [25]; 9 ¼ Freimer et al. [30]; 10 ¼ Stine et al. [31];
11 ¼ Dick et al. [189]; 12 ¼ Cichon et al. [190]; 13 ¼ McInnis et al. [191]; 14 ¼ Adams et al. [192]; 15 ¼ Coon et al. [193]; 16 ¼ Turecki et al. [73];
17 ¼ Seguardo et al. [33]; 18 ¼ Badenhop et al. [194].

a large pedigree from the Old Order Amish popula- Several chromosomal regions have been the focus
tion in southeastern Pennsylvania [25]. However, of particular attention. The original Amish study was
subsequent studies of an expanded version of the of particular interest because the gene for tyrosine
same family demonstrated much reduced evidence hydroxylase (TH) and brain-derived neurotrophic
for linkage [26]. This pattern of inconsistent replica- factor (BDNF) mapped in close proximity to the
tion has plagued linkage studies of bipolar disorder. linked markers. Both of these genes have also been
Figure 17.2 summarizes the major chromosomal reported to be associated with bipolar disorder.
regions that have been reported to be linked to bipolar Though the expansion of the family reduced the evi-
disorder. There have been over 20 linkage genome dence for linkage (LOD score) from over 4.0 to 2.2,
scans conducted in bipolar disorder. Considering all there still remained interesting support for linkage.
of the regions putatively implicated by these scans, a This region, however, has not figured prominently in
substantial portion of the genome shows some evi- subsequent linkage studies. A similar story occurred
dence of linkage. in early reports of linkage to the X chromosome.

201
Chapter 17: Genetics of bipolar disorder

X linkage was reported in a series of Israeli families, enough statistical power to detect linkage. However,
that on re-examination could not be replicated [27, in many cases, in part due to assortative mating, more
28]. Chromosome 18 was first implicated by Berret- than one gene for bipolar disorder might be segregat-
tini et al. [29] who found evidence of linkage to a ing with illness, thereby confounding this strategy.
pericentromeric region. Subsequently, three other This was likely the case for the large Amish family
regions on chromosome 18 were reported to be described above. Later linkage studies focused on
linked. Two of these came from studies of a Costa affected sibling pairs and much smaller families.
Rican population isolate [30], while one demonstrated Though this approach may be more robust to intra-
a maternal pattern of transmission [31]. family heterogeneity, it may have less power to detect
The inconsistency of results between studies has rare genes. Therefore, in interpreting these results, it
made the interpretation of these data challenging. is important to keep in mind that different sizes of
Several attempts have been made to systematically families represent different linkage strategies that are
analyze these combined data. Badner et al. [32] con- powered to detect different kinds of gene effects.
ducted a meta-analysis of 11 bipolar genome scans. Association results as described below are also best
They employed the Multiple Scan Probability (MSP) powered to detect genetic effects in the context of
which was a modification of Fisher’s meta-analysis polygenic transmission, or the common variant
using p-values. Significant evidence of linkage was common disease model. Many studies today are based
obtained for two regions: 13q and 22q. Segurado on the assumption that genes for bipolar disorder are
et al. [33] subsequently examined 18 bipolar genome affected by many different mutations, some of which
scans using the Genome Scan Meta-Analysis (GSMA) are rare and some common. Therefore, some incon-
method. This procedure involved binning the genome sistency in findings may result from the fact that
and ranking each study across all bins. Ranks were different study designs are best powered to detect
then compared across regions and distributions different kinds of gene effects.
examined. No region achieved genome-wide signifi- Another intriguing element is that of natural
cance. However, several regions showed interesting selection. Association relies on the existence of
evidence of linkage including: 9p22, 10q11, and common functional variants. Though many mechan-
14q24. It is interesting that two of these on chromo- isms may exist to lead to the amplification of an allele
somes 9 and 14 were never a major linkage peak in in a population, such as population “bottlenecks” or
any single study. Rather small contributions from random genetic drift, one mechanism may be selec-
many studies led to this result. McQueen et al. [34] tion. If a mutation associated with some aspect of
conducted an analysis of 11 genome-wide linkage bipolar disorder is evolutionarily advantageous, then
scans in which they combined the original genotype it will be selected, thereby increasing the frequency of
data and found genome-wide significant evidence for that mutation in the population. Its ultimate fre-
linkage on chromosomes 6q and 8q. quency may represent balanced selection, where the
There are several possible interpretations of these advantages that increase the frequency are balanced
data and the inconsistency that characterizes them. against its disadvantages so as to achieve a specific
Heterogeneity is frequently invoked as a reason for allele frequency.
nonreplication. It is almost certain that multiple genes
may cause bipolar disorder, but the number of genes
remains unclear. Not only may different genes be Candidate gene studies
involved, but there may be numerous different muta- In a candidate gene study, a specific gene is hypothe-
tions in those genes. Linkage is specific to the gene sized to be involved in bipolar disorder because of its
and linkage tests can detect a gene’s effect even if known function in brain physiology. Known markers
different mutations in the gene are operating in dif- or functional variants can then be tested in a case-
ferent families. Association, however, is specific to the control association study to determine if the gene is
allele, and if a large number of mutations occur in a involved in illness. The advantage of this approach is
gene that may predispose to bipolar disorder, that its physiological relevance and limited number of
effect may be difficult to detect by association statistical tests. The disadvantages are its inability to
methods. Early linkage studies focused on large fam- identify novel previously unsuspected genes and that
ilies such that one extended family might have it does not comprehensively cover the genome.

202
Chapter 17: Genetics of bipolar disorder

Table 17.3 Selected candidate genes reported to be associated to bipolar disorder.

Gene symbol Gene name and function References


SLC6A4 Serotonin transporter, reuptake of serotonin from the synapse [125–139]
BDNF Brain derived neurotrophic factor, neuronal growth and maintenance [42, 140]
COMT Catechol-O-methyltransferase, inactivation of catecholamines [45, 46, 141–143]
TH Tyrosine hydroxylase, rate limiting enzyme in catecholamine biosynthesis [144–150]
NTRK2 Trkb neurotrophin tyrosine kinase receptor for BDNF [43]
ADRBK2 (GRK3) G protein receptor kinase 3, desensitization of G protein coupled receptors [60, 62, 151]
P2RX7 ATP receptor [59, 152–155]
TPH2 Tryptophan hydroxylase 2, rate limiting enzyme in serotonin biosynthesis [156, 157]
GRIK4 Glutatmate receptor, kainate type [158]
DAOA D-amino acid oxidase activator, modulator of glutamate neurotransmission [48, 52]
first identified in schizophrenia
DISC1 Disrupted in schizophrenia, identified in a large Scottish family with a [159–162]
balanced translocation
Sp4 Transcription factor [54, 63, 163, 164]
NRG1 Neuregulin1, signaling molecule originally identified in schizophrenia [165]
CLOCK Clock gene [166, 167]
BMAL1 Clock gene [168]
GNAL G protein in the olfactory bulb [169]
GRIN2B NMDA glutamate receptor subunit [170–172]
SLC6A3 (DAT) Dopamine transporter, reuptake of dopamine [95, 173–177]
CACNG2 Stargazin, AMPA type glutamate receptor associated protein [178]

A large number of such candidate gene association A number of studies have been conducted examining
studies have been conducted. Some of the more the role of this gene in bipolar disorder (Table 17.3),
prominent genes and studies are listed in Table 17.3. though results have been mixed and inconsistent.
See this table for detailed citations. Together, these data do not present a strong case for
The serotonin transporter has been one of the the gene’s role in vulnerability to bipolar disorder.
most widely studied genes in psychiatry [35]. The A limitation of these studies, as with several candidate
transporter plays a key role in serotonin neurotrans- genes, is that the focus of investigation has almost
mission by mediating the active reuptake of serotonin entirely been on the VNTR polymorphism while
at the synapse. It is also the target of selective sero- other functional variants are known [37]. This VNTR
tonin reuptake inhibitor (SSRI) antidepressants that polymorphism, designated 5HTTLPR, has been asso-
block its action. Most attention on the transporter has ciated with risk for depression in response to stress
focused on a variable number of tandem repeats [38]. It has also been associated with response to SSRI
(VNTR) polymorphism 5’ of transcription initiation. antidepressants [39], with poorer response associated
This variant has two predominant forms, long or with the s or short allele which results in a lower
short, based on the inclusion or exclusion of two transcription of the gene.
repeats. This polymorphism has been shown to influ- BDNF has also been extensively studied. It plays
ence transcription of the gene and was initially asso- an important role in brain development and in par-
ciated with the personality trait of neuroticism [36]. ticular in growth and development of catecholamine

203
Chapter 17: Genetics of bipolar disorder

systems implicated in bipolar disorder. BDNF expres- bipolar disorder and schizophrenia in a large Scot-
sion is increased by antidepressants and lithium, and tish family [56]. The breakpoint in DISC1 provided
this increase in expression is necessary for antidepres- an unequivocal causative mutation in this family.
sant-related behavioral responses in animal models This has emboldened biologists to conduct extensive
[40]. The gene contains a Val/Met substitution at studies of DISC1 function. Such studies are very
amino acid 66 that has been shown to affect transport labor intensive and the presence of a clear mutation
of the peptide within the cell and also its release. identifying the gene justified such an investment. As
Several studies have shown BDNF or its receptor a result, DISC1 is probably one of the biggest suc-
NTRK2 to be associated with bipolar disorder [41, cesses of the whole genetic mapping strategy. In
42]. Variation in BDNF and NTRK2 may also influ- addition to both disorders being present in the ori-
ence response to lithium [43]. ginal family, some subsequent studies have also
Several genes related to catecholamine biosyn- shown evidence of association to bipolar disorder
thesis have been examined. TH is the rate limiting providing further support that this gene plays a role
enzyme in catecholamine biosynthesis. This gene in both disorders [57].
contains a tetranucleotide repeat in intron 1 that Several genes have been identified by linkage in
has been shown to regulate transcription and to be bipolar families followed by fine mapping association
associated with bipolar disorder [44]. Catehol-O- studies. Linkage to 12q has been reported in an
methyltransferase (COMT) catabolizes norepineph- isolated population in Quebec [58]. Fine mapping
rine and dopamine thereby leading to their elimin- studies identified nonsynonymous SNPs in the gene
ation from the synapse. A Val/Met substitution in P2RX7 associated with bipolar disorder [59]. P2RX7
COMT has been shown to functionally affect enzyme is a purine receptor that is primarily expressed in
function and has been variably associated with risk immune cells. This has suggested a role for inflam-
for bipolar disorder [45, 46]. mation in depression. G protein receptor kinase 3
Some of the strongest evidence for association of (GRK3 or ADRBK2) is a member of a family of
specific genes to bipolar disorder comes from studies kinases that are involved in the homologous desen-
initially conducted in samples of patients with sitization of G protein coupled receptors. This gene
schizophrenia. D-amino acid oxidase activator was identified first by linkage to 22q in a set of
(DAOA), neuregulin (NRG1), and disrupted in extended bipolar families [50]. There appeared to
schizophrenia (DISC1) are all genes first identified be two distinct linkage peaks on 22q, one peak
by mapping studies in schizophrenia, then later mapped to within 20 kb of the GRK3 gene. Sequen-
shown to be associated to bipolar disorder. Linkage cing studies of the coding and promoter regions
studies of families with schizophrenia first identified revealed a series of SNPs in the proximal promoter.
a region on 13q [47]; subsequent fine mapping asso- One of these, showed significant association to bipo-
ciation studies in this region led to the identification lar disorder in two independent populations [60].
of the DAOA gene [48]. The same region on 13q has Functional studies showed that the promoter variant
also been shown to be linked to bipolar disorder in altered binding of Sp1 family transcription factors
several studies [49–51]. This led investigators to con- and increased expression of the gene in cultured
duct association studies of the DAOA gene in bipolar mouse cortical neurons [61]. GRK3 may interact
disorder and find that the same gene was likely with other possible genes for bipolar disorder. One
involved in both bipolar disorder and schizophrenia such candidate is the gene coding for the transcrip-
[52]. The neuregulin gene (NRG1) was similarly tion factor Sp4; involvement of this gene was first
identified first by linkage to 8p in Icelandic families suggested by anatomical and behavioral changes in a
with schizophrenia followed by fine mapping by knockdown mouse [62] that were suggestive of
association [53]. Subsequent studies demonstrated psychotic illness. A SNP in Sp4 was later shown to
association of the gene to bipolar disorder as well be associated with bipolar disorder and schizophre-
[54]. Neuregulin is a novel signaling molecule nia [63]. The promoter variant in the GRK3 gene
that interacts with its receptor ErbB4 to modulate impacts a transcription factor binding site that can
NMDA receptor mediated glutamate signaling [55]. be occupied by Sp1 or Sp4, suggesting a possible
DISC1 was identified by mapping the breakpoint of a interaction. Furthermore, the P2RX7 gene is phos-
1 : 11 balanced translocation segregating with both phorylated and desensitized by the GRK3 gene [64].

204
Chapter 17: Genetics of bipolar disorder

Table 17.4 Genome-wide association studies of bipolar disorder.

Study Type Sample Ethnicity Sample size Platform Results


(Case/
control)
WTCCC [65] Case WTCCC Caucasian 2000/3000 Affymetrix PALB2, p ¼ 108
control 500 K
Sklar et al. [66] Case STEP-BD Caucasian 1461/2008 Affymetrix MYO5B, p ¼ 107;
control 500 K TSPAN8, p ¼ 107;
CACNA1C, p ¼ 104
Baum et al. Pooling NIMH W1–4 þ Caucasian 461/563 þ Illumina DGKH; p ¼ 108
[69] Bonn/ 772/876 550 K
Mannheim 800/800
Scott et al. Pooling þ NIMH/Pritzker Caucasian 3683/14 507 Illumina 1p31.1, p ¼ 107;
[68] Ind GSK WTCCC 550 K 3p21, p ¼ 107;
genotyping MCTP1, p ¼ 107
Ferreira et al. Meta- WTCCC þ Caucasian 4387/6209 Affymetrix ANK3, p ¼ 109;
[179] analysis STEP-BD 500 K CACNA1AC, p ¼ 108
Smith et al. Case BiGS (NIMH Caucasian 1001/1033 Affymetrix Xq27.1, p ¼ 106;
[67] control W1–5) 6.0 (1 M NAP5, p ¼ 106
SNPs)
African- 345/670 DPY19L3, p ¼ 106;
American NTRK2, p ¼ 106
Hattori et al. Case Hondo area Japanese 107/107 Affymetrix KCNMB2, p ¼ 104;
[70] control 100 K AUTS2, p ¼ 104

Genome-wide association studies architecture and the role of common variants in


GWAS have only recently started in bipolar disorder these disorders.
and other psychiatric disorders. Published studies to Sklar et al. conducted a GWAS using the STEP-BD
date are summarized in Table 17.4. In 2007, the sample and a sample from University College, London
Wellcome Trust Case Control Consortium [66]. STEP-BD was a large naturalistic treatment study
(WTCCC) published the results of a large GWAS of in bipolar disorder. They reported evidence for myosin
seven disorders involving a total of 14 000 cases and 5B (MYO5B) and tetraspanin-8 (TSPAN8); however,
3000 controls [65]. Analysis of 500 000 SNPs identi- this was not replicated in an independent sample. They
fied a region on 16p12 showing the strongest evi- also identified the gene CACNA1C, a member of a
dence for association (p ¼ 108). Several genes are family of L-type calcium channels, as showing evi-
contained in this region including PALB2, which is a dence for association in both their sample and in the
binding partner for BRCA1, and dynactin 5 (DCTN5) aforementioned Wellcome Trust study. Such replica-
that is known to interact with DISC1. One of the most tion made this gene a good candidate even if it did not
striking observations of this study was the difference have the strongest signal in either study alone. Ferreira
in results for the seven disorders. Some disorders et al. reported a combined analysis of these 2 datasets
displayed a few clear strong peaks such as coronary plus additional samples for a total dataset of 4387 cases
artery disease (CAD) or Crohn’s disease, indicating a and 6209 controls. This analysis replicated the earlier
few common variants that made strong contribu- association to CACNA1C, however, the strongest
tions, while others such as bipolar disorder and result was at the gene ankyrin 3 (ANK3). ANK3 codes
hypertension lacked any such strong loci. This sug- for ankyrin G, which is a brain expressed cytoskeletal
gests differences in the underlying genetic protein involved in attaching sodium channels to the

205
Chapter 17: Genetics of bipolar disorder

cytoskeleton. ANK3 is expressed primarily at the axo- genotyped and analyzed. No genome-wide signifi-
nal initial segment where action potentials originate. cant findings were reported.
As some mood stabilizer medications, lamotrigine in
particular, block sodium channels as their likely mech-
anism of action, this gene is of particular interest. Mapping genes for subphenotypes
Smith et al. [67] have reported on a GWAS of the The family data described above are consistent with
NIMH Genetics Initiative for Bipolar Disorder different genes predisposing to different forms of
sample as part of the Bipolar Genome Study (BiGS). illness. A better understanding of the mapping from
In total, 1001 European-American (EA) cases and subphenotype to gene will be invaluable in under-
1033 EA controls, and 345 African-American (AA) standing how different genes result in different forms
cases and 670 AA controls were genotyped at 1-M or courses of illness. Early efforts to do this include a
SNPs. Though no single polymorphism showed study by Potash et al. [71] who conducted a linkage
genome-wide significance, suggestive evidence was study of psychotic bipolar disorder. These investiga-
obtained for several genes including: NAP5, tors identified two regions, 22q12 and 13q31, with
DPY19L3, and NTRK2. NTRK2 is an especially good suggestive evidence for linkage [72]. Both of these
candidate as it codes for Trkb, which is a tyrosine regions have figured prominently in studies of schizo-
kinase receptor for BDNF (see above) that has been phrenia. Turecki et al. [73] attempted to map genes
extensively studied in psychiatry. This study is one of for lithium responsive bipolar disorder. They identi-
the few to date to examine a non-Caucasian popula- fied a region on 15q that segregated with lithium
tion. This enabled a combined analysis and allowed responsive bipolar disorder.
identification of several genes (ROR1, RGS5, and Similar studies have been undertaken using
BTBD16) that showed evidence of association in both GWAS data. Greenwood et al. [74] examined asso-
populations, but to different alleles. This suggests that ciations to the phenotypes of irritable and euphoric
different mutations occurred in these genes in the mania in the BiGS dataset. Initially a case-case
histories of these two populations, and both predis- analysis was conducted comparing subjects with
pose to bipolar disorder. irritable to elated mania. A large region under a
Scott et al. [68] examined three populations for a broad peak on 13q was identified as showing asso-
combined sample of 3683 Caucasian cases and 14 507 ciation in this case-case analysis. Subsequent ana-
controls. A substantial portion of this sample over- lyses showed strong association between irritable
lapped with the Smith et al. study described above. mania and controls, but no association between
Though no single SNP achieved genome-wide signifi- euphoric patients and controls. This suggests that
cance suggestive evidence was obtained for nonsy- whatever gene or regulatory element resides in this
nonymous SNPs in ITIH1, GNL3, NEK4, and ITIH3. region predisposes specifically to irritable mania.
This study also provided support for the ANK3 gene Similar studies of suicide attempts by Willour
described above. et al. [75] in the BiGS and German datasets showed
Two other studies deserve mention. Baum et al. evidence of association to a region on 2p25 that
conducted a pooled GWAS of the NIMH wave 1–4 contains the gene ACP1. Age of onset, as a quanti-
sample with replication in a German Bonn/Mann- tative trait, was also examined in a combined BiGS/
heim sample [69]. In a pooling strategy, samples are German dataset, and evidence of association
pooled across many subjects and genotyped obtained to the oxysterol binding protein-like 1A
together. Signal strength for each of the two alleles (OSBPL1A) and neural cell adhesion molecule 2
is used to estimate control and case allele frequen- (NCAM2) [76]. Alcohol dependence has long been
cies. This is economical and faster, but has less known to have a four-fold greater prevalence in
ability to distinguish smaller differences in allele bipolar disorder as compared to subjects without a
frequencies. This study identified DGKH, a gene mood disorder. Nwulia et al. [77] examined the
involved in lithium’s mechanism of action as having BiGS GWAS dataset to identify genes associated
suggestive evidence of association. The only with the combined phenotype of bipolar disorder
reported study in an Asian population to date is and alcohol dependence. They found evidence
by Hattori et al. [70] in a Japanese population. In implicating the genes WDR59 and alpha N-catenin
total, 107 Japanese cases and 107 controls were (CTNNA2). Sleeplessness at initial mania has been

206
Chapter 17: Genetics of bipolar disorder

Figure 17.3 Partial overlap in familial


16 transmission of bipolar disorder and
schizophrenia. This figure illustrates the
14 familial risk for bipolar disorder,
schizophrenia, and depression in families
12
Morbid risk (%)

with probands of different diagnoses and


10 controls. Rates significantly different from
controls are indicated with an *. Rates
8 Bipolar of bipolar disorder are elevated above
control rates in families of probands with
6 Schizophrenia schizophrenia, and schizophrenia is
4 elevated in the families of those with
Unipolar bipolar disorder. This suggests a partial
2 overlap of causative genes. (Adapted
from [84]).
0
SZ Mania Depression Control

Proband diagnosis

hypothesized to distinguish different forms of ill- disorders. A number of family studies have
ness. Analysis of this phenotype in the BiGS dataset addressed this question. Figure 17.3 illustrates the
yielded three regions of interest: KIAA0143, a results of a study by Tsuang et al. of morbid risk
membrane bound protein; NAP5 which was also for bipolar disorder, schizophrenia and depression in
detected in the overall bipolar analysis; and MEP1B the families of probands with each of these three
a zinc metalloprotease [78]. disorders and controls [84]. In families of bipolar
Affective temperaments have also been examined probands, bipolar disorder displays the largest sig-
as subphenotypes that may index different forms of nificant increase in risk. However, the rate of schizo-
illness. Evans et al. [79] conducted a linkage study in phrenia is also elevated above the rate in control
bipolar families using cyclothymia as a quantitative probands. The mirror image of this is seen in fam-
trait and found evidence of linkage on 18p11. Cor- ilies of probands with schizophrenia. Schizophrenia
responding analysis of the BiGS GWAS dataset shows the largest significant increase in risk, but
found associations for hyperthymia on chromosome bipolar disorder also occurs at a rate greater than
12 and dysthymia on chromosome 4 [80]. Harm in that of control probands, but less than that of
avoidance as a dimension of temperament has been bipolar probands. Together these data suggest that
associated with the D3 dopamine receptor in a portion of the genetic vulnerability for these two
Han Chinese [81]. Another study of a South disorders is shared.
African population of bipolar individuals found the Twin data also support this idea of partial overlap
dopamine transporter to be associated with self- with schizophrenia. Cardno et al. [85] examined the
directedness and the serotonin transporter with medical records of the Maudsley Twin Psychosis
harm avoidance [82]. Registry and found that over time patients presented
differently and there was not strong longitudinal reli-
ability of the two diagnoses. They further found evi-
Overlap with other disorders dence for a significant portion of shared genetic
Since Emil Kraepelin distinguished manic depressive factors for these two disorders. Molecular mapping
illness from dementia praecox [83], the relationship studies have provided further support. Early linkage
between bipolar disorder and schizophrenia has been studies detected similar regions such as 13q and 22q
debated. Though this question has many aspects, in studies of both schizophrenia and bipolar families
genetics may serve as a proxy for differences in leading to the suggestion of shared genetic factors
biology and etiology. As the family studies described [86]. As mentioned above, a large family with a trans-
above began to elucidate the familiality of each dis- location breakpoint in the DISC1 was identified in a
order, investigators also questioned whether each of family segregating for both bipolar disorder and
these “breed true”. That is are different genes and schizophrenia [87]. Linkage was also reported in an
biological abnormalities operating in these two Italian population isolate between a region on 15q

207
Chapter 17: Genetics of bipolar disorder

and both bipolar disorder and schizophrenia [88]. interpreted in the light of the specific family being
Several specific genes have been reported to be asso- counseled. A high familial load suggests a greater
ciated to both disorders such as neuregulin (NRG1) risk as does bilineal transmission. A variety of
and D-amino acid oxidase activator (DAOA) [52, 54]. aspects of illness may be heritable as described
GWAS of both disorders may help illuminate the above. Such factors may be helpful in counseling
common and distinct genes, but such analyses are just families regarding possible course or severity. It is
beginning. important to emphasize the treatable nature of the
A similar overlap may exist with attention-deficit disorder, and be sensitive to the way in which each
hyperactivity disorder (ADHD). Some studies suggest family will hear the data in the light of their own
that probands with ADHD are more likely to have personal experience.
parents with bipolar disorder and parents with bipo- Though a number of gene findings have been
lar disorder are more likely to have children with replicated and likely reflect real effects, together these
ADHD [89–91]. Family studies further suggest that account for only a minority of the genetic variance.
bipolar disorder with ADHD may be a distinct form For this reason, DNA testing currently has limited
of illness [92]. The dopamine transporter which is the predictive and clinical value. However, it is likely in
target for stimulants has been implicated in both the near future that numerous new mutations will
disorders [93–95]. A linkage study of ADHD traits emerge from sequencing studies and these will form
in bipolar families identified a region on 10p14 sig- databases against which an individual’s sequence can
nificantly linked [96]. This might reflect a gene spe- be compared. Analyses of such data may provide a
cific for a form of illness that shares bipolar and useful aid in diagnosis, and in selection of optimal
ADHD traits. medication. As the heritability of the illness is less
In addition to other psychiatric disorders, there than 100% such tests will always be probabilistic in
is much literature to support genetic overlap nature. Yet in addition to aiding the ill patient, such
between bipolar disorder and other medical condi- tests may indicate risks in asymptomatic individuals.
tions. Such evidence for pleiotropy should not be Such information is at risk of being misused in a
surprising and may in fact offer clues as to the discriminatory fashion and requires careful ethical
disease mechanism involving such genes. Odergaard consideration. However, it might also afford an
et al. have reported an increase in the rate of opportunity to prevent or ameliorate the course of
migraine headache in patients with bipolar II dis- illness for some at risk individuals. Little data pres-
order as compared with the general population [97]. ently exists to guide such presymptomatic treatment,
They then went on to attempt to map genes respon- but it is clearly a needed area of research.
sible for the combined phenotype of bipolar and
migraine. Linkage studies of bipolar studies using
bipolar and migraine as the phenotype identified a Conclusions
region on 20p11 [98]. More recently, a similar Bipolar disorder is a highly heritable psychiatric
approach was applied to GWAS data and evidence disorder, and the strong role of genes is supported
of association to a combined phenotype was by family and twin studies. Mapping studies using
obtained for the gene KIAA0564 [99]. linkage and association methods have had modest
success to date despite difficulties in replication
between studies. Linkage studies have shown the
Genetic counseling best support for chromosomal regions: 6q, 8q, 9p,
Families frequently seek advice regarding genetic risk 13q, 14q, and 22q. Several candidate genes first
for bipolar disorder. It is highly heritable and most identified in studies of schizophrenia have shown
affected individuals have affected family members. reproducible association in bipolar disorder (such
Couples from families that have been severely as NRG1, DAOA, and DISC1). GWAS have been
affected may have reservations about having children successful in identifying a few genes with small
and require counseling regarding the risks. The effects on risk. ANK3 and CACNA1C, in particular,
family studies cited above provide a guide for esti- have shown reproducible evidence for association
mates of family risk. It is important to take a careful and may represent the genes with the strongest
family history as the risks from studies must be evidence to date for involvement in bipolar

208
Chapter 17: Genetics of bipolar disorder

disorder. The data overall suggest a high level of availability of complete genomic information. This,
both genic and allelic heterogeneity, as well as, a and large samples now being collected, may provide
complex mode of inheritance. The coming availability the datasets necessary to unravel the genetic complex-
of economical whole genome sequencing promises ities of this illness.

References 21. Kelsoe JR. J Affect Disord 2003;


73(1–2):183–197.
42. Sklar P, et al. Mol Psychiatry
2002;7(6):579–593.
1. Wozniak J, et al. Psychol Med
2009:1–10 [epub ahead of print]. 22. Perugi G, et al. J Affect Disord 43. Bremer T, et al. Mol Diagn Ther
2012;136(1–2):e41–49. 2007;11(3):161–170.
2. Spence MA, et al. Psychiatr Genet
1993;3:143. 23. Evans L, et al. J Affect Disord 44. Meloni R, et al. Lancet 1995;
2005;85(1–2):153–168. 345(8954):932.
3. Cox N, et al. J Psychiatr Res
1989;23:109–123. 24. Risch N, et al. Science 1996; 45. Lachman HM, et al. Psychiatr
273:1516–1517. Genet 1997;7:13–17.
4. Rice J, et al. Arch Gen Psychiatry
1987;44:441–447. 25. Egeland JA, et al. Nature 1987; 46. Biomed European Bipolar
325:783–787. Collaborative Group. Br J
5. Gershon ES, et al. Arch Gen
26. Kelsoe JR, et al. Nature 1989; Psychiatry 1997;170:526–528.
Psychiatry 1982;39:1157–1167.
6. McInnis MG, et al. Am J Hum 342:238–243. 47. Blouin JL, et al. Nat Genet
Genet 1993;53:385–390. 27. Baron M, et al. Nature 1987; 1998;20:70–73.

7. O’Donovan M, et al. Am J Med 326:289–292. 48. Chumakov I, et al. Proc Natl


Genet C Semin Med Genet 28. Baron M, et al. Nat Genet 1993; Acad Sci U S A 2002;99(21):
2003;123C(1):10–17. 3:49–55. 13675–13680.
8. Gershon ES, et al. Biol Psychiatry 29. Berrettini WH, et al. Proc Natl 49. Detera-Wadleigh SD, et al. Proc
1973;7(1):63–74. Acad Sci U S A 1994;91: Natl Acad Sci U S A 1999;96(10):
5918–5921. 5604–5609.
9. Negri F, et al. J Affect Disord
1979;1(4):247–253. 30. Freimer NB, et al. Nat Genet 50. Kelsoe JR, et al. Proc Natl Acad Sci
1996;12:436–441. U S A 2001;98(2):585–590.
10. Merikangas KR, et al. Psychol Med
1982;12(4):753–764. 31. Stine OC, et al. Am J Hum Genet 51. Shaw SH, et al. Mol Psychiatry
1995;57:1384–1394. 2003;8(5):558–564.
11. McMahon FJ, et al. Am J Hum
Genet 1995;56:1277–1286. 32. Badner JA, et al. Mol Psychiatry 52. Hattori E, et al. Am J Hum Genet
2002;7(4):405–411. 2003;72(5):1131–1140.
12. Kornberg JR, et al. J Affect Disord
2000;59(3):183–192. 33. Segurado R, et al. Am J Hum 53. Stefansson H, et al. Am J Hum
Genet 2003;73(1):49–62. Genet 2002;71(4):877–892.
13. Coryell W, et al. Br J Psychiatry
1984;145:49–54. 34. McQueen MB, et al. Am J Hum 54. Green EK, et al. Arch Gen
14. Potash JB, et al. Am J Psychiatry Genet 2005;77(4):582–595. Psychiatry 2005;62(6):642–648.
2001;158(8):1258–1264. 35. Caspi A, et al. Am J Psychiatry 55. Buonanno A. Brain Res Bull
15. Potash JB, et al. Am J Med Genet 2010;167(5):509–527. 2010;83(3–4):122–131.
2003;160:680–686. 36. Lesch KP, et al. Science 56. Blackwood DH, et al. Nat Genet
16. MacKinnon DF, et al. Am J 1996;274:1527–1531. 1996;12:427–430.
Psychiatry 2002;159(1):30–35. 37. Kunugi H, et al. Mol Psychiatry 57. Hodgkinson CA, et al. Am J Hum
17. MacKinnon DF, et al. Bipolar 1997;2:457–462. Genet 2004;75(5):862–872.
Disord 2005;7(5):441–448. 38. Caspi A, et al. Science 2003; 58. Morissette J, et al. Am J Med Genet
18. Schulze TG, et al. Arch Gen 301(5631):386–389. 1999;88:567–587.
Psychiatry 2006; 39. Smeraldi E, et al. Mol Psychiatry 59. Barden N, et al. Am J Med Genet
63(12):1368–1376. 1998;3(6):508–511. B Neuropsychiatr Genet 2006;
19. Saunders EH, et al. Am J Med 40. Nibuya M, et al. J Neurosci 141(4):374–382.
Genet B Neuropsychiatr Genet 1995;15(11):7539–7547. 60. Barrett TB, et al. Mol Psychiatry
2008;147B(1):18–26. 41. Fan J, et al. Novartis Found Symp 2003;8:546–557.
20. Potash JB, et al. Am J Psychiatry 2008;289:60–72; discussion 72–73, 61. Zhou X, et al. Biol Psychiatry
2007;164(8):1229–1237. 87–93. 2008;64(2):104–110.

209
Chapter 17: Genetics of bipolar disorder

62. Zhou X, et al. Mol Psychiatry Manchester, NH: Ayer and 105. Gershon ES, et al. J Psychiatr Res
2005;10(4):393–406. Co.; 1921. 1975;12:283–299.
63. Zhou X, et al. PLoS One 2009;4(4): 84. Tsuang MT, et al. Br J Psychiatry 106. James NM, et al. Br J Psychiatry
e5196. 1980;137:497–504. 1975;126:449–456.
64. Feng YH, et al. Am J Physiol Cell 85. Cardno AG, et al. Arch Gen 107. Johnson GFS, et al. Arch Gen
Physiol 2005;288(6):C1342–C1356. Psychiatry 1999;56(2):162–168. Psychiatry 1977;34:1074–1083.
65. Wellcome Trust Case Control. 86. Berrettini WH. Biol Psychiatry 108. Petterson U. Acta Psychiatr Scand
Nature 2007;447(7145):661–678. 2000;48(6):531–538. 1977;269:1–93.
66. Sklar P, et al. Mol Psychiatry 87. Blackwood DH, et al. Am J Hum 109. Smeraldi E, et al. Acta Psychiatr
2008;13(6):558–569. Genet 2001;69(2):428–433. Scand 1977;56:382–398.
67. Smith EN, et al. Mol Psychiatry 88. Vazza G, et al. Mol Psychiatry 110. Angst J, et al. Hum Genet 1980;
2009;14(8):755–763. 2007;12(1):87–93. 55:237–254.
68. Scott LJ, et al. Proc Natl Acad Sci 89. Kent L, et al. J Affect Disord 111. Dunner DL, et al. Am Coll
U S A 2009;106(18):7501–6. 2003;73(3):211–221. Neuropsychopharm 1980; Abs.
69. Baum AE, et al. Mol Psychiatry 90. Sachs GS, et al. Am J Psychiatry 112. Taylor MA, et al. J Affect Disord
2008;13(2):197–207. 2000;157:466–468. 1980;2:95–109.
70. Hattori E, et al. Am J Med Genet 91. Faraone SV, et al. J Am Acad Child 113. Fieve RR, et al. J Psychiatr Res
B Neuropsychiatr Genet Adolesc Psychiatry 1997; 1984;18:425–445.
2009;150B(8):1110–1117. 36:1378–1387. 114. Tsuang MT, et al. J Psychiatr Res
71. Potash JB, et al. Am J Psychiatry 92. Faraone SV, et al. J Am Acad Child 1985;19:23–29.
2003;60(4):680–686. Adolesc Psychiatry 1997;36(10): 115. Sadovnick AD, et al. Am J Med
72. Lander E, et al. Nat Genet 1378–1387; discussion Genet 1994;54(2):132–140.
1995;11:241–247. 1387–1390.
116. Birmaher B, et al. Arch
73. Turecki G, et al. Mol Psychiatry 93. Waldman ID, et al. Am J Hum Gen Psychiatry 2009;
2001;6(5):570–578. Genet 1998;63:1767–1776. 66(3):287–296.
74. Greenwood T, Jr. In World 94. Cook EH, Jr, et al. Am J Hum 117. Luxenberger H. Zentralblatt fur
Congress on Psychiatric Genetics. Genet 1995;56:993–998. diagensamte Neurologie und
San Diego: WCPG; 2009. 95. Greenwood TA, et al. Mol Psychiatrie 1930;14:56–57.
75. Willour VlS, et al. Mol Psychiatry Psychiatry 2006;11(2):125–133. 118. Rosanoff AJ, et al. Am J Psychiatry
2012;17(4):433–444. 96. Joo EJ, et al. Am J Med Genet 1935;91:725–762.
76. Belmonte PJ, et al. In World B Neuropsychiatr Genet 2010;
119. Slater E. Psychotic and Neurotic
Congress on Psychiatric Genetics. 153B(1):260–268.
Illness in Twins. London: Medical
San Diego: WCPG; 2009. 97. Oedegaard KJ, et al. J Affect Disord Research Council; 1953.
77. Nwulia EH, et al. In World 2005;84(2–3):233–242.
120. Kallman F, et al. In Zubin J et al.
Congress on Psychiatric Genetics. 98. Oedegaard KJ, et al. J Affect Disord (eds.). Depression. New York:
San Diego: WCPG; 2009. 2010;122(1–2):14–26. Grune and Stratton; 1954.
78. Coleman VK, et al. In World 99. Oedegaard KG, et al. In World 121. Harvald B, et al. Genetics and the
Congress on Psychiatric Genetics. Congress on Psychiatric Genetics. Epidemiology of Chronic Diseases.
San Diego: WCPG; 2009. San Diego: WCPG; 2009. Washington, DC: US Department
79. Evans LM, et al. Am J Med Genet 100. Perris C. Acta Psychiatr Scand of Health; 1965.
B Neuropsychiatr Genet 2008; 1966;194:102–117. 122. Allen MG, et al. Am J Psychiatry
147(3):326–332. 1974;131:1234–1239.
101. Winokur G, et al. In Worris J
80. Greenwood T, et al. In World (ed.). Recent Advances in 123. Bertelsen A. Origin, Prevention
Congress on Psychiatric Genetics. Biological Psychiatry. New York: and Treatment of Affective
San Diego: WCPG, 2009. Plenum; 1967. Disorder. London: Academic
81. Lin CI, et al. Psychiatry Res 102. Mendlewicz J, et al. Am J Hum Press; 1979.
2010;177(3):364–366. Genet 1974;26:692–701. 124. Kendler KS, et al.
82. Savitz J, et al. Genes Brain Behav 103. Goetzl U, et al. Arch Gen Arch Gen Psychiatry
2008;7(8):869–876. Psychiatry 1974;31:665–672. 1993;50:699–700.
83. Krapelin E. Manic-Depressive 104. Helzer JE, et al. Arch Gen 125. Collier DA, et al. Neuroreport
Insanity and Paranoia. Psychiatry 1974;31:73–77. 1996;7:1675–1679.
210
Chapter 17: Genetics of bipolar disorder

126. Collier DA, et al. Mol Psychiatry 150. Smyth C, et al. Genomics 172. Martucci L, et al. Schizophr Res
1996;1:453–460. 1997;39:271–278. 2006;84(2–3):214–221.
127. Kelsoe JR, et al. Am J Med Genet 151. Barrett TB, et al. Psychiatr Genet 173. Mick E, et al. Am J Med Genet
1996;67(2):215–217. 2007;17(6):315–322. B Neuropsychiatr Genet 2008;
128. Kunugi H, et al. Lancet 1996;347 152. Shink E, et al. Mol Psychiatry 147B(7):1182–1185.
(9011):1340. 2005;10(6):545–552. 174. Horschitz S, et al. Mol Psychiatry
129. Ogilvie AD, et al. Lancet 1996;347 153. Lucae S, et al. Hum Mol Genet 2005;10(12):1104–1109.
(9003):731–733. 2006;15(16):2438–2445. 175. Keikhaee MR, et al. Am J Med
130. Stober G, et al. Lancet 1996;347 154. Grigoroiu-Serbanescu M, et al. Genet B Neuropsychiatr Genet
(9011):1340–1341. Am J Med Genet B Neuropsychiatr 2005;135B(1):47–49.
131. Bellivier F, et al. Biol Psychiatry Genet 2009;150B(7):1017–1021. 176. Kelsoe JR, et al. Am J Med Genet
1997;41:750–752. 155. McQuillin A, et al. Mol Psychiatry 1996;67(6):533–540.
132. Furlong RA, et al. Am J Med Genet 2009;14(6):614–620. 177. Greenwood TA, et al. Am J Med
1998;81:58–63. 156. Zhang X, et al. Neuron 2005;45 Genet 2001;105(2):145–151.
133. Gutierrez B, et al. Biol Psychiatry (1):11–16. 178. Silberberg G, et al. Pharmacogenet
1998;43:843–847. 157. Chen C, et al. Bipolar Disord Genomics 2008;18(5):403–412.
134. Liu W, et al. Chung Hua I. Hsueh 2008;10(7):816–821. 179. Ferreira MA, et al. Nat Genet
I. Chuan Hsueh Tsa Chih 158. Pickard BS, et al. Mol Psychiatry 2008;40(9):1056–1058.
1998;15:345–348. 2006;11(9):847–857. 180. Nurnberger JI, Jr, et al. Am J Med
135. Seretti A, et al. Mol Psychiatry 159. Maeda K, et al. Biol Psychiatry Genet 1997;74:227–237.
1999;4:280–283. 2006;60(9):929–935. 181. Detera-Wadleigh SD, et al. Am
136. Oliveira JR, et al. Mol Psychiatry
160. Hodgkinson CA, et al. Am J Hum J Med Genet 1997;74:254–262.
2000;5(4):348–349.
Genet 2004;75(5):862–872.
137. Saleem Q, et al. Am J Med Genet 182. Rice JP, et al. Am J Med Genet
161. Macgregor S, et al. Mol Psychiatry 1997;74:247–253.
2000;96(2):170–172.
2004;9(12):1083–1090.
138. Mundo E, et al. Arch Gen 183. Edenberg HJ, et al. Am J Med
Psychiatry 2001;58(6):539–544. 162. Devon RS, et al. Psychiatr Genet Genet 1997;74:238–246.
2001;11(2):71–78.
139. Lotrich FE, et al. Psychiatr Genet 184. Kelsoe JR, et al. Am J Med Genet
2004;14(3):121–129. 163. Maier W. Eur Arch Psychiatry Clin 1998;81:461–462.
Neurosci 2008;258(Suppl 2):
140. Schumacher J, et al. Biol 37–40. 185. Detera-Wadleigh SD, et al. Proc
Psychiatry 2005;58(4):307–314. Natl Acad Sci U S A 1999;96
164. Georgieva L, et al. Biol Psychiatry (10):5604–5609.
141. Kirov G, et al. Mol Psychiatry 2008;64(5):419–427.
1999;4(6):558–565. 186. Blackwood DH, et al. Nat Genet
165. Thomson PA, et al. Mol Psychiatry 1996;12(4):427–430.
142. Shifman S, et al. Am J Med Genet 2007;12(1):94–104.
B Neuropsychiatr Genet 2004;128 187. Straub RE, et al. Nat Genet 1994;8
(1):61–64. 166. Benedetti F, et al. Am J Med Genet (3):291–296.
B Neuropsychiatr Genet 2003;
143. Prata DP, et al. Psychiatr Genet 123B(1):23–26. 188. Pekkarinen P, et al. Genome Res
2006;16(6):229–230. 1995;5(2):105–115.
167. Serretti A, et al. Am J Med Genet
144. Leboyer M, et al. Lancet B Neuropsychiatr Genet 2003; 189. Dick DM, et al. Am J Hum Genet
1990;335:1219–1219. 121B(1):35–38. 2003;73(1):107–114.
145. Gill M, et al. J Psychiatr Res 168. Nievergelt CM, et al. Am J Med 190. Cichon S, et al. Mol Psychiatry
1991;25:179–184. Genet B Neuropsychiatr Genet 2001;6(3):342–349.
146. Byerley W, et al. Hum Hered 2006;141B(3):234–241. 191. McInnis MG, et al. Mol Psychiatry
1992;42:259–263. 169. Vuoristo JT, et al. Mol Psychiatry 2003;8(3):288–298.
147. Inayama Y, et al. Am J Med Genet 2000;5(5):495–501. 192. Adams LJ, et al. Am J Hum Genet
1993;48:87–89. 170. Avramopoulos D, et al. Genet Med 1998;62:1084–1091.
148. Korner J, et al. Psychiatr Genet 2007;9(11):745–751. 193. Coon H, et al. Am J Hum Genet
1994;4:167–175. 171. Szczepankiewicz A, et al. World 1993;52:1234–1249.
149. Smyth C, et al. Am J Psychiatry J Biol Psychiatry 2009;10(4 Pt 2): 194. Badenhop RF, et al. Mol
1996;153:271–274. 469–473. Psychiatry 2002;7(8):851–859.

211
The genetics of major depression
Chapter

18 James B. Potash

Phenotype definition decades later. This study found a higher incidence of


unipolar depressive illness in the families of bipolar
The word “depressed” refers to a symptom, not a probands, but no excess of bipolar disorder in the
diagnosis. A depressed mood generally means a low families of unipolar probands (though subsequent
mood, and that can occur in the setting of many studies have found a slight excess) [4, 5]. This suggests
differing diagnoses. The DSM-IV major depressive both overlapping bipolar/unipolar genes and other
episode diagnosis requires a total of five symptoms, genes that uniquely confer risk for unipolar depres-
one of which must be depressed mood or anhedonia sive illness.
(loss of interest or pleasure), occurring together over A key element of the debate over depression clas-
at least two weeks, to meet criteria for a major depres- sification has focused on the issue of etiology. What is
sive episode. Further, there must be significant dis- the cause of depression? One of the schemes proposed
tress or impairment associated with these symptoms. divided depression into endogenous and exogenous,
It is not uncommon for patients with depressive meaning that some depressions were caused by
episodes to have occasional manic, hypomanic, or factors inside the brain of the ill individual, deriving
mixed symptoms. If these symptoms are few or from genetic predisposition, and some were caused by
fleeting then the patient would still meet criteria for factors in the person’s environment. A portion of the
major depressive disorder (MDD). However, if they etiology of MDD is clearly genetic, and that forms the
reach the threshold for a hypomanic, manic, or mixed basis for the bulk of the discussion in this chapter.
episode, then the patient’s diagnosis becomes bipolar However, given that MDD is only partly inherited,
disorder. The Kraepelinian conceptualization of variations in the makeup of people’s genes cannot
“manic-depressive illness” included all recurrent fully explain the variability in susceptibility to this
severe disorders of mood: those characterized by illness. There are at least two other major kinds of
depression and mania, but also those characterized explanations for this variation. One of these is that
by depression alone [1]. In the 1950s, Karl Kleist environmental factors play a significant role in trig-
proposed a nosological division of mood disorders gering illness, and another is that epigenetic factors
into unipolar and bipolar types, and Karl Leonhard’s are involved. Epigenetic factors are modifications of
1959 classification of psychoses made a clear distinc- genes and chromosomes that can influence the degree
tion between bipolar illness and recurrent unipolar to which genes are turned on or off. These two factors
depression. Carlo Perris reported family findings in a may interrelate as the environment can cause epigen-
1966 monograph suggesting that unipolar and bipolar etic changes.
illness were genetically separate [2]. However, not
long after, Reich et al. [3] reported from their study
of manic probands that there is an excess of unipolar Genetic epidemiology
depressive illness in the families of individuals with The first question in considering the genetics of
bipolar illness. This has been confirmed in several depression is whether or not the illness is heritable.
studies including a family study of over 500 unipolar Only when heritability has been firmly established
and bipolar probands that was completed a couple of does it make sense to begin looking for the particular

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

212
Neurotrophic factors
BDNF EGFR
NT-3
NT-4
IGF-1
TRK EGFR AMPAR LICAM NMDAR EPHB2 DOPAMINE RECEPTORS GPCR

AKAP9
ADENYLATE CYALASE
KALRN
INTEGRIN CAMKK2 SNAP91 CDK5RAP3
RAC
MAP3K71P2

AB12 FLJ13386 CDKS cAMP


SPTBN1
NEF3 NCK
SOS1

SPTAN1 SNX6 PDE4B

TNIK
NDEL1 GNB1
SH3BP5 TRAF3IP1
RAS CDCSL
PAFAH1B1
ANK2
CLU

JNK2 JNK3 JNK1 DST

SEC3L1 DISC1 SRGAP3

TRIM9
MACF1
RHO GTPASES
TRI0
SNAP25

DMO
RHO GTPASES

Figure 7.1 Protein interactome simulated by pathway analyses. Iterative yeast-two hybrid screens, combined with detailed pathway, and
functional analyses suggest DISC1 interacts with proteins involved in key processes involved in neurodevelopment. Deficits in these processes
may underlie decreased dendritic branching, arborisations and neuropil size observed pathologically. (Adapted from [16], with permission.)
(a) Embryo Figure 7.2 Brain region specific gene
manipulations. (a) Schematic
representation of bilateral in utero injection
of constructs followed by their
E14 incorporation by electroporation into
Injection progenitor cells in the ventricular zone at
– + embryonic day 14 (E14). Migrating cells
+ + with green fluorescent protein (GFP) are
visualized at E18 after injection of a GFP
expression construct. (Adapted from [46],
with permission.) (b) Stereotaxic
coordinates and actual injection of
lentivirus-based enhanced GFP (EGFP).
– – Stereotaxic coordinates were determined
from the rat brain in stereotaxic
coordinates. The coordination in Sprague–
E18 Dawley rat at postnatal day 21 was
stereotaxically injected with lentivirus
+ – containing EGFP at the coordinates of
AP = þ2.2; ML = þ0.9; DV = þ2.0, þ1.5,
þ1.0 from the bregma. The crosses
indicate injected sites. Cx, cerebral cortex;
Hip, hippocampus; Th, thalamus; Str,
striatum; Amy, amygdala; Hypo,
hypothalamus; VTA, ventral tegmental
area; P, pontine tegmentum. (Adapted
from [47], with permission.)

(b) Adult brain


medial lateral
0 1 2 3 4 mm

Cx
Cx Hip

Str Th
Amy

AP +2.2 mm AP –1.34 mm
anterior bregma posterior
3 2 1 0 –1 –2 –3 –4 –5 –6 –7 –8 mm
0 dorsal
Cx 1
Hip 2
Str 3
Th
200 μm 4
VTA
Hypo 5
6 ventral
Figure 7.3 Dynamic changes in the adolescent brain. Disturbances generated by susceptibility genes and environmental insults during early
development (three stars on the left-hand side) may impair some of the crucial processes in early development, including progenitor cell
proliferation, neuronal migration and dendritic arborization and outgrowth. Independent of such initial risks/insults, intrinsic disease-associated
factors might also directly affect postnatal brain maturation (two central stars) contributing to the emergence of schizophrenia (SZ) in young
adulthood. (Adapted from [7], with permission.)
(a) Human Mouse

Control SZ 5

Ventricle volume (mm3)


4

0
WT Tg

100 WT
(b)
WT Tg Tg
80

Cells/section
Control SZ
60
PV
40

20
CB
0
PV CB
Figure 7.4 Anatomical and histological abnormalities in schizophrenia patients and in mouse models. (a) Enlarged lateral ventricles as
detected by in vivo magnetic resonance imaging (MRI) in schizophrenia patients (left) and in a DISC1 mouse model (right). (b) Decreased
parvalbumin (PV) expression in the prefrontal cortex of schizophrenia patients (mRNA, left) and a DISC1 mouse model (right). (Adapted
from [60], with permission.)
(a) Histone tail

DNA
H 2B

H4
H3

H2A

A Acetylation
Histone
M Methylation

(b) Active P Phosphorylation


Histones Histone Transcription factor Permissive
tail +
A A A M M
P A A

A Co-Act
A M A A
M
DNA Basal transcription
complex

Inactive
M M
P P
Repressed
Rep Rep
M Rep
M M M Rep A M M
M
A Rep
M
M
?
M

Rep M
M M
M
M M

Figure 8.1 Chromatin remodeling. (a) Picture of a nucleosome showing a DNA strand wrapped around a histone octamer composed
of two copies each of the histones H2A, H2B, H3, and H4. The amino (N) termini of the histones face outward from the nucleosome complex.
(b) Chromatin can be conceptualized as existing in two primary structural states: as active, or open, euchromatin (top left) in which histone
acetylation (A) is associated with opening the nucleosome to allow binding of the basal transcriptional complex and other activators of
transcription; or as inactive, or condensed, heterochromatin where all gene activity is permanently silenced (bottom left). In reality, chromatin
exists in a continuum of several functional states (active; permissive [top right]; repressed [bottom right]; and inactive). Enrichment of
histone modifications such as acetylation and methylation (M) at histone N-terminal tails and related binding of transcription factors and
co-activators (Co-Act) or repressors (Rep) to chromatin modulates the transcriptional state of the nucleosome. Recent evidence suggests that
inactivated chromatin may in some cases be subject to reactivation in adult nerve cells, although this remains uncertain.
(c)
Demethylation Demethylation

HDM HMT HDM HMT

Methylation Methylation
(activating) (repressing)

M M
M M
H3 K4 K27 S28 M
K9 K23 K36
S10 K14 K18 P K79
A A
P A A Histone
tail
Acetylation Phosphorylation
(activating) (activating)

HAT HDAC PK PP

Deacetylation Dephosphorylation

Figure 8.1 (c) Summary of common covalent modifications of H3, which include acetylation, methylation and phosphorylation (P) at several
amino acid residues. H3 phosphoacetylation commonly involves phosphorylation of S10 and acetylation of K14. Acetylation is catalysed by
histone acetyltransferases (HATs) and reversed by histone deacetylases (HDACs); lysine methylation (which can be either activating or
repressing) is catalysed by histone methyltransferases (HMTs) and reversed by histone demethylases (HDMs); and phosphorylation is catalysed
by protein kinases (PK) and reversed by protein phosphatases (PP), which have not yet been identified with certainty. K, lysine residue;
S, serine residue. (From [3].)
Figure 8.2 Regulation of chromatin structure by drugs of abuse. Drug-induced signaling events are depicted for cocaine and amphetamine.
Cocaine and amphetamine can increase cyclic adenosine monophosphate (cAMP) levels in striatum, which activates protein kinase A (PKA)
and leads to phosphorylation of its targets. This includes the cAMP response element binding protein (CREB), the phosphorylation of
which induces its association with the histone acetyltransferase, CREB binding protein (CBP) to acetylate histones and facilitate gene activation.
This is known to occur on many genes including fosB and c-fos in response to psychostimulant exposure. ΔFosB is also upregulated by chronic
psychostimulant treatments, and is known to activate certain genes (e.g. cdk5) and represses others (e.g. c-fos) where it recruits HDAC1.
This repression of c-fos also involves increased repressive histone methylation, which is thought to occur via the induction of specific histone
methyltransferases. It is not yet known how cocaine regulates histone demethylases (HDM) or DNA methyltransferases (DNMTs). Cocaine also
activates the mitogen activated protein kinase (MAPK) cascade, which through MSK1 can phosphorylate CREB and histone H3 at serine 10.
Cocaine promotes H3 phosphorylation via a distinct pathway, whereby PKA activates protein phosphatase 2A, leading to the
dephosphorylation of serine 97 of DARPP32. This causes DARPP32 to accumulate in the nucleus and inhibit protein phosphatase 1 (PP1) which
normally dephosphorylates H3. Chronic exposure to psychostimulants is also known to increase glutamatergic stignaling from the prefrontal
cortex to the NAc. Glutamatergic signaling elevates Ca2þ levels in NAc synapses and activates CaMK (calcium/calmodulin protein kinases)
signaling, which, in addition to phosphorylating CREB, also phosphorylates HDAC5. This results in nuclear export of HDAC5 and increased
histone acetylation on its target genes (e.g. NK1R [NK1 or substance P receptor). (From [3].)
(a) Figure 8.3 Regulation of the bdnf gene
HDAC5 HDAC5 by social defeat. (a) In the absence of
stress, the chromatin state of brain-
A A A derived neurotrophic factor (Bdnf) is at
a basal level, characterized by moderate
+ levels of histone H3 acetylation and
Bdnf expression
virtually no H3K27 dimethylation. In this
state, histone deacetylase 5 (HDAC5)
A A might repress unnecessary activation of
M
M BDNF and maintain a chromatin balance.
(b) Chronic defeat stress induces the
specific and prolonged dimethylation
(b) HDAC5 HDAC5 of histone H3K27. This induces a more
M M M M M “closed” chromatin state at bdnf
A M M A M M A M promoters P3 and P4, and a
corresponding repression of Bdnf

transcripts III and IV expression. H3
Bdnf expression acetylation and HDAC5 regulation are not
affected after chronic defeat stress alone,
corroborating the idea that the main
A A
repressive marker after chronic stress is
M M M M M
M M M M M histone methylation. (c) Chronic
imipramine (antidepressant) treatment
(c) after defeat stress downregulates Hdac5
M M M M
M M M M expression and increases H3 acetylation,
A A A A A with little if any change in H3K27
dimethylation. Imipramine-dependent H3
+ hyperacetylation at the bdnf promoters
Bdnf expression
P3 and P4 allows partial “reopening” of
the repressed chromatin state caused by
M
A A A A A M H3–K27 dimethylation defeat stress, and results in transcriptional
M M M M reactivation of the bdnf gene. K, lysine
M M M M A H3 acetylation residue. (From [3].)

Figure 9.6 Amygdala activity related to masked and unmasked


fear processing in healthy subjects. Enlarged views of the right
amygdala illustrating: (1) the dorsal amygdalar cluster from the
nonmasked fear (F-N) comparison (coronal view at Y = 8 [A] and
axial view at Z = 16 [B]); and (2) the basolateral amygdalar cluster
from the correlation of masked fear-induced activity (FN-NN) with
trait anxiety (coronal view at Y = 8 [C] and axial view at Z = 28
[D]). The color bar indicates the significance, with lighter colors
Non-masked Masked fear indicating a greater difference between the respective fearful and
FEAR (F-N) (FN-NN) vs. STAI-T neutral conditions. (From [200], with permission from Elsevier.)

T
value
6
5 5
4 4
3 3
2 2
1 1
0 0
Figure 9.7 Amygdala activation in response to fearful faces is
modulated by catechol-O-methyltransferase (COMT) genotype.
Whole brain voxel-wise t-map (uncorrected, P > 0.001) of a single
panic disorder patient with a genotype 472G/A (i.e. val-met Figure 9.8 Serotonin 1A receptor genotype moderates fear
heterozygous) comparing activation for the fearful versus the processing in panic disorder patients. Random effects statistical
no-face condition. Note that there is significantly greater amygdala parametric map for the fearful versus neutral faces contrast overlaid
activation (center of cross hair) in the fearful condition. This increase on a three-dimensional canonical Montreal Neurological Institute
was not observed among patients homozygous for the met brain showing right-lateralized activity differences in the prefrontal
encoding allele. The transverse plane is the original one; the coronal cortex between the two patient groups (5HT1A1019GG versus
and sagittal planes are planar reconstructions orthogonal to the CC/CG; p < 0.001, uncorrected). Patients homozygotes for the g
original image. Reader’s right is subject’s right. Emotional face stimuli alleles had reduced activation in orbitofrontal, ventromedial
and no-face control stimuli were controlled for dynamics and prefrontal, and cingulate cortices when viewing unmasked fearful
luminance. (From [81], with permission from Cambridge faces, as compared to patients with the CG or CC genotype.
University Press.) (From [213], with permission from Cambridge University Press.)
STin2 IIe425Val
SNPs VNTR
rs25531,
67 Alternate
(a) Gene rs25532 2 34 5 8 9
1011
1A 1C 1B 12 13 14 polyadenylation sites
5’

3’

5HTTLPR
(LA,LG,SA,SG) G56A
Alternative
splicing
34

(b) SERT SNPs: 8 Extracellular


11 12
6 7
1. T4A1 9. P339L TMS
9 1013
2. G56A2 10. L362M 2 14 15 Intracellular
16
3. S214S 11. L383L
COOH
NH2
4. E215K1 12. A419A 1
5. H235H 13. 2X1425V(/L)
6. L255M 14. T439T
7. S293F 15. K605N1 or 5HT transport in transfected cells
1No response to PKG/p38 MAPK activation
8. G308G 16. P621S1 2No response to 8BrcGMP

Figure 11.2 Human SERT gene organization, with multiple functional variants.

(a)
A189V

Variations in OCD/ R13C A148insGPAGA T523K P606T K910R


TTm patients

Guanylate-Kinase-associated protein (GKAP) domain

Variations in
T156M A189V
controls subjects

(b) Human MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPERSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP


Rhesus MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPERSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Cow MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPERSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Horse MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGPAPGTGTAPERSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Mouse MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGP--GSGAAPEARSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP
Rat MRGYHGDRGSHPRPARFADCQ TLPYQRGPAGAGPGP--GSGAAPEARSE SHSLBAPGKRDY OLPLLAAPAAVSGRP VPPRASPKPPT LLEPKEEXKVPPPIP

R13C A148insGPAGA T156M A189V T523K P606T K910R

Figure 11.3 Identified rare nonsynonymous polymorphisms in synapse-associated protein 90/postsynaptic density-95-associated protein 3
(SAPAP3). (a) Schematic of SAPAP3, which consists of 10 coding exons (boxes). Seven rare changes were identified in trichotillomania and
obsessive–compulsive disorder (OCD) patients (a), but only two in controls (b). Most mutations fell into exon 1; however, three changes
affected the conserved guanylate-kinase-associated protein (GKAP) domain. (b) SAPAP3 is highly conserved between species ( 97% identical
amino acids between human and mouse). Accordingly the identified rare changes affected conserved residues.
Chr 1 Chr 2 Chr 3 Chr 4 Chr 5 Chr 6 Chr 7 Chr 8 Chr 9 Chr 10 Chr 11 Chr X
1.1 5.1 8.1
3.1 10.1 X.1
9.1
7.1

5.2 11.1
2.1
10.2
3.2 5.3
3.3 7.2 9.2 11.2 X.2
4.1 11.3
4.2 10.3
4.3 6.1 7.4 7.3
7.5 X.3
7.6 7.7 9.3 X.4
3.4 7.8 9.4
6.2 9.5
7.10 7.9 X.5
1.2 7.11 X.6
2.2 2.3
2A 3.5
3.6 4.4 12.1

1.3 16.1 15.1


2.5 15.2
17.2 12.2
17.1 15.3 13.1
20.1 19.1 16.2
17.3 16.3
21.1 20.2 14.1
17.4
17.5
22.1 16.4
21.2 15.4
22.2

Chr 22 Chr 21 Chr 20 Chr 19 Chr 18 Chr 17 Chr 16 Chr 15 Chr 14 Chr 13 Chr 12 Chr Y

ID Feature Position Refs ID Feature Position Refs ID Feature Position Refs


1.1 Loss 1p36 3 7.4 RELN 7q22 117–121 15.3 Gain 15q11–15q13 4
1.2 Linkage 1q21–1q23 116 7.5 MET 7q31 70,123 15.4 Linkage 15q22–15q26 5
1.3 DISCI 1q42 122 7.6 Loss 7q31 5 16.1 TSC2 16p13 110
2.1 NRXNI 2p16 5,34 7.7 Linkage 7q32–7q34 15 16.2 Loss 16p11 4,20,35,44
2.2 Loss 2p24 4 7.8 CADPS2 7q31 5,45 16.3 Gain 16p11 20,35,44
2.3 Linkage 2p24–2q31 57,58,TT2 7.9 Linkage 7q34–7q36 14,52,68 16.4 Loss 16q21 5
2.4 SLC 25412 2p24 124–126 7.10 CNTNAP2 7q35–7q36 37–40 17.1 Loss 17p12 5
2.5 Loss 2q37 4 7.11 EN2 7q36 129,130 17.2 Gain 17p12 107
3.1 OTXR 3p25 127,128 8.1 Gain 8p23 5 17.3 SLC6A4 17q11 131–134
3.2 Loss 3p14 4 9.1 Linkage 9p24 5 17.4 Linkage 17q11–17q21 51,54,135
3.3 Gain 3p14 4 9.2 Loss 9q12 5 17.5 ITGB3 17q21 136,137
3.4 Linkage 3q22 15 9.3 Linkage 9q33 5 19.1 Linkage 19p13 140
3.5 Linkage 3q25–3q27 138,139 9.4 Linkage 9q34 15 20.1 Loss 20p13 5
3.6 Loss 3q27–3q28 3 9.5 TSCI 9q34 110 20.2 Loss 20p13 5
4.1 Loss 4q21 3 10.1 Loss 10q14–10p15 4 21.1 Linkage 21q11 55
4.2 Loss 4q21–23 3 10.2 Gain 10q11–10q21 4 21.2 Loss 21q22 3
4.3 Linkage 4q22–4q2S 15 10.3 PTEN 10q23 141 22.1 Loss 22q13 4
4.4 Loss 4q35 5 11.1 Linkage 11q12–11p13 5 22.2 SHANK3 22q13 21,22,142,143
5.1 Linkage 5p15 5 11.2 DHCR7 11q13 108 X.1 NLGN4X Xp22 28
5.2 Linkage 5p13–5q11 140 11.3 Linkage 11q13–11q14 15 X.2 NLGN3 Xq13 28
5.3 Linkage 5q12 5 12.1 CACNAIC 12p13 24 X.3 Linkage Xq21–Xq25 140
6.1 GRK2 6q21 144–146 12.2 AVPRIA 12p14–12q15 147 X.4 Gain Xq24 3
6.2 AHII 6q23 106 13.1 Gain 13p14 5 X.5 FMRI Xq27 105,148
7.1 Loss 7p21 4 14.1 Linkage 14q23 149 X.6 MECP2 Xq28 109
7.2 Loss 7q11 3 15.1 UBE3A 15q11 102,103
7.3 Linkage 7q22–7q32 52,111–113 15.2 GABRB3 15q12 25,114,115

Figure 16.1 Loci implicated in autism spectrum disorder (ASD) etiology. Entries in the ID column of the table map are entries to the
ideograms of individual chromosomes. Red and yellow bars correspond to de novo losses and gains, respectively, which are observed in cases
but not in controls. Green bars correspond to genes that are observed to modulate ASD risk (either through a rare syndrome or through
genetic association): light green and dark green bars represent promising or probable candidate genes, respectively, as defined in the table
map. Regions shaded in purple correspond to linkage peaks. Only human data were considered in the assembly of the table. AHI1, Abelson
helper integration site 1; AVPR1A, arginine vasopressin receptor 1A; CACNA1C, calcium channel voltage-dependent L type a 1C subunit;
CADPS2, Ca2þ-dependent activator protein for secretion 2; CNTNAP2, contactin associated protein-like 2; DHCR7, 7-dehydrocholesterol
reductase; DISC1, disrupted-in-schizophrenia 1; EN2, engrailed homeobox 2; FMR1, fragile X mental retardation 1; GABRB3, g-aminobutyric acid
(GABA) A receptor b3; GRIK2, glutamate receptor ionotropic kainate 2; ITGB3, integrin b3; MECP2, methyl CpG binding protein 2; MET, met
proto-oncogene; NLGN3, neuroligin 3; NLGN4X, neuroligin 4 X-linked; NRXN1, neurexin 1; OXTR, oxytocin receptor; PTEN, phosphatase and
tensin homologue; RELN, reelin; SHANK3, SH3 and multiple ankyrin repeat domains protein 3; SLC25A12, solute carrier family 25 (mitochondrial
carrier, Aralar) member 12; LC6A4, solute carrier family 6 (neurotransmitter transporter, serotonin) member 4; TSC1, tuberous sclerosis 1;
TSC2, tuberous sclerosis 2; UBE3A, ubiquitin protein ligase E3A. (From [8], with permission from Macmillan Publishing Ltd.)
5.5 1.0
5.0 0.9
4.5 0.8
Information content
4.0 Wave 1 0.7
3.5 ZLR= 4 .69, 0.6
Wave 2

Information content
lod= 4 .79,
3.0 p < 0.000002 All famillies 0.5
2.5 HaploSim analysis 0.4
ZLR

2.0 0.3
1.5 0.2
1.0 0.1
0.5 0.0
0.0
–0.5 Mb:
80 85 90 95 99
0 5 10 15 20 25 30 35 40 45
Location (cM)

Figure 18.1 Linkage to chromosome 15q25–26 in the GenRED study. Shown are Z likelihood ratio score statistics from analyses of linkage
of recurrent early-onset major depression to 88 chromosome15q single nucleotide polymorphisms in 631 families of predominantly
European ancestry for wave 1, wave 2, all families, and for information content. Physical locations in megabases from the p-telomere are shown,
with the peak at about 93 Mb. On the deCODE genetic map, the location is from 85.2 to 133.6 cM. Also shown are results of HaploSim analysis
demonstrating the absence of any systematic bias in the multipoint analysis due to linkage disequilibrium. (From [48], with permission.)

M IN TF NR2F2 Cluster 2
IN PIK3R1
IN CUGBP2

IN APOB

M IN PON2

IN C036
M IN GSTP1 M IN EPHX2

IN SPON1

IN ADAMTSL1 M IN RYR2 IN HOMER1

Cluster 1
IN SLC8A1 M IN KCNJ6
IN HAS2

IN EDN1 M IN TNFRSF1A IN DAPK1


M IN PTPRK
M IN PTGER4
M IN SLC1A2
IN TF MYBL2
IN A2BP1 M IN AKR1B1 IN TNFRSP21
IN SLC6A11
IN TF MITF
IN NRG1
IN PTPRS M IN FN1
IN SCNN1A
M IN TF NR2E1 IN CADM1
M IN RPS6KA1 Cluster 3
M IN RAP1B

M IN TGFBR3
IN TPM1
IN CALD1

IN SERPINA4
IN EFNAS

IN EPHAS

Figure 18.3 Results of a literature-based pathway analysis that includes all genes from the Munich Antidepressant Response Signature
project genome-wide association study (GWAS) that correspond to the single nucleotide polymorphisms (SNPs) implicated by the STAR*D
replication GWAS sample. Genes were categorized as related when they were co-cited in the same sentence with a functional descriptor in
between. There are 41 genes that cluster around fibronectin 1 (FN1) (cluster 1), ADAMTS-like 1 (ADAMTSL1) (cluster 2), and endothelin 1 (EDN1)
(cluster 3). Genes with corresponding SNPs that achieved nominal significant replication in the STAR*D sample are shaded in red; green
lines indicate transcription factor (TF) binding site matches in target promoters; the line with the yellow circle indicates annotation by
Molecular Connections experts. IN indicates input gene; M, part of a metabolic pathway. From [105], with permission.
(a) chr15 (q25.1)

(b) rs16969968

76589924 76711042
1.0 80

Recombination rate (cM/Mb)


0.8
60
R-Squared

0.6
40
0.4

20
0.2

0.0 0
CRAB1 IREB2 LCC123088 ADAMTS7
PSMA4
CHRNA5
CHRNA3
CHRNB4

76420 76545 76670 76795 76920


Chromosome 15 position (hg18) (kb)

Figure 23.3 Graphical representation of chromosome 15q25. (a) Chromosome 15 with the region q25.1 demarked by the red line, the
location of CHRNA5 and CHRNA3. This figure was created with the University of California, Santa Cruz (UCSC) genome browser (http://genome.
ucsc.edu). (b) The 100 kb region surrounding rs16969968 (large diamond). Diamonds indicate single nucleotide polymorphisms (SNPs), and the
size of the diamond is directly proportional to the r2 of the SNP with rs16969968. The dashed lines delineate the boundaries for SNPs that
have r2  0.8 with rs16969968. This figure was created using SNAP (http://www.broad.mit.edu/mpg/snap/).
12

(a)
9

Frequency (Hz)
Fz

Cz
6

Pz
3

Theta

-216 -31 153 338 522 706 891 1075 1259


(b) Time (ms)

Chromosome 7
3.5 θ
Fz, Max LOD = 3 .16 at 161 cM
3
Cz, Max LOD = 3 .6 at 164 cM

Pz, Max LOD = 2 .29 at 162 cM


2.5

2
LOD

1.5

CHRM2
0.5
GRM8

0
D7S1790

D7S1802

D7S1838

D7S2846

D7S1830

D7S3046
D7S1870

D7S1797

D7S1796

D7S1799

D7S1817

D7S2847
D7S1809

D7S1804

D7S1824

D7S1805
D7S513

D7S629
D7S673

D7S817

D7S521
D7S691
D7S478
D7S679
D7S665

D7S820

D7S821

D7S490

D7S509

D7S794
NPY2

0 20 40 60 80 100 120 140 160 180


Chromosome Position (cM)
CHRM2
exon

exon

exon

exon

exon

(c)
1 2 3 4 5 6
Coding
5’-UTR 3’-UTR
Sequence

(d) SNPs rs6948054 rs1424548


rs1378646 rs1378650
rs1455858
rs7782965
rs978437

Figure 29.1 Illustration of endophenotype strategy: from neurocognitive endophenotypes to genes. (a) Endophenotype (Theta oscillation),
Theta (y, 3–7 Hz) event-related oscillations (EROs) underlying the processing of target stimuli during P3 production (300–700 ms) in a visual
oddball task as seen in this time-frequency representation (right panel). Head plot (left) displays scalp topography of theta ERO power.
(Theta EROs are correlates of impaired cognitive brain processes in alcoholism and risk.). (b) Genetic linkage analysis scans the entire genome
to assess chromosomal regions that contain polymorphic genetic markers that are linked to a quantitative trait (theta ERO power) within
families. This is a logarithm of odds (LOD) score plot of linkage for theta ERO power at frontal (Fz), central (Cz), and parietal (Pz) electrodes on
chromosome 7 with a significant linkage peak that contains 2 candidate genes under the linkage peak – CHRM2 and GRM8. (c) Candidate
gene studies focus on genes underlying significant linkage peaks and/or genes with relevant biological significance. This is a schematic
diagram of the CHRM2 gene indicating its coding region (gray) and exons. (d) Genetic association tests are performed for each single
nucleotide polymorphism (SNP) in the candidate gene and the trait variable. SNPs are identified within and flanking the candidate CHRM2
gene (exons and introns). SNPs with significant associations (gray boxes) with the endophenotype indicate loci of interest. (The right-most
box with two SNPs lies in the region beyond the 3’ UTR of the gene.)
(a) (b)

Figure 29.2 (a) Linkage plots showing maximum logarithm of odds (LOD) scores with significant linkage peaks over the GABAA receptor gene cluster on chromosome 4 with two resting
EEG phenotypes. Red trace: Resting EEG beta band power – beta 2 (16.5–20.0 Hz). The dataset consists of 1553 individuals from 250 families. Green trace: Resting EEG high theta band
(6–7 Hz) interhemispheric coherence at parieto-occipital bipolar pairs of electrodes (P4_O2-P3_O1). The dataset consists of 1312 individuals from 251 families. (b) Linkage plot showing maximum
LOD scores with significant linkage peaks on chromosome 7, over the region harboring two candidate genes: a cholinergic muscarinic receptor gene (CHRM2) and a glutamate receptor gene
(GRM8). Blue trace: the central midline theta (4–5 Hz) ERO band power (between 300–700 ms, P3 latency window for visual target case during visual oddball task) on chromosome 7. The dataset
consists of 1337 individuals from 253 families; Green trace: Resting EEG theta band (6–7 Hz) interhemispheric coherence at centro-parietal bipolar pairs of electrodes (C4_P4-C3_P3) on
chromosome 7. The dataset consists of 1312 individuals from 251 families.
(a) Adult subjects
Control Alcohol dependent
Frequency (Hz)

Frequency (Hz)
10 10

5 5

0 0
0 200 400 600 0 200 400 600
Time Time

1 2 3 4 1 2 3 4
Amplitude(μV) Amplitude(μV)

(b)
Adolescent subjects
Low risk High risk
12

Frequency (Hz)
12
Frequency (Hz)

θ
7
7 θ
θ
δ 0
δ
0
0 200 400 600 0 200 400 600
Time
Time δ
2 4 6 2 4 6
Amplitude(μV) Amplitude(μV)

Figure 29.3 S-transform derived time-frequency representations of the average instantaneous amplitude that are z-scored for each
frequency. Instantaneous amplitudes were averaged across individual trials of subjects so that nonphase locked or imprecise phase locked
oscillatory energy is preserved. (a) Plots for the target condition at central (Cz) electrode in 120 alcoholic (right) and 120 control (left) subjects.
Center panel – Topographic headplots of the time-frequency regions of interest (TFROIs) indicate that the theta band (4–5 Hz) power at
300–500 ms shows frontal maxima while the delta band (1–3 Hz) power at 350–700 ms has posterior maxima. Note that alcoholics have
weaker responses in both theta and delta bands. (b) Plots for the target condition at frontal (Fz) electrode in visual oddball task in 87 high risk
adolescent offspring of alcoholics from Collaborative Study on the Genetics of Alcoholism (COGA) families (right) and 57 matched low risk
offspring of nonalcoholics from control families (left). Center panel – Topographic head plots for the TFROI that extends from 300 to 700 ms
for each band revealing frontal maxima for theta band power and parietal maxima for delta band power. Note that high risk adolescents
have weaker responses in both theta and delta bands to target stimuli, similar to the alcoholics.
Chapter 18: Genetics of major depression

Table 18.1 Family studies of major depression.

Proband/study Sample size (relatives) Age-adjusted lifetime prevalence in FDRs (%)

BP BP-I BP-II SA MDD


Major depression
Smeraldi et al., 1977 [6] 185 0.6 – – – 8.0
Tsuang et al., 1980 [7] 483 2.0 – – – 13.6
Abrams et al., 1980 [8] 106 4.7 – – – 7.5
Jakimow-Venulet, 1981 [9] 306 0.5 – – – 9.5
Gershon et al., 1982 [10] 166 – 1.5 1.5 0.7 16.6
Baron et al., 1982 [11] 143 2.2 3.0 17.7
Weissman et al., 1984 [12] 810 – 0.9 1.9 0.3 17.6
Andreasen et al., 1987 [4] 1171 – 0.6 2.9 0.2 28.4
McGuffin et al., 1987 [13] 315 – – – – 24.7
Maier et al., 1993 [14] 697 1.8 – – 0.5 21.6
Weissman, 1993 [15] 651 – – – – 24.4
Weighted mean 5033 1.7 0.8 2.4 0.5 20.5
Controls
Tsuang et al., 1980 [7] 541 0.3 – – – 7.5
Gershon et al., 1982 [10] 265 – 0 0.5 0.5 5.8
Weissman et al., 1984 [12] 521 – 0.2 1.1 0.2 5.9
Maier et al., 1993 [14] 419 1.8 – – 0.4 10.6
Weissman et al., 1993 [15] 255 – – – – 5.5
Weighted mean 2001 1.0 0.1 0.9 0.3 7.3
Abbreviations: BP, bipolar disorder; FDR, first-degree relative; MDD, major depressive disorder; SA, schizoaffective disorder; – ¼ not studied
or reported.

genes responsible for this inheritance. Determination An increased relative risk to siblings is observed
of heritability starts with studies of familial aggrega- when two additional clinical features of MDD, recur-
tion and proceeds with twin and adoption studies. rent episodes and an early age at onset, are considered
If a disease has a genetic basis, it is expected to run [16]. Most data suggest that familial risk is greater
in families. Table 18.1 shows that the lifetime preva- when the proband’s age of onset is below 40 and
lence of depression is elevated in first-degree relatives greater still when it is below 30 or 20. Lyons et al.
of depressed probands as compared to relatives of [17] reported that in 3372 male veteran twin pairs,
controls [4, 6–15]. It appears that rates of bipolar heritability of MDD was 0.47 when the age at onset
disorder are slightly elevated as well. A useful way of was below age 30, and 0.10 when it was above 30, and
estimating the strength of the genetic contribution to Bland et al. [18], in a study of 763 first-degree rela-
illness is to apply the concept of relative risk to sib- tives of 75 probands with MDD, found that relatives
lings, which entails comparing the rate of illness in of probands with an early age at onset and recurrent
siblings of ill probands with that in the general popu- depression had a 17.4% risk of depression while rela-
lation. The relative risk for MDD is  2.8, suggesting tives of probands with single-episode depression and
a modest genetic component for the broad phenotype. a late age at onset had only a 3.4% risk. The relative

213
Chapter 18: Genetics of major depression

Table 18.2 Twin studies of major depression.

Findingsa

Proband/study Sample size (pairs) Concordance (%)

MZ DZ Heritabilityb
Bertelsen et al., 1977 [20] 44 54 24 0.39
Torgersen, 1986 [21] 92 27 12 0.54
Andrews, 1990 [22] 82 7 9 0
McGuffin et al., 1996 [23] 177 46 20 0.48
Lyons et al., 1998 [17] 3372 23 14 0.36
Bierut et al., 1999 [24] 2662 33 26 0.30
Kendler et al., 2000 [25] 3790 44 39 0.34
Weighted mean 10 219 34 26 0.34
a
Only studies in which a majority of subjects were directly interviewed are included.
b
Where authors calculate heritability, these figures are provided. If heritability is not provided in a report, Holzinger’s heritability
(MZ concordance  DZ concordance/100  DZ concordance) has been calculated.
Abbreviations: DZ, dizygotic; MZ, monozygotic.

risk to siblings for early-onset recurrent depression is pairs born in Sweden between 1886 and 1958 and
probably at least 4–5 [19]. contacted between 1998 and 2003 for an assessment
Familial aggregation of disease suggests, but does of the presence or absence of depressive epsiodes.
not prove, a genetic contribution to disease. Environ- The investigators reported an overall heritability of
mental factors, such as emotionally traumatic family 0.38, with a significant difference between the sexes.
experiences or other stressors, for example, could also The heritability for women was 0.42 while that for
lead to familial aggregation. Twin studies have been men was 0.29 [26].
the approach used most widely to attempt to disen- Many observers have cited the lack of complete
tangle these contributions. The logic of twin studies is concordance between MZ twins as evidence that
as follows. (Please also see Chapter 1, which discusses environmental factors such as stressful life events
twin, family, and adoption methods in greater detail.) must play a role in the etiology of MDD. More
Identical or monozygotic (MZ) twins are 100% gen- recently, others have suggested that this lack of com-
etically the same, whereas fraternal or dizygotic (DZ) plete concordance could be due to epigenetic factors –
twins share just 50% of their genes; yet the two twin that is, factors that affect the control of gene expres-
types are assumed to be no different in the degree to sion – and that these factors may be influenced by
which they share environments. Therefore, any the environment [27]. In support of this hypothesis,
increased similarity in manifestation of depression there is evidence that MZ twins, despite having
detected in MZ twins as compared with DZ twins identical DNA sequence for all their genes, may
should be due to the greater genetic similarity of the differ in the way their genes are expressed [28].
former. The main measure in these studies is the Patterns of concordance can be converted to a herit-
concordance rate for illness. That is, starting with an ability value – often thought of as the percentage of
ill twin as the proband, what is the rate of illness in the liability to illness that is genetic – in several ways.
the co-twin? As seen in Table 18.2, seven studies [17, Current twin studies employ a liability threshold
20–25] have assessed MDD using primarily a direct model [29], which simultaneously tests the heritabil-
interview approach and reported on concordance ity of and environmental contribution to illness.
rates; they found rates of 34% in MZ twins and 26% When individuals cross the liability threshold, they
in DZ twins, with a heritability of 0.34. The largest develop illness. The model requires the specification
twin study to date looked at 15 493 complete twin of a population prevalence of illness to indicate how

214
Chapter 18: Genetics of major depression

many people in the population lie on the ill side of to be less than that for several other major psychiatric
the threshold. Moreover, heritability estimates will illnesses discussed in this book, such as bipolar dis-
vary with the prevalence figures used, an issue of order, schizophrenia, and autism, but comparable to
particular importance for MDD, for which popula- that for alcohol and other substance dependence.
tion estimates have varied dramatically. For example,
one depression study reported that heritability was
48% when one estimate was used and 75% when Gene-mapping
another was used [23].
A second paradigm for separating genetic from Linkage
environmental effects is the adoption study. The con- Linkage is a method for identifying chromosomal
ceptual basis for this approach is that in adoptees, the regions likely to harbor a disease susceptibility gene.
genetic inheritance occurs through one set of parents, It hinges on the biological process of meiosis and the
while the cultural and environmental experience recombination of homologous chromosomes that is
occurs through a different set. The two sets of factors central to that process. A putative disease gene is
and their potential association with illness can there- linked to a marker if it stays together with the marker
fore be disentangled. in the large chunks of chromosome that are recom-
There are several possible ways to conduct an bined. The resolution of linkage is low, on the order
adoption study. One that has been used for MDD is of 20 Mb, because over the course of a couple of
the adoptees’ relatives method. In this method, ill and generations, there is relatively little recombining.
control adoptees are identified as probands, and rates The first linkage studies of depression were done
of illness are then compared in the biological and at the University of Iowa by Tanna, Go, Winokur, and
adoptive relatives of each group. If genetics plays a others. Using blood group antigens and enzymatic
role, the biological relatives of ill adoptees will have and protein markers, they found several modest link-
elevated rates of illness as compared with the other age signals that could not be replicated [33–36]. Other
three relative groups. teams were also either unable to replicate these early
Adoption studies are not easy to conduct as they linkage findings for depression or obtained conflict-
require access to large databases, and there are bar- ing results with evidence of linkage to other markers
riers of confidentiality involved whose breach can be a [37–39].
highly sensitive matter. Two such studies that have These early studies of the genetics of depression
been done have used national registries – those of usually assumed that the inheritance of MDD was
Sweden and Denmark. In these studies, direct inter- fairly simple, that straightforward Mendelian mech-
views were not performed. In these two studies, more- anisms operated in the transmission of risk genes
over, probands had diagnoses including “affect and that only a few genes would be found to underlie
reaction” and “neurotic depressive reaction”, which the development of depressive illness. It is now clear
are not readily reconcilable with current diagnostic that this is almost certainly not the case. MDD is a
nomenclature. The results of one of these studies genetically complex disease. Genetic mechanisms
supported genetic transmission of depression [30], appear to affect susceptibility to MDD rather than
while those of the other did not [31]. Another adop- causing it more or less directly as is seen in, for
tion study used relatives as probands, and measured example, sickle cell disease, Huntington’s disease, or
rates of depression in the adoptees of those relatives phenylketonuria, where the relative risk to siblings is
who were affectively ill as compared with those who in the vicinity of 2500.
were not. The results supported genetic transmission Because rates of MDD in women are about twice
[32]. Thus of the three adoption studies, two provide those in men, investigators have wondered whether
support for a genetic vulnerability, and one does not. susceptibility genes might differ to some extent by
In sum, MDD has a substantial genetic compon- gender. Because of the potential for sex-specific gen-
ent, suggesting that the identification of susceptibility etic effects in MDD, gene-mapping studies have often
variants should be possible. This encouraged incorporated or even focused on sex-specific analyses.
researchers, beginning in the early 1980s to initiate One study found significant parametric evidence of
molecular studies to search for these variants. How- linkage to markers in 2q33–34 in 170 affected female
ever, the magnitude of the genetic component appears sibling pairs, but not in male pairs [40]. The region

215
Chapter 18: Genetics of major depression

between the markers that yielded the peak logarithm affected relative pairs [47]. Evidence for linkage on
of odds (LOD) score includes the CREB1 gene, which 15q25–26 was reduced to the “suggestive” level.
encodes a cyclic adenosine monophosphate (cAMP) - In secondary sex-specific analyses, nearly significant
responsive element-binding protein (CREB), an evidence for linkage was observed on chromosome
attractive candidate gene because CREB has been 17p12 (LOD ¼ 4.77, excess sharing in male–male and
implicated in depression and antidepressant response. male–female pairs), and suggestive linkage on
Levels of CREB have been found to be abnormally low chromosome 8p22–21.3 (LOD ¼ 3.49, excess sharing
in people with MDD and in the brain tissue of suicide in male–male pairs). We next genotyped 88 single
victims, and are altered by exposure of rat neurons nucleotide polymorphisms (SNPs) across our
to antidepressants [41–43]. A second study found 15q25–26 peak in 631 European-ancestry families
evidence of linkage uniquely in families with at least and carried out multipoint allele-sharing linkage ana-
four affected males [44]. This linkage, in chromo- lyses [48]. The maximum evidence for linkage was
somal region 12q22–23.2, was detected in a sample LOD ¼ 4.78 at 109.8 cM (Figure 18.1). The exact
of Mormon families in Utah in which ascertainment p-value (0.0000014) surpassed the genome-wide signifi-
was restricted to families with a minimum of four cance threshold. It was estimated that the linked locus
affected relatives. In addition to subjects with recur- or loci in this region might account for a 20% or less
rent MDD, individuals with only a single episode of population-wide increase in risk to siblings of cases.
major depression were considered affected, as were In sum, as has been the case for other psychiatric
those with bipolar disorder (who made up about 15% disorders, linkage studies in MDD have not consist-
of the subjects with mood disorder in these families). ently identified the same chromosomal regions,
McGuffin et al. carried out a genome-wide linkage although there has been some encouraging overlap,
scan in a sample of 497 sibling pairs concordant for most prominently on chromosome 15q25–26. Unlike
recurrent MDD [45]. There was suggestive evidence the case for bipolar disorder and schizophrenia, the
for linkage on chromosome 1p36 where the LOD number of genome-wide linkage studies done to date
score for female–female pairs exceeded 3.0 (but has been small, and no meta-analysis of them has yet
reduced to 2.73 when corrected for multiple testing). been carried out.
The region includes a gene – MTHFR – that in previ-
ous studies has been associated with depressive symp-
toms. Two other regions, on chromosomes 12q23.3– Alternative phenotypes
24.11 and 13q31.1–31.3, showed evidence for linkage One approach to potentially improving genetic link-
with a nominal p < 0.01. The 12q peak overlaps with age signals is to attempt to reduce genetic heterogen-
a region previously implicated by linkage studies of eity through reducing clinical heterogeneity. We have
unipolar and bipolar disorders and contains a gene – referred to several alternative phenotypes intended to
DAO – that has been associated with both bipolar capture more homogeneous subsets of patients,
disorder and schizophrenia. The 13q peak lies within including using recurrence of episodes, earlier age at
a region previously linked strongly to panic disorder. onset, and single sex analysis. There are a number of
A fourth modest peak with a LOD of greater than 1.0 other phenotypic distinctions that might help define
on chromosome 15q lies within a region that showed more genetically homogenous subsets of patients.
genome-wide significant evidence of a recurrent For example, Camp et al. [49] examined recurrent,
depression locus in a previous sibling pair study (the early-onset MDD (MDD-RE) and anxiety disorders in
GenRED study, see below). 87 large, extended Utah pedigrees to investigate link-
In the Genetics of Recurrent Early Onset age to three phenotypes: “MDD-RE”, “MDD-RE or
(GenRED) project, we studied 297 families initially anxiety”, and “MDD-RE and anxiety”, where in the
(Wave 1) and carried out a linkage analysis. We latter definition the disorders must appear comorbid
observed a genome-wide significant peak on distal within an individual. Pedigrees ranged in size from
15q (LOD ¼ 3.73) [46], with the peak being even 2–6 generations and contained 3–42 individuals
stronger when the analysis was restricted to the 286 affected with MDD or anxiety (718 total). In their
European-ancestry families (LOD ¼ 4.39). Later we primary analyses, they identified three regions with at
jointly analyzed our Wave 1 plus Wave 2 sample of least suggestive genome-wide evidence for linkage on
656 families, including 1494 informative “all possible” chromosomes 3centr, 7p, and 18q. The best linkage

216
Chapter 18: Genetics of major depression

5.5 1.0
5.0 0.9
4.5 0.8
Information content
4.0 Wave 1 0.7
3.5 ZLR= 4 .69, 0.6
Wave 2

Information content
lod= 4 .79,
3.0 p < 0.000002 All famillies 0.5
2.5 HaploSim analysis 0.4
ZLR

2.0 0.3
1.5 0.2
1.0 0.1
0.5 0.0
0.0
–0.5 Mb:
80 85 90 95 99
0 5 10 15 20 25 30 35 40 45
Location (cM)
Figure 18.1 Linkage to chromosome 15q25–26 in the GenRED study. Shown are Z likelihood ratio score statistics from analyses of linkage of
recurrent early-onset major depression to 88 chromosome15q single nucleotide polymorphisms in 631 families of predominantly European
ancestry for wave 1, wave 2, all families, and for information content. Physical locations in megabases from the p-telomere are shown, with the
peak at about 93 Mb. On the deCODE genetic map, the location is from 85.2 to 133.6 cM. Also shown are results of HaploSim analysis
demonstrating the absence of any systematic bias in the multipoint analysis due to linkage disequilibrium. (From [48], with permission.) See
plate section for color version.

evidence was for a novel locus at 3p12.3–q12.3 (LOD ¼ onset have been studied, but a variety of others might
3.88, “MDD-RE or anxiety”) and 18q21.33–q22.2 be relevant, including chronic course [51], attempted
(LOD ¼ 3.75, “MDD-RE and anxiety”), an established suicide [52], and presence of alcoholism in a family
susceptibility locus for bipolar disorder. along with MDD [53].
We have also employed the phenotype of postpar-
tum mood symptoms in analyses that combined the
GenRED depression sample and the National Institute Association
of Mental Health (NIMH) Genetics Initiative bipolar Association provides an alternative to the linkage
disorder sample [50]. We included women with a method. While linkage is a property of genes or loci
history of pregnancy, any mood disorder, and infor- and occurs within families, association is a property
mation about postpartum symptoms. There were 1210 of alleles and occurs across a population. An associ-
women who met our criteria (30% of whom were ation study can be used for two different purposes.
positive for a history of postpartum mood symptoms). The first is to test directly whether a genetic variant
The maximum linkage peak for postpartum symptoms might be implicated in MDD. An association between
occurred on chromosome 1q21.3–32.1, with a a phenotype and an allele at a locus may mean that
chromosome-wide significant ZLR (an allele-sharing the allele in question leads to susceptibility to the
score much like a LOD score) of 2.93 (permutation phenotype. The second use of association studies is
p ¼ 0.02). This was a significant increase over the to indirectly test the same thing through linkage dis-
baseline ZLR of 0.32 observed at this locus among equilibrium (LD) mapping. This indirect approach
all women with a mood disorder (permutation p ¼ can provide information about the location of a dis-
0.004). Suggestive linkage was also found on 9p24.3– ease allele with resolution that is roughly 1000-fold
p22.3 (ZLR ¼ 2.91). Our results suggest that genetic greater than that of a linkage study.
variations in these regions might increase susceptibil- There are two types of candidate genes – functional
ity to postpartum mood symptoms. candidates and positional candidates. Functional can-
Phenotypic subtypes might shed additional light didate genes are those coding for a protein thought
on linkage results. Comorbid anxiety and postpartum to have some biological role in MDD. Positional

217
Chapter 18: Genetics of major depression

5 Figure 18.2 Meta-analysis of association


studies of functional candidate genes in
major depressive disorder (MDD). Pooled
odds ratios are presented in relation to
GNB3 the total number of cases and controls
included in the meta-analyses. Odds
SLC6A3 ratios are reported for the allelic analyses.
2
The five genes that were significant in
Pooled odds ratio

MTHFR either this or genotypic analyses are


SLC6A4 highlighted. Note that this study was
done before the availability of genome-
1 wide association study data. (From [56],
with permission from Macmillan
Publishers Ltd.)

0.5

APOE
0.2
0 1000 2000 3000 4000 5000 8000 10000 12000
Total number of cases and controls

candidates are those identified through a linkage or A review of association studies [56] found that
LD mapping study. of 102 functional candidate genes that had been stud-
ied for association in MDD, there were sufficient
data for meta-analysis on 18 of them, among which
Functional candidates 11 involved the monoamine neurotransmitters:
Because of neurobiological data implicating the COMT, DRD3, HTR1A, HTR1B, HTR2A, HTR2C,
neurotransmitters, particularly serotonin, norepin- MAOA, SLC6A2, SLC6A3 (DAT1), SLC6A4 (SERT),
ephrine, and dopamine, in MDD, most early studies and TPH1. Five of the 18 showed statistically signifi-
focused on functional candidate genes from these cant evidence of association (Figure 18.2). SLC6A3
systems. While pathophysiological mechanisms of encodes the dopamine transporter, and this gene has
MDD susceptibility are still not known, more recent a 40 bp variable repeat polymorphism, which was
data suggest additional factors that include changes in investigated in three studies. The pooled odds ratio
the hypothalamic-pituitary axis leading to dysregula- for the “9/10” genotype compared to the 10/10 was
tion of cortisol secretion, dysregulation of neuropro- 2.06. APOE was examined in seven studies, and the ε2
tective mechanisms, and inflammatory and immune allele was found to be protective. Guanine nucleotide-
mechanisms. binding protein (GNB3) was tested in three studies
The evidence that the monoamine neurotranmsit- and the T allele of a C825T polymorphism was asso-
ters play a role in MDD first came from drug studies, ciated with increased risk. The MTHFR C677T poly-
when reserpine, a blood pressure medication, was morphism was investigated in six studies and the
found to cause depression [54]. It was later learned T allele was associated with increased risk. SLC6A4
that reserpine inhibits vesicular monoamine trans- encodes the serotonin transporter, the most heavily
porters, thus depleting the brain of dopamine, nor- studied gene in psychiatric genetics. The promoter
epinephrine, and serotonin. These and other length polymorphism has received exceptional atten-
observations led to the theory that depletion of cat- tion because it was shown to be functional, with the
echolamines, particularly norepinephrine, in brain short variant being associated with lower levels of
synapses, causes depression [55]. Later, attention gene transcription [57]. Twenty-four studies with
focused on depletion of serotonin, which led to the 3752 cases and 5707 controls have examined this
development of the serotonin selective reuptake variant and found that the short allele carried a sig-
inhibitors (SSRIs). nificantly increased risk for MDD with an odds ratio

218
Chapter 18: Genetics of major depression

of 1.11, while the homozygous genotype increased A third area of biological interest in relation to
risk by 1.39. We discuss this gene further in the MDD that has generated candidate genes is the hypo-
context of the gene–environment interaction section thalamic-pituitary-adrenal (HPA) axis. The hypothal-
below. An important caveat in these results is that amus produces corticotrophin releasing hormone
among the five associated genes, for only one of them (CRH) which stimulates the pituitary to produce
(SLC6A4) was the number of MDD cases tested adrenocorticotropin hormone (ACTH), which in turn
> 1000. The small sample sizes raise questions about stimulates the adrenals to produce cortisol, the stress
the robustness of the results. hormone. Neurotransmitter systems, including the
Functional candidate gene studies have also more serotonergic system, influence the production of
recently emerged from the neurotrophic model of CRH. Cortisol ultimately docks with the glucocorti-
depression, a newer hypothesis about the nature of coid receptor, which forms a complex that goes into
depression [58]. The neurotrophic model stems from the nucleus of the cell where it binds to specific
research into the mechanism of action of antidepres- DNA sequences and influences gene expression. The
sants, which has found that these medications stimulate first studies showing a relationship between elevated
brain derived neurotrophic factor (BDNF) and also its cortisol and depression were done in the 1950s [68].
cousin neurotrophin 3 (NT3) [59] in the hippocampus, Variations in HPA axis genes may influence the stress
and that injection of these two neurotrophins into the response and susceptibility to MDD. Haplotypes of
hippocampus recapitulates the antidepressant effects the glucocorticoid receptor gene have been associated
in the rat that the medications themselves induce. with MDD [69] as have variations in the CRH recep-
BDNF treatment also reversed two animal paradigms tor 1 gene, as a moderator of the relationship of child
that have been used as models of depression, namely abuse to MDD (see below) [70]. The FKBP5 protein
learned helplessness following exposure to inescapable forms part of a complex with the glucocorticoid
shock and the forced swim test [60]. In humans, serum receptor and can modulate cortisol-binding affinity.
BDNF levels were shown to be lower in depressed Studies have found association of FKBP5 variation
subjects compared to controls [61, 62]. Some genetic with MDD [71], as well as the related phenotypes of
association studies have implicated BDNF variants in differential response to antidepressant drug treatment
MDD [63, 64], although a meta-analysis of 14 studies [71, 72], number of depressive episodes [72], and
found no overall association for the gene [65]. bipolar disorder [73]. Again, an important caveat here
After having reported genome-wide significant is that there are negative reports for each of these
linkage on chromosome 15q25.3–26.2 to recurrent genes as well, and they have not been robustly repli-
early-onset MDD, as described above, we followed cated in genome-wide association studies data.
up with linkage disequilibrium (LD) fine-mapping of
this signal and sequence analysis of NTRK3, which
encodes the receptor for NT3, given its biological Genome-wide association studies
plausibility as a candidate gene [66]. LD mapping Genome-wide association studies (GWAS) provide
showed nominally significant association in nine genes, a method for screening the entire genome at high
including NTRK3, with MDD-RE. In NTRK3, five SNPs resolution in an unbiased way, in a search for
had nominally significant p-values. Sequence analysis disease susceptibility variants. This is an extremely
revealed 35 variants. While the number of rare variants powerful approach, though because tests of so
did not exceed chance expectation, case-control analysis many DNA markers are being done, the statistical
of 13 common variants did show modest nominal bar is raised very high for considering findings true
association of MDD with 3 SNPs (p ¼ 0.008, 0.048, positives. Because these studies perform roughly
and 0.034), which were in partial LD with 4 of 5 1 million tests across the genome, the statistical
associated SNPs from the family based experiment. threshold for calling a finding genome-wide signifi-
We felt that while common variants in NTRK3 or one cant is 5  10–2/1  10–6 ¼ 5  10–8. In addition,
of the other genes identified might play a role in because the effect size for any given variant is
MDD, much larger studies would be required for full expected to be quite low, the number of subjects
evaluation of this region. Another report has found required for adequate power is very large, at least
an association of NTRK3 SNPs with childhood-onset in the thousands, and perhaps in the tens of
mood disorders [67]. thousands.

