+ All Categories
Home > Documents > Copy Number Variation in Neurodevelopmental Disorders

Copy Number Variation in Neurodevelopmental Disorders

Date post: 23-Oct-2021
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
202
i Copy Number Variation in Neurodevelopmental Disorders by Anath Christopher Lionel A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Molecular Genetics University of Toronto © Copyright by Anath Christopher Lionel 2014
Transcript
Page 1: Copy Number Variation in Neurodevelopmental Disorders

i

Copy Number Variation in Neurodevelopmental Disorders

by

Anath Christopher Lionel

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Molecular Genetics University of Toronto

© Copyright by Anath Christopher Lionel 2014

Page 2: Copy Number Variation in Neurodevelopmental Disorders

ii

Copy Number Variation in Neurodevelopmental Disorders

Anath Christopher Lionel

Doctor of Philosophy

Molecular Genetics University of Toronto

2014

Abstract

Recent research on genetic etiologies of different neurodevelopmental conditions such as Autism

Spectrum Disorder (ASD), Attention Deficit Hyperactivity Disorder (ADHD) and schizophrenia

has highlighted a key role for rare copy number variants (CNVs). The study of such rare

deletions and duplications in the genomes of patients has proven to be a powerful strategy for

rapid identification of novel risk genes. Intriguingly, there is an emerging pattern of certain genes

being implicated by rare CNVs across clinically distinct disorders. Findings of genetic overlap

are consistent with comorbidity and shared traits often observed among these conditions and hint

at common underlying biological processes or pathways such as synaptic development. This

trend of overlapping genetic loci predisposing to different phenotypes has not been

systematically investigated in an unbiased fashion on a genome-wide scale.

In this thesis, I describe high resolution microarray analysis of three newly characterized

Canadian cohorts of individuals with different neurodevelopmental disorders (ASD, ADHD and

schizophrenia) in a uniform manner and with two main objectives. First, within each cohort, rare

CNVs were detected and used to identify new genetic risk loci for each disorder. My

prioritization strategy focused on rare CNVs of two types: those of de novo origin and those

significantly enriched in multiple unrelated patients compared to controls. Second, overlapping

Page 3: Copy Number Variation in Neurodevelopmental Disorders

iii

rare CNVs among the three cohorts were further investigated to assess cross-disorder genetic

overlap and pinpoint shared risk genes.

These analyses confirmed regions previously implicated in genetic risk for neurodevelopmental

disorders, and also revealed novel CNV loci such as deletions affecting NRXN3 in ASD,

MACROD2/FLRT3 in ADHD and the 2q13 region in schizophrenia. Cross-disorder rare CNV

comparisons highlighted several shared risk genes including ASTN2/TRIM32 in ASD and ADHD

and GPHN in ASD and schizophrenia. Follow-up clinical and molecular characterization of

these CNVs revealed factors modulating their penetrance including gender and isoform-specific

effects. These results provide support for the role of rare CNVs in the genetic etiologies of ASD,

ADHD and schizophrenia; provide evidence for shared susceptibility genes for different

neuropsychiatric disorders; and identify new risk genes to guide clinical genetic testing as well as

the development of molecular therapeutics.

Page 4: Copy Number Variation in Neurodevelopmental Disorders

iv

Acknowledgments

It is a great pleasure to thank the many wonderful people I have been privileged to work with,

and who have helped me during my time in graduate school.

Firstly, I thank my supervisor Dr. Stephen Scherer for his confidence in me, and for his

mentorship, guidance, advice and financial support during my PhD journey. He has been an

outstanding role model exemplifying research leadership and ‘big picture’ scientific insights, the

initiation and strengthening of powerful collaborative partnerships and effective scientific

communication and writing. My graduate training has been enriched by experiencing first hand

these key facets of research and by the many opportunities and resources that have been provided

in his lab. I will always be proud to have been his graduate student. I am also grateful to the

members of my supervisory committee, Dr. Lucy Osborne and Dr. Gary Bader for their patient

guidance, helpful suggestions and encouragement which have kept me on track during my

research and have been particularly instrumental during key graduate school milestones

including the reclassification exam and completion of this thesis. I am thankful to the other

members of my PhD examination committee including Dr. Paul Pavlidis, Dr. Sean Egan and Dr.

Berge Minassian for their insightful questions and helpful suggestions.

I would like to thank several past and present members of the Scherer academic lab and the

Centre for Applied Genomics (TCAG). In particular, I am indebted to Dr. Christian Marshall for

his research prowess and for the guidance provided, in his own inimitable and humorous way,

from my first days as a rotation student right up to the end of my PhD over the course of many

different projects. I have learned much about genetic research, Canada and life from him. I have

also had the privilege of working closely with Bhooma Thiruvahindrapuram, Dr. Kristiina

Tammimies, Dr. Daisuke Sato, Dr. Andrea Vaags, Dr. Daniele Merico and Matthew Gazzellone

and I am grateful for the invaluable contributions of their excellent scientific skills, support and

kindness to my research training and graduate experience. I thank the other graduate students in

the Scherer lab: Dr. Andrew Carson and Dr. Layla Parker-Katiraee for their guidance during my

early days in the lab and my contemporaries Dr. Andy Pang, Matthew Gazzellone and Lia

Page 5: Copy Number Variation in Neurodevelopmental Disorders

v

D’Abate for their friendship, encouragement, frequent coffee breaks and abundant laughter.

Other current and former members of TCAG that I would like to thank for their assistance at

different times during my research training and friendships that have enriched my time in the lab

include Sanjeev Pullenayegum, Dr. Mary Ann George, Jennifer Howe, Jenny Kaderali, Dr.

Richard Wintle, Dr. Lars Feuk, Dr. Hameed Khan, Dr. Susan Walker, Dr. Ryan Yuen, Dr.

Mohammed Uddin, Dr. Mehdi Seno, Dr. Aparna Prasad, Dr. Chao Lu, Lynette Lau, Dr. Tara

Paton, Elango Cheran, Robert Ziman, Joe Whitney, Deepthi Rajagopalan, Dr. John Wei, Dr.

Zhuozhi Wang, Dennis Chen, Divya Mandyam, Vibha Raghavan and Jeff MacDonald.

It has been a great pleasure and an inspirational experience to work closely with my many

clinical collaborators, of whom those at the Centre for Addiction and Mental Health (Dr. Anne

Bassett, Dr. Gregory Costain, Dr. John Vincent, Dr. Abdul Noor), the Hospital for Sick Children

(Dr. Russell Schachar, Dr. Jennifer Crosbie, Dr. James Stavropoulos, Dr. Wendy Roberts, Dr.

Melissa Carter, Dr. Gregory Handrigan and the Autism Research Unit team) and Toronto

General Hospital (Dr. Candice Silversides) deserve special mention. My research work would

not have been possible without funding and support from different sources including

NeuroDevNet, the SickKids foundation, Autism Speaks and the Ontario Ministry of Education

and Training.

I would also like to mention and thank friends who have been an important source of support and

inspiration during my PhD including those from my undergraduate days - in particular, Sherwin

Abraham, Selvaprabhu Nadarajah and Kaushik Gopal - and those I met during graduate school

including Niladri Chattopadhyay, Kirill Zaslavsky, Santhosh Dhanraj, Bhaskar Chanda, Vivek

Mahadevan and Brige Chugh. I thank my sisters Achsah and Amrita Lionel, and my relatives in

India and elsewhere for their support. Finally, I would like to dedicate this thesis to my parents

Jesuraj and Ranjini Lionel, whose constant encouragement, love and prayers have helped me

through tough times, and have contributed immeasurably to everything I have achieved, and to

the person that I am today. I hope I have made them proud.

Page 6: Copy Number Variation in Neurodevelopmental Disorders

vi

Table of Contents

Acknowledgments.......................................................................................................................... iv

Table of Contents ........................................................................................................................... vi

List of Tables ................................................................................................................................. xi

List of Figures ............................................................................................................................... xii

List of Abbreviations ................................................................................................................... xiii

Chapter 1 Introduction .....................................................................................................................1

1.1 Autism Spectrum Disorder (ASD) .......................................................................................3

1.1.1 Clinical profile of ASD ............................................................................................3

1.1.2 Evidence for the genetic etiology of ASD ...............................................................3

1.1.3 Common genetic variant findings in studies of ASD risk .......................................4

1.1.4 Rare genetic variant findings in studies of ASD risk ...............................................5

1.2 Attention Deficit Hyperactivity Disorder (ADHD) .............................................................8

1.2.1 Clinical profile of ADHD ........................................................................................8

1.2.2 Evidence for the genetic etiology of ADHD............................................................9

1.2.3 Common genetic variant findings in studies of ADHD risk ..................................10

1.2.4 Rare genetic variant findings in studies of ADHD risk .........................................10

1.3 Schizophrenia .....................................................................................................................11

1.3.1 Clinical profile of schizophrenia ............................................................................11

1.3.2 Evidence for the genetic etiology of schizophrenia ...............................................12

1.3.3 Common genetic variant findings in studies of schizophrenia risk .......................13

1.3.4 Rare genetic variant findings in studies of schizophrenia risk ..............................13

1.4 Phenotypic and Genetic Overlap Across NDDs ................................................................15

Page 7: Copy Number Variation in Neurodevelopmental Disorders

vii

1.4.1 Comorbidity and clinical connections between ASD and ADHD .........................15

1.4.2 Comorbidity and clinical connections between ASD and schizophrenia ..............18

1.4.3 Comorbidity and clinical connections between ADHD and schizophrenia ...........20

1.4.4 Evidence for shared genetic risk factors across NDDs ..........................................20

1.5 Thesis Rationale and Overview .........................................................................................29

Chapter 2 Thesis Overview and Methodology ..............................................................................34

2 35

2.1 Thesis Overview ................................................................................................................35

2.2 Uniform CNV Analysis Workflow ....................................................................................37

2.2.1 Microarray genotyping...........................................................................................37

2.2.2 Uniform quality control measures .........................................................................37

2.2.3 CNV detection from microarray data ....................................................................40

Chapter 3 CNV Analysis of ASD Cohort Reveals Novel Risk Gene NRXN3...............................42

3 43

3.1 Abstract ..............................................................................................................................43

3.2 Introduction ........................................................................................................................44

3.3 Results ................................................................................................................................44

3.3.1 Candidate risk loci identified by rare CNVs in ASD cohort .................................44

3.3.2 Follow-up of rare exonic deletions at the NRXN3 locus ........................................48

3.3.3 Clinical information from ASD probands with NRXN3 deletions and their families ...................................................................................................................51

3.4 Discussion ..........................................................................................................................54

3.5 Materials and Methods .......................................................................................................56

3.5.1 Study subjects and methodology ...........................................................................56

Chapter 4 Rare Copy Number Variant Discovery and Cross-Disorder Comparisons Identify Risk Genes for ADHD ..............................................................................................................57

Page 8: Copy Number Variation in Neurodevelopmental Disorders

viii

4 58

4.1 Abstract ..............................................................................................................................58

4.2 Introduction ........................................................................................................................59

4.3 Results ................................................................................................................................59

4.3.1 Detection of rare CNVs in ADHD cohort ..............................................................59

4.3.2 Loci implicated by de novo CNVs in ADHD cases ...............................................65

4.3.3 Rare inherited CNVs at loci previously implicated in ADHD and other NDDs ...66

4.3.4 Overlap between rare CNV findings in the ADHD and ASD cohorts ..................67

4.4 Discussion ..........................................................................................................................69

4.5 Materials and Methods .......................................................................................................73

4.5.1 Study subjects in ADHD cohort ............................................................................73

Chapter 5 Pathogenic Rare Copy Number Variants in Community-Based Schizophrenia Suggest a Potential Role for Clinical Microarrays ....................................................................75

5 77

5.1 Abstract ..............................................................................................................................77

5.2 Introduction ........................................................................................................................78

5.3 Results ................................................................................................................................81

5.3.1 Clinically significant and novel large rare structural variants in schizophrenia ....81

5.3.2 Prevalence of clinically significant CNVs in community sample of schizophrenia .........................................................................................................82

5.3.3 Large rare CNVs of unknown significance in schizophrenia ................................82

5.3.4 Genome-wide CNV burden in schizophrenia ........................................................83

5.3.5 Very rare, smaller CNVs identifying genes of interest in schizophrenia ..............84

5.3.6 Functional networks revealed by pathway enrichment analysis ............................85

5.4 Discussion ..........................................................................................................................98

5.5 Materials and Methods .....................................................................................................104

Page 9: Copy Number Variation in Neurodevelopmental Disorders

ix

5.5.1 Schizophrenia sample ascertainment and assessment..........................................104

5.5.2 Control sample for formal analyses .....................................................................105

5.5.3 CNV prioritization, assessment of pathogenicity and prevalence .......................106

5.5.4 Experimental validation of CNVs ........................................................................107

5.5.5 CNV burden analyses ..........................................................................................107

5.5.6 Functional network from gene association analysis ............................................108

Chapter 6 Gephyrin Deletions in Cross-Disorder Risk for Autism, Schizophrenia and Seizures ...................................................................................................................................109

6 110

6.1 Abstract ............................................................................................................................110

6.2 Introduction ......................................................................................................................110

6.3 Results ..............................................................................................................................112

6.3.1 Detection and inheritance testing of rare microdeletions at the GPHN locus .....112

6.3.2 Frequency comparison of GPHN deletions in cases and controls .......................116

6.3.3 Clinical features of individuals with exonic GPHN deletions .............................116

6.4 Discussion ........................................................................................................................120

6.5 Materials and Methods .....................................................................................................124

6.5.1 Study subjects ......................................................................................................124

6.5.2 CNV validation and inheritance testing ...............................................................124

Chapter 7 Summary and Future Research Directions ..................................................................125

7 126

7.1 Summary and Implications ..............................................................................................126

7.1.1 Implications for clinical genetic testing and molecular diagnostics ....................127

7.1.2 Implications for genetic counseling of patients and their families ......................128

7.1.3 Implications for screening, diagnosis and treatment of NDDs ............................129

7.1.4 Implications for development of novel therapeutics ............................................130

Page 10: Copy Number Variation in Neurodevelopmental Disorders

x

7.2 Future Research Directions ..............................................................................................132

7.2.1 Follow-up of rare CNV findings in clinical genetic cohorts................................132

7.2.2 Functional characterization of rare CNVs and risk genes ...................................141

7.2.3 Higher resolution genomic studies of NDDs .......................................................149

References ....................................................................................................................................151

Page 11: Copy Number Variation in Neurodevelopmental Disorders

xi

List of Tables

Table 1.1. Examples of genomic regions implicated by rare CNVs across different NDDs 23

Table 1.2. Examples of genes implicated by rare CNVs and/or SNVs across NDDs 25

Table 3.1. Examples of risk loci implicated by rare CNVs in ASD probands 46

Table 3.2. Examples of risk loci implicated by rare exonic CNVs across 3 NDD cohorts 47

Table 4.1. De novo and rare inherited CNVs at candidate loci in ADHD probands 60

Table 4.2. Clinical phenotypes of ADHD families with rare CNVs of interest 63

Table 5.1. Very rare, clinically significant CNVs in unrelated schizophrenia probands 86

Table 5.2. Large rare CNVs of uncertain pathogenicity in unrelated probands 89

Table 5.3. Rare autosomal CNV burden in 420 unrelated schizophrenia probands 92

Table 5.4. Candidate genes for schizophrenia overlapped by very rare (<500 kb) CNVs 93

Table 6.1. Genetic and phenotypic details of probands with GPHN deletions 115

Table 7.1. Clinical diagnostic cohorts for screening of rare ASTN2/TRIM32 CNVs 135

Table 7.2. Genetic and clinical information for individuals with rare ASTN2 deletions 138

Table 7.3. Results of CNV enrichment analysis of ASTN2/TRIM32 locus 140

Page 12: Copy Number Variation in Neurodevelopmental Disorders

xii

List of Figures

Figure 1.1. Reports in literature of exonic NRXN1 deletions across a range of NDDs 28

Figure 1.2. Rationale for cross-NDD comparison of rare CNVs 31

Figure 2.1. Overview of thesis project 36

Figure 2.2. Uniform quality control measures 39

Figure 2.3. Uniform CNV analysis workflow 41

Figure 3.1. Rare exonic NRXN3 deletions in ASD probands and controls 49

Figure 3.2. Pedigrees of the families of ASD probands with NRXN3 deletions 50

Figure 4.1. Pedigrees of ADHD families with rare CNVs of interest 62

Figure 4.2. Rare CNVs at the ASTN2/TRIM32 locus in ADHD and ASD probands 68

Figure 5.1. Overview of study design and workflow 80

Figure 5.2. Functional map of schizophrenia from pathway enrichment analysis 96

Figure 5.3. Overlap of schizophrenia pathway analysis with known ASD genes 97

Figure 6.1. Genomic locations of exonic GPHN deletions in the NDD probands 113

Figure 6.2. Pedigrees of individuals with GPHN deletions 114

Figure 7.1. Exonic ASTN2/TRIM32 deletions in clinical and control cohorts 137

Figure 7.2. Impact of ASTN2 deletions on gene expression in lymphoblast cell lines 142

Figure 7.3. Exon conservation analysis of ASTN2 and TRIM32 144

Figure 7.4. Amino acid conservation analysis of ASTN2 protein 145

Figure 7.5. Relative expression levels of ASTN2 transcript isoforms 146

Figure 7.6. Expression profile of ASTN2 across human brain development 147

Page 13: Copy Number Variation in Neurodevelopmental Disorders

xiii

List of Abbreviations

ADHD Attention Deficit Hyperactivity Disorder

ASD Autism Spectrum Disorder

BAP Broader Autism Phenotype

CNV Copy Number Variant

DGV Database of Genomic Variants

DSM The Diagnostic and Statistical Manual of Mental Disorders

FISH Fluorescence In Situ Hybridization

ICD International Classification of Diseases

ID Intellectual Disability

NDD Neurodevelopmental Disorder

NGS Next Generation Sequencing

qPCR Quantitative Polymerase Chain Reaction

SNP Single Nucleotide Polymorphism

SNV Single Nucleotide Variant

WES Whole Exome Sequencing

WGS Whole Genome Sequencing

Page 14: Copy Number Variation in Neurodevelopmental Disorders

1

Chapter 1

Introduction

Parts of this chapter are adapted, with permission for use from Elsevier, from the following

published book chapter:

Marshall CR, Lionel AC, Scherer SW, Chapter 2.2 - Copy Number Variation in Autism

Spectrum Disorders, The Neuroscience of Autism Spectrum Disorders, Academic Press, San

Diego, 2013, Pages 145-154, ISBN 9780123919243

Page 15: Copy Number Variation in Neurodevelopmental Disorders

2

Psychiatric and neurodevelopmental disorders (NDDs) are among the most challenging problems

faced by modern medicine and account for nearly one-third of disability worldwide (Collins et

al., 2011; Sullivan et al., 2012a). Major NDDs such as autism spectrum disorder (ASD),

schizophrenia, intellectual disability (ID) and attention deficit hyperactivity disorder (ADHD)

cause significant suffering to patients and their families, and impose burdens on healthcare

systems and society at large. For many years, the underlying causes of NDDs appeared

intractable but research breakthroughs in psychiatric genetics over the past decade have yielded

unprecedented and long awaited insights on NDD causation (Sullivan et al., 2012a). While

scientific evidence for the genetic underpinnings of NDD etiology has accumulated over the past

century, pinpointing the precise genes responsible has proven to be extremely challenging.

Recent advances in genomics technology have facilitated a landmark accomplishment in the

quest for NDD risk genes: the discovery and characterization of copy number variants (CNVs)

(Cook and Scherer, 2008). While these deletions and duplications of specific DNA fragments are

present in every human genome as part of the natural spectrum of genetic variation, there is

substantial evidence that multiple rare pathogenic CNVs at certain specific genomic locations

contribute to medical conditions (Girirajan et al., 2011; Lee and Scherer, 2010; Weischenfeldt et

al., 2013). The study of these rare CNVs in the genomes of patients has proven to be a powerful

strategy for the rapid identification of novel NDD risk genes and has revealed etiological links

between clinically distinct NDDs (Cook and Scherer, 2008; Malhotra and Sebat, 2012;

Merikangas et al., 2009). Despite recent progress, most NDD cases are still of unknown or

idiopathic causation and the continued elucidation of their genetic architecture is imperative to

guide molecular diagnostics, clinical intervention and therapeutic development. The goal of my

doctoral thesis research is to identify novel candidate risk genes for ASD, ADHD and

schizophrenia, and to systematically investigate cross-disorder genetic overlap to pinpoint shared

risk genes. In this introductory chapter, I present an overview of the current genetic knowledge

for these three NDDs, their emerging clinical and genetic links and the rationale for a cross-

disorder CNV investigation aimed at novel NDD risk gene discovery and characterization.

Page 16: Copy Number Variation in Neurodevelopmental Disorders

3

1.1 Autism Spectrum Disorder (ASD)

1.1.1 Clinical profile of ASD

Autism is the prototypic form of a spectrum of conditions previously known as pervasive

developmental disorder (PDD), but now more widely referred to as autism spectrum disorder

(ASD). ASD is characterized by challenges in communication, impaired social function,

repetitive behaviors, and restricted interests with onset before the age of three. Signs and

symptoms can present with a range of severity and in a variety of combinations often along with

developmental delay / intellectual disability (DD/ID), sleep difficulties, seizure, and other

neurological and behavioral issues (Anagnostou et al., 2014). The Diagnostic and Statistical

Manual of Mental Disorders edition 4 (DSM-IV) defined three major subtypes within ASD

including autism, Asperger syndrome and pervasive developmental disorder not otherwise

specified (PDD-NOS). However, this sub-categorization has not been found to be linked to

replicable differences in etiology or developmental outcomes (Devlin and Scherer, 2012).

Consequently, as of May 2013, the DSM-V has subsumed and replaced the previous categorical

subgroups with a single umbrella term: “autism spectrum disorder”(Lai et al., 2013).

The latest epidemiological data estimate the population prevalence of ASD to be approximately

1 in 88 children (Autism and Developmental Disabilities Monitoring Network, 2012). Incidence

appears to be independent of ancestry and demographics, with similar rates being found on a

global scale when the same diagnostic tools are used (Fombonne, 2009). A feature of ASD is a

gender bias, with males four times more likely than females to receive a diagnosis (Werling and

Geschwind, 2013). However there is variability as it may rise to 11:1 when considering Asperger

disorder and fall to 1:1 when considering severe syndromic cases (Gillberg et al., 2006).

1.1.2 Evidence for the genetic etiology of ASD

Family studies have provided strong evidence for the contribution of complex genetic factors in

ASD etiology. Early studies showed a recurrence risk in siblings of ASD probands of 8% – 10%

(Szatmari et al., 1998; Zwaigenbaum et al., 2005) with more recent studies observing as many as

25% of siblings affected (Constantino et al., 2010; Ozonoff et al., 2011). The latter observation

implies that risk in siblings of an ASD child is roughly 20 times higher than the general

Page 17: Copy Number Variation in Neurodevelopmental Disorders

4

population risk. A recent study looking at the differential in ASD diagnosis in siblings found the

recurrence rate was approximately twice in full siblings compared to that among half siblings

(Constantino et al., 2013) further demonstrating the large genetic component to ASD etiology.

Familial studies cannot distinguish between genetic and environmental factors; however twin

studies point to a substantial genetic contribution that cannot be explained solely by the

environment. Estimates of heritability from twin studies range from 37 to 90% (Bailey et al.,

1995; Hallmayer et al., 2011; Le Couteur et al., 1996; Rosenberg et al., 2009; Steffenburg et al.,

1989). The most recent twin study showed much lower heritability (37%) compared to earlier

studies when using a strict autism cutoff, although the confidence interval was wide (8–84%)

(Hallmayer et al., 2011). When considering a broader spectrum of related cognitive or social

abnormalities, upwards of 92% of monozygotic (MZ) twins were concordant for phenotype in

contrast with ~ 10% in dizygotic (DZ) twins (Bailey et al., 1995; Constantino et al., 2010). This

difference in concordance between MZ and DZ twins suggests ASD is a genetic disorder (Risch

et al., 1999) and that the phenotype likely extends to a subclinical, broader autism phenotype

(BAP) (Losh et al., 2008).

1.1.3 Common genetic variant findings in studies of ASD risk

The role of common genetic variation (e.g. single nucleotide polymorphisms (SNPs) with minor

allele frequency > 5% in the general population) in the genetic etiology of ASD has been the

subject of considerable research and debate (Buxbaum et al., 2010; Devlin et al., 2011). There is

evidence from quantitative genetic studies in support of contributions to ASD heritability from

common variation taken together on a genome-wide scale (Klei et al., 2012). However, there

have been difficulties in pinpointing the precise locations of such contributory variants. The

standard approach for investigation of common genetic variation in disease is the genome wide

association (GWA) study. Aided by the development of SNP based genotyping microarrays,

GWA studies utilize unbiased and simultaneous frequency testing of multiple SNPs across the

genome to detect those enriched in cases relative to controls. Although highly successful in

revealing novel disease genes and biological pathways in complex conditions such as

Alzheimer’s disease and type II diabetes (Manolio, 2013; McCarthy and Hirschhorn, 2008), the

GWA approach has produced mixed results when applied to ASD. To date, there have been five

such major studies (Anney et al., 2012; Anney et al., 2010; Wang et al., 2009; Weiss et al., 2009;

Page 18: Copy Number Variation in Neurodevelopmental Disorders

5

Xia et al., 2013) and significant associations have been reported at different genomic regions

including 1p13.2, 5p14.1, 5p15.2 and 20p12.1. However, there has been no inter-study

replication as each of these findings were reported only by a single study, and targeted attempts

at validation have failed (Curran et al., 2011). In addition, the individual variant findings exhibit

very weak effect sizes and may explain only a modest part of the genetic heritability of this

disorder (Anney et al., 2012; Poot, 2013).

1.1.4 Rare genetic variant findings in studies of ASD risk

The quest for specific genes underlying ASD risk has been far more successful in the domain of

rare genetic variation (variants present at < 1% in the general population) compared with

common variation (Abrahams and Geschwind, 2008; Devlin and Scherer, 2012; Huguet et al.,

2013). The presence of ASD symptoms in rare genetic syndromes which have clear causal links

to a single genetic locus and follow simple Mendelian rules of inheritance revealed the first risk

genes for the disorder. Over 20 such syndromic genes have now been reported in connection

with autism (Abrahams et al., 2013), prominent examples of which include fragile X syndrome

(FMR1 mutations in ~ 1–2% of ASD cases), tuberous sclerosis (TSC1 and TSC2 mutations in ~

1% of ASD cases) and Rett syndrome (MECP2 mutations in ~ 0.5% of ASD cases) (Abrahams

and Geschwind, 2008). Karyotyping studies have observed microscopically visible abnormalities

in close to 5% of ASD cases and individuals with such events often exhibit dysmorphology

and/or severe ID (Shen et al., 2010; Xu et al., 2004). Although large structural changes have

been found on all chromosomes, most are so rare that association with ASD is difficult to prove.

One of the exceptions, and the most common karyotypic change found in individuals with ASD

(at a frequency of ~ 1%), is a 15q11–13 duplication (of the maternal allele) at the Prader–

Willi/Angelman syndrome region (Baker et al., 1994).

The combined observations of evidence for a strong genetic contribution to ASD from family

and twin studies, the increase in cytogenetically visible abnormalities in ASD cases compared to

controls, and the lack of consistently reproducible common variants associated with ASD led to

the hypothesis that rare submicroscopic variants in the form of CNVs contribute to the genetic

architecture of ASD. Early CNV studies using low-resolution bacterial artificial chromosome

(BAC)- CGH (Jacquemont et al., 2006) and SNP (Szatmari et al., 2007) arrays provided some of

the first evidence that CNVs contribute to ASD risk. Subsequently, a series of large studies with

Page 19: Copy Number Variation in Neurodevelopmental Disorders

6

high-resolution microarrays has further probed the role of CNVs in ASD (Bucan et al., 2009;

Christian et al., 2008; Gai et al., 2012; Glessner et al., 2009; Levy et al., 2011; Marshall et al.,

2008; Morrow et al., 2008; Pinto et al., 2010; Rosenfeld et al., 2010; Sanders et al., 2011; Sebat

et al., 2007; Weiss et al., 2008). The design of these studies varies but typically involves

computational CNV prediction from microarray intensity data, followed by rare variant detection

in the probands through filtering out CNVs seen to be present in control individuals or in

databases of copy number polymorphisms in the general population such as the Database of

Genomic Variants (Marshall and Scherer, 2012).

Microarray studies of ASD probands and their parents have identified significant enrichment of

de novo CNVs in ASD (5% - 10% of trios) compared to 1 % of control trios (Levy et al., 2011;

Marshall et al., 2008; Pinto et al., 2010; Sanders et al., 2011; Sebat et al., 2007). This is one of

the most exciting and consistently replicated findings in psychiatric genetics and it is now clear

that these highly penetrant new mutations are an important source of autism causality in

individuals who carry them (Krumm et al., 2014; Ku et al., 2013; Ronemus et al., 2014). In

addition to identifying specific genes underlying ASD etiology, de novo mutations provide a

mechanism by which such a complex, early-onset disorder with reduced reproductive fitness

remains frequent in the population (Veltman and Brunner, 2012). The incidence of de novo

mutations in spermatogenesis increases with age (Hehir-Kwa et al., 2011; Kong et al., 2012),

providing a molecular explanation for the strong epidemiological links observed between

advancing paternal age with ASD risk in children (Hultman et al., 2011). Some studies have

reported a higher rate of de novo CNVs in simplex families (only 1 child with ASD) relative to

multiplex families (two or more children with ASD) (Marshall et al., 2008; Sebat et al., 2007),

although this is not always the case (Pinto et al., 2010). Individuals with multiple de novo

variants often present with more complex and syndromic forms of ASD (Marshall et al., 2008;

Pinto et al., 2010).

Most de novo mutations, with the exception of rare instances of germline mosaicism, are

sporadic events which occur in a single individual in a family. Thus, they are insufficient on their

own to explain the high heritability of ASD evident from family studies. Indeed, rare inherited

CNVs, which often implicate the same genes as rare de novo events, are potentially risk factors

in some cases (Krumm et al., 2013; Matsunami et al., 2013; Salyakina et al., 2011). Although

many inherited CNVs act in an apparently dominant manner with varying penetrance (Fernandez

Page 20: Copy Number Variation in Neurodevelopmental Disorders

7

et al., 2010; Vaags et al., 2012), some transmission is clearly recessive in consanguineous ASD

families with rare homozygous deletions (Morrow et al., 2008). To date, upwards of 10% of

ASD probands have been seen to have rare CNVs of etiological significance. Microarray testing

is now recommended and widely used for ASD patients in clinical diagnostic laboratories in

Canada and other countries (Anagnostou et al., 2014; Miller et al., 2010a; Shen et al., 2010).

CNV analysis focused on rare variants in general, both de novo and inherited, in research and

clinical diagnostic settings has led to discovery of dozens of ASD susceptibility loci (Betancur,

2011). Some of these CNVs are large recurrent microdeletion and microduplication events, each

reported by several independent groups. These include 1q21.1 duplications (Brunetti-Pierri et al.,

2008; Szatmari et al., 2007), 7q11.23 duplications (Cooper et al., 2011; Sanders et al., 2011),

16p11.2 deletions (Kumar et al., 2008; Marshall et al., 2008; Weiss et al., 2008) and 15q13.3

deletions (Ben-Shachar et al., 2009; Sanders et al., 2011) among others (Malhotra and Sebat,

2012). CNVs of this type are created by non-allelic homologous recombination (NAHR) at

segmental duplications flanking CNV breakpoints and typically overlap several genes (Liu et al.,

2012). For some large, multi-genic microdeletions, specific genes that drive the ASD phenotype

have been pinpointed through identification of smaller deletions and point mutations at the same

locus. Examples include SHANK3 within 22q13.3 terminal deletions of Phelan-McDermid

syndrome (Durand et al., 2007; Moessner et al., 2007) and MBD5 within the 2q23.1

microdeletion syndrome (Talkowski et al., 2011).

In addition to multi-genic CNVs, genome-wide findings from high resolution microarrays have

implicated single gene deletions and duplications in ASD. Many of these mutations have

highlighted genes encoding neuronal synaptic complex proteins, and disruption of synaptic

homeostasis is now recognized as one of the key molecular mechanisms underlying ASD

(Bourgeron, 2009; Ramocki and Zoghbi, 2008; Toro et al., 2010). The first synaptic genes to be

implicated were neuroligins and neurexins, which are neuronal transmembrane proteins that bind

to each other and function as major organizers of excitatory glutamatergic synapses (Sudhof,

2008). The detection of rare functional mutations in X-linked genes for neuroligins 3 and 4

(NLGN4X and NLGN3) (Jamain et al., 2003; Laumonnier et al., 2004) was followed by discovery

of rare mutations in ASD individuals affecting all three members of the neurexin gene family:

NRXN1 (Szatmari et al., 2007), NRXN2 (Gauthier et al., 2011) and NRXN3 (Vaags et al., 2012).

Other synaptic risk gene findings from ASD CNV studies include scaffolding proteins such as

Page 21: Copy Number Variation in Neurodevelopmental Disorders

8

those from the SHANK and contactin gene families. In addition to synapse development and

function, additional biological processes highlighted by pathway analyses of genes impacted by

rare CNVs include ubiquitin degradation (Glessner et al., 2009), axon targeting and neuron

motility (Gilman et al., 2011), as well as the GTPase/Ras signaling (Pinto et al., 2010) and

TSC/SHANK (Sakai et al., 2011) networks.

Key insights into ASD risk genetics, first derived from rare CNV findings, have been reinforced

by rare SNV results from recent whole-exome (Iossifov et al., 2012; Neale et al., 2012; O'Roak

et al., 2011; O'Roak et al., 2012; Sanders et al., 2012) and whole-genome sequencing studies

(Jiang et al., 2013; Michaelson et al., 2012). These studies estimate that de novo SNVs contribute

to ASD etiology in 10% to 20% of cases and report significantly higher rates of such mutations

in ASD probands compared to either their unaffected siblings or to controls (Krumm et al., 2014;

Ronemus et al., 2014). This enrichment was particularly strong for de novo loss-of function

(nonsense and frameshift) events which resulted in prematurely truncated proteins. While no

single gene was mutated in more than 1% of cases, several genes were seen to be affected by rare

de novo SNVs across unrelated probands including SCN2A, CHD8, DYRK1A and GRIN2B. In

accordance with earlier results from CNV studies, analysis of pathways enriched for de novo

SNVs in ASD revealed postsynaptic scaffolding proteins to be particularly prominent. Other

gene-sets of note included WNT signaling, chromatin remodeling and known targets of FMR1,

the key protein in fragile X syndrome.

1.2 Attention Deficit Hyperactivity Disorder (ADHD)

1.2.1 Clinical profile of ADHD

First described as a distinct disorder in 1902 by George Still, Attention Deficit Hyperactivity

Disorder (ADHD) has a core triad of phenotypes including inattention, hyperactivity and

impulsiveness (Dalsgaard, 2013). ADHD is estimated to be prevalent in around 4% of school-

age children across the world (Polanczyk et al., 2007). Symptoms of this disorder often arise in

pre-school years before the age of seven, and can persist through adulthood in around half of

cases (Biederman et al., 2008). Skewed gender ratios (estimates range from 3:1 to 8:1) have been

reported for ADHD with boys far more likely than girls to be affected (Biederman et al., 2002).

Phenotype severity is also different between the sexes, with boys tending to be more hyperactive,

Page 22: Copy Number Variation in Neurodevelopmental Disorders

9

and girls more inattentive (Cuffe et al., 2005). Considerable clinical heterogeneity is observed

and the latest DSM version (DSM-5) recognizes four major ADHD subtypes: inattentive

presentation (restrictive), predominantly inattentive presentation, predominantly

hyperactive/impulsive presentation, and combined presentation (Dalsgaard, 2013). Longitudinal

studies spanning childhood to adolescence have shown that ADHD symptoms and subtypes often

change over time, with hyperactive symptoms more prevalent during early childhood and

attention problems more prominent during late childhood and adolescence (Elia et al., 2012b;

Larsson et al., 2011). Additional comorbidities such as depression, anxiety, learning difficulties

and other psychiatric conditions such as conduct and mood disorders are often present in

individuals with ADHD (Geissler and Lesch, 2011). They are also at heightened risk for

substance abuse, school failure, motor vehicle accidents and impaired workplace productivity

(Kessler et al., 2009). Several pharmacological treatments are available and widely used but are

hampered by varying effectiveness and the presence of side effects (Jensen et al., 2005). As a

result, considerable impairment often remains despite treatment.

1.2.2 Evidence for the genetic etiology of ADHD

Converging evidence from family, adoption and twin studies indicate a strong genetic

component underlying ADHD etiology. First degree relatives of children with ADHD are

significantly more likely to have ADHD compared with relatives of non-ADHD control

individuals (Thapar et al., 2007). Epidemiological investigations of parents (Faraone et al.,

2000), siblings (Manshadi et al., 1983) and offspring (Biederman et al., 1995) of adults with

ADHD have also confirmed the familial nature of this disorder. Studies of children with ADHD

who were adopted revealed that their biological parents were 8-10 times more likely to have

ADHD compared with the adoptive parents (Sprich et al., 2000). Cumulative findings from

dozens of pediatric twin studies report a heritability estimate of around 70% for ADHD

(Posthuma and Polderman, 2013). Adult twin studies have been fewer in number and have

generally reported lower heritability estimates of around 30% - 50% (Kan et al., 2013; Larsson et

al., 2013a). One possible reason for the discrepancy between adult and pediatric twin studies

could be that self-rating scales often used to diagnose ADHD in adults are a poorer measure of

the condition than ratings from parents and teachers that are utilized for pediatric diagnosis

(Franke et al., 2012).

Page 23: Copy Number Variation in Neurodevelopmental Disorders

10

1.2.3 Common genetic variant findings in studies of ADHD risk

Simulant based medication such as methylphenidate and amphetamine have been effective in

treating ADHD symptoms. These drugs are known to modify neurotransmission. Consequently,

the first investigations on genetic risk for ADHD focused on testing of SNPs within genes

involved in neurotransmission for association with the disorder (Gizer et al., 2009; Kebir et al.,

2009) These hypothesis based candidate gene studies have found some evidence for a role of

common variants within dopaminergic (e.g. DRD4, DRD5, DAT1) and serotonergic (e.g. 5HTT,

HTR1B) genes in ADHD (Thapar et al., 2013). However, SNPs with reported associations have

very low odds ratios (ranging from 1 to 1.3) and could potentially only account for a very small

percentage (~3%) of ADHD risk (Kuntsi et al., 2006). The candidate gene association approach

suffers from poor replication of results and other methodological issues such as potential bias

during selection of candidate genes, insufficient sample sizes, measurement errors and other

confounding factors (Kebir et al., 2009). In an effort to screen genome-wide for common variant

risk factors in an unbiased, hypothesis free manner, several GWA studies of ADHD have been

conducted (Neale et al., 2008; Neale et al., 2010a; Neale et al., 2010b; Yang et al., 2013). None

of these studies, including a large meta-analysis of close to 3,000 cases (Neale et al., 2010b),

have yielded results that pass the genome-wide threshold for statistical significance.

1.2.4 Rare genetic variant findings in studies of ADHD risk

The paucity of conclusive findings with major effects on ADHD risk from studies of common

genetic variation suggests that rare genetic risk factors could contribute to the high heritability of

the disorder. Previous reports on high incidence of ADHD traits in the phenotype spectrum of

individuals with large microdeletion syndromes also support a role for rare genetic variation. For

instance, around 30% of 22q11.2 DS patients (Arnold et al., 2001), and 70% of individuals with

the 7q11.23 microdeletion (Williams Beuren syndrome) present with ADHD symptoms (Pober,

2010). Following the success of genome-wide microarray based studies in detecting rare

pathogenic CNVs in ASD and schizophrenia, similar investigations of ADHD were launched.

There have been several studies of CNVs in pediatric ADHD cohorts (Elia et al., 2010; Elia et

al., 2012a; Jarick et al., 2014; Langley et al., 2011; Lesch et al., 2011; Lionel et al., 2011;

Stergiakouli et al., 2012; Williams et al., 2012; Williams et al., 2010; Yang et al., 2013)

Page 24: Copy Number Variation in Neurodevelopmental Disorders

11

including one from my thesis work ((Lionel et al., 2011) and Chapter 4 of this thesis) and one

CNV scan of adults with the disorder (Ramos-Quiroga et al., 2014). In four of these studies,

significantly higher numbers of rare CNVs were observed in ADHD cases compared to controls

(Stergiakouli et al., 2012; Williams et al., 2012; Williams et al., 2010; Yang et al., 2013). The

genome-wide CNV burden findings in two of these publications were primarily driven by

enrichment for large (> 500 kb) multi-genic duplications at 16p13.11 (Williams et al., 2010) and

15q13.3 (Williams et al., 2012) regions in individuals with ADHD. Although the other studies

did not find a genome-wide difference in rare CNV numbers between individuals with ADHD

and controls, rare CNV findings highlighted specific genes and biological networks. For

instance, the largest CNV study to date, performed on more than 3,500 individuals with ADHD,

noted a strong enrichment for rare CNVs at loci affecting metabotropic glutamate receptor

(GRM) genes and their interaction partners in 10% of cases, suggesting a possible role for

selective GRM agonists in ADHD treatment (Elia et al., 2012a). Results from two other studies

featuring independent cohorts of German children highlighted involvement of the parkinson

protein 2 (PARK2) and neuropeptide Y (NPY) genes (Jarick et al., 2014; Lesch et al., 2011). An

intriguing trend observed across different studies was enrichment for rare deletions and

duplications in ADHD individuals at genes previously implicated in rare CNV studies of other

NDDs such as ASD and schizophrenia (Elia et al., 2010; Lionel et al., 2011; Williams et al.,

2012; Williams et al., 2010).

Although such efforts are currently underway, to date there have been no publications describing

whole-exome or whole-genome sequencing of individuals with ADHD to discover rare SNVs.

1.3 Schizophrenia

1.3.1 Clinical profile of schizophrenia

Schizophrenia was initially identified and clinically delineated under the name “dementia

praecox” by Emil Kraepelin in 1899, based on longitudinal study of several patients exhibiting

severe decline in behavior and cognition (Jablensky, 2010). Currently, a diagnosis of

schizophrenia is based on a core triad of symptom domains (Moore et al., 2011). The first

comprises symptoms of psychoses featuring delusions and hallucinations. The second set of

symptoms, often closely associated with psychotic episodes, concerns disorganization of

Page 25: Copy Number Variation in Neurodevelopmental Disorders

12

behavior and thought processes (Moore et al., 2011). In addition to these two commonly reported

phenotypes, schizophrenia is often characterized by a third set of symptoms involving reduced

emotional expression, deficits in speech and social function and occupational deterioration

(Costain and Bassett, 2012). The evolution and severity of symptoms are highly heterogeneous

(Moore et al., 2011) and there are often shared clinical features with other psychiatric disorders

such as schizoaffective disorder, delusional disorder and psychotic bipolar disorder (Gejman et

al., 2011). The course of illness is often chronic and marked by fluctuating patterns and frequent

relapses despite modern treatments (Gejman et al., 2011). On average, the life-expectancy of

patients is reduced by 20 to 25 years (Moore et al., 2011). Thus, schizophrenia is a complex

psychiatric condition that results in a severe burden on patients and their families. The lifetime

prevalence estimate for schizophrenia is around 1%. While typical onset is during early

adulthood (17 to 30 years of age), the disorder can also arise in childhood or in the elderly age

range in a small subset of cases (Costain and Bassett, 2012). In contrast to ASD, there is no

striking gender bias in prevalence. However, the average age of onset in males is lower by

around 5 years than that in females (Mulle, 2012).

1.3.2 Evidence for the genetic etiology of schizophrenia

Consistent evidence from twin, family and adoption studies support a predominantly genetic

model of schizophrenia susceptibility. Twin studies conducted in different countries have

consistently found the concordance rate for schizophrenia in monozygotic twins (40%-50%) to

be notably higher than that of 6%-10% in dizygotic twins (Cannon et al., 1998; Cardno et al.,

1999; Franzek and Beckmann, 1998; Klaning et al., 1996). These findings support a high

heritability estimate of around 80% for schizophrenia, second only to ASD among major

psychiatric disorders (Gejman et al., 2011; Sullivan et al., 2003). Familial clustering is a

characteristic of schizophrenia, as siblings and children of individuals with schizophrenia have a

risk for the disorder, almost 10-fold times higher than the general population (Gejman et al.,

2011; Shih et al., 2004). Higher risk for psychosis was observed in foster homes for adopted

children whose biological parents had schizophrenia, but not in offspring of non-schizophrenic

biological parents who were raised by adoptive parents with psychosis (Gejman et al., 2011;

Ingraham and Kety, 2000; Kety et al., 1994).

Page 26: Copy Number Variation in Neurodevelopmental Disorders

13

1.3.3 Common genetic variant findings in studies of schizophrenia risk

Initial attempts to probe the genetic basis of schizophrenia focused on targeted case vs. control

frequency comparisons of SNPs within handpicked genes of known biological function. Such

candidate gene association studies were often prone to false positive findings which suffered

from poor reproducibility, inadequate sample size and failure to correct for multiple testing

(Smoller, 2013). The development of whole-genome microarray technology in the mid 2000’s

enabled simultaneous genome-wide interrogation of millions of SNPs. Subsequent large-scale

GWA meta-analyses on thousands of patients have highlighted the involvement of common

genetic variation (SNPs with allele frequency > 1%) at a handful of loci in risk for schizophrenia

(O'Donovan et al., 2008; Purcell et al., 2009; Shi et al., 2009; Stefansson et al., 2009). These

include genes such as ZNF804A and TCF4, and the major histocompatibility complex (MHC)

region on chromosome 6. Although there is evidence, above the genome-wide threshold for

significance, for their involvement across multiple studies, the odds ratios for the risk alleles at

each of these regions are very small. rs1344706, a SNP within ZNF804A, a gene encoding a zinc

finger protein, was the first genetic variant reported to have a significant association with

schizophrenia on a genome wide level (O'Donovan et al., 2008). This finding has been

subsequently replicated by other studies, with odds ratios consistently around 1.1 (Riley et al.,

2010; Williams et al., 2011b). The cumulative evidence to date is clear that there are no common

genetic variants of large effect size for schizophrenia (Costain and Bassett, 2012; Mulle, 2012).

1.3.4 Rare genetic variant findings in studies of schizophrenia risk

The high heritability of schizophrenia, taken together with the low number of common risk allele

findings and their small odds ratios, implies that rare genetic risk factors with much larger effect

sizes also contribute to the complex genetic architecture of this disorder (Bray et al., 2010;

Costain and Bassett, 2012; Mulle, 2012). Two such high-penetrance rare mutations have been

known for more than a decade: a translocation disrupting the DISC1 gene and the 22q11.2

deletion syndrome (22q11.2 DS). The former was discovered in a large Scottish pedigree where

it co-segregated with schizophrenia and other psychiatric illness such as bipolar disorder (Millar

et al., 2000). Recent functional work has revealed that DISC1 is an important functional hub in

several biological pathways of relevance to schizophrenia including neuronal migration,

neuronal progenitor cell proliferation and neural cell signaling (Mao et al., 2009).

Page 27: Copy Number Variation in Neurodevelopmental Disorders

14

22q11.2 DS, also known as velocardiofacial syndrome or DiGeorge syndrome, features a large

multi-genic microdeletion at the 22q11.2 region and represents the most common genomic

disorder in humans with a prevalence of 1 in 3000-4000 live births (Costain and Bassett, 2012).

These deletions have very large effect sizes and account for ~ 1% of all schizophrenia cases

since individuals with them have a 20-fold increase in risk for schizophrenia relative to the

general population (Bassett and Chow, 2008). In addition to psychoses, 22q11.2 deletions are

characterized by a heterogeneous phenotype spectrum including congenital heart defects,

seizures and ID (Bassett et al., 2011). Usually de novo in origin, 22q11.2 DS arises by NAHR of

flanking segmental duplication regions. In a small number of individuals (5% - 10%), the

deletion is inherited from a parent with a milder neuropsychiatric phenotype (Costain et al.,

2011).

In the years following discovery of the two examples mentioned above, rapid advances in SNP-

based and CGH-based microarrays have accelerated the search for other rare structural variants

involved in schizophrenia risk. The first CNV studies led to the discoveries of large rare

deletions enriched (3-fold) in schizophrenia probands relative to the general population (Walsh et

al., 2008), elevated rates of rare de novo CNVs (8-fold) in individuals with sporadic non-familial

forms of the disorder (Xu et al., 2008) and risk genes involved in glutamate signaling (Wilson et

al., 2006). Other than 22q11.2 DS, no two patients had recurrent mutations in common in these

early studies, likely due to their small sample sizes (Mulle, 2012). Subsequent studies featuring

more sizeable cohorts, including work from my thesis research ((Costain et al., 2013) and

Chapter 5), have confirmed the increased burden of large rare CNVs in patients, and have

reported on recurrent deletions and duplications at specific genomic regions (Buizer-Voskamp et

al., 2011; International Schizophrenia Consortium, 2008; Kirov et al., 2009; Rees et al., 2013;

Stefansson et al., 2008; Stewart et al., 2011; Vacic et al., 2011; Xu et al., 2009). Examples of

such CNVs that are significantly enriched in schizophrenia cases include 1q21 deletions

(International Schizophrenia Consortium, 2008; Stefansson et al., 2008), 1q21 duplications

(Levinson et al., 2011; Vacic et al., 2011), 2q13 deletions (Costain et al., 2013), 3q29 deletions

(Levinson et al., 2011; Mulle et al., 2010), 15q11.2 deletions (Kirov et al., 2009; Stefansson et

al., 2008), 15q13.3 deletions (International Schizophrenia Consortium, 2008; Levinson et al.,

2011), 16p11.2 duplications (Levinson et al., 2011; McCarthy et al., 2009), 17p12 deletions

(Kirov et al., 2009; Vacic et al., 2011) and 17q12 deletions (Moreno-De-Luca et al., 2010; Vacic

Page 28: Copy Number Variation in Neurodevelopmental Disorders

15

et al., 2011). These large multi-genic rare CNV findings show higher inter-study consistency

(Girard et al., 2012) and much larger odds ratios (Mulle, 2012) relative to GWAS results.