219
Table 18.3 Strongest findings in a meta-analysis of three MDD GWAS. Broad phenotype of all MDD.

Band SNP BP A1/ All GAIN-MDD GenRED STAR*D Meta-analysis Annotation

A2 Frq OR p OR p OR p OR p
8p21.3 rs1106634 20110329 A/G 0.12 1.27 3.93E-03 1.27 5.39E-03 1.35 8.18E-05 1.30 6.78E-07 ATP6V1B2; SLC18A1; LZTS1
7p15.3 rs17144465 21470952 A/G 0.04 1.44 4.82E-03 1.82 5.97E-06 1.32 1.47E-01 1.56 7.68E-07 SP4
3p26.1 rs9870680 7504555 C/T 0.43 1.23 1.23E-04 1.22 4.76E-04 1.10 6.89E-02 1.19 1.11E-06 GRM7
7q32.3 rs10265216 130550661 A/T 0.29 1.23 7.64E-05 1.14 3.12E-02 1.16 2.32E-02 1.19 3.02E-06 mRNA AK294384
3q26.32 rs644695 178774391 A/G 0.87 1.61 3.06E-07 1.26 3.20E-02 1.00 9.85E-01 1.35 4.46E-06
2p14 rs724568 67795984 A/C 0.36 1.20 2.58E-04 1.09 1.51E-01 1.19 1.68E-03 1.17 5.32E-06 mRNA BC043421
Xq21.33 rs5990417 94966526 T/C 0.82 1.20 2.90E-03 1.25 3.70E-03 1.27 5.27E-03 1.23 6.11E-06
6p25.1 rs2326810 6557466 C/G 0.92 1.21 7.29E-02 1.48 1.84E-03 1.75 1.07E-05 1.41 6.52E-06 LY86
10p11.23 rs1612122 29331904 A/T 0.48 1.15 4.51E-03 1.20 2.38E-03 1.18 4.34E-03 1.17 6.98E-06
Abbreviations: GWAS, genome-wide analysis studies; MDD, major depressive disorder; OR, odds ratio; SNP, single nucleotide polymorphism.
Modified from Shyn et al. [78].
Chapter 18: Genetics of major depression

The first MDD GWAS was published by Sullivan 18q22.1 (rs17077540, P ¼ 1.83  10–7) in a region
et al. [74], in the context of the Genetic Association that has produced evidence for linkage to bipolar I or
Information Network, or GAIN, a collaborative effort II disorder, within an mRNA detected in human
between the National Institutes of Health (NIH) and brain tissue (BC896490) and approximately 75 kb
several private companies. Using a sample from the upstream of DSEL. Another signal was on chromo-
Netherlands, Sullivan conducted a GWAS study of some 7p15.3 within SP4, a signal that comes almost
435 291 SNPs genotyped in 1738 MDD cases and entirely from females. A member of the Sp family of
1802 controls selected to be at low liability for transcription factors, Sp4 is specific to neurons, forms
MDD. The top signals were found in a 167 kb region complexes with estrogen receptors that influence
overlapping the gene piccolo (PCLO), whose protein regulation of many genes, and could play a role in
product localizes to the cytomatrix of the presynaptic the mediation of neuroprotective enzymes and in
active zone and is important in monoaminergic glutamate-induced neurotoxicity. Zhou et al., who
neurotransmission. P-values did not, however, reach have implicated the gene in bipolar disorder [76],
the threshold for genome-wide significance, as the have also reported that reduced expression of Sp4 in
strongest one, rs2715148, registered at 7.7  10–7. mice leads to hippocampal vacuolization, age-
The investigators undertook replication of SNPs in dependent reduced expression of neurotrophin 3
this region in 5 independent samples, totaling 6079 and deficits in sensorimotor gating and contextual
MDD cases and 5893 controls, but no SNP exceeded memory [77].
the replication significance threshold when all repli- We then collaborated with Dr. Steven Hamilton at
cation samples were analyzed together. the University of California San Francisco (UCSF) to
The second published MDD GWAS report carry out a meta-analysis of three MDD GWAS
involved two independent datasets, one German and samples: GenRED; Dr. Hamilton’s genotypic data
one Swiss [75]. The first experiment was performed for EA cases from the National Institute of Mental
on 1022 recurrent MDD patients and 1000 controls Health (NIMH) -sponsored STAR*D antidepressant
genotyped on the Illumina 550 platform. The second effectiveness clinical trials program; and the Nether-
was conducted on 492 recurrent MDD patients and lands sample of Sullivan et al. described above [78].
1052 controls selected from a population-based col- The top results are summarized in Table 18.3 [78].
lection, genotyped on the Affymetrix 5.0 platform. Even with this much increased sample size, now
Neither GWAS identified any SNP that achieved totaling 3980 cases and 3428 controls, no signal
genome-wide significance. A meta-analysis of the reached genome-wide significance. The strongest
two did not yield genome-wide significant results evidence for association was observed for intronic
either. The most significant p-value (5.8  10–6) from SNPs in ATP6V1B2 (p ¼ 5.69  10–7), GRM7 (p ¼
the meta-analysis of the two studies was observed for 7.11  10–7), and, once again, SP4 (p ¼ 8.38  10–7).
a SNP, rs4238010, on 12p13 located > 260 kb from the The SP4 gene has been discussed above. ATP6V1B2
closest known gene, cyclin D2 (CCND2). Forty-two encodes a subunit for a vacuolar proton pump
SNPs showed meta-analysis association p-values < ATPase. Our signal falls in a distinct LD block within
10  10–5. ATP6V1B2, but SLC18A1 (previously VMAT1), which
In the GenRED study, we carried out a GWAS transports monoamines into synaptic vesicles, lies
using 1110 European-ancestry (EA) MDD cases, 655 very near ATP6V1B2, and could conceivably be impli-
recruited for our linkage study mentioned earlier, cated by this signal.
and 455 recruited for association studies; 1020 cases GRM7 is particularly intriguing as it encodes
were available for analysis after quality control (QC). metabotropic glutamate receptor 7 (mGluR7), which
The genotyping was performed on an Affymetrix prior studies suggest may be involved in mood regu-
6.0 array. Of 2653 EA control subjects available from lation [79]. For example, an mGluR7 agonist had
another project, we selected 1636 after excluding antidepressant-like effects in mice that were blocked
those who met criteria for lifetime MDD or who by knockout of GRM7 [80]. This is the third GWAS to
reported recurrent MDD that missed the threshold report evidence of association to mood disorders in
by one criterion. Unfortunately, no result met this long gene (880 kb), the others being the German/
genome-wide significance criteria. The strongest evi- Swiss recurrent MDD GWAS mentioned above [75],
dence for association was observed on chromosome and the Wellcome Trust Case-Control Consortium

221
Chapter 18: Genetics of major depression

GWAS of bipolar disorder [81], though the strongest is not clear that expression in these white blood cells
SNPs in those studies were in different parts of the mirrors expression in neurons, as control of expres-
gene. Larger samples will be required to determine if sion may be tissue- and cell-type specific. On the
genome-wide significant evidence for association can other hand, some abnormalities that have been
be detected for any variant in GRM7, but the bio- detected in lymphocytes appear to mirror changes
logical evidence suggests that this locus merits further detected in postmortem brain tissue. For example,
investigation. Sullivan et al. tested for correlation across the tran-
Prior functional candidate genes showed little scriptome of lymphocyte to brain region-specific
evidence of association in the meta-analysis. There expression levels and found substantial correlation
were 41 assessed, and as a group the distribution of with several regions including prefrontal cortex and
gene-wide p-values did not deviate significantly from amygdala [82].
the null distribution. Among the genes mentioned A recent review of expression studies in MDD
above, no SNPs in DRD4, APOE, BDNF, SLC6A3, brain demonstrated replicated changes in levels of
SLC6A4, MTHFR, or GNB3 were significant after 15 genes, including decreased expression for genes
correction for the number of SNPs assayed across in glutamate transport (e.g. GLUL and SLC1A2/
the gene. This suggests that if variation in these genes EAAT2) and metabolism, neurotrophic signaling
plays a role in MDD, the effect sizes would have to be (e.g. FGF, BDNF and VGF), and MAP kinase path-
quite small. ways (Table 18.4) [83]. The authors point out that a
A larger meta-analysis is ongoing through the powerful strategy for drug development is to deter-
Psychiatric GWAS Consortium (PGC). It is hoped mine whether genes identified in the brains of
that when the combined sample reaches into the affected patients are altered in a reciprocal way in
10 000–20 000 subjects range, sufficient power will the brains of animals exposed to candidate thera-
be available to definitively identify MDD-associated peutic compounds.
genes and SNPs. Functional studies of the impact of an allele asso-
In sum, as of this writing, there are no associations ciated with MDD can be performed using cell culture.
between specific DNA sequence variants and risk of A copy of the gene containing the candidate allele can
MDD that can be considered definitively established. be inserted (transfected) into cells in culture. These
The gene-mapping approach could produce such cells can then be compared with control cells. Gene
variants as the sample sizes continue to grow. Alterna- expression in these cells can be measured, as can
tively, other approaches could bear fruit, such as relevant consequences of gene function, such as bind-
approaches that investigate epigenetics or gene expres- ing capacity for a receptor. For example, the serotonin
sion, or those that involve study of gene function in cell transporter promoter polymorphism was studied in
culture or animal models. this way. Lymphoblast cell lines were transfected with
genes encoding the long variant and genes encoding
the short variant. The transcriptional activity of the
Gene expression and function long variant was found to be twice that of the short
Studies of gene expression in MDD have been carried variant [57]. The observation that this variant is func-
out in brain samples and in white blood cell samples tionally significant has helped make the serotonin
from MDD patients, as well as in experimental cell transporter promoter polymorphism the object of
lines and in mouse models. Studies focused on the intense study.
brain face a number of important difficulties includ- An even more powerful approach to the study of
ing limited access to tissues and variable quality of gene function is to create mouse models for the gene
available samples. under study. A variety of approaches exist for genetic-
Studies using white blood cells from patients have ally altering the mouse. The knockout approach allows
the great advantage that these samples are much more for the creation of a mouse lacking one or both copies
readily obtainable than brain tissue. Although studies of the gene. Tissue-specific knockout methods allow
have indicated that functional abnormalities exist in even more narrowly defined gene effects to be exam-
the lymphocytes of MDD patients, there have been ined. In the conditional knockout, the mouse carries
few efforts to examine gene expression. The major the gene, but investigators can turn it off at will.
weakness of this approach to studying MDD is that it Transgenic approaches allow for the introduction of

222
Chapter 18: Genetics of major depression

Table 18.4 Replicated gene expression changes in major depressive disorder (MDD) brain.

Gene symbol Gene name Expression change


GLUL Glutamate-ammonia ligase Decreased
SLC1A2/EAA T2 Astrocyte high-affinity glutamate transporter Decreased
AGXT2L1 Alanine-glyoxylate aminotransferase 2-like 1 Decreased
FGF1 Fibroblast growth factor protein 1 (acidic) Decreased
FGF2 Fibroblast growth factor protein 2 (basic) Decreased
FGFR1 Fibroblast growth factor receptor 1 Increased
FGFR2 Fibroblast growth factor receptor 2 Increased
FGFR3 Fibroblast growth factor receptor 3 Increased
NCAM1 Neural cell adhesion molecule Decreased
GPR37 G protein-coupled receptor 37 endothelin receptor type B-like Decreased
GPRC5B G protein-coupled receptor, family C, group 5, member B Decreased
DIMT1L DIM1 dimethyladenosine transferase 1-like Decreased
PRPF19 PRP19/PSO4 pre-mRNA processing factor 19 homolog Decreased
NTRK2 Tropomyosin receptor related kinase 2, TrkB Decreased
AQP4 Aquaporin4 Increased
Adapted from Altar et al. [83].

extra copies of normal genes or copies of altered copy of the GR gene, the GR-heterozygous mutant
genes. The knock-in approach allows for the simul- [85]. These mice exhibit normal baseline behaviors,
taneous knockout of the normal gene and transgenic but demonstrate increased helplessness only after
insertion of an altered version of the gene. With these stress exposure. Similar to depressed patients, these
mouse models, gene effects in the brain and on behav- mutant mice have a disinhibited HPA system and a
ior can be studied. pathological dexamethasone/corticotropin releasing
The glucocorticoid receptor (GR) has been much hormone test. They further have downregulation of
studied in mouse models of depression and anxiety. BDNF protein in the hippocampus, consistent with
Lines of mice generated include the following: those the neurotrophin hypothesis of depression.
with disrupted GR alleles; those with nervous system-
specific knockout of GR; transgenic mice with
increased GR expression; transgenic mice that express Epigenetics
an antisense RNA, which results in decreased GR The modest level of heritability of MDD suggests that
expression by binding to and neutralizing the normal the DNA sequence does not fully explain the variabil-
GR RNA; and knock-in mice, in which the normal ity in susceptibility to this illness. There are at least
gene is knocked out while abnormal GR genes are two other major kinds of explanations for this vari-
introduced and then expressed [84]. Behaviors ation. One of these is that environmental factors such
thought to model depression – such as performance as stressful life events play a significant role in trig-
on the Porsolt forced swim test, the tail suspension gering MDD [86, 87], and another is that epigenetic
test, or the learned helplessness paradigm – can be factors are involved. These may be interdependent
measured in these genetically altered mice as a way of as the environment may cause epigenetic changes.
gauging the impact of the gene, and these behaviors Epigenetic changes, such as those in DNA methyla-
may have relevance for MDD. A particularly interest- tion (DNAm) and in histone modification, are often
ing model is one in which the mouse is missing one correlated with changes in gene expression.

223
Chapter 18: Genetics of major depression

In an animal model of early life stress, reduced itself influenced by genetic factors, and that MDD
maternal care, epigenetic changes, including in DNAm itself constitutes a major stressor.
and in histone modification, were seen in the promoter Several studies have found support for the inter-
region of the glucocorticoid receptor gene, and action of genetic factors and stressful life events in
these persisted into adulthood, where they correlated the MDD etiology. Kendler et al. [87], studying 2164
with disruption of the hypothalamic-pituitary-adrenal female twins, found 12 life event predictors of depres-
(HPA) axis [88]. The same group of investigators sion. The four strongest were death of a close relative,
showed that in DNA from postmortem hippocampus assault, serious marital problems, and divorce/
obtained from suicide victims with a history of child- breakup. In those at lowest genetic risk (identical
hood abuse, as compared to that from either suicide twin, co-twin unaffected with MDD), the probability
victims with no childhood abuse or from controls, of MDD onset per month was 0.5% and 6.2%, for
there was increased DNAm of a glucocorticoid recep- those unexposed and exposed, respectively, to a severe
tor gene promoter [89]. life event. In those at highest genetic risk (identical
There are several examples of epigenetic variation twin, co-twin affected with MDD) the rates were 1.1%
in candidate MDD genes and in DNA treated with and 14.6%. The 14.6% figure represents the interactive
medications used for MDD. For example, early life risk of MDD in those with both genetic vulnerability
adversity increased DNAm in Bdnf in rats [90]. and exposure to high stress [87].
Valproate [91], used to treat bipolar depression, and The most heavily studied interaction between a
haloperidol [92], used for psychotic depression, as specific genetic variant and an environmental factor
well as the antidepressants imipramine [93], tranylcy- in MDD is that between the promoter length poly-
promine [94], and fluoxetine [95] have been shown to morphism of the serotonin transporter gene, known
induce epigenetic changes in rodent brain. Further, as 5HTTLPR, and stressful life events. A study by
administration of a histone deacetylase inhibitor, Caspi et al. [97], published by Science in 2003, found
sodium butyrate, produces an antidepressant effect that in a sample of 800 New Zealanders the short
in an animal model [96]. allele of 5HTTLPR predicted depression only in the
Despite the availability of an essentially complete setting of substantial life stressors. This study was
genome sequence for several years, understanding the called one of the findings of the year across all of
methylome, the full complement of DNA methylation science for 2003, and it has since been cited over
across the genome, has progressed more slowly, 2600 times. There have been many attempts at repli-
largely due to limitations in technology affecting cation, with studies reporting mixed results. In 2009, a
sensitivity, specificity, throughput, quantitation, and meta-analysis of 14 studies on the issue was published
cost among the previously used detection methods. in JAMA and it was disappointing. The study con-
Microarray-based methods can interrogate much larger firmed a three-fold increased risk for MDD in those
numbers of CpGs, the sites of DNAm, than other with a high rate of stressful life events, but it failed to
approaches. But no published study has yet assayed the show any interaction between the 5HTTLPR and these
genome for DNAm in MDD, though several groups, events [98]. However, a more recent meta-analysis that
including our own, are pursuing this work. was far more inclusive – 54 studies included compared
to 14 for the 2009 report – did find that the short allele
was associated with an increased risk of developing
Genes and environment depression under stress [99]. This is, of course, only
We have mentioned the connected hypotheses that one of a great many potential genetic variants that
gene and environment might interact to cause MDD, might be relevant for such interactions.
that stress might be the most relevant environmental Another example of gene–environment interaction
factor, and that epigenetics might be a mechanism comes from a key gene in the HPA axis, CRHR1, for
that mediates between stress and gene. Some have which variants have been reported to predispose to
hypothesized a continuum, with some depressions depression through an interaction with childhood
emerging out of a high degree of stress, others occur- abuse. This was originally reported in two samples
ring in the setting of virtually no stress, and others in [70], and then replicated in a third, but not in a fourth
between. Complicating this scenario are the likeli- sample [100]. Additional work will be needed to clarify
hood that the HPA axis, which mediates stress, is these results.

224
Chapter 18: Genetics of major depression

from the STAR*D sample. The strongest association


Pharmacogenetics of MDD was found for rs6989467 in CDH17 (p ¼ 7.6  10–7),
Pharmacogenetics is a field that first developed in the while the strongest effect, by another measure, was
1950s with clinical observations of inherited differ- observed for rs1502174 in EPHB1 (p ¼ 8.5  10–5).
ences in drug effects. In the early 1960s, a few studies No effect withstood correction for multiple testing.
of familial correlation in response to antidepressants The investigators added two innovative analyses.
were published. One such study found that of 41 pairs First, they defined a binary response allele score based
of relatives treated with the antidepressant imipra- on 46 of the strongest SNPs from the MARS sample
mine, 38 had a concordant response: both responded that also showed nominal significance in the STAR*D
in 34 pairs, neither responded in 4 pairs, and 1 sample. The odds ratio for this response allele score
responded in 3 pairs (Angst, 1961 and 1964: cited in was 2.31 (p ¼ 5  10–8) in the MARS sample and 1.90
[101]). A second study, by Pare et al. [102], found that (p ¼ 5  10–9) in the STAR*D sample. They further
in 8 relative pairs, there was concordance for 6/6 pairs performed a pathway analysis, categorizing genes as
of tricyclic antidepressant trials and for 6/6 pairs of related if they were co-cited in the same sentence of
monoamine oxidase (MAO) inhibitor antidepressant an abstract with a functional descriptor in between.
trials. Pare and Mack [101] later reported concord- Gene clusters were identified in accordance with the
ance in 10/12 new pairs of related patients treated number of co-citations of each pair of genes. This
with antidepressants from the same class, making resulted in three clusters that centered on fibronectin 1,
the total 22/24 (92%) for the Pare et al. [102] and ADAMTS-like 1, and endothelin 1 (Figure 18.3).
Pare and Mack studies [101]. By contrast, analysis of As with the genetics of MDD, the genetics of
relatives’ responses to antidepressants of different antidepressant response is likely to involve large
classes from the two studies revealed concordance in numbers of genetic varations acting jointly to influ-
just 7/18 pairs (39%). ence the phenotype. A simple tallying up of alleles
Many studies have examined antidepressant might one day be of value in a clinical setting, while
response in relation to gene variants, with most of pathway types of analysis that can provide insight into
them focusing on the same serotonin transporter gene the biochemical processes at work in the response
promoter region variant discussed above. A number may guide future studies of the physiology down-
of small studies suggested a better response to select- stream from the relevant genes.
ive serotonin reuptake inhibitors (SSRIs) in patients
with long alleles and a slower or worse response in
subjects with short alleles. However, in what is by far Conclusions
the largest study to date, investigators using the Although familial patterns of depressive illnesses have
STAR*D sample did not find such an association [103]. been noted for over 100 years, a precise description
There have been two antidepressant GWAS of their genetic mechanisms remains an elusive goal.
reports to date. The STAR*D investigators conducted We do not yet have a single defintively identified
a GWAS to systematically assess association with MDD susceptibility gene. However, the field is now
response to citalopram in a sample of 883 responders advancing rapidly and much progress should be
and 608 nonresponders [104]. They employed SNPs expected within the next couple of years. The Psychi-
from the Affymetrix 500K and 5.0 Human SNP atric GWAS Consortium is providing an infrastruc-
Arrays, and association analysis was carried out after ture in which increasingly large sample sizes are being
correcting for population stratification. No results collected. As the sample size grows into the range of
reached genome-wide significance. They identified 3 10 000–20 000 cases and a comparable number of
SNPs associated with response with p-values < 1  10–5 controls, we will acquire the power to detect the
near UBE3C (rs6966038, p ¼ 4.65  10–7), another 100 kb modest gene effects likely to play a role in suscepti-
from BMP7 (rs6127921, p ¼ 3.45  10–6), and a third bility to MDD. Large samples also provide sufficient
in RORA (rs809736, p ¼ 8.19  10–6). The Munich power to employ analytic techniques that can robustly
Antidepressant Response Signature (MARS) project test for interactions between genes.
reported a GWAS on 339 patients with a depressive While GWAS focuses on identifying common
episode [105]. Replication was sought in a further 361 variation associated with MDD, studies searching
German subjects with depression and in 832 subjects for rare variants involved in MDD etiology are just

225
Chapter 18: Genetics of major depression

M IN TF NR2F2 Cluster 2
IN PIK3R1
IN CUGBP2

IN APOB

M IN PON2

IN CD36
M IN GSTP1 M IN EPHX2

IN SPON1

IN ADAMTSL1 M IN RYR2 IN HOMER1

Cluster 1
IN SLC8A1 M IN KCNJ6
IN HAS2

IN EDN1 M IN TNFRSF1A IN DAPK1


M IN PTPRK M IN PTGER4
M IN SLC1A2
IN TF MYBL2
IN A2BP1 M IN AKR1B1 IN TNFRSP21
IN SLC6A11
IN TF MITF
IN NRG1
IN PTPRS M IN FN1
IN SCNN1A
M IN TF NR2E1 IN CADM1
M IN RPS6KA1 Cluster 3
M IN RAP1B

M IN TGFBR3 IN TPM1
IN CALD1

IN SERPINA4
IN EFNAS

IN EPHAS

Figure 18.3 Results of a literature-based pathway analysis that includes all genes from the Munich Antidepressant Response Signature
project genome-wide association study (GWAS) that correspond to the single nucleotide polymorphisms (SNPs) implicated by the STAR*D
replication GWAS sample. Genes were categorized as related when they were co-cited in the same sentence with a functional descriptor
in between. There are 41 genes that cluster around fibronectin 1 (FN1) (cluster 1), ADAMTS-like 1 (ADAMTSL1) (cluster 2), and endothelin 1
(EDN1) (cluster 3). Genes with corresponding SNPs that achieved nominal significant replication in the STAR*D sample are shaded in red; green
lines indicate transcription factor (TF) binding site matches in target promoters; the line with the yellow circle indicates annotation by
Molecular Connections experts. IN indicates input gene; M, part of a metabolic pathway. From [105], with permission. See plate section for
color version.

beginning, making use of powerful next-generation all genes, and potential tissue-specific regulation of
sequencing technology. Such studies can be targeted gene expression may limit their usefulness, the exist-
to particular linkage regions or particular gene sets. ence of such cell lines in the very MDD subjects
As the cost of performing such studies is coming who show association can provide investigators with
down rapidly, studying the whole exome, or even a useful means of studying the expression of some
the whole genome, in this way will soon be feasible. candidate genes.
Findings of statistical association between a gene Epigenetic studies should advance our under-
variant and MDD will need to be complemented by standing of the mechanisms that underlie gene
functional studies of the relevant gene and its poten- expression. As these epigenetic marks are likely to
tially pathogenic alleles. For example, postmortem be partly inherited and partly acquired, they might
brain samples from MDD subjects are a valuable tell us something about the so-called “missing herit-
resource for testing gene expression levels. Brain ability” in MDD, and also something about the bio-
tissue is, of course, not available on subjects who are logical impact of stressful life events. Genome-wide
studied for association, but white blood cells are. studies of DNA methylation variation and of histone
Although these cells express only about one-half of modifications in MDD are ongoing.

226
Chapter 18: Genetics of major depression

Approaches that emphasize clinical insights into could allow clinicians to optimize the use of already
the depressive phenotype may provide increased effective medications by choosing the drug most likely
power by creating more genetic homogeneity. Clinical to work for a given patient with a particular genetic
subtypes such as postpartum depression and comor- profile. The work of establishing these predictors of
bid anxiety, as well as potential endophenotypes response will benefit from the existence of large net-
including variations in HPA function and brain works that can rapidly implement pharmacogenetic
region volumes in patients with MDD and their rela- trials of existing and newly emerging antidepressants.
tives may allow for useful alternative methods for The recently created National Network of Depression
identifying genetic effects. It may be possible to study Centers, involving 21 US academic medical centers, is
genotype–endophenotype correlations. If a gene is one such example.
involved in MDD, it should be possible to demon- Perhaps the most fundamental benefit will come
strate the effect the disease allele has on, for example, from an improved understanding of the pathophy-
HPA abnormalities and reduced hippocampal size. siology of MDD. Finding a gene will lead to an exam-
Given the likelihood of relatively modest effects for ination of the functions of its protein product in the
MDD susceptibility alleles, a conclusive case for the neuron. This examination could lead in turn to the
involvement of an implicated gene may require dem- elucidation of a cascade of neuronal events at work in
onstration of a pathogenic pathway. the disorder. Understanding of these basic processes
There are currently no laboratory methods for would guide the search for novel treatments of MDD.
diagnosing MDD; rather, these diagnoses rest solely The gene product or another protein with which the
on clinical data. Identification of a set of causative gene product interacts could be a target for conven-
variants could clarify the diagnostic process by provid- tional pharmacology, such as receptor blockade or
ing physical evidence of the disorder. Presymptomatic inactivation of an enzyme. Alternatively, an overex-
testing could lead to preventive treatments. Greater pressed gene could be knocked down with siRNA.
precision in prognosis may come from genotype– Work in this area is advancing rapidly [106].
phenotype correlations, whereby particular symptom The next decade should be an exciting one for
clusters and natural course are found to be associated the genetics of MDD as the pieces are all in place
with specific combinations of gene variants. for making enormous progress. This will benefit
Pharmacogenetic studies could also provide excit- researchers who should soon have definitive genetic
ing clinical benefits from genetic investigations of variants involved in MDD that can provide anchor
MDD. If the illness is genetically heterogenous, there points for further work elucidating the role of these
may be some genetic vulnerabilities associated with variants in disease. More importantly, it should pro-
good SSRI response, and other genetic vulnerabilities vide a starting point for discerning ways in which to
associated with a better response to noradrenergic make these results relevant for improving the care of
medications such as nortriptyline. These findings patients suffering with this dreaded disorder.

References 7. Tsuang MT, et al. Br J Psychiatry


1980;137:497–504.
14. Maier W, et al. Arch Gen
Psychiatry 1993;50:871–883.
1. Kraepelin E. Manic Depressive
Insanity and Paranoia. Edinburgh: 8. Abrams R, et al. Am J Psychiatry 15. Weissman MM, et al. Arch
Livingstone; 1899/1921. 1980;137:658–661. Gen Psychiatry 1993;50:
9. Jakimow-Venulet B. Br 767–80.
2. Perris C. Acta Psychiatr Scand
Suppl 1966;194:15–44. J Psychiatry 1981;139:450–456. 16. Sullivan PF, et al. Am J Psychiatry
10. Gershon ES, et al. Arch 2000;157:1552–1562.
3. Reich T, et al. Am J Psychiatry
1969;125:1358–1369. Gen Psychiatry 1982;39: 17. Lyons MJ, et al. Arch Gen
1157–1167. Psychiatry 1998;55:468–472.
4. Andreasen NC, et al. Arch
Gen Psychiatry 1987;44: 11. Baron M, et al. Acta Psychiatr 18. Bland RC, et al. Arch Gen
461–469. Scand 1982;65:253–262. Psychiatry 1986;43:1085–1089.
5. Winokur G, et al. Arch Gen 12. Weissman MM, et al. Arch Gen 19. Levinson DF. Biol Psychiatry
Psychiatry 1995;52:367–373. Psychiatry 1984;41:13–21. 2006;60:84–92.
6. Smeraldi E, et al. Acta Psychiatr 13. McGuffin P, et al. J Psychiatr Res 20. Bertelsen A, et al. Br J Psychiatry
Scand 1977;56:382–398. 1987;21:365–375. 1977;130:330–351.

227
Chapter 18: Genetics of major depression

21. Torgersen S. Arch Gen Psychiatry 45. McGuffin P, et al. Hum Mol Genet 68. Board F, et al. Psychosom Med
1986;43:222–226. 2005;14:3337–3345. 1956;18:324–333.
22. Andrews G, et al. J Affect Disord 46. Holmans P, et al. Am J Hum Genet 69. van West D, et al. Neuro-
1990;19:23–29. 2004;74:1154–1167. psychopharmacology 2006;
23. McGuffin P, et al. Arch Gen 47. Holmans P, et al. Am J Psychiatry 31:620–627.
Psychiatry 1996;53:129–136. 2007;164:248–258. 70. Bradley RG, et al. Arch Gen
24. Bierut LJ, et al. Arch Gen 48. Levinson DF, et al. Am Psychiatry 2008;65:190–200.
Psychiatry 1999;56:557–563. J Psychiatry 2007;164:259–264. 71. Lekman M, et al. Biol Psychiatry
25. Kendler KS, et al. Arch Gen 49. Camp NJ, et al. Am J Med Genet B 2008;63:1103–1110.
Psychiatry 2000;57:94–5. Neuropsychiatr Genet 2005; 72. Binder EB, et al. Nat Genet
135B:85–93. 2004;36:1319–1325.
26. Kendler KS, et al. Am J Psychiatry
2006;163:109–114. 50. Mahon PB, et al. Am J Psychiatry 73. Willour VL, et al. Mol Psychiatry
2009;166:1229–1237. 2009;14:261–268.
27. Mill J, et al. Mol Psychiatry
2007;12:799–814. 51. Mondimore FM, et al. Am 74. Sullivan PF, et al. Mol Psychiatry
J Psychiatry 2006;163:1554–1560. 2009;14:359–375.
28. Kaminsky ZA, et al. Nat Genet
2009;41:240–245. 52. Zubenko GS, et al. Am J Med 75. Muglia P, et al. Mol Psychiatry
Genet B Neuropsychiatr Genet 2010;15:589–601.
29. Falconer DS. Ann Hum Genet 2004;129B:47–54.
1965;29:51–76. 76. Zhou X, et al. PLoS One 2009;
53. Nurnberger JI, Jr., et al. Am 4:e5196.
30. Wender PH, et al. Arch Gen J Psychiatry 2001;158:718–724.
Psychiatry 1986;43:923–929. 77. Zhou X, et al. Mol Psychiatry
54. Freis ED. N Engl J Med 1954; 2005;10:393–406.
31. von Knorring AL, et al. Arch Gen 251:1006–1008.
Psychiatry 1983;40:943–950. 78. Shyn SI, et al. Mol Psychiatry
55. Schildkraut JJ. Am J Psychiatry 2011;16:202–15.
32. Cadoret RJ, et al. J Affect Disord 1965;122:509–522.
1985;9:155–164. 79. Pilc A, et al. Biochem Pharmacol
56. Lopez-Leon S, et al. Mol 2008;75:997–1006.
33. Tanna VL, et al. Neuro- Psychiatry 2008;13:722–785.
psychobiology 1976;2:52–62. 80. Palucha A, et al. Psycho-
57. Lesch KP, et al. Science pharmacology (Berl) 2007;
34. Tanna VL, et al. Neuropsychobiology 1996;274:1527–1531. 194:555–562.
1979;5:102–113. 58. Duman RS, et al. Arch Gen 81. Wellcome Trust Case Control
35. Tanna VL, et al. J Psychiatr Res Psychiatry 1997;54:597–606. Consortium. Nature 2007;
1989;23:99–107. 59. Smith MA, et al. Proc Natl Acad 447:661–678.
36. Wilson AF, et al. Biol Psychiatry Sci U S A 1995;92:8788–8792. 82. Sullivan PF, et al. Am J Med Genet
1989;26:163–175. 60. Siuciak JA, et al. Pharmacol B Neuropsychiatr Genet 2006;
37. Crowe RR, et al. Neuro- Biochem Behav 1997;56:131–137. 141B:261–268.
psychobiology 1981;7:20–25. 61. Karege F, et al. Psychiatry Res 83. Altar CA, et al. Neuro-
38. Hill EM, et al. Biol Psychiatry 2002;109:143–148. psychopharmacology 2009;
1988;24:903–917. 62. Brunoni AR, et al. Int J 34:18–54.
39. Cox N, et al. J Psychiatr Res Neuropsychopharmacol 84. Chourbaji S, et al. Prog Brain Res
1989;23:109–123. 2008;11:1169–1180. 2008;167:65–77.
40. Zubenko GS, et al. Mol Psychiatry 63. Schumacher J, et al. Biol 85. Ridder S, et al. J Neurosci 2005;
2002;7:460–467. Psychiatry 2005;58:307–314. 25:6243–6250.
41. Nibuya M, et al. J Neurosci 64. Licinio J, et al. Arch Gen 86. Paykel ES, et al. Arch Gen
1996;16:2365–2372. Psychiatry 2009;66:488–497. Psychiatry 1969;21:753–760.
42. Dowlatshahi D, et al. Lancet 65. Verhagen M, et al. Mol Psychiatry 87. Kendler KS, et al. Am J Psychiatry
1998;352:1754–1755. 2010;15:260–271. 1995;152:833–842.
43. Dwivedi Y, et al. Arch Gen 66. Verma R, et al. Biol Psychiatry 88. Weaver IC, et al. Nat Neurosci
Psychiatry 2003;60:273–282. 2008;63:1185–1189. 2004;7:847–54.
44. Abkevich V, et al. Am J Hum 67. Feng Y, et al. Am J Psychiatry 89. McGowan PO, et al. Nat Neurosci
Genet 2003;73:1271–1281. 2008;165:610–616. 2009;12:342–348.

228
Chapter 18: Genetics of major depression

90. Roth TL, et al. Biol Psychiatry 96. Schroeder FA, et al. 101. Pare CM, et al. J Med Genet
2009;65:760–769. Biol Psychiatry 2007; 1971;8:306–309.
91. Milutinovic S, et al. Carcinogenesis 62:55–64. 102. Pare CM, et al. Lancet 1962;
2007;28:560–571. 97. Caspi A, et al. Science 2:1340–1343.
92. Shimabukuro M, et al. Behav 2003;301:386–389. 103. Kraft JB, et al. Biol Psychiatry
Brain Funct 2006;2:37. 98. Risch N, et al. JAMA 2009; 2007;61:734–742.
93. Tsankova NM, et al. Nat Neurosci 301:2462–2471. 104. Garriock HA, et al. Biol Psychiatry
2006;9:519–525. 99. Karg K, et al. Arch Gen Psychiatry 2010;67:133–138.
94. Lee MG, et al. Chem Biol 2011;68(5):444–454. 105. Ising M, et al. Arch Gen Psychiatry
2006;13:563–567. 100. Polanczyk G, et al. Arch 2009;66:966–975.
95. Cassel S, et al. Mol Pharmacol Gen Psychiatry 2009; 106. Whitehead KA, et al. Nat Rev
2006;70:487–492. 66:978–985. Drug Discov 2009;8:129–138.

229
The genetics of schizophrenia
Chapter

19 Hugh M. D. Gurling and Andrew McQuillin

Summary have been perturbed by the disease mutations. Trans-


genic mice with behavioral phenotypes that mimic
Schizophrenia is a common disorder which has a high aspects of schizophrenia have been created for the
heritability. Incomplete penetrance of schizophrenia calcineurin (PPP3CC), proline dehydrogenase 2
susceptibility alleles has frequently been observed. (PRODH2), disrupted in schizophrenia (DISC1),
Therefore there are likely to be many unaffected microtubule-associated stable tubule only polypeptide
carriers of these alleles in the population. Tests of (STOP), and neuregulin 1 (NRG1) genes. Cell biology
genetic linkage and allelic association have proven studies to explore pathogenesis have begun for NRG1,
that schizophrenia is extremely heterogeneous with DISC1, and PCM1 using human cell lines and zebra-
many different low frequency disease alleles causing fish. Hundreds of other genes and their proteins will
susceptibility. Recent genome-wide allelic association need to be studied using these methods in order to
studies have further demonstrated heterogeneity even achieve a substantial understanding of the genetic
when comparing sub-samples within a single ances- etiology of schizophrenia.
tral population. In addition it has been shown that a
proportion of schizophrenics have a genetic suscepti-
bility from deletion and insertion mutations, called Introduction
copy number variants (CNVs), some of which are Traditional approaches used in medical genetics, such
de novo and some of which are genetically transmit- as population prevalence, family, twin, and adoption
ted. Because of the presence of extreme heterogeneity studies have shown a very high heritability for schizo-
it will only be possible to determine which modes of phrenia with very little or absent role for the family
genetic transmission exist for schizophrenia by iden- environment. Studies of sex distribution, age of onset,
tifying the actual etiological base pair changes in and clinical variation have been incorporated into
genes implicated by linkage and association studies the family and twin studies. One of these, known
or by defining CNVs. The first few plausible etio- as “anticipation” is where the age of onset becomes
logical base pair changes affecting the DISC1, PCM1, earlier and the severity of symptoms more severe in
ABCA13, and SYNGR1 genes causing schizophrenia the later generations of multiply affected families.
have now been reported. Deletions and duplications We now know that the previous attempts using family
are responsible for a proportion of the genetic suscep- data and segregation analyses to establish the modes
tibility to schizophrenia and there is striking overlap of transmission for schizophrenia were flawed
for some of these mutations with those implicated in because of the underlying heterogeneity. Nevertheless
epilepsy. these older studies are important because they dem-
Pharmacogenomic strategies for the design of new onstrated the great degree of genetic pleiotropism or
drug treatments have now become of great import- variation in phenotypic expression that can arise from
ance because of the identification of susceptibility the same genetic susceptibility within families and
genes. New drugs can be designed by targeting the within identical twins. There is still controversy about
pathogenetic systems with the directionality of drug whether bipolar disorder and schizophrenia can share
effects determined by how the various brain systems the same susceptibility alleles and to what extent they

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

230
Chapter 19: Genetics of schizophrenia

are different disorders. Clarity on this issue has been and association studies can be carried out at much
hampered by the fact that diagnosis of schizophrenia greater speed and schizophrenia has benefitted from
in early-onset cases is often inexact due to the fact that the new microarray genotyping, CNV and sequencing
severe bipolar mania can present with schizophrenic methods. The traditional medical genetic approaches
symptoms. of family and twin studies will need to be carried out
Linkage studies have dominated efforts to find again once the actual etiological DNA base pair
the genes causing schizophrenia for the last 20 years. changes have been detected. Below, the various trad-
The logarithm of odds (LOD) method approach itional and more recent molecular approaches are
requires multiply affected families and the genotyping described in greater detail.
of DNA markers in all the affected and unaffected
members of families. The sib-pair method does not
require the genotyping of unaffected members. The
Age of onset, anticipation, and
LOD method of localizing genes by linkage analysis epigenetic effects in schizophrenia
in families can narrow down any susceptibility genes Age of onset variation is known to be an important
to one of several hundred or several thousand in a consequence of underlying genetic pathology. It
chromosomal region of up to 40 million base pairs appears to be a variable which is well defined, but
of DNA. The LOD score is a statistical test of linkage age of onset is very prone to selection biases when
between schizophrenia to a specific chromosomal comparing cases and valid conclusions may be elusive
region as defined by genetic markers within a sample [1, 2]. Correlations in age at onset between relatives
of families. A LOD of 3.00 (p ¼ 0.001) in two inde- can be used to predict risk to relatives at a given age.
pendent samples has been used as the criterion Kendler and MacLean [3] corrected for various biases
for acceptance of a confirmed linkage. This approach and found that ages at onset for schizophrenia were
was slow to get off the ground in schizophrenia positively correlated in pairs of affected relatives.
research because early linkages took many years to An early age at onset was associated with higher risk
confirm because of the presence of disease locus of illness in siblings and nieces/nephews but not in
heterogeneity. However, these studies eventually children. Neale et al. [4] studied age of onset in a
became very successful with many confirmed link- schizophrenia twin sample. They found that the age
ages with two LOD scores above 3.00 emerging after of onset was correlated within twin pairs. In the
sufficient numbers of family samples had been Roscommon schizophrenia family study, age of onset
studied. was earlier in hebephrenic and catatonic subtypes
The next phase of identifying susceptibility genes compared to the later onset paranoid schizophrenic
employed methods to detect the evolutionary phe- subtype [3].
nomenon of linkage disequilibrium (LD) in unrelated The relationship between age of onset in parental
cases of schizophrenia and normal unrelated controls and offspring generations and genetic effects has been
rather than cases within families. Linkage disequilib- a focus in schizophrenia research for many years.
rium is a generalized finding across the whole human Penrose found what is referred to as “anticipation”
genome and means that genetic variants or mutations in schizophrenia by collecting age of onset data from
that are close to each other tend to be inherited by the a large family sample [5]. He found strong evidence
same individual over evolutionary time. In practice for reduced age of onset in the next generation in
this means that specific marker alleles are found in parent–offspring pairs who had both developed
increased frequency in cases compared to controls. schizophrenia. He also found increasing severity of
This phenomenon is called allelic association and is the illness in successive generations. Penrose was not
caused LD. Linkage disequilibrium implies that during entirely convinced that he had a true finding. How-
evolution, when a new susceptibility mutation arose, ever parent–offspring pairs and aunt/uncle–niece/
it would have occurred in an individual who already nephew schizophrenia pairs both showed strong
had pre-existing genetic variants near the mutation. evidence for anticipation suggesting that selection
From then on as the disease mutation is passed down biases for types of families could not easily account
the generations it will “hitch hike” or co-segregate for anticipation. Anticipation effects were greatest in
with other genetic variants nearby. Advances in pairs with parents who had a late age of onset but
genetic technology now mean that genetic linkage anticipation was also significant in early age of onset

231
Chapter 19: Genetics of schizophrenia

parents as well (p ¼ 0.03). A review of Penrose’s study stable within individuals than is bipolar disorder
[6] concluded that it was a valid finding. The findings which is often misdiagnosed as schizophrenia in
supported further investigations of anticipation early-onset cases particularly in young males [23].
and trinucleotide repeat expansion mutations as well Tsuang [24] found that 92.5% of schizophrenics had
as investigation of methylation abnormalities in the a stable diagnosis over a 30–40 year period whereas
genetics of schizophrenia. only 78.3% of bipolar cases had stable diagnoses. This
Bassett and Honer [7] studied three-generation will mean that more cases of bipolar disorder are
affected families selected for unilineal, autosomal likely to be incorrectly identified as cases of schizo-
dominant-like inheritance. They found strong evi- phrenia than vice versa. All family, twin, and adop-
dence that more subjects were hospitalized with tion studies have observed that the incidence of the
psychotic illness in the youngest generation, that the illness in the relatives of schizophrenic probands is
age of onset became progressively earlier in the later considerably greater than in suitable family or popu-
generations, and that there was increasing severity lation controls [25–34].
of illness. In the study by Asherson et al. [8] of 29 Given that we now know that schizophrenia is
multiply affected pedigrees, an earlier age of onset in highly heterogeneous an important clinical question
later generations was found but all selection biases is to ask to what extent different subtypes of schizo-
could not be excluded. Other studies in multiple phrenia can recur within the same families. Family
ancestral populations also found anticipation [9–19]. studies have shown that, within a family, the same
McInnis [20] reviewed the field and concluded that it disease alleles can give rise to clinical diversity or
was a genuine effect. The main uncertainty in the pleiotropism within the schizophrenia spectrum [31,
study of anticipation is that age of onset for schizo- 35]. The corollary to this is that clinical variability
phrenia may have been getting earlier in the later does not imply underlying genetic heterogeneity
decades [12] purely as an artifact of early intervention within an individual. Although there is little evidence
and increasing awareness of mental illness. However, for the proband schizophrenia subtype to breed true
Penrose’s [12] study was carried out in 1945 and [36, 37] other work has suggested a tendency towards
anticipation has been found in a variety of different homotypia within families for the paranoid and hebe-
cultures. phrenic subtypes [38, 39]. Some families have been
reported with multiple cases of catatonic schizo-
phrenia [40]. Others have put forward the idea that
Familial recurrence of schizophrenia the hebephrenic form of schizophrenia may be more
Behavior genetic approaches have been used to obtain heritable than the paranoid variety [41, 42]. Tsuang
evidence for the relative importance of cultural, and Winokur [38] have proposed that there is a
family environment, and unique (specific) environ- catatonic–hebephrenic–paranoid continuum of illness
mental factors compared to the effects of genes in the with the catatonic type being the most advanced and
etiology of schizophrenia. Several complementary severe form.
behavior genetic approaches such as the twin, family, Family and twin studies[32, 43] show that the risk
and adoption methods need to be compared and for schizophrenia and schizophrenia spectrum dis-
contrasted to come to a balanced view. Genetic ana- orders was not associated with any particular subtype
lyses which are relevant to plant and animal studies, of proband. Bassett and coworkers [44] have tested
rather than human populations, have specific limita- Liddle and Barnes’s three syndrome schizophrenia
tions in relation to the genetics of schizophrenia model [45] in families with schizophrenia employing
[21, 22]. The most often cited limitation is the fact factor analysis. They were able to confirm the “psycho-
that humans do not mate randomly with each other motor poverty” factor but found a combined “reality
and therefore there is no stabilized equilibrium in the distortion/disorganization/inappropriate affect” factor
general population for the distribution of human and a “suspiciousness/stereotyped thinking” factor
traits and their genotypes. instead of Liddle and Barnes’s separate reality distor-
Valid determination of the recurrence risk for tion and disorganization factors.
psychiatric disorders in the relatives of probands is Keefe et al. [46] who carried out a family study,
dependent on accurate diagnosis. An important find- compared severely disabled Kraepelinian type schizo-
ing is that the diagnosis of schizophrenia is more phrenics with other chronic, but intermittently

232
Chapter 19: Genetics of schizophrenia

exacerbating, schizophrenics. They found a higher increased with a significance of p ¼ 0.02. These data
incidence of schizophrenia spectrum disorders in the are not strongly in favor of coaggregation of the two
relatives of the Kraepelinian group (19.3%) compared disorders in the same families and the weak evidence
to the relatives of the intermittently exacerbating is explainable by mis-diagnosis and by assortative
group (10.7%). The Kraepelinian group was also char- mating for schizophrenia and bipolar disorder in the
acterized by asymmetry of the lateral ventricles and parental generation.
a lack of response to neuroleptics. A very large population register study of the
Potential differences between familial and spor- Swedish population entitled “Common genetic deter-
adic schizophrenia were reviewed in 1994 [47]. In minants of schizophrenia and bipolar disorder in
69 studies very few replicated differences were found. Swedish families: a population-based study” claimed
One difference was the increased volume of the len- common genetic effects for schizophrenia and bipolar
ticular nuclei and greater ventricular asymmetry in disorder [57]. Both diseases were diagnosed with a
familial cases compared to sporadic cases and normal nonhierarchical diagnostic structure. Thus, an indi-
control subjects [48]. vidual could have both schizophrenia and bipolar
disorder if the person has been diagnosed with each
disease. There were 2543 such dual diagnosis cases
Schizophrenia and affective disorders accounting for 5086 (6.6%) cases out of 76 860 pro-
Slater and Cowie [36] drew attention to the fact that bands. The chance of an individual having both diag-
family studies of schizophrenia were associated with noses according to the population prevalence of both
what they called attenuated cases of schizophrenia these disorders in Sweden without adjustment for age
that appeared to be cases of neurotic depression. This is 0.0001%. Thus, in this study the expected rate of
idea was later supported by a family study which dual diagnosis is 66 000 times greater than chance.
found an increase in depression but not bipolar dis- The procedure of counting dual diagnosis probands
order in the relatives of schizophrenics [49]. Kendler twice would inflate the cross correlation between
and others found that the risk in all interviewed bipolar disorder and schizophrenia within families.
relatives of schizophrenic probands was 23.6% for The authors also did model-fitting analyses with hier-
all types of severe psychiatric disorder [33, 50–53]. archical diagnoses, in which individuals with two or
The risk for schizophrenia in the parents of probands more schizophrenia diagnoses were classified as
was much less than that found in siblings indicating having schizophrenia and not bipolar disorder, even
partial or incomplete penetrance. These studies also if they have had two or more such diagnoses. How-
showed that the schizophrenia found in these families ever this does not get rid of the problem of misdiag-
did not have any strong familial relationship with nosis because such cases could equally have been
bipolar affective disorder. counted as bipolar probands. It is difficult to explain
Some reviewers of family studies of schizophrenia the differences between the Swedish study [57] with
have drawn the conclusion that schizophrenia and the recent family meta-analysis [56] without consider-
bipolar disorder do not share the same genetic etiol- ing the possibility that the Swedish registry diagnoses,
ogies [54] whereas others argue that they often do which were essentially clinical rather than operation-
[55]. More recently in a meta-analysis of family stud- ally defined research diagnoses, were prone to
ies it was found that the increased rate of schizophre- unstable diagnosis between the two psychoses [56].
nia in the first-degree relatives (FDRs) of bipolar Many other differences between schizophrenia and
subjects was “somewhat equivocal” being marginally bipolar disorder have been found such as voxel-based
significant in the primary analysis but not significant morphometry in schizophrenia compared to bipolar
in the unweighted analysis [56]. When the rate of disorder which found consistent differences with
bipolar disorder in the FDRs of schizophrenic pro- many bipolar studies showing none of the volumetric
bands was examined the increased rate was not signifi- deficits found in schizophrenia [58]. These studies
cant (p ¼ 0.06) in both the weighted and unweighted support the notion that schizophrenia and bipolar
analyses. Only when studies that did not report disorder generally have different genetic underpin-
morbid risk data were excluded from the analyses nings. However, there are cases of schizoaffective
was it found that the risk of schizophrenia in the disorder with clinical features that do cross boundar-
first-degree relatives of bipolar probands was ies between schizophrenia and bipolar disorder but

233
Chapter 19: Genetics of schizophrenia

these cases are rare [59–62]. Maj and coworkers [61] membership and illness status on CT scan frontal
stated that DSM-III-R schizoaffective disorder was horn measurements [69]. Some authors have found
close to schizophrenia and largely corresponds to no difference between those with and those without a
“schizoaffective schizophrenic” disorder in the positive family history [1]. Others have reported sig-
Research Diagnostic Criteria (RDC) [63], whereas nificant results but the direction of the relationship
DSM-III-R mood-incongruent psychotic depression has varied, with some finding larger lateral ventricles
is heterogeneous and corresponds to what could be in the nonfamilial disorders [70–72] and others dem-
called a true “schizoaffective psychosis” state which is onstrating the reverse [73]. Owen et al. [74] described
quite rare. The implication is that the RDC system of a complex curvilinear relationship between family
diagnosis using the Schizophrenia and Affective Dis- history and VBR; the highest frequency of positive
orders Schedule (SADS) [63] is superior to DSM-III- family history scores occurred in those with medium
R and DSM-IV because it subdivides schizoaffective VBRs. Lewis [66] has commented that the distinction
disorders into two types. One is schizoaffective bipo- between familial and sporadic forms of schizophrenia
lar disorder or schizo-mania which is basically a bipo- in terms of CT brain scan appearance has become less
lar manic illness with a schizophrenic coloration tenable. A family CT scan study by Orlova et al. [75]
during acute mania and the other is schizoaffective estimated heritability for CT parameters. Width of
schizophrenia which is a chronic illness more related the frontal horn of the left lateral ventricle, width of
to schizophrenia. This concept gains support from the right lateral ventricle frontal horn, and measures
two familial recurrence studies of the families of of the right lateral horn were found to be transmitted
“schizoaffective” psychosis probands. One study genetically and also to be associated with schizophre-
carried out in New York found that schizoaffective nia. Palmour et al. [76] reported two cases of DSM-
psychosis could be blindly subdivided into either III-R schizophrenia and schizotypal disorder in a
schizoaffective mania (schizomania or schizoaffective family who both had enlarged ventricles and widened
bipolar disorder, SA-BP) or schizoaffective schizo- sulci in scanning. Each person also had a chromosome
phrenia (SA-SCZ) and that the presence of schizo- 4 inversion (inv 4 [p15.2; q21.3]). The most recent
affective bipolar disorder in the proband predicted study of familial versus nonfamilial schizophrenia pre-
the presence of bipolar disorder in the first-degree sented data which showed that familial schizophrenia
relatives whereas schizoaffective schizophrenia in the was associated with more severe structural abnormal-
proband predicted schizophrenia in the relatives [64]. ities than sporadic schizophrenia, especially in the
A second study [65] did not subdivide schizoaffective thalamus [77].
disorder into the two subtypes of SA-BP and SA-SCZ
but found either recurrent bipolar or recurrent
schizophrenia cases in the relatives of schizoaffective
probands.
Neurophysiological deficits with
family genetic control
Cognitive deficits and imaging Blackwood [78] and Holzman [79] have argued for
augmenting the phenotype of schizophrenia to include
changes in family genetic studies smooth pursuit eye movement studies, evoked electro-
X-ray computed tomography (CT) brain imaging encephalogram (EEG) response and cognitive changes.
studies have demonstrated enlargement of the lateral This extended phenotype may then form the basis for
ventricles and the third ventricle, enlargement of the family genetic linkage studies [80]. Holzman et al. [81,
cortical sulci and cortical atrophy. The relationship 82] examined pursuit eye movement dysfunctions in
between CT scan appearance and family history of schizophrenia and found evidence for familial recur-
schizophrenia has been a subject of controversy in the rence in specific families. Siegel et al. [83] studied
past [66]. Weinberger et al. [67] demonstrated that inhibitory gating of auditory-evoked responses in
schizophrenics have larger ventricle-brain ratios schizophrenic patients, their FDRs, and normal sub-
(VBRs) than their healthy siblings. DeLisi et al. [68] jects. The expected deficit in inhibition in people with
investigated families containing two or more siblings schizophrenia was found. Approximately half the
with a history of schizophrenia. A two-way analysis FDRs and one parent had a similar deficit. Deficit in
of variance found significant effects from family the parents was associated with a family history of

234
Chapter 19: Genetics of schizophrenia

schizophrenia. Deficits in P300 and other long latency of 58 : 12 and taking a weighted average from their own
event-related potentials were found in 65 schizo- and four other methodologically similar studies found
phrenic subjects and these had no relationship with a concordance ratio of 46 : 14. The discordance rate of
family history. Auditory P300 and eye tracking dys- over 50% in monozygotic twins supports the import-
function in schizophrenic pedigrees have been studied ance of epigenetic, random and nongenetic biological
in 20 high-density schizophrenic families [78, 84]. or specific environmental effects but which exclude
Abnormalities in one or both measures were found the family or common shared twin environment. Simi-
to be correlated with psychiatric illness in most of the lar MZ/DZ concordance ratios were found by other
families. Normal relatives also had abnormalities. workers using the narrow Research Diagnostic Criteria
Comparison of auditory sensory gating, hippocampal (RDC), DSM-III and Feighner Criteria [42, 116].
volume, and catecholamine metabolism in cases In a study of the Maudsley twin series, Farmer
versus unaffected sibling pairs has been carried out et al. [116] found that the maximum difference in
[85–88]. Abnormal auditory sensory gating was MZ/DZ concordance (i.e. the highest index for herit-
found in all schizophrenics in the 11 families studied ability) was for a definition of being affected which
and in 8 of their 20 siblings. Compared with the included all psychoses that exhibited mood incongru-
schizophrenics, the clinically unaffected siblings with ent delusions, schizotypal personality and atypical
abnormal auditory gating had larger hippocampal psychosis, thus suggesting a common genetic origin
volumes. Cannon et al. [89] utilized a similar design for this spectrum of disorders. Onstad et al. [117]
and found that both schizophrenics and their non- found concordance rates of 48% for MZ twins and
schizophrenic siblings were impaired neuropsycholo- 4% for DZ twins when DSM-III-R criteria were used
gically compared to normal controls. These and other to diagnose schizophrenia. Nonaffective psychotic
results strongly support the hypothesis that impaired disorder such as schizotypal and paranoid personality
information processing and neurophysiological def- disorders were observed in both MZ and DZ co-twins.
icits aggregate in the family members of schizophren- Torgersen et al. [118] studied 176 nonschizophrenic
ics and may serve as an indicator of partial penetrance co-twins and other FDRs of schizophrenic probands
to the disorder [90–93]. and compared them to control twins. Schizotypal
Other studies have demonstrated a deficit in personality disorders were found to be more common
early visual processing in clinically unaffected FDRs among the biological relatives of schizophrenic pro-
of patients with schizophrenia [94, 95]. Functional bands. Odd speech, inappropriate affect, odd behav-
magnetic resonance imaging (MRI) studies have been ior, and excessive social anxiety, were significantly
carried out in cases and their apparently healthy more common among the relatives of schizophrenic
siblings [96, 97] who showed altered activity and con- probands. A biometrical approach to the genetic and
nectivity [98, 99]. Many other endophenotypic deficits environmental effects on schizophrenia using twin
have been found in unaffected siblings [97, 100–112]. data has shown that the family environment effect
Every study finds the same pattern, namely, that normal in the susceptibility to schizophrenia was less than
siblings and relatives of schizophrenic probands who 1% and that the genetic component was 93% [119].
are not diagnosable as cases of schizophrenia but likely Other heritability estimates using twin data were
carriers of schizophrenia susceptibility genes have def- between 67 and 80% [120, 121]. This implies that
icits in multiple imaging, neurophysiological, and psy- the role of mothering and fathering in causing schizo-
chological modalities [113, 114]. phrenia is virtually absent.

Twin studies Co-twin control studies


Higher concordance in monozygotic twins as opposed A very sensitive method of demonstrating structural
to dizygotic twins suggests a genetic effect on schizo- brain changes, cognitive and neurophysiological def-
phrenia. Gottesman and Shields used systematic ascer- icits in schizophrenia is through the use of an MZ
tainment by selecting their twin probands from the twin as a control. Sophisticated measurements of
Maudsley Twin Register where all patients who were brain morphometry can be made with MRI and CT
twins were consecutively ascertained [115]. They found in MZ twins. This method also controls for age and
an monozygotic/dizygotic (MZ/DZ) concordance ratio for polygenic background effects.

235
Chapter 19: Genetics of schizophrenia

CT studies performed with this design show Suddath et al. [99] studied identical twins and
consistent increases in ventricle brain ratio in schizo- described reduced temporal lobe volume as well as
phrenia, and MRI studies document reductions reduced temporal lobe gray matter in schizophrenia.
in temporal lobe structures as well [35, 67, 68, 71, In 12 out of the 15 discordant pairs, the twin with
98, 99, 122, 123]. In twin studies of normal and schizophrenia could be identified by visual inspection.
schizophrenic twin pairs [124, 125] there was a very In two pairs no difference could be seen and in one
high within-twin pair correlation for ventricle brain twin the schizophrenic was misidentified. Quantita-
ratio in healthy MZ twins which was nonsignificant in tive analysis of the hippocampus showed it to be
DZ twins. However, when one MZ twin was schizo- smaller on the left in 14 of the 15 affected twins,
phrenic and the other normal, the correlation as compared with their normal twins, and smaller
between VBRs was reduced with the schizophrenic on the right in 13 affected twins. Another study
twins having the larger ventricles. This has been taken of discordant MZ twins [139] found a focal decrease
to indicate that, in MZ twins discordant for schizo- in gray matter density accompanied by a focal
phrenia, other stochastic or environmental forces are increase in white matter density. Another three stud-
operating to cause both the illness and the increased ies [140–142] found that variation in hippocampal
ventricular size. Suddath et al. [99] used MRI to and ventricular volumes within discordant monozy-
examine MZ twins discordant for schizophrenia and gotic pairs was observed. Hippocampal volumes of
found that, as a group, the affected twins had larger probands were smaller than those of their nonschizo-
ventricles and smaller hippocampi than the phrenic MZ and DZ co-twins and healthy twins.
unaffected twins. Goldberg et al. [126–135] studied Hippocampal volumes of unaffected co-twins were
discordant MZ twins and found within pair correl- smaller than those of healthy twins, but those of
ations between left hippocampal blood flow and unaffected MZ and DZ co-twins were similar.
verbal memory and between Wisconsin card sorting A longitudinal five-year follow-up in MZ and DZ
and prefrontal flow. Co-twin control studies of twin pairs discordant for schizophrenia and healthy
obstetric complications in schizophrenia were carried comparison twin pairs using brain MRI was carried
out by Gottesman and Shields which were not sup- out by Brans et al. [143]. Significant decreases over
portive of an effect on schizophrenia [115]. Cantor- time in whole brain and frontal and temporal lobe
Graae et al. have carried out a similar study in which volumes were found in patients with schizophrenia
they examined the whole neonatal period for neuro- and their unaffected co-twins compared with control
logical hard and soft signs in 22 MZ twin pairs dis- twins. Sommer et al. [144] studied language laterali-
cordant for schizophrenia and 7 normal comparison zation and found that it was decreased in discordant
MZ twin pairs [136]. The neurological impairment twin pairs compared with the healthy twin pairs.
that was found in the healthy discordant MZ twins Functional imaging for verbal tasks in discordant
was significantly related to history of both neonatal twins with schizophrenic probands has been carried
and total obstetric complications. Family history of out [145]. All subjects with schizophrenia showed
psychosis, history of substance abuse and history of increased activation in the right homologue of Broca’s
postnatal cerebral trauma were not related to degree area in contrast to healthy individuals. Ettinger et al.
of neurological impairment in the ill or healthy co- [146] found that thalamic volume was reduced in cases
twins. In the same sample McNeil et al. [137, 138] compared to normal cotwins and that the reduction in
investigated obstetric complications (OCs) during volume was greater in concordant twins than in dis-
pregnancy, labor and the neonatal period. Significant cordant twins. Likewise the unaffected twin had more
differences in OC rates were found in discordant pairs volumetric reduction than normal controls. When gray
compared to normal control pairs. Labor complica- matter volume was examined in the same sample of
tions were more frequent in discordant than concord- twins; frontal, insular, cingulate, medial, parietal, and
ant pairs. OC rates were equal in affected compared to temporal cortical gray matter changes were found that
healthy discordant twins. Trauma at the time of labor correlated within pairs with diagnosis [147].
and delivery and especially prolonged labor appeared A review and meta-analysis of functional scanning
to be of importance for brain structure anomalies with MRI positron emission tomography (PET) and
associated within pairs of twins discordant for single photon emission computed tomography
schizophrenia. (SPECT) in case-control studies, in twins and in cases

236
Chapter 19: Genetics of schizophrenia

with family controls, found that vulnerability to similar results. In the combined national sample of
psychosis was associated with medium to large effect Copenhagen and provincial adoptees, Kety et al. com-
sizes when prefrontal activation was contrasted with ment that chronic schizophrenia in adoptees was
that in controls [148]. Relatives of patients affected found exclusively in the biological relatives of schizo-
with psychosis, the co-twins of patients and subjects phrenics and that the overall prevalence of schizo-
with an “at risk mental state” shared similar neuro- phrenia in adoptees was 10 times greater than the
cognitive abnormalities. The prefrontal and anterior rate in the biological relatives of controls. Later these
cingulate cortex, the basal ganglia, hippocampus, and studies were reviewed and it was noted that these
cerebellum were most affected. Other co-twin control adoption studies showed genetic transmission of
studies [130–133, 149–154] confirm these findings. schizophrenia, nonaffective psychosis, and schizoty-
Clearly within pair differences need an explanation pal rather than any general liability to other forms of
and the recent finding of genomic imprinting differ- psychopathology [161]. Cardno also reviewed the
ences within MZ twin pairs as reflected in differential adoption studies and found that the risk of chronic
methylation effects on DNA is a possible explanation schizophrenia in relatives was predicted by the pres-
[155, 156]. Such effects could combine with other ence of the pervasive negative symptoms of social
random stochastic genetic and biological processes withdrawal, autistic behavior, poverty of thought/
or in utero biological effects to produce discordance. speech, and flat affect and absence of pervasive posi-
tive symptoms [162].
Tienari et al. [163, 164] have carried out adoption
Adoption studies studies of schizophrenia in Finland. The sample was
Adoption and cross-fostering studies can clarify the obtained from a national register. The children of
respective contributions of the genes and environ- schizophrenics given up for adoption were compared
ment because they are more clearly separated than with matched controls who were adopted offspring
in the twin method. The most prominent adoption of nonschizophrenic biological parents. The adoptive
studies come from a Danish–American collaboration families and biological parents were investigated with
which used the Danish National Adoption Register interviews and psychological tests. Among the 155
[43, 157–160]. The studies have been conducted on index offspring, the percentage of schizophrenia and
a Copenhagen sample, then on provincial samples, other psychoses, borderline syndrome, and severe
and finally combined into a national sample. All the personality disorders, was significantly higher than
studies have shown that the adopted-away children of in the 186 matched control adoptees. In adoptive
schizophrenics had a significantly higher risk of families rated as disturbed, genetic effects were more
developing schizophrenia or schizotypal disorder prominent, thus giving an impression that the envir-
and related conditions than the adopted-away chil- onment interacted with the genetic susceptibility to
dren of controls. Most of the children were born increase liability to schizophrenia.
before the first episode of illness in the parent and
about a third of the schizophrenic parents were male,
thus weakening arguments that the schizophrenia Segregation analysis and variance
may have been caused by early mother–child inter-
action or by an intrauterine event. The data obtained
component analysis of twin and
by Kety et al. [158] early in the Copenhagen data family data
found an increased incidence of schizophrenia and As discussed in the twin study section above, data
related disorders in the biological relatives of schizo- from these studies gave heritability estimates of
phrenics who had been adopted (20.3%) compared to between 66 and 93%. Family data indicate that up to
the incidence in adoptive relatives and relatives of 93% of the variance in the etiology of schizophrenia
controls (5.8%). The data were reanalyzed using may be genetic in origin [119, 120, 165–168]. McGue
DSM-III criteria [27] and a 13.5% incidence for et al. [167] estimated the family common environ-
schizophrenia or schizotypal personality in the bio- mental effect to be about 19%. McGuffin et al. [168]
logical relatives of schizophrenics was found com- suggest that stochastic processes may be partly
pared to 1.5% in the controls. In 1994, Kety et al. responsible for the nonfamilial or specific environ-
[160] reported on the provincial sample and found mental source of etiological variance. Henderson [22]

237
Chapter 19: Genetics of schizophrenia

also noted that the specific environmental variance enzymes, peptides and receptors. The third approach
contains an error term, thus potentially diluting the is to detect functional variations in proteins through
amount of true specific environmental variance in the identification of genetic variation at the DNA level
schizophrenia. Nevertheless, the specific environmen- relevant to the functioning of the resultant protein.
tal component may include a proportion of schizo- The following sections describe a selection of
phrenia being due to viral infection, obstetric genes that have been implicated in susceptibility to
complications or other nongenetic causes of brain schizophrenia by cytogenetic, linkage and/or associ-
disorder, but it could be argued that the data as a ation studies. They do not represent every gene impli-
whole shows that the bulk of the schizophrenias are cated by these studies. They do however provide the
genetic in origin. reader with a flavor of the complexity of the issues
Because of the strong evidence of the genetic surrounding the field and indicate where progress has
origins of schizophrenia, attempts have been made been made.
using segregation analysis to determine a mode of
transmission. Baron [169] has reviewed some of this
work and commented that there was considerable
Chromosome 1: NOS1AP, UHMK1,
variation in methodology between studies and con- RGS4, DISC1
flicting results. Baron found that the single major A combined allelic association, cytogenetic and
locus model was compatible with the data in 7 out family linkage study on chromosome 1 using hetero-
of the 12 analyses but was rejected by 5 [170–180]. In chromatic C-band variants in the 1q22.1–23 region
contrast, the multifactorial polygenic model was found co-segregation of a 1qH (C-band) variant and
rejected in only one out of five studies [118, 165, Duffy blood group alleles with schizophrenia in a
181–183], while two studies rejected both the single single family [189]. Later two linkage studies found
major locus and polygenic models [184, 185]. Finally significant evidence for linkage between genetic
two studies that employed the mixed model con- markers on chromosome 1q23.3 and schizophrenia
cluded that major locus transmission with polygenic with LODs of 6.48 and 3.2 [190, 191]. A meta-analy-
background was compatible with the data [180, 186]. sis of linkage studies conducted by Lewis et al. [192]
All of these studies were of course conducted under found consistency in linkage studies on chromosome
the assumption that the etiology of schizophrenia was 1p13.3–1q23.3 with a significance for nonrandom
homogeneous in the samples obtained which is an positive LOD scores of p ¼ 0.014. Other linkage
assumption which we now know is flawed. studies from Finland found significant evidence for
linkage to schizophrenia at 1q32.2 but at a distance
Genetic linkage and association from the DISC1 susceptibility locus [193, 194] nearer
the telomere. Two further analyses from Finland in
studies of schizophrenia independent family samples found a LOD of 3.21
Considerable effort has been focused on genetic link- [195] and then 2.7 at the more distal position of 1q42
age analysis of schizophrenia employing genetic near the DISC1 gene [194]. Other linkage studies
markers in multiply affected families to identify found supportive evidence for linkage to schizophre-
which chromosomal regions harbor susceptibility nia at 1q23.3 [191, 196, 197]. The strongly positive
genes. This approach must take into account the 1q23.3 linkage studies and the positive LODs lower
complication of heterogeneity of linkage in which a down on the chromosome lead to the conclusion
number of susceptibility genes localized to different that there is locus heterogeneity with at least two
chromosomes contribute to the development of loci on chromosome 1. The regions of chromosome
schizophrenia. One strategy in linkage studies has 1 showing linkage to schizophrenia are shown in
been to investigate favored loci implicated by cyto- Figure 19.1.
genetic abnormalities. This older literature was Claims that the NOS1AP/CAPON and RGS4 genes,
reviewed by Bassett [187, 188], but in the last few which are 700 kb apart on chromosome 1q23.3, show
years a large amount of extra data has been accumu- genetic association (LD) with schizophrenia have
lated using microarrays. A second approach is to been published [198–200]. RGS4 has been found
simply study relevant genes such as neurobiological to be associated with schizophrenia in a minority

238
Chapter 19: Genetics of schizophrenia

1p13.3-1q23.3 Figure 19.1 Regions of chromosome


1 showing linkage to schizophrenia,
1q22.1-1q23.3 1q42
the position of the 1q42.1 gene DISC1 that is
disrupted by a translocation in a pedigree
with schizophrenia and the location of
1q21.1 1q23.3 1q32.2 DISC1 the 1q21.1 Copy number variation (CNV)
Deletion CNVs 1q42.1 deletions are shown.
NOS1AP
UHMK1
RGS4

Figure 19.2 Relative genomic positions of schizophrenia associated genes on 1q23.3. The region shown covers 1.13 Mb.

of studies [201, 202] and some doubt exists about [207, 208]. DISC1 is expressed at half the normal
its validity as a schizophrenia susceptibility gene levels in translocation cell lines, indicating that
[201, 203]. Markers at a third gene called UHMK1 haplo-insufficiency may be the cause in this family
in between NOS1AP and NRG1 showed positive allelic [209]. Multiple linkage and genetic association
association with schizophrenia [203]. The marker studies over 20 years [194, 210–222] have provided
D1S1679, which lies between UHMK1 and CAPON evidence for linkage and association between DISC1
has been implicated in two family-based studies of and schizophrenia or bipolar disorder. Endopheno-
schizophrenia using tests of transmission disequilib- typic studies of cognition [212, 215, 223], memory
rium (TDT) [204] and the extended transmission function [211, 224], and imaging variables [211, 212,
disequilibrium test (ETDT) [205], which are both 218] have also provided support for the involvement
methods that attempt to combine linkage and LD of DISC1. A large-scale association study combining
approaches. Further fine mapping of the UHMK1 4 schizophrenia cohorts with 1275 cases and 1236
gene locus was then confirmed in a sample from population-matched controls showed evidence of
Aberdeen [203, 206]. The difficulty in replicating allelic and haplotypic heterogeneity for association
genetic effects from RGS4 and NOS1AP/CAPON in between markers in DISC1 and schizophrenia and
schizophrenia may be because there is LD between bipolar disorder [225]. These findings provided fur-
these genes and the UHMK1 gene which is in the ther support for the role of DISC1 in schizophrenia
middle of RGS4 and NOS1AP/CAPON within a 700 and bipolar disorder and found that there were
kb region. The relative location of the 1q23.3 genes multiple disease haplotypes having an effect within
with evidence for association with schizophrenia are European populations. Animal models of DISC1 are
shown in Figure 19.2. discussed below.
DISC1 located at 1q42 (Figure 19.1) was originally
identified at the breakpoint of a balanced t(1;11)
translocation in a large Scottish pedigree multiply Chromosome 2: ZNF804A
affected with psychiatric illness, including schizophre- A genome-wide association study (GWAS) found good
nia, bipolar disorder, and recurrent major depression evidence of association for the zinc finger 804A

239
Chapter 19: Genetics of schizophrenia

5q11-5q13 5q22-5q34 Figure 19.3 Regions of chromosome


5 showing linkage to schizophrenia.

5q22-5q31 5q33

CLINT1GABRA1
GABRA2
GABRA6

KIAA0171 EBF1 ADRA1B CCNJL ATP10B GABRB2 GABRG2


CLINT1 EBF1 TTC1 SLU7 LOC285629 GABRB2 GABRG2
CLINT1 EBF1 PWWP2A SLU7 GABRB2 GABRG2
CLINT1 EBF1 PWWP2A SLU7 GABRB2 GABRG2
CLINT1 AK123543H PWWP2A PTTG1 GABRB3
RNF145 FABPB ATP10B GABRB2
RNF145 AK097242 ATP10B GABRA6
RNF145 FABP8 ATP10B GABRA6
RNF145 FABP8 KIAA0715 GABRA1
RNF145 CCNJL ATP10B GABRA1
RNF145 CCNJL GABRA1
UBLCP1 CCNJL GABRA1
IL12B C1QTNF2 GABRA1
AK097548K C5orf54 GABRA1
LOC285627 C5orf54 GABRA1
PTTG1
DO658414 k

Figure 19.4 Relative genomic positions of schizophrenia associated genes on 5q33.3–5q34. The region shown covers 4.4 Mb.

(ZNF804A) gene with schizophrenia in a UK sample with p < 0.00005. Further attempts at replication of
[226]. These findings were supported by analysis of this finding were not successful [233–237]. DiBella
16 726 cases and controls drawn from independent et al. [238] found an association between the Msc1
European samples and those from the United States, polymorphism and delusional disorder but no
Australia, Israel, China, and Japan [226]. Independent increased homozygosity with schizophrenia.
replication of these findings with the same marker
and allele (rs1344706 allele T) was found in a schizo- Chromosome 5: CLINT1 (EPSIN4),
phrenia sample from Ireland [227] and in a study of
5164 schizophrenia cases and 20 709 controls from GABRA1, GABRA6, GABRB2, GABRP
Europe and China [228]. The first evidence for a positive LOD score with
genetic linkage markers was claimed by Sherrington
et al. [239] when investigating a chromosome 5 region
Chromosome 3: DRD3 implicated by a cytogenetic abnormality [240] which
Crocq et al. [229] and Kennedy et al. [230] reported co-segregated with schizophrenia. Linkage studies
increased homozygosity or allelic association of the using probes localized a schizophrenia susceptibility
dopamine receptor gene DRD3 in schizophrenics com- locus to the chromosomal area 5q11–5q13 in five
pared to controls. The statistical significance of this large Icelandic and two large British families. Two
result was most marked in positive family history cases entirely new cohorts of families from Iceland and
and in those with good treatment response. Mant et al. the United Kingdom excluded linkage over the
[231] attempted a replication but found significance of 5q11–q13 region [241]. This linkage was unreplicated
only 0.05. Evidence of increased homozygosity was until 2006 when genetic linkage to schizophrenia with
also found in a UK/Icelandic family sample [232]. a LOD of 3.8 was found over the 5q11–13 region
The combined sample of Crocq and Mant nevertheless [242]. Genetic association studies in the 5q11–13
showed a highly significant increase in homozygosity region have since implicated the kinesin interacting

240
Chapter 19: Genetics of schizophrenia

factor 2A (KIF2A) gene in susceptibility to schizo- GABRA1 was found in a Japanese sample [256]. The
phrenia [133], but further tests of association chromosomal region implicated shows a low level of
between KIF2A and schizophrenia are needed to con- genetic recombination. Therefore markers at the
firm this. GABRA1, GABRA6, and GABRG1 genes will show
Other linkage studies have implicated the more LD with each other making it uncertain as to which
telomeric chromosome 5q22–34 region. The first was gene is involved. GABRP is 10 million bases away and
in a sample of Irish families [243] with a heterogeneity will not show LD with the 3 genes nearer the centro-
LOD of 3.35 near the locus D5S804. Subsequently, a mere and will give an independent allelic association
single large kindred from Palau, Micronesia, gave a signal. GABRB2 also on 5q34 has been associated with
LOD of 3.4 showing linkage to schizophrenia with schizophrenia [255] and this gene is not in LD with the
markers in the 5q22–31 region [244]. Later a study of other 5q34 GABA receptor cluster genes. The involve-
Finnish families also found linkage to schizophrenia ment of the GABRB2 gene has also been shown in
with a LOD of 3.56 at 5q22–31 [245]. A British/ Japanese and German populations [257] as well as in
Icelandic study implicated the 5q22–34 region with a an independent Japanese sample [258] and in others
LOD score of 3.6 [246]. A collaborative study of the [259, 260]. A meta-analysis of the involvement of
5q22–31 region [247] found that 4 out of 8 independ- GABRB2 gene [261] confirmed that it was associated
ent schizophrenia family linkage samples gave LOD with schizophrenia across several studies (Figure 19.4).
scores above 1.00 [248]. These studies suggest that
there must be at least 2, and possibly 3 or more,
different regions of chromosome 5 which harbor sus-
Chromosome 6: HLA, DTNBP1,
ceptibility loci for schizophrenia (Figure 19.3). The TRAR4 (TAAR6)
evidence implicating the more distal region of chromo- Early studies of association and linkage between the
some 5 at 5q33 is derived from 4 independent linkage HLA genes and schizophrenia produced consistently
studies reporting LOD scores above 3.00. negative results [262–264] with the exception of asso-
Genetic association studies of the 5q21–22 region ciation between HLA A9 and paranoid schizophrenia.
have now shown that several genes show allelic associ- This has been put into reverse with several GWAS,
ation with schizophrenia. Pimm et al. [249] implicated described below, now implicating at least three haplo-
the clathrin interacting protein 1 (CLINT1 or EPSIN4) types in the HLA region. Evidence from linkage analy-
in susceptibility to schizophrenia. This was strongly sis in families that there is a susceptibility locus for
replicated with several new markers at the CLINT1 schizophrenia on chromosome 6p22.3 was provided
locus in independent samples by the International by several research groups [265–268]. However, other
Schizophrenic Consortium [250]. Liou et al. [251] also research teams were unable to detect the presence of
found some degree of support for involvement of the this locus in family samples [190, 245, 246, 269, 271].
CLINT1 genes in schizophrenia by showing positive The evidence pointed strongly to heterogeneity of
haplotypic association in a Chinese sample. Support linkage and the study by Straub and coworkers [272]
for the finding also derives from a family-based asso- was the first time in psychiatry that an admixture test
ciation study [252] which was positive in a second Han rejected homogeneity in favor of a linked 6p22–25
Chinese population. The methodology of these studies subgroup. Confirmed evidence for linkage at the
has been reviewed [253]. Since then a Latino family- 6p22.3 locus was followed up with allelic association
based association study [254] has also confirmed studies using the family-based methods implemen-
involvement of CLINT1 in schizophrenia. ted in TRANSMIT [272]. When the most stringent
A second study of the 5q23–34 region containing test of allelic association was carried out evidence
the GABA receptor subunit genes GABRA1, GABRA6, for allelic association with schizophrenia was present
GABRB2, and GABRP was carried out [255]. Associ- [272].
ations between single nucleotide polymorphisms The haplotype structure in the original Irish fam-
(SNPs) and haplotypes in GABRA1 with schizophrenia ilies showed that 96% of the variability in the sample
was found as well as associations in the GABRP and was attributed to 6 haplotypes [273]. Only 1 of the
GABRA6 genes. The positive GABRA1 and GABRP 6 haplotypes was identified as a high risk haplotype
findings were replicated in an independent German and this had a frequency of 6% in the whole sample
family-based sample [255] and association with [273, 274]. A subsequent study of German, Hungarian,

241
Chapter 19: Genetics of schizophrenia

and Israeli families also showed positive tests of allelic psychosis. This isolate was unique in that there was
TDT with the DTNBP1 locus but only with a single only one paternal origin for the whole village as a result
SNP and not with any of the SNPs that had been of a bottleneck event that occurred a few centuries ago.
previously found to be associated in the Irish family Identical-by-descent (IBD) haplotype sharing was
sample [275]. The German sample also showed a used to carry out a genome-wide scan with 359 micro-
positive haplotypic association composed of three satellite markers. A positive linkage (p < 0.0001) was
adjacent SNPs. However, the alleles in the associated observed over the DTNBP1 gene on chromosome 6.
haplotype were different to those found in Straub However, such an inbred population cannot distin-
et al.’s Irish sample [272]. guish between linkage and association.
A positive haplotypic association with DTNBP1 Numakawa et al. used a case-control sample of
was also found in a study of Han Chinese trios, where 670 patients and 588 controls and found that four
the risk haplotype was composed of the most SNPs were associated with schizophrenia [284].
common SNP alleles [276]. This was in contrast to Positive allelic and haplotypic associations were also
the haplotype associated in the Irish study which was reported by Li and colleagues [285] in a Han Chinese
composed of minor frequency SNP alleles [272]. population. Positive haplotypic but not allelic asso-
A genetic association study consisting of three case- ciation was reported by Tochigi et al. [286]. How-
control samples from Germany, Poland, and Sweden ever, as with the some other studies [275, 276], the
found no evidence for association with any of the associated haplotypes were not the same as reported
five SNPs previously reported to be positive except in the original study. Li et al. [285] screened a
for one SNP (rs1011313) in the Swedish sub-sample. Scottish sample of 580 schizophrenic cases and 620
Haplotypic associations were also negative for the controls. They were unable to replicate any previously
German and Polish sub-samples. Only the Swedish identified single marker or haplotypic associations
schizophrenic sample showed haplotypic association [278]. However a rare haplotype of SNP A with
with schizophrenia. Several, two, three, or four P1583 remained significant after permutation testing
marker haplotypes and a single five-marker SNP hap- (p ¼ 0.03) [285]. No evidence for allelic or haplotypic
lotype showed association. The strength of association association with these markers was found in an inde-
was increased in the subgroup of cases with a positive pendent UK schizophrenia sample [287]. It seems that
family history of schizophrenia [277]. The Swedish dysbindin is associated with schizophrenia but that
haplotype was yet another one found to be associated there are heterogeneous disease haplotypes in different
with schizophrenia. Van Den Bogaert et al. [277] samples.
suggested that the five marker Swedish haplotype asso- DTNBP1 is a 40-kDa coiled-coil-containing pro-
ciated with schizophrenia is phylogenetically related to tein that shows cross species evolutionary conser-
the original haplotype found to be associated with vation [288]. DTNBP1 binds to both alpha- and
Irish schizophrenia by Straub et al. [272]. beta-dystrobrevin in muscle and brain [288]. In the
A novel DTNBP1 haplotype incorporating SNP A brain DTNBP1 is found in axons, particularly in the
(T allele) has been identified in a case-control sample mossy fiber synaptic terminals of the cerebellum and
of 708 cases and 711 controls of British and Irish hippocampus [288]. Talbot and colleagues [289]
descent that showed association with schizophrenia compared presynaptic dysbindin-1 levels at hippo-
[278]. This haplotype was also found to be associated campal formation sites lacking neuronal dystro-
with schizophrenia in an Irish case-control sample that brevin between schizophrenics and controls and
did not initially show association [279, 280]. The same found that these levels were significantly reduced in
research group also reported a positive replication in a schizophrenics.
sample of 488 Hungarian schizophrenic trios [280]. Other genes on chromosome 6 that have been
Another association study by Funke et al. [281] found implicated in schizophrenia by allelic association
an allelic association that was with the opposite allele include the HLA region genes [250, 290] and the trace
found to be associated by Straub. Hall et al. [282] amine receptor 4 (TRAR4 now known as trace amine
found no significant allelic or haplotypic associations associated receptor 6 [TAAR6] on 6q23.2 [291]). Early
between DTNBP1 and schizophrenia [282]. An Israeli studies of association and linkage between the HLA
isolate was used in a genome-wide scan by Kohn et al. genes and schizophrenia produce consistently negative
[283] to identify loci predisposing to any diagnosis of results [262–264] with the exception of association

242
Chapter 19: Genetics of schizophrenia

6p22.2-6p21.32 Figure 19.5 Regions of chromosome 6


showing linkage to schizophrenia and
6p22.3 (HLA region)
the positions of the schizophrenia
associated genes DTNBP1 and TAAR6.

6p22-6p25 DTNBP1 6q23.2


(TAAR6)

7q12.3 7q22.1 Figure 19.6 The localization of genes


(ABCA13) (RELN) on chromosome 7 that have been shown
to be associated with schizophrenia.

7q21.11-7q21.12
(GRM3)

between HLA A9 and paranoid schizophrenia. This and one was homozygous for missense mutations
has been put into reverse with several genes now whereas controls were heterozygous. This suggests
implicating at least three haplotypes in the HLA recessive inheritance with possible allele dosage
region [262–264]. The low level of recombination effects.
across the HLA region makes it difficult to identify Several studies have implicated the metabotropic
which gene among hundreds is increasing genetic glutamate receptor gene GRM3 in schizophrenia
susceptibility to schizophrenia. It will require deep [295–308]. MRI changes showing abnormal acti-
DNA sequencing of genomic DNA from schizophre- vation patterns in both cortical regions in control
nia cases associated with HLA makers followed by subjects homozygous for a genetic variant have been
humanizing mouse strains with genomic DNA from reported [300]. A transgenic animal for schizophrenia
schizophrenia cases with susceptibility alleles and is described below.
haplotypes to identify which part of the HLA region Markers at the reelin gene (RELN) were reported
is implicated. Allelic association between TAAR6 to be associated with schizophrenia [309] and two
markers and schizophrenia has not been replicated studies find this association only in females with
[292, 293] (Figure 19.5). schizophrenia [310, 311]. Five SNP markers were
nominally associated with schizophrenia in the Inter-
Chromosome 7: ABCA13, GRM3, RELN national Schizophrenia Consortium (ISC) GWAS
[250]. However a number of studies have not shown
A chromosome abnormality affecting the lipid trans- association with RELN [309, 312, 313] (Figure 19.6).
porter gene ABCA13 was found in a person with
schizophrenia. Possible etiological base pair changes
in the form of multiple rare coding variants were
revealed in ABCA13 after sequencing 100 cases
Chromosome 8: NRG1, PCM1,
[294]. ABCA proteins shuttle lipid molecules across PPP3CC, FZD3, DPYSL2
cell membranes. Different ABCA proteins have roles Five genetic linkage analyses of the 8p22–21 region
in vesicular trafficking, signal transduction, and tran- in independent schizophrenia family samples have
scriptional regulation. ABCA13 is strongly expressed confirmed linkage with LODs above 3.00 [242, 246,
in the choroid plexus and the dentate gyrus granule 314–316]. Two other linkage studies have provided
cell layer in the central nervous system (CNS). further support for linkage at 8p23–21 with LODs of
ABCA13 mutations were also found to be present in between 2.00 and 3.00 in Irish and African-American
cases of bipolar disorder, but there were differences family samples [317, 318]. Linkage on the chromo-
between schizophrenia and bipolar disorder in terms somal region 8p22–21 is therefore one of the most
of heterozygosity and homozygosity for rare variants. well-replicated linkages to schizophrenia. Polymorphic
Five cases were found to be compound heterozygotes markers localized within or next to five genes: PCM1,

243
Chapter 19: Genetics of schizophrenia

PPP3CC, DPYSL2, FZD3, and NRG1 have shown percentage of cases and that the frequency of disease
association with schizophrenia in the 8p22–12 region alleles is very low and also variable even with similar
[319–320]. Similar allelic and haplotypic associations ancestral populations.
at the NRG1 locus were found in Scotland [326]. Tests Tests of extended transmission disequilibrium
of haplotypic association did reveal weak evidence (ETDT) in the linked chromosome 8p region in a
for association (p ¼ 0.04) for the original HAPICE sample of large multiplex schizophrenia families
haplotype associated with schizophrenia in Iceland found significant allelic association with schizophre-
[327]. Seventeen microsatellite markers and three nia using the marker D8S261 which is within the
SNPs were analyzed in an Irish case-control sample PCM1 gene [320]. The association with D8S261 was
and no allelic associations were detected [328]. How- then replicated in a sample of 100 US-parent off-
ever a novel two marker haplotype named HapBIRE spring trios [320]. The next significant evidence of
did show association. This haplotype shared some allelic association between schizophrenia and 4 SNPs
alleles also found in “at risk” haplotypes reported in as well as with D8S261 was found in a case-control
the original Icelandic study [323]. The association of sample of 450 cases and 450 controls [320]. Some of
markers at NRG1 with schizophrenia has also been these SNPs were also associated with schizophrenia
detected in the Chinese Han population [325]. An in a sample of 800 cases and 800 controls from
independent attempt at replicating Yang’s findings Aberdeen. In both the Scottish and University College
in different Chinese case-control and trio samples London (UCL) samples, significant haplotypic associ-
did not reveal any significant associations [329]. ations were found with the same alleles being impli-
Li [330] could find no evidence for association with cated [339]. Further evidence for allelic association
the HAPICE haplotype in either a case-control or between PCM1 markers and schizophrenia at the
family-based sample. However, three novel associated PCM1 locus was found independently in the Clinical
haplotypes were identified in these studies. A study Antipsychotic Trials in Intervention Effectiveness
from Holland has also shown support for NRG1 (CATIE) GWAS of schizophrenia [340], the Schizo-
involvement in schizophrenia but when patients were phrenia Genetics Consortium (SGENE) GWAS
split into “deficit” (chronic symptoms with idiopathic [290, 341], and the ISC GWAS [251]. Sixty-nine
negative symptoms) and “nondeficit” subtypes of markers across the PCM1 locus (the PCM1 gene plus
schizophrenia [331] a weak positive association was 50 kb either side) have shown allelic association with
found. A Chinese study took the LD structure into schizophrenia in multiple samples. The ISC data
account when performing haplotypic association ana- showed that 10 PCM1 SNPs were associated with
lyses [332]. Thirteen microsatellite markers were gen- schizophrenia after the UCL sub-sample was excluded
otyped and the investigators observed four different with p-values of 0.005–0.02 [250]. Taken together, the
LD blocks present along the NRG1 region analyzed. association findings provide evidence for replicated
Significant haplotypic associations for two out of the association of PCM1 markers with schizophrenia.
four LD blocks were found [332]. A further positive Abnormal brain morphology was also detected in
replication was successful [333] using a case-control PCM1 associated cases of schizophrenia compared
and trio sample of Han Chinese origin. As with the to normal controls and non-PCM1 associated cases
other studies using Han Chinese populations, the of schizophrenia. Significant reduction in the volume
HAPICE haplotype itself was not associated with of orbitofrontal cortex gray matter was observed in
schizophrenia, but a novel haplotype was found to PCM1 associated patients whereas non-PCM1 associ-
be associated in both the case-control and the family ated schizophrenics showed reductions in temporal
sample. A study based in the United States [334] also pole volumes [320].
found haplotypic association between NRG1 and Sequencing of genomic DNA from schizophrenia
schizophrenia but not for the core HAPICE haplotype. volunteers who had inherited haplotypes associated
Gene expression studies in schizophrenia also sup- with schizophrenia showed a threonine to isoleucine
port a role for NRG1 in schizophrenia susceptibility missense mutation in exon 24 which was likely
[334, 335]. Although there has been substantial sup- to change the structure and function of PCM1
port for the NRG1 association with schizophrenia (rs370429). This mutation was found only as a het-
there have also been negative studies [298, 336–338], erozygote in 98 schizophrenic research subjects and
indicating that the effect of NRG1 is only in a few controls out of 2246 case and control research

244
Chapter 19: Genetics of schizophrenia

subjects [339]. Among the 98 carriers of the threonine population also failed to find association with any of
to isoleucine amino acid change (rs370429) 67 were the markers found within PPP3CC [346]. PPP3CC is
affected with schizophrenia. Another potential etio- an attractive candidate gene for schizophrenia suscep-
logical base pair change in PCM1 was rs445422, tibility, not only because it is highly expressed in the
which altered a splice site signal [339]. A further central nervous system but also because it is thought
mutation, rs208747, was shown by electrophoretic to have a pivotal role in the regulation of dopamine
mobility shift assays to create or destroy a promoter signal transduction [347, 348].
transcription factor binding site [339]. Five further Another gene reported to be associated with
nonsynonymous changes in exons were also found. schizophrenia in the 8p21.1 region is frizzled 3
Genotyping of the new variants discovered in the (FZD3). FZD3 is a seven-transmembrane receptor
UCL case-control sample strengthened the evidence for Wnt glycoproteins in the Wnt signal transduction
for allelic and haplotypic association (p ¼ 0.02– cascade reactions which ultimately play an essential
0.0002). Given the number and identity of the part in neurodevelopment. First reported in a Chinese
haplotypes associated with schizophrenia, further Han sample of 246 trios, 3 SNPs were found to be
etiological base pair changes should exist within and overtransmitted to cases with p-values ranging
around the PCM1 gene [339]. A rare PCM1 variant between 0.003 and 0.00007 [324]. This sample also
that co-segregated with schizophrenia in a single displays association with markers and haplotypes
pedigree that was predicted to cause a premature within the NRG1 locus [325]. Around the same time
truncation of the PCM1 peptide was reported in an a Japanese study identified variants within FZD3
US sample [342]. A 3 Mb deletion that includes the which had allelic, genotypic, and haplotypic associ-
entire PCM1 gene has been reported in a single case ation with schizophrenia [321, 324]. A Chinese study
of schizophrenia from Germany [343]. In the ISC also found association with markers at FZD3 [349].
collaboration a duplication in a single case of schizo- No association at this locus was detected in a British
phrenia from Sweden that included the first 13 exons sample [350], or in Japanese case-control and family-
of the PCM1 gene was discovered [343]. PCM1 has based samples [351, 352].
now been shown to interact directly with the schizo- A further gene in the 8p21–22 region, dihydropyr-
phrenia susceptibility gene DISC1 [342]. An amino imidinase related protein (DPYSL2), 4.1 Mb distal to
acid change in DISC1 changes interaction with PCM1, PPP3CC, has been found to show association with
changes the region of localization of PCM1 at the schizophrenia [322]. A replication was attempted in
centromere, and effects noradrenaline release in two North American samples, one of Caucasian, and
SHY5Y cells [344]. PCM1 is known to be crucial for the other of African-American origin, and the C allele
microtubule function and trafficking in neurons and was associated with schizophrenia only in the Cauca-
could cause the molecular pathology of a subtype of sian sample [353] (Figure 19.7).
schizophrenia.
The PPP3CC gene at 8p21.3, which encodes the
calcineurin gamma A catalytic subunit (CNAg) of Chromosome 11: FXYD6, DRD2
phosphatase 2B, has been implicated by a genetic Genetic linkage analysis carried out as a result of the
association study in schizophrenia. Gerber and col- finding of a cytogenetic abnormality in a rare family
leagues [319] selected SNPs within PPP3CC and gen- has yielded a LOD score of above 3.00 on chromo-
otyped these in a family sample from the United some 11 [208]. The LOD score went to above 4.00
States and South Africa. Allelic and haplotypic associ- when all individuals with psychiatric illnesses were
ation was detected with the two SNP alleles in the counted as cases and showed that schizophrenia and
PPP3CC gamma subunit in the US sample. The South a number of other psychiatric disorders within a
African sample failed to display significant global single large kindred segregated with a chromosomal
transmission distortion, but did show a trend toward translocation, implicating the 11q21–11q22 region on
association with the common five SNP risk haplotype the long arm of chromosome 11 in the vicinity of the
observed in the US sample. dopamine D2 neuroreceptor gene (DRD2). Chromo-
An attempt to replicate these findings was per- some 11q22 was first implicated in schizophrenia in a
formed in a Japanese case-control sample but was linkage study of two Japanese pedigrees in which
negative [345]. A study using an Ashkenazi Jewish positive LOD scores were reported [354]. This was

245
Chapter 19: Genetics of schizophrenia

Figure 19.7 Regions of chromosome 8

8p21.3 (PPP3CC)
8p21.2 (DPYSL2)
showing linkage to schizophrenia and

8p21.1 (FZD3)
8p12 (NRG1)
8p22 ( PCM1) the positions of the schizophrenia
associated genes in the 8p22–12 region.

8p22-8p21

8p23-8p21

11q22.3-11q24.1 Figure 19.8 Regions of chromosome 11


11q23.2 showing linkage to schizophrenia and
11q22 (DRD2) the positions of the schizophrenia
associated genes in the 11q23 region.

11q23.3
11q21-11q22 11q23.3 (FXYD6)

11q22-11q24

confirmed in a separate study of a large Canadian central nervous system during postnatal develop-
pedigree in which a maximum LOD of 3.4 was ment, as well as those in the adult brain. Allelic
obtained [355]. Further evidence of linkage to schizo- association between markers at or near DRD2
phrenia at 11q23.3 was found with a LOD score of 3.1 and schizophrenia has been reported surprisingly
and 1 single family was found to have a LOD score of frequently given that initial studies were negative
3.2 [246]. The chromosome 11q22–24 region has [233, 365–377]. Other studies in several different
been shown to be one of the most well-established populations were negative [378–381]. Two meta-
linkages to schizophrenia by a meta-analysis of 20 analyses of the serine or cysteine at codon 311 of
genome scans which employed nonparametric rank the gene in 2003 were both positive. One of these
order statistics to show that evidence for linkage at included 9152 cases with significance of p < 0.001;
this region was non-random [191]. This region on odds ratio (OR) 1.43 [382]. The other found a pooled
11q was subsequently studied in a group of Welsh/ OR of 1.3 for the G allele that encodes the cysteine
English families and a LOD of approximately 3.40 was residue, which was significant (p ¼ 0.007) [370]. A more
found with a quasi-recessive model [356]. Other Scot- recent meta-analysis of the DRD2 SNPs rs1801028
tish, Irish, Swedish, Icelandic, and Japanese family- and rs6277 was significant for association with both
linkage studies were largely negative [354, 357–359]. SNPs [261] (Figure 19.8).
Markers in the region implicated by the linkage stud-
ies that have shown LD with schizophrenia in case-
control studies include the FXDY6 gene encoding Chromosome 13: DAOA
phosphohippolin and the DRD2 dopamine receptor Linkage evidence for a schizophrenia susceptibility
genes. FXYD6 was associated with schizophrenia in locus mapping to chromosome 13q has been reported
two samples from the United Kingdom [360]. The by several groups [196, 247, 314, 315, 383–386]. Other
gene encodes a sodium ATPase regulator, phospho- research groups were unable to detect the presence of
hippolin, which is highly expressed in most parts of this locus in family-linkage studies [246, 387, 388]
the brain [361–364]. Phosphohippolin may play an once again providing evidence that variability in the
important role in the excitability of neurons in the outcome of linkage studies was due to heterogeneity

246
Chapter 19: Genetics of schizophrenia

13q33.2 Figure 19.9 Regions of chromosome 13


(DAOA) showing linkage to schizophrenia and
13q22-13q32 the position of the schizophrenia
associated gene DAOA in the 13q33.2
region.

13q11-13q21 13q22-13q34

of linkage. The distribution of these linked markers mood disorder group alone [395]. A second UK study
would suggest at least two distinct regions of linkage: found allelic associations with both schizophrenia and
13q11–21 and 13q22–32. The distal 13q22–32 region bipolar disorder [396]. Several studies have been pub-
is more robustly supported by the linkage data. lished where association was not detected between
Chumakov and colleagues [389] initially investigated DAOA and schizophrenia [397, 398]. The DAOA gene
the 13q22–34 region in schizophrenia. Six markers is expressed in the endoplasmic reticulum [389] and
were found to be associated with schizophrenia in a may have a role in glutamate synthesis (Figure 19.9).
French-Canadian sample, two of which were also
associated in a Russian sample [389]. This region
contains two overlapping genes, DAOA (G72) and Chromosome 14: NPAS3, AKT1
G30, which are transcribed in opposite directions A cytogenetic abnormality at 14q13 in a mother and
[389]. A number of groups have subsequently daughter diagnosed with schizophrenia and schizo-
attempted to replicate the associations between phrenia comorbid with mild learning disability con-
DAOA/G30 and schizophrenia. A meta-analysis of sisted of a balanced reciprocal translocation t(9,14)
association findings for the DAOA/G30 locus pub- (q34.2;q13). The 14q13 breakpoint disrupted the
lished up until April 2007 found consistent evidence NPAS3 gene which encodes CNS expressed transcrip-
for association between markers rs947267, rs778293, tion factor of the basic helix-loop-helix PAS gene
and rs1421292 with schizophrenia [390]. Hall et al. family [399–402]. Over 20 SNPs within NPAS3 were
[282], who studied US and Afrikaans trios with associated with schizophrenia in the ISC GWAS
schizophrenia, found association between rs2391191 [251], and association studies in Scotland and Canada
and schizophrenia in the Afrikaans sample. Koros- were positive for this gene [400, 403].
tishevsky et al. [391] found suggestive evidence for AKT1 (V-akt murine thymoma viral oncogene
association with schizophrenia at the DAOA locus in homolog 1) located at 14q32.33 is involved in intra-
a Palestinian-Arab sample. Fallin et al. [346] used an cellular signaling pathways thought to be of etio-
Israeli sample of schizophrenia trios and showed logical importance in schizophrenia.
association with markers at DAOA/G30. Shin et al. Significant AKT1 SNP and/or haplotype associ-
[392] found evidence for association in a Korean ations have been reported with schizophrenia in
schizophrenia case-control sample. Shinkai et al. samples from the United States, Germany, Japan,
[393] report association with schizophrenia in their China, Ireland, and the United Kingdom [404–410].
case-control sample and overtransmission of alleles to In common with many of the other loci discussed, a
affected offspring in their family-based sample. Cor- number of studies have failed to replicate these
vin et al. [394] report a positive association in an Irish findings [411–416] (Figure 19.10).
schizophrenia case-control sample. A study of Pales-
tinians also confirmed involvement of the DAOA
locus in schizophrenia [391].
Chromosome 22: COMT, VCFS REGION,
A British study of allelic association with DAOA PRODH, BRD1
polymorphisms in schizophrenia could not find asso- A genome scan reported linkage between schizophre-
ciation [395]. The authors extended their analyses to nia and markers on chromosome 22q12–13 [417].
include a subset of schizophrenia patients with major Pulver et al. [418, 419] computed linkage analyses
mood disorders in their medical histories as well as and found a positive LOD of 1.54 over the region
grouping together those found with psychosis only 22q12–13.1. Reanalysis by varying parameters in the
[395]. Significance values became more significant dominant model maximized the LOD to 2.82. In a
when this regrouping of cases was applied for the collaborative analysis in a larger sample from five

247
Chapter 19: Genetics of schizophrenia

14q13.1 14q32.33 Figure 19.10 Regions of chromosome 14


(NPAS3) (AKT1) showing the position of the schizophrenia
associated genes NPAS3 and AKT1.
NPAS3 in the 14q13.1 region was
disrupted by balanced reciprocal
translocation in a family with
schizophrenia.