The initial observation (Xu et al., 2008) of a significantly higher de novo CNV rate in

schizophrenia probands relative to control individuals has also been replicated by additional trio-

based studies (Kirov et al., 2012; Malhotra et al., 2011). Strikingly, and just like the situation in

ASD, these findings highlight the presence of rare de novo CNVs in 5% - 10% of schizophrenia

cases as opposed to just 1% of control individuals. In addition to the large multi-genic CNVs

described above, de novo deletions and duplications affecting single genes have also been

identified. Recurrent de novo CNVs shared by unrelated patients affected genes encoding

synaptic proteins such as DLG2, EHMT1 and NRXN1 (Kirov et al., 2012; Rujescu et al., 2009).

Further support for the pathogenic role of de novo mutations in schizophrenia is provided by

recent exome sequencing studies. Significantly higher rates of rare de novo SNVs has been

detected in schizophrenia probands relative to controls, with especially strong enrichment for the

subset of such events which were nonsense mutations (Girard et al., 2011; Xu et al., 2012; Xu et

al., 2011). In addition to highlighting recurrent de novo SNVs across unrelated patients in genes

such as LAMA2, DPYD, TRRAP and VPS39, the sequencing results also revealed de novo SNVs

at genes previously implicated by de novo CNV findings including NRXN1, CIT and two genes

within the 22q11.2 DS region: DGCR2 and TOP3B (Awadalla et al., 2010; Xu et al., 2012).

Statistical analyses integrating significant results from the different de novo CNV, de novo SNV

and GWAS studies of schizophrenia have been performed recently (Gilman et al., 2012). Several

cohesive gene networks related to synaptic function, axon guidance, neuronal cell mobility and

chromosomal remodeling were pinpointed as key underlying molecular mechanisms of

schizophrenia (Gilman et al., 2012).

1.4 Phenotypic and Genetic Overlap Across NDDs

1.4.1 Comorbidity and clinical connections between ASD and ADHD

Historically, the major psychiatric diagnostic classification systems, including The Diagnostic

and Statistical Manual (DSM) and the International Classification of Diseases (ICD), did not

allow concurrent diagnoses of both ASD and ADHD in the same individual (Gargaro et al.,

2011). A diagnosis of ASD was considered sufficient to ‘trump’ and exclude a diagnosis of

Page 29: Copy Number Variation in Neurodevelopmental Disorders

16

ADHD (Martin et al., 2014). Although there are differences between the two conditions in terms

of core symptoms, ages of onset and recommended treatment regimens, there is substantial

evidence in support of overlapping pathophysiology for ASD and ADHD (Gargaro et al., 2011;

Murray, 2010; Russell and Pavelka, 2013; Taurines et al., 2012).

Firstly, ADHD and ASD are both childhood-onset neurodevelopmental conditions with similar

impairments in developmental and cognitive domains and a markedly higher prevalence in males

compared with females (Cooper et al., 2014). Both disorders share a range of developmental

comorbidities (Cooper et al., 2014) including intellectual disability (Ahuja et al., 2013; Matson

and Shoemaker, 2009), reading difficulties (El Zein et al., 2013; Miranda et al., 2013), dyslexia

(Russell and Pavelka, 2013), language delay (Bruce et al., 2006; Mody and Belliveau, 2013),

deficits in executive functions (Rosenthal et al., 2013; Tseng and Gau, 2013), and motor

problems such as developmental coordination disorder (Fournier et al., 2010; Martin et al.,

2010).

Secondly, co-occurrence of comorbid ASD and ADHD symptoms have been repeatedly

observed in individuals from both population-based and clinic-based studies. Several studies

have reported the presence of ADHD symptoms in autistic individuals (de Bruin et al., 2007;

Holtmann et al., 2007; Leyfer et al., 2006; Simonoff et al., 2008; Sinzig et al., 2009; Yoshida and

Uchiyama, 2004), and based on the lowest and highest reported prevalence rates, 30% – 80% of

ASD children meet criteria for ADHD (Murray, 2010; Rommelse et al., 2010). Other studies

have investigated the occurrence of autistic traits in ADHD cohorts and have found such traits to

be present in around 18% – 30% of individuals with ADHD (de Bruin et al., 2007; Grzadzinski

et al., 2011; Kochhar et al., 2011; Kotte et al., 2013; Mulligan et al., 2009; Reiersen et al., 2007).

While most studies investigating clinical overlap of ASD and ADHD have focused on pediatric

cohorts, similar trends have also been seen in adult patients (Anckarsater et al., 2006; Hofvander

et al., 2009).

Thirdly, family and twin studies suggest a heritable etiological link between ASD and ADHD

(Rommelse et al., 2010). ASD symptom levels were seen to be elevated in siblings of ADHD

patients, indicating that ASD symptoms are familial traits within ADHD families (Mulligan et

al., 2009; Nijmeijer et al., 2009; Rommelse et al., 2010). Phenotype correlation analyses suggest

that enrichment of ASD traits in brothers of ADHD probands represented shared genetic rather

Page 30: Copy Number Variation in Neurodevelopmental Disorders

17

than environmental factors (Mulligan et al., 2009; Rommelse et al., 2010). These trends were

observed even after probands and siblings with a full diagnosis of ASD were excluded from

study participation (Mulligan et al., 2009; Nijmeijer et al., 2009). Other studies examining the

correlation of ASD and ADHD symptoms in pairs of twins have found evidence for shared

genetic risk (Rommelse et al., 2010). These findings of overlap were consistently seen in both

pediatric (Lichtenstein et al., 2010; Lundstrom et al., 2011; Ronald et al., 2008; Taylor et al.,

2013) and adult twin studies (Lundstrom et al., 2011; Reiersen et al., 2008) and were robust

across different methods of phenotype assessment (Rommelse et al., 2010). Longitudinal

analysis of symptoms in twins found that early ADHD traits were predictive of later

communication difficulties and development of ASD (Taylor et al., 2013).

Fourthly, neuroimaging studies have detected underlying commonalities among brain

mechanisms implicated in ASD and ADHD. Consistent findings across both disorders include

disruption of both resting and active brain networks, especially within the frontostriatal (Gargaro

et al., 2011) and precuneus regions (Di Martino et al., 2013). Comparisons of EEG profiles of

children with a diagnosis of ADHD and additional autistic features, against profiles from ADHD

children without such features, also provide evidence for comorbidity of these disorders (Clarke

et al., 2011).

The presence of comorbid ADHD and ASD traits in patients has important implications for

prognosis and severity of psychopathology and can influence clinical treatment options.

Individuals with both ADHD and autistic traits were observed to be more severely impaired,

with higher rates of psychopathological, neuropsychological and interpersonal deficits compared

with ADHD individuals who had no autistic traits (Anckarsater et al., 2006; Kotte et al., 2013;

Rommelse et al., 2011; Sprenger et al., 2013). Individuals with traits from both disorders were

also seen to share other phenotypic characteristics such as being more likely to report with co-

morbid oppositional defiant disorder, conduct disorder, anxiety symptoms, lower full-scale IQ,

working memory deficits, language disorder and motor disorder (Cooper et al., 2014; Mulligan et

al., 2009). These findings indicate that individuals with comorbid ASD and ADHD require more

intensive medical and psychosocial support (Mulligan et al., 2009). Drugs such as

methylphenidate and atomoxetine effectively address the ADHD symptoms in individuals with

comorbid ASD and ADHD traits (Mahajan et al., 2012; Murray, 2010; Taurines et al., 2012) and

their use is now widespread (Dalsgaard et al., 2013b) although a lower daily dose is sometimes

Page 31: Copy Number Variation in Neurodevelopmental Disorders

18

required than that used for ADHD individuals without autistic symptoms (Taurines et al., 2012).

Unfortunately, these drugs do not have a substantial effect on ASD symptoms on their own

(Posey et al., 2006; Santosh et al., 2006). As a result, in patients with comorbid ASD and ADHD

traits, there is a particular need for integration of pharmaceutical interventions with behavioral,

educational and social support as well as therapies such as Cognitive Behavioral Therapy and

neurofeedback for optimal treatment of the composite symptoms (Mahajan et al., 2012;

Rommelse et al., 2011).

1.4.2 Comorbidity and clinical connections between ASD and

schizophrenia

Although currently perceived as clinically distinct disorders, schizophrenia and autism were

regarded to be part of the same psychiatric condition prior to the 1970s, with autism considered

to represent manifestation of schizophrenia at an early stage (de Lacy and King, 2013; Stone and

Iguchi, 2011). In fact, autistic behavior such as social withdrawal and developmental delay were

initially classified under “childhood schizophrenia” in the first two editions of the DSM: DSM-I

and DSM-II (Baribeau and Anagnostou, 2013). Differences in ages of onset, treatment methods

and phenotypic heterogeneity between autism and schizophrenia led to their subsequent

demarcation as distinct and non-overlapping conditions in DSM-III and DSM-IV. As a result,

autism could not be diagnosed in an individual if symptoms of schizophrenia such as psychoses,

hallucinations and delusions were also present (de Lacy and King, 2013). However, recent

epidemiological, imaging and genetic studies have highlighted biological and clinical links

between these conditions and suggest overlapping etiologies (Baribeau and Anagnostou, 2013;

de Lacy and King, 2013; King and Lord, 2011; Owen et al., 2011; Pelletier and Mittal, 2012;

Rapoport et al., 2009; Stone and Iguchi, 2011). Both conditions are recognized as

neurodevelopmental abnormalities with several overlapping deficits in motor, cognitive and

social communication domains (Barneveld et al., 2011; Eack et al., 2013; Sheitman et al., 2004;

Vannucchi et al., 2013). Consequently, the DSM-5 released in 2013 now permits concurrent

diagnosis of ASD and schizophrenia in patients (American Psychiatric Association, 2013).

Comorbidity and clinical links to ASD are particularly striking in childhood onset schizophrenia

(COS), a relatively rare form of schizophrenia. COS is characterized by identical diagnostic

criteria as adult-onset schizophrenia but with onset of symptoms before the age of 13 (Baribeau

Page 32: Copy Number Variation in Neurodevelopmental Disorders

19

and Anagnostou, 2013). COS cases often suffer from more severe symptoms and poorer

prognostic outcomes compared to individuals with adult-onset schizophrenia (Clemmensen et al.,

2012). Several studies on COS cohorts have found traits related to ASD to be co-morbid in

around 30% – 50% of individuals (Rapoport et al., 2009; Reaven et al., 2008; Sporn et al., 2004;

Waris et al., 2013). Traits related to ASD often arose 3 – 5 years before the psychotic symptoms

of COS. The unaffected siblings of children with comorbid ASD/COS symptoms were also

found to have a higher rate of ASD symptoms compared to the unaffected siblings of children

with only COS but no ASD traits (Sporn et al., 2004). These findings indicate a heritable

etiology for the comorbid ASD traits observed in COS (Rapoport et al., 2009). Neuroimaging

studies incorporating volumetric, structural and functional data have revealed some

commonalities between COS and ASD (Baribeau and Anagnostou, 2013; Rapoport et al., 2009).

Reduction of volume has been seen in both conditions, particularly in the white matter regions of

the corpus callosum and cingulum (Baribeau and Anagnostou, 2013). Exaggerated grey matter

loss during adolescence has also been reported (Baribeau and Anagnostou, 2013).

In addition to findings in COS, clinical connections to ASD are also observed in individuals with

more typical adult-onset schizophrenia. Relative to controls, individuals with a childhood

diagnosis of ASD were much more likely to experience psychotic symptoms in adolescence

(Bevan Jones et al., 2012; Sullivan et al., 2013) and adulthood (Joshi et al., 2013; Unenge

Hallerback et al., 2012). A longitudinal study over a 30-year timespan found that 35% of

individuals presenting with autism in childhood were later diagnosed with schizophrenia or

schizophrenia spectrum disorder in adulthood (Mouridsen et al., 2008). The overlap between

these conditions has important clinical consequences since the presence of childhood autistic

traits appears to correlate with more severe schizophrenia symptoms and negative long-term

outcomes in adult patients (Esterberg et al., 2008; Vannucchi et al., 2013). Results from family

studies have further underscored the etiologic link between ASD and schizophrenia. The

presence of schizophrenia in parents was correlated with an elevated risk for ASD in their

children in different Swedish cohorts (Sullivan et al., 2012b). The same study also found

evidence for higher risk for ASD in children with a schizophrenic sibling (in independent cohorts

from Sweden and Israel) (Sullivan et al., 2012b). Other large population studies have found

schizophrenia to be more common in parents of children with autism compared to parents of

control children (Daniels et al., 2008; Jokiranta et al., 2013; Larsson et al., 2005).

Page 33: Copy Number Variation in Neurodevelopmental Disorders

20

1.4.3 Comorbidity and clinical connections between ADHD and

schizophrenia

Retrospective investigations have discovered ADHD symptoms to be prevalent in around 15 –

17% of individuals with schizophrenia (Gomez et al., 1981; Peralta et al., 2011). This rate is

higher than the 5% prevalence of ADHD in the general population (Polanczyk et al., 2007).

ADHD was also the most prevalent childhood-onset neurodevelopmental condition detected in a

case series of teens at high risk for psychosis (Mazzoni et al., 2009). Individuals with comorbid

ADHD and schizophrenia had increased rates of suicidal behavior and worse performance in

neuropsychological tests relative to individuals with a single condition (Donev et al., 2011).

Further highlighting the links between these disorders, two Spanish studies found ADHD to be

the most prevalent neurodevelopmental phenotype observed in first degree relatives (siblings and

offspring) of individuals with schizophrenia (de la Serna et al., 2011; de la Serna et al., 2010).

Complementing the aforementioned work on ADHD symptoms in individuals with

schizophrenia and their relatives, other studies have examined the risk for developing

schizophrenia in individuals with ADHD. Children with ADHD are 4 times more likely to

develop schizophrenia (Dalsgaard et al., 2013a; Kim-Cohen et al., 2003) or report a psychotic

episode (Biederman et al., 2006) in adulthood. A large Swedish epidemiological study of more

than 60,000 individuals with ADHD reported substantially increased risk for both schizophrenia

(6-fold increase) and bipolar disorder (24-fold) in adulthood among individuals with a childhood

ADHD diagnosis (Larsson et al., 2013b). This increased risk for psychosis was also seen in first

degree relatives of the ADHD proband group but not in second degree relatives, suggesting that

co-occurrence of ADHD and schizophrenia could be due to shared genetic risk factors.

1.4.4 Evidence for shared genetic risk factors across NDDs

Consistent with the comorbidity and overlapping clinical traits often observed across different

NDDs, there is emerging evidence for the contribution of shared genetic risk factors to etiologies

of these conditions. This intriguing cross-disorder genetic overlap has been noted for both

common and rare forms of genetic variation. In an effort to investigate shared common genetic

variation, the Cross-Disorder Group of the Psychiatric Genomics Consortium (PGC) conducted

a GWAS meta-analysis of genome wide SNP data from 33,332 cases across five major

Page 34: Copy Number Variation in Neurodevelopmental Disorders

21

psychiatric disorders (ADHD, ASD, bipolar disorder, major depressive disorder and

schizophrenia) and 27,888 ancestry-matched controls (Cross-Disorder Group of the Psychiatric

Genomics Consortium, 2013). This was the largest genome-wide analysis of psychiatric illness

to date and provided the first key evidence that individual and aggregate molecular genetic risk

factors are shared between childhood-onset and adult-onset psychiatric disorders that are treated

as distinct categories in clinical practice (Cross-Disorder Group of the Psychiatric Genomics

Consortium, 2013; Smoller, 2013). Four SNPs exceeded the threshold for genome-wide

significance, three of which were involved in shared genetic risk across all five disorders. These

SNPs were within ITIH3 (3p21), AS3MT (10q24) and CACNB2 (10p12). The fourth SNP, within

CACNA1C (12p13.3), contributed to risk for schizophrenia and bipolar disorder. The presence of

two calcium channel genes (CACNA1C and CACNB2) within the top hits and a subsequent

pathway analysis highlighted voltage-gated calcium channel signaling as a key biological

contributor to psychopathology and a potential therapeutic target. The study also revealed other

intriguing shared risk loci for ASD and schizophrenia such as the microRNA mir137 (1p21.3)

and the genes TCF4 (18q21.1) and PCGEM1 (2q32). Earlier GWAS studies with smaller cohorts

have also found evidence for common alleles conferring risk across traditional diagnostic

boundaries in psychiatry (Hamshere et al., 2013; Williams et al., 2011a).

The overlap of genetic risk factors across neuropsychiatric disorders has been especially striking

among rare CNV findings (Carroll and Owen, 2009; Guilmatre et al., 2009; Malhotra and Sebat,

2012; Pescosolido et al., 2013; Rosenfeld et al., 2010; Sahoo et al., 2011). The first such

discoveries involved rare microdeletions and microduplications of sizes large enough (> 500 kb)

to be detected by early low-resolution microarray analyses. These genomic events are often

characterized by recurrent breakpoints determined by flanking segmental duplications, which

mediate CNV formation via non-allelic homologous recombination. In-depth clinical

characterization of individuals with microdeletions and microduplications has established that

several of these events are genetic syndromes, associated with certain core characteristics amidst

a heterogeneous spectrum of NDD phenotypes (Table 1.1). For example, 16p11.2 microdeletions

are the second most prevalent pathogenic CNV (after 22q11.2 microdeletions) in NDD

individuals referred for clinical microarray testing (Cooper et al., 2011; Kaminsky et al., 2011;

Moreno-De-Luca et al., 2013). Initially discovered in ASD cohorts (Kumar et al., 2008; Marshall

et al., 2008; Weiss et al., 2008), follow-up studies have revealed a trend for macrocephaly and

Page 35: Copy Number Variation in Neurodevelopmental Disorders

22

obesity in individuals with this microdeletion, in addition to a varied assortment of traits such as

seizures, speech and motor delays and congenital anomalies (Bochukova et al., 2010; Fernandez

et al., 2010; Jacquemont et al., 2011; Moreno-De-Luca et al., 2013; Walters et al., 2010).

Individuals with the reciprocal 16p11.2 microduplication are at higher risk for schizophrenia,

microcephaly and low BMI (Jacquemont et al., 2011; McCarthy et al., 2009). A similar scenario

of certain mirror phenotypes between deletions and duplications amidst considerable phenotypic

heterogeneity is observed at the 1q21.1 locus where microdeletions and microduplications

present with microcephaly and macrocephaly respectively, together with developmental delay,

congenital anomalies and facial dysmorphisms in both syndromes (Brunetti-Pierri et al., 2008).

Page 36: Copy Number Variation in Neurodevelopmental Disorders

23

Table 1.1. Examples of genomic regions implicated by rare CNVs across different NDDs

Genomic segment Disorders References

1q21.1 deletion ASD, EPY, ID, SZ Brunetti-Pierri 2008, Mefford 2008, Stefansson 2008,

Kirov 2009, ISC 2008, Levinson 2011

1q21.1 duplication ASD, SZ Brunetti-Pierri 2008, Costain 2013

2q13 deletion ADHD, ASD, ID, SZ Yu 2012, Costain 2013

2q13 duplication ASD, SZ Yu 2012, Costain 2013

2q21.1 deletion ADHD, ID, EPY Dharmadhikari 2012

2q21.1 duplication ASD, ADHD, EPY Dharmadhikari 2012

3q29 deletion ASD, ID, SZ Willatt 2005, Mulle 2010, Levinson 2011

7q11.23 deletion ASD, EPY, ID Fusco 2014

7q11.23 duplication ADHD, ASD, SZ Dixit 2013,Sanders 2011, Kirov 2012, Mulle 2013

15q11.2 deletion ASD, EPY, ID, SZ De Kovel 2010, Cooper 2011, Kirov 2009, Levinson

2011

15q11-q13 duplication ADHD, ASD, SZ Pinto 2010, Lionel 2011, Costain 2013

15q13.3 deletion ASD, SZ Ben-Shachar 2009, Costain 2013

16p13.11 deletion ID, SZ Cooper 2011, Ingason 2011, Mefford 2009

16p13.11 duplication ADHD, SZ Williams 2010, Costain 2013

16p11.2 deletion ASD, SZ Weiss 2008, Marshall 2008, Pinto 2010, Costain 2013

16p11.2 duplication ADHD, ASD, SZ Pinto 2010, Lionel 2011, McCarthy 2009

17q12 deletion ASD, EPY, ID, SZ Moreno-De-Luca 2010, Cooper 2011, Kaminsky 2011

22q11.2 deletion ASD, SZ Pinto 2010, Bassett 2008

22q11.2 duplication ASD, ID Lo-Castro 2009, Pinto 2010

Abbreviations: EPY, epilepsy; ID, Intellectual Disability; SZ, schizophrenia.

Page 37: Copy Number Variation in Neurodevelopmental Disorders

24

Since each of the genomic microduplications and microdeletions include several genes, it is not

clear if the spectrum of neurocognitive symptoms associated with them reflects pleiotropic

effects of individual genes or is a result of aberrant dosage of multiple genes (Pescosolido et al.,

2013). Cross-NDD genetic overlap has also been consistently observed in more recent studies

with higher resolution microarrays and sequencing to identify smaller CNVs and SNVs affecting

single genes (Table 1.2). These findings provide gene-level specificity and have been particularly

informative for investigations of the biological pathways involved in etiology of different NDDs.

Such efforts have indeed found evidence for shared underlying molecular mechanisms across

diverse conditions. Independent analyses of functional gene networks impacted by rare CNVs in

ASD (Gilman et al., 2011) and schizophrenia (Gilman et al., 2012) have revealed significant

overlap with convergence around proteins involved in synaptic function. Cristino et al. (2013)

performed a systematic, genome-wide investigation of candidate genes reported for ASD,

schizophrenia, ADHD and ID (from CNV, SNP and sequencing studies). They found evidence

for co-occurrence of genetic factors implicated in these different disorders in the same molecular

pathways and functional domains, with synaptic transmission as a prominent example of such

overlap (Cristino et al., 2013).

Page 38: Copy Number Variation in Neurodevelopmental Disorders

25

Table 1.2. Examples of genes implicated by rare CNVs and/or SNVs across different NDDs

Gene Disorders References

ANK3 ADHD, ASD, ID Bi 2012, Sanders 2012, Iqbal 2013

ARHGEF9 EPY, ID Kalscheuer 2009, Lesca 2011

ARID1B ASD, ID Halgren 2012

ASTN2/TRIM32 ADHD, ASD,

BPD, EPY, ID, SZ

Vrijenhoek 2008, Glessner 2009, Grozeva 2010, Lionel 2011, Van-

Silfhout 2013, Lionel 2014

AUTS2 ADHD, ASD,

EPY, ID Elia 2009, Beunders 2013

BCORL1 ASD, ID, SZ Xu 2012, Schuurs-Hoeijmakers 2013

BIRC6 ASD, SZ Xu 2012

C2CD3 ASD, SZ Xu 2012

CNTNAP2 ADHD, ASD,

EPY, SZ Bakkaloglu 2008, Friedman 2008, Elia 2009, Poot 2010

CSMD3 ASD, SZ Floris 2008, Malhotra 2011

DISC1 ASD, BPD, SZ Millar 2000, Williams 2009, Costain 2013

DLGAP2 ASD, SZ Marshall 2008, Guilmatre 2009, Pinto 2010

DPP6 ADHD, ASD, SZ Marshall 2008, Xu 2008, Elia 2009, Pinto 2010, Liao 2013

DPYD/mir137 ASD, ID, SZ Willemsen 2009, Carter 2011, Xu 2011, Van-Silfhout 2013

EHMT1 ASD, ID, SZ Kleefstra 2006, Kirov 2012, Talkowski 2012

FOXP1 ASD, ID Hamdan 2010, Girirajan 2011, Talkowski 2012

GABRG1 ADHD, ASD Vincent 2006, Lionel 2011

GPHN ASD, EPY, SZ ISC 2008, Lionel 2013

GPRIN3 ASD, SZ Xu 2012

GRIK2 ASD, ID Motazacker 2007

GRIN2B ASD, BPD, ID, SZ Awadalla 2010, Endele 2010, Talkowski 2012

GRM7 ADHD, SZ Walsh 2008, Elia 2009, Costain 2013

H2AFV ASD, SZ Xu 2012

Page 39: Copy Number Variation in Neurodevelopmental Disorders

26

HECTD1 ASD, SZ Xu 2012

ILRAPL1 ASD, ID Piton 2008, Pinto 2010

IMMP2L ADHD, ASD Elia 2009

MACROD2/FLRT3 ADHD, SZ Lionel 2011, Costain 2013

MBD5 ASD, BPD, EPY,

ID Talkowski 2011, Hodge 2013

MEF2C ASD, EPY, SZ Le Meur 2009

MYH10 ASD, SZ Xu 2012

NRXN1 ADHD, ASD,

EPY, ID, SZ Szatmari 2007, Rujescu 2009, Bradley 2010, Duong 2012

PARK2 ASD, ADHD, SZ Elia 2009, Jarick 2012, Glessner 2009, Vacic 2011

PIK3C3 ID, SZ Costain 2013, Van-Silfhout 2013

PTCHD1/PTCHD1AS ADHD, ASD, ID Noor 2010, Filges 2010, Lionel 2011

SCN2A ASD, EPY, ID Ogiwara 2009, Sanders 2012, Jiang 2013

SHANK2 ASD, ID Berkel 2010, Leblond 2012

SHANK3 ASD, EPY, ID, SZ Durand 2007, Moessner 2007, Gauthier 2010, Lesca 2012

SMCHD1 ASD, SZ Xu 2012

SYN1 ASD, EPY Garcia 2004, Fassio 2011

SYNGAP1 ASD, EPY, ID Carvill 2013, Hamdan 2009, Pinto 2010

TCF4 ASD, EPY, ID, SZ Amiel 2007, Talkowski 2012, Van-Silfhout 2013

TOP3B ASD, SZ Xu 2012

VPS39 ASD, SZ Xu 2012

ZNF804A ASD, SZ Steinberg 2011, Griswold 2012, Talkowski 2012, Costain 2013

Abbreviations: EPY, epilepsy; ID, Intellectual Disability; SZ, schizophrenia.

Page 40: Copy Number Variation in Neurodevelopmental Disorders

27

An excellent example of a gene involved in shared risk for different NDDs is NRXN1, which

encodes the neurexin 1 protein, with critically important roles in synaptic cell adhesion and

neurotransmitter secretion (Missler et al., 2003). The first observation of NXRN1 mutations in

ASD was a 300 kb hemizygous de novo deletion that eliminated several NRXN1 exons an ASD

individual (Szatmari et al., 2007). Subsequent studies have found exonic deletions in several

ASD cases (Ching et al., 2010; Glessner et al., 2009; Kim et al., 2008; Pinto et al., 2010; Sanders

et al., 2011; Shen et al., 2010; Zahir et al., 2008). Independent studies have also reported rare de

novo and inherited deletions affecting exons of NRXN1 in cases with schizophrenia (International

Schizophrenia Consortium, 2008; Kirov et al., 2009; Kirov et al., 2008; Rujescu et al., 2009;

Vrijenhoek et al., 2008; Walsh et al., 2008), intellectual disability (Ching et al., 2010; Friedman

et al., 2006; Guilmatre et al., 2009), bipolar disorder (Zhang et al., 2009), ADHD (Bradley et al.,

2010), Tourette syndrome (Sundaram et al., 2010) and autosomal-recessive Pitt-Hopkins

syndrome (Zweier et al., 2009). On the other hand, such deletions are absent or extremely rare in

control individuals (Ching et al., 2010). However the association of NRXN1 deletions with

disease risk often does not reach statistical significance within individual studies due to small

sample numbers and rarity of the deletions. It is only by performing a meta-analysis across

several studies (Figure 1.1) that we observe a highly significant enrichment of exonic NRXN1

deletions (p = 2.4 x 10-18) in cases versus controls. Nrxn1 knockout mice exhibit cognitive

impairment, behavioral changes and aberrant synaptic function, providing additional evidence

for a causal link between NRXN1 and NDD etiology (Etherton et al., 2009).

Page 41: Copy Number Variation in Neurodevelopmental Disorders

28

Figure 1.1. Reports in literature of exonic NRXN1 deletions across a range of NDDs

Overview of rare CNVs detected at 2p16.3 locus, overlapping NRXN1 by different published CNV studies of NDDs. The blue and red bars denote deletions and duplications, respectively. Genomic coordinates and information about transcript isoforms are from Genome Build 36 (hg18).

Page 42: Copy Number Variation in Neurodevelopmental Disorders

29

The overlapping CNV findings described above were derived from incidental observations and

the overall trend of cross-disorder overlap became apparent upon comparing results from

independent CNV scans of different NDDs. Very few studies have systematically and uniformly

assessed different NDD cohorts simultaneously for the presence of CNVs at overlapping risk loci

(Guilmatre et al., 2009; Moreno-De-Luca et al., 2010). In one such effort, Guilmatre et al.

investigated microarray data from 743 individuals with NDDs (247 with ID, 260 with ASD and

236 with schizophrenia) and 236 control individuals for shared risk CNVs. This study was not a

genome-wide microarray scan but rather used a targeted approach (quantitative multiplex PCR of

short fluorescent fragments) restricted to 28 specific candidate loci. These regions were selected

based on previous implication by rare deletions or duplications in individuals with NDDs

including NRXN1, CNTNAP2, SHANK3 and the 16p11.2 and 15q13 regions among others.

Guilmatre et al. observed overlapping CNVs in cases at 39.3% of the 28 selected loci. These risk

CNVs were detected in 4.2% of the schizophrenia cases, 6.2% of the ASD cases and 5.3% of the

ID cases. In contrast, only 0.4% of control individuals were seen to have CNVs at the risk loci,

thus demonstrating a significant excess of CNVs in each disorder cohort compared with controls

(p =.01 for schizophrenia, p < .001 for ASD and p = .001 for ID). Several CNVs in NDD cases

for whom parental DNA was available were seen to be of de novo origin, and overlapped genes

with roles in neurotransmission or in synapse formation and maintenance. The authors concluded

that their observations provided additional support for a causative or contributory role of rare,

and often de novo, CNVs to NDD risk and highlighted the existence of shared biological

pathways underlying etiology of clinically distinct NDDs.

1.5 Thesis Rationale and Overview

The composite genetic picture of NDD risk emerging from the myriad human genetics

investigations over the past decade is highly complex (Sullivan et al., 2012a). Conditions such as

ASD, schizophrenia, ADHD exhibit high heritability and clearly possess substantial genetic

underpinnings. The general failure of common variant studies in identifying specific risk genes

of large effect sizes and consistent reproducibility has led to a paradigm shift in thinking from

the older “common disorder-common variant” model. The new “common disorder-rare variant”

model has been bolstered by numerous rare variant studies investigating CNVs and SNVs. The

exciting results from these studies represent considerable progress in defining specific risk genes

Page 43: Copy Number Variation in Neurodevelopmental Disorders

30

and pathways. However, these findings also highlight challenges going forward. Although

several genomic loci have been consistently implicated in NDD risk, no single locus accounts for

more than 1% of cases and it is now evident that hundreds of risk loci are likely involved, each

responsible for a small subset of individuals with the disorder. The majority of cases still lack a

known genetic cause and the continued cataloguing of bona fide risk loci is essential to expand

the scope of molecular diagnostics, enable more effective clinical intervention and facilitate

therapeutic development The intriguing trend of shared risk genes for clinically distinct but often

overlapping NDDs offers both insight into common causal mechanisms and a promising research

avenue for identifying novel candidate genes.

Prior to the work described in this thesis, no unbiased, simultaneous, genome-wide comparisons

of rare CNVs across different NDDs have been reported. Combining CNV data from multiple

disorder cohorts in such a manner has the advantage of providing increased statistical power

necessary for detection and interpretation of very rare genetic events (Figure 1.2). Such a

strategy has the potential to reveal novel candidate genes contributing to cross-disorder risk for a

spectrum of NDDs and thus explain the genetic basis of shared clinical features often observed

across conditions. For instance, if a rare deletion of a gene is observed in a single individual from

a large cohort of ASD patients but is absent in control individuals, it is difficult to ascertain

whether this CNV is connected to ASD etiology or is merely a benign rare variant. Using a

cross-order analysis, if the same gene is also found to be affected by CNVs and/or point

mutations in unrelated individuals from other NDD cohorts such as schizophrenia or ADHD, this

provides additional support for a likely contributory role for this gene to NDD risk. Follow-up

in-depth phenotyping of individuals with such rare mutations affecting a particular gene can

inform genotype-phenotype correlations and enable detection of subtle shared clinical features

among these individuals.

Page 44: Copy Number Variation in Neurodevelopmental Disorders

31

Figure 1.2. Rationale for cross-NDD comparison of rare CNVs

Page 45: Copy Number Variation in Neurodevelopmental Disorders

32

In Chapter 2, I describe the overview of my thesis project, which involves genome-wide CNV

analysis of three newly characterized cohorts of Canadian individuals with one of three different

NDDs: ASD, ADHD and schizophrenia. My first objective, using intra-cohort analyses, was to

identify new genetic risk loci for each disorder by focusing on de novo CNVs and rare CNVs

significantly enriched in multiple unrelated patients compared to controls. My second objective

was to perform a cross-cohort analysis to detect overlapping rare CNVs among the three

disorders and pinpoint shared risk genes. The CNV calling methodology used for uniform

analysis of the different case cohorts is also described.

In Chapter 3, I present results of a genome wide CNV analysis for the ASD cohort. As an

illustration of my first objective, I outline the discovery of rare de novo and inherited exonic

deletions affecting the NRXN3 gene in multiple unrelated individuals with ASD. This gene, at

chromosome 14q24, encodes a key synaptic structural protein belonging to the neurexin protein

family, whose other members such as NRXN1 and NRXN2 are well-established ASD risk

factors. These deletions exhibited location-dependent penetrance effects, with those overlapping

multiple isoforms of the gene resulting in a more severe phenotype than those affecting a single

isoform. Follow-up characterization of individuals with NRXN3 deletions revealed clinical

complexity including a heterogeneous spectrum of ASD-related phenotypes and variable

expressivity.

In Chapter 4, I present results of a genome-wide CNV analysis for the ADHD cohort. The

availability of parental DNA for these individuals was used to investigate de novo CNV rate in

ADHD for the first time. This was found to be lower than published rates for ASD and

schizophrenia, and individuals with de novo CNVs had complex phenotypes featuring other

neurodevelopmental comorbidities in addition to their ADHD symptoms. Rare inherited CNVs

were observed in 19 of 248 (7.7%) individuals with ADHD, which either overlapped previously

implicated ADHD loci (e.g. DRD5 and 15q13 microduplication) or identified new candidate

susceptibility genes (e.g. MACROD2/FLRT3, CPLX2 and PTPRN2). Genome-wide rare CNV

findings in the ADHD cohort were compared with those from the ASD cohort described in

Chapter 3. Deletions of neuronal ASTN2 and ASTN2-intronic TRIM32 genes at chromosome

9q33.1 yielded the strongest association with ADHD and ASD, but numerous other shared

candidate genes (such as CHCHD3, MACROD2, and the 16p11.2 region) were also revealed.

Page 46: Copy Number Variation in Neurodevelopmental Disorders

33

In Chapter 5, I present results of a genome-wide CNV analysis on the schizophrenia cohort. Rare

CNVs greater than 500 kb in size were highly enriched in individuals with schizophrenia as

compared to a matched control cohort. CNVs at chromosome 2q13, which had been previously

implicated in ASD, were found to be significantly associated with schizophrenia for the first

time. Several new candidate genes for schizophrenia were revealed by studying overlapping rare

CNVs in multiple unrelated individuals. Functional gene-mapping analyses revealed significant

enrichment of exonic deletions in the schizophrenia cohort that impacted biological pathways

previously highlighted in ASD literature including neurodevelopmental and synaptic processes.

In Chapter 6, as an illustration of my second objective, I describe the discovery of exonic GPHN

deletions in individuals from ASD and schizophrenia cohorts. GPHN, at chromosome 14q23.3,

codes for gephyrin, a key scaffolding protein in the neuronal postsynaptic membrane, responsible

for clustering and localization of glycine and GABA receptors at inhibitory synapses. Follow-up

screening of a large clinical microarray dataset revealed additional individuals with GPHN

deletions, who presented with a range of neurodevelopmental diagnoses including ASD,

schizophrenia or seizures. Inheritance testing revealed that the majority of such deletions arose

de novo. All of the deletions shared a common region of overlap affecting a functionally

important G-domain of the gephyrin protein.

In Chapter 7, I summarize my key findings and discuss their implications for genetic screening

and counseling, psychiatric diagnosis and drug development. I conclude with a discussion of

possible extensions and future directions for this research. These include follow-up screening of

candidate genes in clinical microarray datasets; in-depth functional characterization of candidate

genes and molecular mechanisms of disease development; and the higher resolution

investigations of NDDs through whole-genome sequencing approaches.

Page 47: Copy Number Variation in Neurodevelopmental Disorders

34

Chapter 2

Thesis Overview and Methodology

Page 48: Copy Number Variation in Neurodevelopmental Disorders

35

2.1 Thesis Overview

An overview of my thesis project is presented in Figure 2.1. The three patient cohorts used for

the genome-wide CNV analysis were recruited by my clinical collaborators from different sites

across Canada. The phenotyping protocols used and the composition of the patient cohorts are

described in Chapter 3 (ASD cohort), Chapter 4 (ADHD cohort) and Chapter 5 (schizophrenia

cohort). The microarray genotyping, CNV analysis and prioritization of rare CNVs were

conducted in a uniform manner for all patients as described in section 2.2. My first objective was

to discover new candidate risk genes within each disorder dataset. For this purpose, I prioritized

those rare CNVs that affected recurrent genetic regions across multiple unrelated patients within

a cohort, or affected genes previously implicated in literature for NDD risk. In order to address

my second objective of discovering genes involved in shared genetic risk across different NDDs,

I compared the rare CNV findings from all three cohorts. Genes at which rare CNVs were

detected in 2 or more of the cohorts are listed in section 2.3. Follow-up replication studies and

in-depth clinical characterization were performed for two of the novel cross-NDD risk genes

discovered: ASTN2 (Chapter 3 and Chapter 7) and GPHN (Chapter 6).

Page 49: Copy Number Variation in Neurodevelopmental Disorders

36

Figure 2.1. Overview of thesis project

Abbreviations used: ASD, Autism Spectrum Disorder; ADHD, Attention Deficit Hyperactivity Disorder; SCZ, schizophrenia

Page 50: Copy Number Variation in Neurodevelopmental Disorders

37

2.2 Uniform CNV Analysis Workflow

2.2.1 Microarray genotyping

DNA samples from the patients across all 3 cohorts were sent from clinical sites to The Centre

for Applied Genomics (TCAG), Toronto. Samples were genotyped on the Affymetrix Genome-

Wide Human SNP Array 6.0 at the TCAG microarray facility with standard protocols as

provided by the manufacturer. The completion of genotyping experiments for all study samples

at the same facility using identical protocols was designed to minimize laboratory related

artefacts. The Affymetrix 6.0 is a widely used microarray platform that features comprehensive

coverage of the genome with more than 1.8 million probes (McCarroll et al., 2008). Half of these

are at known SNPs (SNP probes) and provide genotype data in addition to intensity data while

the other half are copy number specific probes (CN probes) which provide only intensity data

(McCarroll et al., 2008). The availability of both genotype and CNV data from this microarray is

particularly useful for the different quality control measures described in the next section.

2.2.2 Uniform quality control measures

To ensure high quality microarray data for CNV analysis, different quality control measures

were performed (Figure 2.2). In accordance with recommended quality control guidelines by

Affymetrix, microarrays which did not have values of the contrast QC metric > 0.4 were

excluded from analysis. The contrast QC metric measures the quality of each microarray

experiment by assessing uniformity of probe hybridization and testing for major defects in the

microarray chip. In order to reduce error due to sample mix-up prior to genotyping, the gender of

the individual genotyped on each array was determined using intensities from X and Y

chromosome probes. Those experiments for which the gender indicated by the microarray

contradicted the known gender of the individual were discarded. In order to minimize CNV

artefacts due to batch effects, suitable reference batches for CNV analysis were selected based on

previously validated approaches from the Scherer laboratory (Pinto et al., 2011). Briefly, pair-

wise Pearson correlations of median normalized intensities were calculated between different

arrays to pick suitable analysis batches with Pearson correlation values > 0.88 (Pinto et al.,

2011).

Page 51: Copy Number Variation in Neurodevelopmental Disorders

38

For individuals with parental DNA available for microarray genotyping, as was the case for

some complete trios in the ADHD and ASD cohorts, the PLINK tool set (Purcell et al., 2007)

was used to compute Mendelian error rate of genotypes from SNP probes on the Affymetrix 6.0

platform. Those trios in which Mendelian errors were observed for more than 1% of SNP probes

were considered to have potential issues with non-paternity or sample mix-ups and were

excluded from analysis. The PLINK toolset was also used to systematically check the degree of

genetic relatedness between cases to ensure that recurrent CNV findings across multiple

individuals were not due to hidden relatedness. Pair-wise identity by descent (IBD) was

calculated for every pair of cases from genotypes of microarray SNP probes on the Affymetrix

6.0 platform. Those pairs of samples with PI_HAT values [defined as P(IBD = 2) + 0.5 × P(IBD

= 1)] greater than 0.1, corresponding to first cousins or closer, were further investigated for

relatedness and only one sample from each pair was used in the final dataset. To accurately

estimate ancestry for individuals in patient cohorts, genotypes of 1,120 genome-wide unlinked

SNPs were clustered by the program STRUCTURE (Pritchard et al., 2000) together with

genotypes from 270 HapMap samples, which were used as references of known ancestry during

clustering. Samples were assigned to one of the following three: “European”, “East Asian”,

“African” using a threshold of coefficient of ancestry exceeding 0.9.

Page 52: Copy Number Variation in Neurodevelopmental Disorders

39

Figure 2.2. Uniform quality control measures

The different quality control filters used, together with their associated metrics and thresholds, are depicted. As an illustration of the relative impact of these sequential filters during data cleaning, the number of arrays excluded from analysis in the ADHD project (Chapter 4) were 6 with Contrast QC < 0.4, 3 with gender mismatches and 6 (2 trios) due to high Mendelian errors. A total of 594 arrays (248 from probands and 346 from parents) passed the quality control filters and were used for further analysis in the ADHD project (Chapter 4).

Page 53: Copy Number Variation in Neurodevelopmental Disorders

40

2.2.3 CNV detection from microarray data

Microarray data from all individuals in the clinical cohorts were analyzed for CNVs on a

genome-wide scale using a uniform workflow (Figure 2.3). To maximize sensitivity and

specificity of CNV detection, three different algorithms were used to call CNVs from the

microrray data: Birdsuite (Korn et al., 2008), iPattern (Pinto et al., 2011), and the Affymetrix

Genotyping Console. For each of these algorithms, CNVs were required to be supported by five

or more consecutive array probes. Subsequent analyses focused on “stringent” CNV calls,

defined as those CNVs detected by at least two of three algorithms. All subsequent analyses

focused on these stringent CNVs, which have very high positive validation rates (> 90%) by

independent experimental methods such as quantitative PCR (qPCR) and fluorescence in situ

hybridization (FISH) (Costain et al., 2013; Lionel et al., 2011; Marshall et al., 2008; Pinto et al.,

2010; Prasad et al., 2012; Silversides et al., 2012). Within CNV calls from each individual,

overlapping calls from Birdsuite and iPattern were merged with the outside probe boundaries and

designated as stringent calls. Singleton calls from Birdsuite or iPattern, which overlapped with a

Genotyping Console call from the same individual, were also included in the stringent set.

Samples with number of calls greater than three times the SD from the mean number of calls for

an analysis batch were excluded. Rare CNVs were defined for individuals from the three clinical

cohorts by comparing the total dataset of stringent CNVs against those CNVs identified in two

large population-based control cohorts comprising 2357 individuals of European ancestry from

Ontario (Stewart et al., 2009) and Germany (Krawczak et al., 2006). CNVs were identified in

these control individuals using an identical microarray platform and CNV analysis strategy as

described above for the clinical cohorts. A 50% reciprocal overlap criterion was used; i.e. a case

CNV that was at least 50% unique by length when compared to every CNV in controls was taken

to be rare. In order to identify novel NDD risk genes, those rare CNVs that affected the same

gene across multiple individuals were prioritized for follow-up after confirmation of microarray

calls as true positives. When parental samples were available, as was the case for some of the

probands in the ADHD (Chapter 4) and ASD (Chapter 3), inheritance testing was performed in

order to detect de novo CNVs, which were also prioritized for risk gene discovery. Experimental

validation and inheritance testing of rare CNVs was performed using SYBR Green based qPCR.

Page 54: Copy Number Variation in Neurodevelopmental Disorders

41

Figure 2.3. Uniform CNV analysis workflow

Page 55: Copy Number Variation in Neurodevelopmental Disorders

42

Chapter 3

CNV Analysis of ASD Cohort Reveals Novel Risk Gene NRXN3

Parts of this chapter are adapted, with permission for use from Cell Press, from the following

published journal article:

Vaags AK*, Lionel AC*, Sato D, Goodenberger M, Stein QP, Curran S, Ogilvie C, Ahn JW,

Drmic I, Senman L, Chrysler C, Thompson A, Russell C, Prasad A, Walker S, Pinto D, Marshall

CR, Stavropoulos DJ, Zwaigenbaum L, Fernandez BA, Fombonne E, Bolton PF, Collier DA,

Hodge JC, Roberts W, Szatmari P, Scherer SW. Rare deletions at the neurexin 3 locus in autism

spectrum disorder. American Journal of Human Genetics; 2012; 90(1):133-41. *Joint first

authors

I performed CNV analysis of ASD cohort and controls, prioritized rare CNVs for follow-up,

discovered NRXN3 deletions, as well as drafted and revised the manuscript. Dr. A.K. Vaags

followed up on clinical information for ASD probands with NRXN3 deletions and their families,

performed inheritance testing, and drafted and revised the manuscript.

Page 56: Copy Number Variation in Neurodevelopmental Disorders

43

3.1 Abstract

Autism spectrum disorder (ASD), a common NDD with an estimated prevalence rate of nearly

1% of children, is characterized by impairment in reciprocal social interaction, communication

deficits and a repetitive pattern of behavior. In recent years, several studies have highlighted the

role of rare copy number variants (CNVs) in the genetic etiology of ASD. The study of these rare

deletions and duplications in the genomes of ASD patients has proven to be a powerful tool for

the identification of novel candidate genetic loci for further investigation by targeted gene

sequencing and functional studies. This chapter describes the results from a genome-wide CNV

scan of more than 700 unrelated, newly characterized Canadian ASD patients. Rare exonic

CNVs were detected in the ASD cases that were absent in more than 2,000 ancestry-matched

population based controls and overlapped previously implicated ASD loci (e.g. PTCHD1, 15q11-

q13, 16p11.2), or identified new candidate susceptibility loci for ASD (e.g. NRXN3). Follow-up

investigation of the neurexin 3 (NRXN3) locus in other clinical cohorts revealed additional rare

inherited and de novo deletions overlapping this gene in individuals with ASD. In-depth clinical

characterization of the families of the individuals with these deletions revealed complexities of

penetrance and expressivity at this locus. These findings provide the final piece of evidence for

the involvement of all three members of the synaptic neurexin protein family in ASD risk.

Page 57: Copy Number Variation in Neurodevelopmental Disorders

44

3.2 Introduction

As described in Chapter 1, the emerging picture of ASD risk genetics is highly complex. Despite

strong evidence that ASD has genetic underpinnings, the specific risk genes involved in the

disorder have been hard to pinpoint until recently. The advent of whole-genome microarray

technology has highlighted the role of rare CNVs affecting certain genetic regions in a subset of

ASD cases. Rare CNV detection and characterization has proven to be a powerful tool for the

identification of novel candidate genetic loci for further investigation by targeted gene

sequencing and functional studies. We conducted and present here the results of a genome-wide

CNV scan of more than 700 unrelated, newly characterized Canadian ASD patients.

3.3 Results

3.3.1 Candidate risk loci identified by rare CNVs in ASD cohort

To identify candidate risk genes for ASD using the rare CNV data, the following categories of

variants were prioritized: de novo CNVs, rare CNVs overlapping the same genetic region in

unrelated ASD probands and rare CNVs affecting regions previously implicated in risk for ASD

or other NDD. Table 3.1 shows examples of rare CNV results from each of these categories.

DNA was available from parents of some of the probands for inheritance testing and this enabled

the detection of de novo variants. One de novo deletion was observed to affect several exons of

the autism susceptibility candidate 2 (AUTS2) gene on chromosome 7. Recent studies have

provided support for the role of AUTS2 deletions in ASD and ID (Beunders et al., 2013;

Nagamani et al., 2013). The largest such investigation by Beunders et al. (2013) detected such

de novo deletions at a frequency of 1 in 2,000 in cohorts of ID and ASD individuals but did not

find similar events in more than 15,000 controls. Modeling of the deletion in zebrafish revealed

recapitulation of phenotypes often observed in human patients with such deletions including

microcephaly and craniofacial defects. These deficits in the zebrafish knockdown models were

rescued by introduction of the human mRNA, providing further evidence for the causal role of

such deletions. Another de novo CNV previously implicated in NDD risk was a 22q11.2

duplication. Although marked by variable penetrance and lower severity than the reciprocal

Page 58: Copy Number Variation in Neurodevelopmental Disorders

45

22q11.2DS, this gain has been reported in connection with ASD and ID (Lo-Castro et al., 2009;

Pinto et al., 2010).