22q11.21 22q11.21 22q13.33 Figure 19.11 Regions of chromosome 22


(PRODH) (COMT) (BRD1) showing linkage to schizophrenia and the
22q12-22q13.1 position of the schizophrenia associated
genes PRODH, COMT, and BRD1. The 3 Mb
typically deleted velo-cardio-facial
syndrome (VCFS) region is also shown.

22q11 VCFS 22q12-22q13


deletion region

laboratories [419, 420] this linkage could not be con- genes including proline dehydrogenase (PRODH2)
firmed assuming either homogeneity or hetero- [343]. In the case of PRODH2 evidence of association
geneity. One of the collaborating laboratories [420] with schizophrenia has been published [431, 432] but
also independently found a maximum positive several studies were negative [433–436]. A mouse
LOD of 1.51 between schizophrenia and the marker with a mutation in PRODH2, causing hyperprolinae-
D22S278, assuming homogeneity in their sample. mia, had a deficit in sensorimotor gating accompan-
Kalsi et al. [421] in a UK/Iceland schizophrenia link- ied by regional neurochemical alterations in the brain
age study could not find any evidence over the rele- [437]. Mice genetically changed to have PRODH2 defi-
vant region for linkage to schizophrenia assuming ciency had increased neurotransmitter release at gluta-
either homogeneity or heterogeneity. Polymeropoulos matergic synapses and deficits in associative learning
et al. [422] found a positive LOD of 0.37 for marker as well as responsiveness to psychotomimetic drugs
D22S278 and found that the sibling-pair linkage [438]. The PRODH2 gene encodes for proline dehy-
method also favored linkage. Schwab et al. [266] drogenase, a mitochondrial enzyme that converts pro-
who studied German and Israeli families could not line to 1-pyrroline-5-carboxylate and is involved in
exclude this region using a sibling-pair linkage transfer of redox potential across the mitochondrial
method. In a much larger sample which was, at that membrane. It is widely expressed in the human brain.
time, almost the entire world’s collection of 500 COMT inactivates catecholamines at postsynaptic
schizophrenia families found linkage to the relevant sites in the brain and as such is an ideal candidate
region on chromosome 22 [423]. Several other lines gene for schizophrenia. The most widely studied
of evidence support involvement of chromosome 22. COMT polymorphism is rs4680 which codes for a
A case-control sample from Scotland found support functional variant at codon 108/158 that results in a
for the involvement of the BRD1 gene [424]. A com- change from a valine to a methionine. The valine
bined UK and Danish sample implicated BRD1 and form of COMT has significantly lower enzyme activ-
the neighboring gene ZBED4 in schizophrenia [425]. ity than the methionine [439]. Meta-analyses of pub-
BRD1 is a brain-expressed gene, which has a bromo lished studies initially concluded that the valine allele
domain and a FAM domain, and by sequence simi- was associated with schizophrenia [440]. However,
larity it has been classified as a potential regulator more recent analyses including more stringent quality
of transcription. The function of BRD1 is not yet control has indicated that this finding may not be
understood. robust [441]. A meta-analysis of two additional COMT
The genetic deletion syndrome, velo-cardio-facial polymorphisms that were reported with highly sig-
(VCFS; DiGeorge) syndrome, is known to be associ- nificant association p-values in an Ashkenazi Jewish
ated with schizophrenia in infrequent cases [419, 426– sample of schizophrenics [442] were later shown to
430]. The deletions on 22q11 include the catechol-O- have modest association in a meta-analysis [261]
methyl transferase (COMT) gene as well as 26 other (Figure 19.11).

248
Chapter 19: Genetics of schizophrenia

Cytogenetic abnormalities, deletions collaboration [452]. This group also found an associ-
ation in the 22q11.2 region with a deletion in 0.2%
and duplications in schizophrenia of cases and absent in any controls. An additional new
Linkage markers mapping to a favored locus, impli- deletion was also found on chromosome 1q21.1.
cated by co-segregation of a familial genetic disease Micro-deletions and micro-duplications of greater
with schizophrenia or by cytogenetic abnormalities in than 0.1 Mb were also found in 15% of patients
rare cases of schizophrenia, have been used to choose suffering with schizophrenia but in only 5% of con-
markers in linkage studies of schizophrenia. Cytogen- trols [453]. When the phenotype was altered to
etic abnormalities reported in combination with schizo- include only patients with early-onset disease the gen-
phrenic psychoses include fragile sites [443–445], etic abnormalities were then found in 20% of cases.
deletions, pericentric inversions [446, 447], trisomies Encouragingly, some of the regions identified as
[240, 448, 449], and acentric fragments [450, 451]. having CNVs using the CGH technology were repli-
Many different genetic deletions and duplications, cated by the GWAS studies [453–455]. A CNV study
also known as CNVs have recently been found to by Xu et al. [455] showed that many CNV mutations
contribute to the genetic etiology of schizophrenia. were de novo and not transmitted from parents.
Two general types of microarrays have been used. Confirmed de novo mutations were significantly asso-
One microarray uses the genotyping of SNPs with ciated with schizophrenia (p ¼ 0.00078) and were
special markers incorporated for quantitative assess- 8 times greater in cases than in unaffected controls.
ment of CNV fluorescent intensity differences. The discovery that some CNVs arise de novo in some
The other uses fixed short oligonucleotides or fixed individuals with schizophrenia was also strongly sup-
genomic DNA clones to hybridize with DNA from a ported in the SGENE collaboration where many of
patient in comparison to DNA from a control subject, the 66 de novo CNV mutations found were associated
a technique known as comparative genomic hybrid- with schizophrenia using a case-control design [452].
ization (CGH). The genotypes and fluorescent signal A further study of 2977 schizophrenia cases found a
intensities are then analyzed to detect insertions and 16-fold increase in exonic deletions of the Neurexin 1
deletions of DNA. The CNVs cause disease by either (NRXN1) gene in cases compared to controls [456].
releasing the effect of an unexpressed recessive muta- A strong overlap between deletions and duplications
tion in the homologous gene on the nondeleted piece found in idiopathic generalized epilepsy and schizo-
of chromosome (hemizygosity) or by a direct effect of phrenia can be observed when comparing the ISC
the CNV on the function of the gene. Because of the study [343] and CNV studies of epilepsy based in
low frequency of deletions it is thought that it is the Germany [457, 458] (Table 19.1). The chromosome
actual gene dosage effect from the deletion of a gene 15q13.1–13.3 deletions were found in both diseases
causing hemizygosity which is causing schizophrenia. with equal frequency and only 1 control out of about
In some samples, between 10 and 20% of schizophre- 6000 had a deletion. The chromosome 22q11.2 dele-
nia cases were found to be positive for these types of tions likewise were strongly associated with schizo-
mutation. At present the current methodology used to phrenia and epilepsy with no affected controls. This
find CNVs is only successful in identifying abnormal- deletion was commoner in schizophrenia (13 cases)
ities above 50 kb (5000 base pairs). The results of two than in epilepsy (2 cases). The overlap between
large recent CNV studies of schizophrenia revealed schizophrenia and epilepsy shows that CNVs can
remarkably consistent results. The ISC found that have pleiotropic effects on the clinical phenotype.
CNVs of 100 kb or more were 1.15 times more likely It also shows that the two disorders can be caused
to appear in patients compared to controls [343]. by a single genetic abnormality. The specific CNVs
Further, 2 new regions were identified as having dele- overlapping in schizophrenia and epilepsy are shown
tions of 0.5 Mb or more on chromosome 15q13.1 and in Table 19.1. Although cases with both schizophrenia
chromosome 1q21.1 (Figure 19.1). Additional dele- and epilepsy are not common the presence of the
tions in patients with typical schizophrenia without same CNVs in both disorders may explain why there
VCFS were identified in the critical 22q11.2 VCFS is a complex relationship between the two disorders
region, as well as in the 16p12.2–p12.2 and 12p11.23 when they are comorbid, for example when there
regions. The same two regions (15q13.3 and 1q21.1) is “forced normalization”. When one contrasts the
were also identified as having deletions in the SGENE high yield of positive results for CNVs in individual

249
Chapter 19: Genetics of schizophrenia

Table 19.1 Number of research subjects (N) with deletions and duplications shared by schizophrenia (SCZ) and idiopathic
generalised epilepsy (IGE) [344, 458–461].

Chromosome Position (Mb) Size Total IGE/ CNV CNV CNV Genes affected
(Mb) SCZ (N) IGE (N) SCZ (N) controls (N)
1q21.1 145.0–146.35 1.35 11 1 10 2 ACP6, GJA5, BCL9
15q13.3 28.7–30.3 1.5 41 31 10 1 TRPM1, CHRNA7
16p11.2 27.05–45.6 5.6 14 1 13 2 DOC2A, KIF22,
MAPK3, etc
16p13.1 29.5–30.1 0.6 43 29 14 2 ABCC6, KIAA0430,
ABCC1, etc
22q11.2 17.5–20.5 3.0 15 2 13 0 COMT, SNAP29,
PRODH, etc
All 124 64 60 7

cases of schizophrenia compared to the negligible able to detect CNVs found that the rate of singleton
findings for MRI or CT scans of schizophrenics CNVs in bipolar disorder was 16.2% compared to
it seems that clinical investigation with microarrays 12.3% in controls (p ¼ 0.007). This study however
to detect CNVs will be a much more productive clin- found no overlap of CNVs between schizophrenia
ical investigation than MRI scans both for diagnosis and bipolar disorder. Two further studies found no
and as a guide to more personalized drug and behav- evidence that bipolar disorder was caused by the
ioral treatments. larger sized CNVs above 100 kb and both studies even
found a slight excess of deletions and duplications
in the controls when compared to the bipolar cases
Genetic association and CNV [468, 469].
differences between schizophrenia A further study of schizophrenia claims that a third
of the genetic variance of schizophrenia and bipolar
and bipolar disorder disorder is shared and that it is polygenic. The defin-
Comparison of linkage hot spots from family linkage ition of polygenic transmission is that two or more
studies for bipolar disorder and schizophrenia in disease mutations or susceptibility variants at two or
meta-analyses found little overlap for the most sig- more distinct gene loci need to be inherited by one
nificantly linked chromosome regions [191, 461–463]. person for them to become affected. In this study
Genetic association studies have been given a big genotype relative risk was used to generate “polygenic”
boost from microarray technology and have enabled scores without any attempt to determine whether
most of the human genes to be investigated in large individuals actually had inherited two disease alleles.
samples of schizophrenics and bipolar cases. When a Therefore the use of the term “polygenic” was a mis-
study of 3322 cases of schizophrenia [250] and 4387 nomer and transmission should have been described
cases of bipolar disorder [464] were compared for the as being multigenic and heterogeneous rather than
top 200 most strongly associated SNP markers in each polygenic. The loci which contributed to the genotype
study only one gene (protocadherin 10, PCDH10) was relative risk “polygenic score” were not localized in
genetically implicated in both schizophrenia and terms of their chromosomal positions. These genetic
bipolar disorder. CNVs and regions of homozygosity markers could therefore have clustered around rela-
have also been studied in schizophrenia and bipolar tively few susceptibility loci and not necessarily at
disorder. Lencz et al. [465] and Vine [466] studied “thousands” of loci as stated in the paper. If different
homozygosity in schizophrenia and bipolar cohorts. mutations occurred in the same genes with one set
Findings were discrepant with no overlaps between causing bipolar disorder but a different set causing
schizophrenia and bipolar disorder. A CNV study of schizophrenia then similar multigenic heterogeneous
bipolar disorder [467] employing microarray methods scores for these regions would have been computed

250
Chapter 19: Genetics of schizophrenia

and this would obscure the fact that different muta- [476–481]. Jones et al. [482] found a family with a
tions were responsible for the two disorders. None of b-amyloid precursor protein mutation that was pre-
the heterogeneous multigenic loci showed genome- sent in the genes of a schizophrenic with abnormal
wide levels of association with schizophrenia or bipolar MRI and EEG findings who might have a form of
disorder. Paradoxically markers showing an absent or angiopathy and dementia causing the schizophrenia.
lower level of allelic association with schizophrenia Other research groups have screened for this muta-
contributed more to the overall “polygenic” score. tion in association samples of schizophrenics and
The polygenic score was absent or very much reduced have not been able to find more cases carrying this
in an African-American schizophrenia sub-sample mutation [235, 482–490].
and this was explained by greater recombination in a
population of African descent. Another possibility is Pharmacogenetics of schizophrenia
whether the polygenic score might actually be a very
This field is still in its infancy. Markers at the 5-HT2a
powerful test of hidden genetic stratification between
receptor gene have been associated with influencing
cases and controls within each sample. This might have
response to clozapine [491–493]. This result remains
been created by the tendency of psychotic individuals
poorly replicated despite being offered commercially
to migrate more than controls. On the other hand, the
as a clinical test [494].
markers identified as contributing to the polygenic
score were more often within genes than in gene
deserts making the hypothesis of genetic drift caused Animal and cell biology studies of
by schizophrenia disease alleles unlikely. suscptibility genes in schizophrenia
The claimed sharing of NRG1 association in both
schizophrenia and bipolar disorder also has compli- DISC1
cations when the details are observed. A study of Disrupted in schizophrenia 1 (DISC1) was identified
bipolar disorder [470] found nominal allele-wise sig- as having truncated gene in a large Scottish family. In
nificant association for rs35753505 with the T allele the mouse a deletion in Disc1 was found in the 129S6/
being overrepresented in cases. This is the opposite SvEv strain but it had a normal phenotype. When the
allelic association to the original association study deletion variant was transferred to C57BL/6J it was
where the C allele was associated with schizophrenia. found to induce impairment of working memory and
In another attempt to implicate the NRG1 gene in prepulse inhibition [495]. However, all mice were
bipolar disorder [471], there was little association subsequently found to have this deletion [496]. An
using the standard BP or mood-incongruent psych- N-ethyl-N-nitrosourea (ENU) induced mutation in
otic BP phenotypes and none of the associations Exon 2 of mouse Disc1 showed depressive-like behav-
withstood correction for multiple maker testing. ior with deficits in the forced swim test that were
For the DAOA (G72) locus there is evidence for reversed by an antidepressant. A second mutation
association with both bipolar disorder and schizo- exhibited schizophrenic-like behavior, with obvious
phrenia [396, 472–475]. In one of these studies the deficits in prepulse inhibition and latent inhibition
same marker and allele (rs3918342 allele T) showed which could be reversed by antipsychotic treatment
association with both disorders [396]. There is also [497]. Disc1 transgenic mice expressing two copies of
modest evidence for association with bipolar disorder the truncated Disc1 gene which encodes the first eight
at the ZNF804A locus [228]. The presence of a pro- exons were found to have enlarged lateral ventricles,
portion of misdiagnosed schizoaffective bipolar males reduced cerebral cortex, partial agenesis of the corpus
in the schizophrenia group might be an explanation callosum, and thinning of layers II/III with reduced
for associations that cross-over between schizophre- neural proliferation during neurogenesis. Parval-
nia and bipolar disorder. bumin GABAergic neurons were reduced in the
hippocampus and medial prefrontal cortex, and dis-
Functional polymorphisms and placed in the dorsolateral frontal cortex. In culture,
transgenic neurons grew fewer with shorter neurites.
mutations in other candidate genes Transgenic mice showed immobility and reduced
Significant association between a neurotrophin 3 gene vocalization in depression-related tests as well as
polymorphism and schizophrenia has been reported impaired conditioning of latent inhibition [498].

251
Chapter 19: Genetics of schizophrenia

PPP3C in subicular pyramidal neurons [503]. MRI imaging


of heterozygous mice showed hypofunction in the
Knockout mice engineered to be deficient in calci-
medial prefrontal cortex, the hippocampal CA1, and
neurin CaNB1 subunit were found to have deficits in
subiculum regions. Type III Nrg1 heterozygous mice
working memory which is an endophenotype found in
showed deficits for delayed alternation memory tasks
some cases of schizophrenia [499]. A follow-up study
and had deficits in prepulse inhibition. Chronic nico-
by the same group identified further schizophrenia
tine treatment eliminated differences in between type
related behaviors in these mice such as increased loco-
III Nrg1 heterozygous mice and wild-type mice [503].
motor activity and pre-pulse inhibition and concluded
In another animal model of NRG1 schizophrenia
that further investigation of the role of calcineurin in
it was found that locomotor activity in the open field
schizophrenia should be carried out [500].
or in photocell cages was significantly enhanced in
Nrg1þ/– mice compared to wild-type littermate con-
SLC1A3 trols. Treatment with the 5HT1A receptor agonist
SLC1A3 is a gene encoding a member of a family of 8-hydroxy-dipropylaminotetralin (8-OH-DPAT) showed
high-affinity sodium-dependent transport proteins a differential effect between genotypes, with a disrup-
that regulate neurotransmitter concentrations at the tion of PPI occurring in Nrg1þ/– mice compared to no
excitatory glutamatergic synapses of the mammalian effect in wild-type controls. There was also a signifi-
central nervous system. The SLC1A3 mutant mouse cant reduction of startle response [504].
exhibits behavioral abnormalities consistent with
positive symptoms of schizophrenia [501]. These STOP
symptoms can be reversed in the SLCA13 mouse
Microtubule stabilisation within neurones has been
model by treatment with haloperidol, the mGlu2/3
shown to be dependent on the action of stable tubule-
agonist, LY379268 or the mGlu2/3 agonist, LY379268.
only polypeptide (STOP) proteins. Neurons contain
SLCA13 knockout (KO) mice had poor nesting
many highly stable microtubules that resist depoly-
behavior and impaired sociability. The knockout
merization after exposure to cold. Stable microtubules
heterozygote was also affected but with lowered pref-
have a vital role in neuronal development, as well as
erence for novel social stimuli. Homozygous knock-
their maintenance and function. Mice lacking the
outs exhibited a significantly reduced acoustic startle
STOP gene showed no obvious defects in brain anat-
response and had impaired learning in an instrumen-
omy but they had synaptic deficits, depleted synaptic
tal visual discrimination task [501].
vesicle pools, and impaired synaptic plasticity
together with abnormal behavior that could be
CHNRA7 reversed by antipsychotic drugs [505, 506]. Mutant
The region of the gene encoding the central alpha7- STOP mice showed reduced expression of synapto-
Nicotinic acetylcholine receptor (CHNRA7 or alpha7- physin, VGlut1, GAP-43, and spinophilin mRNAs in
nAChR) has been implicated repeatedly in a range of the hippocampus or cerebellum [507]. Transgenic
cognitive deficits in schizophrenia. CHNRA7 homo- mice lacking the STOP gene showed hyperlocomotion
zygous and heterozygous knockout mice exhibited which could be reversed by clozapine.
significantly higher attentional load omissions and
had impaired odor span ability [502]. Attentional RELN
impairment appeared to be central to the cognitive
Heterozygous reelin knockout mice have been pro-
deficits observed [502].
posed as a genetic model for schizophrenia [508, 509].
Reelin is an extracellular matrix protein secreted by
NRG1 GABAergic interneurons that interacts with pyram-
In the mouse type III Nrg1 is transcribed by a different idal neuron integrin receptors and can alter dendritic
promoter than the other Nrg1 isoforms. It is expressed spine plasticity. Heterozygous reeler mice express
in the medial prefrontal cortex, ventral hippocampus, only about 50% of the normal amount of reelin
and ventral subiculum. Adult heterozygous mutant mRNA and protein and may mimic the dendritic
mice with disrupted type III Nrg1 have enlarged lat- spine, GABAergic and other defects described in
eral ventricles and decreased dendritic spine density schizophrenia [510, 511]. Hypermethylation of the

252
Chapter 19: Genetics of schizophrenia

RELN promoter has been proposed as the cause of those that have schizophrenia uncomplicated by
low Reelin mRNA levels in postmortem human tissue cannabis smoking [516–518]. One study in genetically
from schizophrenics [512, 513] high-risk subjects found no increased cannabis use
associated with development of schizophrenia [519].
GRM3 Life events as a precipitating factor are no longer being
studied largely because life events are likely to be an
mGluR3 receptor knockout mice were given the
epiphenomena of the schizophrenia susceptibility
mGluR2/3 agonist, LY379268 [514]. Phencyclidine
genes rather having any causal effect. Attention has
evoked hyperactivity, circling, falling, stereotypy,
now shifted towards methylation and other epigenetic
and ataxia, as well as amphetamine-induced hyper-
effects to explain the obvious variable penetrance
activity were reduced by LY379268. LY379268
and pleiotropism of the clinical symptoms. Stochastic
reversed phencyclidine induced hyperactivity and
biological processes during central nervous system
behavioral alterations in wild-type and mGluR3
development alone could explain the variable pene-
knockout mice but not in mice lacking mGluR2.
trance. Despite many genes having been implicated
the identification of clinical subtypes of schizophrenia
Conclusions related to specific genes is still in its infancy. DNA
The presence of a strong genetic predisposition is by sequencing of etiological base pair changes will solve
far the most well-confirmed etiological effect in this problem and finding them will lead to new indi-
schizophrenia. The most striking aspect of the new vidualized treatments based on genetic tests. There is
knowledge is that etiological genetic heterogeneity is scope for narrow spectrum drugs based on mutations
very considerable with no single genetic effect being in an individual and medium spectrum drugs based on
present in more than a few percent of cases. Despite which systems are genetically perturbed and also new
this linkage and association results were nevertheless broad spectrum drugs to take over from drugs like
successful with linkage hot spots predicting where clozapine without the side effect of agranulocytosis.
genes would be fine mapped by allelic LD. There are Although systems biology can generate classifications
now plausible etiological base pair changes that of etiological genetic effects on schizophrenia based on
explain some of the linkage and LD data. Twin and evidence of genome-wide association data these studies
adoption studies have shown that the family envir- would best be conducted using the etiological base pair
onment has no influence on the etiology of schizo- changes discovered by DNA sequencing. The genes
phrenia. Many specific environmental factors that implicated in this review could loosely be classified
interact with the etiological base pair changes have into those that are synaptic (NRG1, DAOA, GRM3,
been proposed. These have been reviewed recently DRD2, SYN2, FXYD6, CLINT1, DTNBP1) and those
[515]. None have reached the status of being com- that are microtubule related (DISC1, PCM1, UHMK1,
pelling. Viruses and immunological factors need to RELN, STOP, NPAS3, KIF2A). Once disease mutations
be reappraised given that the HLA region is now well in these genes are well established, then it will be
implicated in schizophrenia. However there are possible to identify the true modes of transmission as
many nonimmune-related genes within the HLA well as any gene–environment interactions. Postmor-
regions. Cannabis use in schizophrenics has not been tem, psychological, neurophysiological, imaging, and
shown to be independent of genetic factors as shown EEG measures will need to be studied afresh in the
by equally positive family histories of schizophrenia for light of the discovery of the actual etiological base
probands who have so-called cannabis psychosis to pair changes.

References 4. Neale MC, et al. Am J Hum Genet


1989;45:226–239.
8. Asherson P, et al. Br J Psychiatry
1994;164:619–624.
1. MacLean CJ, et al. Genet Epidemiol
1990;7:419–426. 5. Penrose LS. Eur Arch Psychiatry 9. Beckmann H, et al. Am J Med
Clin Neurosci 1991;240:314–324. Genet 1996;67:289–300.
2. Chen WJ, et al. Psychiatr Genet
1992;2:219–238. 6. Bassett AS, et al. Am J Hum Genet 10. Chotai J, et al. Psychiatr Genet
1997;60:630–637. 2005;5:181–186.
3. Kendler KS, et al. Genet Epidemiol
1990;7:409–417. 7. Bassett AS, et al. Am J Hum Genet 11. Blinc-Pesek M, et al. Br J
1994;54:864–870. Psychiatry 2007;191:181.

253
Chapter 19: Genetics of schizophrenia

12. Di Maggio C, et al. Am J 35. DeLisi LE, et al. Arch Gen 58. McDonald C, et al. Am
Psychiatry 2001;158:489–492. Psychiatry 1987;44:891–896. J Psychiatry 2006;163:478–487.
13. Laurent C, et al. Am J Med Genet 36. Slater E, et al. The Genetics of 59. Maj M, et al. J Affect Disord
1998;81:342–346. Mental Disorders. London: Oxford 1990;20:71–77.
14. Thibaut F, et al. Psychiatry Res University Press; 1971.
60. Maj M, et al. J Affect Disord
1995;59:25–33. 37. Leboyer M, et al. Psychiatry Res 2000;57:95–98.
15. Imamura A, et al. J Hum Genet 1992;41:107–114.
61. Maj M, et al. Am J Psychiatry
1998;43:217–223. 38. Tsuang MT, et al. Arch Gen 1991;148:612–616.
16. Ohara K, et al. Biol Psychiatry Psychiatry 1974;31:43–47.
62. Maj M, Acta Psychiatr Scand
1997;42:760–766. 39. Scharfetter C, et al. Schizophr Bull 1985;72:542–550.
17. Heiden A, et al. Schizophr Res 1980;6:586–591.
63. Spitzer R, et al. The Schedule for
1999;35:25–32. 40. Stober G, et al. Eur Arch Affective Disorders and
18. Merette C, et al. Am J Med Genet Psychiatry Clin Neurosci Schizophrenia, Lifetime Version,
2000;96:61–68. 1995;245:135–141. 3rd edn. New York: New York
41. Winokur G, et al. J Nerv Ment State Psychiatric Institute; 1977.
19. Husted J, et al. Am J Med Genet
1998;81:156–162. Disord 1974;159:12–19. 64. Baron M, et al. Acta Psychiatr
20. McInnis MG, et al. Am J Med 42. McGuffin P, et al. Arch Gen Scand 1982;65:253–262.
Genet 1999;88:686–693. Psychiatry 1984;41:541–545. 65. Mendlewicz J, et al. J Affect Disord
21. Cavalli-Sforza LL, et al. Monogr 43. Kendler KS, et al. Arch Gen 1980;2:289–302.
Popul Biol 1981;16:1–388. Psychiatry 1994;51:456–468. 66. Lewis S. Br J Psychiatry 1991;
22. Henderson N. Annual Review of 44. Bassett AS, et al. Schizophr Res 159:158–159.
Psychology 1982;33:403–440. 1994;12:213–221. 67. Weinberger DR, et al. Psychiatry
23. Jakubaschk J, et al. Psychiatr Clin 45. Liddle PF, et al. Br J Psychiatry Res 1981;4:65–71.
(Basel) 1979;12:80–91. 1990;157:558–561. 68. DeLisi LE, et al. Arch Gen
24. Tsuang MT, et al. Arch Gen 46. Keefe RS, et al. Am J Psychiatry Psychiatry 1986;43:148–153.
Psychiatry 1981;38:535–539. 1987;144:889–895. 69. Pearlson GD, et al. J Nerv Ment
25. Gottesman II, et al. Schizophrenia: 47. Roy MA, et al. Am J Psychiatry Disord 1985;173:42–50.
The Epigenetic Puzzle. Cambridge, 1994;151:805–814. 70. Oxenstierna G, et al. Br J
UK: Cambridge University Press; 48. Roy MA, et al. Psychiatry Res Psychiatry 1984;144:654–661.
1982. 1994;54:25–36. 71. Reveley AM, et al. Br J Psychiatry
26. Guze SB, et al. Arch Gen 49. Maier W, et al. Arch Gen 1984;144:89–93.
Psychiatry 1983;40:1273–1276. Psychiatry 1993;50:871–883. 72. Turner SW, et al. Psychol Med
27. Kendler KS, et al. Arch Gen 50. Kendler KS, et al. Arch Gen 1986;16:219–225.
Psychiatry 1984;41:555–564. Psychiatry 1983;40:951–955. 73. Nasrallah HA, et al. J Clin
28. Baron M, et al. Am J Psychiatry 51. Kendler KS, et al. Arch Gen Psychiatry 1983;44:407–409.
1985;142:447–455. Psychiatry 1993;50:952–960. 74. Owen MJ, et al. Br J Psychiatry
29. Kendler KS, et al. Arch Gen 52. Kendler KS, et al. Arch Gen 1989;154:629–634.
Psychiatry 1985;42:770–779. Psychiatry 1993;50:781–788. 75. Orlova VA, et al. Zh Nevrol
30. Gershon ES, et al. Arch Gen 53. Kendler KS, et al. Arch Gen Psikhiatr Im S S Korsakova
Psychiatry 1988;45:328–336. Psychiatry 1993;50:645–652. 1994;94:85–90.
31. Kendler KS, et al. Am J Psychiatry 54. Lapierre YD. Can J Psychiatry 76. Palmour RM, et al. J Psychiatry
1988;145:57–62. 1994;39:S59–S64. Neurosci 1994;19:270–277.
32. Onstad S, et al. Acta Psychiatr 55. Craddock N, et al. Br J Psychiatry 77. Lui S, et al. Psychiatry Res 2009;
Scand 1991;83:463–467. 2005;186:364–366. 171:71–81.
33. Kendler KS, et al. Arch Gen 56. Van Snellenberg JX, et al. Arch 78. Blackwood DH, et al. Arch Gen
Psychiatry 1993;50:527–540. Gen Psychiatry 2009;66:748–755. Psychiatry 1991;48:899–909.
34. Maier W, et al. Schizophr Res 57. Lichtenstein P, et al. Lancet 79. Holzman PS. Schizophr Res
1993;9:71–76. 2009;373:234–239. 1994;13:1–9.

254
Chapter 19: Genetics of schizophrenia

80. Blackwood DH, et al. Br 104. Curtis CE, et al. Am J Psychiatry 126. Casanova MF, et al. J Neurol
J Psychiatry Suppl, 1996:85–92. 2001;158:100–106. Neurosurg Psychiatry
81. Holzman PS, et al. Arch Gen 105. Guerra S, et al. Brain Cogn 1990;53:416–421.
Psychiatry 1988;45:641–647. 2009;70:221–230. 127. Casanova MF, et al. Schizophr Res
82. Holzman PS, et al. Arch Gen 106. Kremen WS, et al. Am J Med 1990;3:155–156.
Psychiatry 1984;41:136–139. Genet B Neuropsychiatr Genet 128. Casanova MF, et al. Biol
83. Siegel C, et al. Arch Gen Psychiatry 2007;144B:403–406. Psychiatry 1990;28:83–84.
1984;41:607–612. 107. Leppanen JM, et al. Schizophr Res 129. Goldberg TE, et al. Arch Gen
84. St Clair D, et al. J Psychiatr Res 2008;99;270–273. Psychiatry 1990;47:1066–1072.
1989;23:49–55. 108. Quinones RM, et al. Eur Arch 130. Goldberg TE, et al. Psychiatry Res
85. Waldo MC, et al. Schizophr Res Psychiatry Clin Neurosci 1994;55:51–61.
1988;1:19–24. 2009;259:475–481.
131. Goldberg TE, et al. Schizophr Res
86. Waldo MC, et al. Biol Psychiatry 109. Schulze KK, et al. Bipolar Disord 1995;17:77–84.
2000;47:231–239. 2008;10:377–386.
132. Goldberg TE, et al. Psychol Med
87. Waldo MC, et al. Psychiatry Res 110. Takahashi S, et al. Psychiatry Clin 1993;23:71–85.
1991;39:257–268. Neurosci 2008;62:487–493.
133. Kremen WS, et al. J Clin Exp
88. Waldo MC, et al. Schizophr Res 111. Turetsky BI, et al. Biol Psychiatry Neuropsychol 2006;28:208–224.
1994;12:93–106. 2008;64:1051–1059.
134. Lyons MJ, et al. Twin Res
89. Cannon TD, et al. Arch Gen 112. Turetsky BI, et al. Schizophr Res 2000;3:28–32.
Psychiatry 1994;51:955–962. 2008;102:220–229.
135. Ragland JD, et al. Schizophr Res
90. Clementz BA, et al. J Abnorm 113. Braff DL, et al. Schizophr Bull 1992;7:177–183.
Psychol 1994;103:400–403. 2007;33:21–32.
136. Cantor-Graae E, et al. J Nerv Ment
91. Docherty NM. J Nerv Ment Dis 114. Braff DL, et al. World Psychiatry Dis 1994;182:645–650.
1993;181:750–756. 2008;7:11–18. 137. McNeil TF, et al. Acta Psychiatr
92. Docherty NM. J Nerv Ment Dis 115. Gottesman II, et al. Schizophrenia Scand 1994;89:196–204.
1994;182:443–451. and Genetics: A Twin Study 138. McNeil TF, et al. Am J Psychiatry
93. Hollister JM, et al. Arch Gen Vantage Point. Cambridge, UK: 2000;157:203–212.
Psychiatry 1994;51:552–558. Academic Press; 1972.
139. Hulshoff Pol HE, et al.
94. Frodl T, et al. Int J Psychophysiol 116. Farmer AE, et al. Arch Gen Neuroimage 2006;31:482–488.
2002;43:237–246. Psychiatry 1987;44:634–641.
140. van Erp TG, et al. Arch Gen
95. Yeap S. et al. Arch Gen Psychiatry 117. Onstad S, et al. Acta Psychiatr Psychiatry 2004;61:346–353.
2006;63:1180–1188. Scand 1991;83:395–401.
141. van Erp TG, et al. Psychiatry Res
96. Woodward ND, et al. Schizophr 118. Torgersen S, et al. Am J Psychiatry 2008;159:271–280.
Res 2007;94:306–316. 1993;150:1661–1667.
142. van Haren NE, et al. Biol
97. Woodward ND, et al. Schizophr 119. Fulker DW. Soc Biol 1973;20: Psychiatry 2004;56:454–461.
Res 2009;109:182–190. 266–275.
143. Brans RG, et al. Arch Gen
98. Dauphinais ID, et al. Psychiatry 120. McGue M, et al. Am J Hum Genet Psychiatry 2008;65:1259–1268.
Res 1990;35:137–147. 1983;35:1161–1178.
144. Sommer IE, et al. Br J Psychiatry
99. Suddath RL, et al. N Engl J Med 121. McGuffin P. Psychiatr Prax 2004;184:128–135.
1990;322:789–794. 2004;31 (Suppl 2):S189–S193. 145. Spaniel F, et al. Eur Psychiatry
100. Alfimova M, et al. Int 122. Reveley MA. Br J Psychiatry 2007;22:319–322.
J Psychophysiol 2003;49:201–216. 1985;146:367–371. 146. Ettinger U, et al. Arch Gen
101. Alfimova MV, et al. Span J Psychol 123. Dalton SO, et al. Am J Psychiatry Psychiatry 2007;64:401–409.
2009;12:46–55. 2004;161:903–908. 147. Borgwardt SJ, et al. Biol Psychiatry
102. Calkins ME, et al. Brain Cogn 124. Reveley AM, et al. Psychiatry Res 2010;67:956–964.
2008;68:436–461. 1984;13:261–266. 148. Fusar-Poli P, et al. Neurosci
103. Callicott JH, et al. Biol Psychiatry 125. Reveley AM, et al. Lancet Biobehav Rev 2007;31:
1998;44:941–950. 1982;1:540–541. 465–484.

255
Chapter 19: Genetics of schizophrenia

149. Baare WF, et al. V Arch Gen 171. Slater E. Acta Genetica et Statistica 194. Ekelund J, et al. Mol Psychiatry
Psychiatry 2001;58:33–40. Medica 1958;8:50–56. 2004;9:1037–1041.
150. Hulshoff Pol HE, et al. Biol 172. Garrone G. J Hum Genet 1962; 195. Ekelund J, et al. Hum Mol Genet
Psychiatry 2004;55:126–130. 11:89–100. 2000;9:1049–1057.
151. Hulshoff Pol HE, et al. 173. Heston LL. Science 1970;167: 196. Cai G, et al. Zhonghua Yi Xue
Am J Psychiatry 2002;159: 249–256. Yi Chuan Xue Za Zhi 2002;
244–250. 19:491–494.
174. Elston RC, et al. Behav Genet
152. Bartley AJ, et al. Biol Psychiatry 1970;1:3–10. 197. Shaw SH, et al. Am J Med Genet
1993;34:853–863. 1998;81:364–376.
175. Karlsson J. Br J Psychiatry
153. Weinberger DR, et al. Clin 1988;152:324–329. 198. Brzustowicz LM, et al. Am J Hum
Neuropharmacol 1992;15 Genet 2004;74:1057–1063.
176. Kay DWK, et al. Br J Psychiatry
(Suppl 1 Pt A):122A–123A. 199. Zheng Y, et al. Biochem
1975;127:109–118.
154. Noga JT, et al. Schizophr Res Biophys Res Commun
1996;22:27–40. 177. Elston RC, et al. Neuropsychobiology 2005;328:809–815.
1978;4:193–206.
155. Mill J, et al. Am J Hum Genet 200. Puri V, et al. Biol Psychiatry
2008;82:696–711. 178. O’Rourke DH, et al. Am J Hum 2006;59:195–197.
Genet 1982;34:630–649.
156. Kaminsky ZA, et al. Nat Genet 201. Rizig MA, et al. Am J Med Genet
2009;41:240–245. 179. Tsuang MT, et al. Br J Psychiatry 2006;141B:296–300.
1982;140:595–599.
157. Rosenthal D, et al. Am J Psychiatry 202. Talkowski ME, et al. Biol
1971;128:307–311. 180. Risch N, et al. Am J Hum Genet Psychiatry 2006;60:152–162.
1984;36:1039–1059.
158. Kety SS, et al. The Transmission of 203. Puri V, et al. Biol Psychiatry
Schizophrenia. Oxford, UK: 181. Gottesman II, et al. Proc Natl Acad 2007;61:873–879.
Pergamon Press; 1969. Science U S A 1967;58:199–205.
204. Miranda A, et al. Schizophr Res
159. Kety SS, et al. Am J Psychiatry 182. Tsuang MT, et al. Br J Psychiatry 2006;82:283–285.
1971;128:302–306. 1983;143:572–577.
205. Addington AM, et al. Mol
160. Kety SS, et al. Arch Gen Psychiatry 183. Ungvari G. Acta Psychiatr Scand Psychiatry 2007;12:195–205.
1994;51:442–455. 1983;68:287–296.
206. Puri V, et al. Eur J Hum Genet
161. Ingraham LJ, et al. Am J Med 184. Matthysse SW, et al. Am 2008;16:1275–1282.
Genet 2000;97:18–22. J Psychiatry 1976;133:185–191. 207. Millar JK, et al. Genomics 2000;
162. Cardno AG, et al. Am J Psychiatry 185. Baron M. Acta Psychiatr Scand 67:69–77.
2002;159:539–545. 1982;65:263–275. 208. St Clair D, et al. Lancet 1990;
163. Tienari P, et al. Acta Psychiatr 186. Carter CL, et al. Hum Hered 336:13–16.
Scand Suppl 1985;319:19–30. 1980;30:350–356. 209. Chubb JE, et al. Mol Psychiatry
164. Tienari P, et al. Br J Psychiatry 187. Bassett AS. Br J Psychiatry 2008;13:36–64.
Suppl 2004;184:216–222. 1992;161:323–334. 210. Hennah W, et al. Hum Mol Genet
165. Rao DC, et al. Hum Hered 188. Bassett AS, et al. Am J Med Genet 2003;12:3151–3159.
1981;31:325–333. 2000;97:45–51. 211. Cannon TD. Curr Opin Psychiatry
166. Kendler KS. Am J Psychiatry 189. Kosower NS, et al. Am J Med 2005;18:135–140.
1983;140:1413–1425. Genet 1995;60:133–138. 212. Callicott JH, et al. Proc Natl
167. McGue M, et al. Genet Epidemiol 190. Brzustowicz LM, et al. Science Acad Sci U S A 2005;102:
1985;2:99–110. 2000;288:678–682. 8627–8632.
168. McGuffin P, et al. Br J Psychiatry 191. Blaveri E, et al. Eur J Hum Genet 213. Hodgkinson CA, et al. Am J Hum
1994;164:593–599. 2001;9:469–472. Genet 2004;75:862–872.
169. Baron M. Biol Psychiatry 192. Lewis CM, et al. Am J Hum Genet 214. Maeda K, et al. Biol Psychiatry
1986;21:1051–1066. 2003;73:34–48. 2006;60:929–935.
170. Book JA. Acta Genetica 1953; 193. Hovatta I, et al. Am J Hum Genet 215. Thomson PA, et al. Neurosci Lett
4:345–414. 1999;65:1114–1124. 2005;389:41–45.

256
Chapter 19: Genetics of schizophrenia

216. Liu YL, et al. Schizophr Res 240. Bassett AS, et al. Lancet 264. Campion D, et al. Psychiatry Res
2006;87:15–20. 1988;1:799–801. 1992;41:99–105.
217. Chen QY, et al. J Psychiatr Res 241. Kalsi G, et al. Ann Hum Genet 265. Hwu HG, et al. Am J Med Genet
2006;41:428–434. 1999;63:235–247. 2000;96:74–78.
218. Hashimoto R, et al. Hum Mol 242. Suarez BK, et al. Am J Hum Genet 266. Schwab SG, et al. Nat Genet
Genet 2006;15:3024–3033. 2006;78:315–333. 1995;11:325–327.
219. DeRosse P, et al. Biol Psychiatry 243. Straub RE, et al. Mol Psychiatry 267. Straub RE, et al. Nat Genet
2007;61:1208–1210. 1997;2:148–155. 1995;11:287–293.
220. Kilpinen H, et al. Mol Psychiatry 244. Devlin B, et al. Mol Psychiatry 268. Moises HW, et al. Nat Genet
2008;13:187–196. 2002;7:689–694. 1995;11:321–324.
221. Wood LS, et al. Biol Psychiatry 245. Paunio T, et al. Hum Mol Genet 269. Gurling H, et al. Nat Genet
2007;61:1195–1199. 2001;10:3037–3048. 1995;11:234–235.
222. Palo OM, et al. Hum Mol Genet 246. Gurling HM, et al. Am J Hum 270. Coon H, et al. Am J Med Genet
2007;16:2517–2528. Genet 2001;68:661–673. 1994;54:59–71.
223. Burdick KE, et al. Neuroreport 247. Levinson DF, et al. Am J Hum 271. Straub RE, et al. Cold Spring
2005;16:1399–1402. Genet 2000;67:652–663. Harb Symp Quant Biol 1996;
224. Hennah W, et al. Mol Psychiatry 248. Schwab SG, et al. Mol Psychiatry 61:823–833.
2005;10:1097–1103. 1997;2:156–160. 272. Straub RE, et al. Am J Hum Genet
225. Hennah W, et al. Schizophr Bull 249. Pimm J, et al. Am J Hum Genet 2002;71:337–348.
2006;32:409–416. 2005;76:902–907. 273. van den Oord EJ, et al. Mol
226. O’Donovan MC, et al. Nat Genet 250. ISC. Nature 2009;460: Psychiatry 2003;8:499–510.
2008;40:1053–1055. 748–752. 274. Mutsuddi M, et al. Am J Hum
227. Riley B, et al. Mol Psychiatry 251. Liou YJ, et al. Schizophr Res Genet 2006;79:903–909.
2010;15:29–37. 2006;84:236–243. 275. Schwab SG, et al. Am J Hum Genet
228. Steinberg S, et al. Mol Psychiatry 252. Tang RQ, et al. Mol Psychiatry 2003;72:185–190.
2011;16:59–66. 2006;11:395–399. 276. Tang JX, et al. Mol Psychiatry
229. Crocq MA, et al. J Med Genet 253. Gurling H, et al. Schizophr Res 2003;8:717–718.
1992;29:858–860. 2006;89:357–359. 277. Van Den Bogaert A, et al. Am
230. Kennedy JL, et al. Am J Med Genet 254. Escamilla M, et al. Schizophr Res J Hum Genet 2003;73:1438–1443.
1995;60:558–562. 2008;106:253–257. 278. Williams NM, et al. Arch Gen
231. Mant R, et al. Am J Med Genet 255. Petryshen TL, et al. Mol Psychiatry Psychiatry 2004;61:336–344.
1994;54:21–26. 2005;10:1074–1088. 279. Morris DW, et al. Schizophr Res
232. Kalsi G, et al. Psychiatr Genet 256. Ikeda M, et al. Biol Psychiatry 2003;60:167–172.
1998;8:187–189. 2005;58:440–445. 280. Kirov G, et al. Biol Psychiatry
233. Jonsson EG, et al. Psychiatr Genet 257. Lo WS, et al. Biol Psychiatry 2004;55:971–975.
2003;13:1–12. 2007;61:653–660. 281. Funke B, et al. Am J Hum Genet
234. Jonsson EG, et al. Psychiatr Genet 258. Inada T, et al. Pharmacogenet 2004;75;891–898.
2004;14:9–12. Genomics 2008;18:317–323. 282. Hall D, et al. Genes Brain Behav
235. Nothen MM, et al. J Med Genet 259. Liu J, et al. Biochem Biophys Res 2004;3:240–248.
1993;30:708. Commun 2005;334:817–823. 283. Kohn Y, et al. Am J Med Genet
236. Sabate O, et al. Am J Psychiatry 260. Yu Z, et al. Clin Biochem B Neuropsychiatr Genet
1994;151:107–111. 2006;39:210–218. 2004;128:65–70.
237. Wiese C, et al. Psychiatry Res 261. Allen NC, et al. Nat Genet 284. Numakawa T, et al. Hum Mol
1993;46:69–78. 2008;40:827–834. Genet 2004;13:2699–2708.
238. DiBella D, et al. Psychiatr Genet 262. McGuffin P, et al. Psychol Med 285. Li T, et al. Mol Psychiatry
1994;4:39–42. 1983;13:31–43. 2005;10:1037–1044.
239. Sherrington R, et al. Nature 263. Andrew B, et al. Psychol Med 286. Tochigi M, et al. Neurosci Res
1988;336:164–167. 1987;17:363–370. 2006;56:154–158.

257
Chapter 19: Genetics of schizophrenia

287. Datta SR, et al. Behav Brain Funct 308. Tochigi M, et al. Schizophr Res 330. Li T, et al. Am J Med Genet
2007;3:50. 2006;88:260–264. B Neuropsychiatr Genet
288. Benson MA, et al. J Biol Chem 309. Goldberger C, et al. Am J Med 2004;129B:13–15.
2001;276:24232–24241. Genet B Neuropsychiatr Genet 331. Bakker SC, et al. Mol Psychiatry
289. Talbot K, et al. J Clin Invest 2005;137B:51–55. 2004;9:1061–1063.
2004;113:1353–1363. 310. Shifman S, et al. PLoS Genet 332. Tang JX, et al. Mol Psychiatry
290. Stefansson H, et al. Nature 2008;4:e28. 2004;9:11–12.
2009;460:744–747. 311. Pisante A, et al. Psychiatr Genet 333. Zhao X, et al. J Med Genet 2004;
291. Duan J, et al. Am J Hum Genet 2009;19:212. 41:31–34.
2004;75:624–638. 312. Chen ML, et al. Mol Psychiatry 334. Petryshen TL, et al. Mol Psychiatry
292. Duan S, et al. J Neural Transm 2002;7:447–448. 2005;10:366–374.
2006;113:381–385. 313. Huang CH, et al. Psychiatry Res 335. Hashimoto R, et al. Mol Psychiatry
2006;142:89–92. 2004;9:299–307.
293. Ikeda M, et al. Schizophr Res
2005;78:127–130. 314. Blouin JL, et al. Nat Genet 336. Iwata N, et al. Mol Psychiatry
1998;20:70–73. 2004;9:126–127.
294. Knight HM, et al. Am J Hum
Genet 2009;85:833–846. 315. Brzustowicz LM, et al. Am J Hum 337. Fallin MD, et al. Am J Hum Genet
Genet 1999;65:1096–1103. 2004;75:204–219.
295. Jonsson EG, et al.
Neuropsychobiology 2009; 316. Takahashi S, et al. Psychiatry Res 338. Thiselton DL, et al. Mol Psychiatry
59:142–150. 2005;133:111–122. 2004;9:777–83; image 729.
317. Kendler KS, et al. Am J Psychiatry 339. Datta SR, et al. Mol Psychiatry
296. Albalushi T, et al. Am J Med
Genet B Neuropsychiatr 1996;153:1534–1540. 2010;15:615–628.
Genet 2008;147:392–396. 318. Lin HF, et al. Psychol Med 340. Sullivan PF, et al. Mol Psychiatry
2005;35:1589–1598. 2008;13:570–584.
297. Bishop JR, et al. Psychiatr Genet
2007;17:358. 319. Gerber DJ, et al. Proc Natl 341. Need AC, et al. Hum Mol Genet
Acad Sci U S A 2003;100: 2009;18:4650–4661.
298. Chen Q, et al. Schizophr Res
2005;73:21–26. 8993–8998. 342. Kamiya A, et al. Arch Gen
320. Gurling H, et al. Arch Psychiatry 2008;65:996–1006.
299. Egan MF, et al. Proc Natl Acad Sci
U S A 2004;101:12604–12609. Gen Psychiatry 2006;63: 343. ISC. Nature 2008;455:237–241.
844–854. 344. Eastwood SL, et al. Mol Psychiatry
300. Fujii Y, et al. Psychiatr Genet
2003;13:71–76. 321. Katsu T, et al. Neurosci Lett 2009;14:556–557.
2003;353:53–56. 345. Kinoshita Y, et al.
301. Harrison PJ, et al.
J Psychopharmacol 2008; 322. Nakata K, et al. Biol Psychiatry J Neural Transm 2005;112:
22:308–322. 2003;53:571–576. 1255–1262.
302. Lyon L, et al. Synapse 323. Stefansson H, et al. Am J Hum 346. Fallin MD, et al. Am J Hum Genet
2008;62:842–850. Genet 2002;71:877–892. 2005;77:918–936.

303. Mossner R, et al. Eur 324. Yang JZ, et al. Biol Psychiatry 347. Shibasaki F, et al. J Biochem
Neuropsychopharmacol 2008; 2003;54:1298–1301. 2002;131:1–15.
18:768–772. 325. Yang JZ, et al. Mol Psychiatry 348. Klee CB, et al. J Biol Chem
304. Norton N, et al. BMC Psychiatry 2003;8:706–709. 1998;273:13367–13370.
2005;5:23. 326. Stefansson H, et al. Am J Hum 349. Zhang Y, et al. Am J Med Genet
Genet 2003;72:83–87. B Neuropsychiatr Genet
305. Sartorius LJ, et al.
2004;129:16–19.
Neuropsychopharmacology 327. Williams NM, et al.
2008;33:2626–2634. Mol Psychiatry 2003;8: 350. Wei J, et al. Neurosci Lett
485–487. 2004;366:336–338.
306. Schwab SG, et al. Psychiatr Genet
2008;18:25–30. 328. Corvin AP, et al. Mol Psychiatry 351. Ide M, et al. Biol Psychiatry
307. Tan HY, et al. Proc Natl 2004;9:208–213. 2004;56:462–465.
Acad Sci U S A 2007;104; 329. Hong CJ, et al. Neurosci Lett 352. Hashimoto R, et al. J Neural
12536–12541. 2004;366:158–161. Transm 2005;112:303–307.

258
Chapter 19: Genetics of schizophrenia

353. Hong LE, et al. Am J Med 376. Schindler KM, et al. Mol 398. Goldberg TE, et al.
Genet B Neuropsychiatr Genet Psychiatry 2002;7:1002–1005. Neuropsychopharmacology
2005;136:8–11. 377. Wu SN, et al. Acta Pharmacol Sin 2006;31:2022–2032.
354. Nanko S, et al. Jpn J Psychiatry 2006;27:966–970. 399. Lavedan C, et al. Mol Psychiatry
Neurol 1992;46:155–159. 2009;14:804–819.
378. Chen CH, et al. Am J Med Genet
355. Maziade M, et al. Am J Med Genet 1996;67:418–420. 400. Pickard BS, et al. Mol Psychiatry
1995;60:522–528. 2009;14:874–884.
379. Crawford F, et al. Am J Med Genet
356. Gill M, et al. Psychol Med 1996;67:483–484. 401. Pickard BS, et al. Am J Med
1993;23:27–44. Genet B Neuropsychiatr Genet
380. Kaneshima M, et al. Psychiatry
357. Barr C, et al. Psychiatr Genet Clin Neurosci 1997;51:379–381. 2005;136B:26–32.
1991;2:66–72. 402. Pickard BS, et al. Ann Med
381. Spurlock G, et al. Am J Med Genet
358. Muir WJ, et al. Psychiatr Genet 1998;81:24–28. 2006;38:439–448.
1991;2:18. 403. Macintyre G, et al. Schizophr Res
382. Jonsson EG, et al. Am J Med
359. Su Y, et al. Arch Gen Psychiatry Genet B Neuropsychiatr Genet 2010;120:143–149.
1993;50:205–211. 2003;119B:28–34. 404. Emamian ES, et al. Nat Genet
360. Choudhury K, et al. Am J Hum 383. Camp NJ, et al. Am J Hum Genet 2004;36:131–137.
Genet 2007;80:664–672. 2001;69:1278–1289. 405. Ikeda M, et al. Biol Psychiatry
361. Beguin P, et al. Embo J 384. Faraone SV, et al. Expert Rev Mol 2004;56:698–700.
2002;21:3264–3273. Med 2002;4:1–13. 406. Schwab SG, et al. Biol Psychiatry
362. Crambert G, et al. Sci STKE 385. Lin MW, et al. Psychiatr Genet 2005;58:446–450.
2003;161: RE1. 1995;5:117–126. 407. Bajestan SN, et al. Am J Med Genet
363. Garty H, et al. Ann Rev Physiology 386. Pulver AE, et al. Mol Psychiatry B Neuropsychiatr Genet
2006;68:431–459. 2000;5:650–653. 2006;141B:383–386.
364. Kadowaki K, et al. Brain Res Mol 387. DeLisi LE, et al. Psychiatr Genet 408. Norton N, et al. Schizophr Res
Brain Res 2004;125:105–112. 2000;10:153–158. 2007;93:58–65.
365. Ambrosio AM, et al. Psychiatry 388. Williams NM, et al. Hum Mol 409. Xu MQ, et al. J Clin Psychiatry
Res 2004;125:185–191. Genet 1999;8:1729–1739. 2007;68:1358–1367.
366. Cordeiro Q, et al. Arq 389. Chumakov I, et al. Proc Natl 410. Thiselton DL, et al. Biol Psychiatry
Neuropsiquiatr 2009;67:191–194. Acad Sci U S A 2002;99: 2008;63:449–457.
367. Dubertret C, et al. Schizophr Res 13675–13680.
411. Lee KY, et al. Neurosci Res
2004;67:75–85. 390. Shi J, et al. Schizophr Res 2010;66:238–245.
368. Fan H, et al. Neurosci Lett 2008;98:89–97.
2009;477:53–56. 412. Turunen JA, et al. Schizophr Res
391. Korostishevsky M, et al. Am J Med 2007;91:27–36.
369. Glatt SJ, et al. Mol Psychiatry Genet B Neuropsychiatr Genet
2009;14:885–893. 2006;141B:91–95. 413. Ide M, et al. J Neurochem
2006;99:277–287.
370. Glatt SJ, et al. Mol Psychiatry 392. Shin HD, et al. Schizophr Res
2003;8:911–915. 2007;96:119–124. 414. Liu YL, et al. Psychiatr Genet
2006;16:39–41.
371. Glatt SJ, et al. Am J Med Genet 393. Shinkai T, et al. Neuromolecular
B Neuropsychiatr Genet Med 2007;9:169–177. 415. Ohtsuki T, et al. Mol Psychiatry
2006;141B:149–154. 2004;9:981–983.
394. Corvin A, et al. Am J Med Genet
372. Hanninen K, et al. Neurosci Lett B Neuropsychiatr Genet 416. Sanders AR, et al. Am J Psychiatry
2006;407:195–198. 2007;144B:949–953. 2008;165:497–506.
373. Kukreti R, et al. Neurosci Lett 395. Williams NM, et al. Arch Gen 417. Coon H, et al. Am J Med Genet
2006;392:68–71. Psychiatry 2006;63:366–373. 1994;54:72–79.
374. Monakhov M, et al. Schizophr Res 396. Bass NJ, et al. Behav Brain Funct 418. Pulver AE, et al. Am J Med Genet
2008;100:302–307. 2009;5:28. 1994;54:44–50.
375. Parsons MJ, et al. Psychiatr Genet 397. Mulle JG, et al. Mol Psychiatry 419. Pulver AE, et al. Am J Med Genet
2007;17:159–163. 2005;10:431–433. 1994;54:36–43.

259
Chapter 19: Genetics of schizophrenia

420. Vallada H, et al. Psychiatr Genet 442. Shifman S, et al. Am J Hum Genet 466. Vine AE, et al. Psychiatr Genet
1995;5:127–130. 2002;71:1296–1302. 2009;19:165–170.
421. Kalsi G, et al. Am J Med Genet 443. Rudduck C, et al. Hereditas 467. Zhang D, et al. Mol Psychiatry
1995;60:298–301. 1983;98:297–299. 2009;14:376–380.
422. Polymeropoulos MH, et al. 444. Chodirker BN, et al. Clin Genet 468. McQuillin A, et al. Eur J Hum
Am J Med Genet 1994; 1987;31:1–6. Genet 2011;19:588–592.
54:93–99. 445. Garofalo G, et al. Am J Psychiatry 469. Grozeva D, et al. Arch
423. Gill M, et al. Am J Med Genet 1992;149:1116. Gen Psychiatry 2010;67:
1996;67:40–45. 446. Axelsson R, et al. Hereditas 318–327.
424. Severinsen JE, et al. Mol Psychiatry 1981;95:337. 470. Prata DP, et al. Psychiatr Genet
2006;11:1126–1138. 447. Hong ML. Chinese J Neurol 2009;19:113–116.
425. Nyegaard M, et al. Am J Med Psychiatry 1986;19:188–191. 471. Goes FS, et al. Am J Med
Genet B Neuropsychiatr Genet 448. Turner B, et al. Lancet 1961; Genet B Neuropsychiatr Genet
2009;150B:151–152. ii:49–50. 2009;150B:693–702.
426. Bassett AS, et al. Biol Psychiatry 449. Sperber MA. Biol Psychiatry 472. Hattori E, et al. Am J Hum Genet
1999;46:882–891. 1975;10:27–43. 2003;72:1131–1140.
427. Bassett AS, et al. Curr Psychiatry 450. Kaplan AR. Biol Psychiatry 473. Prata D, et al. Am J Med Genet
Rep 2008;10:148–157. 1970;2:89–94. B Neuropsychiatr Genet
428. Bassett AS, et al. Am J Psychiatry 2008;147B:914–917.
451. Dasgupta J, et al. Indian J Med Res
2003;160:1580–1586. 1973;61:62–70. 474. Zhang Z, et al. Psychiatr Genet
429. Bassett AS, et al. Am J Med Genet 2009;19:151–153.
452. Stefansson H, et al. Nature
2005;138:307–313. 475. Chen YS, et al. Mol Psychiatry
2008;455:232–236.
2004;9:87–92; image 5.
430. Chow EW, et al. Am J Med Genet 453. Walsh T, et al. Science
1994;54:107–112. 476. Hattori M, et al. Am J Med Genet
2008;320:539–543.
2002;114:304–309.
431. Liu H, et al. Proc Natl Acad 454. Kirov G, et al. Hum Mol Genet
Sci U S A 2002;99: 477. Hattori M, et al. Hum Mol Genet
2008;17:458–465.
3717–3722. 1993;2:1511.
455. Xu B, et al. Nat Genet
432. Kempf L, et al. PLoS Genet 2008; 478. Hattori M, et al. Biochem Biophys
2008;40:880–885.
4:e1000252. Res Commun 1995;209:513–518.
456. Rujescu D, et al. Hum Mol Genet 479. Kunugi H, et al. Schizophr Res
433. Fan JB, et al. Neurosci Lett 2009;18:988–996.
2003;338:252–254. 1999;37:271–273.
457. Helbig I, et al. Nat Genet 480. Kunugi H, et al. Neurosci Lett
434. Glaser B, et al. Schizophr Res
2009;41:160–162. 1998;241:65–67.
2006;87:21–27.
458. de Kovel CG, et al. Brain 481. Nanko S, et al. Acta Psychiatr
435. Williams HJ, et al. Mol Psychiatry
2010;133:23–32. Scand 1994;89:390–392.
2003;8:644–645.
459. Dibbens LM, et al. Hum Mol 482. Jones CT, et al. Nat Genet
436. Williams HJ, et al. Am J Med
Genet 2009;18:3626–3631. 1992;1:306–309.
Genet B Neuropsychiatr Genet
2003;120B:42–46. 460. Heinzen EL, et al. Am J Hum 483. Arnholt JC, et al. Biol Psychiatry
437. Gogos JA, et al. Nat Genet Genet 2010;86:707–718. 1993;34:739–740.
1999;21:434–439. 461. Segurado R, et al. Am J Hum 484. Carter D, et al. Hum Mol Genet
438. Paterlini M, et al. Nat Neurosci Genet 2003;73:49–62. 1993;2:321.
2005;8:1586–1594. 462. McQueen MB, et al. Am J Hum 485. Coon H, et al. Am J Med Genet
Genet 2005;77:582–595. 1993;48:36–39.
439. Lachman HM, et al.
Pharmacogenetics 1996; 463. Ng MY, et al. Mol Psychiatry 486. Forsell C, et al. Neurosci Lett
6:243–250. 2009;14:774–785. 1995;184:90–93.
440. Glatt SJ, et al. Am J Psychiatry 464. Ferreira MA, et al. Nat Genet 487. Fukuda R, et al. Jpn J Hum Genet
2003;160:469–476. 2008;40:1056–1058. 1993;38:407–411.
441. Munafo MR, et al. Mol Psychiatry 465. Lencz T, et al. Proc Natl Acad Sci 488. Jonsson E, et al. Biol Psychiatry
2005;10:765–770. U S A 2007;104;19942–19947. 1995;37:135–136.

260
Chapter 19: Genetics of schizophrenia

489. Morris S, et al. Psychiatr Genet 500. Miyakawa T, et al. Proc Natl 509. D’Arcangelo G. Epilepsy Behav
1994;4:23–27. Acad Sci U S A 2003;100: 2006;8:81–90.
490. Mortilla M, et al. Neurosci Lett 8987–8992. 510. Costa E, et al. Mol Interv
1994;165:45–47. 501. Karlsson RM, et al. Neuro- 2002;2:47–57.
psychopharmacology 2009; 511. Eastwood SL, et al. Mol Psychiatry
491. Arranz M, et al. Lancet 1995;
34:1578–1589. 2003;8:769, 821–831.
346:281–282.
502. Young JW, et al. Eur Neuro- 512. Costa E, et al. Curr Opin
492. Arranz MJ, et al. Neurosci Lett
psychopharmacol 2007; Pharmacol 2002;2:56–62.
1996;217:177–178.
17:145–155. 513. Grayson DR, et al. Proc Natl
493. Joober R, et al. J Psychiatry
503. Chen YJ, et al. J Neurosci Acad Sci U S A 2005;102:9341–9346.
Neurosci 1999;24:141–146.
2008;28:6872–6883. 514. Woolley ML, et al. Psycho-
494. Lin CH, et al. Neuroreport 1999; pharmacology (Berl) 2008;
504. van den Buuse M, et al. Int
10:57–60. 196:431–440.
J Neuropsychopharmacol
495. Koike H, et al. Proc Natl Acad Sci 2009;12:1383–1393. 515. Ayhan Y, et al. Behav Brain Res
U S A 2006;103:3693–3697. 2009;204:274–281.
505. Andrieux A, et al. Genes Dev
496. Clapcote SJ, et al. Genetics 2002;16:2350–2364. 516. McGuire PK, et al. Schizophr Res
2006;173:2407–2410. 1995;15:277–281.
506. Fradley RL, et al. Behav Brain Res
497. Clapcote SJ, et al. Neuron 2005;163:257–264. 517. Arendt M, et al. Arch Gen
2007;54:387–402. Psychiatry 2008;65:1269–1274.
507. Eastwood SL, et al. J
498. Shen S, et al. J Neurosci Psychopharmacol 2007; 518. Boydell J, et al. Schizophr Res
2008;28:10893–10904. 21:635–644. 2007;93:203–210.
499. Zeng H, et al. Cell 2001; 508. Brigman JL, et al. Behav Neurosci 519. Phillips LJ, et al. Aust N Z
107:617–629. 2006;120:984–988. J Psychiatry 2002;36;800–806.

261
The genetics of anorexia and
Chapter

20 bulimia nervosa
Andrew W. Bergen, Jennifer Wessel, and Walter H. Kaye

Overview of anorexia nervosa [6]. There has been much conjecture regarding the
processes underlying pathological eating, but little is
and bulimia nervosa understood of their biology [7, 8]. Similar to AN,
Anorexia nervosa (AN) and bulimia nervosa (BN) individuals with AN-BN and BN have a seemingly
are related disorders of unknown etiology that most relentless drive to restrain food intake, an extreme
commonly begin during adolescence in women. fear of weight gain and a distorted view of their body
They are frequently chronic and often disabling con- shape. In contrast to AN, however, AN-BN and BN
ditions that are characterized by aberrant patterns of suffer recurring disinhibition of dietary restraint,
feeding behavior and weight regulation, and deviant resulting in cycles of binge eating and compensatory
attitudes and perceptions toward body weight and actions such as self-induced vomiting. Transitions
shape. In AN, an inexplicable fear of weight gain between AN and BN occur frequently, and these
and unrelenting obsession with fatness, even in the disorders are often cross-transmitted in families
face of increasing cachexia, accounts for a protracted [9–12]. Consequently it has been argued that AN
course, extreme medical and psychological morbid- and BN share some risk and liability factors.
ity, and standardized mortality rates exceeding those Because AN and BN present most often during
of all other psychiatric disorders. BN usually emerges adolescence in women, they are often theorized to be
after a period of food restriction, which may or may caused by cultural pressures for thinness [13] since
not have been associated with weight loss. Binge dieting and the pursuit of thinness are common in
eating is followed by either self-induced vomiting, industrialized countries. Still, AN and BN affect only
or by some other means of compensation for the an estimated 0.3–0.7% and 1.5–2.5%, respectively,
excess of food ingested. Although abnormally low of females in the general population [14]. This dispar-
body weight is an exclusion for the diagnosis of ity between the high prevalence of pressures for thin-
BN, some 25–30% of bulimics have a prior history ness and the low prevalence of eating disorders (EDs),
of AN. combined with clear evidence of AN occurring at least
AN and BN are subdivided by eating behavior several centuries ago [15], the stereotypic presenta-
and psychopathological characteristics [1–5]. AN is tion, substantial heritability, and developmentally
characterized by severe emaciation [6]. Two types of specific age-of-onset distribution, underscores the
consummatory behavior are seen in AN. Restricting- possibility of contributing biological vulnerabilities.
type anorexics (“AN”) lose weight purely by restricted Improvements in the understanding and treat-
dieting and have no history of binge eating or purging. ment of EDs are of immense clinical and public health
Binge-eating/purging-type anorexics (AN-BN) also importance [16] since these are often chronic, relaps-
restrict their food intake to lose weight, but have a ing illnesses [17–19] with substantial and costly med-
periodic disinhibition of restraint and engage in binge ical morbidity [20]. Importantly, for AN, there is
eating and/or purging. Individuals with bulimia ner- no proven treatment that reverses symptoms [16];
vosa (“BN”) do not become emaciated and are able to consequently it has the highest death rate of any
maintain an average body weight (ABW) above 85% psychiatric illness [21].

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

262
Chapter 20: Genetics of anorexia and bulimia nervosa

anxiety and depressive disorders, and obsessive–


ED symptoms compulsive disorder [10, 24–26]. In addition, individ-
The DSM-IV diagnostic criteria for AN and BN focus uals with AN and BN are both consistently character-
on eating behavior and body image distortions. ized by perfectionism, obsessive–compulsiveness,
Because of their unusual and prominent nature, these neuroticism, negative emotionality, harm avoidance,
symptoms tend to capture much attention. The low self-directedness, low cooperativeness, and traits
pathogenesis of the disturbed eating behaviors is associated with avoidant personality disorder. Con-
poorly understood [7, 8]. Individuals with AN rarely sistent differences that emerge between ED groups
have complete suppression of appetite, but rather are high constraint and persistence, low novelty seek-
exhibit an ego-syntonic resistance to feeding drives ing, constriction of affect and emotional expressive-
while simultaneously being preoccupied with food ness, anhedonia and asceticism, and reduced social
and eating rituals to the point of obsession. Individ- spontaneity in restrictor-type AN. Individuals with
uals with AN severely restrict food intake, particularly BN are more likely to have high impulsivity, sensation
fats and carbohydrates, but rarely stop eating com- seeking, novelty seeking, and traits associated with
pletely; rather they restrict their caloric intake to a borderline personality disorder in BN, and substance
few hundred calories a day. They tend to be vegetar- abuse [27].
ians, have monotonous choices in food intake, select Neurocognition: Individuals with AN have an
unusual combinations of foods and flavors, and obsessive, perseverative, and rigid personality style
have ritualized eating behaviors. Similarly, BN is not and have difficulty shifting sets. While those with
associated with a primary, pathological increase in AN do well on goal-directed behavior, they have
appetite; rather, like individuals with AN, individuals difficulties incorporating feedback and modifying
with BN have a seemingly relentless drive to restrain their behavior. For example, they often feel that they
their food intake, an extreme fear of weight gain, and should be able to do things perfectly without making
often have a distorted view of their actual body shape. mistakes, and they have little appreciation for the
Loss of control with overeating in individuals with fact that mistakes are a normal learning experience.
BN usually occurs intermittently and typically only Moreover, they often fail to accurately recognize
some time after the onset of dieting behavior. and incorporate affective and social stimuli in the
Restrained eating behavior and dysfunctional cogni- environment, as confirmed by laboratory tests [28,
tion relating weight and shape to self-concept are 29]. Those ill with AN and those recovered (REC)
shared by all types of patients with EDs. from AN tend [30] to have delayed setshifting, which
AN and BN individuals commonly have clusters normally allows for the adaptation of behavior in
of other puzzling symptoms. Excessive exercise and response to the environment. Furthermore, individ-
motor restlessness are common in AN [22]. While not uals with AN have enhanced ability to pay attention
well studied, excessive exercise is thought to be asso- to detail or use a logical/analytic approach but exhibit
ciated particularly with the purging subtype of AN, as worse performance with global strategies [29, 31].
well as with a constellation of anxious/obsessional State and trait: It has long been debated whether
temperament characteristics. Individuals with AN symptoms in individuals with AN and BN are cause
often have resistance to treatment [23]. In part this or consequence of malnutrition. Recent studies show
is due to the ego-syntonic nature of the disorder, that certain childhood temperament and personality
which is demonstrated by the patient’s denial of being traits [32–35] such as negative emotionality, harm
underweight and refusal to accept the seriousness avoidance, perfectionism, inhibition, drive for thin-
of the medical consequences of the disorder. Conse- ness, altered interoceptive awareness, and obsessive–
quently, few controlled trials of any therapy have been compulsive personality, create a vulnerability for
performed, in part, because it has been difficult to developing AN and BN. In other words, the majority
enlist cooperation of individuals with AN, and in part of people with AN and BN exhibit one or more of
because psychological and pharmacological strategies such traits in childhood, and thus these traits pre-date
that have been successful in other disorders appear to the onset of AN and BN. Moreover, studies done on
be less effective in this illness. three continents have shown that for AN and BN
Mood and impulse control: Individuals with AN individuals with a lifetime history of an anxiety dis-
and BN have elevated rates of lifetime diagnoses of order diagnosis, the anxiety disorder most often

263
Chapter 20: Genetics of anorexia and bulimia nervosa

began in childhood before the onset of the ED [36–39]. a window of vulnerability [47]. Thus these vulnerabil-
The most common [26] premorbid childhood dis- ities, which are biologically mediated, may significantly
orders were obsessive–compulsive disorder (OCD) enhance the risk of onset of an ED, particularly in
and social phobia. In summary, such symptoms may women.
be susceptibility factors that make people vulnerable
to developing an ED. Malnutrition tends to exagger-
ate these premorbid behavioral traits [40] after the Heritability
onset of the illness, with the addition of other symp- Considering that transitions between syndromes
toms that maintain or accelerate the disease process, occur in many individuals, it has been argued that
including exaggerated emotional dysregulation and AN and BN share at least some risk and liability
obsessionality [24, 26]. factors [7, 9]. In fact, AN and BN are cross-transmitted
The process of recovery in AN is poorly under- in families [10, 11]. Moreover there is an increased
stood and, in most cases, protracted. Still, approxi- prevalence of AN and BN as well as subthreshold
mately 50–70% of affected individuals will eventually forms of EDs in relatives, consistent with the possibil-
have complete or moderate resolution of the illness, ity of a continuum of transmitted liability in at risk
often in the early to mid 20s [41–43]. It is important families manifesting a broad spectrum of eating dis-
to emphasize that temperament and personality traits order phenotypes [11]. In fact, AN is highly familial
such as negative emotionality, harm avoidance and [10, 11, 48]. The relative risk for AN in family
perfectionism, obsessional behaviors (particularly members of probands with AN is 11.3 [11].
symmetry, exactness, and order) persist after recovery Twin studies of AN and BN suggest there is
from both AN and BN [41, 43–46] and are similar to approximately a 50–80% genetic contribution to liabil-
the symptoms described premorbidly in childhood. ity [9, 49–52] accounted for by additive genetic factors.
Compared to the ill state, symptoms in REC AN and These heritability estimates are similar to those found
BN tend to be mild to moderate, including elevated in schizophrenia and bipolar disorder, suggesting that
scores on core ED measures. Interestingly, REC AN AN and BN may be as genetically influenced as dis-
and BN tend to be more alike than different on many orders traditionally viewed as biological in nature.
of these measures, although there are some differ- Twin studies on European populations have
ences on factors related to impulse control or stimuli yielded heritability estimates using various strategies.
seeking, such as novelty seeking [34, 42–43]. First, heritability of AN was estimated to be 58% (95%
In summary, individuals with restricting type AN CI 0.33–0.84), in the context of a bivariate twin analy-
are more likely to have restricted eating, constricted sis with major depression [52]. Second, twin analyses
affect and emotional mood expression, and impulse were conducted for a single question of “have you
over-control, as well as personality traits of marked ever had AN?” yielding a heritability estimate of 48%
rigidity, conformity, and reduced social spontaneity. (95% CI 0.27–0.65) [53]. Third, broadening the def-
Individuals with BN may show similar traits, but in inition of AN syndrome, Klump et al. reported the
addition, may exhibit histories of episodic overeating, heritability to be 76% [51]. A Swedish Twin Registry
extremes of intense affect, and impulse dysregulation. study of 31 406 twins born between 1935 and 1958
Thus several domains (eating, affect, impulse control) and diagnosed by clinical interview, hospital dis-
are involved in systematic ways, specifically, over charge diagnosis of AN, or cause of death certificate
control in AN, and switches between over control yielded a heritability estimate of 56% (95% CI 0.00–
and under control in BN, which raises the question 0.87) with the remaining variance attributable to
of whether there is a disturbance of modulation of shared environment and unique environment [49].
multiple systems. AN and BN most commonly This latter possibility is supported by twin studies
develop during adolescence or young adulthood [19] which found essentially no genetic influence on over-
in proximity to puberty. Adolescence [47] is a time of all levels of ED symptoms in 11-year-old twins, but
profound biological, psychological, and sociocultural significant genetic effects (> 50%) in 17-year-old
change, and it demands a considerable degree of twins [54]. These findings collectively imply that
flexibility to successfully manage the transition into puberty may play a role in the genetic diathesis for
adulthood. Psychologically, change may challenge the ED symptoms. The changes associated with adoles-
rigidity of those at risk for AN and BN, and thus open cence differ in males and females and may therefore

264
Chapter 20: Genetics of anorexia and bulimia nervosa

contribute to the sexual dimorphism of AN. Menar- Serotonergic and opioidergic neurotransmitter
che is associated [47] with a rapid change in body system alterations have been observed in people with
composition and neuropeptides modulating metabol- eating disorders; the genes for the serotonin 1D recep-
ism. Little is known about whether the rise in estrogen tor (HTR1D) and the opioid delta receptor (OPRD1)
levels associated with puberty in females is contribu- are found on chr1p36.3–34.3. These candidate genes
tory. Estrogens modulate serotonergic function [55] were resequenced to identify or confirm single nucle-
as well as stress-related neuropeptides such as cortisol otide polymorphisms (SNPs), and four HTR1D
releasing hormone (CRH) [56] via a variety of mech- and five OPRD1 SNPs were tested for linkage and
anisms. Moreover, a major phase of synaptogenesis, association with DSM-IV anorexia diagnosis [67,
pruning and myelination of predominantly frontal 68]. Linkage analysis of these candidate gene SNPs
and limbic areas occurs around the time of puberty with 33 microsatellite markers in N ¼ 37 families
and adolescence and is thought to have a functional including relative pairs concordantly affected with
role in the integration of emotional processing with RAN substantially increased the evidence for linkage
cognition [57]. of this region to restricting AN to an NPL score
The transmitted liability to AN and BN may be of 3.91 (p ¼ 0.00002), which exceeds the Lander/
mediated by a more diffuse phenotype of continuous, Kruglyak threshold for significant linkage [69]. Asso-
heritable behavioral traits related to disordered eating ciation to DSM-IV AN diagnosis was performed
regulation [54, 58–60], often resulting in subthreshold using 196 families in a family-based association test
forms of ED in families [10, 11]. Substantial evidence and with 98 controls in a case-control design. Using
supports that many of these traits exist premorbidly, genotype data on parents and AN probands, 3 SNPs
are heritable, are elevated in unaffected family at HTR1D were found to exhibit significant transmis-
members, persist after recovery from the disorder, sion disequilibrium (p < 0.05). Statistically significant
and are independent of body weight [32, 43, 44, 49, genotypic, allelic, and haplotypic association to AN in
61–64]. Therefore, we postulate that these traits, such the case-control design was observed at HTR1D and
as anxiety, obsessions, or perfectionism, confer liabil- OPRD1 with allelic effect sizes for individual SNPs of
ity to the development of AN. 2.63 (95% CI ¼ 1.21–5.75) for HTR1D rs6300 and 1.46
(95% CI ¼ 1.01–2.13) for OPRD1 rs536706. An inde-
pendent association study [70] of 226 females meeting
Linkage studies of EDs DSM-IV criteria for AN, and 678 matched controls
The Price Foundation, a private, European-based genotyped 4 SNPs in HTR1D and 6 SNPs in OPRD1.
foundation, has supported a multicenter international One HTR1D SNP overlapped between the studies
collaboration to investigate the genetics of AN and BN. (rs674386 or -1123T > C). Three OPRD1 SNPs were
In the first study, 192 families were ascertained. All found to be associated with both RAN and binge-
probands met modified DSM-IV criteria for AN; at purge AN (BPAN), and two HTR1D SNPs exhibited
least one additional affected first- through fourth- significant association with RAN with allelic effect
degree relative met DSM-IV criteria for AN, BN, or sizes for individual SNPs of 1.51 (95% CI ¼ 1.08,
eating disorder not otherwise specified (EDNOS) [65]. 2.10) for HTR1D rs856510 and 1.77 (1.20, 2.62) for
Blood for DNA was collected from all affected individ- OPRD1 rs569356. These data support the hypothesis
uals and available biological parents. Factors poten- that polymorphisms within HTR1D and OPRD1 form
tially affecting susceptibility for AN were assessed with a component of the genetic basis of susceptibility
a battery of standardized and validated instruments. to AN.
Using the Weber screening set, version 9 (Center for In an exploration of how behavioral covariates
Medical Genetics, Marshfield Medical Research Foun- enhanced the linkage signals, Devlin and colleagues
dation) with markers dispersed across the genome at [71] evaluated features of ED for the following
approximately 10 cM, and analyzing families in which at criteria: (1) consistent relationship to eating path-
least two affected relative pairs had AN, restricting sub- ology; (2) heritability; and (3) relationship to sever-
type (RAN) (N ¼ 37 families, 32 sibling pairs of which ity of some aspect of the disorder. Two variables,
11 pairs had data for both parents) we found evidence drive-for-thinness and obsessionality (OBS), each
for linkage, with a nonparametric linkage (NPL) score of yielded a cluster of affected sibling pairs who had
3.03 at marker D1S3721 on chromosome 1p [66]. high and concordant values for these traits, whereas

265
Chapter 20: Genetics of anorexia and bulimia nervosa

other sibling pairs were notably discordant. Incorpor-


ation of these traits into covariate-based linkage ana-
Candidate gene association
lyses [72] yielded a significant additional linkage studies in AN
signal on 1q, with a LOD score of 3.46, p ¼ 0.00003, In addition to the candidate gene studies motivated by
marker D1S1660, as well as two other suggestive link- linkage analysis findings, the use of the neurobio-
age signals, one at 2p (LOD ¼ 2.22, p ¼ 0.0007, logical approach to identify, genotype, and associate
marker D251790), and another at 13q (LOD ¼ 2.50, candidate genes with AN diagnosis has been used
p ¼ 0.00035, marker D135894). since 1997 to explore the potential involvement of
In further exploration of this linkage sample (154 candidate genes with AN susceptibility. These studies
affected sibling pairs) and an additional BN linkage most often included candidate genes encoding pro-
sample (244 affected sibling pairs), Bulik and col- teins involved in monoaminergic function, including
leagues [58] thoroughly explored eating disorder- receptors and interacting proteins (dopamine recep-
related traits. From more than 100 psychiatric, per- tors 2, 3, and 4 and serotonin receptors 1B, 2A,
sonality, and temperament phenotypes, they selected 2C, and 7), transporters (norepinephrine, dopamine,
a parsimonious subset of attributes to incorporate and serotonin), metabolic enzymes (catecholamine
into linkage analyses. Using a multilayer decision transferase, monoamine oxidase A, tryptophan
analysis, they chose variables relevant to eating hydroxylase) or neurotrophic factors (brain-derived
disorder pathology with published evidence for neurotrophic factor). These candidate gene studies
heritability. OBS, age-at-menarche, and a composite are characterized by the selection and genotyping of
anxiety measure (ANX) displayed features of herit- a limited number of polymorphisms, in an era of
able quantitative traits, such as normal distribution limited information on candidate gene variation and
and familial correlation, and thus appeared ideal linkage disequilibrium, and of smaller sample sizes
for quantitative trait locus linkage analysis. By con- that would be considered necessary for discovery of
trast, some families showed highly concordant and novel associations today. Some studies evaluating
extreme values for three variables – lifetime min- dopaminergic and serotonergic candidate genes for
imum body mass index (BMI) (lowest BMI attained association with AN are reviewed below.
during the course of illness), concern over mistakes Schweiger et al. [7] performed a combined case-
(CM), and food-related obsessions (OBF). These dis- control and family-based association analysis of
tributions were consistent with a mixture of popula- DSM-IV AN (excluding binge eating) utilizing 191
tions, and thus the variables were matched with probands with a diagnosis of DSM-IV AN from Euro-
covariate linkage analysis. The most compelling sig- pean–American multiplex families recruited using a
nals arose from the BN cohort [73]. For the BN sibling-pair strategy for linkage analysis, as described
cohort, significant linkage signals arose on 4q21.1 in Kaye et al. [65]. Untransmitted alleles from family
(BMI), 14q21.1 (CM, OBF), 16p13.3 (CM). Suggest- members comprised the controls for family-based
ive linkages were detected at the following chromo- association analysis. For case-control analyses, con-
somal locations: 1q31.1 (ANX), 3p23 (BMI), 4p15.33 trols were European–American females, screened for
(CM), 4q35.2 (ANX), 5p15.3 (BMI), 8q11.23 (CM, Axis I disorders, and with weight for height of
OBF), 10p11.21 (CM), 10p13.1 (BMI, OBF), 16p13.3 < 120%, as defined by the Metropolitan Life tables
(OBF), and 18p11.31 (OBF). For the AN cohort, the of height and weight, to exclude obese individuals.
results for linkage were more modest. No result was Seven DNA polymorphisms at the DRD2 locus
genome-wide significant, although there were some were selected for genotyping: (1) the -141-C or Indel,
suggestive linkage findings: 4q13.1 (BMI), 6q21 rs1799732; (2) IVS2–2730T4C, rs1800498; (3) 932C4G,
(OBS), 9p21.3 (OBS), 11p11.2 (CM), 15q26.2 exon 7, S311C, rs1801028; (4) 939C4T, exon 7, H313,
(OBF), and 17q25.1 (CM, OBF). While substantial rs6275; (5) 957C4T, exon 7, P319, rs6277; (6) Ex8–408
linkage signals were not seen in both cohorts, more 3’UTR, rs6278; and (7) 10620C4>T 3’ of STP or
modest signals did coincide, defining other areas TaqIA, E713K in ANKK1, rs1800497. The most con-
of suggestive linkage. These linkage findings are intri- sistent evidence for association to AN diagnosis is
guing, but they require confirmation before resources exhibited in haplotype analyses. Haplotypes com-
are invested to identify critical genetic variation in the posed of the promoter (-141 Indel) polymorphism
linkage regions. and the exon 7 SNPs (939Y and/or 957Y) exhibited

266
Chapter 20: Genetics of anorexia and bulimia nervosa

statistically significant association in haplotype case- A fixed effects meta-analysis of the HTR2A and
control analyses, differences in linkage disequilibrium SLC6A4 polymorphisms was performed, including
estimates, and statistically significant transmission studies that sampled cases and controls from European
distortion. Confidence in these associations may be ancestry populations, and using a dominant model
greater because deletion of a cytosine at -141 has been with each polymorphism (Tables 20.1 and 20.2) [108,
reported to reduce transcription by an average of 68% 109]. For the HTR2A meta-analysis, we excluded two
in vitro [75], and because the 957T allele has been family-based association studies, based on 45 sibling
shown to exhibit decreased stability and reduced pairs discordant for AN diagnosis and on 313 families
translation of the DRD2 transcript in vitro [76]. The [83, 85], neither of which provided evidence for asso-
957T allele exhibits statistically significant under- ciation with HTR2A, and a case-control association
transmission in combination with the -141 deletion study that used unscreened control samples from a
allele, suggesting that the 957T allele may contribute DNA bank [87], resulting in a final sample of 796 cases
to the association to AN diagnosis via a protective and 1290 controls from 9 sites (Table 20.1). For the
effect. Functional [11C]raclopride neuroimaging SLC6A4 meta-analysis, the final sample consisted of
studies show that individuals who are recovered 489 DSM-IV AN cases and 563 controls from 5 sites
from AN have findings consistent with elevated (Table 20.1). The current literature does not facilitate
D2 or D3 receptor activity [77]. We note that these meta-analyses of diagnostic subtypes of AN, e.g. for
two observations (association of DRD2 SNP alleles the five studies with genotype data on the SLC6A4
with diagnosis and with in vitro gene expression, and polymorphism, only three provide genotype data by
elevated dopamine receptor activity in individuals diagnostic subtype, reducing the sample sizes for meta-
recovered from AN) are superficially consistent, analysis considerably.
however, the comparison is between the behavior of Our meta-analyses of the literature did not iden-
alleles within a cell line and receptors within the tify a significant association of HTR2A -1438G>A
brain, and any mechanistic link between them is with AN diagnosis (Table 20.2), although a much
speculative. larger sample with a similar odds ratio might be
Two serotonin candidate gene polymorphisms significant. Our meta-analysis of the literature did
have received the most attention with respect to inves- identify a significant association of the SLC6A4
tigation of association to DSM-IV anorexia diagnosis, 5-HTTLPR short allele with AN diagnosis. Individ-
i.e. a promoter SNP at the 2A serotonin receptor uals carrying the short allele at the 5-HTTLPR poly-
(5-HT2A) locus (HTR2A –1438G>A) [78–91], and a morphism were 38% (95% CI 6–81%) more likely
promoter polymorphism at the serotonin transporter to fulfill criteria for a diagnosis of DSM-IV AN. There
(SERT) locus (SLC6A4 5-HTTLPR) [84, 90, 92–95]. was no significant evidence for heterogeneity (Q)
The 5HT2A receptor is of interest in AN and BN or publication bias (Egger regression) for either the
because it has been implicated in the modulation of HTR2A or SLC6A4 polymorphism meta-analysis,
feeding and mood, as well as selective serotonin reup- although there was a suggestion of effect size hetero-
take inhibitors (SSRIs) response [96–100]. Post-synaptic geneity and publication bias for the HTR2A literature
5HT2A receptors are in high densities in the cerebral as both estimates had significance values between
cortex and other regions of rodents and humans [101, 0.05 and 0.10 (Table 20.2).
102]. In fact, a number of positron emission tomog-
raphy (PET) imaging studies have found that reduced
binding potential of the 5HT2A receptor persists in Summary and recent findings
individuals who have recovered from AN and BN [96, The candidate gene literature for AN is modest for a
103, 104]. In addition, reduced binding of the 5-HT2A disorder of such morbidity and mortality, and most
receptor has been found in ill AN in one study [105], candidate genes studied have utilized a limited
but not another [106]. In addition, using PET imaging, number of polymorphisms in samples sizes that
recovered restrictor AN had elevated binding potential would now be considered suboptimal for the identifi-
of a SERT ligand compared to recovered bulimia AN cation of significant associations. There is convergent
[107]. Together these imaging studies support the evidence from neuroimaging and molecular genetics
possibility that eating disorder individuals have alter- for involvement of both dopaminergic and serotoner-
ations of 5-HT2A and SERT. gic neurotransmitter systems in the pathophysiology

267
Table 20.1 HTR2A –1438 G>A and SLC6A4 5–HTTLPR case-control association studies to DSM-IV anorexia nervosa diagnosis.

Author Year PMID Ancestry Case N Case Case Case Case Control N Control Control Control Control
major heteroz minor maf* major heteroz minor maf
HTR2A GG GA AA GG GA AA
Cambell 1998 9259661 British 152 45 68 39 0.48 150 53 67 30 0.42
Collier 1997 9259661 British 81 23 33 25 0.51 226 75 117 34 0.41
Enoch 1998 9635956 American 68 16 35 17 0.51 69 25 38 6 0.36
Fuentes 2004 15167698 Spanish 95 30 48 17 0.43 107 39 49 19 0.41
Hinney 1997 9357428 German 100 41 39 20 0.40 355 116 177 62 0.42
Kipman 2002 12231269 French 145 49 69 27 0.42 98 26 50 22 0.48
Ricca 2004 15245785 Italian 77 16 46 15 0.49 115 44 59 12 0.36
Rybakowski 2006 16397402 Polish 148 16 60 55 0.57 89 14 49 26 0.57
Ziegler 1999 10523809 German 78 39 32 7 0.29 170 89 60 21 0.30
SLC6A4 LL LS SS LL LS SS
Di Bella 2000 10889521 Italian 56 17 22 17 0.50 120 48 57 15 0.36
Fumeron 2001 11244478 French 67 17 31 19 0.51 148 44 76 28 0.45
Hinney 1997 9395256 German 96 29 51 16 0.43 112 43 55 14 0.37
Rybakowski 2006 16397402 Polish 132 48 62 22 0.40 93 39 40 14 0.37
Sundaramurthy 2000 10686552 British 138 40 63 35 0.48 90 34 40 16 0.40
* maf, minor allele frequency.
Chapter 20: Genetics of anorexia and bulimia nervosa

Table 20.2 HTR2A –1438 G>A and SLC6A4 5HTTLPR meta-analysis to DSM-IV anorexia nervosa diagnosis.

Author OR 95%CI p Weight Heterogeneity (Q) p Publication bias p


HTR2A
Cambell 1.30 0.80, 2.11 0.288 14.29%
Collier 1.25 0.72, 2.19 0.428 11.18%
Enoch 1.85 0.88, 3.89 0.107 5.08%
Fuentes 1.24 0.69, 2.23 0.467 9.98%
Hinney 0.70 0.44, 1.10 0.123 21.29%
Kipman 0.71 0.40, 1.25 0.230 14.35%
Ricca 2.36 1.21, 4.60 0.012 5.85%
Rybakowski 1.34 0.62, 2.91 0.457 5.39%
Ziegler 1.10 0.64, 1.88 0.731 12.59%
Meta-analysis 1.14 0.95, 1.38 0.166 0.081 0.068
SLC6A4
Di Bella 1.53 0.78, 3.01 0.218 15.07%
Fumeron 1.24 0.65, 2.39 0.512 17.82%
Hinney 1.44 0.81, 2.57 0.217 20.85%
Rybakowski 1.26 0.73, 2.18 0.398 24.97%
Sundaramurthy 1.49 0.85, 2.61 0.167 21.29%
Meta-analysis 1.38 1.06, 1.81 0.017 0.985 0.788

of or susceptibility to AN. However, the diversity of control subjects were analyzed genome-wide for asso-
candidate gene families that has been examined in the ciation of  598 000 SNPs to diagnosis [112]. Both
AN literature is limited and many additional families analyses utilized extensive quality control, assessment
of genes that function in neuronal and other tissues of case-control matching, the Cochran–Armitage
remain to be examined. Fortunately, the development trend test for association testing, and stringent experi-
of genomic technologies enabling exhaustive analyses ment-wide criteria to define significance thresholds
of both common and rare germline variation in the for individual SNPs (p < 9.7EE-6 and p < 1EE-8 for
genome, and an appreciation for the importance of the two studies, respectively). Neither study identified
larger samples sizes in pursuing association studies association findings related to categorical diagnoses
provide opportunities for discovery. Over a period of that reached a priori defined significance levels,
10 years, the Price Foundation multinational collabor- although there were modest associations at biological
ation has ascertained over 1000 unrelated individuals candidate genes in both studies. Additional analyses
with a DSM-IV diagnosis of AN in three separate of the candidate gene dataset for associations with
studies encompassing multiple ascertainment designs, eating disorder related traits previously studied via
and nearly 700 unrelated individuals screened for linkage analyses [113], and for recovery outcomes,
DSM-IV diagnoses, with normal adult weight (19 < with replication [114], resulted in significant associ-
BMI < 27 kg/m2) and matched with ED participants ations between GABA receptor SNPs and recovery.
on 4 factors [65, 110, 111]. Up to 1762 female partici- Investigators concluded that an intronic GABRG1
pants (up to 1085 cases and 677 controls) from this SNP and an excess of GABA receptor SNPs were
dataset have recently been analyzed for association of highly (p = 5EE-6) and nominally (p < 0.05) signifi-
5151 candidate gene SNPs with diagnosis [111], and cantly associated with recovery, respectively [114].
up to 1033 female AN cases and up to 3773 pediatric Using an independent sample of cases and controls,

269
Chapter 20: Genetics of anorexia and bulimia nervosa

investigators identified a nominally significant associ- Further candidate gene and genome-wide analyses
ation of the intronic SNP with trait anxiety, suggest- point to additional candidate genes but also to the
ing a mechanism for the recovery association [114]. need for larger collaborations and alternative molecu-
In summary, a series of linkage and association stud- lar and statistical genetic methods for gene discovery.
ies identified significant linkage in the chr1p34–36 The Price Foundation Collaboration, with the largest
region and provided evidence for association at the and most comprehensive studies in the field, are now
genes coding for the delta opioid and 1D serotonin utilizing next generation sequencing technologies to
receptors in samples ascertained by this collaboration, enable discovery of novel candidate genes and char-
which has been replicated in an independent sample. acterization of existing biological candidates.

References 14. Hoek HW, et al. Am J Psychiatry


1995;152:1272–1278.
31. Lopez C, et al. Int J Eat Disord
2008;41:143–152.
1. American Psychiatric Association.
Diagnostic and Statistical Manual 15. Treasure J, et al. Psychol Med 32. Anderluh MB. Am J Psychiatry
of Mental Disorders, Fourth 1994;24:3–8. 2003;160:242–247.
Edition, Text Revision (DSM-IV- 16. NICE. Core Interventions in 33. Fairburn CG, et al. Arch Gen
TR). Washington, DC: American the Treatment and Management Psychiatry 1999;56:468–476.
Psychiatric Association; 2000. of Anorexia Nervosa, Bulimia 34. Lilenfeld L, et al. Clin Psychol Rev
2. Garner DM, et al. Am J Psychiatry Nervosa and Related Eating 2006;26:299–320.
1985;142:581–587. Disorders (Clinical Guideline 9).
London: National Collaborating 35. Stice E. Pychopharm Bull
3. Halmi KA, et al. Am Acad Child Centre for Medical Health; 2004. 2002;128:825–848.
Psychiatry 1982;21:369–375.
17. Herzog DB, et al. J Clin Psychiatry 36. Bulik CM, et al. J Nerv Ment Dis
4. Herzog DB, et al. N Engl J Med 1992;53:147–152. 1997;185:704–707.
1985;313:295–303.
18. Keel PK, et al. Arch Gen Psychiatry 37. Bulik CM, et al. Acta Psychiatr
5. Strober M, et al. J Nerv Ment Dis 1999;56:63–69. Scand 1997;96:101–107.
1982;170:345–351.
19. Klein D, et al. Int Rev Psychiatry 38. Deep AL, et al. Int J Eat Disord
6. American Psychiatric Association. 2003;15:205–216. 1995;17:291–297.
Diagnostic and Statistical
20. McKenzie JM, et al. Int J Eat 39. Godart NT, et al. Eur Psychiatry
Manual of Mental Disorders,
Disord 1992;11:235–241. 2000;15:38–45.
Fourth Edition. Washington,
DC: American Psychiatric 21. Sullivan PF. Am J Psychiatry 40. Pollice C, et al. Int J Eat Disord
Association; 1994. 1995;152:1073–1074. 1997;21:367–376.
7. Schweiger U, et al. In Jimerson 22. Shroff H, et al. Int J Eat Disord 41. Steinhausen HC. Am J Psychiatry
DC et al. (eds.). Balliere’s Clinical 2006;39:454–461. 2002;159:1284–1293.
Psychiatry. London: Balliere’s 42. Strober M, et al. Int J Eat Disord
23. Halmi K, et al. Arch Gen
Tindall; 1997. 1997;22:339–360.
Psychiatry 2005;62:776–781.
8. Vitousek K, et al. J Abnorm 43. Wagner A, et al. Int J Eat Disord
24. Godart N, et al. J Affect Disord
Psychol 1994;103:137–147. 2006;39:276–284.
2007;97:37–49.
9. Kendler KS, et al. Am J Psychiatry 44. Casper RC. Psychosom Med
25. Godart NT, et al. Int J Eat Disord
1991;148:1627–1637. 1990;52:156–170.
2002;32:253–270.
10. Lilenfeld LR, et al. Arch Gen 26. Kaye W, et al. Am J Psychiatry 45. Srinivasagam NM, et al. Am
Psychiatry 1998;55:603–610. 2004;161:2215–2221. J Psychiatry 1995;152:1630–1634.
11. Strober M, et al. Am J Psychiatry 27. Cassin S, et al. Clin Psychol Rev 46. Strober M. J Psychosom Res
2000;157:393–401. 2005;25:895–916. 1980;24:353–359.
12. Walters EE, et al. Am J Psychiatry 28. Kingston K, et al. Psychol Med 47. Connan F, et al. Physiol Behav
1995;152:64–71. 1996;26:15–28. 2003;79:13–24.
13. Strober M. In Brownell K, 29. Strupp BJ, et al. Neuropsychobiology 48. Strober M, et al. Int J Eat Disord
et al. (eds.). Eating Disorders 1986;15:89–94. 1990;9:239–253.
and Obesity – A Comprehensive
Handbook. New York: Guilford 30. Tchanturia K, et al. Int J Eat 49. Bulik C, et al. Arch Gen Psychiatry
Press; 1995. Disord 2005;37:S72–S76. 2006;63:305–312.