Examples of genetic regions implicated by CNVs previously reported in connected with ASD

include 15q11-13 duplications in two individuals and a 16p11.2 deletion in one individual. As

discussed in Chapter 1, these genetic syndromes often feature ASD as a core phenotype amidst a

spectrum of other varying clinical traits. The PTCHD1/DDX53 region at Xq22 is another bona

fide ASD risk locus affected by a deletion in this cohort. After initial discovery (Marshall et al.,

2008), several other studies have reported deletions at this gene in ASD (Noor et al., 2010;

Sanders et al., 2011) and ID (Filges et al., 2011) Several rare exonic CNVs in the ASD patient

cohort , including CNVs at the ASTN2/TRIM32 locus in three individuals, were observed to

overlap findings from the ADHD cohort presented in Chapter 4 (Table 3.2) and rare exonic

deletions at the synaptic GPHN gene were observed in one individual from each of the ASD and

schizophrenia cohorts. Overlapping rare CNVs between ASD and ADHD are discussed in

Chapter 4, the GPHN deletions are discussed in Chapter 6 and further details about the

ASTN2/TRIM32 locus are presented in Chapters 4 and 7.

One intriguing novel candidate risk gene implicated by overlapping rare CNVs in multiple ASD

probands was neurexin 3 (NRXN3), which is located at chromosome 14 and has two major

transcript isoforms (Figure 3.1).Rare exonic deletions affected the first exon of the longer alpha

isoform in one patient and several terminal exons of both isoforms of the gene in the second

proband. NRXN3, together with the other two members of the neurexin gene family (NRXN1 on

chromosome 2 and NRXN2 on chromosome 11) has been shown to have important roles in

synaptic cell adhesion and neurotransmitter secretion (Missler et al., 2003). While there were no

previous reports of NRXN3 in connection with ASD, the other members of the gene family have

well-established roles in NDD risk (Reichelt et al., 2012; Sudhof, 2008). As described in Chapter

1, rare exonic deletions within NRXN1 are among the most consistently observed findings from

CNV investigations of ASD and other NDDs. A truncating NRXN2 mutation has been reported

in a patient with ASD and was inherited from a father with severe language delay and family

history of schizophrenia (Gauthier et al., 2011). Given the above evidence, the NRXN3 locus was

further investigated as described in the next section.

Page 59: Copy Number Variation in Neurodevelopmental Disorders

46

Table 3.1. Examples of risk loci implicated by rare CNVs in ASD probands

Sample Sex CNV Locus Size (kb) Inh Genes Previous reports1

A. Loci implicated by de novo CNVs in ASD probands

SK0485 F Loss 7q11.22 330 D AUTS2 ASD, ID

124498 M Loss 12q24.33 108 D POLE,PGAM5,PXMP2,ANKLE2

149151 M Gain 22q11.2 2,985 D 64 genes ASD, ID

B. Rare CNVs at loci previously implicated in ASD

153706 M Gain 15q11-13 6,267 U > 100 genes ASD

114963L F Gain 15q11-13 5,639 U > 100 genes ASD

121851 M Loss 16p11.2 610 M 28 genes ASD

91548 M Loss Xp22 124 M PTCHD1/DDX53 ASD, ID

C. Overlapping rare CNVs in unrelated ASD probands

64119 M Loss 9q33.1 90 M ASTN2, TRIM32 ASD, ID, SZ

128963 M Loss 9q33.1 116 M ASTN2, TRIM32 ASD, ID, SZ

136929 M Gain 9q33.1 45 M ASTN2 ASD, ID, SZ

F1-003 M Loss 14q24.3 63 M NRXN3

F2-003 M Loss 14q24.3 292 P NRXN3

Abbreviations: Inh, Inheritance; P, Paternal; M, Maternal; D, De novo; U, Unknown;

1 This column lists the neuropsychiatric disorder(s) in connection to which each locus has been reported by previous studies : ADHD: Attention deficit hyperactivity disorder, ASD: Autism spectrum disorder, ID: Intellectual disability, SZ: Schizophrenia.

Page 60: Copy Number Variation in Neurodevelopmental Disorders

47

Table 3.2. Examples of risk loci implicated by rare exonic CNVs across 3 NDD cohorts

Genetic

locus Gene(s)

ASD

dataset1

ADHD

dataset1

Schizophrenia

dataset1

7q32.2 CHCHD3 1 gain 1 gain -

9q33.1 ASTN2 & TRIM32 2 losses,

1 gain 2 losses

12q24.3 ANKLE2,POLE, PGAM5,PXMP2 1 loss 1 gain -

14q23.3 GPHN 1 loss - 1 loss

16p11.2 28 genes 1 loss 1 gain 3 gains

20p12.1 MACROD2 & FLRT3 1 loss 2 losses -

Xp22.1 PTCHD1/TRIM32 1 loss 1 loss

Page 61: Copy Number Variation in Neurodevelopmental Disorders

48

3.3.2 Follow-up of rare exonic deletions at the NRXN3 locus

Following the detection of the rare exonic NRXN3 deletions in two individuals with ASD (F1-

003 and F2-003 in Figures 3.1 and 3.2), this gene was screened for additional CNVs in other

ASD cohorts including 447 Canadian patients and two groups of individuals referred for clinical

microarray testing to the diagnostic laboratories at the Mayo Clinic, Rochester, Minnesota (n =

1,796) and at Guy’s Hospital, London, UK (n = 1,368). Additional exonic NRXN3 deletions were

identified in one individual from each of the diagnostic laboratory cohorts (F3-003 and F4-003 in

Figures 3.1 and 3.2). Following independent experimental validation of the NRXN3 deletions for

each of the four ASD patients, detailed clinical information was obtained from their referring

physicians together with DNA samples from the parents and several of the immediate relatives of

the probands. Inheritance testing revealed the deletions to be of de novo origin in family 4,

paternally inherited in families 2 and 3 and maternally inherited in family 1(Figure 3.2).

The NRXN3 locus described here are very rare genetic events and not often observed in the

general population. In the 15,122 control individuals examined for CNVs affecting this gene,

only four possessed exonic NRXN3 deletions (Figure 3.1). Taken together, the frequency of

exonic deletions at the NRXN3 locus is significantly higher in ASD cases than in controls

(4/4,322 cases versus 4/15,122 controls; Fisher’s exact test, one-tailed p = 0.039). Three of the

control deletions affected only the alpha isoform while the fourth affected only the beta isoform

(Figure 3.1). Thus, while three of the four deletions in the cases affect both of the known NRXN3

isoforms (alpha and beta), none of the control deletions do so. This difference is also statistically

significant Fisher’s exact test, (one-tailed p = 0.0055). These findings parallel what is observed

in NRXN1, for which deletions, while extremely rare, have been reported in a few control

individuals (Ching et al., 2010; Rujescu et al., 2009). CNVs present in control individuals might

be indicative of reduced penetrance of NXRN3 CNVs (perhaps modified by the non-deleted

allele or other modifier genes), the presence of CNVs within control individuals in regions where

they do not destroy NRXN3 function, a lack of rigorous testing of the neuropsychiatric phenotype

in the controls, or false-positive microarray CNV calls in the controls, since DNA was not

available for their independent experimental confirmation.

Page 62: Copy Number Variation in Neurodevelopmental Disorders

49

Figure 3.1. Rare exonic NRXN3 deletions in ASD probands and controls

The deletions detected at chromosome 14 affecting NRXN3 exons in four ASD probands and four control individuals are shown in red. NRXN3 has two major transcript isoforms – the longer alpha isoform and the shorter beta isoform. Genomic coordinates and isoform information are from hg18 genome build.

Page 63: Copy Number Variation in Neurodevelopmental Disorders

50

Figure 3.2. Pedigrees of the families of ASD probands with NRXN3 deletions

Black-filled symbols represent ASD-affected individuals, gray-filled symbols represent broader autism phenotype (BAP)-affected individuals, and unfilled symbols represent apparently unaffected individuals. Probands are marked with an arrow. Clinical diagnosis and segregation of the NRXN3 deletions are shown. Children are placed left to right in birth order from eldest to youngest. Abbreviations are as follows: ADHD, attention deficit hyperactivity disorder; OCD, obsessive compulsive disorder; and NA, not assessed.

Page 64: Copy Number Variation in Neurodevelopmental Disorders

51

3.3.3 Clinical information from ASD probands with NRXN3 deletions and

their families

Family 1 (Proband with ASD diagnosis; 63 kb maternally inherited deletion):

In Family 1, the male proband (F1-003) was conceived naturally to a 33 year-old (yro) mother

(F1-001) and 35 yro father (F1-002; Figure 3.2). The pregnancy was uncomplicated, although the

mother contracted pneumonia at 7 months of gestation. The proband was delivered by

spontaneous vaginal delivery at 38 weeks with a nuchal cord and birth weight 7 lbs 3 oz. The

parents first noted advanced speech and memorization of detail, as well as social difficulties at

12-24 months of age. At age 9, he was identified as “gifted” following assessment for learning

disabilities. Between ages 9-14, there were concerns of depression. A family stressor, at age 10,

may have contributed to suicidal ideation. At 13 he was diagnosed with Asperger syndrome. He

had difficulties with aggression, anger, transitions and stimulation, including bright lights and

small spaces. His non-verbal cognitive abilities were assessed at age 16 with the Leiter

International Performance Scale - Revised (Leiter-R) measure. He performed above age

expectations compared to his same-aged peers (92nd percentile). In addition, The Vineland

Adaptive Behavior Scales (VABS) was completed with age appropriate outcomes for

communication (39th percentile) and daily living skills (34th percentile), but socialization skills

were severely delayed (below the 1st percentile). Oral and Written Language Scales (OWLS)

testing indicated he had above average (93rd percentile) receptive language and listening

comprehension, but he declined to complete measures of expressive language (Oral Expression

Scale). The Autism Diagnostic Interview- Revised (ADI-R) combined with clinical assessment

was consistent with a diagnosis of Asperger syndrome. The proband reported that his mind races

and he has difficulty sleeping through the night. There have been recurrent episodes of

aggressive outbursts and depression for which he received ineffective treatment with risperidone.

Following complaints of recurrent headaches, he was diagnosed with bilateral chronic migraine

headaches with episodic visual symptoms, likely due to migrainous aura. At 16 yro, he attended

university part-time while maintaining a part-time job.

F1-003 is one of three children of non-consanguineous parents of Irish and English descent. The

mother has self-reported high energy levels, social difficulties, anxiety, as well as auditory and

language processing difficulties. She did not meet criteria for ASD or sub-clinical autism/broader

Page 65: Copy Number Variation in Neurodevelopmental Disorders

52

autism phenotype (BAP) upon clinical interview assessment. Within the extended maternal

family, there is one 3rd degree relative with an ASD diagnosis, as well as several 2nd and 3rd

degree relatives with suspected ASD characteristics. The father, who is apparently healthy, has a

brother with Down syndrome and a 2nd and 3rd degree relative with suspected ASD

characteristics.

The proband’s deletion-negative brother (F1-004) has a normal 46,XY karyotype and normal

FMR-1 repeat. He had a diagnosis of attention deficit hyperactivity disorder (ADHD) at age 12;

testing with the Autism Diagnostic Observation Schedule (ADOS: Module 4) at age 14 did not

meet formal diagnostic criteria for an ASD. He performed at age expectations upon completion

of the Leiter-R test (63rd percentile). His receptive and expressive language skills were above age

expectations (OWLS, 88th percentile) and adaptive behavior was adequate in measures of

communication, daily living skills and socialization as assessed by VABS.

The proband’s sister (F1-005) has been diagnosed with Down syndrome and possesses a

47,XX,+21 karyotype and normal FMR-1 repeat. A research assessment at age 11 revealed she

was below the 1st percentile for her age on the Leiter-R. Her receptive and expressive language

skills also fell below the 1st percentile on the OWLS assessment. She showed mild to moderate

deficits across her adaptive behavior profile for her age. She met criteria for ASD on the ADOS:

Module 2 and ADI-R Behavior dimension only, but did not meet clinical criteria when assessed

by a developmental pediatrician. Her behavior was consistent with her Trisomy 21 status, and

complicated the diagnostic process. At age 16, heightened anxiety as well as increased obsessive

and impulsive behaviors prompted a second clinical assessment. The obsessive and repetitive

components of her behavior were the most challenging and interfered with her adaptive

functioning but an overall diagnosis of non-ASD was given.

Family 2 (Proband with ASD diagnosis; 292 kb paternally inherited deletion):

Family 2 includes trizygous triplets: the proband (F2-003) and an affected sister (F2-004)

inherited the NRXN3 deletion from their father, while an unaffected brother (F2-005; Figure 3.2)

did not inherit the deletion. The triplets were conceived naturally to a 23 yro mother (F2-001)

and 30 yro father (F2-002) and were delivered at 31 weeks of gestation at a weight of 3 lbs, 12

oz. for the proband, and 3 lbs, 4 oz. for the affected sister. Multiple medical interventions were

undertaken at birth due to the premature delivery. The proband was born with a right-side cranial

Page 66: Copy Number Variation in Neurodevelopmental Disorders

53

hemorrhage and the affected sister was also suspected to have suffered a cranial hemorrhage at

birth. All of the siblings have continuing difficulty with asthma, while the proband suffers from

juvenile arthritis and the affected sister has had multiple ear and respiratory infections.

The proband was diagnosed with an ASD through clinical assessment at 3 years, 5 months of

age. During testing for a research protocol at age five years, 8 months, he met the criteria for an

autism diagnosis utilizing the ADI-R and ADOS-Module 1. Non-verbal communication ability

and receptive and expressive language skills could not be assessed, as the proband was unable to

complete the IQ (Leiter-R) and language (OWLS) tests. Testing with the Peabody Picture

Vocabulary Test (PPVT-4) indicated the proband was extremely low functioning in receptive

vocabulary, as well as in his adaptive behavioral profile (VABS-II, below 1st percentile).

The ASD sister (F2-004) was diagnosed at 3 years, 9 months. As part of a research protocol, she

was assessed at the age of five years, 8 months to have autism using the ADI-R and ADOS-

Module 1. Her non-verbal communication ability was assessed with Leiter-R and found to be in

the low average range for her age group (12th percentile), while her VABS-II assessment

indicated extremely low-functioning (below 1st percentile). Assessment of receptive and

expressive language was attempted with the OWLS measure, but could not be completed.

Receptive vocabulary testing (PPVT-4) could also not be completed. The unaffected brother (F2-

005) was assessed at the age of 4 years, 10 months with a Social Communication Questionnaire

(SCQ), which indicated he did not have ASD (total score: 3, non-ASD).

These children were born to non-consanguineous parents of French Canadian descent; neither

has an ASD diagnosis. The mother has asthma and was assessed to not have a BAP, i.e. she was

not non-alexithymic on the Toronto Alexithymia Scale (TAS-20; total score: 29), while the

father is color blind and was also assessed be non-BAP (TAS-20 total score: 41). Clinical

information does not exist on male deletion carrier F2-006.

Family 3 (Proband with ASD diagnosis; 336 kb paternally inherited deletion):

F3-003 was diagnosed to have autism in early childhood and clinical testing with an ADI-R at

age 13 confirmed this finding. He had issues with aggression, self-harm, obsessive behaviors,

persecutory and delusional ideas, suicidal ideation and sleep problems. The father who also

Page 67: Copy Number Variation in Neurodevelopmental Disorders

54

carried the NRXN3 deletion was assessed by clinical review to have BAP as well as learning

disability, depression and alcohol-related issues.

Family 4 (Proband with ASD diagnosis; 247 kb de novo deletion):

In Family 4, the male proband (F4-003) was born at 38 weeks, weighing 8 lbs, 4 oz. The

pregnancy was complicated by insulin controlled gestational diabetes and hypertension. He was

delivered via cesarean section due to maternal hemorrhaging. The neonatal course was

uneventful. During 19 to 24 months of age, the parents first noted their son had difficulty being

consoled, developed temper tantrums, started to bang his head, developed difficulty making eye

contact, and started to develop some stiffness and rigidity. Between age 2-3, he could not keep to

a schedule, developed extreme reactions to taste and touching, and had trouble falling asleep. At

2 years, 10 months he received a clinical diagnosis of ASD, speech language delays and sleep

onset disorder. At 4 year, 6 months, a Caregiver-Teacher Report Form for Ages 1.5-5 was

completed. His scores endorsed internalizing/externalizing problems and the DSM-Oriented

Scales for Boys was in the clinical range for pervasive developmental problems (above the 97th

percentile). At 6 years, 10 months he had near normal speech, no longer banged his head nor had

oppositional defiance. His main concerns included high levels of anxiety and temper tantrums in

unfamiliar situations. An individualized educational plan was in place for math, reading, and

writing and he had a one-on-one educational assistant. Although he was in good general health,

he received occupational therapy for difficulty with thumb flexion and hypotonia of the hands

and chest.

3.4 Discussion

These findings support the hypothesis that hemizygous deletions involving NRXN3 may be

involved in the manifestation of an ASD or sub-clinical autism / BAP phenotype. Perhaps most

compelling, the deletion in Family 4 arose de novo, lending strong support for a causative

relationship between the loss of NRXN3 and the development of ASD.

Family 1 carried a deletion that overlaps exon 1 NRXN3 alpha, which may predispose affected

individuals to a mild ASD or Asperger syndrome phenotype in the absence of confounding

genomic changes, such as trisomy 21, as was observed in the Down syndrome-affected sister. As

Page 68: Copy Number Variation in Neurodevelopmental Disorders

55

NRXN3 alpha is translated from a start codon located in exon 2 (Rowen et al., 2002; Tabuchi

and Sudhof, 2002), it is possible that full-length NRXN3 alpha could be produced despite the

deletion. The effect of deletion of exon 1 may be to decrease transcription of the gene or

diminish the stability of messenger RNA. Interestingly, the same CNV deletion is present within

the mother who does not have a formal ASD diagnosis, but who self-identified as possessing

several ASD-like characteristics, including anxiety, social difficulties as well as auditory and

language processing difficulties. Thus, there may be variable expressivity of the NRXN3 deletion

and/or the necessity for as yet, unidentified, contributing genetic or environmental factors to

cause a clinical ASD phenotype.

In contrast, Family 2, 3 and 4 carried a deletion that removes multiple exons of both the alpha

and beta isoforms of NRXN3. The children carrying these deletions presented with a more severe

form of ASD than that identified in Family 1 wherein only exon 1 of the alpha isoform was

involved. This observation coupled with the absence of deletions affecting both isoforms of

NRXN3 in our controls could imply that the loss of the alpha isoform in addition to the beta

isoform may result in a more severe ASD phenotype. It is perhaps especially interesting to note

that the apparently unaffected carrier father and paternal grandfather in Family 2 carry the same

deletion and missense mutations as the ASD-affected children, yet have not been diagnosed with

an ASD or BAP. Conversely, the proband of Family 3 presented with autism while the carrier

father was diagnosed as having the BAP. The complexities of penetrance and variable

expressivity of NRXN3 in ASD may be partially explained by gene dosage balance and

functional redundancies of the neurexin gene family. Although all three neurexin genes have

similar exonic nucleotide sequences, exon-intron structure and patterns of alternative splicing for

mRNAs, NRXN1 and NRXN3 have the greatest sequence and protein identity. Expression

patterns vary between the neurexin genes and their alpha and beta isoforms. NRXN3 is expressed

throughout the brain, while other isoforms show differential locations and intensities of

expression (Ullrich et al., 1995).

Ultimately, to fully understand genotype and phenotype correlations will require knowledge of

the diploid expression patterns of the neurexin genes, other interacting molecules, and perhaps

non-genetic factors that might impede their homeostasis. The discovery of rare NRXN3 deletions

in ASD presented here provides a set of reference cases for ascertainment and comparison to

Page 69: Copy Number Variation in Neurodevelopmental Disorders

56

other patients, both enabling the molecular diagnosis of autism and providing potential new

targets for therapeutic intervention.

3.5 Materials and Methods

3.5.1 Study subjects and methodology

The cohort of 711 ASD patients used for the CNV analysis described in this chapter was

recruited from four different Canadian sites: The Hospital for Sick Children, Toronto, Ontario;

McMaster University, Hamilton, Ontario; Memorial University of Newfoundland, St. John’s,

Newfoundland and the University of Alberta, Edmonton, Alberta. ASD diagnoses were made

using uniform criteria common to all the sites and were based on a combination of two widely

used and clinically validated diagnostic tools: the Autism Diagnostic Interview-Revised (ADI-R)

and/or the Autism Diagnostic Observation Schedule (ADOS) (Risi et al., 2006). Approval was

obtained for this study from the research ethics boards at The Hospital for Sick Children

(Toronto). All participants provided informed written consent. DNA from the ASD probands,

and from their parents and immediate family members if available, was genotyped and analyzed

for CNVs using the methodology outlined in Chapter 2.

Page 70: Copy Number Variation in Neurodevelopmental Disorders

57

Chapter 4

Rare Copy Number Variant Discovery and Cross-Disorder

Comparisons Identify Risk Genes for ADHD

Parts of this chapter are adapted, with permission for use from AAAS, from the following

published journal article:

Lionel AC, Crosbie J, Barbosa N, Goodale T, Thiruvahindrapuram B, Rickaby J, Gazzellone M,

Carson AR, Howe JL, Wang Z, Wei J, Stewart AF, Roberts R, McPherson R, Fiebig A, Franke

A, Schreiber S, Zwaigenbaum L, Fernandez BA, Roberts W, Arnold PD, Szatmari P, Marshall

CR, Schachar R, Scherer SW. Rare copy number variation discovery and cross-disorder

comparisons identify risk genes for Attention Deficit Hyperactivity Disorder (ADHD).

Science Translational Medicine; 2011; 3(95):95ra75.

I designed and implemented CNV analysis of ADHD cases and controls and cross-disorder

comparison with ASD, prioritized rare CNVs for follow-up, coordinated experimental validation

and collection of information for clinical interpretation of CNVs and drafted and revised the

manuscript.

Page 71: Copy Number Variation in Neurodevelopmental Disorders

58

4.1 Abstract

Attention deficit hyperactivity disorder (ADHD) is a common and persistent condition

characterized by developmentally atypical and impairing inattention, hyperactivity and

impulsiveness. This chapter describes the identification of de novo and rare copy number

variations (CNVs) in 248 unrelated ADHD patients using million feature genotyping arrays. De

novo CNVs were detected in 3 (1.7%) of the 173 ADHD patients for whom DNA was available

from both parents. These CNVs affected brain-expressed genes: DCLK2, SORCS1/SORCS3 and

MACROD2. Rare inherited CNVs, that were absent in 2,357 controls and that either overlapped

previously implicated ADHD loci (for example DRD5 and 15q13 micro-duplication), or

identified new candidate susceptibility genes (ASTN2, CPLX2, ZBBX, PTPRN2) were detected in

19 of 248 (7.7%) ADHD probands. Among these de novo and rare-inherited CNVs there were

also examples of genes (e.g. ASTN2, GABRG1, CNTN5) previously implicated by rare CNVs in

other neurodevelopmental conditions including autism spectrum disorder (ASD). To further

explore the overlap of risks in ADHD and ASD, the rare CNV findings in the ADHD cohort

were compared with those from a newly collected cohort of individuals with a primary diagnosis

of ASD. Deletions of the neuronal ASTN2 and the ASTN2-intronic TRIM32 genes yielded the

strongest association with ADHD and ASD, but numerous other shared candidate genes (such as

CHCHD3, MACROD2 and the 16p11.2 region) were also revealed. These results provide support

for a role for rare CNVs in ADHD risk and reinforce evidence for the existence of common

underlying susceptibility genes for ADHD, ASD and other neuropsychiatric disorders.

Page 72: Copy Number Variation in Neurodevelopmental Disorders

59

4.2 Introduction

As outlined in Chapter 1, despite substantial evidence for ADHD having a genetic basis,

common variant approaches such as GWA have failed to find specific risk genes of large effect

sizes. This could suggest that rare genetic variants have a role in the etiology of the disorder. In

this study of a well-characterized Canadian cohort, we sought to explore the etiologic role of rare

inherited CNVs in ADHD, as well as to determine whether de novo CNV mutations might also

contribute. Our primary study design emphasized the use of identical high-resolution genotyping

microarrays and detection algorithms for all cases and controls, yielding a robust discovery data

set. We also compared our ADHD-CNV data with CNVs from the ASD cohort described in

Chapter 3, which was genotyped with the same microarray platform and analysis design (Chapter

2) to test whether neuropsychiatric risk genes might have pleiotropic effects across—or be

expressed as traits within—disorders.

4.3 Results

4.3.1 Detection of rare CNVs in ADHD cohort

In this study, 248 unrelated ADHD probands (73 female and 175 male) collected in Canada were

genotyped using the Affymetrix single nucleotide polymorphism (SNP) 6.0 microarray and

assessed for CNVs using the methods described in Chapter 2. In 173 of these families, DNA was

genotyped in both parents. Concurrently, 2,357 population-based controls including 1,234 from

Ontario (Stewart et al., 2009) and 1,123 from Germany (Krawczak et al., 2006) were assessed for

CNVs using the same array platform and calling strategy as the cases. A dataset of rare,

stringently defined CNVs was obtained from the ADHD probands that spanned a minimum of

five array probes, had support from two or more calling algorithms and were not found in

controls (less than 50% overlap by length). Using quantitative PCR (qPCR), 33 / 34 (97.1%) rare

stringent calls that were tested were validated as true positives, including all 23 CNVs in Table

4.1. De novo CNVs and rare inherited CNVs which overlapped genetic loci previously

implicated in ADHD or in other NDDs were prioritized to obtained novel ADHD candidate gene

loci (Table 4.1, Figure 4.1 and Table 4.2).

Page 73: Copy Number Variation in Neurodevelopmental Disorders

60

Table 4.1. De novo and rare inherited CNVs at candidate loci in ADHD probands

Sample Sex CNV Locus Size (kb) Inh Genes Previous reports1

A. Loci implicated by de novo CNVs in ADHD probands

27696.3 M Loss 4q31.3 33 D DCLK22 Epilepsy

30600.3 M

Gain 10q25.1 242 D SORCS32 BPD

Gain 10q25.1 318 D SORCS12 BPD

113400.3 M Loss 20p12.1 109 D MACROD23 ASD

B. Rare CNVs at loci previously implicated in ADHD

27027.3 M Gain 4p16.1 482 P DRD5,WDR1,SLC2A9 ADHD

27060.3 M Gain 5q35.2 131 M CPLX2 ADHD

63900.3 F Gain 7q32.3 159 M CHCHD3 ADHD

19896.3 M Gain 7q36.3 1,023 M PTPRN2,WDR60,NCAPG2,ESYT2 ADHD

19929.3 F Gain 11q13.4 147 U4 5 genes ADHD

19752.3 F Gain 15q13 3,666 M 15q13 locus (16 genes) ADHD, ASD, SZ

19839.3 M Gain 16p11.2 789 P 16p11.2 locus (40 genes) ADHD, ASD, SZ

C. Overlapping rare CNVs in unrelated ADHD probands

22512.3 F Loss 3q26.1 165 P ZBBX ADHD

19365.3 F Loss 3q26.1 22 U5 ZBBX ADHD

57300.3 M Gain 6p24.2 262 M TFAP2A,GCNT2,C6orf218 ASD

19698.4 F Loss 6p24.2 92 M GCNT2 ASD

19761.3 M Loss 9q33.1 177 P ASTN2, TRIM32 ASD, BPD, SZ

19812.3 M Loss 9q33.1 148 U4 ASTN2, TRIM32 ASD, BPD, SZ

108300.3 M Loss 20p12.1 542 M MACROD2, FLRT3 ASD

Page 74: Copy Number Variation in Neurodevelopmental Disorders

61

D. Rare CNVs at loci implicated in other NDDs

87600.3 F Loss 4p16 97 M STK32B mild ID

27048.3 M Loss 4p12 64 M GABRG1 ASD

125700.3 M Loss 11q22.1 49 P CNTN52(intronic) ADHD, ASD, SZ

89700.3 M Gain 12q24.33 211 P ANKLE2,POLE,PGAM5,PXMP2 ASD

27075.3 M Loss Xp22.11 388 M DDX53, upstream of PTCHD1 ASD, ID

Abbreviation used: Inh, Inheritance; P, Paternal; M, Maternal; D, De novo; U, Unknown;

1 This column lists the neuropsychiatric disorder(s) in connection to which each locus has been reported by previous studies : ADHD: Attention deficit hyperactivity disorder, ASD: Autism spectrum disorder, BPD: Bipolar disorder, ID: Intellectual disability, SZ: Schizophrenia.

2 Paralogs of these genes have been implicated by previous studies in the neuropsychiatric disorders listed in the last column.

3Loci highlighted in bold had overlapping rare CNVs in both the ADHD and the ASD datasets

4DNA was not available from the father for testing.

5DNA was not available from the mother for testing.

Page 75: Copy Number Variation in Neurodevelopmental Disorders

62

Figure 4.1. Pedigrees of ADHD families with rare CNVs of interest

(A to D) Pedigrees of ADHD families with variants of interest, which were absent in controls and (A) were de novo or (B) occurred at loci previously implicated in ADHD or (C) were present in two unrelated ADHD probands or (D) overlapped loci previously implicated in other NDDs. Circles and squares denote females and males, respectively, whereas arrows highlight the index proband in each family. Black filled objects indicate ADHD diagnosis, unfilled symbols signify unaffected family members, gray symbolizes individuals with non-ADHD neuropsychiatric conditions, and stripes represent individuals with some ADHD traits, but not having a definitive ADHD diagnosis. N/A denotes individuals from whom no DNA was available for testing. Detailed overview of phenotypes is presented in Table 4.2. Proband in family 30600 also has a diagnosis of bipolar disorder.

Page 76: Copy Number Variation in Neurodevelopmental Disorders

63

Table 4.2. Clinical phenotypes of ADHD families with rare CNVs of interest

Family Proband phenotype Maternal phenotype Paternal phenotype Sibling phenotype

27696 ADHD-I;

possible seizure disorder

(No CNV) neg.

(No CNV) neg.

(No CNV) sister: LD,ODD brother: mild tics

30600 ADHD-C

BPD (@ age 15) (No CNV)

neg. (No CNV)

ADHD

(No CNV) brother: ADHD, OCD

traits

113400 ADHD-I; LD (No CNV)

depression, inattention (No CNV)

neg. (No DNA)

2 brothers: Unknown

27027 ADHD-C, SAD

(No CNV) anxiety, depression, post-traumatic stress

disorder

(CNV present) ADHD

(No CNV) Sister: neg.

27060 ADHD-I, transient tic

disorder, LD (CNV present)

Depression (No CNV)

neg.

(No DNA) Brother:

ODD traits, OCD traits

63900 ADHD-I; ODD (CNV present)

unknown (No CNV) unknown

(No DNA) 2 brothers: Unknown

19896 ADHD-HI; ODD (CNV present)

depression (No CNV)

neg.

(CNV present) Brother:

ADHD, anxiety, ODD, query Asperger’s, learning problems

19929 ADHD-I, LD (No CNV)

anxiety, depression, learning problems

(No DNA) ADHD traits, learning problems, CD, problems with law, alcohol

and drugs

No siblings

19752 ADHD, LD, GAD Major depressive episode @ age 15

(CNV present) depression, learning

problems

(No CNV) query learning problems, query

ADHD, query ODD

(No DNA) sister and brother: neg.

for both

19839 ADHD-C; LD (No CNV) Anxiety

(CNV present) neg.

(No DNA) brother: LD. (No CNV) brother: LD

22512 ADHD-C, LD (No CNV)

neg.

(CNV present) query ADHD; anxiety/depressive

traits

(No DNA) sister: ADHD traits,

anxiety trait

19365 ADHD-I, ODD, sleep

disorder (No DNA)

neg. (No CNV)

neg.

(No DNA) 2 brothers: eldest brother - ADHD, language delay

57300 ADHD-I, dysthymia, LD, Crohn's disease

(CNV present) depression

(No CNV) learning problems, ADHD

No siblings

19698 ADHD-I, SAD, LD (CNV present)

neg.

(No DNA) depression, learning problems,

problems with law, alcohol abuse, drug abuse

(CNV present) brother: LD, depressive

traits

19761 ADHD-C, LD (No CNV)

neg. (CNV present)

anxiety (as a teen) (No DNA)

sister: unknown

19812 ADHD-C (No CNV)

neg. (No DNA)

neg. (CNV present) brother: neg.

108300 ADHD-C; LD;

depressive disorder-NOS

(CNV present) anxiety, depression

(No CNV) Anxiety, depression, ADHD traits

(CNV present) Sister: ADHD traits

Page 77: Copy Number Variation in Neurodevelopmental Disorders

64

87600 ADHD-C, ODD (CNV present)

anxiety; depression; inattentive traits

(No CNV) ADHD, ODD, learning problems

No siblings

27048 ADHD-C, ODD, sleep

disorder

(CNV present) anxiety, query learning

problems

(No CNV) ADHD

(No CNV) sister: Neg.

125700 ADHD, LD (No CNV)

neg. (CNV present) ADHD traits

(No DNA) brother: LD, query ODD.

(No CNV) sister: LD

89700 ADHD- C; LD (No CNV)

anxiety; traits of hyperactivity

(CNV present) query ADHD and LD, traits of

ODD and OCD

(No DNA) brother: language delay;

learning problems, ADHD traits

27075 ADHD, LD (CNV present)

anxiety, depression (No CNV)

neg.

(No DNA) 2 sisters: anxious traits in

eldest

Abbreviations: ADHD-C: (ADHD combined type), ADHD-I (ADHD inattentive type), ADHD-HI (ADHD hyperactive/impulsive type), BPD: Bipolar disorder, CD: Conduct disorder, GAD: Generalized anxiety disorder, LD: Learning disability, OCD: Obsessive-compulsive disorder, ODD: Oppositional defiant disorder; SAD: Separation anxiety disorder. “neg.” indicates individuals who did not have a diagnosis for, or any symptoms/traits of, any neuropsychiatric disorder, “CNV present” and “No CNV” indicate presence and absence of CNV respectively as determined by qPCR or FISH testing, “No DNA” indicates samples which could not be tested for CNV status due to unavailability of DNA. “Unknown” indicates individuals whose phenotype information is not known

Page 78: Copy Number Variation in Neurodevelopmental Disorders

65

4.3.2 Loci implicated by de novo CNVs in ADHD cases

Using data from probands and both parents (trio data), the de novo CNV rate in ADHD probands

was determined to be ~2% (3/173) (Table 4.1 and Figure 4.1). De novo CNVs were validated by

either qPCR or fluorescence in situ hybridization (FISH) after confirming parentage in all trios

with array genotypes. In one family, a 33kb de novo deletion at 4q31.3 was detected in a male

proband (27696.3) with ADHD and anxiety traits who also exhibited seizure-like symptoms. The

deletion eliminates two exons of the DCLK2 (doublecortin-like kinase 2) gene, which is

expressed in both proliferating neural cells and post-mitotic neurons (Tuy et al., 2008).

Mutations in the paralogous gene DCX have been associated with epilepsy and periodic limb

movements (Parisi et al., 2010), and the mouse double-knockout model for DCX and DCLK2

exhibits spontaneous seizures and a disorganized hippocampal network with aberrant positioning

of excitatory and inhibitory neurons (Kerjan et al., 2009). The DCLK1 region was one of the

stronger signals in a GWA study of 958 ADHD probands, but did not meet genome-wide

significance (Neale et al., 2008).

In one male ADHD proband (30600.3) of a second family, a pair of adjacent de novo CNVs at

10q25.1 duplicated regions of sizes 242 kb and 318 kb, overlapping the genes SORCS3 and

SORCS1, respectively. SORCS3 and SORCS1 are paralogs encoding sortilin-related VPS10

domain-containing receptor proteins, which are involved in intracellular sorting of surface

membrane proteins and are highly expressed in the developing and mature central nervous

system (Hermey et al., 2004). Upon phenotypic follow-up it was discovered that the proband,

who had met criteria for ADHD at age 8, was later diagnosed with bipolar disorder at age 15.

Indeed, variants at SORCS2 (a paralog of SORCS3 and SORCS1 mapping to 4p16.1) have been

shown by GWA studies to be potential risk factors for both ADHD (Lesch et al., 2008) and

bipolar disorder (Baum et al., 2008; Ollila et al., 2009). Neither the proband’s father nor brother

possessed these CNVs, although both had a diagnosis of ADHD, but not bipolar disorder. Thus,

the de novo duplications may contribute to the bipolar disorder phenotype of the proband.

Lastly, a 109 kb de novo exonic deletion was detected in a male ADHD proband (113400.3)

overlapping MACROD2 at 20p12.1. Further support for the potential pathogenicity of this locus

comes from the presence of a rare, maternally inherited deletion exonic to MACROD2 and

FLRT3, a neuronal cell adhesion gene intrinsic to MACROD2, in another ADHD proband

Page 79: Copy Number Variation in Neurodevelopmental Disorders

66

(108300.3) in this study, and from reports of rare CNVs at this locus in two previous ADHD

CNV scans (Elia et al., 2010; Williams et al., 2010). While little is known about its function, the

MACROD2 gene is expressed in the brain and represented the strongest association in a recent

GWA analysis of ASD (Anney et al., 2010).

4.3.3 Rare inherited CNVs at loci previously implicated in ADHD and

other NDDs

Among rare inherited CNVs that overlapped loci previously reported in genetic studies of

ADHD (Table 4.1 and Figure 4.1) were large inherited duplications at 16p11.2 and 15q13 in

ADHD probands 19839.3 and 19752.3, respectively, consistent with reports of ADHD being a

frequent phenotypic component in patients with microdeletions and duplications at these two loci

(Miller et al., 2009; Shinawi et al., 2010). Other genes of interest implicated by overlap with rare

duplications included DRD5 (at 4p16.1) and PTPRN2 (at 7q36.3). Genetic variants within the

dopamine receptor subtype D5 gene (DRD5) have been associated with ADHD (Barr et al.,

2000b) and in this study, a duplication encompassing DRD5 was transmitted to a male proband

(27027.3) from a father having a presumptive ADHD diagnosis. The duplication at 7q36.3 was

transmitted to two brothers with ADHD (19896.3) from their mother and overlaps PTPRN2,

which encodes a protein tyrosine phosphatase. This genetic locus has been associated with

ADHD traits (Lasky-Su et al., 2008) and with behavioral and learning disturbances in mice with

deletion of the PTPRN2 homolog (Nishimura et al., 2009).

Other CNVs overlapping loci previously reported in ADHD CNV scans (Elia et al., 2010; Lesch

et al., 2011) were detected at ZBBX (two exonic deletions) (Table 4.1 and Figure 4.1), CPLX2,

CHCHD3 and 11q13.4. ZBBX was implicated previously by a de novo deletion in a male ADHD

proband (Lesch et al., 2011), and separately by a CNV gain disrupting ZBBX (Elia et al., 2010).

The observation of CNVs at Complexin 2 (CPLX2) across multiple studies is intriguing given its

co-localization with the SNAP-25 protein in the SNARE complex (McMahon et al., 1995). The

SNAP25 gene has been implicated in ADHD via candidate gene association testing (Barr et al.,

2000a; Kustanovich et al., 2003) and Cplx2 knockout mice exhibited behavioral deficits and

cognitive abnormalities (Glynn et al., 2003).

Page 80: Copy Number Variation in Neurodevelopmental Disorders

67

Overlap was also observed of rare CNVs in the ADHD cohort with those previously implicated

in other neuropsychiatric disorders, notably ASD, including deletions at genes CNTN5,

GABRG1, GCNT2 and STK32B (Table 4.1 and Figure 4.1). There is evidence from multiple

CNV studies for the involvement of contactins in neuropsychiatric disorders (Burbach and van

der Zwaag, 2009). Moreover, disruption of the GABRG1 (Vincent et al., 2006) and GCNT2 (van

der Zwaag et al., 2009) genes have previously been reported in ASD.

4.3.4 Overlap between rare CNV findings in the ADHD and ASD cohorts

To explore further the hypothesis that rare CNVs absent in controls and present in both ASD and

ADHD patients could represent putative candidates for genetic risk across both conditions, the

rare CNV findings in the ADHD cohort were compared with those in the ASD cohort described

in Chapter 3. Usage of an identical microarray platform and CNV detection methodology

(Chapter 2) for the ASD and ADHD cases and the controls allowed for a robust comparison.

Genetic loci highlighted by the occurrence of rare CNVs in both the ASD and ADHD datasets

included those previously mentioned (16p11.2 locus, MACROD2 and CHCHD3) as well as the

12q24.33 region and the X-linked DDX53/PTCHD1 locus.

The most intriguing finding was a significant enrichment of rare CNVs at 9q33.1 overlapping

ASTN2 and TRIM32 in four ASD (~1%) and two ADHD probands (~1%) (Figure 4.2) (frequency

of 6/597 in cases vs. 0/2,357 in controls - Fisher's exact test two tailed p-value = 6.68 x 10-5).

ASTN2 encodes astrotactin 2, a neuronal membrane protein exhibiting abundant expression in the

cerebellum in both the fetal and adult brain (Wilson et al., 2010). Three of the ASD probands

met criteria for ADHD as measured by the Conners’ hyperactivity questionnaire, while the fourth

showed signs of hyperactivity, but was below the threshold for ADHD diagnosis (Figure 4.2).

These findings highlight the potential contribution of this locus to the presence of ADHD

symptoms in ASD. While exceptionally rare CNVs at this locus have been reported in earlier

CNV studies of ASD (Glessner et al., 2009), bipolar disorder (Grozeva et al., 2010), intellectual

disability (Bernardini et al., 2010) and schizophrenia (Vrijenhoek et al., 2008) (Figure 4.2), this

is the first report of CNVs at this locus in ADHD. Following the publication of this ADHD study

(Lionel et al., 2011), newer studies have also identified exonic ASTN2/TRIM32 deletions in two

individuals with Tourette syndrome (Fernandez et al., 2012) and nine individuals with ID (Vulto-

van Silfhout et al., 2013)

Page 81: Copy Number Variation in Neurodevelopmental Disorders

68

Figure 4.2. Rare CNVs at the ASTN2/TRIM32 locus in ADHD and ASD probands

Overview of rare CNVs detected at the 9q33.1 locus, overlapping ASTN2 and/or TRIM32 in six unrelated male probands in this study (two ADHD and four ASD). Variants reported at this locus from other CNV studies of different neuropsychiatric disorders are also shown. The blue and red bars denote deletions and duplications, respectively. No CNVs were seen at this locus in 2,357 control individuals. Genomic coordinates and information about transcript isoforms are from Genome Build 36 (hg18). Pedigrees of four ASD probands with maternally inherited CNVs overlapping ASTN2 and/or TRIM32 are shown. Black filled symbols represent individuals with an ASD diagnosis, whereas black filled symbols with white stripes symbolize ASD probands who also met criteria for ADHD.

Page 82: Copy Number Variation in Neurodevelopmental Disorders

69

4.4 Discussion

These data delineate several rare CNVs identified in ADHD cases, which were not found in

controls, suggesting they may have a contributory role in the clinical phenotype observed in

these individuals. The rate of de novo CNVs observed in the ADHD cohort is ~1.7% (3/173),

just slightly higher than the typical ~1% that has been reported in non-disease control trios (Levy

et al., 2011; Sanders et al., 2011; Xu et al., 2008), but less than the 5 - 10% reported in ASD

(Marshall et al., 2008; Pinto et al., 2010; Sebat et al., 2007) and schizophrenia (Xu et al., 2008).

The similar rate of de novo CNV observed in ADHD cases and controls implies that de novo

events may not simply equate to pathogenicity in ADHD. However, one ADHD proband was

identified with seizure-like symptoms, who was carrying a de novo CNV disrupting DCLK2.

There was also another ADHD proband with de novo CNVs disrupting SORCS1 and SORCS3,

who developed bipolar disorder in adolescence. In both cases, the clinical presentation was

consistent with what might be predicted based on the annotated function of these genes. These

data, therefore, suggest that the contribution of de novo CNV to ADHD is likely small, but that

careful analysis of phenotype and genes affected (and comparison to other cases and controls)

may potentially be clinically revealing. Below, potential scenarios are provided based on case

descriptions for the probands carrying these two de novo CNVs (DCLK2 and SORCS1/SORCS3),

and a third case study of a male with a rare, maternally inherited X-linked CNV (at the PTCHD1

locus). These examples and others from this study support the need for both extensive and

longitudinal phenotype data including family histories in analyses such as these, particularly

when considering subjects exhibiting additional medical complications. Moreover, in all

instances, potential genotype-phenotype correlations will benefit from much larger sample sizes.

Case 1 (27696.3) was first diagnosed at age 7 with ADHD, with combined subtype and sub-

threshold anxiety symptoms that did not warrant separate diagnoses. At age 14, he continued to

show ADHD and social difficulties and the patient and his mother reported episodes during

which the patient stared blankly and was unresponsive to events going on around him, including

someone calling his own name. He also reported having unusual sensory experiences such as

hearing noises (e.g., high-pitched screeches, whispers) and smelling odors. The family was

concerned that he might have psychosis due to schizo-affective disorder in a maternal uncle and

schizophrenia plus depression in the maternal grandfather. There was no family history of

Page 83: Copy Number Variation in Neurodevelopmental Disorders

70

epilepsy or evidence of delusional or disordered thoughts, or mood symptoms. Neurological

examination and EEG found no abnormality. The patient was found to have a de novo deletion at

4q31.3 affecting DCLK2, a gene whose predicted function may account for his seizure-like

episodes.

Case 2 (30600.3), a male, was diagnosed with ADHD, combined subtype, at age 8. There were

no concerns regarding his mood. Family history was positive for ADHD in father and brother,

bipolar disorder and possible schizophrenia in a second degree relative (maternal aunt), possible

childhood schizophrenia in a maternal cousin, and alcoholism in multiple paternal and maternal

second degree relatives. At age 15 the patient was diagnosed with bipolar II disorder, after

several depressive episodes and a single manic episode characterized by inflated self-esteem,

grandiosity, hypersensitivity to environmental stimuli, rapid thought patterns, rapid speech, and

decreased need for sleep. He continued to meet diagnostic criteria for ADHD, combined subtype.

At age 18, both ADHD and bipolar disorder persisted with some escalation of impairment and

symptoms. Subsequent to the third assessment at age 18, de novo duplications at 10q25 were

identified affecting SORCS1 and SORCS3. Variants at SORCS2 have been associated with

bipolar disorder. The existence of these two CNVs in this patient may therefore have influenced

the presentation of bipolar disorder.

Case 3 (27075.3), a male, was diagnosed with ADHD at age 7. He was noted to have poor social

skills and anxious traits. He did not have a history of language delay, social isolation or

obsessive preoccupations. Subsequent to the initial assessment, a community psychiatrist

diagnosed depression, generalized anxiety disorder and learning disability. Assessment at ages

15 and 17 revealed continued ADHD and anxious traits, but no depression. His parents reported

that he preferred to be alone or with adults. Upon entering university, he began to flourish

academically but his anxiety increased with the heightened academic demands. At each

assessment point, his preference for social isolation was noted, but no formal diagnosis of ASD

was made, as he did not meet full criteria. Microarray analysis revealed a maternally inherited

deletion at the PTCHD1 locus at Xp22.11. This CNV has recently been implicated in ASD (Noor

et al., 2010) and could potentially contribute to, and explain the social interaction deficits in this

individual.

Page 84: Copy Number Variation in Neurodevelopmental Disorders

71

Overall, rare-inherited CNVs were detected at loci previously reported in ADHD or in other

NDDs in 8% of the ADHD sample, suggesting that they may be risk factors for ADHD and/or

associated neuropsychiatric phenotypes in the individuals carrying them. CNVs involving

previously implicated-ADHD loci such as DRD5 and 15q13 micro-duplications are most

compelling for further genomic analyses, but new candidate susceptibility genes such as CPLX2,

ZBBX and PTPRN2 were also discovered.

Rare-inherited CNVs were also identified in genes such as ASTN2, GABRG1 and CNTN5, which

are affected by rare CNVs in other NDDs. The apparent overlap of rare-inherited CNVs in the

ADHD cohort with those in cases with other NDDs, in particular ASD, supports the idea that

variants at a common set of genes could be involved in the etiology of several neuropsychiatric

disorders. To further explore this possibility, rare CNVs in a new cohort of unrelated individuals

having a primary diagnosis of ASD were compared to the CNV findings in the ADHD cohort.

Deletions of ASTN2 and the ASTN2-intronic TRIM32 genes yielded the strongest association

with ADHD and ASD. GWA (Lesch et al., 2008) and linkage studies (Arcos-Burgos et al., 2004;

Bakker et al., 2003; Romanos et al., 2008) provide additional evidence for involvement of the

9q33 region in ADHD. GWA studies have also suggested a role for this region in schizophrenia

(Glessner et al., 2010; Wang et al., 2010) and bipolar disorder (Wang et al., 2010).