270
Chapter 20: Genetics of anorexia and bulimia nervosa

50. Bulik C, et al. Biol Psychiatry 72. Devlin B, et al. Genet Epidemiol 94. Matsushita S, et al. Am J Med
1998;44:1210–1218. 2002;22:52–65. Genet B Neuropsych Genet
51. Klump KL, et al. Psychol Med 73. Bacanu S, et al. Am J Med 2004;128B:114–117.
2001;31:737–740. Genet B, Neuropsych Genet 95. Sundaramurthy D, et al. Am J Med
52. Wade TD. Am J Psychiatry 2005;139:61–68. Genet 2004;96:53–55.
2000;157:469–471. 74. Bergen A, et al. Neuropsychopharm 96. Bailer UF, et al. Neuro-
53. Kortegaard LS, et al. Psychol Med 2005;30:1703–1710. psychopharmacology 2004;
2001;31:361–365. 75. Arinami T, et al. Hum Mol Genet 29:1143–1155.
54. Klump KL, et al. J Abnormal 1997;6:577–582. 97. Bonhomme N, et al. J Clin
Psychology 2000;109: 76. Duan J, et al. Mol Psychiatry Psychopharmacol 1998;
239–251. 2003;8:901–910. 18:447–454.
55. Rubinow DR, et al. Biol Psychiatry 77. Frank G, et al. Biol Psychiatry 98. De Vry J, et al. Neurosci Biobehav
1997 8;44:839–850. 2005;58:908–912. Rev 2000;24:341–353.
56. Torpy D, et al. J Clin Endocrinol 78. Ando T, et al. Psychiatr Genet 99. Simansky KJ. Behav Brain Res
Metab 1997;82:982. 2001;11:157–160. 1996;73:37–42.

57. Benes F. Am J Psychiatry 1998; 79. Campbell DA, et al. Lancet 100. Stockmeier CA. Ann NY Acad Sci
155:1489. 1998;351:499. 1997;836:220–232.
58. Bulik C, et al. Am J Med Genet B, 80. Collier DA, et al. Lancet 1997; 101. Burnet PW, et al. Neurochem Int
Neuropsych Genet 2005;139:81–87. 350:412. 1997;30:565–574.
59. Rutherford J, et al. Psychol Med 81. Enoch MA, et al. Lancet 1998; 102. Saudou F, et al. Neurochem Int
1993;23:425–436. 351:1785–1786. 1994;25:503–532.
60. Wade T, et al. Psychol Med 1998; 82. Fuentes J, et al. Psychiatr Genet 103. Frank GK, et al. Biol Psychiatry
28:761–771. 2004;14:107–109. 2002;52:896–906.
61. Bulik C, et al. Int J Eat Disord 2007 83. Gorwood P, et al. Mol Psychiatry 104. Kaye WH, et al. Am J Psychiatry
[epub ahead of print]. 2002;7:90–94. 2001;158:1152–1155.
62. Bulik CM, et al. Int J Eat Disord 84. Hinney A, et al. Life Sci 1997;61: 105. Audenaert K, et al. J Nucl Med
2000;28:139–147. PL 295–303. 2003;44:163–169.
63. Holliday J, et al. Psychol Med 85. Karwautz A, et al. Psychol Med 106. Bailer UF, et al. Biol Psychiatry
2006;36:529–538. 2001;32:317–329. 2007;61:1090–1099.
64. O’Dwyer AM, et al. Psychol Med 86. Kipman A, et al. Eur Psychiatry 107. Bailer UF, et al. Psychopharmacology
1996;26:353–359. 2002;17:227–229. 2007;195:315–324.
65. Kaye WH, et al. Biol Psychiatry 87. Nacmias B, et al. Neurosci Lett 108. Bax L, et al. BMC Med Res
2000;47:794–803. 1999;277:134–136. Methodol 2006;6:50.
66. Grice DE, et al. Am J Hum Genet 88. Nishiguchi N, et al. Biol Psychiatry 109. Bax L, et al. BMC Med Res
2002;70:787–792. 2001;50:123–128. Methodol 2007;7:40.
67. Bergen AW, et al. Mol Psychiatry 89. Ricca V, et al. Neurosci Lett 110. Kaye WH, et al. Int J Eat Disord
2003;8:397–406. 2002;323:105–108. 2004;35:556–570.
68. Bergen AW, et al. Current Drug 90. Rybakowski F, et al. Neuro- 111. Pinheiro A, et al. Am J Med Genet
Targets: CNS Neurolog Disord psychopharm 2006;53: B Neuropsychiatr Genet
2003;2:41–52. 33–39. 2010;153B:1070–80.
69. Lander E, et al. Nat Genet 91. Ziegler A, et al. Lancet 1999; 112. Wang K, et al. Mol Psychiatry
1995;11:241–247. 353:929. 2011;16:945–959.
70. Brown K, et al. Biol Psychiatry 92. Di Bella DD, et al. Mol Psychiatry 113. Root TL, et al. Eur Eat Disord Rev
2007;61:367–373. 2000;5:233–234. 2011;19:487–493.
71. Devlin B, et al. Hum Mol Genet 93. Fumeron F, et al. Mol Psychiatry 114. Bloss CS, et al. Neuropsychopharm
2002;11:689–696. 2001;6:9–10. 2011;36:2222–2232.

271
Genetics and common human obesity
Chapter

21 R. Arlen Price

The prevalence of obesity has increased dramatically case. Many studies have estimated heritability of BMI
over the past several decades. Since 1976 the preva- and related variables [8] and they are consistent in
lence of obesity has more than tripled and by 2008 finding moderate to high heritability. Furthermore,
exceeded one-third of the US population (33.8% with the estimates do not depend on the period of the
body mass index [BMI] > 30), while another third study. For example, 2 studies of twins conducted
was overweight (34.2% with BMI ¼ 25.0–29.9) [1]. almost 20 years apart found virtually identical esti-
Highest obesity rates are found in the United States mates of overall heritability of BMI of about 0.80
and United Kingdom among developed countries, [9, 10]. So, while estimates from particular studies vary,
and in the Middle East and Pacific Islands in there is no trend toward decreasing (or increasing)
the developing world (WHO: https://apps.who.int/ heritability.
infobase/). The increase in obesity rates in developing
countries has coincided with a transition to “West- Gene–environment interaction
ern” diets and lifestyles [2–4]. Life in the developed
So what role do genes play in the presence of
and developing world has become more sedentary
such large environmental effects? Bouchard and col-
just as food has become more widely available. Many
leagues completed a series of landmark studies that
lifestyle factors have been suggested to contribute to
helped to explain the role of inherited variation
the dramatic obesity increase, but the primary cause
in mediating environmentally influenced change.
is an obvious one . . . excess caloric intake [5, 6]. The
Bouchard’s research group studied monozygotic
type of diet does contribute to national differences,
twins exposed to long-term positive or negative
but change within countries appears to have been
energy balance. There were considerable individual
driven primarily by overall food availability [7]. As
differences in weight gain or loss under the different
relatively inexpensive, high caloric foods have become
conditions, but changes were similar in the genetically
readily available in much of the world, we are eating
identical co-twins, both in overall weight and visceral
more, the food is higher in caloric content, and we are
fat. The common genotype of the twins influenced
gaining weight as a consequence.
their similarity in response to the environment [11],
demonstrating that adaptation to environmental
Heritability conditions is a heritable trait.
Obesity has moderate to high heritability, but genetic The major environmental changes that are
variation alone cannot account for the dramatic credited with causing the obesity pandemic have
increase. First of all, gene mutations are rare and allele occurred at a population level, but, as with the study
frequencies do not change over short time periods in of twins, individuals differ in their response. While
large populations. The recent prevalence increases as much as two-thirds of populations of developed
must have an environmental origin. An environmen- countries are overweight or obese, the remaining
tal effect of so great a magnitude raises the question of third, living in the same environment, are of normal
whether the heritability of obesity has declined during weight or thin. At the least, this implies a behavioral
the same period, but this does not appear to be the interaction, and, given the heritability of obesity and

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

272
Chapter 21: Genetics and common human obesity

coordinate changes in twins, gene–environment inter-


action must play a major mediating role. A few stud-
Common obesity/rare gene variation
Overall, candidate gene studies have been unsuccess-
ies have tried to identify environmental interactions
ful in explaining common forms of obesity. Genes
with specific genes, focusing of weight gain or loss
central to energy balance tend to have low variability,
as phenotypes, and diet or exercise as components
presumably because of strong selection pressure.
of the environment. One review identified some 13
Even so, some have argued that mutations in a large
studies that reported associations with some measure
number of genes may account for most human
of exercise and 15 with diet and/or exercise [12].
obesity and other common diseases. This view is
However, most associations have not been replicated.
sometimes called the common trait rare gene hypoth-
The interaction most consistently supported was
esis (CTRV), [15, 16], as opposed the common trait
with the Trp64Arg polymorphism in the adrenergic
common variant (CTCV) hypothesis.
receptor beta 3 (ADRB3) gene [12]. Limited power
Substantial progress in finding rare variants has
due to small sample size may in part contribute to
come with a focus on copy number variation (CNV,
the inconsistency of results. However, in the end most
a variant in a DNA segment of 1 or more kilobase
reported associations will be false positives while
in length). While major deletions, duplications and
a few failures to replicate could be false negatives.
rearrangements of DNA sequence associated with
The pairing of new technologies with larger sample
rare diseases have been know for some time, the
sizes could prove more robust for examining gene–
scale of CNV was not appreciated until the last few
environment interactions, but this possible outcome
years. One whole genome survey found more than
will depend on the nature and magnitude of the
4000 variants, affecting more than 600 Mb of geno-
individual interactions.
mic DNA sequence [17]. Large-scale screening has
identified associations of CNVs with a number of
Candidate genes phenotypes [18] including type 1 diabetes, neuropsy-
The phenotypic response in susceptible individuals chiatric conditions [19], and several other common
must be influenced by variability in genes that influ- disorders [20].
ence energy balance. Energy homeostasis requires Wang et al. [21] completed a genome-wide CNV
the coordination of appetite and satiety with energy survey focused on extreme phenotypes. Larger CNVs
expenditure and storage. A great deal has been were over-represented in obese cases compared with
learned about how energy homeostasis is maintained never-overweight controls. CNVs larger than 2 Mb
[13]. It is a complex process involving genes that were present in 1.3% of the cases but absent in con-
regulate appetite, energy metabolism, and fat depos- trols. The most pronounced effects were associated
ition. Many genes that lie in associated regulatory with rare deletions that disrupt genes (odds ratio
pathways have become candidates for weight gain [OR] ¼ 2.7 for CNVs > 1 Mb). Several CNVs were
and obesity. These have included Leptin, Leptin Recep- found to disrupt known candidate genes for obesity,
tor, MC4R, UCPs, PPARG, NPY, and Ghrelin as well such as a 3.3 Mb deletion disrupting NAP1L5 and
as genes in signaling pathways. a 2.1 Mb duplication disrupting UCP1 and IL15.
Candidate gene studies have identified mutations Other studies have identified CNVs that are asso-
in humans or introduced them in animal models ciated with genes not previously associated with
[14]. The last comprehensive count of human studies obesity. An association between BMI and a chromo-
identified associations with 127 genes, most with at some 10q11 CNV was reported in a Chinese cohort
best mixed records of replication. The positive side of [22]. Two genes in this region are GPRIN2 and
a candidate gene approach is that the genes derive PPYR1. In other studies, a deletion on 16p11.2 was
from an emerging understanding of biology. Any reported to be associated with obesity [23, 24]. More
associations that are detected with common obesity recent studies have identified larger numbers of
fit into a pre-existing framework. Candidate gene CNVs [25, 26], and, although individually rare, a
studies have had their successes. Major gene muta- few have been found in multiple individuals, some
tions have been associated with obesity. However, in independent samples. The functional significance
they tend to be rare and account for a few cases of of most of the genes affected by CNVs is unknown.
extreme obesity [14]. As such they become new candidates for obesity.

273
Chapter 21: Genetics and common human obesity

Linkage studies large sample sizes have been developed through the
cooperation of investigators at multiple sites.
One source of motivation for proposing the CTRV
WGA studies have become widely available only
hypothesis was that attempts, through linkage and
within the past four years. The breakthrough for
association, to identify common genes had been
WGA studies came from the Wellcome Trust Case
unsuccessful, at least until recently. The search for
Control Consortium (WTCCC) study that included
common genes has generally taken a genomic
490 000 SNPs and a total of almost 39 000 individuals,
approach in which the entire genome is screened
although the initial phase utilized a much smaller
without prior hypotheses. Linkage studies were the
number of cases and controls, about 2000 and 3000,
first to take a whole genome approach. There have
respectively [31]. The study was unprecedented in size
been more than 60 of them for obesity-related traits
and in the strength of the association with FTO. An
[14], for example, but the results have been disap-
association with MC4R has also been reported based
pointing. A meta-analysis of 37 of these studies found
on the WTCCC sample [32]. The association with
only 2 regions to be significantly supported at the 1%
FTO has been replicated in most studies that followed.
level [27]: chromosome 13q for BMI and chromo-
MC4R has been replicated as well, although not
some 12q for obesity (BMI  30). The outcome of
as consistently. Recent large-scale meta-analysis of
the meta-analysis helps explain why most comprehen-
multiple GWA studies identified additional genes
sive searches for gene associations under linkage
harboring common SNPs that associate with BMI
peaks have been unsuccessful.
[33–36]. GWA studies have also found associations
with measures of body fat distribution [33, 37, 38].
By far the largest GWA study to date included almost
Whole genome association studies 250 000 individuals and 2.8 million SNPs [39]. Asso-
Whole genome association (WGA) studies made it ciations of BMI with 28 loci reached genome-wide
possible to address the two most serious deficiencies significance. Ten had been reported previously and
of previous approaches in that new genotyping eighteen were newly identified. Four additional loci
technology has been combined with very large sample were associated with body fat distribution, all of
sizes. Moreover, WGA studies have several advan- which had been identified previously. However, even
tages over whole genome linkage scans. The reso- this major expansion of sample size has not explained
lution is 2–3 orders of magnitude greater, 2–5 Mb much variation, 1.39% for BMI and 0.16% for body
in linkage studies compared with 10–100 kb with fat distribution.
association. Cases and controls are much easier to Large sample sizes have helped identify associ-
collect than families, and the sample sizes required ations and improved replication, however, the effect
while large are much smaller than those required for sizes have not grown larger and the increase in
linkage [28] and well within reach for collaborative variance accounted for has been minimal. These
groups, if not individual investigators. The advantage findings on obesity are consistent with those for
of a WGA approach was recognized some time ago stature, a complex trait with an even higher herit-
[29] but the available technology was insufficient at ability of at least 0.80. A large GWA study of stature
that time. involving some 63 000 subjects found 54 associated
The recent spate of GWA studies have depended genes that accounted for only about 5% of the total
upon advances in marker identification and genomic variation in height [40–42]. This finding led to much
technology for high throughput genotyping. The discussion and speculation as to what happened to
International HapMap Project (www.hapmap.org/) the so called “missing heritability” [43]. Suggestions
identified more than 4 million single nucleotide poly- have included gene–environment interaction, as well
morphisms (SNPs) and 550 000 of them provide as epigenetics. As discussed in an earlier section,
about 95% coverage of the genome in most popula- gene–environment interaction can play an important
tions, with about double that number needed for role in the development of obesity, although it
Africans [30]. High throughput technology makes it should be born in mind that this may only compli-
possible to type up to 1 million genotypes in a single cate things further, as environmental response is
pass (Affymetrix and Illumina). Greatly reduced costs itself heritable. Epigenetics will be discussed later
have made the technology widely accessible. Finally, in this chapter.

274
Chapter 21: Genetics and common human obesity

Disparate approaches appear to of imprinting with body size arose due to differential
parental investment in offspring in polyandrous
converge, at least in one aspect animals. Males are invested in larger body size of their
WGA studies results have demonstrated that there are offspring while females have an equal investment in
indeed common variants in genes that increase risk all offspring regardless of the father. The theory is
for obesity. This is particularly true for FTO that has supported by fetus size in deer mice (peromyscus)
been widely replicated. However, the proportion of hybrids of monogamous and polyandrous species [49].
variance in BMI these common genes account for is The best know example relating to obesity is
quite small. Major gene mutations such as those in the Prader–Willi and Angelman syndromes, which
leptin, leptin receptor, and POMC have dramatic are due to imprinting of the paternal or maternal
effects on individuals but are so rare that they account chromosome, respectively, of region 15q11–13.
for essentially no common variance. CNVs are much Another imprinted gene is insulin-like growth factor 2,
more common that major gene mutations, but they and paternal expression is strongly related to
are still relatively rare and account for little variance several measures of fat deposition in pigs [50]. In
overall. While there are marked differences in fre- addition, quantitative trait loci (QTL), inferred genes
quency, each approach has been successful. However, based on linkage, have been identified in mice.
the identified variants have done very little to reduce Imprinting is suggested because linkage depends on
the size of the “missing heritability”. parent of origin. In one study, five QTL were found,
Taken on face value, the results from the different two paternal, two maternal, and one with no parent
approaches suggest polygenic inheritance. The classic of origin effect [51].
polygenic model was devised by R. A. Fisher as a way Parent of origin effects have also been identified
of incorporating Mendelian inheritance into quanti- in humans. A large survey reported parent of origin
tative variation [44]. For convenience he assumed dependent associations of variants in known imprinted
there were multiple causal genes, each with small regions on chromosomes 7q32 and 11p15 with several
and roughly equal effects. The particulars, however, complex disorders, including type 2 diabetes [52].
give a somewhat different picture. It turns out there In our own work, we have found parent of origin
are indeed multiple causal genes, and each variant effects on linkage in chromosome regions 10p12 and
accounts for little overall variance. However, the vari- 12q24, where the linkage signal is due entirely to
ants have a wide range of effects on the individuals maternal transmission [53]. Chromosome 12q24 was
that carry them. There is as yet little evidence that the one of the best supported linkage results in a meta-
effects sum to create the phenotype, although one analysis [27], which seems to indicate that the linkage
study found that individuals who had all 32 risk alleles signal is detectible even if parent of origin is not
associated with obesity did have a somewhat higher modeled in the analyses. The chromosome 10p12
BMI [39]. region (19.4–33.3 Mb) is homologous to a largely over-
lapping segment of mouse chromosome 2A3
(15–23 Mb) that has been predicted to be imprinted
Epigenetic modification based on a machine learning model [54]. Two genes in
There has been much discussion of late about the this region have previously been associated
possible effects of epigenetic changes on risk for with obesity, glutamate decarboxylase 2 (GAD2) and
common disorders [45]. Epigenetic modification refers G protein receptor 158 (GPR158) [55]. The concord-
to changes in gene expression that are heritable, that ance is intriguing, although imprinting mechanisms
is, which are maintained during somatic cell division remain to be identified through molecular studies.
and may in some cases be passed on to offspring. A further suggestion of imprinting effects in humans
Genomic imprinting is the most studied form of is our recent finding of a CNV deletion of a region of
epigenetic modification, and involves the differential chromosome 4 including the NAP1L5 gene [21]. The
marking of parental chromosomes during gameto- gene is normally expressed only on the paternal
genesis. Imprinting appears to occur in all marsupial chromosome, which is deleted, apparently leading to
and placental species, and many of the imprinted an absence of gene expression.
genes are related to body size and/or metabolism Environmentally induced epigenetic modification
[46–48]. The conflict theory suggests the association has been recognized in cancer for some time, but

275
Chapter 21: Genetics and common human obesity

a role in complex disorders such as obesity has only


recently begun to be examined at a genomic level.
Applications: prevention and therapy
One goal of genetic research, whether stated or impli-
However, indirect evidence demonstrating environ-
cit, is that findings will eventually make it possible
mental effects on risk for obesity has been known
to use genotype to make decisions about appropriate
for some time. For example, an early study found
approaches to prevention and therapy. The nature of
increased rates of obesity in men who had been
the genetics of human obesity complicates its applica-
in utero or neonatal during the height of the Dutch
tion, particularly in identifying individuals most at
famine of 1944–45 [56]. Other studies of this type also
risk. Odds ratios for most variants will be even
have found that maternal malnutrition contributes to
smaller than for FTO (about 1.65) and have been
risk for obesity and other aspects of the metabolic
estimated to be 1.2 or lower. Prediction will therefore
syndrome [57]. Another study [58] found that pre-
involve only small increments in risk. In most cases,
natal exposure to maternal diabetes increased the risk
familial obesity will continue to be the best predictor
for obesity in Pima Indians. Animal studies similarly
of risk. This difficulty will not only limit application
have shown that maternal exposure to malnutrition,
but can also raise ethical concerns in providing risk
high fat diets, stress and other factors increase risk
assessments to individuals who may never develop
for obesity and the metabolic syndrome. It is of
obesity or become overweight for different reasons.
some interest that both under- and over-nutrition
While overall heritability is substantial, the contribu-
during fetal development can increase risk [59].
tion of individual genes or genotypes is likely to be
It came as something of a surprise that epigenetic
very small relative to the major environmental influ-
effects induced by prenatal diet can be inherited not
ences of diet and lifestyle.
only in a cell lineage but also across generations, for
The identification of protective genes may have
example, by persistent epigenetic marks in sperm [60]
the earliest application in the form of more individu-
that are associated with body weight and resistance to
alized pharmacological treatment, for example, iden-
dietary obesity.
tifying individuals with resistance to drug-induced
More recent studies of humans have also focused
weight gain. To do so, it is not necessary to identify
on epigenetic changes associated with prenatal expos-
genes involved in etiology, only those genes that dir-
ure. A follow-up study of the Dutch famine cohort,
ectly influence drug effectiveness or side effects.
for example, found that exposure indeed led to
Research in others areas have already made it possible
decreased methylation of the imprinted IGF2 gene
to tailor medication to individual genotype, particu-
[61]. Gene expression differences in monozygotic
larly for cancers. Response to tamoxifen treatment
twins discordant for obesity also suggest the possibil-
for breast cancer, for example, appears to be ineffect-
ity of epigenetic modification [62]. While overall
ive in 5–8% of women with a variant of the CYP2D6
differences in expression could be state dependent,
gene [65]. With regard to obesity, several genes have
mitochondrial DNA copy number differences in adi-
been identified that may influence drug induced
pose tissue of discordant twins are consistent with
weight gain, for example, due to olanzapine, including
epigenetic effects.
PMCH, 5-HT2A, ADRA2A, and PKHD1 [66]. In addi-
The obesity state affects expression of many
tion, SLC6A2 and GRIN1 have been associated with
genes, with perhaps as many as 17 000 transcripts
weight loss in response to norepinephrine/dopamine
related to BMI in adipose tissue according to one
transporter inhibitors [67]. Further research will
estimate [63]. Gene expression in normal weight
be needed before genomic screening is practical on a
animals has also been related to later obesity. Inbred
large scale but applications may be generally available
C57BL/6J mice are susceptible to diet induced
in the not too distant future.
obesity, but there is variation in adiposity from an
early age and the differences are maintained under
both high-fat and restricted low-fat diets [64]. Micro- What lies ahead
array analysis found parallel pre-obesity differences in A focus on clarifying the large discrepancy between
expression of genes in several known metabolic path- apparent heritability and variance accounted for by
ways. The causes of the expression differences are individual gene variants, the so called “missing herit-
unclear but could be due to prenatal or early postnatal ability”, is likely to remain a preoccupation in the
environment. near future. As one might expect, there are differences

276
Chapter 21: Genetics and common human obesity

Table 21.1 Candidate genes for body mass index (BMI)/obesity


and fat distribution (bold) from genome-wide association TFAP2B 6 50,911,009
(GWA) studies. LRRN6C 9 28,404,339
Gene Chromosome Location RPL27A 11 8,561,169
NEGR1 1 72,585,028 BDNF 11 27,682,562
TNNI3K 1 74,764,232 MTCH2 11 47,607,569
1 96,717,385 FAIM2 12 48,533,735
1 176,156,103 MTIF3 13 26,918,180
TMEM18 2 612,827 PRKD1 14 29,584,863
RBJ NRXN3 14 79,006,717
FANCL MAP2K5 15 65,873,892
LRP1B GPRC5B 16 19,841,101
CADM2 SH2B1 16 28,793,160
ETV5 FTO 16 52,361,075
GNPDA2 4 44,877,284 MC4R 18 55,990,749
SLC39A8 4 103,407,732 KCTD15 19 39,001,372
FLJ35779 5 75,050,998 QPCTL 19 50,894,012
ZNF608 5 124,360,002 TMEM160 19 52,260,843
NUDT3 6 34,410,847 Summarized from [39].

in opinion about causes as well as the best concordant and discordant relatives, extreme cases
approaches to identify causal mechanisms [68]. and controls [69, 70]. These methods are less costly
Genomic approaches will surely detect other vari- and in some cases may be applied with existing
ants, both rare and common, that have small, incre- GWA studies data. Environmental influences may
mental influence on risk. Both targeted and whole be better understood by the identification of inter-
genome exon sequencing will identify rare coding actions with specific, measured genotypes. New
and splice variants. Whole genome sequencing shows genes could provide additional targets for pharmaco-
particular promise because of its comprehensive cov- logical intervention. Genotypes at these loci may
erage of noncoding regions. Quantitative methods be used in therapeutic interventions through know-
are constantly under development as well, for ledge of their influence on drug effectiveness or side
example, haplotype and shared segment analyses of effects (Table 21.1).

References 5. Bleich S, et al. Annu Rev Public


Health 2008;29:273–295.
9. Malis C, et al. Obes Res
2005;13:2139–2145.
1. Flegal KM, et al. JAMA
2010;303:235–241. 6. Kumanyika SK. Annu Rev 10. Stunkard AJ, et al. JAMA
Public Health 2008; 1986;256:51–54.
2. Hodge AM, et al. Int J Obes
29:297–302. 11. Bouchard C, et al. J Nutr
Relat Metab Disord 1994;
18:419–428. 7. Silventoinen K, et al. Int J Obes 1997;127:S943–S947.
Relat Metab Disord 2004; 12. Lu Qi YAC. Nutrition Reviews
3. Price RA, et al. Am J Phys
28:710–718. 2008;66:684–694.
Anthropol 1993;92:473–479.
4. Sugarman JR, et al. Am J Clin Nutr 8. Maes HH, et al. Behav Genet 13. Farooqi S, et al. Endocr Rev 2006;
1990;52:960–966. 1997;27:325–351. 27:710–718.

277
Chapter 21: Genetics and common human obesity

14. Rankinen T, et al. Obesity 2006; 33. Lindgren CM, et al. PLoS Genet 52. Kong A, et al. Nature
14:529–644. 2009;5:e1000508. 2009;462:868–874.
15. Bodmer W, et al. Nat Genet 2008; 34. Scherag A, et al. PLoS Genet 53. Dong C, et al. Am J Hum Genet
40:695–701. 2010;6:e1000916. 2005;76:427–437.
16. Iyengar SK, et al. Methods Mol 35. Thorleifsson G, et al. Nat Genet 54. Luedi PP, et al. Genome Res
Biol 2007;376:71–84. 2009;41:18–24. 2005;15:875–884.
17. Cooper GM, et al. Nat Genet 2007; 36. Willer CJ, et al. Nat Genet 55. Bell CG, et al. Nat Rev Genet
39:S22–S29. 2009;41:25–34. 2005;6:221–234.
18. Wong KK, et al. Am J Hum Genet 37. Heard-Costa NL, et al. 56. Ravelli GP, et al. N Engl J Med
2007;80:91–104. PLoS Genet 2009;5: 1976;295:349–353.
e1000539. 57. Hales CN, et al. Br Med Bull
19. Cook Jr EH, et al. Nature
2008;455:919–923. 38. Heid IM, et al. Nat Genet 2001;60:5–20.
2010;42:949–960. 58. Pettitt DJ, et al. N Engl J Med
20. Estivill X, et al. PLoS Genetics
2007;3:e190. 39. Speliotes EK, et al. Nat Genet 1983;308:242–245.
2010;42:937–948. 59. Tamashiro KL, et al. Diabetes
21. Wang K, et al. Diabetes 2010;
59(10):2690–2694. 40. Gudbjartsson DF, et al. Nat 2009;58:1116–1125.
Genet 2008;40:609–615. 60. Yazbek SN, et al. Hum Mol Genet
22. Sha BY, et al. J Hum Genet
2009;54:199–202. 41. Lettre G, et al. Nat Genet 2010;19:4134–4144.
2008;40:584–591.
23. Bochukova EG, et al. Nature 61. Heijmans BT, et al. Proc Natl
2010;463:666–670. 42. Weedon MN, et al. Nat Genet Acad Sci U S A 2008;105:
2008;40:575–583. 17046–17049.
24. Walters RG, et al. Nature
2010;463:671–675. 43. Visscher PM. Nat Genet 62. Pietilainen KH, et al. PLoS Med
2008;40:489–490. 2008;5:e51.
25. Glessner JT, et al. Am J Hum
Genet 2010;87:661–666. 44. Fisher RA. Transactions of the 63. Chen Y, et al. Nature 2008;
26. Jarick I, et al. Human
Royal Society of Edinburg 452:429–435.
1918;52:399–433.
Molecular Genetics 2011; 64. Koza RA, et al. PLoS Genet 2006;
20:840–852. 45. Feinberg AP, et al. Nature 2:e81.
27. Saunders CL, et al. Obesity (Silver 2007;447:433–440.
65. Weinshilboum R. Med Biol
Spring) 2007;15:2263–2275. 46. Reik W, et al. Nat Rev Genet 2008;630:220–231.
28. Sham PC, et al. Am J Human 2001;2:21–32.
66. Muller DJ, et al. Pharma-
Genetics 2000;66: 47. Wilkinson LS, et al. Nat Rev cogenomics 2006;
1616–1630. Neurosci 2007;8: 7:863–887.
29. Risch N, et al. Science 1996; 832–843.
67. Spraggs CF, et al. Pharma-
273:1516–1517. 48. Wood AJ, et al. PLoS Genet cogenet Genomics 2005;
30. International HapMap 2006;2:e147. 15:883–889.
Consortium. Nature 2007; 49. Vrana PB, et al. Nat Genet 68. Eichler EE, et al. Nat Rev Genet
449:851–861. 1998;20:362–365. 2010;11:446–450.
31. The Wellcome Trust Case 50. Nezer C, et al. Nat Genet 69. Thomas A, et al. Annals of
Control Consortium. Nature 1999;21:155–156. Human Genetics 2008;72:
2007;447:661–678. 51. de Koning DJ, et al. Proc 279–287.
32. Loos RJ, et al. Nat Genet Natl Acad Sci U S A 2000; 70. Zhu X, et al. Genetic Epidemiology
2008;40:768–775. 97:7947–7950. 2010;34:171–187.

278
Alcoholism
Chapter

22 Howard J. Edenberg

Background Table 22.1 DSM-IV diagnostic criteria for alcohol dependence


requires meeting 3 or more of the following criteria during
Alcohol use disorders (AUDs) are very common a 12-month period.
around the world, and cause a tremendous burden 1. Tolerance
of disability and death [1]. AUDs can be defined as
2. Withdrawal signs or symptoms
“maladaptive patterns of [alcohol] use leading to clin-
ically significant impairment or stress” [2]. Current 3. Drinking more than intended
diagnostic criteria, both from the Diagnostic and 4. Unsuccessful attempts to cut down on use
Statistical Manual on Mental Disorders (DSM-IV)
5. Excessive time related to alcohol (obtaining it,
[3], used in the United States, and from the Inter-
hangover)
national Classification of Diseases (ICD-10) [4], used
elsewhere, divide AUDs into alcohol dependence and 6. Impaired social or work activities due to alcohol
abuse/harmful use. Criteria for alcohol dependence 7. Use despite physical or psychological consequences
(alcoholism) are very similar in the two diagnostic
systems, based on a syndromic definition [5]. The
criteria include withdrawal, tolerance, loss of control
of drinking, impaired work or social activities, and
Alcoholism is a genetic disease with
continued use despite known problems. A diagnosis a necessary environmental component
of alcohol dependence requires that an individual Alcoholism is a complex genetic disease, with convin-
endorse at least three out of seven DSM-IV criteria cing evidence that variations in both genes and envir-
(Table 22.1) or three out of six comparable ICD-10 onment contribute to differences among individuals
criteria. Fewer individuals meet ICD-10 criteria than in risk. Multiple lines of evidence converge to support
meet DSM-IV criteria [6, 7]. In spite of the hetero- the idea of a genetic contribution to the risk. These
geneity in this syndromic definition, the diagnosis of include adoption studies that demonstrate the adoptees
alcohol dependence is one of the most reliable in the more closely resemble their biological parents than
DSM-IV [6, 8, 9]. The reliability of an abuse diagnosis their adoptive parents in risk [12–15]. They also
is much lower [6, 9, 10]. include twin studies, showing greater concordance in
This article will focus on alcohol dependence alcoholism between monozygotic twins (MZ) (who
(alcoholism), because the diagnosis is more reliable share all of their genes) than between dizigotic twins
and because most of the genetic data focus on (DZ) (who share only half of their genes) [14, 16–18].
dependence. US data from the 2001–2002 National The fact that MZ co-twins of alcoholics are not all
Epidemiological Survey on Alcohol and Related Con- alcoholics is a clear demonstration that the disease is
ditions (NESARC) showed a 12-month prevalence of not determined only by genes. A third line of evidence
3.8% for alcohol dependence defined by DSM-IV: is that aspects of alcoholism, including strong prefer-
5.4% in men and 2.3% in women [11]. Lifetime preva- ence for alcohol over water, willingness to work for
lence of alcohol dependence is 12.5% in the United alcohol, sensitivity to the hypnotic or activating effects
States. of alcohol and to withdrawal, and demonstrations that

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

279
Chapter 22: Alcoholism

alcohol is rewarding even in the presence of food and (followed up with association studies of variants
water, can be modeled in selectively bred rodent lines within the linked regions), and genome-wide associ-
[19–23]. A fourth line of evidence comes from early ation studies. This chapter cannot discuss all of the
genetic studies in humans that demonstrated genetic rapidly growing literature in the area; illustrative
variations in alcohol metabolism affect risk for examples are chosen, often from the work of the
alcoholism [24–27] (and see below). Although there Collaborative Study on the Genetics of Alcoholism
might be caveats to any one line of evidence, the (COGA).
convergence of data provides overwhelming evidence
that genetic variations contribute to individual vari-
ations in the risk for alcoholism. The lack of a simple
Candidate gene studies
The earliest genetic association studies in alcoholism
pattern of inheritance indicates that the genetic risk
were candidate gene studies targeting coding vari-
results from the combined contributions of many
ations in the genes that metabolize alcohol. Most
genes, and probably in part from gene  gene and
ingested alcohol is metabolized in the liver, primarily
gene  environment interactions.
in a two step reaction: oxidation to acetaldehyde,
There is a necessary environmental component to
which is further oxidized to acetate. The first step,
the disease: consumption of alcohol. Absent that,
oxidation to acetaldehyde, can be catalyzed by alcohol
underlying genetic vulnerabilities may surface in
dehydrogenases, cytochrome P450s, and catalase.
other problems or diseases, but not in alcoholism.
The great majority of the alcohol is oxidized by the
Unlike other psychoactive drugs, ethanol is ingested
alcohol dehydrogenases (ADH), with the accompany-
in large amounts; legal intoxication in the United
ing reduction of NADþ to NADH (Figure 22.1). The
States is defined as a blood alcohol concentration of
second step, oxidation of acetaldehyde to acetate,
0.08%. Drinking that allows blood alcohol to reach
is catalyzed primarily by aldehyde dehydrogenases
0.08% or above has been defined as binge drinking
(ALDH), with reduction of another molecule of
[28]. Although there is significant variation among
NADþ to NADH. Humans have seven ADHs and
individuals, a typical 170 pound man would generally
two ALDHs that catalyze most of the ethanol metab-
reach this level after drinking about 5 standard drinks
olism. Variations in the genes encoding enzymes of
within 2 hours and an average women would reach
alcohol metabolism have long been known to affect
this level after ingesting about 4 drinks, due to differ-
the risk for alcoholism [24–27, 30].
ences in weight and body composition. (A standard
A variation in ALDH2 has dramatic effects on
drink in the United States contains 14 g of pure
alcohol metabolism. The ALDH2*2 allele encodes a
ethanol, and is approximately equivalent to 12 ounces
nearly inactive subunit of the mitochondrial aldehyde
of beer, 5 ounces of table wine, or 1.5 ounces of 80-proof
dehydrogenase 2 that is responsible for much of the
spirits.) Alcohol consumption must generally be in
oxidation of ethanol. Presence of a single ALDH2*2
the binge range and frequently repeated to develop
allele renders the ALDH2 catalytic activity (measured
the problems that result in diagnosis of alcohol
in vitro) below the usual limit of detection [31]. In
dependence. The risk of meeting criteria for alcohol
individuals with a single copy of the inactive ALDH2*2
dependence rises nearly linearly with the frequency of
allele consumption of even small amounts of alcohol
binge drinking [29]. Data from the National Epidemio-
causes a dramatic rise in acetaldehyde in blood, which
logic Survey on Alcohol and Related Conditions
triggers a highly aversive reaction similar to that caused
(NESARC) indicate that the number of criteria for
abuse and dependence that are met rise with the fre-
when a patient taking disulfiram (Antabuse ) drinks
alcohol; the reaction includes flushing, tachycardia,
®
quency of binge drinking [2, 29].
This necessary environmental factor, alcohol, inter-
acts with the underlying genetics to result in the disease Alcohol Acetaldehyde Acetate
of alcoholism. Variations in the environment related
ADH ALDH
to alcohol accessibility, price, and social norms, there-
NAD+ NADH + H+ NAD+ NADH + H+
fore, can affect the prevalence of alcoholism.
There are three main approaches to identification Figure 22.1 The primary pathway of alcohol metabolism in the
liver is oxidation by alcohol dehydrogenases (ADH) and aldehyde
of variations in specific genes that affect alcohol dehydrogenases (ALDH) enzymes, with acetaldehyde as the
dependence: candidate gene studies, linkage studies intermediate.

280
Chapter 22: Alcoholism

ADH4 ADH1B homozygous for ALDH2*1) [35, 36, 40, 41]. Neither
of these alleles are common among individuals of
ADH5 ADH6 ADH1A ADH1C ADH7 European ancestry, although ADH1B*2 is found at
Figure 22.2 Genomic arrangement of human ADH genes on moderate frequency in people of Jewish ancestry [42,
chromosome 4q. 43]. The low allele frequency makes studies more diffi-
cult, but ADH1B*2 also appears protective in individ-
and nausea. This aversive reaction greatly reduces uals of European descent (with relative risk about 0.5)
their propensity to drink, the amount consumed [38, 40, 44]. A recent large-scale study has conclusively
per occasion, and their risk for alcoholism [24, 26, demonstrated protective effects in Europeans [45].
27, 31–35]. It is not totally protective: some individ- Recent genetic studies have shown that variations in
uals continue to drink despite the high acetaldehyde ADH4 play a significant role in affecting risk for alco-
levels. But for someone with a single ALDH2*2 allele, holism, as do noncoding variations in ADH1A and
the relative risk of alcoholism range from about 0.13 ADH1B [37]. Other variations in the ADH region were
to 0.40 in different Asian populations [35, 36]. In vivo, also shown to affect alcohol metabolism [46], includ-
having two copies of the ALDH2*2 allele leads to ing variations in and near ADH7 [47] ADH1A,
an even more severe reaction and is nearly totally ADH1B, ADH1C, and ADH4 [48]. Clearly there is
protective against alcoholism; among thousands of more to be learned about the contributions of the
cases reported in the literature, there are only three ADH genes to both alcohol metabolism and alcoholism.
alcoholics homozygous for ALDH2*2 [35, 36]. Despite the fact that blood acetaldehyde levels
Variations in ADHs that catalyze the first step in do not rise substantially in individuals with the
ethanol metabolism also strongly affect the risk for ADH1B*2 or ADH1B*3 alleles, it is generally thought
alcoholism. Humans have seven ADHs that arose that both the more rapid generation of acetaldehyde
from repeated gene duplications; the genes encoding by these ADHs and the reduction in the rate of
them are clustered in a small region of chromosome 4 elimination of acetaldehyde by the inactive ALDH
(Figure 22.2). The kinetic properties of these enzymes affect the risk for alcoholism by at least transiently
suggest that at low levels of alcohol, enzymes encoded increasing acetaldehyde levels in the liver, triggering
by ADH1A, ADH1B, and ADH1C play the major role aversive reactions that reduce excessive drinking.
in metabolism; ADH1B is the ADH present at highest A combination of the strongly protective alleles at
levels in the adult liver and presumably contributes both ADH1B and ALDH2 loci is synergistic, and
most. When ethanol is present at higher levels (intoxi- drops relative risk of alcoholism in Asians to about
cating) the enzyme encoded by ADH4 makes an 1–10% [35, 36]. Thus variations in the ADH and
increasing contribution. ADH7, located in the ALDH genes play strong roles in affecting the risk
esophagus and stomach lining, can contribute to “first for alcoholism. These are among the strongest and
pass” metabolism of ethanol because ethanol is at best replicated findings in the genetics of complex
very high concentrations in stomach during drinking. diseases. Despite their strong impact on some popu-
Variants in ADH1B and ADH1C that increase the lations, the coding variations in ADH and ALDH
rates at which the enzymes they encode oxidize etha- genes do not explain a large fraction of the differences
nol reduce the risk for alcoholism [24, 26, 27, 35–40]. in risk in populations of European ancestry in which
The strongest effects are due to variants in ADH1B the alleles with the strongest effects are rare.
that are relatively common in Asians (ADH1B*2, in There have been many other candidate gene
which arginine 48 is replaced with histidine) and in studies, some with markedly mixed results. A full
people of African ancestry (ADH1B*3, in which description of these is beyond the scope of this article,
arginine 370 is replaced with cysteine). Blood acetal- but several will be discussed as illustrative of the
dehyde levels do not dramatically rise in individuals field. Perhaps the most famous is the association of
with these variants, and there is no severe flushing the Taq1A allele of the dopamine D2 receptor gene
reaction comparable to that in individuals with the (DRD2-Taq1A; rs1800497) with alcoholism, first
ALDH2*2 allele. Nevertheless, the protection against reported in a small study in 1990 [49]. There have
alcoholism provided by the ADH1B*2 allele is very been numerous attempts to replicate this, and
strong, with relative risks approximately 0.18–0.26 for although most have been negative, e.g. [50–52], one
1 allele and 0.10–0.14 for 2 (in Asian subjects recent meta-analysis suggests a small but significant

281
Chapter 22: Alcoholism

effect that might be due in part to publication bias [53]. systems. When variations in one gene is found to affect
The “DRD2-Taq1A” polymorphism actually lies the risk for alcoholism, it is reasonable to examine other
within an adjacent gene, ANKK1 (ankyrin repeat and genes in the system(s) to which it belongs; this has been
kinase domain containing 1) [54]. A recent family useful, for example, in following up initial findings with
study showed that although rs1800497 was not signifi- GABRA2 [73] in other GABA-receptor genes [74, 75].
cant, a different allele at DRD2 (rs6277, a synonymous There are, however, advantages to taking an unbiased
single nucleotide polymorphisms [SNPs]) was [55]. approach toward identifying genes that affect risk, such
Another family-based analysis showed that the associ- as genetic linkage or whole genome association studies.
ation in this region with alcohol dependence is
strongest in the region of ANKK1 that is not in linkage
disequilibrium (LD) with DRD2 [50], and another Genetic linkage studies
family and case-control study suggested the evidence Genetic linkage studies provide an unbiased approach
for association with alcoholism in this region is in the toward identifying genetic variations that affect the
NCAM1, TTC12, and ANKK1 genes [56, 57]. There risk for a complex genetic disease such as alcoholism.
remains much complexity in interpreting these data. Linkage studies follow the inheritance of particular
The opioid system has been implicated in addic- regions of the genome within families. The regions
tions, not only for opiates but also for alcohol [58, are tagged by some readily assayable marker; micro-
59]. There are three main opioid receptors and three satellite markers were the most widely used when the
main genes encoding endogenous ligands. OPRM1 studies on alcoholism began, but SNPs are now easier
encodes the mu receptor (MOR), and POMC encodes to assay. Linkage studies test for significantly excess
its primary ligands; OPRK1 encodes the kappa recep- (or reduced) transmission of particular regions of the
tor (KOR) and PDYN its primary ligands; OPRD1 genome to affected family members. Because nearby
encodes the delta receptor (DOR) and PENK its pri- sites along a chromosome are usually transmitted
mary ligands. The mu receptor has been most widely together, a relatively small number of markers can
studied, particularly a coding variation (Asn40Asp); a report on the genome. This is a technical advantage
recent meta-analysis of those studies concluded that for linkage studies, but the drawback is that linkage
there was no significant evidence for association [60]. studies generally identify relatively large regions, usu-
COGA examined all six genes from this system, and ally between 20 and 50 million base pairs, that contain
found that variations in the kappa system were asso- many genes. Because linkage studies look within indi-
ciated with alcoholism [61], but variations in the mu vidual families, they can identify a region in which
or delta system were not [62], nor were variations in different rare alleles affect the risk in different families.
the related nociceptin system [63]. It was particularly The earliest whole genome linkage studies on
intriguing that variations in the genes encoding alcoholism were reported in 1998 [76, 77]. The Col-
both the receptor (OPRK1) and its ligand (PDYN) laborative Study on the Genetics of Alcoholism
were associated. Follow-up molecular studies demon- (COGA) recruited families of individuals in treatment
strated that one associated polymorphism affected the for alcoholism at six sites across the United States, and
level of expression of OPRK1 [64]. did genetic studies on those families in which at least
NPY is a candidate gene based upon studies of a three first-degree relatives met criteria for alcohol
rat model of alcohol preference [65–67], knockout dependence [77]. The initial linkage study encom-
mice [68], and some but not all human studies, e.g. passed 987 individuals from 105 families, and reported
[69–71]. NPY itself is on human chromosome 7 and suggestive evidence for linkage with alcoholism on
was not significantly associated with alcoholism in the chromosomes 1 and 7, weaker evidence on chromo-
COGA sample [72]. Three NPY-receptor genes on some 2, and linkage with a protective phenotype on
chromosome 4 were also evaluated, and two of these chromosome 4q [77]. Long et al. [76] studied 172
gave evidence of association: NPY2R was associated sibling pairs from a southwest Native American popu-
with alcohol dependence and alcohol withdrawal lation, and reported evidence for linkage on chromo-
symptoms, and NPY5R was associated with alcohol some 11p, 4p and for several markers in the ADH
withdrawal characterized by seizures [72]. region of 4q. A follow-up study of an additional 157
There are many other potential candidate genes, families from the COGA project supported their original
including genes in nearly all neuronal receptor/ligand findings on chromosomes 1 and 7 and also revealed

282
Chapter 22: Alcoholism

some evidence for linkage on chromosome 3 [78]. and results from a linkage study in Southwest Native
A smaller study provided support for the linkage Americans showed the strongest linkage there [76].
on chromosome 1 [79]. Analysis combining the diag- Our interest in the region was greatly enhanced by a
nostic phenotype with an electrophysiological vari- very strong linkage of an electrophysiological pheno-
able (amplitude of the P300 component of the type believed to be related to susceptibility to alcohol-
event-related potential) gave the strongest evidence of ism, the amplitude of the b-EEG [85]. A cluster of 4
linkage in a broad region of chromosome 4q that GABAA receptor genes was located in the middle of
encompassed the ADH genes [80]. A quantitative trait, that linkage peak, and was extensively analyzed. Many
maximum drinks in a 24 hour period, also showed variations and haplotypes in GABRA2, encoding the
linkage on chromosome 4q [81]. Results from the Irish a2 subunit of the GABAA receptor, were associated
Affected Sib Pair Study of Alcohol Dependence again with alcohol dependence as well as with the b-EEG
showed linkage to a broad region of chromosome 4q, phenotype [73]. This has since been confirmed in
and weaker, suggestive evidence on chromosomes 1q, many studies of different populations [86–92]. The
13q, and 22q for alcohol dependence, on 2q, 9q, and association in the COGA sample is with the most
18p for symptom count [82], on chromosome 9 for severe half of the alcoholic subjects, as defined by
age at onset, on chromosomes 1 and 11 for initial comorbid drug dependence [93], which also correl-
response to alcohol, 1, 6, and 22 for tolerance, on ates with early onset of alcoholism and many other
chromosomes 12 and 18 for maximum drinks, and measures of severity [94]. Other studies have shown
on chromosome 2 for withdrawal symptoms [83]. association also with the GABRG1 gene, adjacent to
A more recent linkage study of alcohol dependence GABRA2 and in LD with it [96].
in a set of African-American families originally ascer- One of the large linkage peaks in the COGA
tained for cocaine or opioid dependence showed evi- sample was on chromosome 4q [77, 80]; this region
dence for linkage to chromosome 10q [84]. As noted was also implicated in the Irish study [82]. The ADH
above, these linkage studies point to chromosomal genes lie under this peak. As discussed above, these
regions, rather than specific genes. were targeted by many SNPs, and associations found
with SNPs in ADH4, ADH1A, and ADH1B [37].
These findings were supported and extended by other
Follow-up of linkage studies groups (see above). However, we hypothesized that a
A productive approach to identifying genes that affect broad linkage peak was likely to reflect the combined
the risk for alcoholism is to follow-up linkage studies contribution of several genes, and continued our ana-
with association studies of variants within the linked lyses of other genes within this region. COGA has,
regions. One can either analyze variations spanning thus far, identified several additional genes within
the entire linkage region in a systematic manner, based this linkage peak in which variations affect alcohol
upon LD, or target candidate genes within the region. dependence or a closely-related phenotype. SNCA,
The systematic approach generally requires more gen- encoding a-synuclein, was a top candidate gene based
otyping and correction for more multiple testing, but upon studies in a rat model of alcohol preference [65,
has the advantage of being relatively unbiased. 96, 97] and on data from monkeys [98] and humans
A disadvantage is that the location of the peak of [98–101]. In the COGA subjects, variations in SNCA
linkage is not necessarily centered on the key gene(s) were not significantly associated with dependence per
contributing to the linkage, so it is possible to miss the se, but were associated with craving for alcohol [102].
gene contributing to the linkage signal if too narrow a Continuing our analysis of genes in the chromosome
region is analyzed. The candidate gene approach lever- 4 linkage peak, we demonstrated that TACR3, encod-
ages prior data and biological understanding to reduce ing the tachykinin 3 (neurokinin B) receptor, was
the number of genes that are tested, but might miss associated with alcohol dependence, as was NFKB1
potentially important genes for which there is no prior [103], encoding a subunit of the ubiquitous transcrip-
evidence or hypothesized role in alcoholism. tion factor NF-kB, which regulates many genes in the
These issues will be illustrated by discussing some brain. The association of both TACR3 and NFKB1
results from COGA. There was some evidence from were strongest with the most severely affected subjects
the initial COGA analysis for linkage to markers on [103, 104]. Thus the hypothesis that multiple genes
chromosome 4p around a GABA receptor gene [77], contribute to this linkage peak was supported.

283
Chapter 22: Alcoholism

A second major linkage peak in the COGA


sample was on chromosome 7q [77, 105–107].
Gene × environment interaction
A substantial gene  environment interaction was
A systematic screen across an 18 Mb (2-LOD) inter-
shown dramatically in the Japanese population by
val was carried out, selecting SNPs that efficiently
Higuchi [25]. The degree of protection against alco-
report on the variations within this region based
holism afforded by the ALDH2*2 allele changed sig-
upon patterns of LD [108]. Several SNPs gave
nificantly between the years 1979 and 1992: in that
significant or suggestive evidence of association
time, the fraction of Japanese alcoholics carrying the
with alcoholism; the most consistent evidence was
ALDH2*2 allele in heterozygous form increased
for several SNPs in the ACN9 gene [108], related to
from 2.5 to 13% [25]. This time was far too small
gluconeogenesis and the assimilation of ethanol
for any change in the underlying frequency of the
and acetate. Additional genes within this region are
polymorphism in the Japanese population as a whole,
being examined.
so the only explanation is that its protective effect was
reduced, presumably by sociological changes leading
Genome-wide association studies toward more alcohol consumption. In that same
study, no alcoholics were found to be homozygous
Recently, genome-wide association studies (GWAS)
for ALDH2*2 (although 120 would have been
have largely replaced linkage studies [109, 110]. These
expected based on the allele frequency), suggesting
studies take advantage of modern multiplexed
that the protection afforded by having two copies of
methods for simultaneously genotyping 1 million
the ALDH2*2 allele is much stronger and was not as
SNPs or more. They allow association testing of a
susceptible to the environmental changes [25].
very large fraction of the genome, and have been
successful for several complex diseases [109]. There
are limitations, however. The most obvious is the risk Genes and environment in treatment
of false positives due to the extraordinary amount of It is likely that particular genes and combinations
multiple testing; less obvious is that the corrections of genes will differentially affect risk for different
used to avoid false positives are likely to lead to many manifestations of alcoholism, such as co-occurrence
false negatives. The problem of multiple testing leads with other disorders, and that different combinations
to a requirement for very large sample sizes, with of genes will also affect the response to particular
cases and controls very well matched for ethnicity. treatments. There is already evidence that variations
The use of correction factors using ancestry informa- in some genes primarily affect alcoholism that shows
tive markers can accommodate multiple ethnicities in early onset and comorbidity with other drugs of abuse
the same study. Another limitation is that GWAS are (e.g. GABRA2, ADH4, CHRM2, NFKB1) [73, 93, 95,
designed to find relatively common polymorphisms 103, 104]. As we learn more about the genetic under-
that contribute to the risk for a disease; if a collection pinnings of alcoholism, we are likely to find that differ-
of rare variants in a gene can independently increase ent combinations of genetic variants lead to different
risk, this will often be missed by GWAS (but can be responses to different treatments, which will improve
captured by linkage studies). our ability to design individualized therapies [112].
To date, only one GWAS on alcoholism has been
published. Treutlein et al. [111] studied 487 male
alcoholic inpatients from Germany and 1358 controls, Summary
and followed up the more significant findings in There has been much progress in the genetics of
another group of 1024 patients and 996 controls. No alcohol dependence. Variations in several genes have
SNP met criteria for genome-wide significance in the clearly been shown to affect the risk for developing
initial study, but two did in the combined sample. alcohol dependence. Certain variations in ADH and
Other SNPs provided consistent evidence in the initial ALDH genes have very strong effects on the risk for
sample and follow up, at less significant levels. The alcoholism. Variations in other genes appear to have a
need for large samples and replications means that much smaller effect on risk. In populations of Euro-
results expected within the next few years will be pean ancestry, in which the coding variations in ADH
important for assessing which SNPs are truly associ- and ALDH that have the strongest effects are uncom-
ated with alcohol dependence. mon, most of the individual difference in risk is still

284
Chapter 22: Alcoholism

unexplained, and probably reflects the summation of It should be remembered that although variations in
many genes of small effect, along with gene  gene genes clearly affects an individual’s risk for alcoholism,
and gene  environment interactions. Linkage studies the disease is not determined solely by genes. The envir-
and their follow-up, along with candidate gene studies onment and individual choices plays a major role.
and GWAS, are beginning to fill the gaps. Initial Understanding the genetic contributions to risk should
findings must be confirmed in independent studies, lead to better understanding of the disease processes and
and much work remains to elucidate the mechanisms assist in tailoring treatments to individuals.
involved. Nevertheless, with the new technologies and
larger samples being studied, progress should acceler- Acknowledgements
ate. The future will involve studies of epigenetic Related work in the author’s laboratory is supported
factors, copy number variants, and gene expression, by grants from the National Institute of Alcohol
as well as tests for rare variants of large effect in Abuse and Alcoholism, R37AA006460, U10AA008401,
specific families [113]. U01AA016660, R21AA016301, P60AA007611.

References 15. Sigvardsson S, et al. Arch Gen


Psychiatry 1996;53(8):681–687.
30. Stamatoyannopoulos G, et al.
Am J Hum Genet 1975;27:
1. World Health Organization. 789–796.
The World Health Report 2002: 16. Heath AC, et al. Psychol Med
Reducing Risks, Promoting Healthy 1997;27(6):1381–1396. 31. Crabb DW, et al. J Clin Invest
Life. Geneva: WHO; 2002. 17. Kendler KS, et al. Am J Psychiatry 1989;83(1):314–316.
2. Li TK, et al. Addiction 2007; 1994;151:707–715. 32. Bosron WF, et al. Semin Liver Dis
102:1522–1530. 18. Pickens RW, et al. Arch Gen 1981;1:179–188.
3. American Psychiatric Association. Psychiatry 1991;48:19–28. 33. Chen YC, et al. Pharmacogenet
Diagnostic and Statistical Manual of 19. Buck KJ, et al. J Neurosci 1997; Genomics 2009;19(8):588–599.
Mental Disorders, Fourth Edition. 17(10):3946–3955. 34. Harada S, et al. Lancet 1982;2:827.
Washington, DC: American
Psychiatric Association; 1994. 20. Crabbe JC, et al. Trends Neurosci 35. Luczak SE, et al. Psychol Bull
1999;22(4):173–179. 2006;132(4):607–621.
4. World Health Organization.
International Classification of 21. McBride WJ, et al. Crit Rev 36. Chen YC, et al. Alcohol Clin Exp
Disease, Tenth Edition. Geneva: Neurobiol 1998;12(4):339–369. Res 1999;23:1853–1860.
WHO; 1993. 22. Murphy JM, et al. Behav Genet 37. Edenberg HJ, et al. Hum Mol
5. Edwards G, et al. Br Med J 1976; 2002;32(5):363–388. Genet 2006;15(9):1539–1549.
1(6017):1058–1061. 23. Whatley VJ, et al. Alcohol Clin Exp 38. Macgregor S, et al. Hum Mol
6. Grant BF. Alcohol Clin Exp Res Res 1999;23(7):1262–1271. Genet 2009;18(3):580–593.
1996;20(8):1481–1488. 24. Edenberg HJ. Alcohol Res Health 39. Sherva R, et al. Alcohol Clin Exp
7. Schuckit MA, et al. Addiction 2007;30(1):5–13. Res 2009;33(5):848–857.
1994;89:1629–1638. 25. Higuchi S, et al. Lancet 1994; 40. Whitfield JB. Am J Hum Genet
8. Schuckit MA, et al. Alcohol Clin 343:741–742. 2002;71(5):1247–1250; author
Exp Res 2002;26(7):980–987. 26. Hurley TD, et al. Pharma- reply 1250–1251.
9. Hasin D, et al. Addiction 2006; cogenomics: The Search 41. Li D, et al. Biol Psychiatry
101 S1:59–75. for Individualized Therapies. 2011;70:504–512.
New York: Wiley-VCH; 2002.
10. Grant BF, et al. Drug Alcohol 42. Carr LG, et al. Am J Med Genet
Depend 2007;86(2–3):154–66. 27. Thomasson HR, et al. Am J Hum 2002;112(2):138–143.
Genet 1991;48(4):677–681.
11. Grant BF, et al. Drug Alcohol 43. Neumark YD, et al. J Stud Alcohol
Depend 2004;74(3):223–234. 28. National Advisory Council on 1998;59(2):133–139.
Alcohol Abuse and Alcoholism.
12. Bohman M, et al. Arch Gen Council Minutes: National 44. Borras E, et al. Hepatol 2000;
Psychiatry 1981;38(9):965–969. Advisory Council Meeting, 31(4):984–989.
13. Cloninger CR, et al. Arch Gen February 4–5; 2004. 45. Bierut LJ, et al. Mol Psychiatry,
Psychiatry 1981;38:861–868. 29. Dawson DA, et al. Alcohol in press.
14. Heath AC. Alc Health Res World Clin Exp Res 2005;29(5): 46. Birley AJ, et al. Behav Genet
1995;19:166–171. 902–908. 2005;35(5):509–524.

285
Chapter 22: Alcoholism

47. Birley AJ, et al. Hum Mol Genet 70. Mottagui-Tabar S, et al. Alcohol 91. Lappalainen J, et al. Alcohol Clin
2008;17(2):179–189. Clin Exp Res 2005;29(5):702–707. Exp Res 2005;29(4):493–498.
48. Birley AJ, et al. Hum Mol Genet 71. Zhu G, et al. Alcohol Clin Exp Res 92. Soyka M, et al. J Psychiatr Res
2009;18(8):1533–1542. 2003;27(1):19–24. 2008;42(3):184–191.
49. Blum K, et al. JAMA 1990; 72. Wetherill L, et al. Alcohol Clin Exp 93. Agrawal A, et al. Behav Genet
263:2055–2060. Res 2008;32(12):2031–2040. 2006;36(5):640–650.
50. Dick DM, et al. Alcohol Clin Exp 73. Edenberg HJ, et al. Am J Hum 94. Dick DM, et al. Addiction
Res 2007;31(10):1645–1653. Genet 2004;74(4):705–714. 2007;102(7):1131–1139.
51. Edenberg HJ, et al. Alcohol Clin 74. Dick DM, et al. Alcohol Clin Exp 95. Covault J, et al. Neuro-
Exp Res 1998;22(2):505–512. Res 2004;28(1):4–9. psychopharmacology 2008;33
52. Gelernter J, et al. JAMA 1993; (4):837–848.
75. Dick DM, et al. Alcohol Clin Exp
269:1673–1677. Res 2006;30(7):1101–1110. 96. Liang T, et al. J Neurochem
53. Munafo MR, et al. Mol Psychiatry 2006;99(2):470–482.
76. Long JC, et al. Am J Med Genet
2007;12(5):454–461. B Neuropsych Genet 1998;81: 97. Liang T, et al. Proc Natl
54. Neville MJ, et al. Hum Mutat 216–221. Acad Sci U S A 2003;100(8):
2004;23(6):540–545. 4690–4695.
77. Reich T, et al. Am J Med Genet
55. Hill SY, et al. Am J Med Genet 1998;81(3):207–215. 98. Walker SJ, et al. Alcohol 2006;
B Neuropsychiatr Genet 2008; 38(1):1–4.
78. Foroud T, et al. Alcohol Clin Exp
147B(4):517–526. Res 2000;24(7):933–945. 99. Bonsch D, et al. Alcohol Clin Exp
56. Yang BZ, et al. Hum Mol Genet Res 2005;29(5):763–765.
79. Lappalainen J, et al. Mol
2007;16(23):2844–2853. Psychiatry 2004;9(3):312–319. 100. Bonsch D, et al. Biol Psychiatry
57. Yang BZ, et al. Alcohol Clin Exp 2004;56(12):984–986.
80. Williams JT, et al. Am J
Res 2008;32(12):2117–2127. 101. Zhou W, et al. Brain Res 2002;
Hum Genet 1999;65(4):
58. Kreek MJ. Mol Psychiatry 1996; 926(1–2):42–50.
1148–1160.
1(3):232–254. 102. Foroud T, et al. Alcohol Clin Exp
81. Saccone NL, et al. Am J Med Genet
59. Kreek MJ, et al. Neuromolecular Res 2007;31(4):537–545.
2000;96(5):632–637.
Med 2004;5(1):85–108. 103. Edenberg HJ, et al. Hum Mol
82. Prescott CA, et al. Mol Psychiatry
60. Arias A, et al. Drug Alcohol Genet 2008;17(7):963–970.
2006;11(6):603–611.
Depend 2006;83(3):262–268. 104. Foroud T, et al. Alcohol Clin Exp
83. Kuo PH, et al. Alcohol Clin Exp Res 2008;32(6):1023–1030.
61. Xuei X, et al. Mol Psychiatry Res 2006;30(11):1807–1816.
2006;11(11):1016–1024. 105. Dunn G, et al. BMC Genet 2005;
84. Gelernter J, et al. Biol Psychiatry 6 S1:S122.
62. Xuei X, et al. Am J Med Genet 2009;65(2):111–115.
B Neuropsychiatr Genet 2007; 106. Jones KA, et al. Int J Psychophysiol
144(7):877–884. 85. Porjesz B, et al. Proc Natl Acad Sci 2004;53(2):75–90.
U S A 2002;99:3729–3733.
63. Xuei X, et al. Addict Biol 2008; 107. Wang JC, et al. Hum Mol Genet
13(1):80–87. 86. Covault J, et al. Am J Med Genet 2004;13(17):1903–1911.
B Neuropsychiatr Genet 2004;
64. Edenberg HJ, et al. Hum Mol 129(1):104–109. 108. Dick DM, et al. Biol Psychiatry
Genet 2008;17(12):1783–1789. 2008;63(11):1047–1053.
87. Drgon T, et al. Am J Med Genet
65. Carr LG, et al. Alcohol Clin Exp B Neuropsychiatr Genet 2006; 109. Manolio TA, et al. Annu Rev Med
Res 1998;22:884–887. 141(8):854–860. 2009;60:443–456.
66. Hwang BH, et al. Alcohol Clin Exp 88. Enoch MA, et al. Neuro- 110. Manolio TA, et al. Nat Genet
Res 1999;23(6):1023–1030. psychopharmacology 2009;34 2007;39(9):1045–1051.
67. Spence JP, et al. Neuroscience (5):1245–1254. 111. Treutlein J, et al. Arch Gen
2005;131(4):871–876. 89. Enoch MA, et al. Am J Med Psychiatry 2009;66(7):773–784.
68. Thiele TE, et al. Nature 1998; Genet B Neuropsychiatr Genet 112. Edenberg HJ, et al. Pharmacol
396(6709):366–369. 2006;141(6):599–607. Ther 2005;108(1):86–93.
69. Lappalainen J, et al. Arch Gen 90. Fehr C, et al. Psychiatr Genet 113. Edenberg HJ. Biol Psychiatry
Psychiatry 2002;59(9):825–831. 2006;16(1):9–17. 2011;70:498–499.

286
Nicotine dependence
Chapter

23 Sarah M. Hartz and Laura J. Bierut

Introduction Nicotine dependence is not only a public health


risk due to lung cancer and heart disease, but high
Cigarette smoking is one of the leading causes of comorbidity with psychiatric illness makes it of par-
mortality worldwide. Tobacco use is estimated to cause ticular importance in mental health. There are mark-
12% of vascular disease, 66% of respiratory cancers edly increased rates of lifetime smoking in subjects with
(trachea, bronchus, and lung cancers), and 38% of mental illness as compared to controls (59% versus
chronic respiratory disease. Because of this, 8.8% of 39% respectively, p < 0.001) [4]. The prevalence of
deaths (4.9 million annually worldwide) are attributed nicotine dependence is 13% in the general population
to tobacco use [1]. The relationship between lung and 30% and 70% in the presence of other psychiatric
cancer deaths and US cigarette consumption over time disorders (Table 23.1) [5]. Because of the concentra-
is particularly striking (Figure 23.1). Lung cancer tion of nicotine dependence within the psychiatrically
deaths were practically nonexistent in the early 1900s. ill population, individuals with a psychiatric disorder
As smoking increased, lung cancer increased. In the consume 46% of all cigarettes smoked in the United
United States, an estimated 19.8% of adults are cur- States. Although these data indicate the strong need
rently smokers [2]. Due to the high prevalence and for smoking cessation in the psychiatrically ill popu-
severe health consequences of cigarette smoking, the lation, quit rates are substantially lower for subjects
United States spends $96 billion annually on smoking- with active mental illness as compared to smokers
attributable health-care expenditures [3]. without mental illness (31% versus 43% quit rate) [4].

5000 Figure 23.1 US cigarette consumption


70 and age-adjusted lung cancer death rates
4500 over time. Data is from the public-use
Per capita cigarette consumption

data files, National Vital Statistics System,


4000 60 National Center for Health Statistics,
(age-adjusted, per 100 000)

Centers for Disease Control and


Lung cancer death rate

3500 Per capita


50 Prevention, and the Tobacco Outlook
cigarette
3000 Report, Economic Research Service,
consumption
40 US Department of Agriculture.
2500
Lung cancer
2000 death rate 30
1500
20
1000
10
500
0 0
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
19 5
19 0
95
20 0
05
0
0
1
1
2
2
3
3
4
4
5
5
6
6
7
7
8
8
9

0
19

20

Year

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

287
Chapter 23: Nicotine dependence

Table 23.1 DSM-IV diagnoses and comorbidity with nicotine experiment with smoking; (2) regular smoking,
dependence.
defined as having smoked at least 100 cigarettes during
Sample Prevalence their lifetime; and (3) nicotine dependence, a psychi-
size of nicotine atric disorder defined by symptoms of tolerance,
dependence (%) withdrawal, and loss of control [8]. The criteria for
Entire sample 43 093 12.8
nicotine dependence, specified by the latest version of
the Diagnostic and Statistical Manual for Psychiatric
Alcohol dependence 1484 45.4 Disorders (DSM-IV-TR), parallel the dependence cri-
Drug dependence 549 69.3 teria for other substances including alcohol, cocaine,
Major depression 3119 30.0
and opiate dependence. For research purposes, other
psychometric instruments have been developed to
Mania 724 35.3 measure nicotine dependence. The most commonly
Generalized anxiety 894 32.7 used measure is the Fagerström Test of Nicotine
disorder Dependence (FTND) [9]. Table 23.2 delineates the
Panic disorder with 254 39.8 criteria for the DSM-IV-TR definition of nicotine
agoraphobia dependence as compared to the FTND-based defin-
ition. Since the DSM-IV-TR definition of dependence
Antisocial 1422 42.7
can be applied to any substance, the criteria are less
personality disorder
specific for nicotine dependence. Can something be
Data from Grant et al. [5]. said about the peculiar vulnerability of the adolescent
to nicotine addiction?

Not only is a significant proportion of nicotine


dependence comorbid with psychiatric illness, but a Genetics of nicotine dependence
significant portion of morbidity in psychiatric illness The conceptualization of the different stages of
may be attributed to nicotine dependence. It is smoking is helpful in understanding the varying
striking to note that the greatest risk of mental illness environmental and genetic contributions to each step
is premature death due to heart disease and cancer. In in the development of smoking behaviors. The herit-
an 8-state comparison of deaths of clients at public ability of smoking initiation was estimated to be 44%
mental health clinics, public mental health clients [10], with a sex differential of 37% for males and 55%
lived 13–30 years less than their general public coun- for females [11]. Conversely, the heritability of nico-
terpart [6]. Although suicides and accidental deaths tine dependence was markedly higher at 75% [10]. To
are elevated in this population as compared to con- underscore the social influences of smoking initiation
trols, the primary causes of the increase in deaths are in contrast to nicotine dependence, Saccone et al.
heart disease and cancer. Therefore, the largest modi- found that twins starting to smoke at exactly the same
fiable risk factor for heart disease and lung cancer in time is a good index for shared social influences on
the mental health population is cigarette smoking. smoking initiation, but it is unrelated to nicotine
In addition, nicotine dependence is associated with dependence as defined by DSM-IV and FTND [12].
disease-specific poor outcomes. For example, lifetime This indicates that the significant social influences on
smoking in bipolar disorder is associated with earlier smoking initiation do not appear to influence the
age of onset of symptoms, greater severity of symp- development of nicotine dependence. Multiple twin
toms, poorer functioning, history of suicide attempts, studies have found little common environmental
and comorbid anxiety and substance use disorders influence in risk for nicotine dependence.
[7]. The increased mortality and more severe course Although smoking initiation and nicotine depend-
of illness highlights the importance of understanding ence are related, the genetics of nicotine dependence
the etiology of nicotine dependence in the psychiatric are of primary interest due to its public health ramifi-
population. cations: nicotine dependence predicts difficulty with
In order to better understand smoking behavior, cessation and carries most of the morbidity associated
it is compartmentalized into: (1) initiation, the period with smoking. Specifically, the quantity of cigarettes
during which subjects smoke for the first time and smoked over a lifetime is associated with lung disease

288
Chapter 23: Nicotine dependence

Table 23.2 Definitions of nicotine dependence using DSM-IV-TR and Fagerström Test of Nicotine Dependence (FTND).

DSM-IV-TR [8] FTND [9]


3 (or more) of the following in a 12 month period: Cigarettes per day:
 Tolerance < 10 0 points
 Withdrawal 11–20 1 point
 Using more than intended 21–30 2 points
 Difficulty controlling use > 30 3 points
 Spending a great deal of time obtaining, Time after waking for first cigarette:
using or recovering from substance < 5 min 3 points
 Giving up activities because of use 6–30 min 2 points
 Use despite harm 31–60 min 1 point
>1h 0 points
Smoked more frequently in AM than PM:
Yes 1 point
No 0 points
Smoked where forbidden:
Yes 1 point
No 0 points
Cigarette most valued:
First cigarette after waking 1 point
Any other cigarette 0 points
Smoked when too ill to get out of bed:
Yes 1 point
No 1 point
Nicotine dependence is defined as FTND score ≥ 4

and heart disease, and smoking cessation is a positive addition to linkage studies, isolated candidate gene
prognostic factor for both. Nicotine dependence case-control studies targeted various neurotransmitters
accounts for 13% of the general population but con- and their receptors, but these generally had low
sumes 58% of the cigarettes smoked in the United power and insufficient density of genotypes.
States [5]. Thus, understanding the genetics of nico- The advent of the genome-wide association stud-
tine dependence can lead to targeted treatments and ies (GWAS) brought increased genomic density
ultimately significantly decrease tobacco-associated and large datasets. GWAS relied on the completion
morbidity and mortality. of the International HapMap Project in 2005, where
In the study of nicotine dependence, it is import- common single nucleotide polymorphisms (SNPs)
ant to understand the behavioral progression to nico- were mapped tightly throughout the genome. Using
tine dependence when choosing a control group. a case-control design and hundreds of thousands of
Subjects who have not had adequate nicotine expos- SNPs spanning the entire genome, GWAS look for
ure (have not smoked enough cigarettes) have not association between SNPs and disease. Although there
had opportunities to become nicotine dependent. are more statistical tests of association (one test per
For this reason, genetic studies of nicotine depend- SNP) than there are subjects (typically thousands), the
ence carefully choose their control group as regular combination of setting a sufficiently low threshold
smokers who did not become nicotine dependent. for p-value significance and placing a high emphasis
This minimizes the possibility that the controls would on replication of significant results in independent
be cases if they had adequate nicotine exposure. datasets has resulted in robust associations.
The strong heritability of nicotine dependence has The first GWAS on nicotine dependence was
led to genetic studies. The initial efforts to identify conducted by Bierut et al. [13]. The study compared
susceptibility loci primarily used linkage designs. Unfor- 1050 nicotine dependent cases (FTND  4), and 879
tunately, no associations were robustly replicated. In controls (smoked  100 cigarettes in lifetime and

289
Chapter 23: Nicotine dependence

lifetime FTND ¼ 0). Although no individual SNPs in many other independent studies. Using the quan-
reached genome-wide significance, the data were titative trait of cigarettes per day as a proxy for nico-
reanalyzed for a targeted association study with 3713 tine dependence, Berrettini et al. [15] performed a
SNPs in 348 candidate genes [14]. Several cholinergic GWAS and found an association of cigarettes per
nicotinic receptor genes dominated the top signals day with SNPs in the region of CHRNA3–CHRNA5.
after controlling for multiple testing. In order to further characterize the genetic relation-
ships in this region, these associations were mapped
with rs16969968. All the significant SNPs in the
Moving from association to function CHRNA3–CHRNA5 region formed a single haplotype
It is biologically relevant that the top genetic associ- block, with high and low risk alleles. Thus, in inde-
ations for nicotine dependence were in genes encod- pendent samples, association was seen between nico-
ing for nicotinic receptor subunits. Nicotine produces tine dependence and rs16969968 or related SNPs.
its central and peripheral actions by binding to neur- CHRNA5 and CHRNA3 are nicotinic receptor
onal nicotinic acetylcholine receptors (nAChRs), a subunit genes on chromosome 15q25 coding for the
class of neuronal ligand-gated ion channels expressed a5 nicotinic receptor subunit and the a3 nicotinic
in the nervous system. nAChRs are made of five receptor subunit respectively (Figure 23.3a). The coding
combinations of a and b subunits constructed around sequences are adjacent to one another and SNPs in the
a central pore. The subunits are encoded by nine a two genes are in high linkage disequilibrium (LD)
(a2–a10) and three b (b2–b4) subunit genes, named (Figure 23.3b). rs16969968 is seen in Figure 23.3b in
CHRNA2–CHRNA10 and CHRNB2–CHRNB4 respect- the coding region of CHRNA5. Although rs16969968
ively. The expression of the different subunits in spe- is in the coding sequence of CHRNA5, and a biologic-
cific anatomical areas leads to hypotheses regarding ally plausible relationship exists between CHRNA5
their functional relevance. The addiction of nicotine and nicotine dependence, SNPs in high LD with
is thought to arise in part from the interaction rs16969968 span a large area encompassing the genes
between dopaminergic and nicotinic neurons in the IREB2, PSMA4, CHRNA5, CHRNA3, and CHRNB4.
striatum (Figure 23.2). Multiple nicotinic subunits are Of the SNPs in the region of CHRNA5 and
involved in this interaction including a4, a5, a6, b2, CHRNA3, rs16969968 is the most studied. Although
and b3. This region has been implicated in the reward other SNPs associated with rs16969968 in the region
pathway and is important for the development of have been directly associated with nicotine depend-
substance dependence. ence, rs16969968 is a nonsynonymous SNP associated
Although GWAS are a technological break- with nicotine dependence. Specifically, the minor
through in terms of the density of markers that are variant of rs16969968 results in an amino acid change
tested across the genome, the genome is not com- of aspartic acid to asparagine. The frequency of the
pletely represented by the SNP arrays used in GWAS. minor allele varies between populations, as seen in
For example, the Affymetrix Genome-Wide Human Figure 23.4. It ranges from 0% in African populations
SNP Array 6.0 genotypes 906 600 SNPs out of an to 37% in European populations.
estimated 10 000 000 SNPs in the human genome The biological importance of rs16969968 has
(approximately 2 000 000 of which are known). The been shown in several settings. First, the a5 nicotinic
dense coverage in GWAS, although a vast improve- receptor is expressed in the brain. Its expression in the
ment over previous methodology, is still not dense striatum and direct interaction with the dopaminergic
enough to ensure that associations are found. None- pathway is particularly relevant to addiction. Second,
theless, a SNP reaching genome-wide association sig- the protein sequences of CHRNA5 homologues was
nificance indicates that functional variation is likely examined in multiple species (human, chimpanzee,
to occur through this SNP or one of the correlated, Bolivian squirrel monkey, domestic cow, mouse,
untested SNPs. chicken, and African clawed frog) and the aspartic
The most biologically compelling association with acid residue was present in all species [16]. This
nicotine dependence was found in rs16969968, a non- conservation across species suggests that it has func-
synonymous SNP in the a5 nicotinic receptor subunit tional importance. Third, an in vitro functional study
gene CHRNA5 [13, 14]. This association has been found that a4a5b2 receptors containing the aspara-
replicated with either rs16969968 or correlated SNPs gine amino acid substitution in a5 exhibited an

290
Chapter 23: Nicotine dependence

Figure 23.2 Nicotinic acetylcholine


(a) receptors in the striatum of the rodent
central nervous system. (a) Scheme of the
mesostriatal dopaminergic pathway. (b)
Subunit composition of the functional
nicotinic acetylcholine receptors
(nAChRs) expressed by dopaminergic
nerve terminals (a6b2b3, a6a4b2b3,
a4b2, a4a5b2). (From [29], with
Caudatum-putamen permission from Elsevier.)
and
nucleus accumbens

Substantia nigra
and VTA

(b)

α6β2β3
α4β2

α4α5β2 DA cell
Glu cell
α6α4β2β3

GABA cell

ACh cell

altered response to a nicotine agonist than did the nicotine dependence. Although rs588765 can lead to
receptors with the aspartic acid variant in a5 [16]. This both high expression and low expression of CHRNA5
suggests that decreased nicotinic receptor function is and rs16969968 can lead to high risk and low risk for
associated with increased risk for nicotine dependence. nicotine dependence, only three haplotypes exist in
In-depth analyses of the region led to the discov- the population: high risk (rs16969968) and low expres-
ery of additional SNPs that modify expression of sion (rs588765), low risk (rs16969968) and low expres-
CHRNA5 [17, 18]. The SNP rs588765 was found to sion (rs588765), and low risk (rs16969968) and high
be associated with nicotine dependence and to modify expression (rs588765) (Table 23.3). The lowest risk of
expression of CHRNA5. In addition, together with nicotine dependence occurs with the low expression
rs16969968, it forms a haplotype to alter the risk of allele of rs588765 and the low risk allele of rs16969968,

291
Chapter 23: Nicotine dependence

(a) chr15 (q25.1)

(b) rs16969968

76589924 76711042
1.0 80

Recombination rate (cM/Mb)


0.8
60
R-Squared

0.6
40
0.4

20
0.2

0.0 0
CRABP1 IREB2 LCC123088 ADAMTS7
PSMA4
CHRNA5
CHRNA3
CHRNB4

76420 76545 76670 76795 76920


Chromosome 15 position (hg18) (kb)
Figure 23.3 Graphical representation of chromosome 15q25. (a) Chromosome 15 with the region q25.1 demarked by the red line, the
location of CHRNA5 and CHRNA3. This figure was created with the University of California, Santa Cruz (UCSC) genome browser (http://genome.
ucsc.edu). (b) The 100 kb region surrounding rs16969968 (large diamond). Diamonds indicate single nucleotide polymorphisms (SNPs), and the
size of the diamond is directly proportional to the r2 of the SNP with rs16969968. The dashed lines delineate the boundaries for SNPs that have
r2  0.8 with rs16969968. This figure was created using SNAP (http://www.broad.mit.edu/mpg/snap/). See plate section for color version.

and the highest risk for nicotine dependence occurs


with the low expression allele of rs588765 and the Genetics of lung disease
high risk allele or rs16969968. A similar association The CHRNA5/CHRNA3/CHRNB4 region has also been
was seen in a population of African-Americans with associated with lung cancer and chronic obstructive
the SNP rs555018, a SNP in high LD with rs588765 in pulmonary disease (COPD). Unfortunately, the fact
both European-Americans and African-Americans, that both lung cancer and COPD are nearly completely
suggesting that rs555018 is associated with variable attributable to smoking makes it difficult to determine
expression of CHRNA5 [17]. whether the observed association is independent of
The GWAS results and corresponding biological nicotine dependence. The difficulty with interpretation
experiments suggest that CHRNA5 is involved in the has been emphasized in multiple studies.
development of nicotine dependence, but the relation- Thorgiersson et al. [19] highlighted the import-
ship between changes in structure and expression of ance of gene–environment interactions with a GWAS
CHRNA5 and the development of smoking behavior that found associations between a SNP in LD with
is unknown. Various research groups are evaluating rs16969968, rs1051730, lung cancer, and peripheral
the relationship between these genes and smoking artery disease, in addition to confirming the prior
initiation, the pleasure associated with smoking, and association seen with nicotine dependence. Their
age of onset of smoking. hypothesis was that rs1051730 leads to nicotine

292
Chapter 23: Nicotine dependence

Table 23.3 Risk of nicotine dependence based on haplotypes constructed from rs16969968 and rs588765. Alleles of rs16969968 are
represented as high risk allele for nicotine dependence and low risk allele for nicotine dependence. Alleles for rs588765 are represented
as high expression of CHRNA5 and low expression of CHRNA5. The haplotype of high risk allele (rs16969968) and high expression allele
(rs588765) is not listed because it is extremely rare in the population.

Haplotype risk of nicotine dependence

High Medium Low


rs16969968 High risk allele Low risk allele Low risk allele
rs588765 Low expression allele High expression allele Low expression allele

8
4
7
6
3 5

1
2 9

Figure 23.4 Allele frequency distribution (%) of cSNP rs16969968 in different ethnic populations. The frequency of the minor allele is
the white segment of the circles in the figure. Populations were grouped together on the basis of their genetic structures. Geographic
regions: 1. America; 2. Africa; 3. North Africa; 4. Europe; 5. Middle East; 6. Central/South Asia; 7. Centra/South Asia; 8. East Asia; 9. Oceania.
(From [16] with permission.)

dependence and subsequent smoking behavior. An people who had never smoked, but the risk was not
association with lung cancer may be explained by statistically significant. This evidence led to the con-
the relationship with smoking behavior, but there clusion that the association was unrelated to nicotine
may also be a component of lung cancer that is driven dependence.
independently of nicotine dependence. Another study also found association with lung
Other studies, however, reported evidence that cancer independent of cigarette quantity. Amos et al.
rs1051730 and related SNPs were associated with lung [21] found an association between lung cancer and
cancer independent of smoking behavior. Hung et al. rs1051730 in three independent samples, robust to
[20] found an association between lung cancer and adjustment for pack-years of cigarette exposure. The
rs1051730 with equal risk seen for former smokers as association between rs1051730 and lung cancer was
compared to current smokers. There was an increased absent in nonsmokers. Ideally, if rs1051730 were
risk for lung cancer associated with rs1051730 in related to lung cancer independent of smoking

293
Chapter 23: Nicotine dependence

exposure the effect would be present in nonsmokers. proxy for nicotine and toxin absorption, the associ-
However, the repeated evidence of association of this ation between lung cancer and rs16969968 may still
variant with lung cancer independent of smoking be confounded by the increased nicotine and toxin
exposure argues that there is an association between intake in nicotine dependence even after adjusting
rs1051730 (and related SNPs) and lung cancer beyond for cigarettes per day.
the smoking exposure from nicotine dependence.
The relationship between rs1051730, nicotine
dependence, and lung cancer was more specifically
Genetic association of nicotine
explored by Spitz et al. [22]. This was accomplished dependence with other mental illness
by using three samples: (1) lung cancer cases and The tendency of subjects with mental illness to smoke
controls with a history of smoking; (2) lung cancer has been extensively discussed in the literature.
cases and controls who were lifetime nonsmokers; Hypotheses regarding causation have been presented
and (3) bladder cancer cases and controls with a in both directions. Some researchers argue that
history of smoking. The investigators found a statis- smoking is a means of self-medicating and nicotine
tically significant association of rs1051730 with lung dependence is therefore “caused” by mental illness.
cancer in subjects with a history of smoking. When These studies present evidence that smokers use nico-
stratified for cigarettes per day, the highest risk is seen tine for depressive symptoms. More recently, data has
in the lightest smokers, subjects smoking less than been published that smoking precedes the diagnosis
20 cigarettes per day. There was no association of mental illness and may contribute to the disorder.
between lung cancer and rs1051730 in never smokers One study examined the association between cigarette
and no association between rs1051730 and bladder smoking and a subsequent first depression. It found
and renal cancers. The authors concluded that a four-fold higher risk of first-episode depression for
because the highest risk group was in the subjects that heavy smokers as compared to never smokers [24].
smoke the least, rs1051730 mediated risk to lung The current interpretation of these data is that
cancer beyond the cigarette exposure from nicotine smoking confers additional risk in the development
dependence. of psychiatric disorders.
A recent GWAS for COPD also found association of A third explanation for the comorbidity of nico-
rs1051730 with COPD. This association remained sig- tine dependence and other psychiatric illness is a
nificant after adjusting for pack-years (p ¼ 5.7  10–10). shared etiology. The heritability of major psychiatric
Although suggestive that this is a true association with illness is high with estimates of 64% in schizophrenia
COPD, the authors reported that the lack of further [25], 59% in bipolar disorder [25], and 42% in major
smoking variables lead to difficulty teasing out the full depressive disorder [26]. The combination of the
contribution of nicotine exposure on COPD. heritability of nicotine dependence, the heritability
One explanation for the association of rs16969968 of psychiatric illnesses, and comorbidity of nicotine
with lung disease independent of cigarette quantity is dependence with psychiatric illness indicate that spe-
that the quantity of toxin inhaled when smoking is cific genes may be pleiotropic, i.e. contribute to both
not well explained by the cigarettes per day. To inves- nicotine dependence and other psychiatric illness.
tigate this, Le Marchand et al. [23] conducted a study Due to the comorbidity of alcohol dependence
to evaluate the amount of nicotine and carcinogens and nicotine dependence, the association between
consumed in a sample of 819 smokers. A positive CHRNA5 and alcohol dependence was evaluated.
association was found between the amount of nico- Although alcohol dependence was not associated with
tine metabolites seen in urine, a measure of the rs16969968, it was associated with SNPs related to
amount of nicotine consumed, and the nicotine rs588765, leading to a decreased expression of CHRNA5
addiction risk allele (A) of rs16969968 after adjusting [27]. We expect shared mechanisms between nicotine
for cigarettes per day. This indicates more nicotine dependence and alcohol dependence due to the fact
and associated carcinogens are absorbed by subjects that they both involve addiction, the molecular prop-
with the risk allele than subjects without the risk erties of nicotine and alcohol are unrelated. This
allele. Moreover, the differential absorption is not suggests that expression of CHRNA5 may function
accounted for by cigarettes per day. Because these in addiction at a general level, rather than contrib-
data argue that cigarettes per day is not a complete uting exclusively to nicotine dependence.

294
Chapter 23: Nicotine dependence

Due to the phenotypic complexity of psychiatric biological mechanisms that alter the risk of nicotine
illnesses, the search for pleiotropic genes may help dependence. The first biological mechanism is caused
clarify psychiatric diagnoses and help understand by an amino acid change in CHRNA5, in the non-
biological etiologies of mental illness. Not only is synonymous SNP rs16969968. Functional studies
there a relationship between nicotine dependence have found that expression of the minor variant leads
and other psychiatric illness, but there are comorbid- to reduced response to a nicotinic agonist. The second
ities within and between substance abuse, personality mechanism altering risk of nicotine dependence
disorders and major psychiatric disorders. Suspicion is through altered expression of the a5 mRNA.
of shared genetic susceptibilities with bipolar disorder A second group of SNPs in the region, including
and schizophrenia has been the subject of recent rs588765 and rs555018, is associated with variable
research. A study by Ferreira et al. analyzed both expression leading to decreased expression of
a bipolar GWA dataset and a schizophrenia GWA CHRNA5. The combination of altered protein and
dataset for overlap of top signals, and found evidence variable mRNA expression leads to different levels
of shared association [28]. These studies suggest of addiction risk.
that advances may be made by understanding genetic Associations in this region have also been found in
pleiotropy and interconnections between psychiatric lung disease. Specifically, SNPs correlated to rs16969968
illnesses. are associated with an increased risk of COPD and lung
cancer, after adjusting for cigarette use. Although the
Conclusions interpretation of these studies is difficult due to the
confounding between lung disease and cigarette expos-
Nicotine dependence is a heritable disorder with
ure, there is evidence of a genetic interaction between
major public health ramifications. Nicotine depend-
smoking and rs16969968 that may confer additional
ence accounts for 58% of all cigarettes smoked in the
risk for lung disease beyond cigarette exposure.
United States, and has led to an epidemic of comor-
The comorbidity of nicotine dependence with
bidities including lung cancer and heart disease. The
other heritable psychiatric disorders leads us to investi-
population with psychiatric illness is particularly sus-
gate genetic pleiotropy. Association was found between
ceptible to nicotine dependence. Although psychiatric
alcohol dependence and expression of CHRNA5.
illness is present in 28% of the US population,
Advances in understanding of the genetics in nicotine
the mentally ill consume 44% of cigarettes smoked.
dependence may therefore not only impact the diag-
In addition, public mental health clients had a
nosis and treatment of nicotine dependence, but also
decreased life span of 13–30 years, where the majority
may help us understand shared genetic susceptibilities
of deaths are due to heart disease and cancer. Cigar-
in other mental illnesses.
ette smoking is therefore the largest modifiable risk
factor for morbidity and mortality both in the general
population and in the psychiatric population. Acknowledgements
GWAS have found associations between nicotine The authors are supported by NIH grants UL1RR024992,
dependence and the a5 nicotinic receptor subunit KL2RR024994, K02DA021237, P01CA089392, and
gene. We postulate that there are at least two distinct T32MH014677.

References 5. Grant BF, et al. Arch Gen Psychiatry


2004;61(11):1107–1115.
9. Heatherton TF, et al. Br J Addict
1991;86(9):1119–1127.
1. World Health Organization. The
World Health Report 2002. Geneva: 6. Colton CW, et al. Prev Chronic Dis 10. Vink JM, et al. Behav Genet
WHO; 2002. 2006;3(2):A42. 2005;35(4):397–406.
7. Ostacher MJ, et al. J Clin Psychiatry 11. Kafkafi N, et al. Behav Neurosci
2. Centers for Disease Control and 2003;117(3):464–477.
Prevention. MMWR Weekly 2006;67(12):1907–1911.
2008;57(45):1221–1226. 12. Saccone NL, et al. Am J Med
8. American Psychiatric Association.
Genet B: Neuropsychiatric
Diagnostic and Statistical
3. Centers for Disease Control and Genetics 2009;150B(4):453–466.
Manual of Mental Disorders,
Prevention. MMWR Weekly 13. Bierut LJ, et al. Hum Mol Genet
Fourth Edition, Text Revision
2008;57(45):1226–1228. 2007;16(1):24–35.
(DSM-IV-TR). Washington, DC:
4. Lasser K, et al. JAMA 2000;284 American Psychiatric Association; 14. Saccone SF, et al. Hum Mol Genet
(20):2606–2610. 2000. 2007;16(1):36–49.

295
Chapter 23: Nicotine dependence

15. Berrettini W, et al. Mol Psychiatry 20. Hung RJ, et al. Nature 2008; 25. Lichtenstein P, et al. Lancet
2008;13(4):368–373. 452(7187):633–637. 2009;373(9659):234–239.
16. Bierut LJ, et al. Am J Psychiatry 21. Amos CI, et al. Nat Genet 2008; 26. Edvardsen J, et al. J Affect Disord
2008;165(9):1163–1171. 40(5):616–622. 2009;117(1–2):30–41.
17. Saccone N, et al. Cancer Res 22. Spitz MR, et al. J Natl Cancer Inst 27. Wang JC, et al. Mol Psychiatry
2009;69(17):6848–6856. 2008;100(21):1552–1556. 2009;14(5):501–510.
18. Wang JC, et al. Hum Mol Genet 23. Le Marchand L, et al. Cancer Res 28. Ferreira MA, et al. Nat Genet
2009;18(16):3215–3135. 2008;68(22):9137–9140. 2008;40(9):1056–1058.
19. Thorgeirsson TE, et al. Nature 24. Klungsoyr O, et al. Am J 29. Gotti C, et al. Prog Neurobiol
2008;452(7187):638–642. Epidemiol 2006;163(5):421–432. 2004;74(6):363–396.