ASTN2 has recently been shown to play a vital role in the developing mammalian brain by

forming a complex with its paralog astrotactin 1 (ASTN1) and regulating its expression on the

surface of young cerebellar neuroblasts (Wilson et al., 2010). Although ASTN2 and ASTN1 are

highly homologous, their protein products have been found to function together in a

complementary yet non-redundant manner (Wilson et al., 2010). In vitro studies, using

antibodies against ASTN1, have confirmed its role as a receptor for neuronal migration along

astroglial fibers by facilitating glial-neuron binding (Fishell and Hatten, 1991; Zheng et al.,

1996). Mice deficient in ASTN1 exhibit slowed neuronal migration, altered Purkinje cell

morphology and impairments in movement and co-ordination (Adams et al., 2002). TRIM32

encodes a TRI-partite Motif containing E3 ubiquitin ligase protein, which has been reported to

have a role in deciding the fate of neuronal stem cell lineages (Schwamborn et al., 2009).

Homozygous mutations in this gene have also been reported in a consanguineous family with

Bardet Biedl syndrome and cognitive impairment (Chiang et al., 2006). The observation of

deletions spanning both ASTN2 and TRIM32 in ASD and ADHD patients, could indicate that

Page 85: Copy Number Variation in Neurodevelopmental Disorders

72

disruption of both genes is associated with a higher risk for neuropsychiatric disorder in general

rather than risk for any specific disorder.

There has been recent debate about the possibility that intellectual disability, rather than ADHD

accounts for the large, rare variants previously observed in ADHD (Elia et al., 2011). It is

unlikely that low IQ could account for the rare CNVs reported in the current study, since ADHD

probands with below average IQ were excluded (defined as both a verbal and performance IQ

score below 80) and because the IQ of those subjects with rare CNVs did not differ significantly

from that of the total sample (Unpaired two-tailed t-test p-value = 0.07).

One intriguing finding that emerges from CNV studies of neuropsychiatric conditions such as

ADHD and ASD is that there seem to be common genes and pathways implicated in several

disorders. As detailed in Chapter 1, there is considerable clinical overlap between ADHD and

ASD. Recent genomic hypotheses put forth to explain clinical overlap and pleiotropy suggest

that different human disease-phenotype groups might arise from overlapping molecular causation

(Oti et al., 2008; Rzhetsky et al., 2007). This could involve different functional domains of a

single protein, the interaction between different proteins (such as a ligand and receptor), the

interaction of proteins in a multi-protein complex, or different steps in a cellular pathway. Crespi

and colleagues have also suggested that rare duplications and deletions at dosage sensitive genes

could be responsible for diametric models of neuropsychiatric disorders, e.g. considering

schizophrenia and autism as opposite conditions along a spectrum of social-brain phenotypes

from hypo-development in autism to hyper-development in schizophrenia (Crespi et al., 2010).

Given the extent and types of genes that were identified in this study, it is likely that some or all

of these mechanisms contribute to ADHD etiology. Moreover, all of these issues are exacerbated

in complex and apparently heterogeneous disorders like ADHD, where the genetic liability of

other risk alleles, protective genetic modifiers (both rare and common), and other factors like

gender can influence expression and penetrance of neuropsychiatric phenotypes.

The importance of obtaining detailed longitudinal phenotype data on the ADHD index cases, and

on their family members is illustrated by this work. In the three case studies (27696.3, 30600.3,

27075.3) discussed above and others described in Table 4.1, the original ascertainment diagnosis

of ADHD initially predominated the interpretation of the CNV association results, but these

interpretations were broadened upon receiving new clinical assessments as the participants

Page 86: Copy Number Variation in Neurodevelopmental Disorders

73

reached adolescence. Ascertainment issues will always present challenges in genetic studies of

neuropsychiatric disorders. One example relevant to this work is that the DSM-IV will not allow

a diagnosis of ADHD to a child with ASD, even though data suggest upwards of 50% of children

with ASD would otherwise meet criteria for ADHD. Ultimately then, these new results will serve

as a conservative starting point for much larger studies (such as the follow-up investigation of

the ASTN2/TRIM32 locus presented in Chapter 7) exploring the role of de novo and rare CNVs in

ADHD, ADHD-related disorders and other associated clinical comorbidities.

4.5 Materials and Methods

4.5.1 Study subjects in ADHD cohort

Participants were 175 boys (71%) and 73 girls (29%) of age 5 to 17 years old (mean = 9.5,

minimum = 5.6, maximum = 16.9) who were referred for assessment of attention, learning and

behavior problems to The Hospital for Sick Children, Toronto. Participants were included if they

met criteria for ADHD based on the results of a semi-structured diagnostic interview of the

participant’s parent(s) and of the proband’s teacher, and did not meet any of the exclusion

criteria specified by DSM-IV (mental retardation, pervasive developmental disorder, autism or

co-morbid psychiatric disorder that could better account for the disorder). The Parent Interview

for Child Symptoms (PICS) (Ickowicz et al., 2006) is similar to the Schedule for Affective

Disorders and Schizophrenia (KSADS) (Ambrosini, 2000) with an enhanced module for ADHD

and other disruptive behavior disorders. The PICS was conducted by a social worker with a

master of social work (MSW), a clinical nurse specialist or a clinical psychologist. Reliability of

ADHD diagnoses was assessed through videotaped interviews in 48 cases and was found to be

high (interclass correlation for total symptom score = 0.93). The Teacher Telephone Interview

(TTI) (Charach et al., 2009) is a semi-structured interview conducted by an interviewer with at

least a Master’s degree in psychology. Reliability of the TTI was assessed using audiotapes and

found to be high (interclass correlation for total symptom score = 0.93). Intelligence and

academic attainment were assessed by a clinical psychologist. To receive a best estimate

diagnosis of ADHD arrived at through consensus between the assessing psychiatrist and

psychologist, the participant had to present with impairing and developmentally atypical

symptoms before age 7, meet DSM-IV criteria based on PICS and/or TTI, exhibit evidence of

symptoms and impairment both at home and at school, not present with any of the exclusion

Page 87: Copy Number Variation in Neurodevelopmental Disorders

74

criteria for ADHD as stated in DSM-IV. Individuals were excluded if they had an IQ of less

than 80 on both the verbal and performance subscales of the WISC. The mean full-scale IQ for

the entire sample was 102.12 (SD = 12.74) and 107.57 (SD = 16.25) for the subsample with a

rare CNV of interest (Table 4.1). Presumptive psychiatric status of each proband’s parents and

siblings was established by family history. Parents and children over age 12 years provided

consent and younger participants gave assent. The protocol was approved by the Research Ethics

Board of the Hospital for Sick Children, Toronto. Microarray genotyping, CNV analysis and

prioritization were performed using the uniform methodology presented in Chapter 2.

Page 88: Copy Number Variation in Neurodevelopmental Disorders

75

Chapter 5

Pathogenic Rare Copy Number Variants in Community-Based

Schizophrenia Suggest a Potential Role for Clinical Microarrays

Parts of this chapter are adapted, with permission for use from Oxford Press, from the following

published journal article:

Costain G*, Lionel AC*, Merico D, Forsythe P, Russell K, Lowther C, Yuen T, Husted J,

Stavropoulos DJ, Speevak M, Chow EW, Marshall CR, Scherer SW, Bassett AS. Pathogenic rare

copy number variants in community-based schizophrenia suggest a potential role for clinical

microarrays. Human Molecular Genetics; 2013; 22(22):4485-501. *Joint first authors

My roles in this study were as follows: 1.Co-ordinated experimental batches for microarray

genotyping of DNA samples from schizophrenia patients; 2. Performed quality control of raw

microarray data from cases and controls as described in Chapter 2; 3.Designed and implemented

uniform CNV analysis strategy for cases and controls; filtered and curated final rare CNV dataset

used for all downstream analysis in the study; 4.Assisted with other aspects of data analysis and

interpretation (including annotation of genetic results, CNV burden calculation, interpretation of

gene-set enrichment analysis, and prioritization of specific CNV regions for follow-up in

additional control datasets); 5. Prioritized rare CNV loci for qPCR validation, coordinated and

interpreted results of validation experiments; 6. Drafted and revised the manuscript.

The roles of Dr. G. Costain were as follows: 1. Assisted with primary data collection (recruiting

subjects, performing comprehensive phenotypic assessments, and arranging for the collection of

blood or saliva samples for DNA extraction); 2. Checked, verified, and updated clinical data

(including but not limited to diagnosis, age at onset, family history, and syndromal status) and

assisting with manual curation of CNV data; 3. Created Table 5.1, Table 5.2, and Table 5.3, and

performing the systematic review of all genes included in these tables (>800 in total); 4. Assisted

with and coordinated other aspects of data analysis (including but not limited to the relatedness

calculations, quantitative burden analyses, and the blinded review of all large rare CNVs); 5.

Assisted with data interpretation (particularly in relation to findings concerning the burden of

large, rare CNVs and smaller CNVs implicating candidate genes for schizophrenia); and, 6.

Drafted and revised the manuscript.

Page 89: Copy Number Variation in Neurodevelopmental Disorders

76

Page 90: Copy Number Variation in Neurodevelopmental Disorders

77

5.1 Abstract

Individually rare, large copy number variants (CNVs) contribute to genetic vulnerability for

schizophrenia. Unresolved questions remain, however, regarding the anticipated yield of clinical

microarray testing in schizophrenia. Using high resolution genome-wide microarrays and

rigorous methods, we investigated rare CNVs in a prospectively recruited community-based

cohort of 459 unrelated adults with schizophrenia and estimated the minimum prevalence of

clinically significant CNVs that would be detectable on a clinical microarray. Blinded review by

two independent clinical cytogenetic laboratory directors of all large (>500 kb) rare CNVs in

cases and well-matched controls showed that those deemed clinically significant were highly

enriched in schizophrenia (16.4-fold increase, p<0.0001). In a single community catchment area,

the prevalence of individuals with these CNVs was 8.1%. Rare 1.7 Mb CNVs at 2q13 were

found to be significantly associated with schizophrenia for the first time, compared with the

prevalence in 23,838 population-based controls (42.9-fold increase, p=0.0002). Additional novel

findings that will facilitate the future clinical interpretation of smaller CNVs in schizophrenia

include: (i) a greater proportion of individuals with two or more rare exonic CNVs >10 kb in size

(1.5-fold increase, p=0.0134) in schizophrenia; (ii) the systematic discovery of new candidate

genes for schizophrenia; and (iii) pathway enrichment analysis highlighting a differential impact

in schizophrenia of rare exonic deletions involving diverse functions, including

neurodevelopmental and synaptic processes (4.7-fold increase, p=0.0060). These findings

suggest consideration of a potential role for clinical microarray testing in schizophrenia, as is

now the suggested standard of care for related developmental disorders like autism.

Page 91: Copy Number Variation in Neurodevelopmental Disorders

78

5.2 Introduction

Schizophrenia is a complex neuropsychiatric disease that affects up to 1% of the general

population and shows evidence for neurodevelopmental origins and genetic heterogeneity

(Costain and Bassett, 2012). There is compelling evidence that individually rare, large copy

number variants (CNVs), especially 22q11.2 microdeletions, play a role in the genetic causation

of schizophrenia (Bassett et al., 2010b). As for other major NDDs, such as autism and

intellectual disability, findings from recent genome-wide case-control studies of CNVs support a

multiple rare variant model of causation in schizophrenia (Bassett et al., 2010b; Cook and

Scherer, 2008; Costain and Bassett, 2012). In general, rare and large (e.g., >500 kb) CNVs, often

overlapping multiple genes, are more likely to have a clinically relevant phenotype (Cook and

Scherer, 2008). In schizophrenia, unresolved questions remain regarding the anticipated yield

today on clinical microarray testing of large rare CNVs of potential clinical relevance (Bassett et

al., 2010b; Cooper et al., 2011). Most case-control sampling schemes to date have not been

epidemiologic in nature, and therefore cannot provide reliable estimates (Bassett et al., 2010b).

This uncertainty about yield may contribute to the low rate of uptake of clinical genetic testing in

schizophrenia: only 5 of 38,779 samples processed at Signature Genomics were from patients

with this common and serious mental illness (Sahoo et al., 2011).

Smaller rare CNVs may implicate individual risk genes for schizophrenia (Lionel et al., 2013;

Rujescu et al., 2009), but the clinical interpretation of these variants is more challenging. A

practical consideration for an adult-onset condition like schizophrenia is the typical absence of

parental samples to determine de novo / inheritance status. High resolution genome-wide data

and systematic approaches (both quantitative and descriptive/qualitative) to identify putative

candidate genes are needed to inform a more sophisticated approach to the clinical assessment of

CNVs in schizophrenia. The existence of overarching genetic networks, whose disruption may

contribute to expression of schizophrenia, is a particularly tantalizing possibility (Kirov et al.,

2012; Raychaudhuri et al., 2009).

We prospectively ascertained and systematically assessed a community-based sample of

Canadian patients with schizophrenia to address the above issues. All subjects underwent direct

clinical screening assessments for potential syndromic features using a standardized protocol that

included review of available lifetime medical records and assessment of physical features

Page 92: Copy Number Variation in Neurodevelopmental Disorders

79

(Bassett et al., 2010a; Fung et al., 2008; Silversides et al., 2012). Cases were compared to an

unrelated epidemiological control sample of comparable (European) ancestry (Ontario

Population Genomics Platform; OPGP) (Silversides et al., 2012). We used a high resolution

genome-wide microarray and proven analytic methods for CNV detection and evaluation (Figure

5.1) (Lionel et al., 2011; Marshall et al., 2008; Pinto et al., 2010; Silversides et al., 2012). An

independent, blinded review process identified large rare CNVs deemed "Pathogenic" or of

"Uncertain clinical significance; likely pathogenic" in cases and controls, based on well-

established American College of Medical Genetics (ACMG) guidelines for CNV interpretation

(Kearney et al., 2011). A representative epidemiologic subsample from a single catchment area

(Bassett et al., 2010a) allowed us to estimate for the first time the minimum collective prevalence

of individuals with large, clinically significant CNVs in schizophrenia. To shed light on the

contribution of smaller CNVs to the genetic architecture of schizophrenia, we investigated the

total burden of such CNVs and the potential support for a “multiple hit” model of causation. We

also developed a systematic approach for the discovery of new candidate genes for

schizophrenia, which included pathway enrichment analysis to identify relevant genetic

networks.

Page 93: Copy Number Variation in Neurodevelopmental Disorders

80

Genotyping on Affymetrix 6.0 SNP array

CNV detection

• Comprehensive examination of genes for evidence of relevant expression in human and model animal (mouse, zebrafish) systems

• Overlap with previous genetic findings in SZ• Recurrent and overlapping rare CNVs identifying familiar and novel SZ risk genes• Experimental validation of CNVs (quantitative PCR or clinical microarray)

454 discovery sample SZ cases 416 OPGP controls

Stringent CNVs (SZ cases, OPGP controls)

CNV frequency (rarity) assignment using data from 2,357 independent controls2

CNVs present in > 0.1 % controls2

(<50% unique sequence)

Single algorithm only< 5 consecutive probes

< 10 kb in size

459 adult cases with schizophrenia (SZ)

5 cases with 22q11.2 deletions

Global CNV burden

Stringent rare CNVs (SZ cases, OPGP controls)

Rare CNVs present in both SZ and OPGP control datasets

Stringent very rare CNVs in SZ cases only3

Stringent rare CNV burden analysis1

(SZ vs. OPGP and syndromic SZ vs. non-syndromic SZ)

1CNV burden analysis focused on comparing the subset of unrelated SZ cases of European ancestry (n = 420) with ancestry-matched unrelated adult OPGP controls (n=416).2Population-based adult controls of European ancestry (n=2,357), used in previous studies of autism and attention deficit disorder and independent to OPGP controls, were genotyped on the Affymetrix 6.0 platform and analyzed for CNVs in an identical manner to the SZ cases and OPGP controls.3Included CNVs 4–10 kb in size (n=7) that overlapped CNVs > 10kb in size overlapping SZ candidate genes

CNVs present in any control individual2

(<50% unique sequence)Stringent very rare CNVs (SZ cases, OPGP controls)

Figure 5.1. Overview of study design and workflow

Page 94: Copy Number Variation in Neurodevelopmental Disorders

81

5.3 Results

5.3.1 Clinically significant and novel large rare structural variants in schizophrenia

Two clinical cytogenetic laboratory directors conducted independent blinded examinations of all

rare (present in < 0.1% of 2357 population controls) CNVs 500 kb–6.5 Mb in size in

schizophrenia cases of European ancestry and OPGP controls. This revealed a significant

enrichment in schizophrenia of individuals with ‘Pathogenic’ or ‘Uncertain clinical significance;

likely pathogenic’ (collectively termed ‘clinically significant’) large rare autosomal CNVs [16 of

420 versus 1 of 416; P < 0.0001, odds ratios (OR) 16.44 (95% CI 2.17–124.51)]. In

schizophrenia cases, there were 16 CNVs considered clinically significant at eight loci (Table

5.1). All were very rare (i.e., present in none of 2357 population controls used to adjudicate both

the schizophrenia case and OPGP control samples), including six CNVs at four loci novel to

schizophrenia: 2q13 (Yu et al., 2012), 3q13.31 (Molin et al., 2012), 5p15.33-p15.32 (Zhang et

al., 2005) and 10q11.22-q11.23 (Stankiewicz et al., 2012). None of the corresponding individuals

with schizophrenia had a pediatric diagnosis of moderate or severe intellectual disability,

multiple major congenital anomalies, epilepsy or autism. Notably, five typical 2.6 Mb 22q11.2

deletions in cases (none in controls) were a priori excluded to demonstrate that these findings are

not being driven by those established variants. The single large rare CNV in OPGP control

individuals deemed to be clinically significant was a typical 2.6 Mb 22q11.2 duplication. There

were no significant differences in demographics, years of education, age at onset, family history

of schizophrenia, self-reported history of special education or syndromic designation in these

cases with large, rare, clinically significant CNVs when compared with subjects with no large

rare CNVs 500 kb–6.5 Mb in size. There were also eight very large (>6.5 Mb) anomalies in

schizophrenia cases (none in OPGP controls) that were considered clinically significant (Table

5.1), including a 10 Mb loss (deletion) overlapping FOXP2 (Feuk et al., 2006).

A novel finding was the detection in three of 420 schizophrenia cases of European ancestry of

very rare, recurrent 1.7 Mb CNVs at 2q13 (Table 5.1), a significant enrichment compared with

the prevalence (n = 4) in 23 838 population-based controls (Costain et al., 2013) [P = 0.0002; OR

42.87 (95% CI 9.56–192.14)]. To our knowledge, this CNV has not been previously reported in

other genotyped schizophrenia cohorts [e.g., the International Schizophrenia Consortium dataset

Page 95: Copy Number Variation in Neurodevelopmental Disorders

82

(International Schizophrenia Consortium, 2008)]. None of the three individuals in our case

sample with these 2q13 CNVs had any of the pediatric diagnoses [developmental delay, multiple

congenital anomalies and/or autism spectrum disorder (ASD)] previously reported for samples

assessed at clinical laboratories (Cooper et al., 2011; Yu et al., 2012). Two of the three

individuals reported a history of learning difficulties in school, but all had IQs within the average

range. For the one schizophrenia proband with a positive family history, we were able to confirm

co-segregation of the 2q13 duplication with schizophrenia. Of the 10 genes overlapped by these

2q13CNVs, three are promising candidates for schizophrenia: the neurodevelopmental facilitator

ANAPC1 (Paridaen et al., 2009), the neuronal apoptosis regulator BCL2L11 (Youle and Strasser,

2008), and the TAM receptor component and multiple sclerosis risk gene MERTK (Binder and

Kilpatrick, 2009) (Table 5.1).

5.3.2 Prevalence of clinically significant CNVs in community sample of schizophrenia

We used our schizophrenia catchment sample to estimate the minimum prevalence of large (500

kb–6.5 Mb) rare clinically significant CNVs that would be detectable on a clinical microarray. In

the 248 unrelated case subjects in this catchment area with CNV data, the number of individuals

with these clinically significant CNVs, including two with 22q11.2 deletions, was 15 (6.0%, 95%

CI 3.7–9.8%) (Table 5.1). Nine (60.0%) of these 15 individuals were in the non-syndromic

subgroup (i.e., were a priori found to not meet our established criteria for syndromic features). If

the five anomalies >6.5 Mb were included (Table 5.1), the estimated prevalence would be 8.1%

(95% CI 5.2–12.2%). Notably, none of these 20 variants had been detected prior to study

participation. We estimate that there are a total of 370 unrelated adults with schizophrenia in this

catchment area (i.e., n = 122 not included in this study, in addition to the n = 248 with CNV

data). A conservative estimate of the minimum prevalence of clinically significant CNVs in this

catchment area, based on the assumption of no clinically significant variants in the estimated n =

122 unrelated individuals unavailable for study, would thus be 20 in 370 or 5.4% (95% CI 3.5–

8.3%).

5.3.3 Large rare CNVs of unknown significance in schizophrenia

Table 5.2 shows the remaining large rare CNVs in schizophrenia cases deemed to be of

‘Uncertain clinical significance (no sub-classification)’ (termed ‘variants of unknown

Page 96: Copy Number Variation in Neurodevelopmental Disorders

83

significance’; VUS), or of ‘Uncertain clinical significance; likely benign’ or ‘Benign’

(collectively termed ‘benign’). Of these, VUS [33 of 420 versus 12 of 416; P = 0.0012, OR 2.87

(95% CI 1.46–5.64)], but not benign CNVs, were significantly enriched in schizophrenia cases

of European ancestry compared to OPGP controls. There were three novel loci where there were

overlapping large rare CNVs in unrelated schizophrenia cases: 6q11.1 (two gains), 7p21.3 (one

gain, one loss) and 12q21.31 (one gain, one loss) (Table 5.2). All were VUS, except for the

smaller of the 6q11.1 gains (Table 5.2), making the latter a less likely candidate region for

schizophrenia. The 12q21.31 CNVs overlap LIN7A, an intriguing candidate gene (Figure 5.2)

that interacts with DLG1, DLG2 and GRIN2B, and is implicated in postsynaptic density

functions (Kirov et al., 2012). There was also a 2.4 Mb gain at 15q12–q13.1 (a VUS), nested

within the typical 15q11–q13 duplication region, that overlapped just five genes, including three

gamma-aminobutyric acid (GABA) receptor component genes (Table 5.2; Figure 5.2) (Bassett,

2011). Other large rare CNVs in single case subjects overlapped CNVs previously reported in

schizophrenia or other neuropsychiatric disorders, involving genes of interest including RYR2,

MYT1L, LRP1B, IL1RAPL1 and ZNF804A (Table 5.2) (Kirov et al., 2012; Melhem et al., 2011;

Steinberg et al., 2011; Vrijenhoek et al., 2008; Xu et al., 2008). Interestingly, several genes we

had annotated as candidates were subsequently implicated in recent next-generation sequencing

studies of schizophrenia (Need et al., 2012; Xu et al., 2011) (Table 5.2).

5.3.4 Genome-wide CNV burden in schizophrenia

The proportion of subjects with one or more large (500 kb–6.5 Mb), rare (present in <.1% of

population controls) autosomal CNVs was significantly greater in the schizophrenia cases than in

the OPGP controls [62 of 420 versus 21 of 416; P < 0.0001, OR 3.26 (95% CI 1.95–5.45)]

(Table 5.3). As described above, this finding was driven by clinically significant CNVs and

VUS, and not benign variants. We a priori excluded 22q11.2 deletions and variants > 6.5 Mb

from these to demonstrate that this enrichment is robust to the exclusion of variants that might

reasonably be discovered without the use of chromosomal microarray technology. Case–control

differences were most marked for large rare exonic gains [38 of 420 versus 11 of 416; P <

0.0001, OR 3.66 (95% CI 1.85–7.27)] (Table 5.3). Within the schizophrenia sample, the

subgroup a priori designated as ‘syndromic’ using previously published clinical screening

criteria for adults (Fung et al., 2008; Silversides et al., 2012) was significantly enriched for

subjects with one or more large rare CNVs only when restricting to exonic losses [7 of 73

Page 97: Copy Number Variation in Neurodevelopmental Disorders

84

syndromic cases versus 12 of 347 non-syndromic cases; P = 0.0220, OR 2.96 (95% CI 1.12–

7.80)] (Table 5.3).

The schizophrenia group was also significantly enriched for subjects with two or more rare

CNVs > 10 kb in size that overlapped exons [122 of 420 versus 89 of 416; P = 0.0109, OR 1.50

(95% CI 1.10–2.06)] (Table 5.3). Further restricting to very rare CNVs (i.e., those found in none

of the 2357 adjudication controls) strengthened these results [78 of 420 versus 44 of 416; P =

0.0012, OR 1.93 (95% CI 1.30–2.87)]. In contrast, as expected (6,18,19), the overall CNV

profile (unrestricted, e.g., by rarity or size) was similar for schizophrenia cases and OPGP

controls. The results were also non-significant for subjects with one or more, or with two or

more, rare CNVs unrestricted by exonic status (Table 5.3). The results for all analyses were

similar if subjects of non-European ancestry were included.