296
Human molecular genetics
Chapter

24 of opioid addiction
Mary Jeanne Kreek, Dmitri Proudnikov, David A. Nielsen,
and Vadim Yuferov

Introduction in the receptor protein and/or in the gene expression


due to sequence differences. After successful cloning of
Addiction to opiates is a chronic, relapsing brain the MOPr gene (OPRM1) in 1995 [3], our laboratory,
disease that, if left untreated, can cause major med- in collaboration with the laboratory of Dr. Lei Yu,
ical, social and economic problems. There are at least started studies on functionality of the gene variants in
three different categories of factors that contribute to the coding region. Among several single nucleotide
the vulnerability of developing a specific addiction, polymorphisms (SNPs), we found two SNPs of high
once self-exposed: (1) environmental factors, includ- allelic frequency, each of which alters an amino acid in
ing cues, conditioning, external stressors, and the the N-terminus of the receptor. The 17T (rs1799972)
stress they cause; (2) drug-induced factors, that lead variant results in an amino acid change of alanine to
to a variety of molecular neurobiological changes valine at position 6 (A6V) (overall allelic frequency
resulting in altered behaviors; (3) genetic factors, 6.6%), and the A118G, variant, with allelic frequency
which represent approximately 40–60% of the risk range around 2% in African-Americans to over 40% in
of developing an addiction to opioids [1]. In addition, Japanese, which results in amino acid change from
addiction to opioids may sometimes arise from asparagine to asparatic acid at position 40 (N40D).
opioid treatment of chronic pain [2]. One of these, the 118G (rs1799971), removes one of
In this review we present several experimental five potential N-glycosylation sites, which may affect
approaches performed in the Laboratory of the Biol- the trafficking of MOPr into the plasma membrane,
ogy of Addictive Diseases to characterize the relation- and an interaction of ligands with the receptor.
ship of gene variations with heroin addiction and In the first report of in vitro functional studies
pharmacogenomics. of MOPr encoded by either 118A or 118G in the
hamster fibroblastoid tumor cells AV-12, we found
The mu-opioid receptor that the MOPr 118G variant or the more common
The mu-opioid receptor (MOPr) is the primary site of prototypic 118A receptor variant bound most of
action of several of the endogenous opioid peptides, the exogenous and endogenous opioid ligands with
including beta-endorphin. This receptor is also the similar affinity [4]. However, when the longest of the
major target for clinically important opioid analgesic endogenous opioids, beta-endorphin, was studied,
agents, including morphine, methadone, fentanyl, we found striking differences. Beta-endorphin bound
and related drugs, as well as a major site of action with a three-fold higher affinity to the 118G variant as
for heroin. MOPr belongs to the seven transmem- compared to the 118A variant. Also, in Xenopus
brane receptor family and it is coupled to an inhibi- oocytes co-expressing the 118G OPRM1 variant with
tory G protein (Gi) that inhibits adenylate cyclase the G protein-activated inwardly rectifying potassium
activity, resulting in a reduction of intracellular cyclic channel, a three-fold greater channel activation was
adenosine monophosphate (cAMP) production, which found when activated by beta-endorphin compared to
acts as a second messenger within cells. Clinical activation of the prototype 118A variant.
observations have shown that individuals have varied In the second study, we assessed the effects of the
sensitivity to opioids, suggesting potential variability A118G SNP of the OPRM1 and cell type on cell-surface

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

297
Chapter 24: Human molecular genetics of opioid addiction

expression, agonist binding affinity, and receptor sig-


naling through inhibition of adenylate cyclase [5]. In Epigenetics of the MOPr
cell lines AV-12 and HEK293, which stably expressed Epigenetics is the study of inherited changes in
either 118A or 118G variants, the [3H]-DAMGO gene expression that are not encoded in the sequence
whole-cell saturation binding assay showed lower of the DNA. Epigenetic regulation provides a layer of
cell-surface expression of the MOPr 118G variant. control that can regulate gene expression. The two
Affinity (KD) of the 118G variant was found to be main processes that control epigenetics are DNA
significantly greater in HEK293 cells whereas there methylation and histone modifications. Alterations
was no difference in the binding affinity between the in these processes may disregulate gene expression
variant and the prototype MOPr in AV-12 cells. To with significant clinical outcomes. These alterations
characterize the agonist-dependent receptor signaling may be passed to daughter cells and also to successive
properties, a decrease of forskolin-induced cAMP generations.
levels were assayed in AV-12 and HEK293 cell lines Cytosine methylation by DNA methyltransferases
expressing the 118A or 118G variants of MOPr at cytosine:guanine dinucleotide (CpG) sites in geno-
following stimulation with opioid receptor agonists mic DNA represent a key epigenetic pathway modify-
morphine, methadone, DAMGO, and beta-endor- ing gene expression. The methylation of specific CpG
phin. All agonists except beta-endorphin displayed sites in genomic DNA may alter transcription factors
greater potency (EC50) in both cell lines expressing binding to DNA regulatory sequences (e.g. [11, 12]).
the 118G variant. In contrast to stable cell lines, the DNA methylation of CpG dinucleotide sites in the
prototype MOPr or the 118G variant transiently promoter regions of genes is usually associated with
expressed in either AV-12 or HEK293 showed no downregulation of gene expression [13].
difference in EC50 of DAMGO or beta-endorphin by Our laboratory has recently found and reported
the cAMP assay in both cell lines. that two CpG sites in the promoter region of the
A lower in vitro expression of the OPRM1 118G MOPr gene OPRM1, the site of action of opiates and
variant was also reported by other laboratories in opioids, have increased levels of methylation in
stably expressed HEK293 cells [6] and transiently former severe heroin addicts well stabilized in metha-
expressed CHO cells [7]. Moreover, Sadee and col- done maintenance treatment [14]. Of the 16 CpG sites
leagues tested the A118G SNP for allele-specific gene examined in the OPRM1 promoter region, the –18
expression in human postmortem brain tissues from and the þ84 CpG sites had hypermethylation in
cortical lobe and pons, and showed two-fold lower former addicts. Both of these sites are located in
OPRM1 mRNA levels of the G allele than the A allele potential Sp1 transcription factor-binding sites. It is
[7]. Also, they found that the overall OPRM1 mRNA possible that methylation of these CpG sites could
expression in pons in subjects homozygous for 118G lead to reduced OPRM1 expression in former heroin
was lower compared to the 118A variant. It is very addicts. The DNA hypermethylation of these specific
likely that the 118G variant has altered MOPr expres- CpG sites in the former heroin addicts could be due
sion at the level of mRNA transcription as well as at to: (1) inheritance through genomic imprinting; (2)
cell surface presentation. In vitro functional analyses changes occurring earlier in life which predispose to vul-
of A118G have shown different results in some stud- nerability to develop heroin addiction; (3) chronic heroin
ies [6, 8], probably due to differences in experimental use; or (4) long-term methadone pharmacotherapy.
condition. However, a dramatically different func- Other studies have shown that other drugs
tional effect of the A118G SNP has been demonstrated of abuse can alter DNA methylation or histone modi-
by many studies in our laboratory and others. These fications. Alcoholism has been shown to increase
studies have demonstrated an association of this overall DNA methylation [15, 16], decrease DNA
receptor variant with both opioid and alcohol addic- methyltransferase expression [17], and increase spe-
tions, and, further, documented functional differ- cific DNA methylation at specific genes [15, 18]. In
ences, including pharmacodynamic responses to rodents, cocaine administered on an acute or chronic
mu-opioid receptor antagonist, naltrexone, medica- basis has been shown to cause chromatin remodeling
tion (pharmacogenetics), and physiological responses, through histone modification at the cFos, or at the
including stress responsivity (physiogenetics) [reviewed bdnf and cdk5 promoters, respectively [19], which
in 1, 9, 10]. may be due to a decrease in the histone deacetylase

298
Chapter 24: Human molecular genetics of opioid addiction

HDAC5 function [20]. Cocaine has also been shown sequence [28]. Studies of the gene structure in our
to decrease histone methylation in the prefrontal laboratory defined an additional 50 exon of the
cortex [21], while maternal cocaine administration untranslated region and several transcription initi-
increases cytosine methylation in the promoter of ation sites [29]. The redefinition of the exon/intron
the protein kinase Cepsilon (PKCE) gene in fetal organization of the human OPRK1 gene has facili-
heart [22]. tated the localization of the proximal promoter of
Several pharmacological agents such as azacitidine the gene, and potential regulatory polymorphisms.
and valproic acid (discussed in [14]) which alter DNA Two studies have evaluated variants of the OPRK1
modification or histone modification have been used gene for possible association with opiate addiction.
in clinical therapies for cancer treatments. Perhaps Yuferov and colleagues [29] analyzed 12 single
these may be used to modify epigenetic states in nucleotide polymorphisms (SNPs), including 5 novel
addiction. variants in 145 former heroin addicts and 146
control subjects. A logistic regression analysis showed
a point-wise significant association of the synonym-
The kappa-opioid receptor ous SNP G36T (rs1051660) with opiate addiction
Although opioid receptors often subserve similar (p ¼ 0.016). The number of haplotypes of eight high
physiological functions, activation of the kappa-opioid frequency SNPs seen in the three ethnic groups were
receptor (KOPr) by exogenous agonists produces dys- nine, six, and five for African-Americans, Caucasians,
phoria in humans and aversive effects in experimental and Hispanics, respectively, with corresponding sig-
animals, in contrast to activation of MOPr [23]. KOPr nificance levels for differences in haplotype frequen-
agonists, including dynorphin, the primary endogen- cies between cases and controls of p ¼ 0.0742, 0.1015,
ous peptide ligand for this receptor, decrease basal and 0.0041. In Hispanics, the AGCTCGTC haplotype
and also drug-induced increases of dopamine levels is found only in control subjects (frequency of 0.06),
in several areas of the dopaminergic nigrostriatal and while the GGCGTGCC (frequency of 0.07) is found
mesolimbic–mesocortical systems [24]. The KOPr- only in subjects with opiate addiction. An association
dynorphin system may therefore be considered to be of the SNP G36T with opiate addiction was also
a part of the counter-modulatory mechanisms of the shown in a Caucasian population of heroin addicts
brain following direct or indirect drug-induced dopa- [30]. The function of the SNP G36T is not known,
minergic stimulation [1]. The ability of dynorphin and it may be in linkage disequilibrium (LD) with
peptides to modulate dopaminergic tone in humans other cis-regulatory polymorphisms in OPRK1 gene.
has been demonstrated in studies conducted in normal Recently, a novel high frequency insertion of 830 bp
volunteers and also in methadone-maintained former in a putative promoter region of OPRK1 was reported
heroin addicts. These studies have measured prolactin [31]. In a reporter gene expression assay, the presence
release, which is under tonic inhibition by dopamine in of the insert reduced transcription activity of the
the tuberoinfundibular region of the hypothalamus. OPRK1 promoter to 53% of that in its absence.
Administration of dynorphin A(1–13), a shortened form
of the natural peptide dynorphin A(1–17), in humans
leads to an abrupt rise in serum prolactin levels caused Melanocortin receptor type 2
by a lower dopaminergic tone through activation of The melanocortin receptor type 2 (MC2R or adreno-
KOPr [25]. Prolactin response to dynorphin A(1–13) corticotropic hormone, ACTH receptor) belongs to
was significantly attenuated in methadone (mu-opioid the superfamily of G-protein-coupled receptors and is
agonist)-maintained volunteers compared to control involved in regulation of adrenal cortisol secretion,
subjects, suggesting either alterations in the KOPr important in the physiological response to stressors.
system or, more likely, a lowering of tuberoinfundibular Physical and psychological conditions have been found
dopaminergic tone in former heroin addicts [26]. Also, to be associated with dysregulation of the hypothal-
clinical studies have shown that dynorphin A(1–13) amic-pituitary-adrenal (HPA) axis: post-traumatic
attenuates opioid-withdrawal symptoms in humans [27]. stress disorder, fibromyalgia, Alzheimer’s disease,
The human OPRK1 gene is located on chromo- major depression, and specific stressors [32]. We found
some 8q11.2. The initial cloning of OPRK1 indenti- that specific addictive diseases also associated with
fied three exons that contained the complete coding dysregulation of the HPA axis: hyperresponsivity to

299
Chapter 24: Human molecular genetics of opioid addiction

removal of glucocorticoid negative feedback was COMT has been found in peripheral and central
found in cocaine addicts [33]; HPA hypoactivity was tissues [40]. A common SNP 472G>A alters amino
found in medication-free illicit drug-free former acid Val to Met at position 158 which results in
heroin addicts [34]. a four-fold decrease of activity of COMT [42]. In
Adrenocorticotropic hormone (ACTH) is the major heterozygous human lymphoblast cell lines and
hormone derived from proopiomelanocortin (POMC), brains, allele 158Met was overexpressed compared to
and regulates adrenal glucocorticoid and androgen syn- 158Val in all samples studied [43]. In a functional
thesis in the zonae fasciculata and reticularis in the magnetic resonance imaging (fMRI) study, amphet-
adrenal cortex. ACTH binds to its specific melanocortin amine administration enhanced the prefrontal func-
2 receptor (MC2R or ACTH receptor) [35]. MC2R, tioning in individuals homozygous for the Val allele
which is involved in steroidogenesis, is expressed mostly during a working memory task, while those homozy-
in adrenal cortex and was also found in human skin and gous for the Met allele showed no enhancement of
ovarian steroid cell tumor [32]. cortical efficiency at low to moderate working memory
In most human genetics studies, variants in the load [44].
MC2R gene have been linked to familial glucocorti- Variants of the COMT have been used in genetic
coid deficiency (e.g. [36, 37]). Recent studies showed studies of association with various neuropsychiatric
possible involvement of MC2R in stress mechanisms. conditions including schizophrenia, panic disorder, sui-
Substitution of an A to G in –179A>G (–2T>C) of cide, major depression, bipolar disorders, obsessive–
the MC2R gene results in lower promoter activity of compulsive disorder, attention deficit hyperactivity
this gene in vitro and was associated with impaired disorder (ADHD), and efficacy of response to treat-
cortisol response to ACTH stimulation in vivo [38]. ment of Parkinson’s disease [40].
A clinical study with a six-hour ACTH stimulation The high-activity 158Val allele variant has been
test showed that subjects with the –179AA genotype have found to be associated with abuse and addiction to
a higher dehydroepiandrosterone (DHEA) response several drugs in Caucasians [45], abuse of metham-
than –179GG subjects, while baseline DHEA concen- phetamine in Han Chinese [46], and heroin addiction
trations did not differ between groups [39]. in Caucasians [47]. In contrast, the 158Met variant of
In our studies [32], we found an experiment-wise COMT has been found to be associated with alcohol-
significant association of the minor A allele of the ism in Caucasians [48]. A number of different specific
polymorphism –184G>A of the MC2R gene and the haplotypes of COMT have been found to be associ-
haplotype AACT bearing –184A variant with a pro- ated with nicotine dependence in African-Americans
tective effect from the development of heroin addic- and Caucasians [1] and with cocaine dependence
tion in Hispanics. We also found a strong effect in in African-Americans [49]. A study of dopamine trans-
association of individual genotype frequencies of this porter (DAT)–COMT gene–gene interaction [50]
polymorphism or combination of GA þ AA geno- showed that subjects homozygous for 158Met/Met
types with a protective effect from heroin addiction, with lower COMT activity and greater dopamine tone,
providing genetic evidence supporting our hypothesis presumably, showed greater activation in both pre-
that dysregulation of the HPA axis contributes to the frontal cortex and ventral putamen measured in an
development of drug addiction. fMRI guessing task sensitive to reward.
The result of an association study might be
gender-specific: in one study the 158Met allele was
Catechol-O-methyltransferase associated with obsessive–compulsive disorder in
Catechol-O-methyltransferase (COMT) takes an males, but not in females [40]; in another study,
important part in metabolism of catechol neurotrans- the allele 158Val was associated with alcoholism
mitters including dopamine. Prolonged administration in American-Indian females, but not males [51].
of drugs of abuse can lead to alterations in the dopami- COMT homozygous knockout female mice develop
nergic system. In animal models, chronic administra- increased anxiety in a light/dark model compared to
tion of cocaine results in reduction of striatal dopamine COMT knockout males; also in male mice only an
and dopamine D2 receptors levels [40]. In subjects increased aggressive behavior in COMT heterozy-
addicted to drugs of abuse, brain imaging shows reduc- gous knockouts compared to other genotypes was
tions in striatal dopamine D2 receptors [41]. found [40]. In in vitro cellular studies, physiological

300
Chapter 24: Human molecular genetics of opioid addiction

concentrations of 17-beta-estradiol were shown to Recently our group has tested a number of SNPs
downregulate COMT gene transcription and protein of the HTR1B gene for association with heroin addic-
expression [40]. tion [53]. A protective effect from the development of
In recent studies performed by our group [40], we heroin addiction by the minor allele 1180G in Cauca-
tested four SNPs located in the coding region of the sians was found. Association studies of a haplotype
COMT gene for association with heroin addiction; containing a G at position 1180 showed stronger
an association of the Val158Met SNP in Hispanic association effects with protection from heroin addic-
subjects with opiate addiction was found. In this tion, suggesting an involvement of HTR1B in the
study, stratification of the data by gender has shown mechanisms that underlie the development of addic-
an association of 158 Val/Met and 158Met/Met tion to heroin.
genotypes with opiate addiction in women, but not
men [40].

5-hydroxytryptamine (serotonin)-1B Tryptophan hydroxylase 1 and 2


Various facets of impulsivity and mood are regulated
receptor by serotonin. Levels of cerebral spinal fluid
The 5-hydroxytryptamine (serotonin)-1B receptor 5-hydroxyindolacetic acid (CSF 5-HIAA), a degrad-
(HTR1B) is involved in many neuropsychiatric ation product of serotonin, are low in subjects with
functions including thermoregulation, locomotion behaviors characterized by deficits in impulse control
and feeding [52]. Serotonin receptor knockout mice [59, 60]. The rate-limiting enzyme in the biosynthesis
showed increased impulsive aggression, decreased of serotonin is tryptophan hydroxylase (TPH). There
anxiety, increased exploratory activity, increased are two known isozymes of TPH, TPH1 and TPH2,
special memory performance, increased locomotor which are encoded by separate genes, TPH1 and
response to cocaine administration, and increased TPH2, respectively [61]. TPH1 is expressed primarily
alcohol consumption [53]. Serotonin 1B receptor in the pineal gland, the developing raphe nuclei
knockout mice showed increased cocaine self- (during the late developmental stage), and the enter-
administration, and also administration of serotonin ochromaffin cells of the gut, while TPH2 is expressed
1B receptor agonists increased serotonin 1B function primarily in the raphe nuclei of the brain [62].
leading to reduced cocaine self-administration in TPH2 is also expressed in the brain in the hippo-
rats [53]. campus, hypothalamus, cortex, thalamus, amygdala,
In human genetics studies, polymorphisms of and cerebellum at levels slightly lower than TPH1
HTR1B gene were used in association studies with mRNA [63].
feeding disorders, ADHD, childhood-onset mood dis- Variants in the TPH1 gene have been found to
order, schizophrenia, bipolar disorder, major depres- be associated with CSF 5-HIAA concentrations [64],
sion, aggressive behavior, and suicide [38]. SNPs of the linked to alcoholism [65, 66], and associated with
HTR1B gene were also tested for an association with other addictive phenotypes, including age of onset
substance abuse and alcoholism [e.g. 53–56]. of alcoholism [67], drinking-related antisocial behav-
Gene expression studies reported that the T-161 ior [68], smoking initiation [69, 70], and nicotine
variant is expressed in a higher level than A-161 in dependence [71]. However, in our study, variants in
both BeWo and COLO 320 DM cell lines [55]. The TPH2 have only been found to be associated with one
haplotype variant consisting of –261G and –161A addictive disease, heroin addiction [72]. In our study,
enhances transcriptional activity 2.3-fold compared one TPH1 variant, which had previously been found
to the haplotype consisting of –261T and –161A. to be associated with alcoholism and smoking, and
The substitution of A with T in position –161 reverses six common TPH2 variants were genotyped. At the
this effect, making transcriptional activity of G-261/ two-locus genotype pattern level in Hispanics, the
T-161 equal to the major haplotype T-261/A-161 [57]. interaction of the TPH1 rs1799913 variant with
A common polymorphism rs13212041 in the 3’ the TPH2 rs7963720 variant was found to signifi-
untranslated region of the HTR1B gene was found to cantly interact and to be associated with heroin
attenuate microRNA miR-96 function and was asso- addiction. Also, a second interaction of the TPH1
ciated with aggressive behavior [58]. rs1799913 variant and the TPH2 rs4290270 variant

301
Chapter 24: Human molecular genetics of opioid addiction

was significantly associated with heroin addiction.


In the African-Americans, a significant association
Hypothesis-driven genes
of a specific TPH2 haplotype with heroin addiction association study
was found [72]. The haplotype blocks containing the In our recent study [74], we have used a SNP hypoth-
TPH2 and TPH1 variants may contain other variants esis-driven array that was designed by the group
that may differ based on ethnic background. The of D. Goldman at the National Institute of Alcohol
expression of the two TPH genes may coordinately Abuse and Alcoholism (NIAAA). Variants totalling
regulate serotonergic functions. This combined expres- 1350 in 130 candidate genes were scanned in a
sion may be dissimilar in the different ethnic groups, Caucasian population of 412 former severe heroin
and may affect an individual’s vulnerability to develop addicts in methadone treatment, and 184 healthy
heroin addiction with unique ethnic characteristics. controls. The majority of the subjects were from
the United States (New York City and Las Vegas)
and the minority from Israel. A total of 178 ancestry
P-glycoprotein gene (ABCB1) variants informative markers (AIMs) were employed to
exclude population stratification. The nine variants
and methadone dose requirements that showed the most significant associations were
Opiate addiction is a chronic relapsing disorder that in noncoding regions of the genes encoding the
is treated world-wide with methadone. Successful mu-, kappa-, and delta-opioid receptors; the neuro-
treatment relies, in part, on individual methadone peptide galanin; the serotonin receptor 3B; and the
dose optimization. The inter-individual variability in casein kinase 1 epsilon. The two OPRM1 variants
methadone dose-effectiveness may be determined in (rs510769 and rs3778151, intron 1) are in strong
part by genetic factors. The ABCB1 gene is highly LD and belong to a haplotype block of 32 kb. They
polymorphic and a few common variants have been are adjacent to rs1799971 (A118G, exon 1). However,
shown to be associated with P-gp expression, drug no evidence for association was found with SNP
response, and disease susceptibility. Since methadone A118G in this cohort. Galanin and its receptors have
is a substrate of P-glycoprotein (P-gp) 170 and few been shown to be involved in behavioral processes,
common ABCB1 variants were associated with pro- morphine withdrawal, and effects of opiates and high
tein expression and drug response, we assessed the stress response, among other functions. Casein
association between ABCB1 SNPs and methadone kinase 1 epsilon regulates the circadian clock gene
dose requirements [73]. P-gp is a member of the PER1, which may be involved in drug dependence and
ATP-binding cassette (ABC) superfamily. It is com- reward, and participates in important signaling path-
posed of two homologous sequences, each containing ways including DARPP-32 (the dopamine- and cyclic-
six transmembrane domains and an ATP binding AMP-regulated phosphoprotein) phosphorylation
domain. P-gp is expressed in tissues with barrier [74].
function, including the endothelial cells lining the
brain capillaries. The stabilizing methadone doses
were normally distributed with a mean and median Genome-wide association studies
dose of 160 mg/day (range 30–280 mg/day). Nine ABCB1 Genome-wide association studies (GWAS) allow
common SNPs (rs1045642, rs6949448, rs2235067, the screening of large numbers of variants in a single
rs2032583, rs2032582, rs1922242, rs1128503, rs2520464, study to identify variants and genes that may be
and rs3789243) were genotyped. A significant differ- involved in determining a particular phenotype.
ence in genotype frequencies was found between the A recent study on vulnerability to develop heroin
“higher” (> 150 mg/day) and “lower” ( 150 mg/day) addiction was conducted using the Affymetrix 10K
dose groups for the synonymous SNP 1236C>T GeneChip, which simultaneously genotyped 10 000
(rs1128503). Furthermore, individuals bearing the SNPs [75]. In that study, DNA from former severe
three-locus genotype pattern TT-TT-TT (rs1045642, heroin addicts (meeting Federal criteria for metha-
rs2032582, and rs1128503) required higher doses, done maintenance) and control subjects, all of Cauca-
while individuals heterozygous for these three SNPs sian ethnicity, was analyzed. Separate analyses were
stabilized at lower doses. done for the autosomal and X chromosomal variants.

302
Chapter 24: Human molecular genetics of opioid addiction

The strongest association of allele frequency with on vulnerability to develop heroin addiction [80–82].
heroin addiction was with the autosomal variant Two studies by the Tsuang group have found a link-
rs965972, located in a Unigene cluster of unknown age peak with point-wise significance on chromosome
function on chromosome 1q31.2. The variant with 4q31 [81, 82]. Two recent linkage studies have exam-
the strongest association by genotype frequency with ined opioid dependence using SNPs [83, 84]. Lachman
heroin addiction was in the transcription factor myo- et al. [83] found a single chromosomal region at
cardin MYOCD gene at chromosome 17p12. The two chromosome 14q in their Hispanic group “suggestive”
variants with the next strongest association by geno- of genome-wide evidence for linkage. In the other
type frequency with heroin addiction were located in study, the Gelernter group [84] identified eight
intergenic regions at chromosome 17p12 and 1q31.2, variants with point-wise significance of association
respectively. A strategy was employed in which the with opiate dependence. These studies have provided
most significant variants identified by association of evidence that a number of genes act in a complex
genotype frequency with heroin addiction were ana- cooperative and synergistic manner to influence the
lyzed together to identify common genotype patterns vulnerability and development of heroin addiction.
of unlinked alleles associated with heroin addiction. It will require additional studies to delineate the
When evaluated for genotype patterns using these role(s) of these genes and genetic variants.
three variants that had the strongest significance for
association of genotype frequency with heroin addic-
tion, one genotype pattern was found to be signifi- Summary
cantly associated with vulnerability to develop heroin
Many studies now have been conducted to address a
addiction. This pattern had an odds ratio (OR) of 6.25
possible association of specific variants of specific
and explained 27% of the population-attributable risk
genes with opioid addiction. The genes studied are
for heroin addiction. Another genotype pattern of
usually those which have been shown to be involved
these variants was found to be significantly associated
in some aspects of development of addiction. In some
with protection from developing heroin addiction
studies, multiple genes, all hypothesized to be poten-
(OR 0.13). Lacking this genotype pattern explained
tially involved in opioid addiction, have been studied.
83% of the population-attributable risk for developing
In other cases, GWAS have been performed to iden-
heroin addiction. Evaluation of 393 genes that were
tify regions of chromosomes which may influence
hypothesized or known to be involved in heroin
vulnerability to the development of addiction. These
addiction or affective disorders identified 5 genes that
studies usually do not identify specific genes within
may be involved in the development of heroin addic-
chromosome regions. Here we have reviewed our
tion. These were genes coding for: the mu-opioid recep-
recent studies and a few selected other studies which
tor; the metabotropic receptors mGluR6 and mGluR8;
have used some of these approaches.
nuclear receptor NR4A2; and cryptochrome 1
(photolyase-like).
Other association studies of drug and alcohol
addiction have found evidence of specific genes which Acknowledgements
may be involved in vulnerability to develop an addic- Funding support for this paper was received from the
tion. In an early study, the Uhl group [76] identified a National Institutes of Health (NIH) – National Insti-
number of variants that were associated with drug tute on Drug Abuse (NIDA) Research Center Grant
abuse vulnerability using a DNA pooling approach P60-DA05130; the NIH National Institute of Mental
and a 1494 variant chip. Additional chromosomes Health (NIMH) grant MH 79880; the NY State Office
regions were found using a similar pooling approach of Substance Alcohol and Substance Abuse Services
with a larger cohort and the 10K GeneChip [77]. (OASAS) grant C001839 (Kreek); and the NIH grant
Further pooling studies have used the 100K and UL1RR024143 and the National Center for Research
500K GeneChips to narrow down the significant asso- Resources (NCRR) (Coller).
ciations to 89 genes suggestive of playing a role in The authors also wish to thank Dr. Orna Levran
addiction vulnerability [78, 79]. Short tandem repeat for reading this chapter, and Kitt Lavoie and Susan
(STR) markers have also been used in linkage studies Russo for help in the preparation of the manuscript.

303
Chapter 24: Human molecular genetics of opioid addiction

References 23. Mucha RF, et al. Psycho-


pharmacology (Berl) 1985;
43. Zhu G, et al. Psychopharmacology
(Berl) 2004;177:178–184.
1. Kreek MJ, et al. Pharmacol Rev 86:274–280.
2005;57:1–26. 44. Mattay VS, et al. Proc Natl Acad
24. Zhang Y, et al. Psycho- Sci U S A 2003;100:6186–6191.
2. Savage SR, et al. Addict Sci Clin pharmacology (Berl) 2004;
Pract 2008;4:4–25. 45. Vandenbergh DJ, et al. Am J Med
172:422–429. Genet 1997;74:439–442.
3. Mestek A, et al. J Neurosci 25. Kreek MJ, et al. J Pharmacol Exp
1995;15:2396–2406. 46. Li T, et al. Am J Med Genet B
Ther 1999;288:260–269. Neuropsychiatr Genet 2004;
4. Bond C, et al. Proc Natl Acad Sci 26. Bart G, et al. J Pharmacol Exp Ther 129:120–124.
U S A 1998;95:9608–9613. 2003;306:581–587. 47. Horowitz R, et al. Am J Med Genet
5. Kroslak T, et al. J Neurochem 27. Specker S, et al. Psycho- 2000;96:599–603.
2007;103:77–87. pharmacology (Berl) 1998; 48. Tiihonen J, et al. Mol Psychiatry
6. Beyer A, et al. J Neurochem 137:326–332. 1999;4:286–289.
2004;89:553–560. 28. Simonin F, et al. Proc Natl 49. Lohoff FW, et al. Neuro-
7. Zhang Y, et al. J Biol Chem Acad Sci U S A 1995;92: psychopharmacology 2008;
2005;280:32618–32624. 7006–7010. 33:3078–3084.
8. Befort K, et al. J Biol Chem 29. Yuferov V, et al. Pharmacogenetics 50. Yacubian J, et al. Proc Natl
2001;276:3130–3137. 2004;14:793–804. Acad Sci U S A 2007;104:
30. Gerra G, et al. Am J Med 8125–8130.
9. Kreek MJ, et al. Mol Interv
2007;7:74–78. Genet B Neuropsychiatr Genet 51. Enoch MA, et al. Alcohol Clin Exp
2007;144B:771–775. Res 2006;30:399–406.
10. Kreek MJ. Clin Pharmacol Ther
2008;83:615–618. 31. Edenberg HJ, et al. Hum Mol 52. Barnes NM, et al. Neuro-
Genet 2008;17:1783–1789. pharmacology 1999;38:
11. Douet V, et al. Biochem Biophys
1083–1152.
Res Commun 2007;354:66–71. 32. Proudnikov D, et al. Neurosci Lett
2008;435:234–239. 53. Proudnikov D, et al. Pharmacogenet
12. Michelotti GA, et al. FASEB J Genomics 2006;16:25–36.
2007;21:1979–1993. 33. Schluger JH, et al. Neuro-
psychopharmacology 2001; 54. Huang YY, et al. Neuro-
13. Jaenisch R, et al. Nat Genet psychopharmacology 2003;
2003;33:245–254. 24:568–575.
28:163–169.
14. Nielsen DA, et al. Neuro- 34. Kreek MJ, et al. Neuropeptides
1984;5:277–278. 55. Sun HF, et al. Biol Psychiatry
psychopharmacology 2009; 2002;51:896–901.
34:867–873. 35. Mountjoy KG, et al. Science
1992;257:1248–1251. 56. Cigler T, et al. Am J Med Genet
15. Bleich S, et al. Alcohol Clin Exp B Neuropsychiatr Genet
Res 2006;30:587–591. 36. Tsigos C, et al. J Clin Invest 2001;105B:489–497.
16. Bönsch D, et al. J Neural Transm 1993;92:2458–2461.
57. Duan J, et al. Mol Psychiatry
2004;111:1611–1616. 37. Elias LL, et al. J Clin Endocrinol 2003;8:901–910.
17. Bönsch D, et al. J Neural Transm Metab 1999;84:2766–2770.
58. Jensen KP, et al. Mol Psychiatry
2006;113:1299–1304. 38. Slawik M, et al. J Clin Endocrinol 2009;14:381–389.
18. Bönsch D, et al. Neuroreport Metab 2004;89:3131–3137.
59. Brown GL, et al. In Sandler M.
2005;16:167–170. 39. Reisch N, et al. Eur J Endocrinol (ed.). Psychopharmacology of
19. Kumar A, et al. Neuron 2005;153:711–715. Aggression. New York: Raven
2005;48:303–314. 40. Oosterhuis BE, et al. Am J Med Press, 1979.
20. Renthal W, et al. Neuron Genet B Neuropsychiatr Genet 60. Linnoila M, et al. Life Sci 1983;
2007;56:517–529. 2008;147B:793–798. 33:2609–2614.
21. Black YD, et al. J Neurosci 41. Volkow ND, et al. Mol Psychiatry 61. Cooper JR, et al. J Pharmacol
2006;26:9656–9665. 2004;9:557–569. Exp Ther 1961;132:265–268.
22. Zhang H, et al. Mol Pharmacol 42. Lotta T, et al. Biochemistry 62. Walther DJ, et al. Science 2003;
2007;71:1319–1328. 1995;34:4202–4210. 299:76.

304
Chapter 24: Human molecular genetics of opioid addiction

63. Zill P, et al. J Psychiatr Res 2007; 71. Reuter M, et al. Am J Med 78. Johnson C, et al. Am J M
41:168–173. Genet B Neuropsychiatr Genet Genet B Neuropsychiatr Genet
64. Nielsen DA, et al. Arch Gen 2005;134B:20–24. 2006;141B:844–853.
Psychiatry 1994;51:34–38. 72. Nielsen DA, et al. Behav Genet 79. Liu QR, et al. Am J Med Genet B
2008;38:133–150. Neuropsychiatr Genet 2006;
65. Nielsen DA, et al. Arch Gen
141B:918–925.
Psychiatry 1998;55:593–602. 73. Levran O, et al. Hum Mol Genet
2008;17:219–227. 80. Gelernter J, et al. Am J Hum
66. Hsu YP, et al. Mol Psychiatry
Genet 2006;78:759–769.
1998;3:213–214. 74. Levran O, et al. Genes Brain Behav
81. Glatt SJ, et al. Am J Med Genet B
67. Chung IW, et al. Alcohol 2005; 2008;7:720–729.
Neuropsychiatr Genet 2006;141B:
36:1–3. 75. Nielsen DA, et al. Mol Psychiatry 648–652.
68. Ishiguro H, et al. J Neural Transm 2008;13:417–428. 82. Glatt SJ, et al. Drug Alcohol
1999;106:1017–1025. 76. Uhl GR, et al. Am J Hum Genet Depend 2008;98:30–34.
69. Lerman C, et al. Am J Med Genet 2001;69:1290–1300. 83. Lachman HM, et al. Hum Mol
2001;105:518–520. 77. Liu QR, et al. Proc Natl Acad Genet 2007;16:1327–1334.
70. Sullivan PF, et al. Am J Med Genet Sci U S A 2005;102: 84. Yu Y, et al. Hum Genet 2008;
2001;105:479–484. 11864–11869. 123:495–506.

305
Genetics of stimulant dependence
Chapter

25 Joseph F. Cubells and Yi-Lang Tang

Introduction States in 2007 [2]. Research on genetic mechanisms


influencing liability to dependence on psychostimu-
Central nervous system (CNS) stimulants, also called lants will hopefully allow development of better treat-
psychostimulants, are pharmacological agents that ments and preventive strategies for addressing those
temporarily enhance many aspects of physiological disturbingly high rates of psychostimulant abuse.
and psychological function. The term psychostimu-
lant encompasses a broad array of substances, includ-
ing those prescribed for medical conditions (e.g. Evidence supporting the importance
d-amphetamine and methylphenidate, used in the
treatment of attention-deficit hyperactivity disorder
of genetic inheritance in stimulant
and narcolepsy); those used primarily illicitly for the dependence
“high” they induce (e.g. cocaine and both legally and Before considering molecular genetic mechanisms, it is
illegally manufactured amphetamine-type stimulants important first to establish that genetic inheritance con-
[ATS]) and those found in over-the-counter decongest- tributes to risk for a particular phenotype. Approaches
ants (e.g. psuedoephedrine), coffee and tea (caffeine and for establishing the role of genes in human use of
theophylline), and tobacco (nicotine). Repeated admin- psychostimulants have included family, twin, and adop-
istration of many psychostimulants, including cocaine, tion studies. The most basic requirement for establish-
ATS, nicotine, and caffeine, leads to dependence, which ing the role of genes in any disorder is to demonstrate
can be broadly defined as the repeated and compulsive that the disorder is familial, or that it tends to run in
use of a substance despite adverse consequences. This families. Familiality, however, does not imply a genetic
chapter will focus mainly on genetic mechanisms mechanism, as family environment could account for
underlying liability to dependence on cocaine and traits shared among family members. Disentangling the
ATS, which are commonly used illicitly in the United role of genes from environment relies on twin studies
States and in many other parts of the world. The genet- and adoption studies. To date, the only adoption studies
ics of nicotine dependence is the subject of a separate to focus on the offspring of psychostimulant-using
chapter (Chapter 23). parents are those examining the neurodevelopment of
Both cocaine and ATS are commonly abused young children (e.g. [3]), rather than the risk of abuse or
drugs. According to the World Drug Report 2008 [1], dependence. The focus here will therefore be on family
the numbers of abusers for ATS and cocaine were and twin studies.
25.6 and 16.0 million respectively, accounting for Family studies: Family studies have analyzed
0.8% and 0.4% of the global population (age 15–64) transmission of substance abuse disorders from gen-
respectively. According to estimates published by eration to generation through families. The basic
the US Substance Abuse and Mental Health Services approach is to determine if family members of sub-
Administration (SAMHSA), there were approxi- stance abusers are at greater risk for substance abuse
mately 1.6 million users of cocaine aged 12 or older, than individuals in the general population. Family
and an additional 400 000 users of other stimulants studies consistently find elevated rates of drug abuse
(largely amphetamines and derivatives) in the United and dependence (regardless of the specific substance)

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

306
Chapter 25: Genetics of stimulant dependence

in families of drug-dependent probands. For example, to substance use disorders contributes to dependence
Luthar and Rounsaville studied first-degree relatives for stimulants and other classes of illicit drug with the
of 298 cocaine abuser probands and found that 40% exception of opiates, which appeared to associate with
of brothers and 22% of sisters of cocaine users, as well a substantial substance-specific risk. For stimulant
as 5% of parents, displayed drug abuse or dependence dependence, 33% of the risk was attributable to gen-
[4]. Similarly, data from the Collaborative Study on etic factors, and most of that genetic risk was due to
the Genetics of Alcoholism (COGA) [5] showed that vulnerability to substance dependence, whereas only
relatives of alcoholics are at substantially higher risk approximately 9% of the total variance in risk could
for cocaine dependence (10.7% of first-degree rela- be accounted for by stimulant-specific genetic effects.
tives of alcoholic probands versus 1.1% of first-degree This pattern, in which genetic factors common to all
relatives of control probands) and other stimulant substance dependence disorders accounted for most
dependence (5.8% versus 0.8% in those same classes of the genetic risk for psychostimulant dependence
of relatives). Merikangas and colleagues [6] observed was also reported by Kendler and colleagues [8]. In
that among 27 probands with a cocaine use disorder that study, which examined cocaine dependence and
as the predominant disorder, 14.9%, 18.1%, and other stimulant dependence separately in a popula-
78.3% of 94 first-degree relatives were regular users tion-based cohort of male twins from the Common-
of illicit drugs, alcohol, and tobacco, respectively. wealth of Virginia, genetic factors associated with a
Those data demonstrate clear familial clustering of general liability to substance dependence accounted
substance use among relatives of cocaine users. In for 63% or 57% of the risk for cocaine or stimulant
addition to familial clustering of all substance use dependence, respectively, while an additional term for
disorders, some evidence also supports substance- substance-specific risk did not improve the model.
specific clustering. In the Merikangas study, 7.5% of Considering only cocaine or stimulant use (but not
the first-degree relatives of cocaine users had a abuse/dependence), there appeared to be a small
cocaine use disorder. Family data thus suggest that proportion of risk that was substance-specific (7% or
substance use disorders generally, as well as cocaine 4% for cocaine or stimulants, respectively), and the
use disorders specifically, cluster in families. Family majority of the genetic risk (accounting for 49–51% of
studies, however, cannot separate the influence of the total variance in risk) was attributable to a genetic
genes from that of family environment. Thus, twin influence on general liability to substance use. How-
studies of substance dependence are critical for estab- ever, one of the limitations of that study was that
lishing the role of genes in substance use disorders. there were very few cocaine-addicted individuals
Twin studies: Twin studies compare the degree to in the total sample, so the power to detect cocaine-
which traits are shared between dizygotic (DZ) twins specific effects was quite limited. Kendler and col-
(who share on average 50% of their nuclear genes leagues examined substance use disorders in a
identical by descent) to those shared by monozygotic broader context of psychiatric disorders, examining
(MZ) twins (who share 100% of their nuclear genes samples of both male and female twins [9]; a set of
identical by descent). To the degree that MZ concord- genetic factors increasing risk for substance use dis-
ance in a given trait is greater than DZ concordance, orders explains most of the risk for all substance use
there is evidence for the effect specifically for a role of disorders in both males and females.
genes. Twin studies have consistently supported gen- In summary, data from family and twin studies
etic inheritance as an important explanation of the clearly support an important role for genes in the
risk of substance dependence in general, and psychosti- risk for substance use disorders. A general risk for
mulant dependence in particular. A study of male twins substance dependence, rather then substance-specific
who served in the US military during the Vietnam era genetic risks, appears to account for most of the
found higher concordance rates for stimulant depend- genetic risk for substance dependence. Those findings
ence (cocaine, amphetamines, and amphetamine-like make sense given clear epidemiological findings that
drugs) among MZ than DZ twins, with 26.2% and the abuse of one drug or class of drugs strongly
16.5% of each twin type respectively sharing diag- associates with abuse of other drugs [10]. As molecu-
noses of stimulant dependence [7]. Further analysis lar-genetic studies implicating specific genes and
of the same twin cohort found that the best model for genomic regions emerge, it may well be useful to
explaining the data was that a common vulnerability evaluate findings, in part, by their applicability to

307
Chapter 25: Genetics of stimulant dependence

addiction to more than one class of substance [11]. the investigators observed “suggestive” linkage signals
Molecular studies may also introduce unexpected (i.e. those in which the logarithm of the odds [LOD]
complexities into the principle that a common set scores were positive, but fell short of genome-wide
of genetic factors modifies risk for multiple substance significance) on chromosome 10 and two distinct
use disorders. Thus, a recent report [12] provides regions of chromosome 3, in European-American
convincing evidence that a particular allele of a non- (EA) families only. For CIP, they observed a genome-
synonymous single nucleotide polymorphism (SNP) wide-significant LOD score of 3.65 on chromosome 9,
in the CHNRA5 gene, encoding the a-5 subunit of the in African-American (AA) families. Their strongest
nicotinic cholinergic receptor, associates negatively results were observed for the cluster membership
with risk for cocaine dependence, whereas prior stud- traits, including a LOD score of 4.66 for membership
ies by the same investigators clearly support a positive in the “Heavy Use, Cocaine Predominant” cluster on
association of the same variants with nicotine chromosome 12 (in EAs only) and a LOD score of
dependence [13]. 3.35 for membership in the “Moderate Cocaine and
Opioid Abuse” cluster on chromosome 18 in AA
Molecular genetic studies of families only. These findings should set the stage
for fine-mapping studies to identify specific genes
psychostimulant dependence accounting for the observed linkage signals.
in humans
Linkage studies Association studies
Linkage analyses examine the co-segregation of Association studies seek to identify specific alleles that
molecular markers with phenotypic traits to identify consistently co-occur with a trait of interest within a
regions of the genome causally related to the traits. population. In case-control association analysis, such
Genetic markers at a single gene locus are said to be co-occurrence is demonstrated by showing a statistic-
genetically linked with a trait if specific molecular ally significant difference in the frequency of a given
markers (which tag specific places, or loci, in the allele in a large group of cases (who all exhibit the
genome) are consistently co-inherited within families trait) as compared to controls. As in all case-control
together with the trait of interest. Another method for designs in epidemiology, a fundamental assumption
detecting linkage is to estimate the proportion of underlying the approach is that cases and controls are
relative pairs with a given trait who share a genetic sampled from the same population. Particularly in
marker (or set of markers) identical by descent; if genetic case-control studies it is essential that com-
such sharing occurs to a degree greater than expected parison groups be from the same geographical ances-
by chance, then positive evidence exists for linkage of try. When comparison groups differ in known or
the trait to the genomic region represented by the unknown ways with regard to population background
markers. (e.g. if different ethnic groups are represented in the
To date, only one genome-wide linkage analysis of two groups), and coincidentally, the phenotype being
cocaine dependence and related traits has been pub- examined differs in frequency between the population
lished [14]. This study, of small nuclear families ascer- group(s) represented by cases versus the group(s)
tained through polysubstance-using probands who represented by controls, population stratification can
were dependent on cocaine and identified cocaine as give rise to statistically significant, but biologically
their drug of choice, is noteworthy for several reasons. meaningless, results. In family-based association
First, rather than focusing only on the presence or analysis, differential transmission of an allele from
absence of cocaine dependence, a variety of cocaine- parents to affected children identifies association in
related phenotypes were examined, including CIP, the presence of linkage. An important advantage of
which occurs in 60–80% of chronic cocaine users family-based designs is that they are robust to the
[15–17], as well as a series of traits defined by statistical effects of population stratification, as population
analysis of drug-use characteristics such as heaviness of background is inherently controlled within each
use, preferred drug, and route of administration. The family, and transmissions within families are the
clusters derived from those characteristics were measure of interest. A detailed treatment of popula-
shown to be heritable. For cocaine dependence itself, tion stratification, and strategies for addressing this

308
Chapter 25: Genetics of stimulant dependence

potential confound, is beyond the scope of this and the degree of LD to adjacent markers, for
chapter. However, Hamer and Sirota provide an approximately 4 million SNPs, in samples from sev-
entertaining and easily understood overview of the eral different populations (Japanese, European-
problem [18], while Devlin and colleagues provide a American, British, Yoruban/Nigerian, and Chinese).
more detailed discussion [19]. Using software tools such as Haploview (http://www.
It is important to note, for readers who are not broad.mit.edu/mpg/haploview/), researchers can esti-
geneticists, that genetic association implies more mate how many SNPs will be required to capture
than simple statistical association: When evidence most of the common variation present in a region
across studies consistently supports association within a population of similar ancestry to the
between a specific marker and a phenotype, we HapMap reference populations, and which SNPs to
take that as evidence of linkage disequilibrium (LD) target for genotyping. Thus, for example, if two SNPs
between the marker and a casual variant(s) nearby in exhibit extremely high LD, as indicated by the square
the genome. LD refers to the correlation between of their correlation coefficient (r2) approaching unity,
nearby markers in the genome, such that a causal it is only necessary to genotype one of them to capture
allele derived from an ancient common ancestor has the genetic variation represented by both SNPs,
been inherited through generations together with because they yield redundant information.
other nearby (noncausal) alleles that are close Numerous published studies have attempted to
enough to the causal variant that meiotic recombin- evaluate association between molecular polymorphic
ation between them has occurred only infrequently. markers (usually SNPs) and stimulant abuse or
Thus, even over dozens or hundreds of generations dependence. Until recently, such association studies
between a mutation event and inheritance of the have almost always been guided by candidate-gene
disease allele by a given case subject, the alleles of hypotheses generated from knowledge about the
polymorphisms near the casual allele are likely to be neurobiology of cocaine dependence, with a particu-
derived from the same ancestral chromosome, and lar focus on the role of dopamine as an important
therefore “tag” the allele exerting a causative effect on mediator of the reinforcing properties of cocaine and
the phenotype. In the presence of such LD, associ- other substances of abuse. Candidate-gene studies of
ation between an allele at a particular marker and a cocaine dependence have therefore been dominated by
phenotype cannot be taken as evidence supporting a studies of loci encoding key proteins mediating the
direct biological role of the allele in the phenotype, synaptic actions of dopamine (DA). Attention has also
without other lines of biological evidence to support been paid to some genes encoding proteins involved
such a conclusion. in serotonin (5HT)- or norepinephrine (NE) -mediated
A key difference between LD and linkage is that neurotransmission, based on the fact that cocaine is a
only a few recombination events are likely to occur in potent inhibitor of the uptake of those monoamines as
a given chromosome across the two to four gener- well. Finally, a variety of neuropeptides are co-localized
ations typically examined in linkage studies. Evidence with, or play a role in regulating, DA, NE, and 5HT, so
of linkage to a given marker therefore could be due to some studies have examined genes relevant to neuro-
a causal allele quite far (up to many megabases) from peptides and neurotrophic factors.
the linked marker. In contrast, LD between a marker As in many genetic-association studies of other
and a phenotype typically implicates a much smaller complex disorders, many initial “positive” findings
candidate region, usually within  100 000 bp in regarding genetic associations to cocaine dependence
European populations, although this average “LD have not been replicated in subsequent studies.
distance” varies widely across the genome, and differs Although the reasons vary and are beyond the scope
in populations of differing geographical ancestry. of this chapter, one very important factor is that
A key step in evaluating most putative genetic associ- until very recently, almost all association studies of
ations is therefore to evaluate the “LD structure”, or cocaine-dependence suffered from extremely low
pattern of marker-to-marker correlation, surrounding statistical power, with early studies often having
the implicated allele. That process, of evaluating LD fewer than 100 cocaine-dependent or control subjects
structure of a candidate gene or genomic region has of a single ethnic group per comparison group, and
been greatly facilitated by the HapMap project (http:// many studies examined only a single variant (usually
www.hapmap.org/), which reports allele frequencies, a SNP). In such studies, even in the context of theoretical

309
Chapter 25: Genetics of stimulant dependence

considerations that might support the hypothesis anxiety. Given phenomenology similarities between
that a given polymorphism associates with cocaine anxiety and paranoia, it is possible that associations
dependence, the a priori chance of a negative result between genetic variation and the two phenotypes
is so high that levels of nominal statistical significance represent the same biological pathway. Additional
that might be indicative of true positive findings in evidence regarding association between variation at
simpler settings are likely to be due to chance. The SLC6A3 and human responses to psychostimulants
selected results reviewed below therefore must be derives indirectly from genetic studies of attention-
viewed as preliminary and in need of replication. deficit hyperactivity disorder (ADHD). Linkage and
It seems fair to say at present that no locus has yet association implicate SLC6A3 in risk for ADHD in
been identified that can be said to show a confirmed some (but not all) studies [27]. Psychostimulants such
genetic association with cocaine dependence or as methylphenidate and d-amphetamine are front-
cocaine-associated behavioral phenotypes. line treatments for ADHD symptoms, and several
studies have suggested that variation at SLC6A3 asso-
ciates with altered therapeutic response to psychosti-
Studies of dopamine-related genes mulants in ADHD [reviewed in 28]. In summary,
Dopamine transporter gene (SLC6A3): The DA there is no evidence to support a direct association
transporter protein, encoded by the locus SLC6A3, between variation at SLC6A3 and cocaine or amphet-
modulates DAergic tone by transporting DA into amine dependence, but some evidence supports the
the presynaptic neuron following synaptic release, hypothesis that variation at that locus associates with
thereby terminating its action. Cocaine’s primary subjective response to stimulants.
addiction-related pharmacological action is thought DRD2: Because DA receptors mediate many of
to be inhibition of DA by the DAT. At least three lines the rewarding effects of stimulants, genetic alterations
of evidence suggest that variation at SLC6A3 influ- influencing dopamine receptors are obvious candi-
ences human responses to cocaine: (1) several poly- date genes for cocaine and amphetamine dependence.
morphisms identified in SLC6A3, such as T265C Many early genetic association studies of substance-
(Val55Ala) and T1246C (Val382Ala), are associated use-related phenotypes focused on DRD2, the locus
with deficient DA uptake in in vitro studies [20]; encoding the dopamine D2 receptor, with a particular
(2) the SNP, G2319A, and a variable-number of focus on a SNP of no known function, located
tandem repeats (VNTR) polymorphism found in the approximately 10 000 bp 3’ to the DRD2 coding
30 -untranslated region of exon 15 of the SLC6A3 gene region. This variant, known as TaqI A is a SNP that
have been reported to associate with differences was typed as a restriction fragment length poly-
in availability of DA transporter binding sites as esti- morphism (RFLP), and was named for the restriction
mated by single proton emission tomography enzyme, Taq1, used for such assays. Claims of associ-
imaging (SPECT) [21]; and (3) the 9-repeat allele of ation of this SNP have been made for polysubstance
the foregoing VNTR, which associated in the SPECT abuse [29], age of onset of substance abuse [30], and
study with higher ligand-binding potential (presum- psychostimulant [31] and cocaine dependence [32].
ably reflecting higher levels of available neuronal DA The RFLP TaqI B, another SNP of no clear function,
transporter molecules) was reported to associate with has also been claimed to associate with polysubstance
CIP [22, 23]. Some, but not all, subsequent studies abuse in Caucasians [29] and with cocaine depend-
have supported those conclusions [23]. Thus, the ence [33]. All of those studies were very underpow-
9-repeat allele associated with methamphetamine- ered, and therefore must be viewed with extreme
induced psychosis in a Japanese sample [24] but not caution. More recently, as more detailed knowledge
in Chinese cocaine users [25]. A human-laboratory has emerged on the genomic region surrounding the
study suggested that individuals homozygous for Taq1 A polymorphism, evidence suggests that the
the 9-repeat allele showed diminished subjective variant may actually be more relevant to loci other
responsiveness to oral d-amphetamine, as compared than DRD2. Most notably, that Taq1 A SNP is actually
to individuals homozygous for the 10-repeat allele, a nonsynonymous variant in the coding region of
or heterozygotes [26]. Although that study did not ANKK1, encoding the ankrin repeat and kinase
specifically assess amphetamine-induced paranoia, it domain-containing protein 1. Gelernter and col-
did show an association to amphetamine-induced leagues have conducted several association studies

310
Chapter 25: Genetics of stimulant dependence

of the Taq1 A SNP, along with several dozen addi- acid position 158 (in the long, membrane-bound
tional SNPs, to generate a detailed picture of linkage form of the protein) or 108 (in the shorter, soluble
and association of that genomic region to substance protein). This SNP has been intensively studied as
dependence phenotypes including nicotine depend- it has been shown to alter the thermal stability of
ence and alcohol dependence [34, 35]. Thus, the evi- the enzyme, such that the methionine-containing
dence for an association to cocaine dependence of allelomorph is relatively unstable at physiological
variants in DRD2 remains scant, despite early claims temperature, leading to lower enzyme activity [44].
to the contrary. Lower COMT enzyme activity could lead to slower
DRD3: Dopamine D3 receptor (D3R) has been rates of DA and NE degradation, thus altering synap-
reported as a mediator in cocaine use. Some studies tic activity of these catecholamines. Lohoff and col-
in humans also suggest a potential role of the D3R in leagues [45] recently reported an association study of
cocaine dependence. For example, the D3R is present cocaine dependence examining the val-met SNP,
in high density in the nucleus accumbens; the prime together with two additional SNPs at COMT, located
site of the dopamine reward pathway. Its density in at opposite ends of the coding region. The study
this area is two-fold higher in victims of accidental examined African-American subjects who were either
cocaine overdose [36, 37]; of note, such decedents dependent on cocaine or free of addictive disorders.
often exhibit psychotic excited delirium immediately They found evidence for an association between
prior to death. These observations suggest that cocaine the val-met SNP and cocaine dependence, with the
use leads to a neuroadaptive elevation in D3R density low-activity associated met allele occurring more
in response to increased synaptic dopamine and that frequently in dependent individuals than in controls
the D3R may be one mediator of both the reinforcing (35% versus 29%, respectively; Bonferroni-corrected
and psychosis-promoting effects of cocaine [38, 39]. p ¼ 0.014). A study of the val-met SNP in Chinese
Several early studies tested for associations methamphetamine abusers versus healthy controls
between the DRD3 SNP rs6280 (a nonsynonymous found the opposite direction of association, with the
SNP also known as ser9gly because it encodes either val allele occurring more frequently in abusers than
serine or glycine at amino acid position 9) and controls [46]. A study of methamphetamine abuse in
cocaine dependence, either by itself [40, 41] or in Japan found no association to the val-met SNP [47].
the context of comorbid schizophrenia [42]. None of The reasons for the apparent discrepancies among
those studies reached definitive conclusions, all these three studies are not entirely clear, but notable
suffered from extremely low statistical power (fewer differences among the studies include the ethnicities
than 100 cocaine-dependent subjects of a single ethnic of the comparison groups and the diagnoses exam-
group per study), and each study examined only a ined (abuse of methamphetamine versus dependence
single SNP (albeit one of clear biological function, as on cocaine). The evidence supporting a role for gen-
it alters the predicted amino acid sequence of the etic variation at COMT in psychostimulant depend-
protein). A more recent larger study [43] of rs6280 ence requires clarification and replication.
(N ¼ 730 hospitalized cocaine addicts; 781 healthy Dopamine b-hydroxylase (DbH): Plasma DbH
blood donors who denied drug use) found no associ- activity is a genetic trait controlled by only a few genes
ation between genotypes or alleles at the locus and [48, 49]. Most of the variance in plasma DbH activity
cocaine dependence. However, that study can be criti- is explained by variation in plasma levels of DbH
cized because its comparison groups were ethnically protein, as shown by strong correlations (r2  0.80)
mixed, with different proportions of self-identified between plasma DbH activity and plasma levels of
racial groups, and the analysis did not control for that DbH immunoreactive protein. DbH activity is a
potential source of stratification. Overall, however, the highly heritable trait. Abundant evidence indicates
evidence supporting a clear association between vari- that the structural gene encoding the protein, DbH,
ation at DRD3 and cocaine dependence is quite weak. regulates much of the genetic variation in the trait
COMT: COMT encodes catechol-O-methyltrans- (reviewed in [16]). Prior work has shown several
ferase, a major enzyme of DA and NE catabolism. polymorphisms at DBH to be two polymorphisms,
A nonsynonymous SNP at COMT, rs4680, is often an insertion/deletion polymorphism (DBH*5’-ins/del
referred to as the “val-met” SNP, owing to the fact [a biallelic 19-nucleotide insertion/ deletion approxi-
that it encodes either a valine or methionine at amino mately 4.7 kb 5’ upstream from the transcription

311
Chapter 25: Genetics of stimulant dependence

initiation site]) and a SNP (DBH*444g/ a, rs1108580), Cannabinoid-related genes: A trinucleotide repeat
respectively, are in strong LD and show similar asso- polymorphism in the 30 flanking region of the canna-
ciation with plasma levels of DbH. A haplotype of binoid receptor 1 (CNR1) gene is associated with
these two polymorphisms not only correlated with intravenous drug abuse (cocaine, amphetamine, or
low DbH levels but also was associated with CIP in heroin) [60]. A study of 22 polymorphisms in CNR1
a sample of American-Caucasians who abuse cocaine identified a haplotype in an intronic 50 region of the
[50]. Interestingly, one polymorphism, –1021C/T gene that is associated with substance (cocaine, opiate,
(rs1611115), at the promoter region is found to alcohol, or other drug) abuse [61]. Recently, Zuo and
be associated with low plasma DbH activity in colleagues [62] reported an association between a
European-Americans, African-Americans, and Japanese two-SNP haplotype at CNR1 and cocaine dependence,
[51, 52]. This functional polymorphism may also have originally observed in a European-American family
pharmacogenomic significance since disulfiram, which sample, and then replicated in a European-American
inhibits DbH, could be more effective in individuals case-control sample. Fatty amide acid hydrolase,
with this SNP [53], and subjects homozygous for the encoded by the FAAH gene, is an enzyme that metab-
“very low activity” T allele at –1021C–>T show an olizes endogenous ligands of the cannabinoid recep-
increased propensity to paranoia over time during tors. A functional SNP that alters the sensitivity of the
cocaine self-administration [54]. enzyme to protease in vitro is associated with drug
and alcohol abuse [63]. Taken together, the evidence
supporting a role for variation in genes related to
Other candidate genes cannabinoid-mediated neurotransmission in cocaine
Prodynorphin gene: Pharmacological and clinical dependence appears as strong or stronger than for
studies have implicated the dynorphin peptides and any other locus.
the kappa-opioid receptor are important in the
rewarding properties of cocaine, heroin, and other
drugs of abuse. Prodynorphin is an opioid peptide
precursor molecule which has been shown to be Genome-wide association studies
associated with cocaine, opiate, and other drug abuse A genome-wide association study (GWAS), also
in animals. A repeat polymorphism in the promoter known as whole genome association study (WGA
region of the preprodynorphin gene may cause study), is defined as an association study that surveys
changes in cellular functioning with importance for most of the genome for causal genetic variants, with-
cocaine addiction. The polymorphism is a 68-nucle- out the requirement of prior pathophysiological
otide tandem repeat element in the putative promoter knowledge about the disease or any prediction of
region of the gene. This polymorphism, which con- candidate genes. As in candidate gene association
tains a putative AP-1 transcription complex (c-Fos/ studies, GWAS also requires two groups of partici-
c-Jun) binding site, is found in one to four copies. pants: people with the disease and similar people
Some have hypothesized that alleles of this gene may without the disease. With the advancement of
cause an increase in dynorphin levels and, therefore, molecular genetic technologies, systematic associ-
may be protective against cocaine addiction [55]. Sev- ation study of the whole genome can now be con-
eral studies have examined an association of this ducted using DNA chips or arrays. After obtaining
polymorphism with drug dependence with conflicting samples from each participant, the set of markers
results. One study showed that Hispanic individuals such as SNPs are scanned and data are analyzed. If
with three or four copies of the repeat have a lower genetic variations are more frequent in people with
risk for development of cocaine dependence [56]. the disease, the variations are said to be “associated”
However, two subsequent studies using more strin- with the disease. The associated genetic variations
gent diagnostic criteria showed increased risk for are then considered pointers to the region of the
cocaine dependence and cocaine/alcohol co-depend- human genome where the disease-causing problem
ence in African-Americans with three or four repeats resides. Since the entire genome is analyzed for
[57, 58]. More recently, Yuferov et al. [59] reported a the genetic associations of a particular disease, this
functional haplotype to be associated with reduced technique allows the genetics of a disease to be inves-
PDYN expression in human brain. tigated in a nonhypothesis-driven manner. When a

312
Chapter 25: Genetics of stimulant dependence

GWAS detects association between a SNP and the transmission makes the possible association a very
disease, this signal usually represents association interesting new lead in understanding cocaine
with a set of several highly correlated SNPs in dependence.
strong LD. At the time of this writing, there has been one
A family-based low-density SNP scan, in which GWAS published on methamphetamine dependence
families ascertained for sibling pairs affected by [66]. Two independent case-control comparisons were
either cocaine dependence or opioid dependence evaluated by pooled genotyping up to  500 000 SNPs
[64], was recently reported. That study included across the genome. The first compared 140 metham-
339 African-American families and 334 European- phetamine-dependent individuals of Han Chinese
American families, and examined a total of 1699 ethnicity to 240 healthy control Han Chinese individ-
individuals. A variety of phenotypes were investi- uals. The second set compared 100 methampheta-
gated, including cocaine dependence (with a total of mine-dependent and 100 healthy control individuals
1284 individuals affected) and CIP (873 affected). from Japan. The investigators used a pooled genotyp-
The investigators made the a priori decision to ing strategy, in which microarray-based genotyping
focus on polymorphisms that yielded nominally methods are applied to pooled samples of DNA from
significant results (i.e. p < 0.05, uncorrected for 20 subjects per sample. The relative signal intensities
multiple testing) in both African-American and representing the respective alleles within each pooled
European-American families. Several interesting sample are, in theory, proportional to the frequency
findings emerged. A SNP in the MANEA locus, of the alleles. The results of the analysis provided
encoding a-endomannosidase, associated with CIP interesting suggestive evidence for SNPs differing
(p ¼ 0.00005). Although the finding did not with- in allele frequency between cases and controls. The
stand strict correction for multiple testing (which investigators focused on differences that occurred in
required, in this study, p < 0.00001), such a correc- both the Taiwanese and Japanese comparisons. In an
tion is arguably overly conservative, due to the cor- intriguing strategy for statistical analysis, the investi-
relation among at least some of the SNPs examined. gators evaluated not just the degree of allelic signal
A follow-up study examined additional SNPs across difference between cases and controls, but also the
the MANEA locus, first in the original family clustering of differences along the genome. Clusters
samples, and then in an independently ascertained of multiple SNPs showing differential signals between
case-control sample. Several SNPs showed nominally cases and controls within specific chromosomal
significant association to CIP in the denser SNP regions were shown by Monte-Carlo simulation to
analysis. In the replication sample, several of these be unlikely to be due to chance. The analysis impli-
also showed significant association to CIP in African- cated a number of interesting candidate genes as
Americans, and an additional SNP showed asso- associated to methamphetamine dependence, includ-
ciation in European-Americans [65]. While not ing genes encoding several enzymes likely to regulate
definitive, those studies suggest that MANEA and its intracellular signaling (e.g. PDE6C, encoding a cyclic-
product, a-endomannosidase, a lysosomal enzyme GMP phosphodiesterase and PRKG1, encoding a
deficiency which results in rare neuronal storage cGMP-dependent protein kinase), as well as a variety
diseases, merit further investigation with regard to of transcription factors and cell-adhesion molecules.
CIP, even though no obvious mechanism suggests The results of the study require replication in larger
itself that would explain why variation in MANEA independent samples, as the sample sizes used in
would alter risk for CIP. the study were small. In addition, the pooled geno-
In the original family-based study by Yu et al. typing strategy used is subject to errors in estimating
[64], a SNP within the 3’ untranslated region of the relative “allele frequencies” because even small errors
STY13 locus was nominally associated with cocaine in relative concentrations of DNA from individual
dependence. The gene encodes synaptotagmin VIII, subjects can distort signal intensities. Nevertheless,
a protein involved in regulation of synaptic neuro- the study represents an important step in association
transmission by calcium. While the level of statistical analysis of stimulant dependence. It is interesting
significance of this association to cocaine dependence to note that none of the loci implicated by the study
(p ¼ 0.003 in the combined samples) was modest, were “obvious” candidate genes such as those dis-
the fact that the gene product regulates synaptic cussed above.

313
Chapter 25: Genetics of stimulant dependence

Summary and conclusions dependence comparable in genotyping density to


those performed on other complex disorders has yet
In this chapter, we have reviewed evidence supporting
been performed. A GWAS of methamphetamine
the hypothesis that genetic inheritance plays a sub-
dependence, while yielding some interesting leads,
stantial role in dependence on cocaine and (to a less
requires replication in light of its small size, and reli-
well-studied degree) other illicit psychostimulants.
ance on pooled genotyping. While several intriguing
The role of genes in cocaine dependence, however,
candidate-gene associations between specific loci and
may largely reflect a more general liability to develop
cocaine dependence have been reported, to date there
dependence on a variety of substances. Studies of
has yet to be a definitively replicated result reported.
molecular genetic mechanisms in cocaine dependence
Clearly, more work is required in the human genetics of
remain in an early stage of development. Only one
stimulant dependence, to identify and characterize how
genome-wide linkage study of cocaine dependence
specific genes influence risk for this set of disorders.
has been reported to date, and no GWAS of cocaine

References 15. Brady KT, et al. J Clin Psychiatry


1991;52:509–512.
32. Young RM, et al. Addict Behav
2004;29:1275–1294.
1. United Nations Office for Drug
Control and Crime. World Drug 16. Cubells JF, et al. Drug Alcohol 33. Noble EP, et al. Drug Alcohol
Report 2008. Vienna: United Depend 2005;80:23–33. Depend 1993:33:271–285.
Nations; 2008. 17. Satel SL, et al. Am J Psychiatry 34. Gelernter J, et al. Hum Mol Genet
2. Substance Abuse and Mental 1991;148:495–498. 2006;15:3498–3507.
Health Services Administration 18. Hamer D, et al. Mol Psychiatry 35. Yang BZ, et al. Hum Mol Genet
(SAMHSA). Results from the 2007 2000;5:11–13. 2007;16:2844–2853.
National Survey on Drug Use and 19. Devlin B, et al. Genet Epidemiol 36. Segal DM, et al. Brain Res Mol
Health: National Findings. 2001;21:273–284. Brain Res 1997;45:335–339.
Rockville, MD: SAMHSA; 2008.
20. Lin Z, et al. Pharmacogenomics 37. Staley JK, et al. J Neurosci
3. Nulman I, et al. Clin Invest Med 2003;3:159–168. 1996;16:6100–6106.
2001;24:129–137.
21. Jacobsen LK, et al. Am J Psychiatry 38. Caine SB, et al. Science
4. Luthar SS, et al. Int J Addict 2000;157:1700–1703. 1993;260:1814–1816.
1993;28:415–434.
22. Gelernter J, et al. Neuro- 39. Rodriguez De Fonseca F, et al. Eur
5. Nurnberger JI Jr., et al. Arch Gen psychopharmacology 1994; J Pharmacol 1995;274:47–55.
Psychiatry 2004;61:1246–1256. 11:195–200. 40. Comings DE, et al. Mol Psychiatry
6. Merikangas KR, et al. Arch Gen 23. Haile CN, et al. Behav Genet 1999;4:484–487.
Psychiatry 1998;55:973–979. 2007;37:119–145. 41. Freimer M, et al. Addict Biol
7. Tsuang MT, et al. Am J Med Genet 24. Ujike H, et al. Pharmacogenomics 1996;1:281–287.
1996;67:473–477. 2003;3:242–247. 42. Krebs MO, et al. Mol Psychiatry
8. Kendler KS, et al. Am J Psychiatry 25. Hong CJ, et al. J Neural Transm 1998;3:337–341.
2003;160:687–695. 2003;110:345–351. 43. Messas G, et al. Psychiatr Genet
9. Kendler KS, et al. Arch Gen 26. Lott DC, et al. Neuro- 2005;15:171–174.
Psychiatry 2003;60:929–937. psychopharmacology 2005; 44. Lachman HM, et al. Pharma-
10. Kessler RC, et al. Arch Gen 30:602–609. cogenetics 1996;6:243–250.
Psychiatry 1997;54:313–321. 27. Friedel S, et al. Mol Psychiatry 45. Lohoff FW, et al. Neuro-
11. Li MD, et al. Nat Rev Genet 2007;12:923–933. psychopharmacology 2008;
2009;10:225–231. 28. McGough JJ. Biol Psychiatry 33:3078–3084.
12. Grucza RA, et al. Biol Psychiatry 2005;57:1367–1373. 46. Li T, et al. Am J Med Genet B
2008;64:922–929. 29. O’Hara BF, et al. Hum Hered Neuropsychiatr Genet 2004;
13. Bierut LJ, et al. Am J Psychiatry 1993;43:209–218. 129B:120–124.
2008;165:1163–1171. 30. Comings DE, et al. Drug Alcohol 47. Suzuki A, et al. Psychiatr Genet
14. Gelernter J, et al. Am J Med Depend 1994;34:175–180. 2006;16:133–138.
Genet B Neuropsychiatr Genet 31. Persico AM, et al. Biol Psychiatry 48. Oxenstierna G, et al. J Psychiatr
2005;136B:45–52. 1996;40:776–784. Res 1986;20:19–29.

314
Chapter 25: Genetics of stimulant dependence

49. Weinshilboum RM. Hum 55. LaForge KS, et al. Eur J Pharmacol 61. Zhang PW, et al. Mol Psychiatry
Genet Suppl, 1978;101–112. 2000;410:249–268. 2004;9:916–931.
50. Cubells JF, et al. Mol Psychiatry 56. Chen AC, et al. Am J Med Genet 62. Zuo L, et al. Neuropsycho-
2000;5:56–63. 2002;114:429–435. pharmacology 2009;34:
51. Tang YL, et al. Eur J Hum Genet 57. Dahl JP, et al. Am J Med Genet B 1504–1513.
2007;15:878–883. Neuropsychiatr Genet 2005; 63. Sipe JC, et al. Proc Natl Acad Sci
52. Zabetian CP, et al. Am J Hum 139B:106–108. U S A 2002;99:8394–8399.
Genet 2001;68:515–522. 58. Williams TJ, et al. Addict Biol 64. Yu Y, et al. Hum Genet 2008;
53. Carroll KM, et al. Arch Gen 2007;12:496–502. 123:495–506.
Psychiatry 2004;61: 59. Yuferov V, et al. Neuropsycho- 65. Farrer LA, et al. Arch Gen
264–272. pharmacology 2008;34:1185–1197. Psychiatry 2009;66:267–274.
54. Kalayasiri R, et al. Biol Psychiatry 60. Comings DE, et al. Mol Psychiatry 66. Uhl GR, et al. Arch Gen Psychiatry
2007;61:1310–1313. 1997;2:161–168. 2008;65:345–355.

315
Genetics of personality disorders
Chapter

26 C. Robert Cloninger

Personality disorders provide an excellent illustration IV [5], but taxonomic analyses have long shown that
of the challenges facing contemporary psychiatric such categorical descriptions are inadequate [6].
genetics. In several ways, what we know about the There is no rigorous evidence for discrete boundaries
genetics of personality disorders actually allows us to between categories or even clusters [4], and most
distinguish approaches to psychiatric genetics that patients with any personality disorder satisfy criteria
will be fruitful more readily than do studies of clinical for two or more putative categories [7, 8].
syndromes like schizophrenia or mood disorders. Individual differences in personality can be meas-
I say that because it was recognized that personality ured well in terms of multiple quantitative dimen-
disorders are complex quantitative traits earlier than sions of personality [2, 8]. One of the most robust
for other psychiatric syndromes [1, 2]. Several lines of findings about personality, but one that is surprising
research document the inadequacy of phenotypic to many psychiatrists, is that the same dimensions of
assessment based on categories or linear dimensions personality are observed whether one begins with
of pathological symptoms for understanding the gen- normal personality variation in the general commu-
etics, development, and evolution of personality dis- nity, with abnormal personality traits, or with symp-
orders. Recognition of the functional complexity of toms of personality disorders in treatment samples.
personality disorders may indicate solutions that will There are three main lines of evidence indicating
be fruitful for psychiatric genetics in general. that normal and abnormal personality traits and
First I will briefly outline the basic observations symptoms all rest on the same foundation. First,
that must be faced in order to understand the genetics people with personality disorders or other forms of
of personality disorders. Then I will suggest a general psychopathology have extreme values on one or
functional approach based on the evolution of human more personality dimensions, but the dimensional
brain functions that may be of general utility [3]. In the structure is the same in samples from the general
past, psychiatric geneticists have often based their ana- community and from psychiatric treatment facilities
lyses on unrealistic linear assumptions about pheno- [7, 9, 10]. Second, the genetic structure of normal
typic traits, thereby failing to describe or appreciate the and abnormal personality traits in studies of twins is
nonlinear dynamics of the evolution and development indistinguishable, suggesting that influences on
of human brain systems [4]. More realistic approaches normal and abnormal personality act through
are now possible, however, and can be informed by systems common to both, whether the twins are
combining an understanding of the genetics of person- reared together [10] or apart [11]. Third, variability
ality with the evolution of the brain systems that in normal personality traits that are descriptive of
underlie healthy personality and its disorders. the general population are diagnostic of personality
disorders [7, 12]. Therefore, normal personality traits
Structure of abnormal personality is the share a common functional foundation with abnor-
mal personality traits and most, if not all, mental
same as that of normal personality disorders. The same biopsychosocial systems influ-
Traditionally personality disorders have been ence individual differences in normal personality and
described as a set of discrete categories, as in DSM- its disorders.

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

316
Chapter 26: Genetics of personality disorders

Healthy personality is more than the extensive gene–gene and gene–environment inter-
action [19]. Nonlinear dynamic systems tend to main-
absence of personality disorder tain intermediate adaptive optima so that extreme
Health is a state of physical, mental, social, and spirit- high or low values on traits are disadvantageous
ual well-being, rather than merely the absence of [15]. The liability to illness is proportional to the
disease according to standard definitions [13]. Mental inverse of the prevalence for semi-quantitative traits
health is a state of well-being characterized primarily (that is, those that can be ranked by severity: rare
in terms of a well-integrated personality [14]. A well- traits have greater familial loading than common
integrated personality is characterized by a person traits) [20]. More information is available about com-
being self-directed, cooperative, and self-transcendent plex traits when they are measured quantitatively than
[15, 16]. Clearly the absence of particular abnormal when they are reduced to categorical disorders. In
traits, such as anxiousness or callousness, does not other words, categorical, semi-quantitative, and quan-
guarantee that personality is well-integrated. titative measurements provide increasing amounts of
Consequently, a narrow focus on the genetics of information about the underlying causes of personal-
abnormal traits or personality disorders is likely to ity and its disorders. These theoretical expectations
ignore the fact that evolution always operates on the can be and have been tested with available informa-
adaptive functioning of whole individuals in popula- tion about personality and its disorders.
tions, not on particular structures or traits that differ
between individuals. To understand the genetics of
personality it is essential to characterize the dynamics
Evidence of quantitative multifactorial
of adaptive functioning within an individual, which variability
makes the person more or less reproductively fit [3]. The personality disorder that has been most exten-
sively studied is antisocial personality. Family, twin,

Well-functioning of personality
and adoption studies have been carried out showing
the moderate heritability of antisocial personality dis-
depends on integration of all aspects orders. A wide variety of diagnostic criteria have been
of health used, but regardless of the criteria heritability has
been consistently moderate. Adoption studies of
It is surprising to many specialty trained physicians psychopathy [21], antisocial personality [22, 23], and
that health depends on the integration of all aspects of criminality [24, 25] were carried out in the United
a human being – sexual, corporeal, emotional, intel- States and Scandinavia, all showing substantial herit-
lectual, and spiritual. Yet it is well established that able influences. Interactions between genetic predis-
there is strong interdependence among physical position and childhood rearing in an unstable or
health, personality disorder, emotional distress, social hostile environment were also demonstrated for petty
dysfunction, and spiritual dysfunction [14, 15, 17]. criminality and antisocial personality. Monozygotic
Common genetic factors account for 45–60% of the (MZ) twins were also more often concordant for
associations among subjective well-being, perceived criminality than were dizygotic (DZ) twins, as
health, and somatic illness [18]. Therefore, the inte- reviewed in detail elsewhere [26]. A recent meta-
grated functioning of the whole individual needs be analysis of 51 twin and adoption studies estimated
considered to understand heritable functions, rather that there were moderate proportions of variance
than focusing on particular traits or disorders [3]. due to additive genetic influences (0.32), nonadditive
genetic influences (0.09), shared environmental influ-
Implications for research on the ences (0.16), and nonshared environmental influences
(0.43) [27]. Diagnostic criteria, assessment method,
genetics of personality disorder zygosity determination method, and age, but not
The robust observation that personality is a complex gender, were significant moderators of the magnitude
biopsychosocial matrix of functions of whole persons of genetic and environmental influences on antisocial
in populations has strong implications for how to behavior.
approach the study of personality and its disorders. However, when consistent diagnostic criteria and
Complex or quantitative traits always involve ascertainment methods are applied, gender differences

317
Chapter 26: Genetics of personality disorders

Table 26.1 A study of categorical personality disorders assessed by Structured Clinical Interview for DSM-IV-TR Axis II Personality
Disorders (SCID-II) interviews in 92 monozygotic (MZ) twin pairs and 129 dizygotic (DZ) twin pairs in Norway.

Diagnoses MZ correlation 100 (r  SE) DZ correlation 100 (r  SE) Heritability (% SE)
Any personality disorder 58  10 36  10 44  20
Any Cluster A 37  14 9  11 37  25
Any Cluster B 60  11 31  12 59  23
Any Cluster C 61  9 23  11 59  20
Adapted from Torgersen et al. [29].

have a substantial impact in both twin and family explained without taking genetic variability into
studies. The concordances for criminality and anti- account. Estimates of heritability for the clusters and
social personality were well predicted by differences specific categories were moderate (i.e. between 40 and
in severity of liability by a function proportional to 60%), much as observed for quantitative measures of
the inverse of the prevalence in the general population, normal [30, 31] and abnormal personality traits [32].
indicating that susceptibility was to an underlying The correlations between DZ pairs were usually less
quantitative variable-like personality [20, 26, 28]. For than half of those of MZ pairs, suggesting that gene–
example, antisocial personality disorder is more rare in gene and gene–environment interactions are import-
women than in men, and antisocial women carry a ant for categorical diagnoses as they are for personality
greater genetic loading (measured by more antisocial dimensions in twins reared apart [33–36]. The small
relatives) than do antisocial men. number of individuals with specific diagnoses made
Such findings about the quantitative nature of any conclusions about the heritability of the categor-
variation in personality traits provides an important ies imprecise, as shown by the large standard errors
justification for analysis of the number of symptoms in Table 26.1. Nevertheless, the consistency of the
of personality disorders, rather than focusing on cat- overall results with those of studies of personality
egorical classifications based on a more or less arbi- dimensions support the moderate heritability of per-
trary number as a diagnostic threshold. As we will see, sonality disorders.
studies using only categorical cut-offs have much less Even stronger evidence for the moderate heritabil-
information and also fail to consider the full range of ity of personality disorders has been provided by
variation in the study population. more recent studies of the inheritance of the number
of symptoms in each of 10 categories of personality
Inheritance of personality disorder disorders. The sample included 1386 Norwegian twin
pairs aged 19–35 as assessed by SIDP-IV interviews.
symptoms and categories The total heritability and the heritability unique to
Some studies have been carried out using categorical symptoms of each disorder within the three DSM-IV
diagnoses despite the theoretical and practical disad- clusters are shown in Table 26.2. The data are com-
vantages of categorical approaches. Svenn Torgersen piled from separate reports for the number of symp-
and his colleagues [29] carried out a twin study based toms of disorders in personality cluster A [37], cluster
on a wide range of categorical diagnoses in Norway. B [38], and cluster C [39]. The total heritability of
They ascertained 92 MZ and 129 DZ twin pairs in liability to the number of symptoms in each category
which at least one proband had a diagnosis of person- was weak to moderate, ranging from 28% for para-
ality disorder, and divided the cases according to noid personality symptoms to 38% for antisocial per-
DSM clusters and categories. The probandwise con- sonality symptoms.
cordances for any definite personality disorder were The unique heritability in Table 26.2 indicates
40% for MZ pairs and 29% for DZ pairs, indicating variability for specific disorders that are not shared
substantial genetic influences (p < 0.01), as summar- by other disorders in the same cluster. For example,
ized in Table 26.1. Concordance for membership in obsessive–compulsive personality disorder has 85%
personality disorder clusters also could not be unique heritability, indicating that its genetic

318
Chapter 26: Genetics of personality disorders

Table 26.2 Heritability of the number of symptoms in determinants are largely distinct from those of
personality categories in a sample of 1386 Norwegian twin pairs
aged 19–35 years as assessed by Structured Interview for DSM-IV
avoidant and dependent personality disorders. This
Personality (SIDP-IV) interviews. Total heritability and the supports the recommendation that obsessional traits
heritability unique to each disorder within the three clusters constitute a fourth cluster with high persistence or
are shown.
anankastic traits [7]. Likewise unique genetic deter-
Clusters of Total Unique minants predominate for dependent personality
personality disorder heritability heritability (52%) in cluster C and for schizoid personality
symptoms (%) (%) (74%) and paranoid personality (54%) in cluster A.
Cluster A (“odd”)
Cluster B is relatively more homogeneous but even
Schizotypal 26 0 there antisocial and borderline personality disorders
Paranoid 21 57 are more closely related to each other than to histri-
Schizoid 28 74 onic and narcissistic disorders.
The extent of sharing of common genetic factors
Cluster B (“dramatic”)
Antisocial 38 15
among the various personality disorders was further
Histrionic 31 0 evaluated by estimating the number of common
Narcissistic 24 10 factors and unique determinants of all 10 DSM-IV
Borderline 35 0 personality disorders simultaneously by path analysis
[40]. Three latent genetic factors were inferred as
Cluster C (“anxious”)
Avoidant 35 17
summarized in Table 26.3. The first factor is a
Dependent 31 52 common determinant of all personality disorders.
Obsessive–compulsive 27 85 This common personality factor corresponds to what
is measured as neuroticism or as low self-directedness
Compiled from Kendler et al. [37] for Cluster A, Torgersen et al. [38]
for Cluster B, and Reichborn-Kjennerud et al. [39] for Cluster C. (i.e. irresponsible, blaming) using the Temperament
and Character Inventory (TCI) [2, 7]. The second
genetic determinant contributes most strongly to
antisocial and borderline personality disorder, and
most likely corresponds to high TCI Novelty Seeking

Table 26.3 Path coefficients for contributions to heritability of numbers of personality disorder symptom groupings measured by the
Structured Interview for DSM Personality (SIDP) from 3 latent genetic factors (GF) and unique contributions for each grouping of
symptoms in 2794 Norwegian twins.

Clusters of Path from GF1 Path from GF2 (? High Path from GF3 (? High Path from
personality (? Low self- novelty seeking/low harm avoidance/low unique or
disorder directedness) cooperativeness) reward dependence) unexplained
symptoms GF
Cluster A
Schizotypal 11 22 22 31
Paranoid 30 13 22 28
Schizoid 2 7 37 34
Cluster B
Antisocial 6 63 9 0
Histrionic 49 23 14 0
Narcissistic 35 12 7 33
Borderline 32 43 16 24
Cluster C
Avoidant 14 7 59 0
Dependent 29 15 23 37
Obsessive 29 6 13 41
Adapted from Kendler et al. [40].

319
Chapter 26: Genetics of personality disorders

(i.e. impulsive, quick-tempered) and/or low TCI Table 26.4 Heritability and concordances in 236 monozygotic
(MZ) and 247 dizygotic (DZ) twin pairs for 18 basic scales of the
Cooperativeness (i.e. hostile, revengeful). The third Differential Assessment of Personality Pathology in Canada.
genetic determinant contributed most strongly to
schizoid and avoidant personality disorders, and most Scale label MZ DZ Heritability
likely corresponds to high TCI Harm Avoidance (i.e. correlation correlation (%)
anxious, shy) and/or low TCI Reward Dependence (r  100) (r  100)
(i.e. aloof, detached). Affective 49 12 45
In addition, most of the disorders have substantial lability
genetic determinants that are not explained by these Anxiousness 42 25 44
three latent variables, so the description of the per-
sonality disorders by a few genetic factors is far from Callousness 56 32 56
complete. The specification of personality disorders is Cognitive 48 31 49
better explained by comprehensive multidimensional distortion
personality inventories that measure normal and/or Compulsivity 40 19 37
abnormal personality traits [2, 7, 12].
Conduct 53 36 56
problems
Inheritance of general personality Identity 51 28 53
dimensions problems
Multidimensional inventories for assessing compon- Insecure 45 27 48
ents of personality disorders have been developed to attachment
measure both abnormal traits and also traits that char- Intimacy 47 24 48
acterize the full range of variability in the general problems
population. John Livesley and his colleagues developed
the Differential Assessment of Personality Pathology Narcissism 51 22 53
(DAPP) to measure the self-report of a hierarchy of Oppositionality 41 29 46
dimensions of personality problems. The questionnaire Rejection 33 19 35
is composed of 560 items measuring 18 factorially
derived dimensions, each with at least three specific Restricted 48 26 41
expression
facet scales [10, 32]. The heritability and concordances
in 236 MZ twin pairs and 247 DZ twin pairs is shown Self-harm 39 26 41
for all 18 DAPP traits in Table 26.4. The heritability Social 52 27 53
ranged from 35% for rejection sensitivity to 56% for avoidance
conduct problems and callousness, once again showing
Stimulus- 38 21 40
that personality traits are moderately heritable. seeking
Factor analysis of the 18 basic scales yields
four higher-order factors labeled “emotional lability Submissiveness 41 29 45
or dysregulation”, “antagonism or dissocial behavior”, Suspiciousness 42 29 45
“interpersonal responsiveness or inhibition”, and Adapted from Jang et al. [32].
“compulsivity” [10]. These higher-order dimensions
of abnormal personality resemble the dimensions
of normal personality. For example, emotional dys- TCI self-directedness and cooperativeness, or low
regulation is similar to high neuroticism in the five- DAPP emotional dysregulation and dissocial behavior,
factor model (measured by the NEO) [41] and low or low NEO neuroticism and agreeability) and person-
self-directedness in the seven-factor model of tempera- ality disorder at the other extreme (e.g. low TCI self-
ment and character (measured by TCI) [2]. Antagon- directedness and cooperativeness, etc.) [7, 10, 41].
ism or dissocial behavior is similar to low agreeability Twin studies of personality dimensions in the
in the five-factor model and low cooperativeness in the general population also demonstrate moderate herit-
seven-factor model. Hence these dimensions define ability of normal personality traits. For example, the
healthy personality at one extreme (namely, high TCI was developed based on specific neurobiological

320
Chapter 26: Genetics of personality disorders

Table 26.5 Total heritability of each of the 7 temperament and studies indicate that the narrow heritability of person-
character inventory personality dimensions estimated in 2517
twins in Australia. Unique effects exclude genetic contributions
ality is about 20–30% [36, 48], rather than the 40–50%
shared with other personality dimensions. estimated by twin studies. The prominence of
gene–gene and gene–environment interaction for
Personality Total Unique personality traits confirms the expectation of non-
dimension heritability (%) heritability (%)
linear dynamic interactions from evolutionary theory
Harm avoidance 42 29 [19, 49].
Novelty seeking 39 32 The prominence of genetic complexity has several
practical consequences. Meta-analyses of candidate
Reward 35 20 genes may grossly underestimate the importance of
dependence
individual genetic polymorphisms when interactions
Persistence 30 23 with other genes and/or environmental variables are
Self-directedness 34 25 not measured. There is now substantial direct evi-
dence that personality development depends on the
Cooperativeness 27 16
nonadditive effects of gene–gene interactions [50–52],
Self- 45 26 as discussed in detail elsewhere [15]. For example, TCI
transcendence novelty seeking depends on the three-way interaction
Adapted from Gillespie et al. [47]. of DRD4 with COMT and the serotonin transporter
locus promoter’s regulatory region (5HTTLPR). In
the absence of the short 5HTTLPR allele (5-HTTLPR
L/L genotype) and in the presence of the high activity
and psychosocial data [1, 2, 42–44]. It measures four COMT Val/Val genotype, novelty seeking scores
dimensions of temperament (describing behavioral are higher in the presence of the DRD4 seven-repeat
biases in response to basic emotional stimuli) and allele than in its absence [53]. Furthermore, within
three dimensions of character (describing higher cog- families, siblings who shared identical genotype
nitive processes influencing the maturity of a person’s groups for all three polymorphisms (COMT, DRD4,
goals and values). Initial twin studies were carried out and 5HTTLPR) had significantly correlated TCI nov-
using only measures of temperament [45, 46]. More elty seeking scores (r ¼ 0.4 in 49 subjects, p < 0.01).
recent twin studies have used the TCI and show that In contrast siblings with dissimilar genotypes in at
each of the seven TCI dimensions has a unique gen- least one polymorphism showed no significant correl-
etic variance that is not explained by the other dimen- ation for novelty seeking. Similar interactions were
sions [47]. The heritability of each of the seven TCI observed between these polymorphisms and novelty
dimensions in a sample of 2517 Australian twins is seeking in an independent sample of unrelated sub-
summarized in Table 26.5. Total heritability varied jects [53] and have been replicated by independent
from 27 to 45% without correcting for the reduced investigators [51].
reliability of the short form of the TCI used in this Likewise there is also substantial evidence for spe-
study. Both temperament and character traits are cific gene–environment interactions on personality
roughly equally heritable, and each dimension had development [54–63]. For example, TCI novelty seek-
genetic determinants unique to it (that is, not overlap- ing scores in adulthood are associated with particular
ping with the genetics of other dimensions). Thus DRD4 polymorphisms only if the children were
personality is moderately heritable whether measured reared in a hostile childhood environment with meas-
as normal traits, abnormal traits, number of symp- ures during childhood of maternal reports of emo-
toms, or categorical diagnoses. Alternative measure- tional distance and punitive discipline [61].
ment methods are highly convergent with one another. Assessments of childhood environment were based
on maternal reports and were made prior to the
independent assessment of adult personality.
Epigenetics of human personality Such complex nonlinear interactions present
The estimates of heritability in twin studies are great challenges for identification and replication of
inflated by contributions from gene–gene and gene– relationships. Although the heritability of personality
environment interactions. Adoption and linkage traits is substantial, genome-wide association studies

321
Chapter 26: Genetics of personality disorders

(GWAS) of personality traits have failed to discover neuroanatomy and anthropology allows the specifica-
specific genetic variants [64]. Nearly all genetic vari- tion of a testable model of the emergent brain struc-
ants contributing to variation in personality traits tures and functional capacities along the line of
remain “hidden” for all measures that have been ancestors leading to human beings [3].
tested by GWAS, including measures developed by An evolutionary approach to understanding
Verweij et al. [64]. Resolution of the hidden genetic human functioning has many potential advantages
variation may be impossible if investigators lack an over traditional descriptive approaches, but much
understanding of the functional systems underlying work will need to be done by many people in many
personality development and vulnerability to psycho- fields to delineate and explore each of the components
pathology. It may therefore be useful to develop a of human functioning. A renewed focus on adaptive
testable model of the functional structure of brain functioning in evolution and development is likely to
systems that is relevant for the development of per- be more fruitful than biologically blind testing of
sonality and its underlying brain functions. candidate genes or polymorphic markers of uncertain
A model of the evolution of brain functions may function. It should not be surprising that such an
be the only way to make sense of the psychobiology of evolutionary model is needed for us to begin to
personality and psychopathology. Progress in understand psychiatric genetics in a way that corres-
molecular phylogenetics now allows the specification ponds to natural functional systems. As Dobzhansky
of the ancestral line from the earliest eukaryotes to said, “Nothing in biology makes sense except in the
human beings. Likewise progress in comparative light of evolution” [65].

References 9. Krueger RF. Arch Gen Psychiatry


1999;56(10):921–926.
and Biometric Foundations.
Chicago: University of Chicago
1. Cloninger CR. Arch Gen Psychiatry Press; 1984.
1987;44(6):573–588. 10. Livesley WJ, et al. Arch Gen
Psychiatry 1998;55:941–948. 20. Cloninger CR, et al. Br J
2. Cloninger CR, et al. Arch Gen Psychiatry 1975;127: 11–22.
Psychiatry 1993;50(12):975–990. 11. Markon KE, et al. J Pers 2002;
70(5):661–693. 21. Schulsinger F. Int J Mental Health
3. Cloninger CR. Aust N Z 1972;1:190–206.
J Psychiatry 2009;43(11): 12. Grucza RA et al. J Pers Assess
994–1006. 2007;89(2):167–187. 22. Crowe RR. Arch Gen Psychiatry
1972;27:600–603.
4. Cloninger CR. In O’Donohue WT, 13. World Health Organization. The
et al. (eds.). Personality Disorders: Constitution of the World 23. Crowe RR. Arch Gen Psychiatry
Toward the DSM-V. Los Angeles: Health Organization. 1974;31:785–791.
Sage Publications; 2007. Official Records. Geneva: WHO; 24. Cloninger CR, et al. Arch Gen
5. American Psychiatric Association. 1946. Psychiatry 1982;39(11):1242–1247.
Diagnostic and Statistical Manual of 14. World Health Organization. 25. Mednick SA, et al. Science
Mental Disorders, Fourth Edition Mental Health: New 1984;22:891–894.
(DSM-IV). Washington, DC: Understanding, New Hope. 26. Cloninger CR, et al. In Mednick
American Psychiatric Association; Geneva: WHO; 2001. SA, et al. (eds.). The Causes of
1994. Crime: New Biological Approaches.
15. Cloninger CR. Feeling Good: The
6. Eysenck HJ. In Millon T, et al. Science of Well-Being. New York: Cambridge: Cambridge
(eds.). Contemporary Directions in Oxford University Press; 2004. University Press; 1987.
Psychopathology. New York: 27. Rhee SH, et al. Psychol Bull
16. Cloninger CR. World Psychiatry
Guilford; 1986. 2002;128(3):490–529.
2006;5(2):71–76.
7. Svrakic DM, et al. Arch Gen
17. Wilkinson RG. The Impact of 28. Cloninger CR, et al. Arch Gen
Psychiatry 1993;50(12):
Inequality: How to Make Sick Psychiatry 1978;35(8):941–951.
991–999.
Societies Healthier. New York: 29. Torgersen S, et al. Comp
8. Widiger TA, et al. Dimensional New Press; 2005.
Models of Personality Disorders: Psychiatry 2000;41:416–425.
Refining the Research Agenda 18. Roysamb E, et al. J Pers Soc 30. Eaves LJ, et al. Genes, Culture and
for DSM-V. Washington, DC: Psychol 2003;85(6):1136–1146. Personality: An Empirical
American Psychiatric Association; 19. Wright S. Evolution and the Approach. London: Academic
2006. Genetics of Populations: Genetics Press; 1989.

322
Chapter 26: Genetics of personality disorders

31. Loehlin JC. Genes and 43. Bohman M, et al. J Psychiatr Res 55. Caspi A, et al. Science 2003;301
Environment in Personality 1987;21(4):447–452. (5631):386–389.
Development. Newbury Park: Sage 44. Cloninger CR. Psychol Assess 56. Hintsanen M, et al. J Psychosom
Publications; 1992. 2008;20(3):292–299; discussion Res 2009;67(1):77–84.
32. Jang KL, et al. Acta Psychiatr 300–304.
57. Jokela M, et al. J Affect Disord
Scand 1996;94:438–444. 45. Heath AC, et al. J Pers Soc Psychol 2007;100(1–3):191–197.
33. Pedersen NL, et al. J Pers Soc 1994;66(4):762–775.
58. Keltikangas-Jarvinen L, et al.
Psychol 1999;55:950–957. 46. Stallings MC, et al. J Pers Soc Am J Med Genet B
34. Tellegen A, et al. J Pers Soc Psychol Psychol 1996;70(1):127–140. Neuropsychiatr Genet 2009;
1988;54:1031–1039. 47. Gillespie NA, et al. Pers Individ 150B(3):389–394.
35. Bergeman CS, et al. J Personal Diff 2003;35:1931–1946. 59. Keltikangas-Jarvinen L, et al.
1993;61:159–179. 48. Cloninger CR, et al. Am J Med Genes Brain Behav 2007;
36. Plomin R, et al. J Pers Soc Psychol Genet 1998;81(4):313–317. 6(4):305–313.
1998;75:211–218. 49. Wright S. Annu Rev Genet 60. Keltikangas-Jarvinen L, et al.
37. Kendler KS, et al. Psychol Med 1982;16:1–19. Genes Brain Behav 2006;
2006;36:1583–1591. 5(1):11–18.
50. Benjamin J, et al. Molecular Genetics
38. Torgersen S, et al. Psychol Med and the Human Personality. 61. Keltikangas-Jarvinen L, et al.
2008;38:1617–1625. Washington, DC: American Mol Psychiatry 2004;9(3):308–311.
Psychiatric Publishing; 2002. 62. Keltikangas-Jarvinen L, et al.
39. Reichborn-Kjennerud T, et al.
Psychol Med 2007;37:645–653. 51. Strobel A, et al. Mol Psychiatry Scand J Psychol 2009;
2003;8(4):371–372. 50(6):574–582.
40. Kendler KS, et al. Arch Gen
Psychiatry 2008;65(12): 52. Van Gestel S, et al. Mol Psychiatry 63. Kim-Cohen J, et al. Mol Psychiatry
1438–1446. 2002;7(5):448–450. 2006;11(10):903–913.
41. Costa PTJ, et al. J Pers Disord 53. Benjamin J, et al. Mol Psychiatry 64. Verweij KJH, et al. Biol Psychology
1990;4:362–371. 2000;5(1):96–100. 2010;85:306–317.
42. Cloninger CR. Psychiatr Dev 54. Caspi A, et al. Nat Rev Neurosci 65. Dobzhansky T. American Biology
1986;4(3):167–226. 2006;7(7):583–590. Teacher 1973;35:125–129.

323
Ethical issues in behavioral genetics
Chapter

27 Stephen H. Dinwiddie, Jinger Hoop, and Elliot Gershon

Introduction The use (and misuse) of genetic


The principles of Mendelian genetics became widely information for social ends
known around 1900; over the next 20 years the
theoretical basis of quantitative genetics was estab- Eugenics
lished. Molecular investigation followed statistical Artificial selection by controlled breeding can be
genetics as Watson and Crick characterized the traced back to the beginnings of pastoralism and
structure of DNA in 1953. By 1990 the Human agriculture, and thus pre-dates by millennia any sci-
Genome Project had began. Now, two decades later, entific understanding of the mechanisms of biological
the Personal Genome Project has made genetic and inheritance. Behavioral characteristics cannot have
health data on its first volunteers available and hopes been ignored in this endeavor; presumably domestic
to eventually recruit 100 000 subjects. The tempo of animals were selected very early on for docility at
discovery and the power of technology will continue least, and hunting dogs (to take but one example)
to increase synergistically, and advances in genetics have long been bred for very specific traits of tem-
will pose in concrete terms ethical questions which a perament and behavior.
handful of years ago were only speculative and Family resemblance for many behaviors and attri-
theoretical. butes among humans, such as personality or intellect,
It should be acknowledged that advances in genet- has also been widely acknowledged, and while select-
ics have not so much raised new ethical questions as ive mating to influence these was proposed at least as
simply refined and rephrased established ones. More- far back as Plato’s Republic, philosophers have more
over, few if any of the major ethical questions now commonly emphasized alterations in environmental
being raised have relevance solely to behavioral gen- factors to enhance the well-being of society. Popular
etics, while some, such as those touching on issues acceptance of the proposition that the social good
such as gene therapy or genetic engineering of might be substantially improved by differential repro-
humans, are unlikely to be relevant to considerations duction of selected citizens appears to be of much
of behavior and mental illness in the short term. On more recent origin. The invention of the term “eugen-
the other hand, the role of genetics goes far beyond ics” is attributed to Sir Francis Galton, who in 1883
the determination of physical characteristics or risk of wrote:
medical disease – it touches on how we must now
Whenever a low race is preserved under conditions of
think about such fundamental concepts as personal
life that exact a high level of efficiency, it must be
responsibility.
subjected to rigorous selection. The few best specimens
In this chapter, we will concentrate mainly on of that race can alone be allowed to become parents, and
three areas: The use (and misuse) of genetic infor- not many of their descendants can be allowed to live. On
mation for social ends; issues that arise at the inter- the other hand, if a higher race be substituted for the low
face of medicine and other social institutions such as one, all this terrible misery disappears. The most
the law; and issues of privacy and control over infor- merciful form of what I ventured to call “eugenics”
mation about one’s genotype. would consist in watching for the indications of superior

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

324
Chapter 27: Ethical issues in behavioral genetics

strains or races, and in so favoring them that their manifestly unfit from continuing their kind . . . Three
progeny shall outnumber and gradually replace that of generations of imbeciles are enough.1
the old one . . . ([1], p. 199).
Prior to Buck v. Bell, sterilization laws in the United
The historical context of this passage must be kept in States had been legally problematic but the decision
mind: It was written a quarter-century after Darwin’s cleared the way for their expansion such that, for
On the Origin of Species was published, and Galton (a example, the state of Oklahoma could mandate steril-
cousin of Darwin’s) was a Victorian Englishman par ization of “habitual criminals”, defined as individuals
excellence: a geographer, meteorologist, mathemat- convicted of repeated felonies involving “moral turpi-
ician, and biologist who seemed to have had in tude” [4, 5]. The focus of eugenics had shifted from
adequate measure that comfortable certainty of his encouraging the “fit” to reproduce to sterilizing the
own superiority and that of his society so characteris- “unfit” as attention moved from disease or physical
tic of the age – an intellectual and cultural chauvinism weakness to constructs such as “feeblemindedness”
that, in tandem with what was later called “social and habitual criminality, with the implicit assumption
Darwinism”, was to shape the eugenics movement. that all of these constructs were equally biological and
Galton’s scientific achievements lent considerable heritable. But, as with kindred “sexual psychopath”
weight to his views and he can therefore be properly laws that proliferated at about the same time, by
considered a founder of “scientific racism”. uncritically accepting scientific-appearing assertions
The propositions that the human race could be and using them as justification for policy, these stat-
improved by consciously selecting for (or against) utes served social aims: surely no reasonable person
certain behavioral characteristics and that many such could oppose governmental efforts to lighten the
characteristics were biologically heritable fit well also burden on upstanding, tax-paying citizens who other-
with the intellectual climate of the US Progressive era: wise would continue to pay ever-increasing amounts
In an 8–1 decision by the US Supreme Court, in Buck for the support of unproductive individuals who were
v. Bell [2] (challenging a Virginia statute that allowed at best doomed to spend their lives in misery and at
compulsory sterilization of the mentally retarded) worst would imperil the public by their acts of
Justice Oliver Wendell Holmes, Jr. described the case immorality or violence. The punitive subtext was
as follows: not as readily apparent, either in the United States
Carrie Buck is a feeble-minded white woman who was or other countries (including Canada and a number
committed to the State Colony . . . She is the daughter of in Europe) that adopted eugenic measures.
a feeble-minded mother in the same institution, and the But it was in Nazi Germany that the process
mother of an illegitimate feeble-minded child . . . The reached its end-point. As in the United States and
Commonwealth [of Virginia] is supporting in various elsewhere, eugenic arguments were used in Germany
institutions many defective persons who if now first to justify involuntary sterilization of individuals
discharged would become a menace but if incapable of with a variety of mental or neurological illnesses
procreating might be discharged with safety and become including schizophrenia, manic-depressive illness,
self-supporting with benefit to themselves and to society; alcoholism, epilepsy, and hereditary blindness or
and that experience has shown that heredity plays an
deafness, among other conditions; over time, those
important part in the transmission of insanity,
imbecility, etc.
deemed to be unworthy of life were not merely steril-
ized but executed, supposedly to promote the welfare
Holmes went on to write (no doubt thinking of his of society as a whole [6]. Soon, entire ethnic groups
own military experiences in the Civil War): were decreed to be inferior, indeed not fully human
We have seen more than once that the public welfare and a threat to the Reich – and so deserving of death.
may call upon the best citizens for their lives. It would be It had taken little more than half a century to go from
strange if it could not call upon those who already sap
the strength of the State for these lesser sacrifices, often 1
As it happened, conclusions about Carrie Buck’s
not felt to be such by those concerned, in order to “imbecility” and that of her daughter were at best
prevent our being swamped with incompetence. It is overstated. There is also convincing evidence that the
better for all the world, if instead of waiting to execute pregnancy was the result of rape rather than “promiscuity”
degenerate offspring for crime, or to let them starve for and that the commitment was motivated by an effort to
their imbecility, society can prevent those who are protect the family’s reputation [3].

325
Chapter 27: Ethical issues in behavioral genetics

encouraging reproduction of the “fit” to discouraging By virtue of training and orientation, mental
and preventing reproduction by the “unfit” to whole- health professionals may tend to concentrate on the
sale murder. role of explanatory, putatively causal factors in think-
What can be learned from these atrocities? The ing about behavior, particularly that which seems to
history of the eugenics movement shows how easily arise from mental illness; both psychoanalytic theory
social value judgments can be taken for incontrovert- and biological psychiatry, still the two dominant para-
ible scientific fact. It also demonstrates how easily the digms in clinical practice, tend to be philosophically
rights of the defenseless and marginalized can be deterministic, though of course the causal factors
violated – even by individuals many of whom would proposed are radically different. It would seem, at
(we suspect) honestly assert they had nothing but least at first glance, that a better characterization of
altruistic motives. It is easy in retrospect to trace this the underlying biology would provide a whole new
evolution – or devolution – and condemn it; but it is range of explanations, and hence excuses at least for
far too simplistic to ascribe intellectual dishonesty undesirable or self-defeating behavior. (At least in this
and malevolent motives to all those who supported formulation the father as well as the mother would be
the eugenics movement: History amply shows that blamed!)
even the most meritorious of ideas and programs But of course to do so would be to commit a
can be hijacked and used to support depraved and fundamental error. A distinction must be made
corrupt ends. Rather, the tragedy stemmed from the between finding an explanation for a given behavior
fact that from early on, constructs of variable and and assigning that behavior a moral weight: It is not
often minimal validity were taken out of proper con- clear that a better understanding of the underlying
text and uncritically used to justify the victimization biology must necessarily impact that value judgment.
of powerless individuals, by so doing serving social Regardless of the biology behind it, we subjectively
agendas ranging from dubious to overtly evil. have a sense of our own “agency”, and we assign
moral responsibility based on that perception. Any
mechanistic account (regardless of the extent to which
Issues at the interface of medicine the factors involved are heritable) must include that
and the law experience. After all, as Dennet has noted:
What we want when we want free will is the power to
Genetics and responsibility decide our courses of action, and to decide them wisely,
Personality traits appear substantially biologically in the light of our expectations and desires. We want to
transmissible. Though estimates vary, it is likely that be in control of ourselves, and not under the control of
roughly half of observed variance in normal person- others. We want to be agents, capable of initiating, and
ality traits can be explained by genetic factors [7]. taking responsibility for, projects and deeds. All of this is
For specific psychiatric illnesses, additive genetic ours, I have tried to show, as a natural product of our
factors (assuming no radical environmental changes, biological endowment, extended and enhanced by our
of course), may account for as much as 80% of the initiation into society ([14], p. 169).
observed phenotypic variance in schizophrenia and Dennet also points out that the role of development in
bipolar illness [8–10] and perhaps around 65% for becoming a moral agent cannot be ignored. The pro-
alcohol dependence and conduct disorder [11, 12], cess begins, after all, with an infant who has a unique
though the latter two appear to have some common set of innate (potential) qualities but minimal self-
genetic determinants [13]. awareness or volition; the end result is the contingent
Progress in identifying specific alleles that have product of reciprocal interactions with the environ-
differential influence on normal personality variance ment (a point made, after all, by the biometric mod-
has to date lagged behind identification of loci eling that indicates that roughly half of personality
involved in moderating risk for psychiatric illness, variance can not be attributed to heritable factors).
an endeavor that itself can as yet boast few successes. Over time, the actor more and more actively shapes the
But such alleles may certainly be identified, using surrounding environment as well as being shaped by it
presently available methods. Is it too early to ask what (there is compelling evidence that gene expression can
effect characterizing such influences might have on be powerfully affected by social environment, for
how we think of individual choice and responsibility? example [15]). In short, there is a profound distance

326
Chapter 27: Ethical issues in behavioral genetics

between genotype and the final product of that devel- views regarding the developmental process of respon-
opmental process; indeed, given the powerful inter- sibility, Dennet has argued:
play between environment and gene expression it may Any finite control system (such as a human brain) will
be misleading to talk of the “final product” of a always be prone to making mistakes or arriving at
dynamic process. This is consistent with a position decisions that a more leisurely analysis would condemn;
sometimes called “compatibilism” or “soft determin- it is an inevitable feature of human character, even
ism”, or perhaps “conditional free will”: perfected to its limit. Original sin, naturalized. It is wise,
[S]ocial human behavior is contingent on a countless however, to adopt policies that minimize the bad effects of
number of possible decisions from among which the these inevitable defects of character . . . [B]y somewhat
individual may choose. Not all of those decisions are arbitrarily holding people responsible for their actions, and
feasible, however, nor are the resources available that are making sure they realize that they will be held responsible,
required to act on them. Choosing a course of action, we constrain the risk-taking in the design (and redesign)
therefore, is limited by preset boundaries . . . [which] of their characters within tolerable bounds ([14], p. 165).
include current circumstances and opportunities, It should be kept in mind that the legal system makes
learning experiences, physiological abilities, and genetic no pretense at being scientific; ultimately it is simply a
predispositions ([16], pp. 30–31).
social means of settling disputes short of bloodshed,
To observe that individuals’ behaviors are influenced and it can be argued that the criminal law is at base a
and constrained by a host of factors (some more and means of ensuring public safety and order, backed by
some less accessible to reflection or responsive to the armed power of the state. (In that view, fairness or
environmental factors) is hardly a revolutionary state- at least the public perception of fairness might be seen
ment. More importantly, to assume that once one has as no more than the easiest way to have citizens
an understanding (from some perspective) of the acquiesce to legal restrictions on their behavior.)
“cause” of an action that the actor is relieved of The usual justifications for criminal sanction
responsibility is simply to commit a category error include reformation of the wrongdoer, restraint (or
(why should knowledge of a causal process necessarily incapacitation) of further wrongdoing (for example
impact moral judgment?). There is, moreover, no by imprisonment), retribution (“an eye for an eye”),
logical reason that biological characteristics which and deterrence of the actor or others by the example
might make some choices for a given actor harder of punishment. Though the relative emphasis on
should be privileged over well-known environmental these factors shifts over time, all have in common
factors that similarly influence behavior. Perhaps a the idea that the criminal is a moral agent, responsible
better way of thinking of how to assign moral respon- for his actions, and as a result, for most serious crimes
sibility for an action, determined or not, would be to both the commission of a wrongful act (actus reus)
consider the extent to which: (a) the act follows a and the concurrent presence of intent (mens rea)
rational as opposed to irrational deliberative process; must be demonstrated. It is generally (though not
and (b) the extent to which the actor him- or herself universally) accepted that severe mental illness may,
accepts and endorses the desires motivating the act under some circumstances, negate or at least minim-
[17]. In this formulation, genetic factors are of no ize punishment (i.e. by undermining mens rea entirely
greater or lesser import than any other causative or via the insanity defense in the former case, or by
influences. diminished capacity or the equivalent in the latter). It
would seem obvious that, to the extent inborn char-
Implications for criminal responsibility acteristics are beyond the control of the individual,
and punishment they should be taken into account when holding indi-
Individuals are held accountable for their actions, viduals to account for their behavior, and certainly it
despite knowledge that some behavioral choices are would be better (contra Dennet) if this process were
difficult, every day. The preeminent social mechanism scientifically based and not “somewhat arbitrary”.
for assigning responsibility for behavior is the legal But tempting as this approach might be, as dis-
system. We will concentrate on criminal law here, cussed above, it is problematic, and, indeed, an analo-
though of course the issue might arise in a variety of gous approach has been tried – and failed. In 1954 the
other contexts, for example in personal injury litiga- US Court of Appeals for the District of Columbia
tion. As a matter of social control, in keeping with his federal circuit found, in Durham v. US, that a criminal

327
Chapter 27: Ethical issues in behavioral genetics

defendant should not be held accountable if it could that genetic variants that appear to be associated with
be shown that the “unlawful act was the product of violence seem either to be quite rare in the general
mental disease or defect” [18], the hope then being population [23] or to result in violence only in the
that mental health professionals could identify to a context of adverse social circumstances [24]. The
sufficient degree of accuracy and reliability, at least in variability in expression even in the case of very
selected cases, the psychological factors motivating specific mutations again underscores the complexity
specific criminal acts and could therefore assist the of the interplay between genetic factors and specific
court in assigning the appropriate degree of criminal triggering environmental factors. Thus, while it is
responsibility and hence the appropriate severity of conceivable that (in some environments) rare genetic
punishment. In the event, there was sufficient dis- variants might be strongly associated with specific
agreement among the experts regarding diagnosis, criminal acts, it seems more likely that heritable
causation, and even the boundaries of “mental disease factors that explain a substantial portion of observed
or defect” that, despite legal holdings which attempted criminal behavior will moderate risk in a rather non-
to bring some degree of order to the process, this specific way, perhaps through decreasing intelligence
approach was abandoned less than 20 years later [19]. or heightening impulsivity.
Substitute “genetic factors causing unlawful activ- Any exculpatory power that such findings might
ity” for “mental disease or defect”, and the same confu- have is likely also to vary with the degree of associated
sion would likely eventuate (though the tautological risk: if a given genotype is associated with a markedly
nature of this approach is more explicit in this formu- elevated risk of violence compared to the population
lation). No matter what phraseology is used, what base rate, but many or most of the individuals with
might properly be included (or excluded) is a matter that genotype still do not act violently, it is not likely
of perspective. The characterization of genetic factors to persuade the justice system very strongly. On
associated with, for example, antisocial personality the other hand, if the genotype were associated with
disorder or perhaps elevated risk for violent behavior violence in a large majority [17], particularly in any
would undoubtedly lead to defense attorneys arguing condition whose expression is strongly influenced by
for clemency or exculpation on this basis – as has gene–environment interaction, there is the concern
indeed been done sporadically at least since Clarence that individuals with a “high risk” genotype might be
Darrow’s defense of Leopold and Loeb in 1924 [20]. inappropriately labeled as violence-prone, to their harm.
But given the many processes operating in the space It is not clear that the identification of such gen-
between genotype and behavior, many “causes” of etic factors would necessarily be exculpatory in any
behavior could be adduced, operative at many levels. event. While the presence of a genotype associated
Note that this concern has nothing to do with the with low monoamine oxidase A (MAOA) has been
scientific foundation upon which a diagnostic con- associated with markedly increased (in the setting of
struct should be based; the concern is over the misap- childhood maltreatment) risk of violent behavior
plication of such a construct. Surely the identification [24], such behavior was far from universal in the
of specific genetic factors that promote the develop- cohort studied. Clearly other factors (heritable, envir-
ment of a given behavioral syndrome would count onmental, or very likely both) further moderate risk,
toward the validation of that syndrome, and it has and there is nothing in the nature of genetic factors that
long been established that some heritable factors do, makes them inherently inaccessible to amelioration
indeed, appear to be associated with an elevated pro- after birth. Moreover, the presence of a statistical
pensity for crime and violence [21], though few such association between a given genotype and an elevated
factors have so far been identified. However, given likelihood of expressing a class of behaviors (violence
that “criminality” is a meaningless concept except in in this example) says nothing of the subjective experi-
the context of the society that defines specific acts as ence of choice and congruence with values in the
unlawful, and the observation that at the clinical level context of a specific antisocial act [25]. Thus, it would
a number of psychiatric disorders appear to be asso- be a tremendous oversimplification to equate pres-
ciated with elevated risk for criminal and/or violent ence of a given genotype with lack of moral responsi-
behavior [22], it would seem unlikely that heritable bility. To parse out the relative contribution of
factors will prove to map very precisely onto legal specific genetic factors to a given antisocial act would
concepts such as criminality. It is intriguing to note be to go far beyond the limits of science.

328
Chapter 27: Ethical issues in behavioral genetics

Indeed, it could be argued that, to the extent that be quite nonspecific in their effects. But as happened
such potentially violent individuals see their acts as with the eugenics movement, there is the risk that
consistent with their values and as appropriate means such information might well be taken out of context
to achieve their ends, they might reasonably be held to and misapplied to support specific social agendas.
be more, rather than less, culpable [17]. Regardless of
any biological roots of antisocial attitudes and behav-
iors, there is a general sense that individuals are given Privacy and control of genetic
multiple opportunities in life to observe and model information
their behaviors along more acceptable lines: their
failure to do so is precisely what society feels is worthy Beyond the offender
of condemnation. The use of DNA testing on biological samples has
Sentencing might conceivably become harsher proven to be of great value in identifying criminals;
rather than more lenient. Certain groups of offenders also of great importance has been its use in excul-
(e.g. those diagnosed with antisocial personality dis- pating individuals wrongly convicted. There are many
order or who have elevated scores on measures of other uses of genetic information in forensic settings,
psychopathy [26]) are known to be less likely to learn including elimination of suspects during the investi-
from experience and have higher risk for reoffense. gative process, identification of human remains, and
To the extent that any genetic findings overlap with establishment of paternity. In the United States, the
such clinical constructs, genetic results could be used Combined DNA Index System (CODIS), based on
by the State to increase the penalty or deny parole for States’ databases, has information at 13 loci on more
a given offense, under the theory that such individuals than 5.5 million people [28]; in the United Kingdom,
are more unlikely to respond to less harsh punish- the National Criminal Intelligence DNA database
ment or that they present a greater danger to the (NDNAD) has information at 10 loci on more than
public and so should be incarcerated longer. If the 4 million – about 5% of the population [29]. In neither
condition is relatively rare, any such test, even with system is acquisition and/or nor storage of genetic
relatively high specificity and sensitivity, will have information restricted to those convicted of a crime.
quite a high “false positive” rate [27], thus risking an In the United Kingdom, hair and samples from
unjust result. mouth swabs can be taken without consent if the
Moreover, even if it were accepted that such indi- individual is arrested or detained in a police station
viduals might not be morally culpable, it seems likely even if the sample is not relevant to the crime being
that such individuals would be treated much the same investigated. In both the US and UK databases, many
as insanity acquitees – who are on average deprived of samples are from juveniles or those who have not
their liberty longer than those simply found guilty, by been accused or convicted of crimes.
virtue of being perceived by society as dangerous. In theory, genetic information from a crime scene
Legislatures might well craft statues allowing indefin- could result in a partial match to a profile already in
ite incarceration of such individuals, much as many the database, thus raising the possibility that the
US states now allow “sexually violent predators” (and offender might be a close relative of that individual.
in England and Wales, individuals found to have This has been done on occasion [30], though since the
“dangerous and severe personality disorders”) to be pool of suspects can be quite large (there might be
indefinitely civilly committed for purposes of “treat- numerous partial matches), in practice the utility of
ment” – even if no clearly effective treatment exists this approach is limited.
and (if there were) even though there is no obvious Despite the value of these databases to criminal
way to determine whether it might be effective until investigators, their establishment has raised concern.
after the release of such individuals back into the One concern is privacy: in the case of partial matches,
community. there is the possibility of uncovering family secrets
On balance, it seems unlikely that genetic infor- such as a child fathered outside of the primary rela-
mation (at least if properly weighed) will contribute tionship or the discovery of nonpaternity. Because
much to the legal determination of blameworthiness samples may be stored there remains the possibility
or its absence. Genetic factors are likely to be too far of expanding the genotypic information beyond –
removed from causation of behavior, and are likely to possibly far beyond – the loci currently assessed, with

329
Chapter 27: Ethical issues in behavioral genetics

unclear ramifications. There is also the issue of the But autonomy can be undermined by limitations
proper use of this information in the case of juveniles in rational thinking (for example the presence of an
or those not convicted perhaps not even of a crime. active psychotic illness) or by limited understanding
Finally, despite legal limitations on the use and dis- of the nature and consequences of the decision to be
closure of this information, requests to obtain identi- made. A related principle is therefore that of
fying information about profiles have been granted informed consent, which is generally held to require
(at least in the case of the UK database) for unclear three elements: voluntariness, possession of sufficient
reasons: information upon which to base a decision, and com-
The first of the two approved operational requests petence – the ability to make decisions for oneself in
was made by police to check for “named individuals”, both the legal sense (i.e. an adult not under guardian-
but it is not clear what this might mean. Further ship) and in the clinical sense of having sufficient
clarification was provided which explained that such mental ability to process information and communi-
requests related to “seeking named suspects in a specific cate a choice. This “decisional capacity”, in turn,
inquiry at the police’s request”. This is still far from might exist only at a very basic level, demonstrated
explaining such a use of the NDNAD. If the police seek for example by the ability to make a choice, whether
a DNA match on the NDNAD and one is found, then
rational or not. A somewhat higher threshold would
that individual’s name will be readily known by the
police. If the police sought a named individual’s DNA
be demonstration of the ability to comprehend the
profile for purposes other than making a match to one facts of the situation; higher yet would be evidence
found at a crime scene, this might signal a departure that the individual was able to meaningfully appreci-
from the purposes for which the NDNAD may be ate the risks and likely consequences. Finally (a very
lawfully used ([30], p. 83). high standard) would be demonstration of an ability
This may relate to another area of concern, the possi- to rationally manipulate the information and inte-
bility of “function creep” –using the information for grate it with his or her individual values.
purposes other than criminal investigation, for The issues of autonomy and informed consent
example the establishment of paternity so that child may arise in the context of testing for genetic risk. It
support payments can be made. Other uses with sig- appears generally accepted that if genetic testing is
nificant potential social benefit could easily be identi- offered, it should be voluntary. The individual should
fied; but no matter how laudable the goals, such use be given relevant information regarding the nature of
would be to go beyond the initial rationale for the the information to be obtained and the risks of the
databank’s establishment. procedure, should have sufficient time to consider the
options, and should be free to withdraw from testing,
even after the fact. But it is less clear what level of
Issues of consent decisional capacity would be appropriate. Presumably
Beneficence (the principle that the physician should a more conservative standard would be superior in
act for the good of the patient) and nonmaleficence cases where risks and benefits are less clear-cut.
(the principle that harmful acts should be avoided and How does this translate into practice? A current
if harm is unavoidable it should be balanced against example is the recent offering of a test that screens
potential good) are two widely accepted foundations for several single-nucleotide polymorphisms (SNPs)
for medical ethics. Other foundational principles associated with elevated risk for breast cancer [32].
include respect for autonomy and fairness or justice As opposed to BRCA1 and 2, which are relatively
[31]. Subsumed under these are the additional prin- less common but associated with substantial risk of
ciples of honesty, dignity, and confidentiality. cancer, in this test the SNPs identified appear to be
Autonomy is generally privileged; that is, assum- associated with relatively small increases in risk of
ing sufficient decisional ability, it is generally believed disease individually, and it is likely that the variants
that an individual should be free to make his or her so far identified constitute only a small subset of the
own choices without external constraint, even if the total [32]. Thus, negative results cannot be equated
decision leads to an undesirable consequence (the with lower overall risk, while positive results are open
paradigmatic case is the Jehovah’s Witness who to misinterpretation as to the real degree of risk.
refuses a blood transfusion, risking death but adher- It might be assumed that the physician, by virtue of
ing to his or her religious principles). training and access to information, would have a

330
Chapter 27: Ethical issues in behavioral genetics

better understanding of the probabilistic nature of for employment or health insurance reasons. Would
any such tests and their implications. Unfortunately, it make sense to identify higher risk individuals and
it is not clear that this is the case [33] and physicians adjust premiums accordingly? Perhaps not, since ini-
are not immune to common cognitive errors [34]. tiation of substance use is frequently at a young age.
One can easily imagine situations in which patients Thus the individual paying the premiums (or the
request testing but neither the patient nor the phys- company paying the premiums for the employee)
ician fully understands the implications. Given the would be financially penalized, not for his or her
complexity of the information and the probabilistic own actions or even her own risk status but strictly
nature of any testing results, how likely is it that the based on genotype.
patient will be able to meaningfully understand the It would appear that in the United States, with the
implications of the test? What level of understanding signing of the Genetic Information Nondiscrimina-
on the part of the patient is acceptable, and how tion Act (GINA) on May 21, 2008 (when it became
should one assess it? Who should pay for the time Public Law 110–233) these issues may have been
spent educating the patient? How would sufficient resolved. GINA prohibits denial of insurance cover-
competence to order, interpret, and explain the results age to healthy individuals or requiring payment of
on the part of the physician be assured? increased insurance premiums by healthy individuals
based on genetic screening; it also prohibits use of
genetic information by employers when making deci-
Testing and addictive disorders sions on hiring, promotion, or termination.
The same difficulties regarding the interpretation of
positive and negative test results could easily arise in
the context of psychiatric illness. The use of such testing
For-profit testing for genetic susceptibility
might lead to a false sense of inevitability and fear of a to psychiatric illness
“worst-case” scenario, even though many mental ill- Commercial testing for susceptibility to severe mental
nesses are eminently treatable. The situation could be illness using molecular genetic methods has been
further muddled if testing were applied to relapsing available since at least 2007, when a test for variants
and remitting rather than chronic conditions. In fact, in the GRK3 gene was offered directly to consumers
particularly if test sensitivity were low, one might have by an internet laboratory [39]. According to the com-
a test that missed many severe cases and identified pany, positive test results indicate a doubling to trip-
numerous “cases” who through much of their lives ling of the risk of bipolar disorder as compared to the
needed little if any intervention [35]. general population – though of course this still
This appears to be the case with addiction to (at amounts to only a 2 or 3% chance of developing the
least) nicotine and alcohol, both of which appear to disorder.
have substantial heritable components [36, 37]. Both Other companies have recently announced plans
are relapsing and remitting conditions; for alcohol- to market similar tests, (e.g. for susceptibility to
ism, only about half those diagnosed are likely to be schizophrenia, autism, and antidepressant-induced
drinking and having problems at any one time [38]. suicidal thinking) [40, 41]. As of the time of this
Its course is quite variable and treatment quite effect- writing, one company offers testing (at a cost of
ive. The clinical utility of any genetic test (even if $399 plus shipping and handling) for some 91 traits
accurate) is unclear since screening for early diagno- and conditions – though it notes that of those, only 23
sis, which consists of asking a few questions, is quick have been found to be “supported by multiple, large,
and cheap. Moreover, health risks (cancer, heart dis- peer-reviewed studies” while the remaining 68 are
ease, and the like) are due to repeated use regardless “research reports” with less strong support. The “dis-
of whether the individual merits a clinical diagnosis of eases, traits, and conditions” assessed range from eye
dependence, and in the case of tobacco and illicit color, bitter taste sensation, and earwax type to resist-
drugs presumably all patients should be counseled to ance to HIV/AIDS, celiac disease, and diabetes melli-
avoid initiation or continuation of use regardless of tus, while assessments in the “research report” range
genetic liability to addiction. from odor detection and freckling to alcohol depend-
More likely, screening would be seen as most ence, schizophrenia, and progressive supranuclear
useful for purposes of identifying “at-risk” individual palsy [42].

331
Chapter 27: Ethical issues in behavioral genetics

In the United States at this time, there is little afflicted – if that understanding is properly and con-
government regulation of genetic testing, and test sistently communicated to the public.
providers are not required to provide information
supporting clinical validity [43–45]. There is valid
concern that commercial interests may exploit fears A duty to warn?
of psychiatric patients and their families by offering Clinicians are generally expected to maintain the
tests with limited benefit and significant potential confidentiality of health information. However, this
for harm. In the case of some Mendelian disorders duty is not absolute, and must in some situations be
and hereditary cancers, potential harm may include violated if other ethical obligations are to be met.
complications of preventive treatment as well as Reporting is mandated for certain communicable
psychological stress, family strain, and social stigma- diseases or injuries, where presumably the harm
tization [46–51]. Fortunately, the risk associated due to breach of confidentiality is balanced by the
with testing for these disorders appears to be low, benefit accrued by society by tracking and interven-
at least if is accompanied by safeguards such as ing in identified cases. In psychiatry, it is now gen-
informed consent, confidentiality protection, genetic erally accepted also that confidential information
counselling, and pretest screening. However, it is may be released if there is reason to believe that
possible that the impact of testing for common psy- doing so would avert significant danger (for example
chiatric disorders may be greater. Mental illnesses if a patient disclosed an intention to harm an identi-
more directly affect emotions, cognition, and behav- fied [or identifiable] victim [57]), and various courts
ior – key aspects of an individual’s “personhood”. have held that under some circumstances the
Learning that one has a genetic susceptibility to such treater’s duty can extend to such a third party (some-
disorders may threaten one’s sense of self and future times referred to as a “duty to warn” or “duty to
prospects in ways different from the threats posed to protect”).
those at risk for cancer or other nonpsychiatric med- Knowledge of an individual’s genotype gives one
ical conditions [46]. In addition, those who seek some knowledge of the genotype of family
testing for psychiatric disorders because they have members. Even if only probabilistic and uncertain,
affected relatives may (partially because of know- there is an evolving body of malpractice case law
ledge of that family history) have higher baseline that indicates that health professionals must under
levels of anxiety and depression [52, 53], and this some circumstances disclose such information
pre-test distress has been associated with post-test under a “duty to warn” theory. In one such case
distress [51]. On the other hand, a person identified [58], after developing medullary thyroid carcinoma,
by testing as at elevated risk might be more inclined the plaintiff sued the physicians who had treated
to seek monitoring (and preventive treatment if her mother for the same condition, alleging that
justified by future research). given the heritable nature of risk for the disorder
The social risks of susceptibility testing may be she should have been warned so that she could have
greater for psychiatric disorders than for other condi- taken preventive measures. The Florida Supreme
tions, because psychiatric disorders are more stigma- Court held that the physicians did indeed have a
tized. It is not clear currently whether progress in duty to inform a third party (the child of the
psychiatric genetics will improve or worsen stigma. patient) under such circumstances.
It may be that further evidence of the biological Similar reasoning was used in a New Jersey case
nature of psychiatric illness will reduce negative social [59], which also involved heritable risk factors for
attitudes (though the massive amount of evidence to neoplastic disease. It has been suggested that impos-
date seems not to have sufficed). On the other hand, ition of this “duty to warn” in the presence of a
some reports seem to indicate that elucidation of the known risk might eventually lead to the creation of
genetic bases for major mental illnesses may make an affirmative obligation for physicians to routinely
these disorders appear more permanent and less test for certain genetic conditions, even though the
amenable to treatment [54–56]. Nonetheless, it is to consequences of providing such information are
be hoped that better understanding of the causation unknown [60]. Ideally, it might prompt the (poten-
(including genetics) of psychiatric illness will lead to tially) affected individual to take steps to increase
better treatment and social acceptance of those monitoring or prevention efforts. However, the

332
Chapter 27: Ethical issues in behavioral genetics

concern has been voiced that knowledge of this risk


might prompt parents to overestimate its degree, with Testing in the workplace
unknown but potentially adverse effects on the While the potential for misuse of genetic information
parent–child relationship [61], though the basis for has been recognized and opposed at least since the
this concern and the likely magnitude of harm has Declaration of Bilbao in 1993 and addressed in UNES-
been disputed [62]. CO’s Universal Declaration on the Human Genome
The impact of such cases on behavioral genetics is and Human Rights (1997) and its International Dec-
unclear. For traits or disorders with substantial gene– laration on Human Genetic Data (1993), this has not
environment interaction it is likely that imposition generally resulted in passage of statutory protections
of a duty to warn is at best premature [60]. However, internationally; in the United States, while workplace
for conditions such as certain forms of mental retard- discrimination on the basis of genetic risk has been
ation a similar duty has been established [63]. In a alleged many times, it does not appear that this has led
recent Minnesota case, the plaintiff had a child with to significant litigation. The best-known (and to date
fragile X syndrome. Testing for the condition was perhaps the only) such case involved an employer that
considered, but not done. A second child who also surreptitiously tested employees for a rare genetic vari-
had the syndrome was subsequently born to the ant associated with the development of carpal tunnel
plaintiff. In this case there was a legal finding syndrome and threatened to fire a worker who declined
that the “duty” extended past the patient (the first testing. In this case, the Equal Employment Opportun-
child) to the mother (the carrier of the condition) ity Commission (EEOC) filed suit against the employer
on the theory that it was foreseeable that she would (Burlington Northern Santa Fe Railroad) in 2001; the
have other children. The court did not address case was quickly settled and testing suspended. Since
whether a similar duty might be established to other passage of GINA, it would appear that in the United
relatives. States genetic information cannot be used for purposes
But what of the situation of an individual who, for of hiring, promotion, or retention. Nonetheless,
reasons of personal privacy or even malicious reasons, markers indicating elevated liability for various illnesses
refuses permission for such sharing? This issue would likely be of interest to employers for planning
appears to have been addressed in 1982 by the US purposes, to the extent that these illnesses may be
President’s Commission for the Study of Ethical associated with absenteeism, decreased productivity,
Problems in Medicine and Biomedical and Behavioral or increased health service utilization. Given the high
Research [64], which concluded that confidentiality prevalence of mental illness and its impact on worker
could be breached if: (a) reasonable attempts to obtain productivity, should tests for liability to common psy-
voluntary consent to disclose the information failed; chiatric conditions become available, they would
(b) there were a high probability of harm if the infor- undoubtedly be of great interest to employers.
mation were withheld and disclosure could prevent The availability of genetic screening would raise
the harm; (c) the potential harm were serious; and (d) issues not only of privacy but perhaps of benefit to the
disclosure was limited to the genetic information employee. How might (or should) a test that indicated
necessary for diagnosis and treatment of the disease. heightened risk for mood or anxiety disorders be used if
However, this “duty to warn” might conflict with the an employee is considering a high-stress position?
Health Insurance Portability and Accountability Act Under GINA, the company would not be able to bar
(HIPAA), which places broad restrictions on release the employee from taking the position, but the
of medical information. Under current regulations employee might factor that information into the deci-
(Section 164.512[j] of the final HIPAA privacy rules), sion as to whether or not to accept the position. Perhaps
one circumstance under which protected health infor- such information might encourage him/her to make
mation may be released without the individual’s lifestyle changes or investigate preventive therapies.
authorization would be to “prevent or lessen a serious Conversely, if he or she does accept the position, should
and imminent threat to the health or safety of a the company then be liable if the employee becomes ill?
person or the public” [65]. Fortunately, with the kind Might the company be expected to provide additional
of information at issue here, it is difficult to see how support or training for stress management?
the harm to be prevented could be considered Under the Americans with Disabilities Act (ADA),
“imminent”. a “disability” is a physical or mental impairment that

333
Chapter 27: Ethical issues in behavioral genetics

substantially limits one or more major life activity; Investigators are currently localizing and characteriz-
individuals can qualify as disabled based on having a ing the genetic factors involved. We are confident that
record of such impairment or being regarded as this endeavor will lead to a greater understanding of
having such impairment. The ADA prohibits (among the neurobiology of psychiatric illness as well as
other things) job discrimination against qualified indi- normal personality variation, and to radically new
viduals who have disabilities as so defined. However, and better ways of relieving the suffering caused by
the US Supreme Court has held that companies may these disorders.
refuse to employ an individual with a pre-existing However knowledge can be put to the service of
condition in a position that might worsen that condi- bad ends, and the history of the eugenics movement
tion. In that case, Chevron U.S.A., Inc. v. Echazabal demonstrates the danger of uncritical acceptance
[66], a worker who was discovered to have hepatitis of constructs that contain social value judgments.
C was fired from a maintenance job at an oil refinery There is nothing inherent in the study of behavioral
on the basis of the employer’s concern that the liver genetics that should alter our concepts of personal
disease could be worsened by exposure to toxins in responsibility or our respect for individual autonomy.
the workplace. It appears, therefore, that a company But this knowledge, if not carefully employed, can be
would be permitted to bar a vulnerable employee from misused – to marginalize unpopular individuals or
a position that would worsen (or perhaps precipitate?) groups as, perhaps, deserving of harsher punishment;
a health condition. This holding potentially would to discriminate on the job; or to increase profits by
conflict with GINA should a genetic test become offering tests of unclear benefit. These concerns are
available that would indicate, for example, heightened perhaps more salient given the complex, nuanced,
risk of psychiatric illness in the presence of job stress. and probabilistic nature of genetic information,
which is susceptible to oversimplification and errone-
Conclusion ous application – or, as H. L. Mencken is quoted as
Many behavioral traits and psychiatric illnesses are saying, “For every complex problem there is a solu-
known to have a substantial heritable component. tion that is simple, neat – and wrong”.

References 10. Cardno AG, et al. Arch Gen


Psychiatry 1999;56:162–168.
Information in Court.
Washington, DC: American
1. Galton F. Inquiries into Human Psychological Association; 1999.
Faculty and Its Development. 11. Heath A, et al. Psych Medicine
London: Macmillan and Co.; 1997;27(6):1381–1396. 21. Dinwiddie SH. AAPL Bulletin
1883. 12. Slutske WS, et al. J Abn Psychology 1994;22(3):327–342.
2. Buck v. Bell, 274 US 200 (1927). 1997;106(2):266–279. 22. Guze SB: In Botkin JR, et al. (eds.).
13. Slutske WS, et al. J Abn Psychology Genetics and Criminality: The
3. Lombardo PA. New York Law
1998;107(3):363–374. Potential Misuse of Scientific
Review 1985;60(1):50–62.
Information in Court.
4. Skinner V. Oklahoma, Ex Rel 14. Dennet DC. Elbow Room. Washington, DC: American
Williamson 316 US 535 (1942). Cambridge, MA: MIT Press; Psychological Association; 1999.
1984.
5. Nourse VF. In Reckless Hands: 23. Brunner HG, et al. Science
Skinner v. Oklahoma and the Near 15. Robinson GE, et al. Science 1993;262:578–580.
Triumph of American Eugenics. 322:896–900.
24. Caspi A, et al. Science
New York: WW Norton and Co.; 16. Fishbein DH. Criminology
2002;297:851–854.
2008. 1990;28;27–72.
6. Sofair AN, et al. Ann Intern Med 25. Dinwiddie SH. AAPL Bulletin
17. Wasserman D. J Law Med Ethics
2000;132:312–319. 1996;24(1):95–108.
2004;32;252–256.
7. Van Gestel S, et al. Mol Psychiatry 26. Salekin RT, et al. Clin Psychol – Sci
18. Durham v. US 214 F.2d 862
2003;8:840–852. Pr 1996;3(3):203–215.
(1954).
8. Sullivan PF, et al. Arch Gen 27. Dinwiddie SH. In Fishbein DH
19. US v. Brawner 471 F.2d 969 (1972).
Psychiatry 2003:60(12):1187–92. (ed.). The Science, Treatment, and
20. Summer DA. In Botkin JR, et al. Prevention of Antisocial Behaviors.
9. Kendler KS, et al. Behav Genetics (eds.). Genetics and Criminality: Kingston, NJ: Civic Research
1995;25:127–232. The Potential Misuse of Scientific Institute, Inc.; 2000.

334
Chapter 27: Ethical issues in behavioral genetics

28. Information from http://www.fbi. 42. 23AndMe, Inc. Available at 53. Poobalan AS, et al. Br J Psychiatry
gov/hq/lab/html/codis1.htm, https://www.23andme.com/. 2007;191:378–386.
accessed 11/6/08. Accessed April 5, 2012. 54. Phelan JC. Trends Neurosci 2002;
29. Graham EAM. Forensic Science, 43. Genetics and Public Policy Center. 25(8):430–431.
Medicine and Pathology Who regulates genetic tests? 55. Link BG, et al. Ann Rev Sociol
2007;3:285–288. Available at http://www. 2001;27:363–385
dnapolicy.org/policy.issue.php?
30. Nuffield Council on Bioethics. 56. Corrigan PW, et al. Schizophr Bull
Accessed April 5, 2012.
The Forensic Use of 2004;30(3):477–479.
Bioinformation: Ethical 44. Javitt GH, et al. Oklahoma Law
Issues. London: Nuffield Rev 2004;57(2):251–302. 57. American Psychiatric Association.
Council; 2007. The Principles of Medical Ethics
45. Secretary’s Advisory Committee With Annotations Especially
31. Beauchamp T: In Bloch S, et al. on Genetic Testing. Enhancing the Applicable to Psychiatry.
(eds.). Psychiatric Ethics, third Oversight of Genetic Tests. Washington, DC: American
edition. New York: Oxford Bethesda, MD: National Institutes Psychiatric Association; 2008.
University Press; 1999. of Health; 2000.
58. Pate v. Threlkel (1995), 661 So. 2d
32. Couzin J. Science 2008;322:357. 46. Hoop JG. Harva Rev Psychiatry 278 (Fla.).
2008;16(6):332–338.
33. Greendale K, et al. Am J Med Gen 59. Safer v. Pack (1996), 677 A.2d 188
2001;106:223–232. 47. Clayton EW. N Engl J Med (App. Div. NJ).
2003;349(6):562–569.
34. Graber ML, et al. Arch Intern Med 60. Petrila J. Behav Sci Law 2001;
2005;165(13):1493–1499. 48. Committee on Energy and 19:405–421.
Commerce. The Potential for
35. Quaid KA, et al. Alcohol Clin Exp 61. Wertz DC, et al. JAMA 1994;
Discrimination in Health
Res 1996;20(8):1430–1437. 272:875–880.
Insurance Based on Predictive
36. Munafo M, et al. Addict Biol Genetic Tests: Hearing Before the 62. Malpas PJ. J Med Ethics 2008;
2001;6:109–117. Subcommittee on Commerce, 34:275–278.
37. Enoch M-A, et al. Curr Psychiatry Trade and Consumer Protection of 63. Molloy v. Meier (2004), 679 N.W.
Rep 2001;3(2):144–151. the Committee on Energy and 2d 711, 714 (Minn.).
Commerce. Washington, DC: US
38. Helzer JE, et al. In Robis LN, et al. 64. President’s Commission for the
Government Printing Office;
(eds.). Psychiatric Disorders in Study of Ethical Problems in
2001.
America. New York: Free Press; Medicine and Biomedical and
1991. 49. Anderlik MR, et al. Annu Rev Behavioral Research. Making
Genomics Hum Genet 2001; Health Care Decisions.
39. Psynomics. https://psynomics.
2:401–433. Washington, DC: US Government
com/. Accessed April 5,
2012. 50. Lapham EV, et al. Science Printing Office; 1982.
1996;274(5287):621–624. 65. Department of Health and
40. Couzin J. Science 2008;319
(5861):274–277. 51. Lerman C, et al. J Consult Clin Human Services. Federal Register
Psychol 2002;70(3):784–797. 2000;65(250):82813–82817.
41. SureGene, LLC. Available at:
http://www.suregene.net. 52. Burke L. Int Rev Psychiatry 66. Chevron USA, Inc. v. Echazabal
Accessed April 5, 2012. 2003;15(3):243–255. (2002), 536 US Supreme Court 73.

335
Genetics of Tourette syndrome
Chapter

28 and related disorders


Maria G. Motlagh, Thomas V. Fernandez, and James F. Leckman

Introduction Over the past 25 years, TS has emerged as a model


developmental disorder at the interface of neurology
Tourette syndrome (TS) is a developmental neurop- and psychiatry. The identification of abnormalities
sychiatric disorder of childhood onset. TS is a involving the basal ganglia in neuropathological
chronic, potentially disabling condition characterized [6, 7] and neuroimaging studies [8, 9], the possibility
by the presence of both motor and vocal tics. Tics of a post-infectious form of the disorder [10, 11] and
are characterized by brief, stereotypical but non- the growing appreciation of the impact of environ-
rhythmic movements and vocalizations. Common mental, neurobiological, and developmental factors in
tics include eye blinking, grimacing, jaw, neck, disease expression have all contributed to making TS
shoulder or limb movements, sniffing, grunting, a model disorder for understanding developmental
chirping, or throat clearing. In the natural history psychopathology more broadly [1, 12].
of TS, motor tics often begin between the ages of 3 Three decades of research has led to widespread
and 8, several years before the appearance of vocal agreement that genes play a central role in the eti-
tics. Tics typically follow a waxing and waning ology of this disorder. TS and related conditions not
course [1, 2]. Tic severity usually peaks early during only aggregate within families but cluster in patterns
the second decade of life with many patients showing that suggest a high degree of heritability. Moreover,
a marked reduction in severity by the end of adoles- high rates of comorbid obsessive–compulsive dis-
cence [3, 4]. Only 20% or less of individuals with TS order (OCD) and attention-deficit hyperactivity dis-
continue to experience a moderate level of impair- order (ADHD) appear to be accounted for, at least in
ment of global functioning by the age of 20 years [3]. part, by a common genetic diathesis. Yet perhaps the
However, tic disorders that persist into adulthood most important conclusion from nearly 30 years of
can be associated with the most severe symptoms work in the field is that the pathogenesis of TS
including episodes of self-injurious motor tics (sec- involves the reciprocal interaction of genes and envir-
ondary to hitting or biting) or socially stigmatizing onment. There has been increasing excitement and
coprolalic utterances (e.g. shouting obscenities or anticipation as an international consortium and other
racial slurs). major laboratories have begun to collect the pheno-
The characterization of tics as intermittent trains typic and genetic data needed to take full advantage of
of involuntary motor discharge is incomplete. Many the accelerating advances in the genomic sciences,
tics are under partial voluntary control, evidenced by from gene microarray studies, to genome-wide asso-
patients’ capacity to suppress them for brief periods ciation studies, to the study of noncoding RNA and
of time. A key feature of tics is that they are frequently gene copy number variation.
associated with antecedent sensory phenomena, This review summarizes the data suggesting that
including either a general sense of inner tension or the liability to TS and related disorders is heritable. It
focal “premonitory urges” or both. These urges are will then review nearly three decades of efforts to
often experienced as nearly irresistible, can be a major identify the specific genes involved and chronicle the
source of impairment, and likely reflect an underlying shifting strategies. First, we will examine the evidence
deficit in sensorimotor gating [5]. for the general nature and extent of genetic risk in TS

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

336
Chapter 28: Genetics of Tourette syndrome

and comorbidities. Next, we will briefly examine the their DNA) [15–17]. MZ concordance rates have been
results of traditional parametric linkage analyses, both shown to be approximately 53–56% versus less than
in extended families and within genetic isolates that 10% in DZ twins [16]. When study methodology has
likely share the same subset of vulnerability genes allowed for direct patient examination and included
by descent. Here we focus on the apparent role of the diagnosis of chronic tics (CT) in addition to TS,
L-histidine decarboxylase (HDC) in a two-generation MZ concordance has been shown to approximate
family with nine affected individuals. Third, we con- 100% [17].
sider the results and future promise of nonparametric In addition to family and twin studies, segregation
and other “model-free” methods of gene identifica- analyses provide circumstantial evidence for the role
tion. Fourth, we consider the merits of cytogenetic of genes in TS. These studies examine pedigrees with
methods to search for rare genetic variants based on affected individuals and compare the actual patterns
chromosomal rearrangements, with particular atten- of transmission to hypothetical patterns that would be
tion to Slit and Trk-like family member 1 (SLITRK1). expected under varying modes of genetic transmis-
Fifth, we summarize the investigations of various sion. Using this methodology, several studies have
candidate genes. Sixth, we review some of the putative suggested that TS is transmitted in an autosomal
environmental risk factors which may contribute dominant fashion with partial penetrance [18–20].
toward the development of TS. Finally, we consider Alternatively, Walkup et al. [21] found that TS was
the future prospects for TS genetic research and close most likely the result of a gene of major effect confer-
with a discussion of the clinical and research implica- ring more than half of the overall risk for the dis-
tions of the increasing understanding of the genetic order, with the remainder accounted for by genetic
underpinnings of this syndrome. background and environmental factors. Hasstedt et al.
[22] completed a segregation analysis incorporating
assortative mating in a single large pedigree contain-
TS is a heritable condition ing 182 members. The analysis provided evidence of a
Multiple lines of evidence indicate that TS is an major locus with an intermediate inheritance pattern
inherited disorder. The first and most important piece with a significant assortative mating correlation.
of evidence is that TS and related tic disorders tend to However, when assortative mating was not included
aggregate across multiple generations within families. in the model, intermediate inheritance was not
The risk of a first-degree relative of someone with TS inferred. In addition, other studies have been unable
also having TS or a lesser variant is substantially to demonstrate a convincing Mendelian pattern of
greater than the risk to the general population. For inheritance, even taking into account TS spectrum
instance, the frequency with which siblings are phenotypes [23]. In sum, although a consistent pic-
affected has consistently been found to be on the ture has emerged from nearly three decades of
order of 5% for sisters and more than 10% for broth- research which strongly suggest that genes play a
ers [12]. Judged against the best available epidemi- major role in etiology of TS, the exact mode of inher-
ological data [13, 14], this represents at least a 10-fold itance remains elusive. As believed for other neurop-
increase in risk for first-degree relatives compared to sychiatric disorders, TS, in most instances, will likely
the overall prevalence of the disorder. be found to be the result of multiple interacting gene
The evidence for familial aggregation of TS, while alleles, each with relatively small contributions
suggestive, does not itself confirm a role for genes in compared to the genetic effects observed in Mendel-
the etiology of the disorder. There are a variety of ian disorders.
reasons why the syndrome might run in families,
including an infectious etiology or a clustering of
other nongenetic (e.g. environmental) risk factors.
Twin studies complement heritability data, providing
Evidence for genetic relationships between
direct evidence concerning the relative contributions TS and other psychiatric disorders
of inherited versus environmental factors. Several When Georges Gilles de la Tourette published his
studies have shown that monozygotic (MZ) twins first descriptions of TS more than a century ago
are far more likely to also suffer from TS than dizy- (1885), he made several prescient observations: first,
gotic (DZ) twins (who share on average just 50% of that tics and TS aggregate in families; and second,

337
Chapter 28: Genetics of Tourette syndrome

that patients with the disorder also suffer from The issue of whether ADHD and TS may be
obsessions and compulsions. In fact, clinicians variable expressions of a shared genotype has been
familiar with TS have consistently noted a wide controversial [36–38]. Although there is a substantial
range of psychiatric comorbidity in their patients. body of data indicating that these disorders are trans-
Indeed, simple and transient tics in the absence of mitted independently within families, the large
comorbid conditions are common and occur in at comorbidity rate leaves open the possibility that in
least 5% of children [13]. In clinical and population- some instances TS and ADHD may share a common
based samples, TS alone is the exception rather than genetic vulnerability. Indeed, a recent latent class
the rule [22]. analysis (LCA) of 952 individuals from 222 TS fam-
In family studies, as many as 30–60% of TS ilies indicated that individuals with TS, OCD, and
probands will meet diagnostic criteria for OCD. ADHD had the most heritable form of the disorder
These rates compare to population frequencies for [39]. In addition to LCAs, there have been several
this disorder between 2–3% [20, 24]. In addition, factor analytic studies that have sought to refine her-
when relatives of TS patients are examined, there is itable phenotypes [40–43].
a significant increase in the risk for OCD in add- The co-occurrence of depression and anxiety
ition to the increased risk for tic disorders. How- symptoms with TS also may reflect either shared
ever, such an increase in the risk for a disorder genetic vulnerability or the cumulative psychosocial
among relatives does not necessarily imply an alter- burden of having tics or other shared biological diath-
native expression of a single genetic diathesis. The eses [31]. A co-occurrence with autism has also been
same finding could be explained by comorbidity reported. Indeed, among autistic subjects the preva-
due to common pathophysiological (but not gen- lence of TS has been reported to be 6.2%, about 10
etic) pathways, or the clustering of environmental times the prevalence in the general population
risk factors. However, multiple studies have found [44, 45]. Furthermore, a recent study of genomic copy
that relatives of patients with TS (but without OC number variation (CNV) in TS showed a significant
symptoms) have higher rates of OCD (without tics) amount of overlap among rare CNVs in TS and
than would be found in unaffected families [17, 20, autism spectrum disorders (ASD), but not schizo-
24–26]. Conversely, family studies of OCD pro- phrenia or intellectual disability [46].
bands have consistently observed elevated rates of
tic disorders [26–28]. In addition, several clinical
case series have documented that individuals with a Parametric linkage analysis
tic-related form of OCD are more likely to report Once family and twin studies have demonstrated that
obsessions of symmetry and exactness and a need genetic factors are likely to play a role in the patho-
to redo activities to achieve a sense of completion genesis of a disorder, there are several different means
or a sense of things looking, feeling or sounding toward identification of specific genetic loci involved.
“just right” [29, 30]. In sum, the high rates of OCD Studies suggesting an autosomal dominant mode of
in TS patients and their families, the patterns with transmission made TS a clear candidate for paramet-
which obsessive–compulsive symptoms are ric linkage analysis. However, with one exception,
expressed in TS pedigrees, and the presence of a parametric linkage studies over three decades have
distinctive natural history in tic-related OCD all failed to identify and confirm a specific genetic locus
suggest that in some families a shared genetic involved in the etiology of this disorder.
diathesis may represent itself as either TS, OCD, L-histidine decarboxylase (HDC). The one excep-
or both disorders. tion is the report by Ercan-Sencicek et al. [47] that
ADHD is frequently diagnosed in children with described a two-generation family with nine affected
TS, with a prevalence as high as 60–70% [6, 31]. This members with TS. The logarithm of odds (LOD)
co-occurrence of TS and ADHD can be associated score of 2.05 for a region on chromosome 15
with disruptive behaviors such as aggression, explo- approximated the maximum theoretical LOD score
sive behavior, low frustration tolerance, and noncom- given a model of dominant transmission. They iden-
pliance. When comorbid ADHD is present, it is tified a heterozygous G-to-A transition at nucleotide
frequently associated with academic difficulties, peer position 951 in exon 9 of the HDC gene, resulting in
rejection, and family conflict [32–35]. a W317X substitution, predicted to result in a

338
Chapter 28: Genetics of Tourette syndrome

truncated protein lacking key segments of the active


domain. Studies of mRNA from patient cells indi-
Nonparametric linkage analysis
With a few exceptions, there is a growing recognition
cated that the mutation escaped nonsense-mediated
that TS is unlikely to be inherited in a predominantly
decay. The mutation was not found in 3000 control
Mendelian fashion. As a result, researchers have
chromosomes from northern and western Europe.
favored the use of nonparametric approaches. Such
In vitro studies in Escherichia coli indicated that the
investigations do not require the specification of a
mutant protein acted in a dominant-negative
hypothesis regarding the mode or character of inherit-
manner, resulting in loss of enzyme activity. Of the
ance. An initial affected sibling pair analysis by The
nine affected individuals, four also had OCD and
Tourette Syndrome International Consortium for
one also had Asperger syndrome. HDC is the rate-
Genetics [59] identified two regions of the genome
limiting enzyme in histamine (HA) biosynthesis sug-
with evidence suggestive of linkage, one on chromo-
gesting that histaminergic neurotransmission is
some 4q and another on chromosome 8p, that
involved in the pathobiology of TS in this family
achieved LOD scores of greater than 2. The same
and perhaps more generally. HA signaling in the
consortium more recently reported the results of the
central nervous system is mediated by four
largest TS genetic linkage study yet undertaken. The
G protein-coupled receptors, located both presynap-
sample included 238 nuclear families yielding 304
tically (predominantly H3 as well as H4) and post-
independent sibling pairs and 18 separate multigenera-
synaptically (H1–H3). Presynaptic HA receptors
tional families, totaling 2040 individuals [60]. Suggest-
regulate not only the release of HA, but also a variety
ive evidence of linkage was observed for a region on
of other neurotransmitters, including dopamine. Sev-
chromosome 2p in the analyses that included individ-
eral lines of evidence suggest that HA acts in a
uals with TS or CT disorder as affected. Of note,
counter-regulatory fashion, with increased HA
neither of the two earlier reported sites on chromo-
resulting in decreased DA signaling and vice versa
some 4q or 8p was confirmed despite the fact that the
[48, 49]. H2 and H3 receptors are enriched in the
same families were a subset of the larger study.
striatum and cortex, regions of the brain implicated
Regions on chromosome 17q have also provided
in TS [50], and studies of rodents with decreased
suggestive evidence for linkage [61, 62]. For example,
brain HA show increased sensitivity to stereotypies
an initial scan of chromosome 17 in two large pedi-
when administered DA agonists [51]. The finding of
grees by Paschou et al. [61] provided a nonparametric
a loss of function of HDC has the potential to lead to
LOD score of 2.41. Fine mapping with 17 additional
the development of animal models and eventually
microsatellite markers increased the LOD score to
the development of novel therapeutics for TS.
2.61. Genotyping data from 25 single nucleotide poly-
Other than this important finding, assuming auto-
morphisms (SNPs) within 3 candidate genes in the
somal dominant transmission and genetic homogen-
region of interest on 17q was obtained from the
eity, at least 90% of the remaining genome has been
original families plus additional independent families
excluded [21, 52–58]. In light of the substantial body
with one or two affected children. These data yielded
of evidence supporting a major genetic contribution
several SNPs and three-marker haplotypes signifi-
to TS, these initial linkage results suggest that genes
cantly associated with TS.
conferring the phenotype are positioned in the
remaining 10% of the genome or that the methods
by which investigators have sought to identify these Population isolates and special
loci have been flawed or that the view of TS as a
unitary genetic entity is erroneous. populations
Historically, the failure to discover a significant TS Well-powered genome-wide association studies
genetic locus using a traditional parametric linkage have proven to be invaluable in identifying vulner-
strategy has lead to a number of alternative ability genes for a number of complex traits, but no
approaches including nonparametric linkage analyses, such study has yet been undertaken for TS. Efforts
the examination of population isolates, cytogenetic toward such studies are underway and investigators
methods to search for rare genetic variants, and asso- have so far focused genome-wide association efforts
ciation studies, including candidate gene studies and on genetically isolated populations. Examining
family-based tests of association. population isolates restricts the number of potential

339
Chapter 28: Genetics of Tourette syndrome

vulnerability genes, as linkage disequilibrium families including sites at 2p12, 3p21.3, 7q35–36,
should extend for much greater distances than in 8q21.4, 9pter, 13q31, and 18q22.3 [69–72].
outbred populations. For example, Simonic et al. Boghosian-Sell et al. [68] described a family where
[63, 64] completed an early genome-wide associ- TS was segregating with a balanced 7;18 translocation.
ation study for TS in a sample from the Afrikaner The breakpoint on chromosome 7 was mapped to
population of South Africa. This population was within chromosomal bands 7q22 and 7q31. Subse-
founded in the seventeenth century with some quently, Kroisel et al. [73] described a de novo dupli-
admixture from English, south Asian, and indigen- cation on chromosome 7 (7q22.1–31.1) observed in a
ous populations. The authors found evidence of an 13-year-old boy with TS, moderate mental retard-
association of TS to markers on chromosomes 2p, ation, and minor physical anomalies. Both break-
8q, and 11q. Subsequently, Merette et al. [65] geno- points were within or close to the breakpoint region
typed the 24 markers that achieved a nominally described by Boghosian-Sell et al. [68], suggesting that
significant association in the Afrikaners in a large a gene located at or near this region may be involved
kindred from an isolate of French Canadians in in the pathogenesis of TS in some cases. Further
Quebec and found suggestive evidence of linkage molecular analysis of this same proband by Petek
on chromosome 11 (11q23). Several additional et al. [70] revealed that his de novo abnormality was
genome-wide association studies of TS are under- an inverted duplication which resulted in disruption
way in samples that are now available from several of the IMMP2L gene, a human homologue of the
isolated populations in Costa Rica, Antioquia yeast mitochondrial inner membrane peptidase sub-
[Colombia], Quebec, and Ashkenazi Jewish unit 2. More recently, this research group screened 39
diaspora. TS patients and 95 multiplex autistic disorder (AD)
The contribution of recessive loci as a mechanism families (due to the localization of IMMP2L in the
of genetic vulnerability in TS with reduced penetrance critical region for an AD candidate locus on chromo-
is one possibility that has been difficult to explore. some 7q, AUTS1), for sequence and CNV in IMMP2L
Homozygosity mapping has been successfully used to [74]. No coding mutations were found in either TS or
detect recessive loci within populations with high AD patients and expression studies provided no evi-
rates of consanguinity. Using this technique, even dence of parental imprinting at this gene locus, sug-
quite small inbred families can be informative due gesting that IMMP2L may not be a common
to autozygosity in which the two alleles at an auto- etiological factor in either disorder.
somal locus are identical by descent (i.e. copies of a State et al. [71] reported on a young man with
single ancestral gene). Motlagh et al. [66] identified 12 CT and OCD who was found to carry a paracentric
consanguineous Iranian families with TS. Remark- inversion involving chromosome 18q22. This team
ably, these families presented with an unusual natural mapped the telomeric end of the inversion to a
history characterized by the early onset of vocal tics genomic location that was within 1 Mb of a trans-
and coprolalia and frequent comorbidity with OCD. location breakpoint previously described by Bogho-
Genotyping the affected and unaffected members of sian-Sell et al. [68]. Although no genes were
these pedigrees has the potential to identify rare reces- structurally disrupted by this inversion, functional
sive contributions to this disorder. studies of two transcripts in the region showed rep-
lication timing dysregulation. This finding suggested
the possibility that epigenetic influences could affect
Cytogenetic abnormalities gene expression and influence phenotype, and left
Cytogenetic techniques such as karyotyping, fluores- open the possibility that genes in this region may
cence in situ hybridization (FISH), and array com- serve as candidates in TS.
parative genomic hybridization (aCGH), have been Verkerk and colleagues [72] reported a shared
used to identify patients with chromosomal abnor- complex rearrangement in a father with TS and his
malities (i.e. translocations, deletions, duplications). affected children, disrupting the contactin-associated
Identifying genes at or near the disrupted chromo- protein 2 gene (CNTNAP2) on chromosome 7q35.
somal regions may flag potential susceptibility genes This gene encodes a membrane protein located at
that warrant further investigation. A number of cyto- nodes of Ranvier of axons that may be important
genetic abnormalities have been reported in TS for the distribution of potassium channels, which

340
Chapter 28: Genetics of Tourette syndrome

would affect signal conduction along myelinated quite uncommon, functional mutations in this gene
neurons. However, Belloso et al. [75] recently have been identified in patients with TS, trichotillo-
described a familial balanced reciprocal translocation mania, and ADHD [67, 76, 77]. SLITRK1 contains
t(7;15)(q35;q26.1) in phenotypically normal individ- one exon and encodes 696 amino acids. There are
uals. In this family the 7q35 breakpoint disrupted six known members in the SLIT and TRK-like gene
CNTNAP2, indicating that disruption of this gene family, SLITRK1 is unique in that it lacks tyrosine
does not necessarily lead to TS. phosphorylation sites in its short intracellular domain
Cuker et al. [69] reported a 14-year-old girl with [78]. SLITRK family proteins are characterized as
severe OCD and a CT disorder with a t(2;18)(p12; integral membrane proteins that have two leucine-
q22) translocation. The patient’s chromosome 18 rich repeat (LRR) domains and a carboxy-terminal
breakpoint localized to the same chromosomal band domain that is partially similar to trk neurotrophin
as two previously reported rearrangements associated receptor proteins [79]. SLITRK1 demonstrates a
with TS, OCD, and CT disorder [68, 71], and mapped developmentally regulated pattern of expression
to a genomic position approximately 5 Mb from these within cortical and striatal structures, particularly
rearrangements. The clustering of these three break- within the striosomal compartment and the so-called
points within a relatively small genetic interval sug- direct pathway of the CTSC loop, characterized at the
gests that 18q22 may be a promising region for molecular level in part by dopamine D1 receptors.
containing a gene or genes of etiological importance A number of studies [80–85] have sought to repli-
in the development of the TS/OCD phenotypic cate the findings reported in Abelson et al. [67] and
spectrum. have identified var321 of SLITRK1 in a total of nine
Slit and Trk-like family member 1 (SLITRK1). TS families, with the variant segregating along with
Abelson et al. [67] identified and mapped a de novo affected status in five of these families. Concerns that
chromosome 13 inversion in a patient with TS. The the original SLITRK1 findings may have been errone-
gene designated as Slit and Trk-like family member 1 ous as a result of occult ethnic differences in cases
(SLITRK1) was identified as a brain expressed candi- versus controls (population stratification) have not
date gene mapping approximately 350 kb from the been supported by recent evidence [76].
13q31 breakpoint. Mutation screening of 174 unre-
lated TS patients of European ancestry identified one
SLITRK1 identification of a truncating frame-shift Candidate gene studies
mutation in a second family affected with TS. In Current theories of the pathogenesis of TS have
addition, two patients were identified with a rare guided the selection of several candidate genes for
variant (var321) in a highly conserved region of the association testing in individuals with TS. These
30 untranslated region (3’UTR) of this gene, corres- candidate genes have included various dopamine
ponding to a brain expressed micro-RNA (miRNA) receptor genes (DRD2, DRD3, and DRD4 [86–89]),
binding domain. miRNAs are short, 20–22 bases, the dopamine transporter (DAT [90]), catecol-O-
noncoding RNAs that typically suppress translation methyltransferase (COMT [89, 91]), three noradre-
and destabilize messenger RNAs that bear comple- nergic receptor genes (ADRA1C, ADRA2A, and
mentary target sequences. Many miRNAs are ADRA2C [92, 93]), dopamine b-hydrolyase (DBH
expressed in a tissue-specific manner and may con- [90]), monoamine oxidase A (MAOA [87]), and a
tribute to the maintenance of cellular identity. In vitro few serotonergic genes including tryptophan hydox-
studies showed that both the frame-shift and the ylase 2 (TPH2), the serotonin receptor 3 (5-HT3),
miRNA binding site variants had functional potential and the serotonin transporter (5-HTTLR) [91, 94–
and were consistent with a loss-of-function mechan- 96]. Several additional candidate genes involved in
ism. Studies of both SLITRK1 and the miRNA pre- neuronal development, neuroendocrine, and
dicted to bind in the variant-containing 30 region immunological function have also been studied with
showed expression in basal ganglia and deep layers negative results [90, 97–100]. Genetic variation at
of cortex in both mouse and human. any one of these loci is unlikely to be a major source
As a result of the Abelson et al. [67] study, of vulnerability to TS, but in concert, certain alleles
SLITRK1 has emerged as a strong candidate gene for could have cumulative effects and contribute to
rare cases of TS and related conditions. Although phenotypic variability.

341
Chapter 28: Genetics of Tourette syndrome

1800s [111] and has recently become an intense and


Environmental factors: perinatal controversial area of research [112]. It is well estab-
events, psychosocial stress, infection, lished that group A beta hemolytic streptococci
(GABHS) can trigger immune-mediated disease in
and immune response genetically predisposed individuals. Rheumatic fever
A number of environmental factors have been impli- is characterized by inflammatory lesions involving the
cated in the pathogenesis of TS including psycho- joints, heart, and/or central nervous system. The cen-
social stress, gestational and perinatal insults, tral nervous system manifestations are referred to as
exposure to androgens, heat and fatigue, and post- Sydenham’s chorea (SC). In addition to chorea, some
infectious autoimmune mechanisms. For example, SC patients display motor and phonic tics as well as
perinatal hypoxic/ischemic events appear to increase OCD and ADHD symptoms, suggesting the possibil-
the risk of developing TS [14, 101, 102], and one ity that at least in some instances these disorders share
recent retrospective study added prenatal maternal a common etiology [113]. Case reports have also
smoking as a risk factor for TS [103]. Maternal stress implicated other infectious processes in TS etiology
during pregnancy may also influence the individual’s including Lyme disease [114] as well as mycoplasma
tic severity in adolescence [104]. pneumonia [115].
Male sex is a risk factor for TS. While this could be In 1998, Swedo and colleagues proposed that Pedi-
understood by genetic mechanisms, frequent male-to- atric Autoimmune Neuropsychiatric Disorder Asso-
male transmissions within families appear to rule out ciated with Streptococcal infection (PANDAS)
the presence of a predominant X-linked vulnerability represents a distinct clinical entity and includes some
gene. The increased prevalence of TS in males has led cases of TS and OCD. In PANDAS, it is postulated
to the hypothesis that the presence of androgenic that although GABHS is the initial autoimmunity-
steroids during critical periods in fetal development inciting event, viruses, other bacteria, or even nonin-
may play a role in the later development of the illness fectious immunological responses are capable of trig-
[105]. This notion is supported by the observation gering subsequent symptom exacerbations via
that gender-related behaviors in children and adults molecular mimicry, such that antibodies directed
with TS are similar to such behaviors in children with against GABHS attack cells in the brain because of a
known elevated prenatal androgens, and these behav- similar structure [116].
iors correlate with tic severity [106]. While these
effects may be due to androgenic steroids expressed
early in development, it is likely that there are sex- Future directions for TS genetics
specific patterns of gene expression in male versus In this final section, we briefly consider a few of the
female brains that influence their differentiation and emerging areas of great promise in TS genetics,
function [107]. including the availability of increasingly high-reso-
Patients with TS report higher levels of psycho- lution DNA microarrays, expanded use of genome-
social stress, and latent class modeling of prospective wide association studies, the importance of noncod-
longitudinal data indicate that antecedent stresses can ing RNAs, gene copy number variation, and epigen-
increase future tic and obsessive–compulsive symp- etic programming.
tom severity [108]. Microarray studies and the future of genome-
Temperature dysregulation involving some wide association studies. Since the complete sequen-
change in hypothalamic function has also been pro- cing of the human genome, it is now possible to
posed as a factor in the pathobiology of some individ- genotype millions of SNPs and monitor expression
uals with TS [109]. In a case series [110], an increase levels of thousands of genes simultaneously using
in ambient temperature as well as core body tempera- DNA microarrays. Gene expression profiling of per-
ture was associated with a transient increase in tics in ipheral blood from a small number of TS patients has
some patients. This increase in tics was correlated led to the preliminary identification of a subset of
with patients’ local sweat rate – via a dopamine medi- cases in which altered expression of an interrelated
ated pathway in the hypothalamus. set of immune genes discriminated between TS and
Speculation concerning a post-infectious auto- age- and gender-matched controls that included indi-
immune etiology for TS and OCD dates from the late viduals with a variety of other neuropsychiatric

342
Chapter 28: Genetics of Tourette syndrome

disorders [117–121]. Many of these genes are associ- in the tonically active cholinergic interneurons that
ated with natural killer (NK) cell function. NK cells regulate the activity of the fast spiking GABAergic
are a type of cytotoxic lymphocytes that constitute a interneurons in the striatum [1, 125].
major component of the innate immune system. If Copy number variation. The emergence of
replicated, further work needs to be done to assess the microarray technologies that can detect sub-micro-
heritability of these NK gene variants and expression scopic structural variation revealed extensive copy
profiles within TS families. Longitudinal studies will number variation (CNV) across the human genome
also be important to determine whether the expres- [126, 127] and provided opportunities for genome-
sion of any of these genes is associated with fluctu- wide assessment of rare variation. Studies in schizo-
ations in symptom severity. phrenia [128–134] and ASD [135–138] demonstrated
Work using DNA microarrays and mRNA expres- an over-representation of rare CNVs, particularly
sion in key brain regions has begun, and in the future genic de novo variants [131, 135, 136, 139], and high-
this method will permit investigators to use hundreds lighted molecular mechanisms that likely play a role
or perhaps thousands of susceptibility genes to assess in these conditions. Fernandez et al. [46] conducted a
an individual’s genetic vulnerability and to explore genome-wide analysis of rare (< 1% frequency)
developmental change and continuity, comorbidity CNVs in 460 individuals with TS, including 148
with other disorders, gene–gene, and gene– parent–child trios and 1131 matched controls. The
environment interactions [122]. results support recent findings implicating histami-
Noncoding RNAs. Another recent discovery with nergic neurotransmission in the etiology or modula-
far-reaching implications for future genetic research tion of tics and highlight the potential involvement of
in TS is the importance of noncoding RNAs. The GABAergic mechanisms as well. In addition, the
discovery that one of the rare variants of SLITRK1 results reinforce the notion of shared genetic risks
alters the binding region for a miRNA [67] has among ASD and TS, and identify three novel, large,
focused the field’s attention on the potentially rare, genic, de novo CNVs that are likely carrying risk
important role of noncoding RNAs in the genesis in the individuals in which they were identified, based
of TS and related disorders. Remarkably, recent stud- on their de novo status and high gene content relative
ies suggest that tissue-specific miRNAs may function to controls.
at multiple hierarchical stages within gene regulatory Epigenetic programming. Epigenetics is another
networks, from targeting hundreds of effector genes emerging field of potential promise for understanding
to controlling the levels of key transcription factors the origins of TS. Epigenetic programming is a fun-
[123]. This multilevel regulation may permit individ- damental part of eukaryotic biology, involving the
ual miRNAs to profoundly affect the gene expression modification of DNA and the chromatin proteins that
program of differentiated cells. It is possible that associate with it during key periods of development.
specific miRNAs could alter cellular identities. Frequently these modifications have enduring effects
Although unlikely, it is conceivable that variant miR- on gene expression in key brain regions. In addition,
NAs could affect the number of fast spiking some epigenetic alterations can be passed from one
GABAergic interneurons in the striatum of individ- generation to the next. For example, there is now
uals who go on to develop TS [6]. Specifically, the compelling data for the presence of developmental
number of fast spiking GABAergic interneurons in windows during which the genetically determined
the striatum in an animal model of idiopathic dysto- microcircuitry of key limbic–hypothalamic–midbrain
nia appears to decrease and then return to normal structures are susceptible to early environmental
levels over the course of development [124]. The influences and that these influences powerfully shape
changes in the number of these interneurons in the an individual’s responsivity to psychosocial stressors
striatum closely parallel the severity of their move- and their capacity to parent the next generation [140].
ment disorder. The molecular mechanisms under- These early environmental influences have been
lying this change in interneuron number are shown to alter the pattern of methylation of the pro-
unknown but might include cellular reprogramming moter region of the glucocortoid receptor gene in the
mediated in part by miRNAs. It is also of interest to hippocampus, stably altering the level of glucocortoid
note that the one area in the adult brain where receptor gene expression and hypothalamic-pituitary-
SLITRK1 continues to be expressed in adulthood is adrenal (HPA) responses to stress. This finding could

343
Chapter 28: Genetics of Tourette syndrome

be relevant to the observation that high levels of analyses, and association studies, as well as the study
maternal stress during pregnancy are associated with of rare genetic variants. The currently available data
a more severe form of TS [104]. Cross-sectional and suggest that, with rare exceptions, TS is typically a
longitudinal studies of TS and early-onset OCD have multidimensional, polygenic disorder in which the
consistently suggested that these disorders are sensi- effects of individual genes are much smaller than
tive to psychosocial stress [108], and that TS patients previously believed. This implies that in a majority
show a heightened stress response via the HPA axis of cases, TS vulnerability is determined by the com-
[141, 142] and higher levels of cerebrospinal fluid bined effect of multiple genes interacting with specific
(CSF) corticotropin releasing hormone [143]. These environmental risk factors, such as maternal stress
findings are consistent with an epigenetic effect of the and maternal smoking during pregnancy, as well as
sort reported by Meaney and colleagues. In addition inflammatory processes and psychosocial stressors
to the evidence summarized above regarding gene affecting the child. We believe that there is reason
expression changes of the glucocorticoid receptor for considerable optimism, as much of the conven-
gene in the hippocampus, it also appears that the tional wisdom about psychiatric genetics needs to be
expression of the estrogen receptor alpha gene in the re-examined in light of recent discoveries such as the
hypothalamus is under epigenetic control. If such importance of CNV and the effect of environment on
changes occur in two genes, it seems possible that DNA methylation and gene expression. One may
the expression of multiple other central nervous anticipate that the fast pace of genetic discoveries will
system and immune genes relevant to TS can be continue and will increasingly affect research in all of
altered through epigenetic programming. medicine and psychiatry, especially developmental
neuropsychiatry.
Conclusion
The last three decades of research have demonstrated Acknowledgements
that TS is more genetically heterogeneous than ini- Portions of the research described in this review
tially thought. Important progress toward the under- were supported by grants from the National Insti-
standing of genetic influences in TS has been made by tutes of Health (NIH): MH49351, MH30929, and
the combination of family and twin studies, segrega- K05MH076273, and by the Tourette Syndrome
tion analyses, parametric and nonparametric linkage Association.

References 9. Bloch MH, et al. Neurology


2005;65(8):1253–1258.
19. Curtis D, et al. Br J Psychiatry
1992;160:845–849.
1. Leckman JF. Lancet 2002;
360(9345):1577–1586. 10. Mell LK, et al. Pediatrics 2005; 20. Pauls DL, et al. N Engl J Med
116(1):56–60. 1986;315(16):993–997.
2. Lin H, et al. J Am Acad Child
Adolesc Psychiatry 2002; 11. Swedo SE, et al. Am J Psychiatry 21. Walkup JT, et al. Am J Hum Genet
41(9):1070–1077. 1998;155(2):264–271. 1996;59(3):684–693.
3. Bloch MH, et al. Arch Pediatr 12. Leckman JF, et al. J Child Neurol 22. Hasstedt SJ, et al. Am J Hum Genet
Adolesc Med 2006;160(1):65–69. 2006;21(8):642–649. 1995;57(3):682–689.
4. Leckman JF, et al. Pediatrics 13. Khalifa N, et al. Dev Med Child 23. Seuchter SA, et al. Genet
1998;102(1 Pt 1):14–19. Neurol 2003;45(5):315–319. Epidemiol 2000;18(1):33–47.
5. Swerdlow NR, et al. 14. Khalifa N, et al. Acta Paediatr 24. Hebebrand J, et al. J Psychiatr Res
Psychopharmacology (Berl) 2005;94(11):1608–1614. 1997;31(5):519–530.
2006;186(2):246–254. 15. Hyde TM, et al. Neurology 1992; 25. Pauls DL, et al. Am J Hum Genet
6. Kalanithi PS, et al. Proc Natl Acad 42(3 Pt 1):652–658. 1991;48(1):154–1630.
Sci U S A 2005;102(37): 16. Price RA, et al. Arch Gen 26. Rosario-Campos MC, et al. Am
13307–13312. Psychiatry 1985;42(8):815–820. J Med Genet B Neuropsychiatr
7. Kataoka Y, et al. J Comp Neurol 17. Walkup JT, et al. Psychopharmacol Genet 2005;136B(1):92–97.
2010;518(3):277–91. Bull 1988;24(3):375–379. 27. Nestadt G, et al. Arch Gen
8. Peterson BS, et al. Arch Gen 18. Baron M, et al. Am J Hum Genet Psychiatry 2000;57(4):
Psychiatry 2003;60(4):415–424. 1981;33(5):767–775. 358–363.

344
Chapter 28: Genetics of Tourette syndrome

28. Pauls DL, et al. Am J Psychiatry 48. Ferrada C, et al. 68. Boghosian-Sell L, et al. Am J Hum
1995;152(1):76–84. Neuropharmacology Genet 1996;59(5):999–1005.
29. Kwak C, et al. Mov Disord 2003; 2008;55:190–197. 69. Cuker A, et al. Am J Med Genet A
18(12):1530–1533. 49. Munzar P, et al. 2004;130A(1):37–39.
30. Woods DW, et al. J Dev Behav Neuropsychopharmacology 70. Petek E, et al. Am J Hum Genet
Pediatr 2005;26(6):397–403. 2004;29:705–717. 2001;68(4):848–858.
31. Coffey BJ, et al. J Am Acad Child 50. Haas HL, et al. Physiol Rev 71. State MW, et al. Proc Natl Acad
Adolesc Psychiatry 2000; 2008;88:1183–1241. Sci U S A 2003;100(8):4684–4689.
39(5):562–568. 51. Kubota Y, et al. J Neurochem 72. Verkerk AJ, et al. Genomics
32. Carter AS, et al. J Child Psychol 2002;83:837–8451. 2003;82(1):1–9.
Psychiatry 2000;41(2):215–223. 52. Barr CL, et al. Am J Med Genet 73. Kroisel PM, et al. Am J Med Genet
33. Hoekstra PJ, et al. J Clin Psychiatry 1999;20:437–445. 2001;101(3):259–261.
2004;65(3):426–431. 53. Curtis D, et al. Psychiatr Genet 74. Petek E, et al. Mol Genet Genomics
34. Peterson BS, et al. J Am Acad 2004;14(2):83–87. 2007;277(1):71–81.
Child Adolesc Psychiatry 2001;40 54. Heutink P, et al. Adv Neurol 75. Belloso JM, et al. Eur J Hum Genet
(6):685–695. 1992;58:167–172. 2007;15(6):711–713.
35. Sukhodolsky DG, et al. J Am Acad 55. McMahon WM, et al. Adv Neurol 76. O’Roak B, et al. Mol Psychiatry
Child Adolesc Psychiatry 2003; 1992;58:159–165. 2010;15(5):447–450.
42(1):98–105. 56. Pakstis AJ, et al. Am J Hum Genet 77. Zuchner S, et al. Mol Psychiatry
36. Comings DE, et al. J Am Acad 1991;48(2):281–294. 2006;11(10):887–889.
Child Psychiatry 1984; 57. Pauls DL, et al. J Am Acad Child
23(2):138–146. 78. Aruga J, et al. Gene 2003;
Adolesc Psychiatry 1990; 315:87–94.
37. Pauls DL, et al. J Am Acad Child 29(2):195–203.
Adolesc Psychiatry 1993; 79. Aruga J, et al. Mol Cell Neurosci
58. Verkerk AJ, et al. Mol Psychiatry
32(5):1044–1050. 2006;11(10):954–964.
2003;24(1):117–129.
38. Stewart SE, et al. J Am Acad Child 80. Keen-Kim D, et al. Hum Mol
59. Tourette Syndrome Association
Adolesc Psychiatry 2006; Genet 2006;15(22):3324–3328.
International Consortium
45(11):1354–1362. for Genetics. Am J Hum 81. Deng H, et al. Acta Neurol Scand
39. Grados MA, et al. Biol Psychiatry Genet 1999;65(5): 2006;114(6):400–402.
2008;64(3):219–225. 1428–1436. 82. Chou IC, et al. Pediatr Neurol
40. Alsobrook JP, II, et al. Am 60. Tourette Syndrome Association 2007;37(6):404–406.
J Psychiatry 2002;159(2):291–296. International Consortium for 83. Scharf JM, et al. Neurology
41. Mathews CA, et al. Biol Psychiatry Genetics. Am J Hum Genet 2008;70(16 Pt 2):1495–1496.
2007;61(3):292–300. 2007;80(2):265–272.
84. Orth M, et al. Mov Disord 2007;
42. Robertson MM, et al. Psychiatr 61. Paschou P, et al. Am J Hum Genet 22(14):2090–2096.
Genet 2007;17(3):143–152. 2004;75(4):545–560.
85. Fabbrini G, et al. Mov Disord
43. Cavanna AE, et al. J Neurol 62. Zhang H, et al. Am J Hum Genet 2007;22(15):2229–2234.
Neurosurg Psychiatry 2011;82 2002;70(4):896–904.
86. Cruz C, et al. Neurosci Lett
(12):1320–1323. 63. Simonic I, et al. Am J Hum Genet 1997;231(1):1–4.
44. Baron-Cohen S, et al. Psychol Med 1998;63(3):839–846.
87. Diaz-Anzaldua A, et al.
1999;29(5):1151–1159. 64. Simonic I, et al. Am J Med Genet Mol Psychiatry 2004;9(3):
45. Kadesjo B, et al. J Am Acad Child 2001;105(2):163–167. 272–277.
Adolesc Psychiatry 2000;39 65. Merette C, et al. Am J Hum Genet 88. Grice DE, et al. Am J Hum Genet
(5):548–555. 2000;67(4):1008–1013. 1996;59(3):644–652.
46. Fernandez TV, et al. Biol 66. Motlagh MG et al. Mov Disord 89. Tarnok Z, et al. Am J Med
Psychiatry 2012;71(5):392–402. 2008;23(14):2079–83. Genet B Neuropsychiatr
47. Ercan-Sencicek AG, et al. N Engl 67. Abelson JF, et al. Science 2005; Genet 2007;144B(7):
J Med 2010;362(20):1901–1908. 310(5746):317–320. 900–905.

345
Chapter 28: Genetics of Tourette syndrome

90. Yoon DY, et al. Am J Med Genet 107. Dewing P, et al. Curr Biol 2006;16 125. Stillman A, et al. J Comp Neurol
B Neuropsychiatr Genet (4):415–420. 2009;513(1):21–37.
2007;144B(5):605–610. 126. Sebat J, et al. Science
108. Lin H, et al. J Child Psychol
91. Cavallini MC, et al. Psychiatry Res Psychiatry 2007;48(2):157–166. 2004;305:525–528.
2000;97(2–3):93–100.
109. Kessler AR. J Child Neurol 127. Eichler EE. Nature Genetics
92. Chou IC, et al. Psychiatr Genet 2002;17(10):738–744. 2006;38:9–11.
2007;17 (6):359.
110. Scahill L, et al. Percept Mot Skills 128. Stefansson H, et al. Nature
93. Xu C, et al. Am J Med Genet 2001;92(2):419–432. 2008;455:232–236.
B Neuropsychiatr Genet
111. Kushner HI. A Cursing Brain? The 129. Walsh T, et al. Science 2008;320
2003;119B(1):54–59.
Histories of Tourette Syndrome. (5875):539–543.
94. Brett PM, et al. Am J Psychiatry Cambridge, MA: Harvard
1995;152(3):437–440. 130. Wilson G, et al. Hum
University Press; 1999. Mol Genet 2006;15:
95. Mössner R, et al. Mol Psychiatry 112. Hoekstra PJ, et al. Cell Mol Life Sci 743–749.
2007;12(7):617–619. 2004;61(7–8):886–898. 131. Xu B, et al. Nat Genet
96. Niesler B, et al. Psychiatr Genet 113. Maia AS, et al. Rev Hosp Clin 2008;40:880–885.
2005;15(4):303–304. Fac Med Sao Paulo 1999;54(6): 132. Mulle JG, et al. Am J Hum Genet
97. Comings DE, et al. Neuroreport 213–221. 2008;87:229–236.
1999;10(7):1589–1592.
114. Riedel M, et al. Lancet 1998;351 133. Consortium IS. Nature
98. Kindler J, et al. Psychiatry Res (9100):418–419. 2008;455:237–241.
2008;15:235–9.
115. Muller N, et al. Psychiatry Res 134. McCarthy S, et al. Nat Genet
99. Laurin N, et al. Am J Med Genet 2004;129(2):119–125. 2009;41:1223–1227.
B Neuropsychiatr Genet
116. Snider LA, et al. Mol Psychiatry 135. Marshall C, et al. Am J Hum Genet
2009;150B(1):95–103.
2004;9(10):900–907. 2008;82:477–488.
100. Miranda DM, et al. Am J Med
Genet B Neuropsychiatr Genet 117. Liao IH, et al. Pharmacogenomics 136. Sebat J, et al. Science 2007;316
2008;147B(1):68–72. 2010;11(12):1733–41. (5823):445–449.
101. Burd L, et al. J Perinat Med 118. Lit L, et al. Am J Med Genet 137. Szatmari P, et al. Nat Genet
1999;27(4):295–302. B Neuropsychiatr Genet 2007;39:319–328.
2007;144B(7):958–963. 138. Glessner J, et al. Nature
102. Whitaker AH, et al. Arch Gen
Psychiatry 1997;54(9):847–856. 119. Tang Y, et al. Arch Neurol 2005;62 2009;459:569–573.
(2):210–215. 139. Sanders SJ, et al. Neuron
103. Mathews CA, et al. Am
J Psychiatry 2006; 120. Tian Y, et al. Brain Res 2011;70:863–885.
163(6):1066–1073. 2011;1381:228–236. 140. Kaffman A, et al. J Child
104. Leckman JF, et al. J Am Acad 121. Tian Y, et al. Am J Med Genet Psychol Psychiatry 2007;48
Child Adolesc Psychiatry 1990;29 B Neuropsychiatr Genet (3–4):224–244.
(2):220–226. 2011;156B(1):72–78. 141. Chappell P, et al. Biol Psychiatry
105. Peterson BS, et al. J Clin 122. Johnson MB, et al. Neuron 1994;36(1):35–43.
Psychopharmacol 1998; 2009;62(4):494–509. 142. Corbett BA, et al.
18(4):324–331. 123. Makeyev EV, et al. Science Psychoneuroendocrinology
106. Alexander GM, et al. Dev 2008;319(5871):1789–1790. 2008;33(6):810–820.
Psychopathol 2004; 124. Hamann M, et al. Brain Res 143. Chappell P, et al. Biol Psychiatry
16(2):407–420. 2007;1150:190–199. 1996;39(9):776–783.

346
Endophenotypes in psychiatric genetics
Chapter

29 Andrew C. Chen, Madhavi Rangaswamy, and Bernice Porjesz

It has been more than a half century since Nobel and nongenetic contributions to a continuous pheno-
laureates Drs. Watson, Crick, and Wilkin discovered type on the basis of observations of seed colors in
the structure of DNA. Since that time, rapid advance- crosses of oats and wheat [1]. As we now know, a
ment of molecular biological technology has facili- genotype is the composition of DNA sequences; in
tated progress in the field of genetics, with the contrast, a phenotype represents observable charac-
complete sequencing of the human genome by the teristics of an organism, and is the joint product of
Human Genome Project accomplished in 2003. It is both genotypic and environmental influences. In gen-
now possible, not only to identify single gene Men- etic diseases that are transmitted through the classic
delian disorders, but also to understand “complex” or Mendelian model, genotypes are usually indicative
disorders that are the result of multiple small gene of phenotypes in terms of the presence or absence of
effects that predispose individuals to these disorders, the disease. However, this certain correlation between
as they interact with environmental factors. As we try genotypes and phenotypes does not exist in complex
to envision the future direction of human genetics genetic diseases.
research, particularly the genetics of complex genetic Recently, there has been a surge of interest in the
diseases such as the majority of psychiatric diseases, it use of endophenotypes in psychiatric research,
would be helpful to briefly review some important although the concept was first introduced to the field
milestones for the development of genetics in terms of psychiatry by Gottesman and Shields more than
of phenotypes before the era of molecular biology. three decades ago [2]. This has been driven by con-
A few years after the term “genetics” was first cerns about the limited success and relatively poor
introduced by William Bateson in 1902, the Danish reproducibility of the current psychiatric genetic
botanist Wilhelm Johanssen attempted to provide the research approach, and the fact that current psychi-
clarifying distinction we now take for granted atric diagnostic systems, DSM or ICD, are primarily
between the concepts of “genotype” and “phenotype”. based on phenomenology and lacking of justification
He also introduced the word “gene” in 1909, a half by etiology. Psychiatry’s classification systems
century before the discovery of the structure of the describe heterogeneous disorders [3]. The brain is
genetic molecule, DNA. Through his research on self- the most complex of all organs. It is subject to com-
fertilized lines of beans, Johanssen found that the plex interactions not only among genes, proteins,
phenotype is often an imperfect indicator of the geno- cells, and circuits of cells, but also across individuals
type: the same genotype may give rise to a wide range and their changing experiences [4]. As such, the
of phenotypes. Similarly, the same phenotype may phenotypic output from the brain, i.e. behavior, is
have arisen from different genotypes. This important not simply a sum of all its parts. Therefore, it would
observation was contradictory to the genetic theory at be a more reasonable approach to apply more opti-
that time, which endorsed the idea that there was a mally reduced measures of neuropsychiatric function-
one-to-one correlation between genotype and its ing than a behavioral complex, the diagnosis, in
“consequential” phenotype. About the same time, studies of the biological and genetic components of
Herman Nilsson-Ehle, a Swedish expert in plant psychiatric disorders. It has been suggested that
breeding and genetics, provided evidence for genetic ideally, molecular genetic studies should not be

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

347
Chapter 29: Endophenotypes in psychiatric genetics

performed on psychiatric diagnoses alone, which to apply endophenotypes to study the functional con-
reflect distal and variable effects of genes, but on sequences of risk alleles.
quantitative neurobiological measures or markers
that reflect more proximal effects of genes involved
in the genetic predisposition for developing psychi-
Characteristics useful for the
atric disorders [5]. identification of endophenotypes
in psychiatric genetics
The endophenotype concept There have been a number of attempts to devise
criteria to define the optimal characteristics of an
An endophenotype is typically an unobserved pheno-
endophenotype [6, 8–10]. As the diagnostic categor-
type, such as metabolite level, cell or organ activity,
ies of psychiatric disorders in use today were initially
biosignal, or other “biomarker”, that is thought to
formulated in the late nineteenth and early twentieth
contribute to the etiology of a visible phenotype or
centuries by a small number of psychiatrists
disease susceptibility. It is associated with a disease
who relied on perceived similarities in behavioral
but can be measured independently of disease status.
syndromes and clinical outcomes, we should be
Gottesman and Gould have defined endophenotype
aware that such categories reflected only observable
as “measurable components unseen by the unaided
behaviors rather than dysfunctions in distinct ana-
eye along the pathway between disease and proximal
tomical substrates or patho-physiological processes.
genotype” [6]. In their writings summarizing genetic
Any given disorders with singular labels, such as
theories in schizophrenia 30 years ago, Gottesman
schizophrenia or attention-deficit hyperactivity dis-
and Shields described “endophenotypes” as internal
order (ADHD), are probably better thought of as a
phenotypes that lie on the pathway between genes and
heterogeneous group of dysfunctions whose final
disease, and can be observed only by a “biochemical
pathways eventually lead to similarity in symptoms,
test or microscopic examination” [2]. These quantita-
course of illness, and other clinical features. For
tive biological markers serve as covariates that correl-
endophenotypes to be useful in psychiatric genetic
ate with the main trait of interest (diagnosis) and
analyses, they should have several properties [6, 11].
serve to better define that trait or its underlying
We summarize the six properties below. The first
genetic mechanism. The term “endophenotypes” was
two are most necessary for any endophenotype to
adapted from a 1966 paper explaining concepts in
be useful in medical genetics. The last four, although
evolution and insect biology by John and Lewis [7];
not strictly necessary, are properties that can aid
they wrote that the geographical distribution of grass-
in the successful genetic dissection of complex
hoppers was a function of some feature not apparent
behaviors.
in their “exophenotypes” that are obvious and exter-
nal. In contrast, these features were “microscopic and
internal”, hence the term, “endophenotype”. Thus, the The endophenotype is heritable
term endophenotype appears to fit the needs of psy- Complex traits or phenotypes are the outcome of an
chiatric genetics, and this concept bridges the gap interaction between genetic and environmental
between the gene and the complexity of psychiatric factors. One of the most important criteria defining
disease processes. Typically, endophenotypes are an endophenotype is that it be a heritable trait. For
quantitative traits that can be measured in a general endophenotypes to aid in the genetic dissection of
population. That is, the traits are present in individ- complex traits, there must be some evidence of herit-
uals who have manifested the symptoms of the disease ability of the phenotype. It is not very useful to link a
as well as in those who do not have the disease, trait to genes if genetic factors do not contribute to
including those at risk prior to the onset of the the individual differences in the trait, i.e. when it is
disease. largely influenced by environmental factors. How-
Generally speaking, the utility of the endopheno- ever, it should be kept in mind that using endophe-
types is relevant for two types of studies. First, as notypes of relatively low heritability may still be
originally proposed [2, 6], they are used to aid in the important, particularly when there is a potential for
discovery of novel genes that are associated with the gene  environment interaction. As environmental
disorders. Second, an increasingly popular method is effects may be crucial in altering the effects of even

348
Chapter 29: Endophenotypes in psychiatric genetics

highly heritable disorders, these endophenotypes can Within families, endophenotype


also aid in the development of pharmacological or
behavioral treatments and interventions. and illness co-segregate
Affected members of families manifest the endophe-
The endophenotype is associated notypic marker significantly more than unaffected
members of the families.
with the disorder
There should be reasonable evidence that the endo-
phenotype is associated with the pathophysiology of The endophenotype found in affected family
the illness. Ideally, it should be a part of the causal members is found in nonaffected family
pathway from gene to disorder rather than effects
(sequelae) of disorders or their treatment. This (to
members at a higher rate, or at a higher level,
be associated with the causes of the disorders) is than in the general population
important when the endophenotype is being used to This will allow small to moderate effect sizes between
highlight the risk genes, and relatively less important biological relatives of affected patients and commu-
when the endophenotype is applied to study the func- nity controls to be observed more easily [13].
tional consequences of known risk alleles.
Application of endophenotypes
The endophenotype is present in both
in psychiatric genetics
affected and nonaffected individuals, and it A fundamental assumption for the application of
varies continuously in the general population endophenotype in psychiatric genetic research is that
Although the diagnosis of mental disorders is dichot- the number of genes required to produce variations in
omous, i.e. whether a certain disorder is present or these traits is fewer than those involved in producing
not, it is unlikely that common mental disorders vary a psychiatric diagnostic entity, which may reflect a
in a discrete manner between individuals. Statistical more distal end-point of the combined contributions
analyses of continuous traits that are artificially cat- from multiple genetic and environmental factors, as
egorized into discrete groups have substantially less well as the interactions among them. Quantitative
power than analyses conducted on the original con- neurobiological measures of risk as endophenotypes
tinuous scale. Therefore, it would be more powerful are closer to gene action involved in the predispos-
to analyze endophenotypes that can be measured on ition for the disorder. Therefore, in searching for
continuous scales. This also highlights another genetic variations that are involved in complex
important advantage of analyzing endophenotypes (non-Mendelian) psychiatric diseases, endopheno-
that vary continuously in the nonaffected population types provide a more powerful strategy than the
because it can simplify the sampling process and dichotomous diagnosis of the disease [6, 14, 15]. In
reduce ascertainment biases in population-based psychiatry, a number of attempts have been made to
studies [12]. develop and determine the feasibility of candidate
endophenotypes. Although few candidates have
The endophenotype is primarily met all the criteria listed earlier, some linkage and
association studies using the endophenotype strategy
state-independent have had moderate success.
As mentioned, to be useful for genetic analysis, the This chapter is organized into two main sections.
endophenotype should be present in both affected In the first part of this chapter, we summarize repre-
and unaffected individuals. To aid in the discovery sentative published studies that have successfully used
of novel genes that are associated with the disorders, various endophenotypes to find candidate genes
the endophenotype should also manifest in an indi- involved in several psychiatric disorders, including
vidual at a similar quantitative level regardless of schizophrenia, mood disorders, Alzheimer’s disease
whether or not the illness is active. It should (AD), ADHD, and suicidal behavior. The second half
demonstrate adequate test–retest stability and of this chapter focuses on alcoholism, and more fully
reliability. illustrates the successful use of the endophenotype

349
Chapter 29: Endophenotypes in psychiatric genetics

strategy in the Collaborative Study on the Genetics of promoter region of the CHRNA7 gene have been
Alcoholism (COGA) project, where we have used found to be associated with schizophrenia and/or
brain oscillations as endophenotypes in the identifi- P50 suppression abnormalities [27, 28].
cation and understanding of genes involved in alco-
holism and related disinhibitory disorders. Eye-tracking dysfunction as endophenotypes
Sensory motor gating as endophenotypes for schizophrenia and bipolar disorder
Eye movements are generally of two forms: saccadic
for schizophrenia (brief and extremely rapid movements) and smooth
Deficits in sensory motor gating are consistent neuro- (“smooth pursuit” eye movements occur only when
psychological findings in individuals suffering from the subject is following an object moving at a constant
schizophrenia [16]. The hypothesized association has velocity, such as a pendulum or bright dot on a
face validity primarily on the basis of patients’ reports computer monitor). Initiation and maintenance of
that they may have difficulty filtering information smooth pursuit eye movements (SPEMs) involve inte-
from multiple sources that occur in everyday life gration of functions of the prefrontal cortex frontal
[16–18]. Neurophysiological tests, including assess- eye fields, visual and vestibular circuitry, thalamus,
ments of P50 suppression and pre-pulse inhibition and cerebellum, as well as the muscles and neural
of the startle response, have been developed to discern circuitry directly responsible for eye movement [29].
efficiencies in these capabilities. Studies have found that patients with schizophrenia
In tests of pre-pulse inhibition, startling sensory and their unaffected relatives have increased rate of
stimuli (e.g. loud noise, bright light) are used to elicit deficiencies in SPEMs [30, 31]. Similar results have
an unconditional reflexive startle response in individ- also been reported in bipolar disorder [31, 32]. Stud-
uals. If a weaker pre-stimulus is provided before the ies have reported evidence for linkage of SPEM
startling stimulus, the subsequent startle response is phenotypes to chromosome 6p23–21 [33, 34]; the
generally diminished. A relatively reproducible find- region harbors two genes associated with risk of
ing is that this diminution of the second response is schizophrenia, ATXN1 (SCA1) and NOTCH4 [32].
attenuated in patients with schizophrenia [16, 17]. Of
note, abnormal pre-pulse inhibition is not specific to Neurocognitive endophenotypes
schizophrenia; studies have identified this abnormal-
ity in obsessive–compulsive disorder [19] and Hun- for schizophrenia and ADHD
tington’s disease [20]. Family and twin studies have suggested moderate
The P50 suppression test uses two auditory stim- heritability of working memory deficits in schizo-
uli presented at 500 ms intervals. Event-related poten- phrenia [35–37]. Recently, researchers such as the
tials (ERPs) for both stimuli are measured by Consortium on the Genetics of Schizophrenia
electroencephalogram (EEG). In normal individuals, (COGS), a seven-site collaboration that examines the
the ERP to the second stimulus is of lower amplitude genetic architecture of quantitative endophenotypes
than the first. However, suppression of P50 amplitude in families with schizophrenia, have attempted to
is compromised in patients with schizophrenia [16, select neurocognitive tasks as endophenotypic meas-
17, 21, 22] as well as in their unaffected first-degree ures in genetic studies [38, 39]. The COGS neurocog-
relatives [22–24]. Studies on the heritability of P50 nitive assessment includes measures of attention,
suppression have strongly suggested that the variation verbal memory, working memory, and a computer-
in P50 is under the influence of genetic factors and ized neurocognitive battery that also includes facial
that P50 suppression is a good candidate endopheno- processing tasks [40].
type [25]. Freedman and colleagues [26] used P50 A study of Finnish twins [41] using the sum of
suppression to identify a potential susceptibility locus performance scores on four neuropsychological tests
for schizophrenia on chromosome 15, a chromo- as an endophenotype suggested linkage and associ-
somal region where the CHRNA7 gene encoding the ation to a region of chromosome 1q, a region previ-
alpha-7 nicotinic acetylcholine receptor resides. Link- ously suggested in traditional linkage studies of
age disequilibrium in this region [27] has subse- schizophrenia [42]. Studies [43] have also shown an
quently been reported, and variations in the association of poorer performance on a working

350
Chapter 29: Endophenotypes in psychiatric genetics

memory task with the gene encoding the enzyme (PFC) was needed to accomplish the same task in
catechol-O-methyltransferase (COMT) located at the fMRI assessment, suggesting that activation of
chromosome 22q and increased risk for developing the dorsolateral prefrontal cortex is less efficient in
schizophrenia. This chromosomal region has been those subjects [50]. This inefficiency was observed in
reported to be linked to both schizophrenia and bipo- schizophrenic subjects as well as unaffected siblings of
lar disorder and overlaps with a deletion that has been schizophrenic patients [51].
associated with velo-cardio-facial syndrome In agreement with this discovery, variation in
(DiGeorge syndrome) and schizophrenia [44]. COMT also modulates other PFC-dependent neuro-
A functional polymorphism, Val158Met, which psychological performance [45] and the cortical
causes a valine to methionine mutation at peptide response to amphetamine, which increases synaptic
158 in the gene encoding COMT, results in four-fold dopamine [52]. The latter finding suggests that the
decreased activity in the rarer Met allele, hence pref- COMT genotype links prefrontal dopamine stimula-
erentially increasing prefrontal extra-synaptic dopa- tion and neuronal activities, in which homozygotes
mine, because COMT provides the major clearing for the Val-encoding allele presumably possesses less
step for dopamine released from the synapse in this synaptic dopamine due to maximal COMT activity
region [45]. The more common Val allele is associ- than the Met allele carriers. Additional evidence for
ated with less efficient cognitive processing by the this comes from a positron emission tomography
brain, and is transmitted at a higher rate than the (PET) study [53] showing that the COMT genotype
Met allele to patients with schizophrenia than to their has an impact on the prefrontal regulation of mid-
nonaffected siblings [46]. brain dopamine synthesis.
Similarly, many neurocognitive measures of In addition to fMRI, phenotypes measured by
executive functions, for example, Attention Network anatomical imaging such as structural MRI have also
Task, Continuous Performance Test, Matching been proposed as endophenotypes for some psychi-
Familiar Figures Test, Span of Apprehension Test, atric disorders. For example, reduction in orbitofron-
Spontaneous Selective Attention Task, and Wisconsin tal volume in ADHD [54], decreased gray matter in
Card Sorting Test, have been proposed as endophe- insular cortex, and temporal lobe gray matter abnor-
notypes for ADHD [47, 48]. Studies have shown that malities in schizophrenia [55]. Recent studies using
deficits in neurocognition are a correlate of ADHD diffusion tensor imaging (DTI) and magnetization
and show preliminary evidence of heritability and transfer ratio (MTR) protocols have begun to shed
association with relevant candidate genes. Nonethe- more light on the white matter deficits among schizo-
less, studies that have assessed the familial and genetic phrenia patients, and studies on their heritability
overlap of neurocognitive impairments with ADHD (including their potential as endophenotypes) are
have yielded inconsistent results. In order for execu- under way [55, 56].
tive function deficits to be used as an endophenotype
for ADHD, more studies with greater attention to the Neurochemical metabolites as
neurocognitive heterogeneity of this disorder and to
the precision of measurement of the neuropsycho- endophenotypes
logical tests are required [49]. In addition to soft neurological signs and neurocog-
nitive performance, many studies have focused on
neurochemicals, neurohormonal substances, and
Neuroimaging endophenotypes their metabolites in search of candidate endopheno-
Another powerful and rapidly evolving technique types for psychiatric disorders. Here we use a few
used in psychiatric genetic studies is functional mag- recent studies in AD and suicide as examples.
netic resonance imaging (fMRI), which combines the Aggregation and deposition of amyloid beta (Ab)
neurocognitive tasks with the anatomical location of in the brain is thought to be central to the pathogen-
brain activities. Studies have demonstrated that the esis of Alzheimer’s disease (AD). Recent studies sug-
more common Val allele of the functional poly- gest that cerebrospinal fluid (CSF) Ab levels are
morphism Val158Met in the COMT gene is associated strongly correlated with AD status and progression.
with less efficient cognitive processing by the brain, Using CSF Ab levels as a phenotype for AD, a recent
and that more blood flow in the prefrontal cortex study has identified a DNA sequence variation in the

351
Chapter 29: Endophenotypes in psychiatric genetics

(a) 12

Frequency (Hz)
Fz

Cz
6

Pz
3

Theta
–216 –31 153 338 522 706 891 1075 1259
(b) Time (ms)

Chromosome 7
3.5 θ
Fz, Max LOD = 3 .16 at 161 cM
3
Cz, Max LOD = 3 .6 at 164 cM

Pz, Max LOD = 2 .29 at 162 cM


2.5

2
LOD

1.5

0.5 CHRM2
GRM8

0
D7S1790

D7S1802

D7S1838

D7S2846

D7S1830

D7S3046
D7S1870

D7S1797

D7S1796

D7S1799

D7S1817

D7S2847
D7S1809

D7S1804

D7S1824

D7S1805
D7S513

D7S629
D7S673

D7S817

D7S521
D7S691
D7S478
D7S679
D7S665

D7S820

D7S821

D7S490

D7S509

D7S794
NPY2

0 20 40 60 80 100 120 140 160 180


Chromosome position (cM)
CHRM2
(c)
exon

exon

exon

exon

exon

1 2 3 4 5 6
Coding
5’-UTR 3’-UTR
Sequence

(d)

SNPs rs6948054 rs1424548


rs1378646 rs1378650
rs1455858
rs7782965
rs978437

Figure 29.1 Illustration of endophenotype strategy: from neurocognitive endophenotypes to genes. (a) Endophenotype (Theta oscillation),
Theta (y, 3–7 Hz) event-related oscillations (EROs) underlying the processing of target stimuli during P3 production (300–700 ms) in a visual
oddball task as seen in this time-frequency representation (right panel). Head plot (left) displays scalp topography of theta ERO power.

352
Chapter 29: Endophenotypes in psychiatric genetics

presenilin 1 (PSEN1) gene that is a novel disease- spectrum of disinhibitory disorders, which include
causing mutation in clinically characterized research externalizing and other substance use disorders.
subjects with extreme values of CSF [57]. Many of the same genetic risk factors are postulated
Another example is cortisol response to psycho- to underlie these disinhibitory co-occurring dis-
social stress with regard to suicidal behavior. Disturb- orders which are explained by a common under-
ances in hypothalamic-pituitary-adrenal (HPA) axis lying genetic liability involving impulse control
function have been observed in suicide attempters [65]. These findings are in keeping with electro-
using various indices, including CSF corticotrophin physiological findings reflecting a similar electro-
releasing hormone (CRH), cortisol levels following physiological profile in these related disinhibitory
dexamethasone challenge, and urinary cortisol [58, disorders [15]. Brain dysfunction is likely to be
59]. Twin studies of cortisol levels in blood, urine, involved in a genetic predisposition to develop alco-
and saliva estimated heritability to be approximately holism and other psychiatric disorders, and neuro-
60% [60]. A twin study of cortisol response to psy- electric events may serve as excellent biological
chosocial stress (Trier Social Stress Test [TSST]) markers or endophenotypes.
estimated heritability of cortisol response with repeti-
tion of the stressor to be between 56 and 97% [61].
Several polymorphisms that have been shown to be Brain oscillations as endophenotypes
associated with cortisol response to TSST include the Brain oscillations provide a rich source of potentially
mineralocorticoid and glucocorticoid receptor genes useful endophenotypes for psychiatric genetics, as
[62], the 5HTTLPR [63], and the g-aminobutyric they represent important neural correlates of infor-
acid (GABA)-A a-6 receptor gene [64]. Studies are mation processing and cognition in humans. These
necessary to establish if the same deficits in cortisol oscillations are dynamic indices of the millisecond
response are more frequent in unaffected relatives by millisecond balance between excitation and inhib-
of suicidal individuals compared with the general ition in brain neural networks and have exquisite
population. temporal resolution. They reflect ensembles of
neurons firing in synchrony and represent the basic
mechanism of neural communication. High frequency
Neurocognitive phenotypes in alcoholism (beta [12–28 Hz], gamma [28þ Hz]) synchronizations
are involved in short-range communication, while low
and related disorders: brain oscillations as frequencies (delta [1–3 Hz], theta [4–7 Hz], alpha [8–
endophenotypes 12 Hz]) are involved in longer range communication
In this section, we more fully illustrate the endo- between brain areas [66]. Brain oscillations represent
phenotype approach – from endophenotype to traits less complex and more proximal to gene function
genes, using brain oscillations as endophenotypes than either diagnostic labels or traditional cognitive
in the search for genes predisposing to alcoholism measures. Therefore, these oscillations may be utilized
and related disinhibitory disorders in the COGA as phenotypes of cognition, and as valuable endophe-
project (Figure 29.1). Alcoholism is a common notypes for the understanding of some complex gen-
complex (non-Mendelian) disorder with contribu- etic brain disorders. Several brain oscillations meet
tions from both genetic and environmental influ- criteria for endophenotypes (see above); they are highly
ences and their interactions. Recent evidence heritable: delta 76%, theta 89%, alpha 89%, and beta
suggests that alcohol dependence is part of a 86% [67].

Figure 29.1 (cont.) (Theta EROs are correlates of impaired cognitive brain processes in alcoholism and risk.). (b) Genetic linkage analysis scans
the entire genome to assess chromosomal regions that contain polymorphic genetic markers that are linked to a quantitative trait (theta
ERO power) within families. This is a logarithm of odds (LOD) score plot of linkage for theta ERO power at frontal (Fz), central (Cz), and parietal
(Pz) electrodes on chromosome 7 with a significant linkage peak that contains 2 candidate genes under the linkage peak – CHRM2 and GRM8.
(c) Candidate gene studies focus on genes underlying significant linkage peaks and/or genes with relevant biological significance. This is a
schematic diagram of the CHRM2 gene indicating its coding region (gray) and exons. (d) Genetic association tests are performed for each
single nucleotide polymorphism (SNP) in the candidate gene and the trait variable. SNPs are identified within and flanking the candidate
CHRM2 gene (exons and introns). SNPs with significant associations (gray boxes) with the endophenotype indicate loci of interest. (The right-
most box with two SNPs lies in the region beyond the 3’ UTR of the gene.) See plate section for color version.

353
Chapter 29: Endophenotypes in psychiatric genetics

In our strategy for finding susceptibility genes for As increased resting beta power is already observed
alcohol dependence and related disorders, we select in offspring before the onset of alcohol dependence,
brain oscillations that not only differentiate between it is considered to be a trait rather than a state measure.
alcoholics and controls, but also between high risk Significant linkage and linkage disequilibrium between
offspring of alcoholics and controls, to be sure that we the EEG beta frequency and a g-aminobutyric acid
are selecting “trait” rather than “state” measures that (GABAA) receptor gene on chromosome 4 has been
meet criteria for good endophenotypes. COGA is a reported by COGA (Figure 29.2) [79]. With the use
genetic study of densely affected families ascertained of multiple single nucleotide polymorphisms (SNPs)
through affected probands in treatment, where there across this cluster of GABAA receptor genes on
were at least three affected (DSM III-R and Feighner chromosome 4, that includes GABRA2, GABRA4,
definite for alcohol dependence) first-degree relatives. GABRB1, and GABRG1, we were able to specifically
We focused on neuroelectric endophenotypes (such identify that it was variations only in the GABRA2
as resting eyes-closed EEG [e.g. power and coher- receptor gene that accounts for the linkage /linkage
ence], ERPs [e.g. the P3 component], and event- disequilibrium findings with the beta frequency. Thus,
related oscillations [EROs] [e.g. theta and delta variations in GABRA2 (the gene encoding the a-2
event-related oscillations]) recorded during sensory subunit of the GABAA receptor) affect brain oscilla-
and cognitive tasks. As will be shown in this chapter, tions and are directly involved in the level of neural
we have successfully used this approach with brain excitability (balance between excitation and inhib-
oscillations as endophenotypes first to target chromo- ition). There is a strong relationship between the most
somal regions using linkage analysis, and secondly significant SNP (rs279836) in the GABRA2 receptor
identifying potential candidate genes that may be gene and beta EEG power. Of note individuals who
involved in both brain function and diagnosis of are homozygous for the rarer genotype (15%) of this
alcohol dependence and related disorders contained SNP have significantly increased EEG beta compared
in the chromosomal regions under that linkage peak. to individuals with all other genotypes, indicating
These genes are then more closely investigated using underlying central nervous system (CNS) disinhibi-
association analyses (Figure 29.1). tion. The same GABRA2 gene associated with the
EEG beta endophenotype was subsequently found
to be associated with DSM-IV diagnosis of alcohol
Resting EEG beta power: GABRA2 gene dependence [80], substance dependence [81], ASP
EEG signatures of the resting state of the brain have [82], and childhood conduct disorder [83]. The associ-
revealed characteristic patterns in individuals with ation between GABRA2 and alcohol dependence has
alcoholism and also those with high risk for develop- been replicated by a number of groups throughout the
ing alcoholism and related spectrum of externalizing world [84–89].
conditions [15]. Across most studies published in this Fast synaptic inhibition in the mammalian CNS is
field, increased beta power in the EEG has emerged as mediated largely by activation of GABAA receptors
one such important feature; it is noted in affected [90]. GABAA actions are a fundamental requirement
[68–73] and high risk offspring of alcoholics [74, for both gamma and beta oscillations to occur, and
75], including a large sample of offspring of alcoholics blockade of these receptors results in loss of syn-
at high risk from the COGA project [76]. Increased chronization [91]. Although the recording of elec-
resting beta power has been reported at frontal leads trical oscillations from a neural population reflects
in those who also have a diagnosis of antisocial per- the firing of multiple excitatory pyramidal cells, the
sonality disorder (ASP) [77]. Also, increased beta mechanism underlying beta and gamma oscillations
power may be associated with increased vulnerability, depends on the firing patterns of a network of inhibi-
as female high risk subjects with a larger number of tory interneurons [92, 93], that are gated by their
affected first-degree relatives displayed significantly mutually induced GABAA action [94].
elevated beta power compared to those with just one These genetic findings relating EEG beta and the
affected parent [72]. Thus evidence of elevated beta GABRA2 gene suggests that variations in the GABRA2
power provides strong support for the excitation– gene affect the inhibitory tone and thus affect the level
inhibition imbalance model proposed to underlie the of neural excitability, which may underlie the predis-
predisposition to alcohol dependence [78]. position to develop alcohol dependence and related

354
(a) Chromosome 4 (b) Chromosome 7

3.5

2.5

LOD
1.5

CHRM2
GRM8
0.5
GABA-A
0
D4S2639
D4S2382
D4S1627
D4S3025
D4S1536

D4S1630

D4S1645
D4S2432
D4S3253
D4S2367
D4S1558
D4S2393
D4S2630
D4S3088
D4S2361
D4S1544
D4S2404
D4S1559

D4S1570
D4S1651

D4S1625

D4S1629
D4S1626

D4S2374

D4S1652
GABRB1

D7S1790

D7S1802

D7S1838

D7S2846

D7S1830
D7S3046
D7S1870

D7S1797

D7S1796

D7S1799

D7S1817

D7S2847
D7S1809

D7S1804

D7S1824

D7S1805
D4S400

D4S171

D7S513

D7S629
D7S673

D7S817

D7S521
D7S691
D7S478
D7S679
D7S665

D7S820

D7S821

D7S490

D7S509

D7S794
FABP2
ADH3

NPY2
50 100 150 200 0 20 40 60 80 100 120 140 160 180
Chromosome position (cM) Chromosome position (cM)

Figure 29.2 (a) Linkage plots showing maximum logarithm of odds (LOD) scores with significant linkage peaks over the GABAA receptor gene cluster on chromosome 4 with two resting
EEG phenotypes. Red trace: Resting EEG beta band power – beta 2 (16.5–20.0 Hz). The dataset consists of 1553 individuals from 250 families. Green trace: Resting EEG high theta band (6–7
Hz) interhemispheric coherence at parieto-occipital bipolar pairs of electrodes (P4_O2-P3_O1). The dataset consists of 1312 individuals from 251 families.
(b) Linkage plot showing maximum LOD scores with significant linkage peaks on chromosome 7, over the region harboring two candidate genes: a cholinergic muscarinic receptor gene
(CHRM2) and a glutamate receptor gene (GRM8). Blue trace: the central midline theta (4–5 Hz) ERO band power (between 300–700 ms, P3 latency window for visual target case during visual
oddball task) on chromosome 7. The dataset consists of 1337 individuals from 253 families; Green trace: Resting EEG theta band (6–7 Hz) interhemispheric coherence at centro-parietal
bipolar pairs of electrodes (C4_P4-C3_P3) on chromosome 7. The dataset consists of 1312 individuals from 251 families. See plate section for color version.
Chapter 29: Endophenotypes in psychiatric genetics

disorders. Alcoholics and offspring at high risk mani- indicative of altered functional thalamocortical and
fest increased power in EEG beta oscillations, suggest- cortico-cortical connectivity.
ing an imbalance between excitation–inhibition (CNS We conducted a whole genome linkage analysis in
disinhibition). This provides a biological hypothesis COGA using the high theta coherence at parietal-
relating the underlying CNS disinhibition to the gen- occipital leads as the phenotype. Highly significant
etic risk for alcohol dependence and related disorders linkage was found on chromosome 4 in the same region
[78]. The involvement of the GABAergic system in spotlighted by the EEG beta power phenotype [104;
alcoholism is supported by neuroimaging studies, Rangaswamy et al., unpublished]. Family-based associ-
which report specific deficits in GABA benzodia- ation analyses with the cluster of GABAA receptor
zepine receptor function in the brains of alcoholics genes under this linkage peak revealed strong associ-
[95–97] and individuals at risk [98]. Taken together, ation with a large number of SNPs (several at p < 0.001)
these findings suggest GABA deficits probably con- genotyped in GABRA2 for the Caucasian-only subset,
tribute to a state of hyperexcitability in the brains and not with other genes in the cluster (Figure 29.2).
of alcoholics and individuals at risk, and this may Another earlier study [105] reports that three
underlie the predisposition to develop alcoholism. exonic variants of the gene encoding the human
GABAB receptor on chromosome 6 modify scalp-
recorded EEG coherence (cortical synchronization).
Interhemispheric theta coherence: GABRA2 Parietotemporal coherence showed statistical signifi-
and CHRM2 genes cance associated with exon 7, suggesting that this
Another EEG measure that provides a good endo- exon may be functionally meaningful and impact on
phenotype for genetic study is EEG coherence – a cortical EEG oscillations.
measure of cortical synchronization in neural net- These results suggest that GABRA2 may indeed
works, indexing communication in populations of influence susceptibility to alcohol and drug depend-
neurons. Recent studies have suggested a significant ence, not just by modulating level of neural excitation,
role for theta frequency coherence in normal and but also by influencing functional connectivity of
aberrant thalamocortical interactions [99, 100]. interhemispheric networks. In their model of thala-
Impairments in neural synchrony have been reported mocortical dysrhythmia, Linas and co-workers have
in several psychiatric conditions, including alcohol proposed that the enhanced low-frequency (theta)
dependence. There is evidence in the literature that oscillations in the thalamocortical module can affect
not only do alcoholics manifest differences in EEG the lateral inhibitory drive in the cortex and eventu-
power in specific frequency bands, but they also ally result in high frequency coherent activation of
manifest increased interhemispheric coherence [101, cortical module [106]. This is particularly significant
102]. It has been reported that bilateral intrahemi- in light of our genetic findings where two resting state
spheric coherences in alpha and beta frequency electrophysiological signatures – beta power (high
bands were increased in both long-term abstinent frequency activity associated with arousal) and theta
and nonabstinent alcoholics compared to controls coherence (low frequency synchrony) are both linked
[103], particularly for alpha when depressiveness to the same GABAA receptor gene.
was included as a covariate; there was no effect of The same interhemispheric high theta coherence
length of abstinence on these findings. In our labora- phenotype at centroparietal leads indicated significant
tory we have observed significant increases in resting linkage on chromosome 7 and significant associations
EEG interhemispheric high theta (6–7 Hz) coherence with a muscarinic acetylcholine receptor (M2) gene
in alcohol-dependent subjects when compared to (CHRM2), underlying the linkage peak using family-
normal controls, particularly posteriorly at parietal- based association tests [104] (Figure 29.2). Significant
occipital and centroparietal regions [Rangaswamy linkage and association for evoked theta band
et al., unpublished]. Peak theta band coherence was responses at the same CHRM2 gene have been previ-
highest for posterior electrode pairs in alcohol ously reported in COGA [107, 108] (see below).
dependent subjects, displaying a shift from a more Both GABAergic and cholinergic systems interact
frontocentral prominence as seen in normal control significantly in the functions of local inhibitory cir-
subjects. Hence, this increased EEG coherence (cor- cuits, thus affecting network functions and influen-
tical synchronization) observed in alcoholics may be cing cortical synchronization. Increased GABAergic

356
Chapter 29: Endophenotypes in psychiatric genetics

inhibition is likely to be a mechanism underlying reciprocal synchronization has been observed in


impaired synaptic plasticity observed with M2 knock- the theta range between hippocampus and frontal
out mice, who demonstrate impaired behavioral flexi- and parietal regions in the brain during attentional
bility and memory deficits [109]. A recent study tasks [66]. Delta is related to signal detection and
suggested that M2 receptor activation produces a decision-making, and is generated by cortico-
presynaptic inhibition of GABA release by long-range cortical interactions and is prominent after the
inhibitory neurons of the perirhinal cortex projecting target stimuli.
to the entorhinal cortex [110]. In a visual oddball paradigm, alcoholics mani-
fest significantly less evoked theta and delta ERO
Theta and delta event-related oscillations underlying P3 amplitudes while processing the target stimuli
during target detection: CHRM2 and GRM8 genes [121]; these findings are most significant anteriorly
One of the most consistent robust findings in the for theta, and posteriorly for delta. In order to
literature is the reduced P3 amplitude in alcoholics determine whether these deficits in theta and delta
and in offspring at high risk prior to alcohol exposure oscillations antecede the development of alcoholism
[111] (for review see [15]), providing a good endo- we examined adolescent high risk children of alco-
phenotype for genetic studies. However, the P3 com- holics compared to normal children, using the same
ponent is not a unitary phenomenon, but emanates paradigm [122]. The results showed that the adoles-
from multiple sources in the brain with contributions cent offspring of alcoholics have reduced delta and
from frontal cortex (including anterior cingulate), theta band ERO amplitude (underlying P3) while
amygdala and hippocampus [112–115]. Low P3 processing the target stimuli compared to controls;
amplitudes coupled with weaker and less well-organ- differences were most prominent centroparietally
ized sources in alcoholics and offspring at risk suggest for theta, and parietally for delta. The ERO meas-
inefficient allocation of resources during neural pro- ures have provided robust group differentiation and
cessing. This undifferentiated neurophysiological pat- at times function as more sensitive measures than
tern suggests a level of cortical disinhibition in P3 in differentiating between high risk and low risk
alcoholics and individuals at risk. The low P3 ampli- offspring. Thus, the results of these two studies
tude is not only observed in abstinent alcoholics and demonstrate that decreased theta and delta EROs
offspring of alcoholics, but is also present in various to target stimuli may antecede the development of
disinhibitory conditions, such as substance abuse, alcoholism and represents a strong trait marker
ASP, conduct disorder, and ADHD. Moreover, indi- (Figure 29.3).)
viduals with low P3 amplitudes manifest a signifi- We have reported significant linkage and associ-
cantly higher incidence of externalizing disorders ation between the CHRM2 gene on chromosome 7
and disinhibitory traits than those with high P3 and frontal theta oscillations to target stimuli in a
amplitudes [15]. visual oddball task; association analyses using both
More recently, event-related activity recorded population-based tests (Measured Genotype) and
during cognitive tasks, including paradigms pedigree-based tests (Quantitative Pedigree Disequi-
eliciting the P3 component, has been examined in librium Test, QPDT) indicate significant association
the frequency domain (i.e. EROs). Several studies of the frontal theta band ERO phenotype with several
have demonstrated that P3 responses are primarily SNPs surrounding exon 4 of CHRM2. Further, an
the outcome of theta and delta EROs elicited during examination of the slower frequency parietal-occipital
cognitive processing of stimuli [116–120]. Topo- delta band EROs also revealed significant association
graphically, delta ERO power peaks at the posterior with several SNPs [107, 108] (Figure 29.3).
region, while the theta power peak is located in the These findings implicate a role of CHRM2 in the
frontocentral region [119]; theta oscillations also generation and modulation of evoked oscillations.
contribute strongly to N2 components of ERPs. Theta and delta EROs depend on the level of acetyl-
Theta and delta EROs underlying P3 are related to choline (muscarinic activation). M2 receptors inhibit
different cognitive functions: Theta is associated presynaptic release of acetycholine, leading to inhib-
with memory processes and attention, and involves ition of irrelevant networks. Muscarinic receptors are
fronto-limbic or cortico-hippocampal interactions, especially concentrated in the forebrain and possibly
and is taken as an index of frontal processing; serve to maintain the effective balance of relevant/

357
Chapter 29: Endophenotypes in psychiatric genetics

(a)
Adult subjects
Control Alcohol dependent
Frequency (Hz)

Frequency (Hz)
10 10
θ θ θ
5 δ 5
δ
0 0
0 200 400 600 0 200 400 600
Time δ Time

1 2 3 4 1 2 3 4
Amplitude (μV) Amplitude (μV)

(b) Adolescent subjects


Low risk High risk
12

Frequency (Hz)
12
Frequency (Hz)

θ
7
7 θ
θ
δ 0
δ
0
0 200 400 600 0 200 400 600
Time
Time δ
2 4 6 2 4 6
Amplitude (μV) Amplitude (μV)

Figure 29.3 S-transform derived time-frequency representations of the average instantaneous amplitude that are z-scored for each
frequency. Instantaneous amplitudes were averaged across individual trials of subjects so that nonphase locked or imprecise phase locked
oscillatory energy is preserved. (a) Plots for the target condition at central (Cz) electrode in 120 alcoholic (right) and 120 control (left)
subjects. Center panel – Topographic headplots of the time-frequency regions of interest (TFROIs) indicate that the theta band (4–5 Hz)
power at 300–500 ms shows frontal maxima while the delta band (1–3 Hz) power at 350–700 ms has posterior maxima. Note that
alcoholics have weaker responses in both theta and delta bands. (b) Plots for the target condition at frontal (Fz) electrode in visual oddball task
in 87 high risk adolescent offspring of alcoholics from Collaborative Study on the Genetics of Alcoholism (COGA) families (right) and 57
matched low risk offspring of nonalcoholics from control families (left). Center panel – Topographic head plots for the TFROI that extends
from 300 to 700 ms for each band revealing frontal maxima for theta band power and parietal maxima for delta band power. Note that
high risk adolescents have weaker responses in both theta and delta bands to target stimuli, similar to the alcoholics. See plate section
for color version.

irrelevant networks, hence, having a direct influence Wechsler Adult Intelligence Scale-Revised (WAIS-R).
on P3 generation [123]. Genetic underpinnings of These results remain significant after taking into
evoked oscillations are likely to stem from regulatory account alcohol dependence and depression diagnoses
genes that control the neurochemical processes of the in the sample [126].
brain and, therefore influence neural function. The Evidence from the COGA project indicates that
three major neurochemical substrates contributing to the CHRM2 gene is not only associated with brain
theta and delta rhythms and P3 involve strong oscillations and cognition, but also clinical diagnosis.
GABAergic, cholinergic, and glutamatergic system Significant linkage and association were reported for
interactions [123]. the CHRM2 gene and a diagnosis of alcohol depend-
Moreover, the cholinergic muscarinic genes have a ence and depression [128], comorbid alcohol and
major role in memory and cognition [124, 125]. Sev- drug dependence (a more severe addiction profile)
eral studies, including the COGA study, have found [129], as well as a spectrum of externalizing disorders
evidence that the CHRM2 gene may be involved in [130]. Other groups have replicated these findings,
intelligence [125–127]. In the COGA study, we found reporting that the CHRM2 gene predisposes to alco-
evidence of association with multiple SNPs across hol dependence, drug dependence and affective dis-
CHRM2 and Performance IQ, as measured by the orders [131], and major depression in women [132].

358
Chapter 29: Endophenotypes in psychiatric genetics

Thus genes important for the expression of the endo- targeted by mGluR8-positive boutons. The postsy-
phenotype (brain oscillations) help in identification of naptic interneuron type-specific expression predicts
genes that increase the susceptibility for risk of alco- a role in adjusting the activity of interneurons
hol dependence and related disorders [68, 133]. depending on the level of network activity.
Another likely candidate gene located under the
same theta ERO linkage peak on chromosome 7 is a
metabotropic glutamate receptor (GRM8) gene. The Conclusions
glutamatergic system is one of the major players Psychiatric disorders result from a complex inter-
modulating CNS electrophysiological networks, action of changing genetic and environmental liabil-
and in particular is also involved in theta oscilla- ities across development, possibly with greater genetic
tions and P3 [123]. Family-based association ana- loading in individuals who have early onset of the
lyses of theta EROs revealed significant associations disorders. The use of quantitative endophenotypes
with several SNPs in the GRM8 gene and theta provides the power to more easily localize and char-
EROs to target stimuli at frontal, central, and par- acterize disease susceptibility genes than diagnostic
ietal regions [134]. An interesting finding is that categories. The measures of these endophenotypes
several SNPs were also significantly associated with reflect the genetic liability of the disorder among
the diagnosis of alcohol dependence using ICD-10 nonaffected relatives of affected individuals. Many of
diagnostic criteria. the same genes important for the expression of the
A 3-T proton magnetic resonance spectroscopy endophenotypes help in identification of genes that
(1H-MRS) study has suggested the involvement of increase the susceptibility for risk of the disease.
glutamatergic neurotransmission in integrative Hence, as this chapter has illustrated, the utility of
frontal-hippocampal processing [135] and the sensa- quantitative biological endophenotypes for the study
tion-seeking personality dimension [136]. The study of genetic risk of psychiatric disorders continues to be
demonstrated a robust relationship between glutam- very promising.
ate levels in the hippocampus and frontal theta Endophenotypes may have additional uses in
activity during auditory stimulus processing. Gluta- psychiatry, including uses in diagnosis, classification,
matergic neurotransmission and its neuroadaptive and the development of animal models. The lack of a
changes have been proposed as important molecular biological basis for the classification of psychiatric
determinants of craving and relapse [137, 138]. In disorders has led to limited success in research in
particular, it is suggested that a hyperglutamatergic the neurobiology and genetics of psychiatric dis-
state mediates, at least in part, alcohol relapse behav- orders. Endophenotype-based analysis would be
ior and maintenance of alcoholism [139]. Several useful for establishing a biological underpinning for
studies have suggested the involvement of glutamate diagnosis and classification; a net outcome would be
receptors – NMDA and metabotropic, in alcohol improved understanding of the neurobiology and
relapse [68, 140, 141]. Acamprosate, a drug used to genetics of psychopathology.
prevent relapse in alcoholic patients [142], has While the endophenotype approach is not a new
been suggested to act through a suppression of a idea, it is an approach whose time has come. Because
hyperglutamatergic state created by alcohol addiction of the rapidly evolving technology in the fields of
[143, 144]. molecular and statistical genetics, new methods are
In light of the theta oscillations showing strong emerging that will facilitate the search for genes
association to both CHRM2 and GRM8 genes, one underlying complex traits in the near future. With
could speculate a synergistic genetic mechanism the advent of SNP genotyping technology, genome-
underlying this electrophysiological phenotype, thus wide association studies using dense sets of SNPs
opening doors to future research in this direction. across the genome, together with recent advances in
Interestingly, in studies on the rat hippocampus, novel statistical genetic techniques and computational
authors have reported that a majority of interneurons power, have facilitated the identification of genes that
strongly immunopositive for the muscarinic M2 or are associated with heritable, genetically influenced
the mGlu1 receptors were the primary targets of quantitative traits that are risk factors involved in
mGluR8-containing terminals [145]. Rare neurons the etiology of various psychiatric disorders. Once
coexpressing calretinin and M2 were consistently genes are identified and understood, risk genotypes

359
Chapter 29: Endophenotypes in psychiatric genetics

and haplotypes can be studied in prospective studies of how genes contribute to susceptibility, which in
of young individuals who have not yet developed the turn can lead to the design of well-targeted prevention
disease, and can lead to an improved understanding initiatives.

References 21. Freedman R, et al. Biol Psychiatry


1983;18:537–551.
42. Hovatta I, et al. Am J Hum Genet
1999;65:1114–1124.
1. Akerberg E. Hereditas
1986;105:1–5. 22. Siegel C, et al. Arch Gen Psychiatry 43. Egan MF, et al. Proc Natl Acad Sci
1984;41:607–612. U S A 2001;98:6917–6922.
2. Gottesman II, et al. Schizophrenia
and Genetics: A Twin Study 23. Clementz BA, et al. Am 44. Sklar P. Annu Rev Genomics Hum
Vantage Point. New York: J Psychiatry 1998;155:1691–1694. Genet 2002;3:371–413.
Academic Press; 1972. 24. Myles-Worsley M. Am 45. Tunbridge EM, et al. Biol
3. Lewis DA. Am J Psychiatry J Psychiatry 2002;159:2007–2012. Psychiatry 2006;60:141–151.
2002;159:1467–1469. 25. Myles-Worsley M, et al. Biol 46. Weinberger DR, et al. Biol
4. Kandel ER. Am J Psychiatry Psychiatry 1996;39:289–295. Psychiatry 2001;50:825–844.
1998;155:457–469. 26. Freedman R, et al. Proc Natl Acad 47. Doyle AE, et al. J Child Psychol
5. Tsuang MT. Am J Psychiatry Sci U S A 1997;94:587–592. Psychiatry 2005;46:774–803.
2000;157:489–491. 27. Freedman R, et al. Am J Med 48. Doyle AE, et al. Biol Psychiatry
6. Gottesman, II, et al. Am Genet 2001;105:20–22. 2005;57:1324–1335.
J Psychiatry 2003;160:636–645. 28. Leonard S, et al. Arch Gen 49. Rommelse NN. Expert Rev
7. John B, et al. Science Psychiatry 2002;59:1085–1096. Neurother 2008;8:1425–1429.
1966;152:711–721. 29. Munoz DP. Prog Brain Res 50. Meyer-Lindenberg A, et al. Mol
8. Munafo MR. Genes Brain Behav 2002;140:89–96. Psychiatry 2006;11:867–877, 797.
2006;5 Suppl 1:3–8. 30. Hong LE, et al. Arch Gen 51. Callicott JH, et al. Am J Psychiatry
9. Skuse DH. Br J Psychiatry Psychiatry 2006;63:259–264. 2003;160:709–719.
2001;178:395–396. 31. Kathmann N, et al. Am 52. Mattay VS, et al. Proc Natl Acad
10. Waldman ID. Biol Psychiatry J Psychiatry 2003;160:696–702. Sci U S A 2003;100:6186–6191.
2005;57:1347–1356. 32. Lin PI, et al. Schizophr Bull 53. Meyer-Lindenberg A, et al. Nat
11. Walters JT, et al. Mol Psychiatry 2008;34:791–797. Neurosci 2005;8:594–596.
2007;12:886–890. 33. Arolt V, et al. Am J Med Genet 54. van ‘t Ent D, et al. Neuroimage
12. Almasy L, et al. Am J Med Genet 1996;67:564–579. 2007;35:1004–1020.
2001;105:42–44. 34. Matthysse S, et al. Am J Med Genet 55. Prasad KM, et al. Schizophr Bull
13. Leboyer M, et al. Trends Neurosci B Neuropsychiatr Genet 2008;34:774–790.
1998;21:102–105. 2004;128B:30–36.
56. Burmeister M, et al. Nat Rev Genet
14. Almasy L, et al. Am J Hum Genet 35. Cannon TD, et al. Am J Hum 2008;9:527–540.
1998;62:1198–1211. Genet 2000;67:369–382.
57. Kauwe JS, et al. Ann Neurol
15. Porjesz B, et al. Clin Neurophysiol 36. Conklin HM, et al. Am 2007;61:446–453.
2005;116:993–1018. J Psychiatry 2000;157:275–277. 58. Lindqvist D, et al.
16. Braff DL, et al. 37. Park S, et al. Arch Gen Psychiatry Psychoneuroendocrinology
Psychopharmacology (Berl) 1995;52:821–828. 2008;33:1061–1068.
2001;156:234–258. 38. Calkins ME, et al. Schizophr Bull 59. Mann JJ, et al. Biol Psychiatry
17. Braff DL, et al. Arch Gen 2007;33:33–48. 2009;65:556–563.
Psychiatry 1990;47:181–188. 39. Gur RE, et al. Schizophr Bull 60. Bartels M, et al.
18. Geyer MA, et al. Brain Res Bull 2007;33:49–68. Psychoneuroendocrinology
1990;25:485–498. 40. Greenwood TA, et al. Arch 2003;28:121–137.
19. Swerdlow NR, et al. Biol Gen Psychiatry 61. Federenko IS, et al. J Clin
Psychiatry 1993;33:298–301. 2007;64:1242–1250. Endocrinol Metab
20. Swerdlow NR, et al. J Neurol 41. Gasperoni TL, et al. Am J Med 2004;89:6244–6250.
Neurosurg Psychiatry Genet B Neuropsychiatr Genet 62. Derijk RH, et al. Eur J Pharmacol
1995;58:192–200. 2003;116B:8–16. 2008;585:492–501.

360
Chapter 29: Endophenotypes in psychiatric genetics

63. Gotlib IH, et al. Biol Psychiatry 85. Drgon T, et al. Am J Med Genet Oscillatory Properties of
2008;63:847–851. B Neuropsychiatr Genet Mammalian Central Neurons In
64. Uhart M, et al. 2006;141:854–860. Vivo. London: Macmillan; 1984.
Neuropsychopharmacology 86. Enoch MA, et al. Am J Med Genet 107. Jones KA, et al. Behav Genet
2006;31:2255–2263. B Neuropsychiatr Genet 2006;36:627–639.
65. Kendler KS, et al. Arch Gen 2006;141:599–607. 108. Jones KA, et al. Int J Psychophysiol
Psychiatry 2003;60:929–937. 87. Fehr C, et al. Psych Genet 2004;53:75–90.
66. von Stein A, et al. Int 2006;16:9–17. 109. Seeger T, et al. J Neurosci
J Psychophysiol 2000;38:301–313. 88. Lappalainen J, et al. Alcohol Clin 2004;24:10117–10127.
67. van Beijsterveldt CE, et al. Am Exp Res 2005;29:493–498. 110. Apergis-Schoute J, et al. J Neurosci
J Hum Genet 1996;58:562–573. 89. Soyka M, et al. J Psychiatr Res 2007;27:4061–4071.
68. Bachteler D, et al. 2008;42:184–191. 111. Begleiter H, et al. Science
Neuropsychopharmacology 90. Tobler I, et al. Proc Natl Acad Sci 1984;225:1493–1496.
2005;30:1104–1110. U S A 2001;98:6464–6469. 112. Ardekani BA, et al. Brain Res Cogn
69. Bauer LO. 91. Haenschel C, et al. Proc Natl Acad Brain Res 2002;14:347–356.
Neuropsychopharmacology Sci U S A 2000;97:7645–7650. 113. Halgren E, et al. Science
2001;25:332–340. 1980;210:803–805.
92. Faulkner HJ, et al. Br
70. Costa L, et al. Drug Alcohol J Pharmacology 1999;128:1813– 114. Kiehl KA, et al. Schizophr Res
Depend 1997;46:87–93. 1825. 2001;48:159–171.
71. Propping P, et al. Hum Genet 93. Kopell N, et al. Proc Natl Acad Sci 115. Menon V, et al. Neuroreport
1981;59:51–59. U S A 2000;97:1867–1872. 1997;8:3029–3037.
72. Rangaswamy M, et al. Biol 94. Whittington MA, et al. Int 116. Basar-Eroglu C, et al. Int
Psychiatry 2002;52:831–842. J Psychophysiol 2000;38:315–336. J Psychophysiol 1992;13:161–179.
73. Winterer G, et al. Psychiatry Res 95. Abi-Dargham A, et al. Am 117. Basar E, et al. IEEE Eng Med Biol
1998;78:101–113. J Psychiatry 1998;155:1550–1555. Mag 1999;18:56–66.
74. Finn PR, et al. Alcohol Clin Exp 96. Krystal JH, et al. Arch Gen 118. Karakas S, et al. Neuroscience Lett
Res 1999;23:256–262. Psychiatry 2006;63:957–968. 2000;285:45–48.
75. Gabrielli WF, Jr, et al. 97. Lingford-Hughes AR, et al. Br 119. Karakas S, et al. Clin Neurophysiol
Psychophysiology J Psychiatry 1998;173:116–122. 2000;111:1719–1732.
1982;19:404–407. 120. Yordanova J, et al. Neuroreport
98. Volkow ND, et al. Alcohol Clin
76. Rangaswamy M, et al. Neuroimage Exp Res 1995;19:510–516. 1996;8:277–280.
2004;21:329–339. 121. Jones KA, et al. Clin Neurophysiol
99. Sarnthein J, et al. J Neurosci
77. Bauer LO, et al. J Stud Alcohol 2007;27:124–131. 2006;117:2128–2143.
1993;54:577–589. 100. Sarnthein J, et al. Int 122. Rangaswamy M, et al. Int
78. Begleiter H, et al. Alcohol Clin Exp J Psychophysiol 2005;57:87–96. J Psychophysiol 2007;63:3–15.
Res 1999;23:1125–1135. 101. Kaplan RF, et al. J Stud Alcohol 123. Frodl-Bauch T, et al.
79. Porjesz B, et al. Proc Natl Acad Sci 1985;46:122–127. Neuropsychobiology
U S A 2002;99:3729–3733. 1999;40:86–94.
102. Michael A, et al. Acta Psychiatr
80. Edenberg HJ, et al. Am J Hum Scand 1993;87:213–217. 124. Calabresi P, et al. Eur
Genet 2004;74(4):705–714. 103. Winterer G, et al. Acta Psychiatr J Neuroscience
81. Agrawal A, et al. Behav Genet Scand 2003;108:51–60. 1998;10:3020–3023.
2006;36(5):640–650. 104. Porjesz B, et al. Scientific World J 125. Comings DE, et al. Mol Psychiatry
82. Dick DM, et al. J Stud Alcohol 2007;7:131–141. 2003;8:10–11.
2006;67:185–194. 105. Winterer G, et al. Am J Med Genet 126. Dick DM, et al. Behav Genet
83. Dick DM, et al. Behav Genet B Neuropsychiatr Genet 2007;37:265–272.
2006;36:577–590. 2003;117:51–56. 127. Gosso FM, et al. BMC Med Genet
84. Covault J, et al. Am J Med Genet 106. Linas R. Rebound Excitation as the 2007;8:66.
B Neuropsychiatr Genet Physiological Basis for Tremor: 128. Wang JC, et al. Hum Mol Genet
2004;129:104–109. A Biophysical Study of the 2004;13:1903–1911.

361
Chapter 29: Endophenotypes in psychiatric genetics

129. Dick DM, et al. Addiction Genet 2009;150B: 140. Holter SM, et al. Pharmacol
2007;102:1131–1139. 359–368. Biochem Behav 2000;
130. Dick DM, et al. Arch Gen 135. Gallinat J, et al. 66:143–151.
Psychiatry 2008; Psychopharmacology 141. Krystal JH, et al. Pharmacol Ther
65:310–318. 2006;187:103–111. 2003;99:79–94.
131. Luo X, et al. Hum Mol Genet 136. Gallinat J, et al. Neuroimage 142. Mann K, et al. Alcohol, Clin
2005;14:2421–2434. 2007;34:671–678. Experiment Res 2004;
132. Comings DE, et al. Am 137. Cornish JL, et al. J Neurosci 28:51–63.
J Med Genet 2002; 2000;20:RC89. 143. Dahchour A, et al. Prog Neurobiol
114:527–529. 138. Tzschentke TM, et al. Crit 2000;60:343–362.
133. Begleiter H, et al. Int Rev Neurobiol 2000; 144. Spanagel R. Addiction
J Psychophysiol 2006;60:162–171. 14:131–142. 2005;100:1813–1822.
134. Chen AC, et al. Am J Med 139. Tsai G, et al. Annu Rev Med 145. Ferraguti F, et al. J Neurosci
Genet B Neuropsychiatr 1998;49:173–184. 2005;25:10520–10536.

362
Developmental disorders
Chapter

30 Craig A. Erickson, Khendra I. Peay, and Christopher J. McDougle

Introduction dementia among other disorders [3]. Despite being


associated with language impairment and intellectual
This chapter will review recent evidence describing disability, many persons with Down syndrome show
the genetic underpinnings of several developmental good social communication skills [1]. However, there
disorders. While a full discussion of the genetics of is also a subgroup of Down patients with autism
autistic disorder occurs elsewhere in this text, the spectrum disorders [1]. About 7–15% of persons with
phenotypic overlap between autism and other devel- Down syndrome additionally meet diagnostic criteria
opmental disorders whose genetic underpinning are for an autism spectrum disorder (ASD), a rate that is
more readily understood will be highlighted in this more than 10 times the rate of ASD in the general
chapter. Discussion of Down syndrome, Rett’s dis- population [1, 4–7].
order, Prader–Willi syndrome, Angelman syndrome, Down syndrome is caused by genomic-dosage
Smith–Magenis syndrome, and fragile X syndrome imbalance [8]. Since there are 3 copies of chromo-
will place the focus of this chapter on disorders with some 21 rather than 2, it would be expected that in
well characterized genetic abnormalities combined affected persons, 1.5-fold gene expression would
with phenotypic overlap with autism. The genetics occur for genes coded on that chromosome [8]. While
of these well characterized developmental disorders on average gene expression on chromosome 21 in
provide a window of understanding to aide investi- Down syndrome is upregulated near the expected
gation into specific genetic contributions to autism. 1.5-fold rate, great variation in gene expression has
Additionally, given the relative genetic homogeneity been found in Down syndrome cell line analysis [8].
of these disorders, investigators can begin to better This gene expression variability is likely related to the
understand the link from gene(s) to neurochemistry/ significant phenotypic variability that is characteristic
neuroanatomy to behavior. of the disorder [8]. In Down syndrome cell line analy-
sis, only between 39 and 62% of genes on chromo-
Down syndrome: disorder of gene some 21 showed significant upregulation compared
to control samples [8].
dosage Utilizing cell line gene expression analysis in
Down syndrome represents the most common known Down syndrome, researchers have been able to char-
cause of developmental disability. Down syndrome acterize genes on chromosome 21 as: (1) potentially
occurs in 1 in 1000 live births and results from tri- characteristic of core features of Down syndrome; (2)
somy of chromosome 21 [1]. Down syndrome rates characteristic of phenotypic variability seen in Down
increase with advancing maternal age. Physical syndrome; and (3) not likely involved in the presen-
markers of the disorder include hypotonia, flat facies, tation of Down syndrome [8]. Genes whose expres-
slanted palpebral fissures, small ears, and single or sion is consistently higher in Down syndrome cell
Simian palmar crease [2]. Individuals with Down lines compared to controls are genes that likely
syndrome are also at increased risk for cardiac anom- involve aspects of the Down syndrome phenotype
alies, thyroid dysfunction, acute leukemia, seizures, that exist in all affected persons [8]. Genes in this
celiac disease, diabetes mellitus, and early onset of group are potentially involved in nuclear signaling

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

363
Chapter 30: Developmental disorders

and apoptosis, post-translational modification of pro- In 1999, mutations at Xq28 within the gene encoding
teins, regulation of DNA replication, and interferon methyl-CpG-binding protein 2 (MECP2) were identi-
signaling pathways [8]. Genes with partially overlap- fied as a major cause of Rett’s disorder [10]. In about
ping expression between Down syndrome and control 70–80% of clinical cases of Rett’s disorder, mutations
samples may represent genes that are associated with in this gene exist [10]. Rare familial cases of Rett’s
the variable phenotypic expression of Down syn- disorder do exist when carrier females with signifi-
drome [8]. Many genes representing between one-half cantly skewed X inactivation patterns remain asymp-
and one-third of genes expressed from chromosome tomatic or mildly affected and then potentially pass
21 fall into this category [8]. Genes with similar cell on their MECP2 mutation to their children [11].
line expression in Down syndrome cells and controls Rett’s disorder is marked by apparently normal
are thought to likely not contribute to the disorder. prenatal and perinatal development, normal psycho-
The gene encoding for amyloid precursor protein motor development through the first 5 months of life,
(APP), which has been shown to be dosage sensitive normal head circumference at birth, and deceleration
in the brain and is known to confer risk for early- of head growth between ages 5 and 48 months [12].
onset dementia, is included among genes whose Rett’s disorder is also defined by losses in previously
dosage mirrors control samples in Down syndrome gained skills including loss of purposeful hand move-
fibroblast and lymphoblast cell line analysis [8]. This ments between ages 5 and 30 months with the subse-
result is contrary to the expectation that the gene for quent development of stereotypic hand movements
APP would be upregulated in Down syndrome cell (classic hand ringing or hand washing) and loss of
lines given the association between Down syndrome social engagement early in the disorder [12]. Severe
and early dementia. This finding points to the likely psychomotor retardation, severely impaired language
need for cell line analysis in Down syndrome beyond skills, and poor coordination also mark the clinical
utilization of fibroblast and lymphoblast cells [8]. presentation of the disorder [12]. Individuals with
Given the developmental nature of some aspects of Rett’s disorder frequently also have seizures, a dis-
the Down syndrome phenotype, in particular the ordered sleep pattern, and bruxism [13, 14].
onset of dementia, future cell line work in this field With allelic heterogeneity marked by over 200
may also need to utilize cell lines at different develop- specific possible mutations identified in the MECP2
mental stages/ages [8]. gene accounting for the majority of Rett’s disorder
Down syndrome is a phenotypically variable, rela- cases, some work has been done to test for association
tively common cause of developmental disability. of specific mutations with clinical features [13]. In a
Much remains to be done in mapping the contribu- study of 135 females with Rett’s disorder in a cohort
tion of specific genes on chromosome 21 to the from the United Kingdom, the overall mutation effect
pathopysiology of the disorder. Future work in this on clinical presentation was found to be minimal [13].
area will likely aid understanding not only of the Other reports have noted that mutations more prox-
cause of core features of Down syndrome, but also imal in the MECP2 gene are associated with more
help to define potential risk genes involved in the significant neurological impairment and gastrointest-
phenotypic variability of the disorder. Identifying inal problems [14]
those genes which may define variable presentations MECP2 functions in initiating and/or maintaining
of the disorder will likely aid in the understanding of neuronal maturation [15]. In animals, expression of
the genetics of autistic disorder given the over-repre- MECP2 has been shown to correlate with neuronal
sentation of autism in persons with Down syndrome. maturation immediately prior to synaptogenesis [16,
17]. These animal findings correlate well with the
Rett’s disorder: persistence of genetic clinical course of Rett’s disorder with symptom onset
during a period of rapid human neuronal develop-
heterogeneity ment and growth marked by synaptogenesis [14].
Rett’s disorder is an X-linked neurodevelopmental While the genetic underpinnings of Rett’s disorder
disorder that affects approximately 1 in 10 000 most often relate to mutations in a single gene, the
females [9]. Rett’s disorder rarely occurs in males disorder is still marked by some significant pheno-
and it is thought that genetic mutations associated typic variation, and a significant minority of cases,
with Rett’s are often not compatible with life in males. 20–30%, lack classic MECP2 mutations. While variant

364
Chapter 30: Developmental disorders

X inactivation could play a part in this presentation, Angelman syndrome generally results from
other factors not yet quantified (variable gene expres- maternally derived chromosomal deletion also in the
sion, epistasis, gene–environment interactions, etc.) 15q11–13 region [18]. The prevalence of Angelman
likely contribute to this occurrence. It is likely that syndrome is between 1 in 10 000–40 000 births [24].
MECP2 deficits represent only one means by which a The majority (68–75%) of Angelman syndrome cases
neurobiological effect leads to the classic clinical pre- are due to maternally derived gene deletions that
sentation of the disorder. impact expression of the gene UBE3A [25]. Approxi-
mately 2–7% of cases are due to paternal uniparental
Prader–Willi and Angelman syndromes: disomy, 2–5% due to imprinting center mutations,
and 8–11% due to specific maternally inherited muta-
a study in genetic imprinting tions in the UBE3A gene [25]. UBE3A codes for a
Prader–Willi and Angelman syndromes represent dis- HECT-domain ubiquitin ligase involved in synapto-
orders of genomic imprinting. Although people genesis whose substrates and function are not com-
inherit one seemingly equivalent gene copy from each pletely understood [22, 26]. Work utilizing knockout
parent, in some cases gene pairs are not functionally drosophila and mouse models is underway to better
equivalent (i.e. they are “imprinted”) as gene expres- understand the implication of deficient UBE3A
sion is based upon parent of origin. Imprinting is expression [26, 27]. Angelman syndrome also repre-
termed epigenetic because the production of func- sents an example of potential locus heterogeneity
tional protein is governed by factors that regulate because approximately 12% of people with a clinical
protein expression and not the actual nucleotide Angelman syndrome diagnosis do not have identifi-
sequence [18]. Genomic imprinting has been identi- able chromosome 15 abnormalities [28]. This points
fied in about 30 genes [18], including those on to the possibility that other proteins involved with
chromosome 15q11–13 that result in Prader–Willi UBE3A could manifest mutations that also result in
and Angelman syndromes. the Angelman syndrome phenotype.
Prader–Willi syndrome is a rare genetic disorder Clinical characteristics found in all persons with
with a prevalence of 1 in 10 000–15 000 births [19]. Angelman syndrome include developmental delay,
Newborns with Prader–Willi syndrome show hypo- speech impairment, movement disorder, and “behav-
tonia and failure to thrive followed within the first few ioral uniqueness” defined by a combination of fre-
years of life by hyperphagia, food preoccupations, and quent laughter/smiling, apparent happy demeanor,
frequent obesity [18]. Prader–Willi syndrome is also easily excited personality, hypermotoric behavior,
marked by mild to moderate mental retardation and and short attention span [29]. Frequent characteris-
frequently with temper tantrums, aggression, and tics found in about 80% of affected persons include
obsessive–compulsive behaviors including skin microcephaly, seizures, and abnormal electroenceph-
picking [20, 21]. Prader–Willi syndrome significantly alogram (EEG) [29]. Sleep problems and feeding dif-
overlaps with autism: approximately one in four per- ficulty are also frequently seen [30]. Angelman
sons with Prader–Willi syndrome have a comorbid syndrome also has significant overlap with autism
autism spectrum disorder (ASD) [22]. Prader–Willi with estimates utilizing standard diagnostic instru-
syndrome is caused by the absence of paternally ments reporting an ASD comorbidity rate of 42–
derived imprinted genes within the 15q11–13 region 81% [31, 32]. However, ASD rates in pooled case
[23]. While 70% of Prader–Willi syndrome cases are series using clinical ASD diagnoses are as low as
derived from paternally derived chromosomal dele- 1.9–3.6% [22]. The wide range of estimated ASD
tions, the rest of nondeletion cases are caused by comorbidity in Angelman syndrome may be due in
uniparental disomy in which both copies of chromo- some cases to potential ASD over-diagnosis due to the
some 15 are derived from the mother [23]. Addition- very low mental age of many persons with Angelman
ally, a small minority of cases are derived from syndrome [31].
mutations in a separate locus controlling gene Given the phenotypic overlap between ASD, Pra-
imprinting [23]. The final common effect of all der–Willi, and Angelman syndrome, it is not surpris-
abnormalities leading to Prader–Willi syndrome is a ing that in some otherwise idiopathic ASD cases,
lack of paternally derived gene expression from the chromosome 15q11–13 abnormalities have been
15q11–13 region. implicated. In particular, the reports of maternally

365
Chapter 30: Developmental disorders

derived duplications at 15q11–13 in ASD has gener- marked by spasmodic tensing of the upper body and
ated the hypothesis that, among cases of Prader–Willi clasping of the hands at the chest while tightly squeez-
or Angelman syndrome, the rate of ASD would be ing the arms is thought to be a behavior potentially
highest in persons with uniparental disomy Prader– unique to SMS [40]. Eighty percent of SMS patients
Willi syndrome due to increased expression of mater- exhibit self-injury, including repetitive fingernail
nal 15q11–13 in these cases [22]. In a pooled analysis picking, wrist-biting, head-banging, and insertion of
assessing reports on the comorbidity of ASD in these foreign objects into the ear [6]. Additional maladap-
disorders, Veltman et al. noted a significantly tive behavior frequently seen in SMS includes affect-
increased prevalence of ASDs in uniparental disomy ive lability, physical aggression, nervousness, and
Prader–Willi syndrome (38%) compared to deletion intense attention-seeking behavior [41]. It is esti-
Prader–Willi cases (18%) [22]. Additionally in this mated that 80–100% of persons with SMS meet cri-
report, uniparental disomy Prader–Willi syndrome teria for an ASD [6].
cases had a significantly increased rate of comorbid Analysis of the impact of RAI1 haploinsufficiency
ASD compared to a combined group of deletion has been conducted in a cell line study [33]. Testing
Prader–Willi cases and all cases of Angelman syn- utilizing lymphoblastic cell lines from SMS patients
drome (group ASD rate of 20%) [22]. and HEK293T transfected cells with 50% knockdown
Overall Prader–Willi and Angelman syndromes of RAI1 were utilized to assess for altered gene expres-
represent developmental disorders whose genetic sion associated with insufficient RAI1 [33]. Many
underpinnings are better understood than autism. candidate genes, both upregulated and downregu-
However, despite their relative genetic homogeneity, lated, were identified in this manner. These included
the molecular basis of these disorders remains some- genes implicated in lipid biosynthesis, circadian
what undefined. Molecular understanding has begun rhythm, gene expression, cell growth, and neuronal,
in Angelman syndrome where a specific deficit cardiovascular, renal, and skeletal function [33].
(UBE3A) provides the basis for animal modeling. Overall, RAI1 insufficiency led to dysregulation of
genes impacting a wide variety of homeostatic func-
Smith–Magenis syndrome: a complex tions. Genome-wide expression findings from SMS
also showed significant overlap with genes dysregu-
congenital disorder lated in other disorders including fragile X syndrome,
Smith–Magenis syndrome (SMS) is a complex devel- Prader–Willi syndrome, and Down syndrome [33].
opmental disorder encompassing over 30 clinical fea- This suggests common mechanisms for the overlap-
tures affecting neurological, behavioral, and ping phenotypes of these disorders. Given the broad
metabolic functioning [33]. About 90% of SMS cases clinical presentation of SMS marked by numerous
are due to a 17p11.2 deletion containing the retinoic metabolic, neurologic, medical, and behavioral mani-
acid induced 1 (RAI1) gene [34]. The reported preva- festations, the finding of numerous dysregulated
lence of SMS, 1 in 25 000 births, may be an under- genes in cells with RAI1 haploinsufficiency is not
estimate since the diagnosis is frequently missed surprising. Overlap between SMS and other develop-
because of the phenotypic similarity of the disorder mental disorders, both phenotypically and genetically,
to Down syndrome, Prader–Willi syndrome, and will allow for a better future understanding of the link
fragile X syndrome [33]. between specific gene dysregulation and phenotypic
Physical markers of SMS include short stature, expression.
small hands, obesity, and subtle facial dysmorphism
marked by a short, tented philtrum and relative prog-
nathism [6]. Medical problems common to persons Fragile X syndrome: understanding the
with SMS include hypercholesterolemia/hypertrigly- molecular basis of a developmental
ceridemia, renal abnormalities, thyroid dysfunction,
hypotonia, failure to thrive as an infant, vocal cord disorder
nodules/polyps, hearing loss, and congenital cardiac Fragile X syndrome (FXS) is an example of genetic
anomolies [35–37]. A severe sleep disturbance findings leading to increased molecular understand-
marked by an altered circadian rhythm is commonly ing of disease pathogenesis and then potential dis-
seen in persons with SMS [33, 38, 39]. Self-hugging order-specific treatment modalities. FXS represents

366
Chapter 30: Developmental disorders

the most common inherited form of intellectual dis- significant overlap between FXS and ASDs, under-
ability. It is the result of a cysteine-guanine-guanine standing the molecular basis of FXS may contribute
(CGG) trinucleotide repeat expansion (> 200 repeats) significantly to the understanding of the neurobiology
located at the 5’ untranslated region of the fragile of ASDs.
X mental retardation gene (FMR1) located on the Following the discovery of the FMR1 gene in 1991,
long arm of the X chromosome at Xq27.3. The cyto- a large body of research has demonstrated that FMRP
sines on the CGG islands can be methylated (epigen- is an RNA binding protein associated with actively
etically impacting the pattern of gene expression) translating dendritic ribosomes [51]. This finding has
[18]. The increase in methylation associated with the led to the hypothesis that FMRP may play a role in
expanded CGG repeats correlates directly with the activation-mediated protein synthesis, and thus may
extent of the loss of functional fragile X mental be a regulator of synaptic plasticity [52]. Further work
retardation protein (FMRP) [18]. FXS is inherited in understanding the function of FMRP has focused
from pre-mutation carrier parents (most often on the relationship between FMRP and metabotropic
mothers) who have between 55 and 200 CGG repeats. glutamate (mGluR) receptor activity. Synthesis of
The increase in repeat numbers with ensuing gener- FMRP is enhanced after group 1 mGluR (includes
ations is an example of genetic anticipation (in which mGluR1 and mGluR5) receptor activation [53].
subsequent generations potentially suffer from more Excess activation of mGluR5 receptors has been asso-
significant impairment). On rare occasion, FXS has ciated with a form of protein synthesis-dependent
been associated with FMR1-point mutations instead long-term depression (LTD) [54], a phenomenon
of CGG repeat expansion with subsequent methyla- enhanced in the hippocampus of FMR1 knockout
tion [18]. This is another example of allelic hetero- mice [55]. Activation of mGluR-mediated LTD
geneity in which different mutations in a gene can results in weakening of synaptic connections leading
lead to similar phenotypes [18]. to internalization of AMPA receptors and what
FXS occurs in approximately 1 in 4000–6000 live appear to be structurally immature, elongated den-
births. Among all persons with intellectual disability, dritic processes [56]. Weak, elongated, and immature
between 1.9% [42] and 6.0% [43] are thought to synaptic connections which would be expected with
have FXS. As an X-linked disorder, FXS occurs more enhanced mGluR-mediated LTD have been docu-
frequently in males and the impact of the disorder mented in the FMR1 knockout mouse and in FXS
is more marked in this gender. Due to variable post-mortem brain tissue [52]. These findings have
X inactivation patterns, the full fragile X mutation led to the mGluR theory of FXS which postulates that
in females can be associated with variable phenotypic FMRP normally acts as a brake on mGluR neuro-
expression ranging from the full FXS phenotype to transmission and that excessive mGluR activation
persons with normal cognition with little, if any, clear (specifically mGluR5) leads to the FXS phenotype
clinical impact from the mutation. The clinical including increased risk for seizures, hypersensitivity
phenotype of full mutation FXS is frequently marked to tactile stimuli, intellectual disability, and hyper-
by physical features including a long, narrow face, activity among other characteristics [56].
high arched palate, enlarged ears, joint laxity, and The treatment implications of the mGluR theory
macroorchidism [44]. Behaviorally, persons with FXS of FXS have been evaluated to date in animal models
frequently suffer from attention problems, hyperactiv- and have recently entered human study. Administra-
ity, aggression, self-injury, and anxiety [45]. tion of the mGluR5 antagonist MPEP has been
FXS has significant overlap with ASDs. While reported to reverse phenotypic characteristics associ-
only approximately 2% of persons with autism will ated with the drosophila and mouse models of FXS
have FXS [46], ASDs impact the majority of males [57, 58]. Fenobam, a nonbenzodiazepine anxiolytic
with FXS with about 1 in 3 males meeting criteria with demonstrated mGluR5 antagonist activity [59],
for autism and an additional 1 in 3 meeting dia- is the first mGluR5 antagonist to be studied in
gnosis of Pervasive Developmental Disorder Not humans with FXS [60]. Fenobam was recently studied
Otherwise Specified (PDD-NOS) criteria [47–50]. in six males and six females with full mutation FXS in
The rate of ASDs in females with full mutation a single-dose design. Reportedly, 4 (67%) males and 2
FXS is assumed to be much lower given the overall (33%) females had a response to treatment and the
phenotypic variability of this population. Given the drug was generally well tolerated [60]. Currently, a

367
Chapter 30: Developmental disorders

Table 30.1 Characteristics of selected developmental disorders.

Disorder Genetic Incidence Overlap Physical Medical Characteristic


findings with PDDs appearance comorbidities behavior
Down Trisomy 21 1 in 1000 7–15% meet Flat facies, Leukemia, thyroid Majority exhibit
syndrome live births PDD criteria slanted dysfunction, early strong social
palpebral dementia, cardiac skills for level of
fissures, small anomalies cognitive
ears, single functioning
palmar crease
Rett’s Mutation at 1 in A PDD by Small head Seizure disorder Stereotypic
disorder Xq28 in gene 10 000 definition hand
coding for females movements
MECP2 (classic hand
ringing or
hand washing)
Prader–Willi Paternally 1 in 1 in 4 Obese Hypotonia Obsessive–
syndrome derived gene 10 000– compulsive
deletion within 15 000 behavior (skin
15q11–13 births picking)
Angelman Maternally 1 in 42–81% Frequent Seizure disorder, Frequent
syndrome derived gene 10 000– meet PDD smiling, severe sleep laughter,
deletion within 40 000 criteria microcephaly disturbance hyperactive,
15q11–13 births short attention
span
Smith– Gene deletion 1 in 80–100% Short stature, Severe sleep Self-hugging,
Magenis at 17p11.2 25 000 meet PDD obesity, disturbance, severe self-
syndrome containing RAI1 births criteria prognathism, hyperlipidemia, injury
gene short/teneted renal anomolies,
philtrum hearing loss,
hypotonia, thyroid
dysfunction
Fragile CGG repeat 1 in 4000– 2 in 3 meet Long face, Otitis media, seizure Gaze aversion,
X syndrome expansion in 6000 live PDD criteria large ears, disorder inattention/
FMR1 gene on births large testicles, hyperactivity,
long arm of high arched anxiety,
X chromosome palate aggression/
self-injury
FMR1, fragile X mental retardation gene; MECP, methyl-CpG-binding protein 2; PDD, pervasive developmental disorder; RAI1, retinoic acid
induced 1 gene.

proprietary mGluR5 antagonist from Novartis is feasible. While other developmental disorders such
under study in subjects with FXS in Europe [61]. as autism have a much more complex and heteroge-
FXS stands as an example of a genetic finding neous etiology, lessons can be learned from FXS when
leading to neurobiological understanding with further investigating the genetic underpinnings and resulting
progress linking neurobiology to phenotypic expres- neurobiology of developmental disorders.
sion. This translational approach, resulting in a new The phenotypic overlap between FXS and autism
targeted treatment strategy is a major goal of research gives investigators a solid foundation upon which to
in all neuropsychiatric disorders. The status of FXS as build investigation into autism susceptibility. Future
a single gene disorder has made this progression work may start with thorough genetic analysis of

368
Chapter 30: Developmental disorders

persons with FXS with and without autistic disorder. reason to use findings from the better characterized
All subjects with FXS appear to suffer from a lack of genetic disorders highlighted in this chapter as a
FMRP; the research community has not yet identified potential key to unlock further understanding of
additional factors that may lead to some persons with seemingly more complex idiopathic disorders
FXS expressing classic autism, while others maintain (ASDs).
a modicum of social ability. Increased understanding
of glutamatergic neurotransmission on the basis of Acknowledgements
findings in FXS may also hold promise for future
This work was supported, in part, by research grants
study of the neurobiology and treatment of ASDs.
from the National Institute of Mental Health (NIMH)
including U10 MH066766 (McDougle), R01
Conclusion MH072964 (McDougle); by a FRAXA Research
A review of the genetics of developmental disorders Foundation Grant (McDougle & Erickson); by an
highlights several concepts in psychiatric gene- Indiana Clinical Research Center grant UL1
tics including gene dosage, genetic heterogeneity RR025761 from the National Institutes of Health
(different mutations leading to the same/similar (NIH) to Indiana University; by a Indiana University
phenotypes), imprinting, epigenetics, X inactivation Clinical and Translational Sciences Institute (CTSI)
patterns, and anticipation. Interestingly, among the Career Development Award KL2 (Erickson); and by a
disorders discussed (Table 30.1), while clear genetic Indiana Bureau of Developmental Disabilities
differences exist, there is phenotypic overlap between Services Fellowship in Autism and Developmental
many of the disorders. This overlap gives researchers Disorders (Peay).

References 12. American Psychiatric Association.


Diagnostic and Statistical Manual
23. Milner KM, et al. J Child
Psychol Psychiatry 2005;
1. Feinstein C, et al. Child Adolesc of Mental Disorders, Fourth 46(10):1089–1096.
Psychiatr Clin N Am 2007;16 Edition, Text Revision (DSM-IV-
(3):631–647. 24. Buckley RH, et al. Am J Med Genet
TR). Washington, DC: American 1998;80(4):385–390.
2. Jones KL. Smith’s Recognizable Psychiatric Association; 2000.
Patterns of Human Malformation. 25. Williams CA, et al. Am J Med
13. Robertson L, et al. Am J Med Genet 2001;101(1):59–64.
Philadelphia: W. B. Saunders Genet B Neuropsychiatr
Company; 1988. Genet 2006;141B(2): 26. Wu Y, et al. Proc Natl Acad Sci
3. Schieve LA, et al. Pediatrics 177–183. U S A 2008;105(34):
2009;123(2):e253–260. 12399–12404.
14. Naidu S, et al. J Child Neurol
4. Lowenthal R, et al. J Autism 2003;18(10):662–668. 27. Heck DH, et al. Hum Mol Genet
Dev Disord 2007;37(7): 2008;17(14):2181–2189.
15. Shahbazian MD, et al. Hum Mol
1394–1395. Genet 2002;11(2):115–124. 28. Clayton-Smith J, et al. J Med
5. Kent L, et al. Dev Med Child Genet 2003;40(2):87–95.
16. Jung BP, et al. J Neurobiol 2003;
Neurol 1999;41(3):153–158. 55(1):86–96. 29. Williams CA, et al. Am J Med
6. Cohen D, et al. J Autism Dev 17. Cohen DR, et al. Mol Cell Neurosci Genet 1995;56(2):237–238.
Disord 2005;35(1):103–116. 2003;22(4):417–429. 30. Horsler K, et al. J Intellect Disabil
7. Dykens EM. Ment Retard Dev 18. Venkitaramani DV, et al. Child Res 2006;50(Pt 1):33–53.
Disabil Res Rev 2007; Adolesc Psychiatr Clin N Am 31. Trillingsgaard A, et al. Autism
13(3):272–278. 2007;16(3):541–556. 2004;8(2):163–174.
8. Prandini P, et al. Am J Hum Genet 19. Cassidy SB. J Med Genet 1997; 32. Peters SU, et al. Clin Genet
2007;81(2):252–263. 34(11):917–923. 2004;66(6):530–536.
9. Leonard H, et al. Eur Child 20. Dykens EM, et al. J Child Psychol 33. Girirajan S, et al. Clin Genet
Adolesc Psychiatry 1997;6 S1:8–10. Psychiatry 1996;37(8):995–1002. 2009;75(4):364–374.
10. Amir RE, et al. Nat Genet 1999; 21. Dykens EM, et al. Am J Med Genet 34. Girirajan S, et al. Genet Med
23(2):185–188. 1995;60(6):546–549. 2006;8(7):417–427.
11. Wan M, et al. Am J Hum Genet 22. Veltman MW, et al. Psychiatr 35. Greenberg F, et al. Am J Med
1999;65(6):1520–1529. Genet 2005;15(4):243–254. Genet 1996;62(3):247–254.

369
Chapter 30: Developmental disorders

36. Di Cicco M, et al. Int J Pediatr 45. Bailey DB, Jr., et al. Am J Med 54. Huber KM, et al. J Neurophysiol
Otorhinolaryngol 2001;59(2): Genet A 2008;146A(6):720–729. 2001;86(1):321–325.
147–150. 46. Kielinen M, et al. Autism 2004;8 55. Huber KM, et al. Proc Natl
37. Sweeney E, et al. J Med Genet (1):49–60. Acad Sci U S A 2002;
1999;36(6):501–502. 47. Clifford S, et al. J Autism Dev 99(11):7746–7750.
38. Dykens EM, et al. J Autism Dev Disord 2007;37(4):738–747. 56. Bear MF, et al. Trends Neurosci
Disord 1997;27(2):203–211. 48. Garcia-Nonell C, et al. Am 2004;27(7):370–377.
39. Hicks M, et al. J Dev Behav Pediatr J Med Genet A 2008; 57. McBride SM, et al. Neuron
2008;29(1):42–46. 146A(15):1911–1916. 2005;45(5):753–764.
40. Greenberg F, et al. Am J Hum 49. Bailey DB, Jr., et al. J Autism Dev 58. Yan QJ, et al. Neuropharmacology
Genet 1991;49(6):1207–1218. Disord 2001;31(2):165–174. 2005;49(7):1053–1066.
41. Dykens EM, et al. J Intellect 50. Rogers SJ, et al. J Dev Behav 59. Porter RH, et al. J Pharmacol Exp
Disabil Res 1998;42 (Pt 6): Pediatr 2001;22(6):409–417. Ther 2005;315(2):711–721.
481–489. 51. Jin P, et al. Trends Biochem Sci 60. FRAXA.org. First Fenobam Trial
42. Gerard B, et al. Ann Genet 1997; 2003;28(3):152–158. Results are Positive. 2008; http://
40(3):139–144. 52. Beckel-Mitchener A, et al. Ment www.fraxa.org/, accessed April 6,
43. Florencia G, et al. J Biochem Mol Retard Dev Disabil Res Rev 2012.
Biol 2006;39(6):766–773. 2004;10(1):53–59. 61. ClinicalTrials.gov. http://www.
44. Erickson CA, et al. Curr 53. Weiler IJ, et al. Proc Natl Acad Sci clinicaltrials.gov. Accessed April 6,
Psychiatry 2006;5(10):80–92. U S A 1997;94(10):5395–5400. 2012.

370
Alzheimer’s disease
Chapter

31 Carlos Cruchaga, John S. K. Kauwe, and Alison M. Goate

In 2006 there were an estimated 26 million people which are derived from the β-amyloid precursor pro-
worldwide with Alzheimer’s disease (AD) [1]. The tein (APP) [8]. NFTs are intracellular deposits of
most recent studies of AD predict a rapid increase hyper-phosphorylated tau protein.
in prevalence, projecting that by 2050 there will be a
nearly four-fold increase in AD cases worldwide [1]. Genetic epidemiology
This complex neurodegenerative disorder is charac- Familial aggregation of AD was clearly described in
terized by gradual and progressive memory loss as 1934 [9]. Unfortunately, the late onset of AD can
well as deficits in executive functioning, language, make family studies difficult. In most families direct
visuo-spatial abilities, personality, behavior, and self- examination of parents is impossible. Older siblings
care. Individuals with AD live from 8 to 20 plus years may already be dead, while younger siblings and
after the onset of symptoms. In 2007, the US national children may not have reached the risk period for
costs of caring for individuals with AD were calcu- AD. Despite these difficulties, risk for first-degree
lated to be in excess of 148 billion dollars [2]. Just a relatives has been reported to be 10–40% greater than
five-year delay in the onset of disease could result in unrelated individuals [10–13]. Sibling relative risk
in half as many AD cases in just one generation [3]. ratios are estimated to be between 4 and 5 [11, 14, 15].
The identification of genetic risk factors for AD can Twin studies indicate the presence of a genetic com-
elucidate novel disease-related biological pathways ponent to disease risk, with monozygotic twins show-
and drug targets, making it possible to develop new ing greater concordance (0.49) than dizygotic twins
approaches to prevention and treatment. (0.18) [16, 17].
Standard diagnostic criteria have been developed
by The National Institute of Neurological and Com- Genetic risk factors for familial Alzheimer’s
municative Disorders and Stroke-Alzheimer’s Disease
and Related Disorders Association (NINCDS- disease
ADRDA) [4]. AD can be divided into two categories Mutations in three genes, APP, presenilin 1 (PSEN1),
based on age of onset and familial aggregation, famil- and presenilin 2 (PSEN2), cause FAD by affecting Aβ
ial (FAD), and late onset (LOAD). Cases with evi- levels [18–20] (Figure 31.1). APP is encoded by 18
dence of Mendelian inheritance (autosomal exons spanning 290 kb on chromosome 21 [21]. Each
dominant) and early onset (< 60 years) are categor- of the seven known transcripts encodes a multi-
ized as FAD. Less than 1% of all AD cases fall into this domain protein with one transmembrane domain
category [5–7]. LOAD is characterized by onset after [22]. APP is expressed in all tissues. The major tran-
approximately 60 years of age and complex patterns script in most cell types is longer (APP770) than the
of inheritance. FAD and LOAD share the same clin- transcript in neurons (APP695). APP can be proteo-
ical and pathological features. The pathological lytically processed via one of two pathways. In most
changes include neuronal loss, beta-amyloid (Aβ) cell types, APP is cleaved first by a-secretase, then by
plaques and neurofibrillary tangles (NFTs) g-secretase. In neurons APP is cleaved by β-secretase
(Figure 31.1). Plaques are extra-cellular deposits of and g-secretase resulting in a variety of Aβ species
insoluble proteins composed mainly of Aβ peptides, varying in length from 37–43 amino acids. The most

Principles of Psychiatric Genetics, eds John I. Nurnberger, Jr. and Wade H. Berrettini. Published by Cambridge University
Press. © Cambridge University Press 2012.

371
Chapter 31: Alzheimer’s disease

a-secretase
PSEN1
b-secretase g-secretase
-
PSEN2
APP

Ab
Ab degradation
APOE
SORL1

CALHM1
CALHM1 ACE
TF
Synapses impairment Ab Oligomers
Ab40–42 APOE APOE
Plaques

Neuronal death

TF
Oxidative stress Kinases/
& Inflammation Phosphatases
P
P
P
MAPT
Neurofibrillary P P P PP
P
tangles GAB2 P Tau P Tau
PPPP
Dementia
Fig 31.1 Role of Alzheimer’s disease (AD) -associated genes in likely pathways of neurodegeneration. Genes implicated in risk for AD are
marked with black-border rectangles. The amyloid precursor protein is encoded by APP. APP gives rise to Aβ through serial cleavage by
a-secretase, g-secretase and β-secretase. Mutations in APP, PSEN1 and PSEN2 (g-secretases) are found in early-onset familial AD. APOE, SORL1,
CALHM1 or TF could be implicated in amyloid precursor protein (APP) processing and risk for disease. APOE may also be involved in
amyloid-beta (Aβ) oligomerization and Aβ plaques formation. On the other hand, it is thought that the angiotensin-converting enzyme (ACE)
pathogenic mechanism is related to the Aβ oligomerization and plaques cleavage. According to the amyloid cascade hypothesis,
Aβ pathology would trigger other AD pathogenic events such as tau deposition. Genetics variants in the MAPT gene and GAB2,
favor the formation of NTF, decreasing age at onset or increasing risk for AD.

common fragment is 40 amino acids in length disorder characterized by hemorrhagic strokes as well
(Aβ40); but the major component of amyloid plaques as dementia and pathological evidence of neuritic
observed in AD is the 42 amino acid peptide (Aβ42). plaques and CAA [25]. It appears that elevated levels
The 32 known mutations in APP account for about of total Aβ lead to both AD and CAA pathology,
10% of FAD cases (http://www.molgen.ua.ac.be/ while elevation of Aβ42/Aβ40 ratios leads to AD
ADMutations). Most of these mutations occur at pathology. Mutations in APP may lead to a spectrum
codons near the β and g-secretase cleavage sites in of clinical phenotypes including both dementia and
APP. Mutations near the g-secretase cleavage site hemorrhagic strokes.
result in an increase in the ratio of Aβ42 to Aβ40 After linkage studies provided evidence for a locus
[23, 24]. The “Swedish mutation” (KM670/671N) on chromosome 14, Sherrington et al. identified five
occurs at the β-secretase cleavage site and results in missense mutations in PSEN1 that segregated with
an increase in total Aβ species but does not affect the FAD in their samples [20]. The PSEN1 contains 12
Aβ42/Aβ40 ratio. Individuals with this mutation exons spanning 84 kb on chromosome 14q24.2. The
show pathological evidence of both amyloid plaques full-length protein includes nine transmembrane
and cerebral amyloid angiopathy (CAA). Duplica- domains [26–29]. With some exceptions [30], muta-
tions of APP also lead to disease, causing a familial tions in PSEN1 cause a very early-onset form of the

372
Chapter 31: Alzheimer’s disease

disease, with a mean age at onset (AAO) of 26–60 risk while homozygotes exhibit an eight-fold increase
years [31–33]. To date 182 mutations in PSEN1 have [45]. Nearly all individuals homozygous for the APOE
been identified, accounting for a large proportion of ε4 allele will develop LOAD by 80 years of age [46].
FAD cases (http://www.molgen.ua.ac.be/ADMuta- It is also clear that the APOE ε2 allele has a protec-
tions). These mutations occur throughout the mol- tive effect [45]. APOE ε4 also explains some of the
ecule but all result in an increased Aβ42/Aβ40 ratio variance in AAO in families with a known FAD
[31–33]. mutation [47].
Levy-Lahad et al. identified a mutation in PSEN2 Several hypotheses of the mechanism by which
that segregated with AD in the Volga German kin- APOE affects risk for AD have been proposed. As is
dreds [19]. PSEN2 is homologous to PSEN1 and is the case with the known FAD mutations, it appears
encoded by a gene on 1q42.13. It also has 12 exons but that APOE affects risk for LOAD via an Aβ-related
spans just 25 kb. Presenilins may have other functions mechanism (Figure 31.1). Patients carrying at least
but their role in g-secretase activity is of particular one APOE ε4 allele have more Aβ plaques than non-
relevance to AD. PSEN1 and PSEN2 have distinct but carriers [48]. In vitro, APOE ε4 binds to Aβ with
overlapping g-secretase activities [34] and are thought higher affinity than APOE ε3 [49]. APOE and Aβ
to be the catalytic core of the g-secretase complex, may also compete for clearance through the same
which also includes APH1, NCT, and PEN2 [35, 36]. receptor [50]. Mice over-expressing human APP con-
Fourteen FAD mutations in PSEN2 have been identi- taining an FAD mutation show fibrillar Aβ deposition
fied. These mutations result in a later AAO (40–75 only when the mouse apoE gene is expressed: Aβ does
years) than homologous mutations in PSEN1 and not form amyloid in the absence of apoE [51]. In
may exhibit incomplete penetrance [37]. humans APOE ε4 may also influence fibril formation
and clearance of Aβ, causing increased Aβ deposition
[52]. Indeed, mice expressing a human APOE ε4 allele
Risk factors for LOAD have earlier and more severe pathology than mice
While much is known about genetic risk factors for expressing APOE ε3 [53]. The APOE ε2 allele is pro-
FAD, the vast majority of AD cases are LOAD. The tective for AD and appears to have the opposite effect
single most important known risk factor for LOAD is on amyloid deposition; mouse studies suggest that
age [38–40]. APOE genotype is the strongest genetic although Aβ is deposited it cannot form amyloid in
risk factor for LOAD. The APOE epsilon 4 allele the presence of APOE ε2 [53]. It is also possible that
(APOE ε4) has been genotyped in samples of many additional variation in APOE, such as promoter variants,
racial and ethnic origins, and consistently shows evi- may alter risk for LOAD [54].
dence for association with LOAD [41, 42]. The APOE
gene is located in a region of chromosome 19 that has
been identified as a risk region in genetic linkage Disease mechanisms
studies in LOAD families [43]. It spans less than 4 Our knowledge of the mechanisms by which the
kb and has 4 exons. known genetic risk factors for AD alter risk clearly
APOE exists in three isoforms (APOE ε2, -ε3, and indicates that amyloid metabolism and deposition is
-ε4) which differ from each other at two amino acid central to the pathology of AD. There is some dispute
positions (codon 112 and codon 158). The APOE ε3 as to whether Aβ plaques are themselves neurotoxic
allele (Cys112, Arg158) is the most frequent isoform or whether the toxic species of Aβ may be soluble
in all populations [41]. In most populations the oligomers [55]. This evidence is the basis for the
second most common allele is APOE ε4 (Arg112, substantial body of research investigating mechan-
Arg158) and the third allele is known as APOE ε2 isms of Aβ production, clearance, degradation, and
(Cys112, Cys158). APOE mediates the binding, regulation. Drugs targeting g-secretase and several
internalization, and catabolism of lipoprotein par- aspects of plaque formation and clearance are cur-
ticles, and has been implicated in cardiovascular dis- rently in various stages of clinical trials (for a sum-
ease as well as LOAD (for review see [44]). mary of drugs in clinical trials see www.alzforum.org/
APOE ε4 shows a dose-dependent increase in risk drg/drc/).
for AD. European-Americans who are heterozygous While it does appear that Aβ is central to the
for the APOE ε4 allele exhibit a three-fold increase in pathology of AD, it is clear that tau is also important.

373
Chapter 31: Alzheimer’s disease

The microtubule-associated protein tau, is encoded Hispanic patients [64, 65] and approximately 60% of
by a single gene (MAPT, on 17q21) containing 15 Caucasian AD patients do not carry an APOE
exons. Under normal conditions, tau plays a role in ε4 allele. The discovery of novel genetic risk factors
microtubule stabilization and neuronal integrity [56]. may provide us with unique insight into disease-
Six different isoforms are known to exist in the related biological pathways and thus, novel thera-
human brain and result from alternative splicing of peutic targets.
the MAPT gene. Exon 10 of MAPT can undergo
alternative splicing resulting in a protein with 3 (iso-
form 3R) or 4 (isoform 4R) microtubule-binding
The search for novel risk loci
repeats. The ratio of these 2 isoforms is generally 1 : Linkage and candidate gene studies
1 but appears to be disrupted in specific frontotem- Linkage studies have led to the identification of many
poral dementia (FTD) subtypes involving tau path- potential disease genes that are found in regions of
ology and is detectable upon brain autopsy. For linkage, or positional candidate genes. Studies of the
example, Pick’s disease is characterized by an increase role of Aβ, tau, and even cholesterol processing in AD
in 3R tau, whereas progressive supranuclear palsy have led to the identification of many biological can-
(PSP), corticobasal degeneration (CBD), and most didate genes. Many of these genes have been studied
cases with MAPT mutations share an overabundance using genetic association studies. More than 1000
of 4R tau. In contrast, there is no alteration in the 4R/ candidate gene studies examining hundreds of genes
3R isoform ratio in AD brains [57]. To date, 44 and SNPs have been published. A detailed summary
mutations have been identified in the MAPT gene of these studies and linkage regions can be found on
(http://www.molgen.ua.ac.be/ADMutations/) causing AlzGene (www.alzgene.org [43]). AlzGene is a “com-
FTD. Mutations generally occur within exons 1 and prehensive, unbiased and regularly updated collection
9–13. Missense and intronic splice site mutations are of genetic association studies performed on Alzhei-
most common and involve either disruption of the mer’s disease phenotypes”. This website provides a
normal 4R/3R isoform ratio, altered microtubule meta-analysis of published data for each SNP. Linkage
assembly, or promotion of fibril aggregation [58]. and candidate gene studies have not successfully iden-
However, no mutations in MAPT have been found tified polymorphisms that provide consistent evi-
to cause AD. dence for association with LOAD across multiple
In vitro experiments show that APOE ε3 binds to studies [66].
tau with a greater affinity than APOE ε4, suggesting
that APOE may also have some effect on NFTs [59].
Recent studies of genetic variation in microtubule- Genome-wide association studies
associated protein tau (MAPT) suggest that changes Genome-wide association studies (GWAS) have had a
in tau expression may alter risk and/or AAO of profound effect on the search for genetic risk factors
LOAD [60–62] (Figure 31.1). for human disease. To date over 190 GWAS have
Transgenic mouse studies have also shown that been published. Several studies using very large
Aβ pathology is influenced by levels of tau expres- samples (> 10 000), including exploratory and repli-
sion [63]. Several tau-based therapeutics have been cation datasets, have successfully used GWAS to iden-
studied extensively, including treatments that work tify novel genetic risk factors for complex diseases like
to inhibit tau aggregation or target tau phosphatases diabetes and breast cancer [67–73].
or kinases. Results of Phase II clinical trials for the Several groups have reported the results of a
drug RemberTM, reported at the International Con- GWAS for LOAD [74, 75]. Li et al. carried out a
ference on Alzheimer’s Disease, were very promis- GWAS by genotyping approximately 500 000 SNPs
ing. The active compound of RemberTM is methyl in a series of 753 AD cases and 756 elderly controls
thioninium chloride (MTC), commonly known as from Canada. An additional 418 AD cases and 249
methylene blue. The drug developers suggest that elderly controls from the United Kingdom were used
MTC interferes with tau aggregation and even for follow-up studies. Only SNPs in high linkage
works to clear existing aggregates. There are likely disequilibrium (LD) with APOE ε4 showed associ-
to be genetic effects independent of APOE ε4, as this ation with risk of AD after multiple test correction.
shows only a modest effect on risk in Amish and A SNP (rs10519262) located in an intergenic region

374
Chapter 31: Alzheimer’s disease

on chromosome 15 between ATP8B4 and SLC27A2, was of APP or the enzymes that cleave APP. Finally, several
found to be associated with AAO of AD. Li et al. [76] observations suggest that pathways involving CR1
suggest that rs7176805, located in the ATP8B4 distal products are involved in Alzheimer’s disease pathogen-
promoter region and in strong LD with rs10519262, esis, particularly in Aβ clearance.
could be the functional variant. Rs7176805 potentially It is important to note that CLU, PICALM, and
modifies a CCAAT box transcription factor binding site CR1 participate in other processes not related to Aβ
affecting the expression of ATP8B4, which is involved in fibrillogenesis, processing or clearance, and therefore
phospholipid transport within the cell membrane, with studies of the role of these genes in the brain may
low levels of expression in hippocampus, caudate, sub- reveal evidence for additional disease mechanisms,
stantia nigra, and cerebellum. which go beyond Aβ accumulation. For example,
Coon et al. [74] also genotyped approximately 500 two of the identified AD susceptibility genes (CLU,
000 SNPs in histopathologically verified AD cases (n ¼ CR1) have known functions in the immune system,
664) and elderly controls (n ¼ 442). Only rs4420638, in suggesting a possible role for the immune system in
LD with APOE ε4, was associated with AD after mul- the risk for AD. It is hoped that the new association
tiple test correction. However, in a subsequent study, evidence for these genes leads to a better understand-
rs2373115, located within the gene encoding GRB- ing of the pathological processes implicated in AD.
associated binding protein 2 (GAB2), was associated The proportion of AD cases associated with these
with AD (p ¼ 9  10–11; OR ¼ 4.1; CI ¼ 2.8–14.7) in risk genes has been calculated to be approximately
APOE ε4 carriers [77]. GAB2 is involved in multiple 25.5% for APOE, 8.9% for CLU, 5.8% for PICALM,
signaling pathways, and may be related to tau, Aβ, and 3.8% for CR1. In addition, it is clear that none of
and other aspects of AD pathology and cell survival these genes is pathogenic by itself. Although these are
(Figure 31.1). The association between GAB2 and only crude estimates it is very likely that other genetic
LOAD in APOE ε4 carriers was independently repli- risk factors remain to be identified. The identification of
cated in a Belgian LOAD case-control series [78]. such factors may be possible by combining the results
To obtain results similar to those in other complex from previous studies (meta-analyses) or by the collab-
diseases, GWAS of LOAD will require the use of oration of several research groups and consortiums in
much larger datasets. order to carry out larger studies. It is expected that
For this reason, several groups, including the additional genetic variants will be identified in future
AD Genetics Consortium (AGDC) have combined studies and that these variants will explain only a small
existing GWAS data and/or performed large collab- percentage of the risk for AD. The importance of these
orative GWAS. Two studies published in 2009, 16 studies will not be the identification of specific variants,
years after the identification of APOE as a risk factor with small OR, but the identification of new genes and
for LOAD, found 3 new candidate genes that showed pathways implicated in AD that will enable the identifi-
consistent and compelling evidence for association cation of new therapeutic targets.
with risk for LOAD [79, 80]. These three new genes The most recent technologies used for genotyping
are Clusterin (CLU, also called APOEJ; chromosome 8), the complete genome also allow the quantification of
the gene for the phosphatidylinositol binding clathrin gene copy number. Recent data suggest that copy
assembly protein (PICALM; chromosome 11) and the number variants (CNVs) are surprisingly common;
gene for complement component (3b/4b) receptor 1 it is estimated that between 5 and 12% of the genome
(CR1; chromosome 1). could have CNVs [81]. Association between CNVs
Because pathogenic mutations in APP, PSEN1, and and several diseases have already been identified
PSEN2 directly affect the processing of β-amyloid and [82–86]. Rapid progression of technology for the
isoforms of APOE could also increase risk for AD by detection, quantification and analysis of CNVs has
modifying Aβ aggregation and clearance, researchers made it possible to begin to understand the extent to
have tried to find evidence that links CLU, PICALM, which common CNVs affect risk for human disease.
and CR1 with Aβ. It has been proposed that variants in
CLU could affect risk for AD by promoting the aggre-
gation of Aβ. PICALM, is involved in the intracellular LOAD endophenotypes
trafficking of proteins and lipids, and variants in this The idea of using endophenotypes in the study of
gene could influence Aβ levels by affecting trafficking psychiatric disease was introduced by Gottesman

375
Chapter 31: Alzheimer’s disease

and Shields in 1973 [87]. They defined endopheno- Several studies have also evaluated common variation
types as “internal phenotypes discoverable by a bio- for association with CSF Aβ levels. An assortment of
chemical test or microscopic examination”. An ideal SNPs in SORL1 are associated with risk for AD in
endophenotype should be a risk “factor” (a factor that several association studies [116–121]. Association
increases risk of developing a disease and plays a role between these SNPs and CSF Aβ levels has been
in the development of the disease) and not solely a observed in one small sample but has not been repli-
risk “marker” (a marker of incidental changes associ- cated [113].
ated with disease but not part of the causal pathway) In 2003 Kehoe et al. published data suggesting that
[88–90]. It should also be both inherited and state- SNPs in ACE were associated with CSF Aβ42 [122].
independent [91, 92]. The use of an endophenotype These findings were recently replicated in a large,
for studying the genetics of complex disease may independent sample [112]. The study by Kauwe
provide greater power because it is less heterogeneous et al. went on to identify SNPs in BDNF, DAPK1,
than clinical diagnoses and more directly affected by and TF that show significant association with LOAD
genetic variation. Endophenotypes may also provide a in AlzGene.org and CSF Aβ levels [112]. Rs4878104
biological model of disease and of the possible effects (DAPK1) and rs6265 (BDNF) showed association
of the associated genetic variation. with total Aβ levels while rs1800764 (ACE) and
There is consistent evidence that levels of Aβ42 in rs1049296 (TF) showed association with CSF Aβ42/
cerebrospinal fluid (CSF) correlate with LOAD status Aβ40 levels. Data showing that cell lines overexpres-
and may be a useful endophenotype for LOAD [17, sing a TF cDNA containing the minor allele of
93–103]. Furthermore, Fagan et al. showed that CSF rs1049296 have a significantly higher Aβ42/Aβ40
Aβ42 levels vary inversely with Aβ deposition as ratio than those overexpressing wild-type TF cDNA
measured by positron emission tomography (PET) and are consistent with the CSF associations [112].
using the amyloid imaging tracer Pittsburgh com- TF encodes transferrin, which is the major circulating
pound B (PIB) [104]. Recent animal studies have also glycoprotein involved in iron metabolism and is
illustrated the potential importance of Aβ40 levels: highly expressed in the brain. Iron levels are increased
elevated Aβ40 levels in a transgenic model of Aβ in the brains of AD patients and several reports
deposition substantially delayed Aβ42 deposition in suggest that iron misregulation may influence
these animals [105]. These results and the observa- neurodegeneration by increasing oxidative stress
tions from FAD mutations suggest that risk for AD [123] (Figure 31.1).
may be increased by either increasing Aβ42 levels or It has been proposed that polymorphisms that
decreasing Aβ40 levels. Given that the known genetic modify CSF tau levels may modify the AAO or the
risk factors for AD affect Aβ processing, CSF Aβ severity of disease [62]. Laws et al. found association
levels appear to be useful endophenotypes for genetic between rs242557 (MAPT) and CSF tau levels [124].
studies of AD risk. CSF tau and tau phosphorylated at In 2008, Kauwe et al. found SNPs in MAPT that
residue 181 (phospho-tau) levels are also useful endo- showed association with both CSF tau and phospho-
phenotypes for AD. CSF tau and phospho-tau levels tau181. The minor allele of rs3785883 was associated
are increased in AD patients compared with controls with higher CSF tau and phospho-tau181 levels in AD
and phospho-tau levels correlate with the number of cases and with higher mRNA in AD patients. The
NFTs and the load of hyperphosphorylated tau pre- authors suggest that the increase in tau expression
sent in brain [106]. APOE ε4 shows replicable associ- accelerates the formation of tau deposits and as a
ation with decreased CSF Aβ42 [30, 62, 107–114] as result of this, the minor allele carriers have an earlier
well as tau and phospho-tau levels [30, 62, 107–114]. AAO than the noncarriers [62]. In a candidate gene
Finally, sensitive and specific ELISA tests to measure study analyzing SNPs in tau and in 33 other genes
Aβ42 as well as tau and phospho-tau181 in the CSF related to tau phosphorylation, dephosphorylation,
are readily available [115]. and other tau post-translational modifications, Cru-
Several recent studies illustrate the potential of chaga et al. identified a SNP (rs1868402) associated
this endophenotype-based approach to identify both with higher CSF phospho-tau181 levels. People with
rare and common variants that influence disease. In either one or two copies of the minor allele had higher
2007, Kauwe et al. used CSF Aβ levels to identify an levels of CSF phospho-tau181 than those homozygous
individual with an FAD mutation in PSEN1 [30]. for the common allele of this SNP. Cruchaga et al.

376
Chapter 31: Alzheimer’s disease

followed this up by looking for a connection between molecules include single and multicomponent pro-
this SNP and disease parameters such as risk, age of teins in the cerebrospinal fluid, blood, and urine as
onset, and rate of progression. They found that people well as biomarkers based on gene expression (for
who carried the allele associated with higher CSF review see [132]). Efforts to use proteomics and other
phospho-tau181 levels had a six-fold faster cognitive novel methods to identify novel biomarker proteins in
decline (as measured by change in the clinical demen- various fluids are also underway [133–136]. These
tia rating per year) than those homozygous for the new endophenotypes may be useful in elucidating
common allele. In brains with Aβ pathology, but not and understanding novel pathways that contribute
in normal brains, the harmful SNP was also associ- to risk for AD.
ated with lower levels of protein phosphatase PET imaging of the amyloid-binding agent, Pitts-
B mRNA, and with more neurofibrillary tangles burgh Compound-B (PIB) has also emerged as a
[125]. The authors of this study suggest that genetic possible endophenotype for AD studies. PIB retention
variants associated with CSF Aβ42 levels may have appears to be a good marker of amyloid deposition in
higher influence on risk and AAO (e.g. APOE) but the brain and may be detected before the presence of
variants associated with CSF phospho-tau181 levels clinical symptoms. There is a strong inverse relation
have a greater impact on rate of progression. between in vivo amyloid imaging load and CSF Aβ42
Regular case-control studies are not designed to in humans [104, 109, 137, 138]. As larger samples of
identify genetic variants associated with the rate of individuals with PIB scans are procured these data
progression. For this reason endophenotypes may be may be useful in genetic studies of AD.
useful to identify such variants. Several studies have To date, studies using this endophenotype-based
reported that the ratio of CSF tau/Aβ42 is a useful approach have been biased toward the study of
biomarker to predict progression to Alzheimer’s from existing candidate genes. This can be easily addressed
mild cognitive impairment [126–131] and also rate of by the performance of GWAS in samples with endo-
progression among AD patients [128]. Large-scale phenotype information. The success of these studies,
studies to identify genetic variants associated with even with biased approaches and small sample sizes,
disease progression have not been performed because highlights the potential of this endophenotype-based
it requires longitudinal data collection on a large approach to help us understand the genetic etiology
number of subjects. The use of the ratio of CSF tau/ and pathobiology of LOAD.
Aβ42 as a proxy for disease progression would allow
GWAS in a large sample without the need for longi-
tudinal data. A drawback of this approach is that Gene expression levels
there are few large CSF series available for such stud- New high-throughput technologies for measuring
ies. The genes and pathways implicated in disease gene expression make it possible to interrogate up to
progression may be different from the genes impli- 47 000 transcripts from more than 23 000 genes and
cated in risk for disease; by using disease progression to quantify the different splicing isoforms. Genome-
as a phenotype additional genes and pathways may be wide expression profiles have been combined with
identified. Variants associated with disease progres- complete genome sequence data in order to detect
sion may be useful in predicting more accurately the SNPs that modify the expression of genes (or eQTLs)
time from diagnosis to functional impairment that implicated in AD pathology. This approach has been
may require nursing home placement. Stratification used to show significant association of the H1 haplo-
of samples by such biomarkers will enable cheaper type in MAPT with increased tau expression [61],
and more efficient clinical trials by selecting individ- which could increase the risk for AD [60] or modify
uals expected to have faster rates of progression. By the AAO [62].
targeting different facets of AD biology, this approach The genome-wide expression profiles have also
may identify a broader range of potential therapeutic been used to detect genes that show differential
targets than a conventional case-control design. expression in AD patients and identify potential gen-
While CSF, Aβ, and tau levels are the current etic risk factors for AD. The APOE, BACE1, STUB1
focus of endophenotype-based approaches, several (CHIP), FYN, GGA1, and SORL1 genes, have been
other molecules are being studied as potential bio- identified as differently expressed in AD [139].
markers/endophenotypes for AD. These candidate A recent study used genome-wide expression data to

377
Chapter 31: Alzheimer’s disease

identify AD associated variation in a novel candidate can be also present in AD [148, 149] or in
gene. A variant in CALHM1, which codes for a trans- amyotrophic lateral sclerosis [150]. In FTLD-U, the
membrane glycoprotein that controls cytosolic Ca2þ, presence of TDP-43 deposits are associated with
was identified by screening genes that showed expres- mutation in the GRN gene [151–153]. Most GRN
sion in the hippocampus and are located in AD pathogenic mutations are nonsense, splice-site or
linkage regions. The associated SNP, rs2986017, is a frameshift mutations that generate a loss of GRN
P86L substitution and confers a partial loss of func- function. It has also been reported that GRN expres-
tion which reduces permeability of CALHM1 for sion is lower in sporadic AD cases and that the minor
Ca2þ. The partial loss of function results in lower allele of rs5848, located in the 3’ UTR of GRN, is
levels of cytosolic calcium, increasing Aβ levels by associated with lower GRN gene expression [154]
an APP-dependent mechanism. In the same study and protein levels [155]. These data suggest that vari-
the minor allele frequency of rs2986017 was higher ation in GRN may modify risk for AD. Individuals
in AD cases than in age matched controls in five with low plasma or serum GRN levels can easily be
independent populations [140]. detected using an ELISA assay, allowing early
The ability to measure differences in total expres- identification of at risk asymptomatic subjects [156].
sion, splicing, or transcript ratios between AD cases
and controls will further inform our future studies,
making it possible to identify genetic variants which
Conclusion
have direct effects on expression. Progress in understanding the genetic etiology of AD
has been slow. It appears clear that genetic risk for
LOAD is the result of numerous variants with small
Common themes in neurodegenerative effects as well as one variant with a moderate to large
diseases effect (ApoE4). The task of unraveling this complex
Loci that cause autosomal dominant forms of many genetic architecture is daunting. However, results from
neurodegenerative diseases also carry common gen- recent GWAS and approaches incorporating the use of
etic variants that are associated with increased risk for expression data and quantitative endophenotypes are
the sporadic forms of the same disease [141]. For promising. As scientists come together in the ADGC
example, mutations in PRNP have been shown to and move forward with the application of new tech-
segregate with Creutzfeldt–Jakob disease [142] and nologies to investigate SNPs, structural variation, and
polymorphisms upstream of PRNP exon 1 and also expression at a genome-wide scale, the future is bright.
a missense polymorphism (met/val129) are associated
with the sporadic form of Creutzfeldt–Jakob disease Web-site resources
[143]. In Parkinson’s disease (PD), mutations in a- AD and FTD mutation database:
synuclein and duplications and triplications of a- http://www.molgen.ua.ac.be/ADMutations
synuclein are associated with familial PD [144], and Alzgene database: http://www.alzforum.org/
polymorphisms that alter a-synuclein expression have Database of genotype and phenotype:
been shown to contribute to risk for sporadic forms of http://www.ncbi.nlm.nih.gov/gap
PD [145]. Especially interesting is the case of MAPT. National Center for Biotechnology information:
Tau deposits are found in several neurodegenerative http://www.ncbi.nlm.nih.gov
diseases including AD, FTD, progressive supranuclear Summary of drugs in clinical trials:
palsy (PSP), and corticobasal degeneration (CDB). http://www.alzforum.org/drg/drc/
MAPT missense and splicing mutations can cause
familial FTD, but the H1-MAPT haplotype is associ-
ated with sporadic FTD, PSP, CDB, and even AD [60, Abbreviations
146, 147]. Other proteins such as TARDBP (TDP-43) AAO, age at onset
are also part of pathogenic brain protein deposits Aβ, β-amyloid
found in several neurodegenerative diseases. TDP-43 Aβ40, Aβ of 40 amino acids in length
is found primarily, in frontotemporal lobar deg- Aβ42, Aβ of 42 amino acids in length
eneration with ubiquitin-positive, tau- and alpha- AD, Alzheimer’s disease
synuclein negative inclusion bodies (FTLD-U), but it AD LOAD, late onset AD

378
Chapter 31: Alzheimer’s disease

AGDC, AD Genetics Consortium MAPT, microtubule-associated protein tau


APOE ε4, APOE epsilon 4 allele MTC, methyl thioninium chloride
APP, β-amyloid precursor protein NFTs, neurofibrillary tangles
dbGaP, genotype and phenotype NINCDS-ADRDA, National Institute of Neurological
CAA, cerebral amyloid angiopathy and Communicative Disorders and Stroke–Alzheimer’s
CDB, corticobasal degeneration Disease and Related Disorders Association
CNVs, copy number variants PD, Parkinson’s disease
CSF, cerebrospinal fluid PET, positron emission tomography
FAD, familial AD Phosphor-tau, tau phosphorylated at residue 181
FTD, frontal temporal dementia PIB, Pittsburgh compound B
GAB2, GRB-associated binding protein 2 PSEN1, presenilin 1
GWAS, genome-wide association studies PSEN2, presenilin 2
LD, linkage disequilibrium PSP, progressive supranuclear palsy

References 16. Bergen A. Clinical Genetics


1994;46:144–149.
33. Dermaut B, et al. Am J Hum Genet
1999;64(1):290–292.
1. Brookmeyer R, et al. Alzheimers
Dement 2007;3(3):186–191. 17. Bergen A, et al. Arch Gen 34. Lai MT, et al. J Biol
Psychiatry 1997;54:264–270. Chem 2003;278(25):
2. Alzheimer’s Association. 22475–22481.
Alzheimer’s Disease: Facts and 18. Goate A, et al. Nature 1991;349
Figures. Chicago: Alzheimer’s (6311):704–706. 35. Kimberly WT, et al. Proc Natl
Association; 2007. 19. Levy-Lahad E, et al. Science Acad Sci U S A 2003;
1995;269(5226):973–977. 100(11):6382–6387.
3. Brookmeyer R, et al. Am J Public
Health 1998;88(9):1337–1342. 36. Schroeter EH, et al. Proc Natl
20. Sherrington R, et al. Nature
Acad Sci U S A 2003;100
4. McKhann G, et al. Neurology 1995;375(6534):754–760.
(22):13075–13080.
1984;34(7):939–944. 21. Hattori M, et al. Nature 2000;405
37. Ezquerra M, et al. Arch Neurol
5. Lopera F, et al. JAMA 1997;277 (6784):311–319.
2003;60(8):1149–1151.
(10):793–799. 22. Yoshikai S, et al. Gene 1990;87
38. Jorm AF, et al. Acta Psychiatr
6. Crook R, et al. Nat Med 1998;4 (2):257–263.
Scand 1987;76(5): 465–479.
(4):452–455. 23. Citron M, et al. Nature 1992;360 39. Rocca WA, et al. Ann Neurol
7. Ezquerra M, et al. Neurology (6405):672–674. 1991;30(3):381–390.
1999;52(3):566–570. 24. De Strooper B. et al. J Cell Sci 40. Ritchie K, et al. Lancet 1995;
8. Selkoe DJ. Annu Rev Neurosci 2000;113(Pt 11):1857–1870. 346(8980):931–934.
1994;17:489–517. 25. Rovelet-Lecrux A, et al. Nat Genet 41. Finckh U. J Neural Transm
9. Lowenberg K. Arch Neurol 2006;38(1):24–26. 2003;110(3):253–266.
Psychiatr 1934;31:737. 26. Li X, et al. Neuron 1996;17 42. Bertram L, et al. The AlzGene
10. Mohs RC, et al. Arch Gen (5):1015–1021. Database. Alzheimer Research
Psychiatry 1987;44(5): 27. Li X, et al. Proc Natl Acad Sci Forum. http://www.alzgene.org.
405–408. U S A 1997;94(22):12204–12209. 43. Strittmatter WJ, et al. Proc Natl
11. Breitner JC, et al. Am J Epidemiol 28. Li X, et al. Proc Natl Acad Sci Acad Sci U S A 1993;
1998;128(3):536–548. U S A 1998;95(12):7109–7114. 90(5):1977–1981.
12. Huff FJ, et al. Neurology 1988;38 29. Laudon H, et al. J Biol Chem 44. Stampfer MJ. J Intern Med
(5):786–790. 2005;280(42):35352–35360. 2006;260(3):211–223.
13. Mayeux R, et al. Arch Neurol 30. Kauwe JS, et al. Ann Neurol 45. Bertram L, et al. Nat Genet
1991;48(3):269–273. 2007;61(5):446–453. 2007;39(1):17–23.
14. Sadovnick AD, et al. Genet 31. Xia W, et al. J Biol Chem 1997;272 46. Corder EH, et al. Science 1993;
Epidemiol 1989;6(5):633–643. (12):7977–7982. 261(5123):921–923.
15. Hirst C, et al. Genet Epidemiol 32. Gustafson L, et al. Hum Genet 47. Pastor P, et al. Ann Neurol
1994;11(4):365–374. 1998;102(3):253–257. 2003;54(2):163–169.

379
Chapter 31: Alzheimer’s disease

48. Schmechel DE, et al. Proc Natl 71. Scott LJ, et al. Science 2007;316 95. Galasko D, et al. Arch Neurol
Acad Sci U S A 1993; (5829):1341–1345. 1998;55(7):937–945.
90(20):9649–9653. 72. Skol AD, et al. Genet Epidemiol 96. Shoji M, et al. J Neurol Sci
49. Strittmatter WJ, et al. Proc Natl 2007;31(7):776–788. 1998;158(2):134–140.
Acad Sci U S A 1993;90(17):
73. Zeggini E, et al. Science 2007;316 97. Mehta PD, et al. Arch Neurol
8098–8102.
(5829):1336–1341. 2000;57(1):100–105.
50. Kounnas MZ, et al. Cell 1995;82
74. Coon KD, et al. J Clin Psychiatry 98. Kawarabayashi T, et al. J Neurosci
(2):331–340.
2007;68(4):613–618. 2001;21(2):372–381.
51. Bales KR, et al. Nat Genet 1997;
75. Li H, et al. Arch Neurol 2008; 99. Strozyk D, et al. Neurology
17(3):263–264.
65(1):45–53. 2003;60(4):652–656.
52. Holtzman DM, et al. Ann Neurol
2000;47(6):739–747. 76. Nagase T, et al. DNA Res 2001;8 100. Sunderland T, et al. JAMA
(6):319–327. 2003;289(16):2094–2103.
53. Fryer JD, et al. J Neurosci 2003;23
(21):7889–7896. 77. Reiman EM, et al. Neuron 2007; 101. Hampel H, et al. Mol Psychiatry
54(5):713–720. 2004;9(7):705–710.
54. Wang JC, et al. Neurology 2000;55
(11):1644–1649. 78. Sleegers K, et al. Hum Mutat 102. Jia JP, et al. Neurosci Lett 2005;383
2009;30(2):E338–E344. (1–2):12–16.
55. Cerpa W, et al. Curr Alzheimer
Res 2008;5(3):233–243. 79. Harold D, et al. Nat Genet 2009; 103. Schoonenboom NS, et al. Ann
41(10):1088–1093. Neurol 2005;58(1):139–142.
56. See TM, et al. J Geriatr Psychiatry
Neurol 2010;23(4):260–268. 80. Lambert JC, et al. Nat Genet 104. Fagan AM, et al. Ann Neurol
2009;41(10):1094–1099. 2006;59(3):512–519.
57. Cairns NJ, et al. J Pathol 2004;204
(4):438–449. 81. Redon R, et al. Nature 2006; 105. Kim J, et al. J Neurosci 2007;27
444(7118):444–454. (3):627–633.
58. Gasparini L, et al. Neurogener Dis
2007;4(2–3):236–253. 82. Breunis WB, et al. Blood 2008; 106. Buerger K, et al. Arch Neurol
111(3):1029–1038. 2002;59(8):1267–1272.
59. Strittmatter WJ, et al. Proc Natl
Acad Sci U S A 1994; 83. Cusco I, et al. BMC Med Genet 107. Bouwman FH, et al. Neurobiol
91(23):11183–11186. 2008;9:27. Aging 2007;28(7):1070–1074.
60. Myers AJ, et al. Hum Mol Genet 84. Marshall CR, et al. Am J Hum 108. Brys M, et al. Neurobiol Aging
2005;14(16):2399–2404. Genet 2008;82(2):477–488. 2009;30(5):682–690.
61. Myers AJ, et al. Neurobiol Dis 85. McCarroll SA, et al. Nat Genet 109. Fagan AM, et al. Arch Neurol
2007;25(3):561–570. 2008;40(9):1107–1112. 2007;64(3):343–349.
62. Kauwe JS, et al. Proc Natl Acad Sci 86. Stefansson H, et al. Nature 110. Hansson O, et al. Dement Geriatr
U S A 2008;105:8050–8054. 2008;455(7210):232–236. Cogn Disord 2007;23(5):316–320.
63. Roberson ED, et al. Science 87. Gottesman II et al. Br J Psychiatry 111. Wiltfang J, et al. J Neurochem
2007;316(5825):750–754. 1973;122(566):15–30. 2007;101(4):1053–1059.
64. Pericak-Vance MA, et al. Ann 88. Kannel WB, et al. Ann Intern Med 112. Kauwe JS, et al. Neurogenetics
Neurol 1996;39(6):700–704. 1961;55:33–50. 2009;10(1):13–17.
65. Tang MX, et al. JAMA 1998;279 89. Nadella RK. Lancet 1979;1 113. Kolsch H, et al. Neurosci Lett
(10):751–755. (8130):1354. 2008;440(1):68–71.
66. Bertram L. et al. Hum Mol Genet 90. Hanefeld M, et al. Przegl Lek 114. Smach MA, et al. Neurosci Lett
2004;13(Spec. No. 1):R135–R141. 1989;46(7):588–594. 2008;440(2):145–149.
67. Easton DF, et al. Nature 2007; 91. Gershon ES, et al. Acta Psychiatr 115. Vanderstichele H, et al. Amyloid
447(7148):1087–1093. Scand 1986;74(2):113–118. 2000;7(4):245–258.
68. Hunter DJ, et al. Nat Genet 92. Gottesman II, et al. Am 116. Lee JH, et al. Arch Neurol 2007;
2007;39(7):870–874. J Psychiatry 2003;160(4):636–645. 64(4):501–506.
69. McPherson R, et al. Science 93. Motter R, et al. Ann Neurol 117. Li Y, et al. Neurobiol Dis 2008;
2007;316(5830):1488–1491. 1995;38(4):643–648. 29(2):293–296.
70. Saxena R, et al. Science 2007;316 94. Tamaoka A, et al. J Neurol Sci 118. Liu F, et al. Am J Hum Genet
(5829):1331–1336. 1997;148(1):41–45. 2007;81(1):17–31.

380
Chapter 31: Alzheimer’s disease

119. Rogaeva E, et al. Nat Genet 132. Lonneborg A. Mol Diagn Ther 145. Chiba-Falek O, et al. Hum Genet
2007;39(2):168–177. 2008;12(5):307–320. 2003;113(5):426–431.
120. Bettens K, et al. Hum Mutat 133. Hu Y, et al. Proteomics Clin App 146. Hutton M, et al. Nature 1998;
2008;29(5):769–770. 2007;1:1373–1384. 393(6686):702–705.
121. Lee JH, et al. Neurology 2008; 134. Ray S, et al. Nat Med 2007;13 147. Pittman AM, et al. J Med Genet
70(11):887–889. (11):1359–1362. 2005;42(11):837–846.
122. Kehoe PG, et al. Hum Mol Genet 135. Sonnen JA, et al. Expert Rev 148. Hu WT, et al. Acta Neuropathol
2003;12(8):859–867. Neurother 2007;7(8):1021–1028. 2008;116(2):215–220.
123. Shcherbatykh I. et al. J Alzheimers 136. Craig-Schapiro R, et al. Biol 149. Uryu K, et al. J Neuropathol
Dis 2007;11(2):191–205. Psychiatry 2010;68(10):903–912. Exp Neurol 2008;67(6):
124. Laws SM, et al. Mol Psychiatry 137. Fagan AM, et al. Neurobiol Dis 555–564.
2007;12(5):510–517. 2002;9(3):305–318. 150. Rutherford NJ, et al. PLoS Genet
125. Cruchaga C, et al. PLoS Genet 138. Fagan AM, et al. Ann Neurol 2008;4(9):e1000193.
2010;6(9):e1001101. 2000;48(2):201–210. 151. Baker M, et al. Nature 2006;
126. Kester MI, et al. Neurology 139. Liang WS, et al. Physiol Genomics 442(7105):916–919.
2009;73(17):1353–1358. 2008;33(2):240–256. 152. Cruts M, et al. Nature 2006;
127. Petrie EC, et al. Arch Neurology 140. Dreses-Werringloer U, et al. Cell 442(7105):920–924.
2009;66(5):632–637. 2008;133(7):1149–1161. 153. Mukherjee O, et al. Ann Neurol
128. Snider BJ, et al. Arch Neurol 141. Hardy J. Biochem Soc Trans 2006;60(3):314–322.
2009;66(5):638–645. 2005;33(Pt 4):578–581. 154. Fenoglio C, et al. J Neurology
129. Landau SM, et al. Neurology 142. Owen F, et al. Brain Res Mol Brain 2008;255:47.
2010;75(3):230–238. Res 1990;7(3):273–276. 155. Rademakers R, et al. Hum
130. Wallin AK, et al. Neurology 143. Mead S, et al. Am J Hum Genet Mol Genet 2008;17(23):
2010;74(19):1531–1537. 2001;69(6):1225–1235. 3631–4642.
131. Okonkwo OC, et al. Arch Neurol 144. Zarranz JJ, et al. Ann Neurol 156. Ghidoni R, et al. Neurology
2011;68(1):113–119. 2004;55(2):164–173. 2008;71(16):1235–1239.

381
Index

1000 Genomes Project, 32 major depressive disorder, 215 environmental factors, 279–280
mood disorders, 4–5 future research, 284–285
AAV (adeno-associated virus) post-traumatic stress disorder, 136 genome-wide association
technology, 73 ADORA2A (adenosine studies, 284
ABCA13, and schizophrenia, 243 receptor 2A), 95 heritability, 5–6
ABCB1 (P-glycoprotein gene), ADRA2A, and obesity, 276 neurocognitive phenotypes,
variants, and methadone ADRB3, and obesity, 272–273 353–359
dose, 302 adrenocorticotropic hormone and opioid system, 282
Ab (amyloid beta), and Alzheimer’s (ACTH) risk factors, 279–280
disease, 351–353 and opioid addiction, 300 twin studies, 279–280
absolute risk, 8 roles, 300 see also alcohol dependence
ACE, and Alzheimer’s disease, 376 affected sibling pair tests, 18 alcohol use disorders (AUDs), 279
acetylation, histones, 79–81, 88 affective disorders, and schizophrenia, aldehyde dehydrogenases
ACN9, and alcoholism, 284 233–234 (ALDH), 280
ACTH see adrenocorticotropic Affymertrix GeneChip Mapping ALDH2, variations, 280–281
hormone (ACTH) Arrays, 25 ALDH2*2, and alcoholism, 280–281,
actus reus, 327 aggression, 145 284
AD see Alzheimer’s disease (AD) agoraphobia (AG) ALDH, and alcoholism, 281, 284–285
addictions classification, 112 ALDH (aldehyde
testing issues, 331 diagnosis, 112 dehydrogenases), 280
vulnerability, 303 familial transmission, 113 AliBaba2, 40
see also alcohol dependence; genomic studies, 118 alignments, 41–43
alcoholism; drug addiction; nicotine and panic disorder, 118 categorization, 42
dependence prevalence, 90, 112–113 global, 42
adeno-associated virus (AAV) AGP see Autism Genome local, 42–43
technology, 73 Project (AGP) motif-based, 42
adenosine receptor 2A agranulocytosis, antipsychotic- multiple, 42
(ADORA2A), 95 induced, 59 pairwise, 42
adenosine system, genes AGRE see Autism Genetic Resource sequence, 31, 41–43
involved in, 95 Exchange (AGRE) structural, 43
ADH AKT1, and schizophrenia, 247 alleles, 8–9
and alcoholism, 281, 283–285 alcohol, 280 allelic association, concept of, 231
arrangement, 281 alcohol dehydrogenases (ADH), 280 Allen Brain Atlas, 44
ADH (alcohol dehydrogenases), 280 alcohol dependence AlzGene, 37–38
ADHD see attention-deficit diagnostic criteria, 279 Alzheimer’s disease (AD)
hyperactivity disorder genetics, 295 and amyloid beta, 351–353
(ADHD) linkage plots, 355 candidate genes, 372, 374
admixture mapping, 19 and nicotine dependence, 294 categorization, 371
adoption studies see also alcoholism databases, 37–38
antisocial behavior, 145–146, 156 alcoholism deterministic alleles, 8
gene–environment candidate genes, 280 diagnostic criteria, 371
interactions, 152 and DNA methylation, 298–299 disease mechanisms, 373–374
antisocial personality disorder, endophenotypes, 352 future trends, 378
148, 317 family studies, 282–283 gene expression, 377–378
attention-deficit hyperactivity follow-up linkage studies, 283–284 genetic epidemiology, 371–373
disorder, 149–150, 168 gene–environment interactions, 284 genetics, 371–378
and disease etiology, 2–3 in treatment, 284 genome-wide association studies,
and disorder transmission studies, 3 genetic linkage studies, 282–283 374–375
genetic epidemiology, 2–3 genetics, 279 linkage studies, 374

382
Index

murine studies, 374 diagnosis, 262 studies, 152–156


prevalence, 371 diagnostic criteria, 263 twin studies, 152
risk loci, 374–378 and exercise, 263 gene identification, 153–154
web site resources, 378 family studies, 264–267 genetics, 145–146
see also familial Alzheimer’s genetics, 262–270 externalizing factors, 146–148
disease (FAD); late-onset heritability, 264 and heart rate, 155–156
Alzheimer’s disease (LOAD) linkage studies, 265–266 heritability, 146
American Civil War, post-traumatic malnutrition and, 263–264 interventions, 156–157
stress disorder, 134 and mood disorders, 263 measured risk factors, 154–156
amiodarone, drug interactions, morbidity, 262 molecular genetic studies, 152
53–54 neurocognitive deficits, 263 and psychiatric disorders, 145
amphetamines overview, 262 and substance use disorders, 150
attention-deficit hyperactivity prevalence, 262 symptoms, 148
disorder studies, 176, 179 recovery, 264 treatment, 157
epigenetic mechanisms, 83–85 risk factors, 264 twin studies, 145–148, 156
metabolism, 176 and setshifting, 263 types of, 145
amphetamine-type stimulants (ATS), state–trait characteristics, 263–264 use of term, 145
abuse, 306 symptoms, 262 variance, 147
amyloid beta (Ab), and Alzheimer’s twin studies, 264–265 antisocial personality disorder, 145
disease, 351–353 anterior cingulate, roles, 102 adoption studies, 148, 317
AN see anorexia nervosa (AN) anterior frontal lobes, functions, family studies, 317
Angelman syndrome, 365–366 192–193 heritability, 148
animal models anterior temporal lobes, functions, prevalence, 148
advantages, 72–73 192–193 sex differences, 317–318
attention-deficit hyperactivity anticipation symptoms, 148
disorder, 173, 176–179 and bipolar disorder, 198 twin studies, 317
pharmacogenetics, 179 and schizophrenia, 230–231 anxiety disorders
behavioral assays, 75, 76 anticonvulsants, 59–61 comorbidities, 115–116
depression studies, 85 antidepressants genetic epidemiology, 5
drug-addiction studies, 83–85 adverse reactions, pharmacogenetic genetic variance, 116
epigenetic mechanisms, 88 studies, 56 and Tourette syndrome, 338
gene–environment interactions, drug interactions, 53–54 see also generalized anxiety
178–179 genome-wide association disorder (GAD);
gene expression studies, 49 studies, 225 obsessive–compulsive
genetic susceptibility factor studies, pharmacodynamics, genetic effects disorder (OCD)
72–77 on, 54 APOE E2 (apolipoprotein-E E2) allele,
grooming disorders, 128 pharmacogenetics, 53–57, 225 139–140
hemizygous deletions, 178 pharmacokinetics, genetic effects APOE E4 (apolipoprotein-E E4) allele,
knockdown, 178 on, 53–54 8
knockout, 176–177 see also tricyclic antidepressants APOE
locomotor activity changes, 177 (TCAs) and Alzheimer’s disease, 374
noradrenergic pathway, 177 antipsychotic drugs, 58–59 and late onset Alzheimer’s
obsessive–compulsive antisocial behavior disease, 373
disorder, 128 adoption studies, 145, 156 and major depressive disorder,
post-traumatic stress and attention-deficit hyperactivity 218–219
disorder, 141 disorder, 145, 149 apolipoprotein-E E2 allele, 139–140
psychiatric disorders, 73 common pathways model, 147 apolipoprotein-E E4 allele, 8
schizophrenia, 73, 76, 251–253 comorbidities, 145, 157 APP, and Alzheimer’s disease, 375
selected inbred strains, 178 DSM-IV disorders related to, Applied Biosystems
transgenic, 178 heritability, 148–150 SNPplex assays, 25
animal phobias, sex differences, endophenotypes, 154–156 SOLiD, 28, 29
115 environmental factors, 156 TaqMan assays, 25
ANK3, and bipolar disorder, factors affecting, 156 arginine vasopressin (AVP), 95
205–206 family studies, 145, 153–154 Ariadne Genomics Pathway
ANKK1, and alcoholism, 281–282 future research, 157 Studio, 46
anorexia nervosa (AN) gene–environment interactions, ArrayExpress, 44
candidate gene association studies, 145–157 artificial selection, 324
266–267 adoption studies, 152 ASDs see autism spectrum disorders
future research, 267–270 approaches, 152–153 (ASDs)

383
Index

association studies copy number variations in, 191–192 suicidal, 353


major depressive disorder, 217–218 diagnostic criteria, 183 see also antisocial behavior
psychostimulant dependence, etiology, 183 behavioral assays, animal
308–310 biochemical, 192–193 models, 75
schizophrenia, 238 family studies, 186–187 behavioral disorders, anticonvulsant
assortative mating, bipolar disorder genetics, 183 therapy, 59
and, 198 genetic studies, historical behavioral endophenotypes, 188
atomoxetine, attention-deficit background, 184 behavioral genetics
hyperactivity disorder genetic variants, 191–192 and criminal responsibility,
studies, 179 genome-wide association studies, 327–329
ATP2C2, and language impairment, 190–191 developments, 324
164 genome-wide linkage studies, ethical issues, 324–334
ATP6V1B2, and major depression, 221 186–187 and law, ethical issues, 326–329
ATS (amphetamine-type stimulants), heterogeneity, 187 and medicine, ethical issues,
abuse, 306 Ingenuity network analyses, 192 326–329
attention-deficit hyperactivity disorder linkage studies meta-analyses, 146
(ADHD), 148, 168–179 historical background, 183–188 and punishment, 327–329
adoption studies, 149–150, 168 pathways, 192–193 Bell, Buck v., 325
animal models, 173, 176–179 and Prader–Willi syndrome benzodiazepines, 59
pharmacogenetics, 179 compared, 365 panic disorder treatment,
and antisocial behavior, 145, 149 sex differences, 188, 191 95–96
and autism, 193 twin studies, 183 binary SAM (BAM), 31
and bipolar disorder, 208 Autism Genetic Resource Exchange BIND (Biomolecular Interaction
brain imaging, 176 (AGRE), 186–187 Network Databank), 45
candidate genes, 163, 169–173 endophenotype studies, 188 BIOBASE, 40
catecholaminergic pathways, 172 genome-wide association studies, biochemical bases, for autism,
cholinergic pathways, 172 190–191 192–193
comorbidities, 126–127, 175 linkage studies, 187 BioGPS, 44
as complex phenotype, 173–176 Autism Genome Project (AGP) bioinformatics, 44
dopaminergic pathways, 169–171 genome-wide associations BiologicalNetworks, 46
endophenotypes, neurocognitive, studies, 191 biomarkers, 49–50
350–351 linkage studies, 187 see also genetic markers
family studies, 168 studies, 186 Biomolecular Interaction Network
gene–environment interactions, 173 autism spectrum disorders (ASDs), Databank (BIND), 45
genetic risk factors, 179 183–193 Biomolecular Object Network
genome-wide association studies, 173 endophenotypes in, 188–189 Databank (BOND), 45
genome-wide linkage studies, and fragile X syndrome BioNetSQL, 46
168–169, 171 compared, 367 BioPAX, 45
glutaminergic neurotransmission, genetics, 183 bipolar disorder, 4
172 loci, 186 and anticipation, 198
heritability, 149–150, 168, 169 autonomic dysregulation and assortative mating, 198
heterogeneity, 173–175 model, 101 and attention-deficit hyperactivity
neurocognitive tests, 175 avoidant personality disorder, and disorder, 208
neuroimaging studies, 176 social phobia compared, 117 candidate genes, 202–204
pharmacogenetics, 176 AVP (arginine vasopressin), 95 comorbidities, 207–208
prevalence, 149, 168 AVPR1A, and autism, 190 familial transmission, 207
serotonergic pathways, 172 AVPR1B, and panic disorder, 95 family epidemiology, 196–199
structural variants, 173 azacitidine, 299 family studies, 196
symptoms, 149, 168 future research, 208
and Tourette syndrome, 338 balanced selection, 202 gene mapping methods, 199–200
twin studies, 149–150, 168, 176 BAM (binary SAM), 31 genetic counseling, 208
attributable risk, 8 BDNF, 85–86 genetics, 209
AUDs see alcohol use disorders and cocaine administration, 83–84 genetic susceptibility factors, 69
(AUDs) and histone acetylation, 87–88 genome-wide association studies,
autism, 183–193 and suicidal ideation, 56–57 200, 205–206
and attention-deficit hyperactivity BDNF see brain-derived neurotrophic linkage studies, 200–202
disorder, 193 factor (BDNF) lithium treatment, 59–60
candidate genes, 189–190, 192 behavior and migraine, 208
comorbidities, 192 and environmental factors, 327 pharmacotherapy, 57

384
Index

and schizophrenia, 230–231, heritability, 264 catecholaminergic pathways,


233–234, 208 linkage studies, 265–266 attention-deficit hyperactivity
copy number variant differences, malnutrition and, 263–264 disorder, 172
250–251 and mood disorders, 263 catechol-O-methyltransferase
genetic association differences, overview, 262 (COMT)
250–251 prevalence, 262 and antisocial behavior, 154
subphenotypes, 199 risk factors, 264 and attention-deficit hyperactivity
genetics, 198–199 state–trait characteristics, disorder, 172
mapping genes, 207 263–264 and bipolar disorder, 204
transmission mode, 196–198 symptoms, 262, 263 localization, 300
twin studies, 196 twin studies, 264–265 and opioid addiction, 300–301
blepharospasm, 127 panic disorder studies, 94, 105
BMI see body mass index (BMI) C2ORF3, and reading disability, 163 polymorphisms, 95
BMP7, and major depressive CACNA1C, and bipolar disorder, roles, 94–95, 300
disorder, 225 205–206 CBZ see carbamazepine (CBZ)
BN see bulimia nervosa (BN) CALHM1, and Alzheimer’s disease, CCK (cholecystokinin), 95
body mass index (BMI) 377–378 CDH17, and major depressive
heritability, 272 candidate gene association disorder, 225
variance, 275 studies, 28 cdk5, and drug addiction, 83
BOND (Biomolecular Object Network anorexia nervosa, 266–269 cell biology
Databank), 45 post-traumatic stress disorder, 137, in genetic susceptibility factor
brain 138 studies, 71–72
adolescent, dynamic changes, 75 psychiatric disorders, 49 schizophrenia studies, 251–253
gene expression in, research candidate genes cell cultures, primary, 71
issues, 87 alcoholism, 280 cellular assays, in vitro, 72
human postmortem, gene Alzheimer’s disease, 372, 374 Center for Information Biology
expression profiling, 49 attention-deficit hyperactivity Gene Expression Database
structural changes, and panic disorder, 163, 169–173 (CIBEX), 44
disorder, 103 autism, 189–190, 192 CFG see Convergent Functional
brain abnormalities, and panic bipolar disorder, 202–204 Genomics (CFG)
disorder, 102 cocaine dependence, 309–310 c-fos
brain alterations, gene-induced, 156 language impairment, 164–165 and drug addiction, 83
BrainArray, 36 major depressive disorder, 218–219 and histone phosphorylation, 81
brain-derived neurotrophic factor obesity, 273, 277 Chevron USA, Inc. v. Echazabal, 334
(BDNF), 56–57 obsessive–compulsive disorder, 121, childhood apraxia of speech, 160
genes, 140, 201–202 123–126 children
roles, 85–86, 203–204, 219 personality disorders, 321 conduct disorder, 145
brain dysfunction, and genetic reading disability, 161–163 schizophrenia, 4
susceptibility factors, 77 schizophrenia, 204, 231, 238–248 chlorpromazine, 57
brain function, regulatory animal and cell biology studies, CHNRA5, and stimulant dependence,
mechanisms, 82 251–253 307–308
brain imaging functional polymorphisms and CHNRA7, and schizophrenia, 252
attention-deficit hyperactivity mutations, 251 cholecystokinin (CCK), 95
disorder, 176 stimulant dependence, 312 cholinergic pathways, attention-deficit
humans, genetic susceptibility factor Tourette syndrome, 341 hyperactivity disorder, 172
studies, 77 types of, 217–218 CHRH1, and panic disorder, 95
schizophrenia studies, 234 candidate gene studies, versus CHRM2, and alcoholism, 356–357
brain maturation, postnatal, 75 genome-wide studies, 18 CHRNA3
brain oscillations cannabinoid receptor 1 gene (CNR1) and lung disease, 292
as endophenotypes, 353–359 see CNR1 and nicotine dependence, 290
brain responses, to fear, 103–104 cannabinoid-related genes, and CHRNA4
BRD1, and schizophrenia, 247–248 stimulant dependence, 312 in attention-deficit hyperactivity
BRITE, 45 carbamazepine (CBZ), 59 disorder, 172
Buck v. Bell, 325 adverse reactions, 60–61 and lung disease, 292
bulimia nervosa (BN) applications, 60–61 CHRNA5
diagnosis, 262 Stevens–Johnson syndrome and alcohol dependence, 294
diagnostic criteria, 263 induction, and nicotine dependence, 290–292,
family studies, 264, 265 pharmacogenetics, 61 295
genetics, 262–270 carbon dioxide hypersensitivity, 96 CHRNA7, and schizophrenia, 350

385
Index

CHRNB4, and lung disease, 292 and alcoholism, 282–283 chromosome 19, 19q13, and reading
chromatin and antisocial behavior, 154 disability, 164
epigenetic mechanisms, 79–81, 88 and schizophrenia, 243 chromosome 20, 20p13, and
future research, 88 and Tourette syndrome, 340 autism, 187
remodeling, 80, 87 chromosome 8, and schizophrenia, chromosome 21, and Down
structure, 84, 87 243, 246 syndrome, 363–364
cocaine-induced alterations, 84 chromosome 9, 9p24, and obsessive– chromosome 22
chromosomal anomalies, in compulsive disorder, 122 22q11–13, and autism spectrum
obsessive–compulsive chromosome 10 disorder, 192
disorder, 130, 130 10p12, and diabetes, 275 22q12, and bipolar disorder,
chromosomal haplotypes, 16 10q, and alcoholism, 282–283 206–207
chromosomal regions, in chromosome 11 and schizophrenia, 247–248
obsessive–compulsive 11p14, and autism, 187 chromosome 22q deletion syndrome,
disorder, 123 11p15 and obsessive–compulsive
chromosome 1 bipolar disorder, 200–201 disorder, 127–128
1p36, and major depressive and diabetes, 275 chromosomes, and meiosis, 14–15
disorder, 216 and schizophrenia, 245–246 chronic obstructive pulmonary disease
1p36.3–34.3, and eating chromosome 12 (COPD), genetics, 292–294
disorders, 265 12p13, and major depressive chronic social defeat stress, 85–86
and alcoholism, 282–283 disorder, 221 CIBEX (Center for Information
and schizophrenia, 238–239, 12q24, and diabetes, 275 Biology Gene Expression
350–351 chromosome 13 Database), 44
chromosome 2 13q21, and language cigarette consumption, and lung
2p, and Tourette syndrome, 339 impairment, 164 cancer, 287
2q, and autism spectrum disorders, 13q31, and schizophrenia, 206–207 cigarette smoking
186, 188 and schizophrenia, 246–247 mortality, 287
and alcoholism, 282–283 chromosome 14, and schizophrenia, see also nicotine dependence
and schizophrenia, 239–240 247 cimetidine, drug interactions, 53–54
chromosome 3 chromosome 15 citalopram, studies, 55–56
3p12.3, and reading disability, 163 15q11–13 CJD (Creutzfeldt–Jakob disease), 378
3q27–28, fine mapping, 122 and Angelman syndrome, 365 clinical depression see major
and alcoholism, 282–283 and autism spectrum disorder, depressive disorder (MDD)
and schizophrenia, 240 192, 365–366 CLINT1, and schizophrenia, 240–241
chromosome 4 and obesity, 275 clozapine, 57–59
4p, and alcoholism, 283 15q21, and reading disability, 162 CLSA (Cooperative Linkage Study in
4q 15q25 26, and major depressive Autism), 186–187
and alcoholism, 283 disorder, 216, 217 CLU, and Alzheimer’s disease, 375
arrangements, 281 15q25, and nicotine dependence, ClustalW, 42
chromosome 5 290, 292 Cluster-Buster, 36, 40
5p15, and autism, 187 and Prader–Willi syndrome, 365 CMIP, and language impairment, 164
and schizophrenia, 240–241 chromosome 16 CNPs see copy number
chromosome 6 16p11, and autism spectrum polymorphisms (CNPs)
6p22, and reading disability, disorder, 184, 185, 192 CNR1
162–163 16q23–24, and language and post-traumatic stress
6q27, 187 impairment, 164 disorder, 139
and schizophrenia, 241–243 chromosome 17 and stimulant dependence, 312
chromosome 7 17p11.2, and Smith–Magenis CNTNAP2
7p15.3, and major depressive syndrome, 366 and language impairment, 165, 189,
disorder, 221 17q12, and major depressive 192–193
7q31, and language disorder, 216 and Tourette syndrome, 340–341
impairment, 164 17q CNVs see copy number variations
7q32, and diabetes, 275 and autism, 186 (CNVs)
7q35 and depression, 54–55 cocaine
and language impairment, 165 and Tourette syndrome, 339 epigenetic mechanisms, 83–85
and Tourette syndrome, chromosome 18 and histone phosphorylation, 82
340–341 18p11, and bipolar disorder, 207 cocaine dependence, 306, 314
7q 18q22, and Tourette syndrome, candidate genes, 309–310
and alcoholism, 284 340, 341 genome-wide association studies,
and autism, 186, 187 and bipolar disorder, 201–202 308, 313, 314

386
Index

cognitive deficits, in in autism, 191–192 single nucleotide polymorphisms,


schizophrenia, 234 identification, 26, 175 36–39
cohort effect, 198 in obesity, 273 data mining, 40
color space, 29 schizophrenia, 249–250 DAT (dopamine active transporter)
common environment, use of and bipolar disorder compared, gene, 139
term, 114 250–251 DBH
common genetic variants, 23 in Tourette syndrome, 343 and attention-deficit hyperactivity
autism, 191–192 use of term, 173 disorder, 172
common pathways model, antisocial cortico-striato-thalamocortical and post-traumatic stress disorder,
behavior, 147 (CSTC) pathways, 137–139
common trait rare gene hypothesis alterations, 127 and stimulant dependence, 311–312
(CTRV), 273 corticotropin releasing hormone dbSNP (Single Nucleotide
comorbidities, psychiatric disorders, 7 (CRH), 95 Polymorphism Database), 39
Complete Genomics, next-generation cortisol levels, and suicidal DCDC2, and reading disability,
sequencing, 29 behavior, 353 162–163
COMPOUND database, 45 co-twin control studies, schizophrenia, DCTN5, and bipolar disorder, 205
compulsive hoarding, 122 235–237 DCX, and reading disability, 162
genetics, 130–131 CR1, and Alzheimer’s disease, 375 debrisoquin, pharmacokinetics, 53–54
computer-based analysis and CREB1, and major depressive deletions, schizophrenia, 249–250
resources, for psychiatric disorder, 215–216 Dennet, D. C., 326–327
genetics research, 34–47 Creutzfeldt–Jakob disease (CJD), 378 depression
COMT see catechol-O- CRH (corticotropin releasing animal models, 85
methyltransferase (COMT) hormone), 95 definition, 212
COMT CRHR1, major depressive disorder epigenetic mechanisms, 79–88
murine studies, 300–301 studies, 224 etiology, 212
and personality disorders, 321 criminal responsibility, and behavioral neurotrophic model of, 219
and schizophrenia, 247–248, genetics, 327–329 and nicotine dependence, 294
350–351 criminals, sterilization, 325 pharmacotherapy, 53
and stimulant dependence, 311 CSTC (cortico-striato- symptoms, 85
variants, 128, 300 thalamocortical) pathways, and Tourette syndrome, 338
conditioning, heritability, 118 alterations, 127 treatment, 85
conduct disorder, 148–149 CTNNA2, and bipolar disorder, see also major depressive
children, 145 206–207 disorder (MDD)
conscience, development CTRV (common trait rare gene deterministic alleles, 8
failure, 156 hypothesis), 273 developmental disconnection,
consent issues, genetic information, cultural risk factors, migration studies hypothesis, 193
330–331 and, 3 developmental disorders
controls, selection criteria, 7–8 CYP1A2, 57–58 characteristics, 368
Convergent Functional Genomics CYP2D6, polymorphisms, 57, 276 future research, 369
(CFG), 50 cytochrome p450 system, 53–54 genetics, 363–369
applications, 49 amphetamine metabolism, 176 developmental dyspraxia, 160
approaches, 51 pharmacokinetic studies, 57 DGKH, and bipolar disorder, 206
biomarkers, 50 cytogenetic abnormalities diabetes, 275
development, 49 schizophrenia, 249–250 Diagnostic and Statistical Manual of
and genome-wide association Tourette syndrome, 340–341 Mental Disorders (DSM)
studies, 49 Cytoscape, 45–46 DSM-I, traumatic neurosis,
p-values and, 51 cytosine, methylation, 298 134–135
Cooperative Linkage Study in Autism DSM-III
(CLSA), 186–187 Danio rerio (zebrafish), as animal generalized anxiety disorder, 119
COPD (chronic obstructive model, 73 post-traumatic stress disorder, 134
pulmonary disease), genetics, DAO, and bipolar disorder, 216 social phobia, 112–113
292–294 DAOA, and schizophrenia, 204, trauma exposure, 135
copy number analysis, 26 246–247 DSM-IV
copy number polymorphisms (CNPs) Darrow, Clarence, 328 alcohol abuse disorders, 279
analysis, 26 Darwin, Charles, On the Origin of alcohol dependence, 279
and attention-deficit hyperactivity Species, 325 anorexia nervosa, 263, 266–267
disorder, 173 databases antisocial behavior related
copy number variations (CNVs), 4, in linkage studies, 35 disorders, 148–150
30–31, 375 mutations, 36–37, 39 autism, 183

387
Index

Diagnostic and Statistical Manual of DNA demethylases, existence of, DRD4, 154
Mental Disorders (DSM) (cont.) 82–83 animal models, 178
bulimia nervosa, 263 DNAm see DNA methylation and personality disorders, 321
nicotine dependence, 288 (DNAm) polymorphisms, 171
panic disorder, 90 DNA methods, 23 variants, 175
phobias, 112–113 DNA methylation (DNAm), 79–82, DRD5, polymorphisms, 171
post-traumatic stress disorder, 134 224 Drosophila melanogaster (fruit fly),
trauma exposure, 135 and alcoholism, 298–299 as animal model, 73
DSM-V, 107 catalysis, 82–83 drug addiction
diathesis–stress model, and current research, 224 epigenetic mechanisms, 79–88
post-traumatic stress and gene expression repression, 82 treatment, 83
disorder, 134 and opioid addiction, 298–299 see also cocaine dependence, heroin
dietary factors, drug metabolism, 53–54 roles, 87 addiction; opioid addiction;
DISC1, 73 DNA samples, whole-blood, 23 psychostimulant dependence;
and schizophrenia, 204, 238–239, DNA sequence variations, 21, 36 stimulant dependence
251 analysis tools, 37 drug-addiction studies, animal
structure, 70–71 future research, 46–47 models, 83–85
studies, 69, 72 DNA sequencing DRUG database, 45
animal models, 76–77 analysis, 35 drug interactions, antidepressants,
murine, 76 focused, 35 53–54
disease etiology, adoption studies and genetic studies, 20–21 drug metabolism, dietary factors,
and, 2–3 tools and resources, 36 53–54
disease mapping, single nucleotide dopamine active transporter (DAT) drug relapse, mechanisms, 83
polymorphisms, 23–24 gene, 139 drug responses, environmental factors,
diseases dopamine beta-hydroxylase (DBH) 61–65
complex, 27 gene see DBH DSM see Diagnostic and Statistical
genetic variants, 35 dopamine receptor gene, 137–139 Manual of Mental Disorders
genome-wide association studies, post-traumatic stress disorder (DSM)
34–35 studies, 140 DTNBP1, and schizophrenia, 241–243
disorder transmission studies, dopamine-related genes, and duplications, schizophrenia, 249–250
adoption studies and, 3 stimulant dependence, Durham v. US, 327–328
disrupted in schizophrenia 1 (DISC1) 310–312 dynamic programming algorithms, 42
see DISC1 dopaminergic pathways, attention- dyslexia see specific reading disability
dizygotic (DZ) twins, 2, 14 deficit hyperactivity (SRD)
antisocial behavior studies, 145 disorder, 169–171 dystonia myoclonus syndrome (DMS),
antisocial personality disorder dopaminergic system, 58 and obsessive–compulsive
studies, 317 dopamine system disorder, 127
autism studies, 183 genes DYX loci, 161
bipolar disorder studies, 196 in obsessive–compulsive DYX1, 162
carbon dioxide hypersensitivity disorder, 126 DYX1C1, and reading disability, 162
studies, 96 in post-traumatic stress disorder, DYX2, 162–163, 165–166
extraversion studies, 117 137–139 DYX5, 163
major depressive disorder, genes involved in, 94–95 DYX8, 165–166
214–215 dopamine transporter gene (SLC6A3)
neuroticism studies, 117 see SLC6A3 EAAC1 see SLC1A1
obsessive–compulsive disorder Down syndrome, 363–364 eating disorders (EDs)
studies, 121 DPYSL2, and schizophrenia, 243 linkage studies, 265–266
panic disorder studies, 92 DRD1, polymorphisms, 171 morbidity, 262
personality disorder studies, 318 DRD2 prevalence, 262
post-traumatic stress disorder alleles, 154 symptoms, 263
studies, 136 and alcoholism, 281–282 see also anorexia nervosa (AN);
schizophrenia studies, 235 animal models, 178 bulimia nervosa (BN); obesity
co-twin controls, 235–237 polymorphisms, 58, 171, 266–267 EBI see European Bioinformatics
stimulant dependence studies, and schizophrenia, 245–246 Institute (EBI)
307–308 and stimulant dependence, 310–311 EBI Dali, 43
Tourette syndrome studies, 337 DRD3 Echazabal, Chevron U.S.A.,
DMS (dystonia myoclonus syndrome), polymorphisms, 58 Inc. v., 334
and obsessive–compulsive and schizophrenia, 240 egr-1, and cocaine withdrawal,
disorder, 127 and stimulant dependence, 311 83–84

388
Index

electroencephalography (EEG) EPHB1, and major depressive autism studies, 188–189


interhemispheric theta coherence, disorder, 225 resources, 44
356–357 EPIC (Events Preceding Interstitial externalizing disorders
resting, beta power, 354–356 Cystitis), 99–100 heritability, 146–148
Elston–Stewart algorithm, 26–27 epidemiology twin studies, 116–117
EMBOSS, 42–43 phobias, 112–113 use of term, 145
E-MSD (Macromolecular Structure see also genetic epidemiology extraversion, 117
Database), 43 epigenetic mechanisms, 79–83 eye-tracking dysfunction, and
EN2, and autism, 189–190 animal models, 88 schizophrenia, 350
ENCODE Project, 20, 32, 47 in depression, 79–88
regions, 40–41 in drug addiction, 79–88 FAAH, and stimulant dependence, 312
endophenotypes future research, 88 FACS (fluorescence activated cell
alcoholism, 352 and gene expression, 79 sorting), 87
for antisocial behavior, 154–156 epigenetic modifications, 51 factor analysis, personality
attention-deficit hyperactivity obesity, 275–276 disorders, 320
disorder, 350–351 epigenetics Fagerström Test of Nicotine
in autism spectrum disorders, definition, 79 Dependence (FTND), 288
188–189 field of study, 79 familial aggregation, 3
behavioral, 188 future research, 88 indicators, 1–2
brain oscillations as, 353–359 major depressive disorder, familial Alzheimer’s disease (FAD),
concept of, 348 223–224 376
conceptual development, 347–348 future research, 226 diagnostic criteria, 371
definition, 187–188 mu-opioid receptor, 298–299 genetic susceptibility factors, 371
family studies, 349 personality, 322 familial recurrence, schizophrenia,
heritability, 348–349 post-traumatic stress disorder, 232–233
late onset Alzheimer’s disease, 140–141 familial specificity, phobias, 114
375–377 research issues, in psychiatry, 87 familial transmission
neurochemical metabolites as, schizophrenia, 231–232 bipolar disorder, 207
351–353 Tourette syndrome, 343–344 phobias, 113
neuroimaging, 351 EPSIN4, and schizophrenia, 240–241 schizophrenia, 207
obsessive–compulsive disorder, 188 eQTL see expression quantitative trait family epidemiology, bipolar disorder,
for psychiatric disorder locus (eQTL) 196–199
classification, 9 EROs (even-related oscillations), theta family history screen (FHS), 96–97
psychiatric disorders and, 349 and delta, 357–359 Family Interview for Genetic Studies
in psychiatric genetics, 347–360 escitalopram, studies, 55–56 (FIGS), 96–97
applications, 349–359 ethanol, consumption, 280 family studies
future trends, 359–360 ethical issues advantages, 2
identification criteria, 348–349 behavioral genetics, 324–334 alcoholism, 282–283
schizophrenia, 350–351 eugenics, 324–326 anorexia nervosa, 264
state independence, 349 sterilization, 325 antisocial behavior, 145, 153–154
variations, 349 N-ethyl-N-nitrosourea (ENU) antisocial personality disorder, 317
ENU (N-ethyl-N-nitrosourea) mutants, 178 attention-deficit hyperactivity
mutants, 178 eugenics, 324–326 disorder, 168
environmental factors European Bioinformatics autism, 186–187
alcoholism, 279–280 Institute (EBI) bipolar disorder, 196
antisocial behavior, 156 GeneWise, 35 bulimia nervosa, 264
and behavior, 327 InterProScan, 43 early, 34
drug responses, 61–65 Macromolecular Structure endophenotypes, 349
identification, 9 Database, 43 generalized anxiety disorder, 119
in obsessive–compulsive even-related oscillations (EROs), theta genetic epidemiology, 1–2
disorder, 130 and delta, 357–359 major depressive disorder,
Tourette syndrome, 342 Events Preceding Interstitial Cystitis 213–214
see also gene– (EPIC), 99–100 obsessive–compulsive disorder,
environment interactions exercise, and anorexia nervosa, 263 121, 123
(G  E) expanded spectrum approach, panic disorder, 90–92
environmental risk factors, migration 97–102 post-traumatic stress disorder,
studies and, 3 ExPasy Translate, 35 135–136
environmental variance, phobias, twin expression quantitative trait psychiatric disorders, 69
studies, 114 locus (eQTL) sampling, 7

389
Index

family studies (cont.) fruit flies, 73 animal models, 178–179


schizophrenia FTND (Fagerström Test of Nicotine antisocial behavior, 145–157
cognitive deficits and imaging Dependence), 288 adoption studies, 152
changes, 234 functional analysis, computer- approaches, 152–153
segregation analysis, 237–238 based, 35 studies, 152–156
subtypes, 232–233 functional assessment, genes, 38 twin studies, 152
variance component analysis, functional imaging studies attention-deficit hyperactivity
237–238 limitations, 103 disorder, 173
single nucleotide polymorphism panic disorder, 102–104, 107 definition, 151–152
genotyping, 26–27 functional magnetic resonance major depressive disorder, 224
social phobia traits, 117–118 imaging (fMRI), 351 obesity, 272–273
stimulant dependence, 306–307 functional polymorphisms, and post-traumatic stress
substance use disorders, 5, 6 schizophrenia, 251 disorder, 142
Tourette syndrome, 337 functional validation, of genetic substance use disorders, 150
FASTA, 42 susceptibility factors, for gene expression, 44
fastq, 31 psychiatric disorders, 69–77 Alzheimer’s disease, 377–378
fear FXS see fragile X syndrome (FXS) in brain, research issues, 87
brain responses to, 103–104 FXYD6, and schizophrenia, 245–246 databases, 44
conditioning, susceptibility to, 118 FZD3, and schizophrenia, 243 drug-induced, 83
irrational, 113 epigenetic mechanisms and, 79
and post-traumatic stress disorder, GABA-ergic system, genes involved in, major depressive disorder,
103–104 95–96 222–223
prevalence, 113 GABA system see gamma- future research, 226
processing, in panic disorder, 107 aminobutyric acid (GABA) repression, 82
stimuli, masked versus unmasked, system Gene Expression Omnibus
104 GABHS (group A beta hemolytic (GEO), 44
unconditioned, 102–103 streptococci), 342 gene expression profiling, 50
fearful faces, amygdala GABRA1, and schizophrenia, 240–241 gene expression studies
activation, 106 GABRA2, and alcoholism, 282–283, animal models, 49
FHS (family history screen), 96–97 354–357 developments, 51
FIGS (Family Interview for Genetic GABRA4, and alcoholism, 354–356 future trends, 51
Studies), 96–97 GABRA6, and schizophrenia, 240–241 human postmortem brain, 49
FKBP5 GABRB1, and alcoholism, 354–356 large-scale, viability, 49
and major depressive GABRB2, and schizophrenia, 240–241 in psychiatric disorders, 49–51
disorder, 219 GABRG1, and alcoholism, 283, p-values in, 50–51
and post-traumatic stress disorder, 354–356 GeneGo MetaCore, 46
139–140 GABRP, and schizophrenia, 240–241 GeneHunter program, 26–27
fluorescence activated cell sorting GAD2, and obesity, 275 gene identification
(FACS), 87 GAD see generalized anxiety disorder antisocial behavior, 153–154
fluoxetine, 86, 179, 224 (GAD) future trends, 49
pharmacokinetics, 53–54 GAD (glutamic acid decarboxylase), genetic epidemiology and, 7–10
FMR1, and fragile X syndrome, 366–367 95–96 gene localizations, in reading
fMRI (functional magnetic resonance Galton, Sir Francis, 324–325 disability, 161
imaging), 351 gamma-aminobutyric acid (GABA) gene mapping, major depressive
focal dystonias, 127 system disorder, 215–222
FosB, and drug addiction, 83 panic disorder studies, 95–96 gene mapping methods
FOXP2, and language receptor genes, and alcoholism, 283 bipolar disorder, 199–200
impairment, 164 GENDEP (Genome-Based subphenotypes, 207
fragile X mental retardation protein Therapeutic Drugs for Gene Ontology (GO) database, 45
(FMRP), 366–367 Depression), 55–56 generalized anxiety disorder (GAD),
fragile X syndrome (FXS) gender differences see sex differences 119–120
and autism spectrum disorders gene association studies, hypothesis- comorbidities, 117
compared, 367 driven, opioid addiction, 302 genetic epidemiology, 5
etiology, 366–367 GENE database, 45 genetics, 112–120
future research, 368–369 gene dosage, disorders, 363–364 generalized social phobia, 117
genetics, 366–369 gene–environment correlations, 6 genes
and mGluR theory, 367–368 gene–environment interactions additive
prevalence, 367 (G  E), 6, 151–152 generalized anxiety disorder, 120
sex differences, 367 alcoholism, 284 and phobias, 114–115, 117

390
Index

in adenosine system, 95 bipolar disorder, 200 epigenetics; pharmacogenetics;


in dopamine system, 94–95 schizophrenia, 238 psychiatric genetics
functional assessment, 38 genetic markers, 34–35 genetic studies
in GABA-ergic system, 95–96 see also biomarkers autism, historical background, 184
in G-protein signaling regulators, 96 genetic risk factors see genetic DNA sequencing and, 20–21
in neuropeptide systems, 95 susceptibility factors limitations, 61–65
in obsessive–compulsive disorder, genetics panic disorder, 90
121, 123–126 alcohol dependence, 295 issues, 107
rare, in obsessive–compulsive alcoholism, 279 limitations, 106
disorder, 130, 130 alcohol use disorders, 279 post-traumatic stress disorder,
related to carbon dioxide Alzheimer’s disease, 371–378 135–140
hypersensitivity, 96 Angelman syndrome, 365–366 see also molecular genetic studies
in serotonin system, 93–94 anorexia nervosa, 262–270 genetic susceptibility factors
see also candidate genes antisocial behavior, 145 animal models, 72–77
GENESEQ, 45 autism, 183 assays, 74–75
gene set expression analysis autism spectrum disorders, 183 DISC1 studies, 76–77
(GSEA), 44 bipolar disorder, 209 readouts relevant to psychiatric
genes-to-brain-to-antisocial-behavior subphenotypes, 198–199 disorders, 74–75
model, 156 bulimia nervosa, 262–270 biochemical studies, 69
genetic analysis, issues, 19–20 chronic obstructive pulmonary bipolar disorder, 69
genetic association disease, 292–294 and brain dysfunction, 77
bipolar disorder, 200 compulsive hoarding, 130–131 cell biology studies, 71–72
schizophrenia and bipolar disorder developmental disorders, 363–369 during neurodevelopment, 75
compared, 250–251 Down syndrome, 363–364 familial Alzheimer’s disease, 371
genetic bases, of psychiatric disorders, fragile X syndrome, 366–369 for-profit testing, 331–332
methods, 23 generalized anxiety disorder, human studies, with brain
genetic case-control association 112–120 imaging, 77
studies, post-traumatic stress grooming disorders, 128–130 late onset Alzheimer’s disease, 373
disorder, 137–140 heroin addiction, 301, 302–303 pathways, 70–71
genetic counseling, bipolar disorder, 208 historical background, 324 protein chemistry studies, 69–71
genetic engineering, murine, 73 lung cancer, 292–295 for psychiatric disorders, functional
genetic epidemiology, 1–3 lung disease, 292–295 validation, 69–77
adoption studies, 2–3 major depressive disorder, 212–227 schizophrenia, 4, 69
Alzheimer’s disease, 371–373 Mendelian, 324 genetic variance
applications, to gene identification, nicotine dependence, 288 anxiety disorders, 116
7–10 and psychiatric disorders, phobias, twin studies, 114
early studies, 1 294–295 genetic variants
family studies, 1–2 obesity, 272–277 autism, issues, 191
field of study, 1 obsessive–compulsive disorder, diseases, 35
migration studies, 3 121–131 functional assessment, 38
panic disorder, 90 future research, 130–131 germ-line, 23
phobias, 113–118 opioid addiction, 297–303 identification, 61–65
future research, 119 origin of term, 347 linkage and linkage disequilibrium,
of psychiatric disorders, 1–10 personality disorders, 316–322 14–17
risk estimation, 8 phobias, 112–120 multiple comparisons, 19–20
schizophrenia, 3–4 Prader–Willi syndrome, 365–366 types of, 23
twin studies, 2 psychostimulant dependence, see also common genetic variants;
genetic factors, interactions, 20 308–310 copy number polymorphisms
genetic heterogeneity, 6, 20 reading disability, 161 (CNPs); copy number
genetic imprinting, 365–366 and responsibility, 326–329 variations (CNVs); rare genetic
genetic information Rett’s disorder, 364–365 variants
applications, social, 324–326 schizophrenia, 230–253 gene transfer, in utero, 73–74, 76–77
consent issues, 330–331 Smith–Magenis syndrome, 366 GeneWise, 35
control of, 329 speech sound disorder, 165–166 GenMAPP, 46
forensic, privacy issues, 329–330 stimulant dependence, 306–314 Genome-Based Therapeutic Drugs for
health professionals and, 332–333 Tourette syndrome, 126–127, Depression (GENDEP), 55–56
privacy of, 329 336–344 genome scan meta-analysis
genetic linkage weight loss, 276 (GSMA), 202
alcoholism, 282–283 see also behavioral genetics; genome sequencing, complete, 35

391
Index

genome-wide association studies glucocorticoid receptor (GR), hemizygous deletions, animal


(GWAS), 19–20, 55–56 polymorphisms, 139 models, 178
alcoholism, 284 glutamate system genes, 123 HEP (Human Epigenome Project), 47
Alzheimer’s disease, 374–375 glutamic acid decarboxylase (GAD), heritability
antidepressants, 225 95–96 alcoholism, 5–6
applications, 27 glutaminergic neurotransmission, in anorexia nervosa, 264
disease studies, 34–35 attention-deficit hyperactivity antisocial behavior, 146
attention-deficit hyperactivity disorder, 172 antisocial personality disorder,
disorder, 173 P-glycoprotein gene (ABCB1), 148
autism, 190–191 variants, and methadone attention-deficit hyperactivity
bipolar disorder, 200, 205–206 dose, 302 disorder, 149–150, 168–169
case-control, 27 GNB3, and major depressive disorder, body mass index, 272
cocaine dependence, 308, 314 218–219 bulimia nervosa, 264
and Convergent Functional GO (Gene Ontology) database, 45 conditioning, 118
Genomics, 49 GPR158, and obesity, 275 conduct disorder, 148–149
development, 24–25 GPRIN2, and obesity, 273 of DSM-IV disorders related to
errors, sources of, 28 G-protein signaling regulators, genes antisocial behavior, 148–150
late onset Alzheimer’s disease, involved in, 96 endophenotypes, 348–349
374–375 GR (glucocorticoid receptor), establishment, 13
limitations, 35, 49 polymorphisms, 139 estimates, 3
major depressive disorder, 219–222 GRIN1, and weight loss, 276 externalizing disorders, 146–148
nicotine dependence, 289–292, GRK3, and bipolar disorder, 204 migration studies, 14
295 GRM3, and schizophrenia, 243, 253 missing, 274–275
obesity, 274 GRM7 mood disorders, 196
obsessive–compulsive disorder, and attention-deficit hyperactivity nicotine dependence, 289
122–123, 130 disorder, 176 obesity, 272
opioid addiction, 302–303 and major depressive disorder, oppositional defiant disorder, 149
panic disorder, 93, 106 221–222 panic disorder, 92
permutational analysis, 28 GRM8, and alcoholism, 357–359 personality disorders, 318–321
personality disorders, 321–322 GRN, and Alzheimer’s disease, 378 personality traits, 326
post-traumatic stress disorder, grooming disorders, 128–130 phobias, sex differences, 114
141–142 group A beta hemolytic streptococci psychopathic traits, 150–151
single nucleotide polymorphism (GABHS), 342 schizophrenia, 230–231
genotyping, 27–28 GSEA (gene set expression analysis), 44 smoking initiation, 288
stimulant dependence, 312–313 GSMA (genome scan meta- stimulant dependence, 306–308
Tourette syndrome, 342–343 analysis), 202 substance use disorders, 150
use of term, 23 GWAS see genome-wide association Tourette syndrome, 337
genome-wide linkage studies studies (GWAS) traits, 113
attention-deficit hyperactivity G  E see gene–environment twin studies, 14
disorder, 168–169, 171 interactions (G  E) heritability estimates, psychiatric
autism, 186–187 disorders, 3
genome-wide studies, versus candidate haloperidol, 57, 224, 252 heroin addiction, genetics, 301–303
gene studies, 18 haplotype blocks, 39 heterogeneity
genomic imprinting disorders, 365 haplotypes, chromosomal, 16 autism, 187
genomics HapMap Project see International phenotypic, 20
future research, 47 HapMap Project see also genetic heterogeneity
impacts, on psychiatric science and Haseman–Elston linkage algorithm, HGMD (Human Gene Mutation
practice, 9–10 118–119 Database), 37–38
see also Convergent Functional HATs (histone acetyltransferases), 81 HGVBase, 37–38
Genomics (CFG) HD see Huntington’s disease (HD) high-density single nucleotide
genomic studies, phobias, 118–119 HDACs see histone deacetylases polymorphism (SNP)
genotype likelihood format (GLF), 31 (HDACs) genotyping
genotyping HDC, and Tourette syndrome, arrays, 23
concept of, 347 336–339 technologies, 24–25
future research, 46–47 HDMs (histone demethylases), 82 high-throughput assays, 44
GEO (Gene Expression Omnibus), 44 health professionals, and genetic L-histidine decarboxylase (HDC)
Gibbs sampling, 42 information, 332–333 see HDC
GLF (genotype likelihood format), 31 heart rate, and antisocial behavior, histone acetyltransferases (HATs),
global alignments, 42 155–156 81

392
Index

histone deacetylases (HDACs), 81 sequencing, 347 Ingenuity network analyses,


cocaine and, 84 variant identification, 13 autism, 192
inhibitors, 86 strategies, 34 Ingenuity Pathway Analysis, 46
histone demethylases (HDMs), 82 Human Genome Project, 324 inheritance
histone methyltransferases (HMTs), human studies, genetic susceptibility general personality dimensions,
82 factors, with brain imaging, 77 320–321
inhibition, 85 huntingtin (htt), and Huntington’s Mendelian laws of, 1–2
histones disease, 69 modes of, 6
acetylation, 79–81, 88 Huntington’s disease (HD), personality disorders, 318–320
and bdnf, 87–88 etiology, 69 interhemispheric theta coherence,
genome-wide studies, 81 hybridization-based target 356–357
and epigenetic mechanisms, 79–81 enrichment, 32 interleukin receptors, encoding,
methylation, 79–82, 88 6-hydroxydopamine (6-OHDA), 177 56–57
localization, 82 5-hydroxytryptamine-1B receptor internalizing disorders, twin studies,
regulatory mechanisms, 82 (5-HT1B), and opioid 116–117
phosphorylation, 79–82, 85 addiction, 301 International Classification of Diseases
localization, 81 hydroxytryptamine (5-HT), 154 (ICD-10), alcohol abuse
regulatory mechanisms, 82 hypercapnia, 96 disorders, 279
HLA, and schizophrenia, 241–243 hypothalamic-pituitary-adrenal International HapMap Project, 18,
HLA-B*1502 allele test, 65 (HPA) axis, 224 24–25
HMTs see histone methyltransferases and major depressive disorder, 219 databases, 39
(HMTs) International Molecular Genetic Study
hoarding see compulsive hoarding IC see interstitial cystitis (IC) of Autism Consortium
Holmes, Oliver Wendell, 325 ICD-10 (International Classification of (IMGSAC), 183–186
HPA axis see hypothalamic-pituitary- Diseases), alcohol abuse two-stage genome scan, 186
adrenal (HPA) axis disorders, 279 International Schizophrenia
5-HT1A receptor, 93–94 idiopathic generalized epilepsy Consortium, 46–47
5-HT1B (5-hydroxytriptamine-1B (IGE), and schizophrenia, InterProScan, 43
receptor), and opioid 249–250 interstitial cystitis (IC), 99
addiction, 301 IGE (idiopathic generalized epilepsy), case-control studies, 98, 99–100
5-HT2A receptors and schizophrenia, 249–250 in utero gene transfer, 73–74, 76–77
and major depressive disorder, 55 IGF2, and obesity, 276 inversions, 30–31
and obesity, 276 IL15, and obesity, 273 in vitro cellular assays, 72
polymorphisms, 58, 94 IL28RA, and suicidal ideation, iPS cells see induced pluripotent stem
5-HT2C receptors, 58–59 56–57 (iPS) cells
5-HT (hydroxytryptamine), 154 Illumina irrational fears, and phobias, 113–115
HTR1B, 301 BeadArray Mapping arrays, 25
HTR1D, and eating disorders, 265 Genome Analyzer, 28, 29 JAligner (software), 42
HTR2A HiSeq, 29
anorexia nervosa studies, 267–269 procedures, 25 kappa-opioid receptor (KOPr), 299
and attention-deficit hyperactivity imaging changes, in KE family, 164
disorder, 172 schizophrenia, 234 KIAA0319, and reading disability,
and major depressive disorder, imaging genetics, 104–106 162–163, 166
55–56 IMGSAC see International Molecular knockdown animal models, 178
htt (huntingtin), and Huntington’s Genetic Study of Autism knock-in mice, 73
disease, 69 Consortium (IMGSAC) knockout animal models, 176–177
5-HTTLPR see serotonin-transporter- imipramine, 224–225 knockout mice, 73, 176–177
linked polymorphic region pharmacokinetics, 53–54 KOPr see kappa-opioid receptor
(5-HTTLPR) IMMP2L, and Tourette (KOPr)
Human Epigenome Project (HEP), 47 syndrome, 340 Kyoto Encyclopedia of Genes and
Human Gene Mutation Database inbred strains, selected, 178 Genomes (KEGG), 45–46
(HGMD), 37–38 incomplete penetrance, 6
human genetics research induced pluripotent stem (iPS) lactate dehydrogenase (LDH), 96
analysis strategies, overview, 13–21 cells, 71 Lander–Green algorithm, 26–27
developments, 13 infections language impairment (LI), 164–165
field of study, 13 and immune response, Tourette deficits, 166
future trends, 21 syndrome, 342 future research, 166
human genome streptococcal, 130 see also specific language
properties, 13 viral vectors, 71 impairment (SLI)

393
Index

large deletions, 30–31 schizophrenia, 231 malnutrition, 263–264


large insertions, 30–31 single nucleotide polymorphism MANEA, and stimulant
late onset Alzheimer’s disease genotyping, 26–27 dependence, 313
(LOAD) surrogate markers, 26 manic-depression see bipolar
diagnostic criteria, 371 linkage variables, identification, disorder
endophenotypes, 375–377 188–189 MAOA see monoamine oxidase A
genetic susceptibility factors, 373 lithium, 59–60 (MAOA)
genome-wide association studies, LOAD see late onset Alzheimer’s MAPT, and Alzheimer’s disease,
374–375 disease (LOAD) 373–374, 376–378
law, and behavioral genetics, ethical local alignments, 42–43 MAST, 40
issues, 326–329 loci, linkages, 14–15 MATCH, 40
LD see linkage disequilibrium (LD) locomotor activity changes, animal mate-pair sequencing, 29
LDH (lactate dehydrogenase), 96 models, 177 mating, non-random, 7
learning disabilities, 160–166 logarithm of odds (LOD) scores, 18, Maudsley Twin Register, 121
definitions, 160–161 24, 26–27, 122, 231 MC2R see melanocortin receptor
diagnostic issues, 160–161 lung cancer type 2 (MC2R)
prevalence, 160–161 and cigarette consumption, 287 MDD see major depressive disorder
leptin gene, 58–59 genetics, 292–295 (MDD)
LI see language impairment (LI) prevalence, 287 measured risk factors, 154–156
linkage analysis, 17 lung disease, genetics, 292–295 MECP2, and Rett’s disorder, 364–365
Tourette syndrome, 338–339 lymphocyte protein studies, 50 medicine, and behavioral genetics,
linkage disequilibrium (LD), lyonization, 79–81 ethical issues, 326–329
200, 309 MedScan, 46
biological significance, 20 Macromolecular Structure Database meiosis, chromosomes and, 14–15
concept of, 27, 231 (E-MSD), 43 melanocortin receptor type 2 (MC2R)
functional assessments, 20 major depressive disorder (MDD), 4 and opioid addiction, 299–300
genetic variants, 14–17 adoption studies, 215 roles, 299–300
mapping, 17–18 association studies, 217–218 MEME, 40
single nucleotide candidate genes, 218–219 Mendel, Gregor Johann, 112
polymorphisms, 24 citalopram studies, 55 Mendelian genetics, 324
and variant segregation in diagnosis, 227 Mendelian laws, of inheritance, 1–2
populations, 15–17 diagnostic criteria, 212 Mendelian ratio, 2
linkages epigenetics, 223–224 mens rea, 327
biological information, leveraging future research, 226 mental disorders see psychiatric
strategies, 35 etiology, 212 disorders
biological significance, 20 family studies, 213–214 mental illnesses see psychiatric
functional assessments, 20 future research, 225–227 disorders
genetic variants, 14–17 gene–environment interactions, 224 MERLIN program, 26–27, 122
mapping, 17–18 gene expression, 222–223 MET, and autism, 189–190, 192–193
special designs, 18–19 future research, 226 methamphetamine dependence, 314
and recombination, 14–15 gene function, 222–223 genome-wide association
linkage studies, 26–27 future research, 226 studies, 313
Alzheimer’s disease, 374 murine models, 222–223 methadone, opioid addiction
analysis, 26–27 gene mapping, 215–222 treatment, 302
anorexia nervosa, 265–266 genetic epidemiology, 212–215 methylation
autism genetics, 212–227 cytosine, 298
historical background, 183–188 genome-wide association studies, histones, 79–82, 88
overview, 187–188 219–222 see also DNA methylation
bipolar disorder, 200–202 linkage studies, 215–216 (DNAm)
bulimia nervosa, 265–266 pathophysiology, 227 methylphenidate, attention-deficit
databases in, 35 pharmacogenetics, 53, 225 hyperactivity disorder studies,
early, 34 future trends, 227 176, 179
eating disorders, 265–266 pharmacokinetics, 54 methylthioninium chloride
follow-up, alcoholism, 283–284 phenotypes (MTC), 374
major depressive disorder, 215–216 alternative, 216–217 mGluR theory, and fragile
obesity, 274 definition, 212 X syndrome, 367–368
obsessive–compulsive disorder, 122 future research, 227 mice
panic disorder, 92, 99 sex differences, 215–216 in animal models, 72–73
psychostimulant dependence, 308 twin studies, 214–215, 224 COMT studies, 300–301

394
Index

DISC1 studies, 76, 251 mood stabilizers, pharmacogenetics, neurocognitive deficits, anorexia
knock-in, 73 59–61 nervosa, 263
knockout, 73, 176–177 MOPr see mu-opioid receptor (MOPr) neurocognitive phenotypes,
major depressive disorder studies, motif-based alignments, 42 alcoholism, 353–359
222–223 MRPL19, and reading disability, 163 neurocognitive tests, attention-deficit
over-grooming behavior, 128–130 MSK1, 85 hyperactivity disorder studies,
schizophrenia studies, 251–253 MSP (multiple scan probability), 175
transgenic, 73 202 neurodegenerative diseases, common
microarray platforms, 49 MTC (methylthioninium themes in, 378
microarray studies, Tourette chloride), 374 neuroimaging, endophenotypes, 351
syndrome, 342–343 MTHFR, and major depressive neuroimaging studies
migraine, and bipolar disorder, 208 disorder, 216, 218–219 attention-deficit hyperactivity
migration studies multiple alignments, 42 disorder, 176
genetic epidemiology, 3 multiple comparisons, genetic panic disorder, 102–106
heritability, 14 variants, 19–20 neurons, olfactory, 71
missing heritability, 274–275 multiple scan probability (MSP), 202 neuropeptide systems, genes involved
mitral valve prolapse (MVP), 101 multi-threshold multifactorial model, in, 95
MoD Tools, 40 196–198 neuropsychological deficits,
molecular genetic studies multivariate genetic models, measured schizophrenia, 234–235
antisocial behavior, 152 risk factors, 155–156 neuropsychopharmacology, field of
panic disorder, 92–97 mu-opioid receptor (MOPr) study, 53
molecular manipulations, 71, 73–74 epigenetics, 298–299 neuroticism, 117
molecular profiling, 75 and opioid addiction, 297–298 neurotrophic model, of depression, 219
monoamine oxidase A (MAOA) roles, 297 next-generation sequencing, 23
antisocial behavior studies, MUSCLE, 42 analysis, 30–31
152–154 Mutation Discovery, 37–38 data analysis, 30–31
attention-deficit hyperactivity mutations data standards, 31
disorder studies, 172 analysis tools, 37 emerging platforms, 29
panic disorder studies, 94 databases, 36–37, 39 sequencing technologies, 31–32
monozygotic (MZ) twins, 2, 14 and non-uniform recombination standards, 30–31
antisocial behavior studies, 145 rates, 19 technologies, 28–29
antisocial personality disorder schizophrenia, 251 NFKB1, and alcoholism, 283
studies, 317 MVP (mitral valve prolapse), 101 NHE (Naples high-excitability) rats,
attention-deficit hyperactivity MYOCD, and heroin addiction, 178–179
disorder studies, 176 302–303 nicotine dependence, 287–295
autism studies, 183 and alcohol dependence, 294
bipolar disorder studies, 196 nAChRs (nicotinic acetylcholine allele frequency distribution, 293
carbon dioxide hypersensitivity receptors), 290–291 comorbidities, 287–288, 294–295
studies, 96 NAP1L5, and obesity, 273, 275 shared etiologies, 294
extraversion studies, 117 Naples high-excitability (NHE) rats, definitions, 289
major depressive disorder, 214–215 178–179 and depression, 294
neuroticism studies, 117 Naples low-excitability (NLE) rats, 178 diagnostic criteria, 288
obesity studies, 272 National Center for Biotechnology etiology, 287–288
obsessive–compulsive disorder Information (NCBI), genetics, 288
studies, 121 39, 42, 44 future research, 295
panic disorder studies, 92 National Institute of Diabetes and genome-wide association studies,
personality disorder studies, 318 Digestive and Kidney Diseases 289–290, 295
post-traumatic stress disorder (NIDDK), 100–101 limitations, 290
studies, 136 Nazis, eugenics, 325–326 heritability, 289
schizophrenia studies, 235–237 NCAM1, and alcoholism, 281–282 prevalence, 287–288, 295
stimulant dependence studies, NCBI Blast, 42 and psychiatric disorders, genetics,
307–308 NCBI (National Center for 294–295
Tourette syndrome studies, 337 Biotechnology Information), stages, 288
mood disorders 39, 42, 44 twin studies, 288
adoption studies, 4–5 NCBI VAST, 43 nicotinic acetylcholine receptors
and anorexia nervosa, 263 Needleman–Wunsch algorithm, 42 (nAChRs), 290–291
and bulimia nervosa, 263 Neuregulin-1, 72–73 NIDDK (National Institute of
genetic epidemiology, 4 neurochemical metabolites, as Diabetes and Digestive and
heritability, 196 endophenotypes, 351–353 Kidney Diseases), 100–101

395
Index

NLE (Naples low-excitability) animal models, 128 risk factors, 297


rats, 178 biological bases, 121 and tryptophan hydroxylase 1,
noncoding RNA, Tourette syndrome candidate genes, 121, 123–126 301–302
studies, 343 chromosomal anomalies, 130, 130 and tryptophan hydroxylase 2,
nonparametric linkage (NPL) method, chromosomal regions, 301–302
122 candidate, 123 opioid system, and alcoholism, 282
Tourette syndrome, 339 and chromosome 22q deletion oppositional defiant disorder, 145
nonpsychiatric disorders, and syndrome, 127–128 heritability, 149
psychiatric disorders, 101–102 and dystonia myoclonus prevalence, 149
nonrandom mating, 7 syndrome, 127 symptoms, 149
nonsynonymous polymorphisms, endophenotypes, 188 twin studies, 149
39–40 environmental factors, 130 OPRD1
noradrenergic pathway, animal family studies, 121, 123 and alcoholism, 282
models, 177 genes and eating disorders, 265
nortriptyline rare, 130 OPRK1
pharmacokinetics, 53–54 SAPAP3, 128–130 and alcoholism, 282
studies, 55–56 SLC1A1, 123–124, 130 and opioid addiction, 299
NOS1AP, and schizophrenia, 238–239 genetic epidemiology, 5 OPRM1
NPAS3, and schizophrenia, 247 genetics, 121–131 and alcoholism, 282
NPL (non-parametric linkage) formal, 122 and opioid addiction, 297–298
method, 122 future research, 130–131 ordered subset analysis
NPY genome-wide association studies, (OSA), 188
and alcoholism, 282 122–123, 130 ORegAnno, 40–41
and cocaine administration, 83–84 heterogeneity, 130 ORs see odds ratios (ORs)
NRG1, and schizophrenia, 204, 243, linkage studies, 122 OSA (ordered subset analysis), 188
252 onset, 121 OSBPL1A, and bipolar disorder,
NTRK2 prevalence, 121 206–207
and bipolar disorder, 203–204, 206 and Tourette syndrome, 338 OXTR, and autism, 190
and suicidal ideation, 56–57 twin studies, 121
NTRK3, and major depressive OCD see obsessive–compulsive P2RX7, and bipolar disorder, 204
disorder, 219 disorder (OCD) P50 suppression test, 350
odds ratios (ORs), 18 PAG see periaqueductal gray region
obesity in generalized anxiety disorder, 5 (PAG)
candidate genes, 273, 277 in genome-wide association pairwise alignments, 42
copy number variations in, 273 studies, 27 PALB2, and bipolar disorder, 205
disparate approaches, 275 offending, age of onset, 154 PANDAS (pediatric autoimmune
epigenetic modifications, 275–276 6-OHDA (6-hydroxydopamine), 177 neuropsychiatric disorders
gene–environment interactions, olanzapine, 57–58, 276 associated with streptococcal)
272–273 olfactory neurons, 71 infections, 130
genetic expression differences, 276 oligogenic modes, of inheritance, 6 panic attacks (PAs)
genetics, 272–277 Online Mendelian Inheritance in Man induction, 96
applications, 276 (OMIM), 36–37 and panic disorder, 91
future research, 276 opioid addiction panic disorder (PD), 90–107
genome-wide association studies, 274 and catechol-O-methyltransferase, and agoraphobia, 118
heritability, 272 300–301 and brain abnormalities, 102
linkage studies, 274 current research, 303 case control studies, 98
parent of origin effects, 275 gene association studies, comorbidities, 90–92
polygenic model, 275 hypothesis-driven, 302 diagnosis, 90
prevalence, 272 genetics, 297–303 diagnostic criteria, 90
prevention, 276 genome-wide association studies, variance, 92
protective genes, 276 302–303 epidemiology, 90
rare genetic variants, 273 and 5-hydroxytryptamine-1B family studies, 90–92
resistance to, 276 receptor, 301 fear processing, 107
therapy, 276 and kappa-opioid receptor, 299 functional imaging studies,
twin studies, 272 and melanocortin receptor type 2, 102–104, 107
obsessive–compulsive disorder (OCD) 299–300 future research, 107
age of onset, as secondary methadone treatment, 302 genes
phenotype in linkage and mu-opioid receptor, in adenosine system, 95
studies, 122 297–298 in dopamine system, 94–95

396
Index

in GABA-ergic system, 95–96 PCL-R (Psychopathy Checklist- definition, 53


in G-protein signaling regulators, Revised), 150–151 lithium, 59–60
95–96 PCM1, and schizophrenia, 243 major depressive disorder, 53, 225
in neuropeptide systems, 95 PD see panic disorder (PD) future trends, 227
related to carbon dioxide PDGene, 37–38 mood stabilizers, 59–61
hypersensitivity, 96 PDYN, and alcoholism, 282 profiling, 65
in serotonin system, 93–94 pediatric autoimmune in psychiatry, 53–65
genetic epidemiology, 5, 90 neuropsychiatric disorders schizophrenia, 251
genetic studies, 90 associated with streptococcal see also psychiatric
conclusions, 96–97 (PANDAS) infections, and pharmacogenetics
issues, 107 obsessive–compulsive pharmacokinetics
limitations, 106 disorder, 130 antidepressants
molecular, 92–97 PENK, and alcoholism, 282 genetic effects on, 53–54
genome-wide association studies, periaqueductal gray region (PAG), antipsychotic drugs, genetic
93, 106 102–103 factors, 57
heritability, 92 perinatal events, Tourette definition, 53
imaging genetics, 104–106 syndrome, 342 pharmacotherapy
linkage studies, 92, 99 Perlegen, 25 applications, 57
neuroimaging studies, 102–106 personality psychiatric disorders, 53
and panic attacks, 91 abnormal vs. normal compared, 316 and single nucleotide
and phobias, 118–119 epigenetics, 322 polymorphisms, 61–65
prevalence, 90 general dimensions, inheritance, phenotypes
probands, 101 320–321 concept of, 347
and structural brain changes, 103 well-functioning, requirements, 317 see also endophenotypes
structural imaging studies, 102 personality disorders phenotypic heterogeneity, 20
symptoms, 90, 97 candidate genes, 321 phenytoin, 59
twin studies, 90–92 common genetic factors, 319–320 phobias
panic disorder syndrome definitional issues, 317 and additive genes, 114–115, 117
expanded spectrum approach, factor analysis, 320 classification, 112
97–102 genetics, 316–322 comorbidities, 113
study characteristics, 99 genome-wide association studies, diagnosis, 112
tests, 100 321–322 epidemiology, 112–113
panic phenotype, studies, 106–107 heritability, 319–321 familial specificity, 114
papilin, encoding, 56–57 inheritance, 318–320 familial transmission, 113
PAPLN, and suicidal ideation, 56–57 multidimensional inventories, 320 genetic epidemiology, 5, 113–118
parent of origin effects, in obesity, 275 quantitative multifactorial future research, 119
Paris Autism Research International variability, 317–318 genetics, 112–120
Sibpair Study (PARIS), twin studies, 318, 320–321 genomic studies, 118–119
183–186 see also antisocial personality heritability, sex differences, 114
Parkinson’s disease, databases, 37–38 disorder and irrational fears, 113–115
paroxetine personality traits, heritability, 326 meta-analyses, 103–104
panic disorder studies, 93 PET (positron emission and panic disorder, 118–119
pharmacokinetics, 53–54 tomography), 102 prevalence, 112–113
PAs see panic attacks (PAs) Pfam, 43 twin studies, 114–115
Pathguide, 45 pharmacodynamics, 53–54 variance, 116
PathSys database, 46 pharmacogenetics see also agoraphobia (AG), social
PATHWAY, 45 anticonvulsants, 59–61 phobia, specific phobia
pathway analysis, 44–46, 70 antidepressants, 53–57 phobic disorders see phobias
commercial tools, 46 adverse reaction studies, 56 phonologic disorders, 160
future trends, 46 antipsychotic drugs, 57 phosphoacetylation, 81
information sources, 44 adverse reaction studies, 58–59 phosphorylation, histones, 79–81
pathway data repositories, 45 attention-deficit hyperactivity PICALM, and Alzheimer’s disease,
pathways disorder, 176 375
autism, 192–193 animal models, 179 PKHD1, and obesity, 276
as networks, 45 of carbamazepine-induced plasmids, transfection, 71
pathway visualization, 45–46 Stevens–Johnson syndrome, 61 Plato, Republic, 324–325
PATIKAweb, 46 in clinical practice, 62 pleiotropic models, 101
PCLO, and major depressive commercial tests, 65 pleiotropy, 6
disorder, 221 conceptual issues, 61–65 PLINK (software), 27

397
Index

P-MATCH, 40 genetics, 365–366 pharmacotherapy, 53


PMCH, and obesity, 276 and genomic imprinting risk factors, 3
PMut, 39–40 disorders, 365 risk ratios, 3
polygenic model, obesity, 275 prevalence, 365 transmission, complex patterns of, 6
polygenic modes, of inheritance, 6 predictive algorithms, 40 psychiatric genetics
polymorphisms PRINTS, 43 endophenotypes in, 347–360
analysis tools, 37 privacy issues, genetic applications, 349–359
nonsynonymous, 39–40 information, 329 future trends, 359–360
regulatory, 40–41 PRNP, mutations, 378 identification criteria, 348–349
see also copy number PRODH, and schizophrenia, psychiatric genetics research,
polymorphisms (CNPs); single 247–248 computer-based analysis and
nucleotide polymorphisms (SNPs) prodynorphin gene, and stimulant resources, 34–47
Polyphen, 39–40 dependence, 312 psychiatric pharmacogenetics
POMC, and alcoholism, 282 PROSITE, 43 field of study, 53
population attributable risk, 8 protein chemistry, in genetic future trends, 61–65
population isolates, Tourette susceptibility factor studies, implications for clinical practice,
syndrome studies, 339–340 69–71 61–65
population prevalence, 1–2 protein kinases, 82 psychiatric phenomics, field of
population stratification, 308–309 protein phosphatases, 82 study, 49
positron emission tomography proteins, 43 psychiatric phenotypes,
(PET), 102 overexpression, 71 limitations, 49
postnatal brain maturation, 75 protein sequencing, 35 psychiatry
post-traumatic stress disorder (PTSD), protein structures, 43 epigenetic research, issues, 87
134–143 PRPF31, and autism, 188–189 pharmacogenetics in, 53–65
adoption studies, 136 PSEN1 psychic trauma, 134–135
animal models, 141 and Alzheimer’s disease, 375 see also post-traumatic stress
comorbidities, genetic factors, 139, and familial Alzheimer’s disease, disorder (PTSD)
140 371, 376 psychopathy, 150–151
conceptual development, 134–135 PSEN2 and antisocial behavior, 145
diagnosis, 134 and Alzheimer’s disease, 375 Psychopathy Checklist-Revised
epidemiology, 135 and familial Alzheimer’s (PCL-R), 150–151
diagnostic criteria, issues, 142 disease, 373 psychopharmacology, limitations,
and diathesis–stress model, 134 pseudo single-molecule sequencing, 61–65
epigenetics, 140–141 concept of, 28 psychosocial stress, Tourette
family studies, 135–136 PSI-MI, 45 syndrome, 342
and fear, 103–104 psychiatric disorders psychostimulant dependence, 308–310
future trends, 140–142 animal models, 73 psychostimulants
and gene–environment and antisocial behavior, 145 use of term, 306
interactions, 142 biomarkers, identification, 49–50 see also stimulant dependence
genetic studies, 135–140 candidate gene association psychotic depression,
candidate gene approach, 137, studies, 49 pharmacotherapy, 57
138 classification PTSD see post-traumatic stress
case-control association, 137–140 endophenotypes, 9 disorder (PTSD)
future trends, 142–143 lack of validity, 6 PubMed, 46
methodological issues, 142 comorbidities, 7 punishment, and behavioral genetics,
molecular, 137–140 complexity in, sources of, 6 327–329
quantitative, 135–136 and endophenotypes, 349 PupaSuite, 36
genome-wide association studies, family studies, 69 p-values, 50–51
141–142 gene expression studies, 49–51
historical background, 134–135 genetic bases, methods, 23 QTL (quantitative trait loci), 275
prevalence, 135 genetic epidemiology of, 1–10 quantitative multifactorial variability,
sex differences, 135 genetic susceptibility factors for, personality disorders, 317–318
trauma exposure, 135–137 functional validation, 69–77 quantitative trait loci (QTL), 275
twin studies, 136, 140 heritability estimates, 3
PPP3C, and schizophrenia, 252 impacts of genomics on, 9–10 RAI1, and Smith–Magenis
PPP3CC, and schizophrenia, 243 and nicotine dependence, genetics, syndrome, 366
PPYR1, and obesity, 273 294–295 rare genetic variants, 23
Prader–Willi syndrome, 79–81, 275 and non-psychiatric disorders, autism, 191–192
and autism compared, 365 101–102 obesity, 273

398
Index

rare variant hypothesis, 191–192 SAD (social anxiety disorder), Kraepelinian type, 232–233
RAVEN (Regulatory Analysis of 100–101, 104 linkage studies, 231
Variation in Enhancers), 40 sampling, family studies, 7 loci, 239
RCSB PDB, 43 SAM (sequence alignment/MAP), 31 murine studies, 251–253
RD see reading disability (RD) Sanger sequencing technology, 28–29 neuropsychological deficits, with
Reactome, 45 SAPAP3, 128–130 family genetic control, 234–235
read depth (RD), 30–31 SBML, 45 pharmacogenetics, 251
reading disability (RD), 161–163 schizophrenia (SZ) pharmacotherapy, 57
comorbidities, 165 adoption studies, 237 subtypes, family studies, 232–233
deficits, 166 and affective disorders, 233–234 twin studies, 235, 350–351
future research, 166 age of onset, 231–232 segregation analysis, 237–238
gene localizations, 161 anatomical and histological variance component analysis,
genetics, 161 abnormalities, 76 237–238
and language impairment animal models, 73, 76 SCOPE, 40
compared, 164 anticipation and, 230–231 second-generation sequencing
sex differences, 161 association studies, 238 see next-generation sequencing
read pairs (RPs), 30–31 and bipolar disorder, 230–231, segregation analysis
recombination, and linkages, 14–15 233–234, 208 schizophrenia, 237–238
recombination rates, non-uniform, copy number variant differences, Tourette syndrome, 337
and mutations, 19 250–251 selection, artificial, 324
RefSeq, 45 genetic association differences, selective serotonin reuptake inhibitors
regulator of G-protein signaling 250–251 (SSRIs)
(RGS), 96 candidate genes, 204, 231, 238–248 adverse reaction studies, 56
Regulatory Analysis of Variation in animal and cell biology studies, depression treatment, 85
Enhancers (RAVEN), 40 251–253 major depressive disorder
regulatory motifs, 41 functional polymorphisms and treatment, 225
regulatory polymorphisms, 40–41 mutations, 251 panic disorder treatment, 93
regulatory RNAs, 51 children, 4 pharmacokinetics, 53–54
relative risk, 8 copy number variations, 249–250 SEMA5A, and autism, 187
RELN, and schizophrenia, 243, 252 co-twin control studies, 235–237 sensory motor gating, and
responsibility, and genetics, 326–329 cytogenetic abnormalities, 249–250 schizophrenia, 350
Rett’s disorder, 364–365 deletions, 249–250 sequence alignment/MAP (SAM), 31
Rett syndrome, etiology, 82 diagnostic criteria, 233–234 sequence alignments, 41–43
RGS4, and schizophrenia, 238–239 diagnostic issues, 232 sequence analysis, 35
RGS (regulator of G-protein duplications, 249–250 see also DNA sequencing
signaling), 96 endophenotypes sequence conservation, 41
risk alleles, 8 eye-tracking dysfunction, 350 Sequenced Treatment Alternatives to
risk estimation, 8 neurocognitive, 350–351 Relieve Depression (STAR*D)
risk factors sensory motor gating, 350 Study, 55–56, 226
antisocial behavior, 154–156 environmental risk factors, 4 treatment emergent suicidal
see also measured risk factors epigenetics, 231–232 ideation studies, 56–57
risk ratios, psychiatric disorders, 3 familial recurrence, 232–233 sequence read format (SRF), 31
risperidone, 58 familial transmission, 207 sequencing, 28
RNA interference (RNAi), 71, family studies future research, 46–47
73–74 cognitive deficits and imaging see also DNA sequencing, next-
RNA sequencing, analysis, 35 changes, 234 generation sequencing
ROBO1, and reading disability, 163 segregation analysis, 237–238 Sequenom, 25
Roche, 454 Sequencing System, variance component analysis, serotonergic pathways, attention-deficit
28–29 237–238 hyperactivity disorder, 172
Roche Diagnostic AmpliChip CYP450 functional polymorphisms, 251 serotonergic system, 58, 105–106
test, 65 future research, 46–47 serotonin, in carbon dioxide
rodents genetic epidemiology, 3–4 hypersensitivity, 96
depression studies, 85 genetic linkage, 238 serotonin norepinephrine reuptake
drug-addiction studies, 83–85 genetics, 230–253 inhibitors (SNRIs)
molecular manipulations, 73, 74 future research, 253 adverse reaction studies, 56
see also mice genetic susceptibility factors, 4, 69 depression treatment, 85
RORA, and major depressive heritability, 230–231 pharmacokinetics, 53–54
disorder, 225 and idiopathic generalized epilepsy, serotonin receptor gene (HTR2A)
RPs (read pairs), 30–31 249–250 see HTR2A

399
Index

serotonin receptors, 107 linkage studies, 26–27 SLI2 locus, 164, 166
polymorphisms, 93–94 see also high-density single SLI see specific language
serotonin receptor subunit genes, and nucleotide polymorphism (SNP) impairment (SLI)
antipsychotic drugs, 58 genotyping SLITRK1, and Tourette syndrome,
serotonin system, genes involved in, single nucleotide polymorphisms 336–337, 341
93–94, 123–126 (SNPs), 13 Slit and Trk-like family member 1
serotonin transporter gene (SLC6A4) applications (SLITRK1) see SLITRK1
see SLC6A4 disease mapping, 23–24 Smith–Magenis syndrome
serotonin-transporter-linked early, 24–25 (SMS), 366
polymorphic region classic, 24 Smith–Waterman algorithm, 42
(5-HTTLPR), 54–55 classification, 24 smoking initiation, heritability, 288
and adverse reactions, 56 databases, 36–39 SMS see Smith–Magenis
anorexia nervosa studies, 267 definition, 23–24 syndrome (SMS)
bipolar disorder studies, 203 functional, 24 SNAP-25, and attention-deficit
major depressive disorder linkage disequilibrium, 24 hyperactivity disorder,
studies, 224 mapping, 24 172–173
obsessive–compulsive disorder nonfunctional, 24 SNPExpress, 44
studies, 123–126 and pharmacotherapy, 61–65 SNP genotyping see single nucleotide
and personality disorders, 321 site-directed mutagenesis, 69–70 polymorphism (SNP)
polymorphisms, 93, 105–106 SJS (Stevens–Johnson syndrome), genotyping
post-traumatic stress disorder carbamazepine-induced, SNPs3D, 39–40
animal studies, 140 pharmacogenetics, 61 SNPs see single nucleotide
studies, 140 SLA6A2, animal models, 177 polymorphisms (SNPs)
see also SLC6A4 SERT, SLC1A1 SNPs (single nucleotide
organization, 125 in obsessive–compulsive disorder, polymorphisms), 13
setshifting, and anorexia nervosa, 263 123, 130 SNRIs see serotonin norepinephrine
sex differences studies, 124 reuptake inhibitors (SNRIs)
animal phobias, 115 SLC1A3, in attention-deficit social anxiety disorder (SAD),
antisocial personality disorder, hyperactivity disorder, 100–101, 104
317–318 172–173 social phobia
autism, 188, 191 and schizophrenia, 252 and avoidant personality disorder
fragile X syndrome, 367 SLC6A2 compared, 117
major depressive disorder, and attention-deficit hyperactivity classification, 112
215–216 disorder, 172 diagnosis, 112
phobias and major depressive disorder, familial transmission, 113
heritability, 114 55–56, 218–219 genomic studies, 118
prevalence, 113, 115 and weight loss, 276 prevalence, 112–113
post-traumatic stress disorder, 135 SLC6A3, 137–139 traits, family studies, 117–118
reading disability, 161 animal models, 176–178 see also generalized social
Tourette syndrome, 342 in attention-deficit hyperactivity phobia
shank3 point mutation, 192 disorder, 169, 173, 176 sodium butyrate, 224
shell shock, 134–135 and stimulant dependence, 310 SP4, and major depressive
see also post-traumatic stress variants, 169–171, 175 disorder, 221
disorder (PTSD) SLC6A4 spasmodic torticollis, 127
SHRs (spontaneously hypertensive anorexia nervosa studies, 267–269 SPCH1 locus, 164
rats), 178 and autism, 190 specific environment, use of
Sibling-pair designs, autism studies, bipolar disorder studies, 203 term, 114
183 major depressive disorder studies, specific language impairment (SLI),
SIFT, 39–40 218–219 160
simple phobia see specific phobia obsessive–compulsive disorder specific phobia
Single Nucleotide Polymorphism studies, 123–126 classification, 112
Database (dbSNP), 39 panic disorder studies, 93 diagnosis, 112
single nucleotide pharmacogenetic studies, 54–55 familial transmission, 113
polymorphism (SNP) post-traumatic stress disorder genomic studies, 118
genotyping, 23–28 studies, 140 prevalence, 112–113
copy number analysis, 26 see also serotonin-transporter- specific reading disability (SRD), 160
family studies, 26–27 linked polymorphic region speech sound disorder (SSD), 165–166
genome-wide association studies, (5-HTTLPR) definition, 160
27–28 SLI1 locus, 164, 166 prevalence, 160

400
Index

spontaneously hypertensive rats subPSEC, 39–40 tobacco use


(SHRs), 178 substance dependence, and antisocial prevalence, 287
SRD see specific reading behavior, 145 see also nicotine dependence
disability (SRD) substance use disorders, 148 Tourette syndrome (TS)
SRF (sequence read format), 31 and antisocial behavior, 150 and anxiety disorders, 338
SSD see speech sound family studies, 5, 6 and attention-deficit hyperactivity
disorder (SSD) gene–environment interactions, disorder, 338
SSRIs see selective serotonin reuptake 150 candidate genes, 341
inhibitors (SSRIs) genetic epidemiology, 5 comorbidities, 337–338
stable tubule-only polypeptide heritability, 150 copy number variations, 343
(STOP) proteins, and twin studies, 5–6 cytogenetic abnormalities, 340–341
schizophrenia, 252 suicidal behavior, and cortisol and depression, 338
STAR*D Study see Sequenced levels, 353 environmental factors, 342
Treatment Alternatives to suicidal ideation, 56–57 epigenetics, 343–344
Relieve Depression (STAR*D) Systematic Treatment Enhancement family studies, 337
Study Program for Bipolar Disorder genetics, 126–127, 336–344
state–trait characteristics (STEP-BD), 205–206 future research, 344
anorexia nervosa, 263–264 SZ see schizophrenia (SZ) future trends, 342–344
bulimia nervosa, 263–264 genome-wide association studies,
stem cells, 71 TAAR6, and schizophrenia, 342–343
see also induced pluripotent 241–243 heritability, 337
stem (iPS) cells TACR3, and alcoholism, 283 infection and immune response, 342
STEP-BD (Systematic Treatment TAMAL, 36 linkage analysis, 338–339
Enhancement Program for tamoxifen, 276 microarray studies, 342–343
Bipolar Disorder), 205–206 TaqI, 172 noncoding RNA studies, 343
sterilization, 325 tardive diskinesia (TD), antipsychotic- non-parametric linkage
Stevens–Johnson syndrome (SJS), induced, 58–59 method, 339
carbamazepine-induced, target enrichment, hybridization- and obsessive–compulsive
pharmacogenetics, 61 based, 32 disorder, 338
stimulant dependence TCAs see tricyclic antidepressants perinatal events, 342
candidate genes, 312 (TCAs) population isolates, 339–340
and dopamine-related genes, TCI (Temperament and Character prevalence, 337
310–312 Inventory), 320–321 psychosocial stress, 342
family studies, 306–307 T-Coffee, 42 research, conclusions, 336
future research, 314 TD see tardive diskinesia (TD) segregation analysis, 337
genetics, 306–314 TDTs see transmission disequilibrium sex differences, 342
genome-wide association studies, tests (TDTs) studies, 336
312–313 Technical University of Denmark, symptoms, 336
heritability, 306–308 Center for Biological Sequence twin studies, 337
markers, 309–310 Analysis, 43 toxic epidermal necrolysis (TEN),
prevalence, 306 Temperament and Character carbamazepine-induced,
twin studies, 306–307 Inventory (TCI), 320–321 60–61
see also psychostimulant TEN (toxic epidermal necrolysis), TPH1, 301–302
dependence carbamazepine-induced, TPH2, 301–302
STOP (stable tubule-only polypeptide) 60–61 TPH (tryptophan hydroxylase), 94
proteins, and TESIs (treatment emergent suicidal TRa1, mutations, 178
schizophrenia, 252 ideations), antidepressant- training data, 40
stratification, 20 induced, 56–57 traits, heritability, 113
streptococcal infections, and testing TRANSFAC database, 40
obsessive–compulsive for-profit, for genetic susceptibility, transgenic animals, 178
disorder, 130 331–332 transgenic mice, 73
structural alignments, 43 issues, addictions, 331 Alzheimer’s disease studies, 374
structural imaging studies, panic in workplace, 333–334 translocations, 30–31
disorder, 102 TF, and Alzheimer’s disease, 376 transmission disequilibrium tests
structural variants TH (tyrosine hydroxylase), 201–202, (TDTs), 18–19
attention-deficit hyperactivity 204 advantages, 27–28
disorder, 173 tics overview, 27–28
types of, 30–31 characterization, 336 tranylcypromine, 86, 224
STY13, and cocaine dependence, 313 comorbidities, 337–338 TRAR4, and schizophrenia, 241–243

401
Index

trauma exposure panic disorder, 90–92 VCF (variant consensus format),


assaultive versus non-assaultive, personality disorders, 318, 31
136–137 320–321 velo-cardio-facial syndrome (VCFS),
epidemiology, 135 phobias, 114–115 4, 249–250
genetic factors, 136–137 environmental variance, 114 genetics, 127–128, 247–248
non-genetic factors, 137 genetic variance, 114 venlafaxine, pharmacokinetics,
traumatic neurosis, 134–135 prevalence, 113 53–54
see also post-traumatic stress disorder post-traumatic stress disorder, 136, ventricle-brain ratios (VBRs), 234
(PTSD) 140 vesicular neurotransmission,
treatment emergent suicidal ideations psychopathy, 151 regulatory mechanisms,
(TESIs), antidepressant- schizophrenia, 235, 350–351 172–173
induced, 56–57 segregation analysis, 237–238 violence, 145
trichotillomania, genetics, 128–130 variance component analysis, genetic factors, 328–329
tricyclic antidepressants (TCAs), 53 237–238 viral vectors, infection, 71
TRRD database, 40 stimulant dependence, 306–307 VisANT, 46
tryptophan hydroxylase 1 (TPH1) substance use disorders, 5–6 VISTA Enhancer Browser, 40–41
see TPH1 Tourette syndrome, 337 VMP, and reading disability, 163
tryptophan hydroxylase 2 (TPH2) see also dizygotic (DZ) twins, VNTRs (variable number tandem
see TPH2 monozygotic (MZ) twins repeats), 94, 203
tryptophan hydroxylase (TPH), 94 tyrosine hydroxylase (TH), 201–202,
TS see Tourette syndrome (TS) 204 warfarin, pharmacogenetic
TStag, 44 profiling, 65
TTC12, and alcoholism, 281–282 UBE3A, and Angelman syndrome, 365 war neurosis, 134–135
T variant, 15–17 UBE3C, and major depressive see also post-traumatic stress
twin studies disorder, 225 disorder (PTSD)
alcoholism, 279–280 UCP1, and obesity, 273 WDR59, and bipolar disorder,
anorexia nervosa, 264–265 UCSC Genome Browser, 40–41 206–207
antisocial behavior, 145–148, UHMK1, and schizophrenia, 238–239 WebLogo, 41
156 unconditioned fear, 102–103 weight gain
gene–environment unipolar depression see major antipsychotic-induced, 58–59
interactions, 152 depressive disorder (MDD) drug-induced, 276
antisocial personality disorder, 317 U.S., v. Durham, 327–328 weight loss, genetics, 276
attention-deficit hyperactivity Weka, 40
disorder, 149–150, 168, 176 Val66Met, polymorphisms, 140 WGAS see genome-wide association
autism, 183 val158met, polymorphisms, 105 (GWA) studies
bipolar disorder, 196, 197 valproate, 59, 224 WGA studies see genome-wide
bulimia nervosa, 264–265 valproic acid, 299 association (GWA)
carbon dioxide hypersensitivity, 96 Val/Val genotype, 172 studies
conduct disorder, 149 variable expressivity, 6 Wistar Kyoto (WKY) strain, 178
externalizing disorders, 116–117 variable number tandem repeats word methods, 42
extraversion, 117 (VNTRs), 94, 203 workplace, testing in, 333–334
generalized anxiety disorder, 119–120 variance component analysis, WorldWide Protein Data
genetic epidemiology, 2 schizophrenia, 237–238 Bank, 43
heritability, 14 variant consensus format (VCF), 31
internalizing disorders, 116–117 variant identification X-inactivation, 79–81
major depressive disorder, 214–215, in human genome, 13
224 strategies, 34 yeast two-hybrid assays, 69–71
neuroticism, 117 variant segregation, and linkage
nicotine dependence, 288 disequilibrium, 15–17 zebrafish, as animal models, 73
obesity, 272 VBRs (ventricle-brain ratios), 234 zinc finger proteins, 87–88
obsessive–compulsive disorder, 121 VCFS see velo-cardio-facial syndrome ZNF804A, and schizophrenia,
oppositional defiant disorder, 149 (VCFS) 239–240

402

You might also like