5.3.5 Very rare, smaller CNVs identifying genes of interest in schizophrenia

Table 5.4 shows 15 candidate genes identified, independent of gene enrichment mapping, using a

strategy to determine very rare CNVs that overlapped the same gene in two or more

schizophrenia cases and in no controls. Five unrelated schizophrenia cases had loss CNVs

overlapping the RBFOX1 gene (previously A2BP1) (Melhem et al., 2011) and four had intronic

loss CNVs overlapping the SOX5 gene (Walsh et al., 2008). Arguably, the most persuasive of

candidates may involve genes of smaller genomic extent (< 200 kb), where CNVs overlap exons;

five genes met these conservative criteria: DNM1L, HIST3H3, JAK2, LIMS1 and PPP3CC.

Singleton very rare CNVs (Table 5.4) implicated other potential candidates, including DISC1,

GRK4, GRM4, PIK3C3 and RELN with overlapped exons, GRIK1, GRIN2A, and GRM7, with

introns overlapped only, and individual genes within large CNVs (Tables 5.1 and 5.2), e.g.,

IL1RAPL1 and TRPM1 (Melhem et al., 2011; Xu et al., 2008). Several of these candidates

(DISC1, IL1RAPL1, JAK2, PIK3C3, RELN and TRPM1) converged with functional gene

enrichment mapping results using loss CNVs (Figure 5.2). There was also convergence with

multiple genes implicated in two recent next-generation sequencing studies of schizophrenia

(Need et al., 2012; Xu et al., 2011), including 10 genes (CACNA2D1, GRIN2A, GRM7, JAK2,

NRXN1, PIK3C3, PTGER3, PTPRT, SLC1A1 and USH2A) implicated by multiple rare single

nucleotide variants (SNVs) (Need et al., 2012), and one gene (PTPRM) implicated by a de novo

rare SNV (Xu et al., 2011) (Table 5.4).

Page 98: Copy Number Variation in Neurodevelopmental Disorders

85

5.3.6 Functional networks revealed by pathway enrichment analysis

Figure 5.2 summarizes the pathway enrichment analysis results. For 70 (2.9%) of 2456 pathways

tested (Silversides et al., 2012), we found a significant enrichment of very rare, exonic loss

CNVs in schizophrenia cases compared with OPGP controls (OR 1.9, 95% CI 1.0–3.6),

involving 73 support genes (Figure 5.2). Considering just the three functional clusters

comprising eight pathways most clearly related to neurodevelopmental and synaptic processes

(Figure 5.2), the effect was greater (OR 4.7, 95% CI: 1.3–26.0), involving 19 support genes for

schizophrenia (Figure 5.2). The other clusters having a less obvious functional scope may

suggest more novel aspects of a gene network for schizophrenia (Figure 5.2). Notably, there was

a significant degree of functional cluster overlap in a formal cross-disease comparison of this

schizophrenia network with known genes for ASD (Figure 5.3). This included not only the eight

pathways annotated as neurodevelopmental and synaptic, but also novel clusters of pathway

enrichment analysis involved in cell motility, signal transduction and transcription (Figure 5.3).

Page 99: Copy Number Variation in Neurodevelopmental Disorders

86

Table 5.1. Very rare, clinically significant CNVs in unrelated schizophrenia probands

Subject CNV characteristics Candidate gene(s)

Case

# Catchment

area

Cytoband Start Size (bp) CN Flanking segmental

duplications

# of genes

Large, very rare CNVs <6.5 Mb in size, excluding 22q11.2 deletionsb (n=16 subjects)

2 ● 1q21.1 144,472,163 1,839,252 gGain ● 16

BCL9, GJA5, GJA8, PDZK1e,f, PRKAB2

3 ● 1q21.1 144,643,825 1,653,983 gGain ● 14

8 ● 2q13 111,105,101 1,727,363 Loss ● 10

ANAPC1, BCL2L11, MERTK

7c ● 2q13 111,105,101 1,727,363 Gain ● 10

9 2q13 111,105,101 1,727,363 Gain ● 10

13c ● 3q13.31 115,308,450 2,062,410 Loss 7 DRD3, GAP43, LSAMP, ZBTB20

17 ● 5p15.33-p15.32 1,864,574 3,687,431 Loss 9 IrxA cluster (IRX1, IRX2, IRX4), NDUFS6

40c ● 10q11.22-q11.23 45,905,767 5,423,684 Gain ● 52 CHAT, ERCC6, GDF2, GPRIN2, MAPK8,

SLC18A3

48 ● 15q11-q13 20,224,763 6,498,447 dGain ● 115

GABA receptor gene clusterf (GABRB3, GABRA5, GABRG3), MAGEL2, NDN, UBE3A

49 ● 15q11-q13 21,192,955 5,680,224 dGain ● 106

50 ● 15q11-q13 21,192,955 5,015,049 dGain ● 101

52 ● 15q13.2-q13.3 28,608,929 1,690,584 Loss ● 10 CHRNA7f, TRPM1f

Page 100: Copy Number Variation in Neurodevelopmental Disorders

87

55c ● 16p11.2 29,474,810 624,599 Gain ● 28

DOC2A, MAPK3, PRRT2, QPRT, SEZ6L2, TBX6

57c 16p11.2 29,474,810 624,599 Gain ● 28

56 ● 16p11.2 29,474,810 659,635 Gain ● 38

58 16p11.2 29,474,810 659,635 Gain ● 38

Very rare CNVs >6.5 Mb in size (n=8 subjects; one subject also appears above)

247c 6p25.3-p25.1 94,661 6,836,704 gLoss 41 FOXC1, GMDS, NRN1, TUBB2B

206 ● 7q22.2-q31.1 105,304,955 10,037,597 Loss ● 37 COG5, DOCK4, FOXP2, GPR85, IMMP2L,

LAMB1, NRCAM, PIK3CG, PNPLA8

115c ● 8p23.3-p23.1 148,062 6,828,275 gLoss ● 25 ANGPT2, CLN8, CSMD1f, DLGAP2, MCPH1

277c ● 19p13.3-p13.2 2,705,548 9,595,341 gGain 280 Various, including DNMT1, DOCK6

173 ● X chr (47, XXX) - - Gain - 975

Various, including IL1RAPL1f, SYN1

278 X chr (47, XXX) - - Gain - 975

57c X chr (47, XXY) - - Gain - 975

279c ● Y chr (47, XYY) - - gGain - 34 IL9R, SPRY3, SRY, TSPY, UTY, VAMP7

Case #, case number for subjects (n=454) with schizophrenia (n=2 were not of European ancestry: cases 278, 279); Catchment area, subjects originating from the only community mental health clinic in a specific catchment area of ~150,000 people (�), see text for details; Cytoband, cytogenetic location of CNV; CNV start, hg18 (NCBI Build 36.1, March 2006); CNV size, in base pairs; CN, type of copy number aberration; Flanking segmental duplication(s), known flanking segmental duplication(s) (from the UCSC Genome Browser hg18 version) that cover at least 20% of the CNV length (�); # of genes, number of known genes overlapped by CNV as annotated in the Database of Genomic Variants (http://projects.tcag.ca/variation/; September 2011); Candidate gene(s), selected based on reported neuropsychiatric/neurodevelopmental phenotype identified from systematic searches of human (e.g., Online Mendelian Inheritance in Man; http://www.omim.org/) and model organism (e.g., Mouse Genome Informatics; http://www.informatics.jax.org/) databases (for CNVs overlapping >250 genes, all genes were not systematically searched). aCNVs assessed independently by two clinical cytogenetic lab directors to be "Pathogenic" or of "Uncertain clinical significance; likely pathogenic" according to established guidelines (Kearney et al., 2011). bRecurrent 1.5 to 3 Mb 22q11.2 deletions (n=5, all meeting syndromic criteria) are not shown; of these, 2 are from the community catchment population, and 1 is not of European ancestry. Bolded cytoband indicates a novel association with schizophrenia cMet criteria for syndromic subgroup

Page 101: Copy Number Variation in Neurodevelopmental Disorders

88

dMaternal origin of 15q11-q13 gain per results from Multiplex Ligation-dependent Probe Amplification analysis to determine methylation status of the imprinted gene SNRPN eGene overlapped by 1.8 Mb gain in Case 2 only fGene(s) overlapped by one or more rare CNVs in other unrelated probands with schizophrenia in our sample: PDZK1, 826 kb exonic gain; GABA receptor gene cluster, 2.4 Mb exonic gain; CHRNA7, 54 kb exonic loss; TRPM1, 28 kb exonic gain (also implicated in a previous CNV study of schizophrenia (Kirov et al., 2009); CSMD1, 65 kb exonic, and 13 kb, 17 kb, 92 kb, and 117 kb intronic, losses (also implicated in previous CNV studies of schizophrenia (Kirov et al., 2012; Melhem et al., 2011); IL1RAPL1, 3.1 Mb exonic loss and 60 kb intronic loss (also implicated in a previous CNV study of schizophrenia (Melhem et al., 2011). gPreviously reported by our group: 1q21 gains (Dolcetti et al., 2013), 6p25 loss (Caluseriu et al., 2006), 8p23 loss (Bassett et al., 2010a), 19p13 gain (Bassett et al., 2010a), XYY (Bassett et al., 2010a).

Page 102: Copy Number Variation in Neurodevelopmental Disorders

89

Table 5.2. Large rare CNVs of uncertain pathogenicity in unrelated schizophrenia probands

Subject CNV characteristics Candidate gene(s)

Case

# Catchment

area

Cytoband Start Size (bp) CN Very rare

Flanking segmental

duplications

# of genes

VUS

1 1q21.1 143,677,127 826,295 Gain ● 21 ● PDE4DIPe, PDZK1c, PEX11B

4 1q43 235,349,594 726,228 Gain ● 1 ● RYR2b

5 ● 2p25.3 1,377,419 745,146 Gain ● 3 ● MYT1Lb

6 2q11.1 94,691,614 735,772 Gain ● ● 8 ● KCNIP3

10 ● 2q22.1 139,898,997 1,322,193 Loss ● 1 ● LRP1Bb,e

11 2q32.1 185,043,646 790,869 Loss ● 1 ● ZNF804Ac

12 3p12.1 83,940,054 1,038,151 Loss 1 ● -

14 ● 4q25-q26 113,634,704 783,442 Gain ● 10 ● ANK2e, NEUROG2

15 ● 4q33-q34.1 171,556,613 2,693,129 Loss ● 1 ● -

16 ● 4q34.3-q35.1 179,878,817 4,343,697 Gain ● 6 ● ODZ3

18 ● 5p15.1 15,134,588 628,993 Gain ● 1 ● -

19 5p13.2 37,294,917 520,350 Gain 2 ● -

20 5q13.3-q14.1 74,706,210 2,764,437 Gain ● 20 ● OTP, PDE8B, SV2Ce

22 ● 6q11.1 61,944,399 1,031,545 Gain ● 2 ● KHDRBS2, MTRNR2L9

23 6q11.1 61,944,399 502,986 Gain 1 MTRNR2L9

24 ● 6q14.3 86,804,893 577,156 Gain ● 0 -

25 ● 6q16.2-q16.3 98,969,135 1,408,033 Gain ● 9 ● POU3F2

26 ● 7p21.3 9,580,685 819,384 Loss 1 ● PER4

27 ● 7p21.3 9,573,572 1,080,196 Gain ● 1 ●

28 7q31.31 118,515,093 1,204,095 Gain 1 ● KCND2

29 ● 8p23.3 439,294 733,309 Gain ● ● 2 ● -

30 8p11.21-p11.1 43,132,979 772,981 Gain 2 HGSNAT

Page 103: Copy Number Variation in Neurodevelopmental Disorders

90

31 8q12.1 60,901,679 766,789 Gain ● 2 ● CA8, RAB2A

32 ● 8q21.3 89,566,045 1,073,381 Gain ● 0 -

33 9p21.1 30,791,693 549,623 Loss ● 0 -

34 9p13.3-p13.2 36,173,229 862,436 Loss ● 5 ● CLTA, GNE, PAX5e

35 9p12 41,440,153 535,035 Gain ● ● 6 -

36 ● 9p11.2 44,936,301 570,215 Gain ● ● 0 -

37 9q12 65,370,766 902,773 Loss ● ● 1 -

38 ● 9q21.11-q21.12 72,107,555 964,557 Loss ● 4 ● KLF9, TRPM3e

39 ● 10p11.22 33,636,052 651,356 Gain ● 1 ● NRP1

42 ● 11q22.1 97,224,044 626,127 Loss ● 0 -

43 ● 12p11.1 34,206,228 539,538 Gain 0 -

39 ● 12q21.31 79,255,527 568,008 Gain ● 6 ● LIN7A

44 ● 12q21.31 79,679,461 1,397,367 Loss ● 5 ●

45 13q13.1-q13.2 32,108,632 2,434,026 Loss ● 5 ● KLe, NBEAe, PDS5B

46 ● 13q21.32-q21.33 66,742,848 993,434 Gain ● 0 -

47 14q21.1-q21.3 40,497,844 2,771,925 Gain ● 1 ● LRFN5

51 ● 15q12-q13.1 23,766,255 2,437,700 Gain ● ● 5 ● GABRB3c, GABRA5c, GABRG3c

36 ● 15q13.2 28,153,539 564,254 Gain ● ● 7 CHRFAM7A

53 16p13.11 14,805,302 1,498,712 Gain ● 24 ● NDE1, NOMO3, NTAN1

54 16p13.11 15,032,942 1,388,198 Gain ● 20 ●

59 ● 16p11.1-p11.2 34,324,072 816,656 Gain 4 -

60 17p13.1 10,552,154 607,279 Gain ● 5 ● -

61 ● 19q12 32,648,879 2,916,074 Loss ● 10 ● C19orf12e, ZNF536

62 20p12.1 14,305,685 1,108,941 Loss ● 1 ● MACROD2b,c

63 ● 22q12.1 25,696,811 1,304,432 Loss ● 6 ● PITPNB

64d ● Xp21.3-p21.2 27,619,068 3,129,173 Loss ● ● 11 ● IL1RAPL1b,c

65 Xq11.1 63,283,656 721,753 Gain ● ● 3 ● -

66 ● Xq21.31 86,334,886 839,884 Gain ● 1 ● -

Page 104: Copy Number Variation in Neurodevelopmental Disorders

91

Case #, case number for subjects (n=454) with schizophrenia (n=6 were not of European ancestry: Cases 35, 37, 43, 60, 278, 279); Catchment area, subjects originating from the only community mental health clinic in a specific catchment area of ~150,000 people (�), see text for details; Cytoband, cytogenetic location of CNV; CNV start, hg18 (NCBI Build 36.1, March 2006); CNV size, in base pairs; CN, type of copy number aberration; Very rare, not found in 2,357 population controls and absent in OPGP controls (�), see text for details; Flanking segmental duplication(s), known flanking segmental duplication(s) (from the UCSC Genome Browser hg18 version) that cover at least 20% of the CNV length (�); # of genes, number of known genes overlapped by CNV as annotated in the Database of Genomic Variants (http://projects.tcag.ca/variation/; September 2011); VUS, variant of unknown significance (�) as assessed independently by two clinical cytogenetic lab directors (variants not so annotated were considered likely benign and not reportable clinically); Candidate gene(s), selected based on reported neuropsychiatric/neurodevelopmental phenotype identified from systematic searches of human (e.g., Online Mendelian Inheritance in Man; http://www.omim.org/) and model organism (e.g., Mouse Genome Informatics; http://www.informatics.jax.org/) databases. There were no such candidate genes identified for 11 genic CNVs. aFive large CNVs in schizophrenia cases of European ancestry that were present in the OPGP controls, using a 50% reciprocal overlap criterion are not shown: 3p14.2 gain in Case 263, 9q12 loss in Case 172, 11q11 gain in Case 41, and two 14q32.33 losses in Cases 156 and 280 bGene implicated in a previous CNV study of schizophrenia: RYR2 (Kirov et al., 2012), MYT1L (Vrijenhoek et al., 2008), LRP1B (Xu et al., 2008), ZNF804A (Steinberg et al., 2011), MACROD2 (Kirov et al., 2012), IL1RAPL1 (Melhem et al., 2011) cGene(s) overlapped by one or more additional rare CNVs in unrelated probands with schizophrenia in our sample: PDZK1, 1.8 Mb exonic gain (Table 5.1); GABA receptor gene cluster (GABRB3, GABRA5, GABRG3), 5.0 Mb, 5.7 Mb, and 6.5 Mb exonic gains (Table 5.1); MACROD2, 14 kb gain and 15 kb, 25 kb, and 63 kb losses; IL1RAPL1, X chromosome aneuploidies (Table 5.1) and 60 kb intronic loss dFemale subject, therefore judged to be a VUS instead of "Uncertain clinical significance; likely pathogenic" eGene implicated in a previous next-generation sequencing study of schizophrenia: PDE4DIP (Xu et al., 2011), LRP1B (Need et al., 2012), ANK2 (Need et al., 2012), SV2C (Need et al., 2012), PAX5 (Need et al., 2012), TRPM3 (Need et al., 2012), KL (Need et al., 2012), NBEA (Need et al., 2012), C19orf12 (Need et al., 2012).

Page 105: Copy Number Variation in Neurodevelopmental Disorders

92

Table 5.3. aRare autosomal CNV burden in 420 unrelated schizophrenia probands Schizophrenia cases vs. OPGP controls Schizophrenia cases

OPGP controls

All schizophrenia cases

Analysis Non-syndromic

cases Syndromic

cases Analysis

(n=416) (n=420) (n=347) (n=73) n (%) n (%) p OR (95% CI) n (%) n (%) p OR (95% CI)

Large rare CNVs (>500 kb)

Subjects with one or more large rare CNVs

Loss or gain 21 (5.05) 62 (14.76) <0.0001 3.26 (1.95) (5.45) 46 (13.26) 16 (21.92) NS 1.84 (0.97) (3.47) Exonic loss or gainc 17 (4.09) 57 (13.57) <0.0001 3.69 (2.11) (6.45) 42 (12.10) 15 (20.55) NS 1.88 (0.98) (3.61) Exonic lossc 7 (1.68) 19 (4.52) 0.0180 2.77 (1.15) (6.66) 12 (3.46) 7 (9.59) 0.0220 2.96 (1.12) (7.80) Exonic gain 10 (2.40) 38 (9.05) <0.0001 4.04 (1.98) (8.22) 30 (8.65) 8 (10.96) NS 1.30 (0.57) (2.97)

All rare CNVs (any size)

Subjects with one or more rare CNVs

Loss or gainb 372 (89.42) 376 (89.52) NS 1.01 (0.65) (1.57) 310 (89.34) 66 (90.41) NS 1.12 (0.48) (2.63) Exonic loss or gainc 249 (59.86) 264 (62.86) NS 1.14 (0.86) (1.50) 217 (62.54) 47 (64.38) NS 1.08 (0.64) (1.83)

All rare CNVs (any size)

Subjects with two or more rare CNVs

Loss or gainb 264 (63.46) 282 (67.14) NS 1.18 (0.88) (1.56) 232 (66.86) 50 (68.49) NS 1.08 (0.63) (1.85) Exonic loss or gainc,d 89 (21.39) 121 (28.81) 0.0134 1.49 (1.08) (2.04) 97 (27.95) 24 (32.88) NS 1.26 (0.73) (2.17)

aRare = CNVs with <0.1% prevalence in 2,357 population controls of European ancestry (used to adjudicate both schizophrenia cases and OPGP controls). Rare CNVs <10 kb or >6.5 Mb in size, 22q11.2 deletions and all sex chromosome CNVs were excluded CNV burden was measured using proportion of subjects with one or more, or with two or more, rare CNVs; measuring rare CNV burden (all sizes) for cases and controls using total number of genes overlapped by CNVs or total genomic extent of CNVs showed similar results. bResults were similarly non-significant for rare CNVs involving losses only and gains only, and for non-exonic losses and non-exonic gains cFive syndromic subjects of European ancestry in the schizophrenia group with typical 1.5 to 3 Mb 22q11.2 deletions (i.e., exonic losses) were not included in the study. NS = non-significant (p>0.05). Significant results are shown in bold font

Page 106: Copy Number Variation in Neurodevelopmental Disorders

93

Table 5.4. Candidate genes for schizophrenia overlapped by very rare (<500 kb) CNVs

Candidate gene Cytoband Gene size

(nt) Case CNV start CNV size

(bp) CN Exonic

Very rare CNVs overlapping the same CNS-related gene in two or more unrelated schizophrenia cases

CAMTA1 1p36.31 984,383 67 7,063,417 33,667 Loss �

68 7,545,052 44,523 Loss

DNM3 1q24.3 571,237 87 170,223,915 47,814 Loss �

168 170,450,307 4,136 Loss

HIST3H3 1q42.13 481 69 226,634,235 54,736 Gain �

70 226,673,588 109,976 Gain �

COMMD1 2p15 230,403 66 62,049,330 36,475 Loss �

32 62,106,806 24,384 Loss

LIMS1 2q13 152,892 71 108,619,802 61,009 Loss �

72 108,619,802 61,583 Loss �

DPP6 7q36.2 679,607

2 153,637,309 9,260 Loss

73 153,737,928 6,935 Loss

74 154,060,964 31,082 Loss �

PPP3CC 8p21.3 100,043 75 22,265,902 166,778 Gain �

76 22,265,902 166,778 Gain �

JAK2a,b 9p24.1 142,939 7 4,527,834 486,659 Loss �

77 5,014,345 82,361 Gain �

ERC1 12p13.33 504,696 78 865,867 157,163 Gain �

50 1,255,099 369,587 Loss �

SOX5a 12p12.1 1,030,150

79 23,826,011 11,461 Loss

80 23,828,090 9,382 Loss

81 24,063,813 25,105 Loss

82 24,561,569 35,937 Loss

DNM1L 12p11.21 66,448 83 32,679,398 45,218 Gain �

84 32,679,398 47,929 Gain �

RBFOX1a 16p13.2 1,694,209

54 6,171,253 32,382 Loss

85 6,754,460 50,650 Loss �

86 6,766,601 9,285 Loss

87 6,973,749 104,964 Loss �

62 6,992,360 143,753 Loss �

PRKCAa 17q24.2 507,937 61 61,941,341 277,949 Loss �

88 62,145,527 7,081 Loss

DOK6 18q22.2 448,040 89 65,614,643 10,824 Loss

90 65,639,534 10,209 Loss

Page 107: Copy Number Variation in Neurodevelopmental Disorders

94

PTPRTb 20q12 1,117,166

91 40,639,325 86,602 Loss

92 40,778,749 187,266 Gain �

93 40,887,741 7,643 Loss

Very rare CNVs overlapping promising candidate genes for schizophrenia found in a single case

PTGER3b 1p31.2 195,456 141 71,150,238 44,186 Loss �

USH2Ab 1q41 800,503 104 213,914,685 64,806 Loss �

DISC1 1q42.2 414,458 38 230,150,208 121,262 Loss �

NRXN1a,b 2p16.3 1,114,032 170 503,371,141 18,515 Loss

GRM7a,b 3p26.1 880,417 146 7,058,814 35,895 Loss

GRK4 4p16.3 77,132 100 2,994,521 31,371 Loss �

DOK7a 4p16.2 31,177 169 3,439,241 13,595 Gain �

FGF2 4q26 71,528 198 123,980,237 58,004 Loss �

GRM4 6p21.31 111,816 95 34,175,770 43,039 Gain �

RUNX2 6p21 222,766 231 45,129,193 298,662 Loss �

GABRR1 6q15 40,274 182 89,894,102 52,079 Loss �

CALN1a 7q11.22 632,885 223 71,331,369 131,729 Loss �

CACNA2D1b 7q21.11 493,614 166 81,829,583 187,332 Gain �

SEMA3A 7q21.11 236,559 195 83,380,795 87,642 Gain �

RELN 7q22.1 517,733 131 103,178,239 90,728 Loss �

SLC1A1a,b 9p24 97,043 7 4,527,834 486,659 Loss �

ABLIM1 10q25 253,546 200 116,351,384 17,427 Loss �

RHOG 11p15.4 14,006 188 3,817,748 11,292 Loss �

MTNR1B 11q21 13,160 240 92,336,224 10,613 Loss �

WNT5B 12p13.3 30,157 50 1,255,099 369,587 Loss �

ATN1c 12p13.31 17,859 208 6,837,052 74,565 Loss �

ENO2c 12p13.31 9,246 208 6,837,052 74,565 Loss �

GRIN2Ab 16p13.2 429,347 32 9,982,679 10,953 Loss

PTPRMa,b 18p11.2 839,546 119 7,899,131 55,764 Loss �

PIK3C3b 18q12.3 126,250 132 37,880,681 27,546 Loss �

AP3D1a 19p13.3 50,564 195 2,015,619 180,285 Loss �

JAK3 19p13.11 23,251 271 17,754,374 68,190 Gain �

GRIK1 21q21.3 403,029 166 30,082,515 11,811 Loss

S100B 21q22.3 6,505 181 46,843,667 16,998 Loss �

Page 108: Copy Number Variation in Neurodevelopmental Disorders

95

Candidate gene were selected based on reported neuropsychiatric/neurodevelopmental phenotype identified from systematic searches of human (e.g., Online Mendelian Inheritance in Man; http://www.omim.org/) and model organism (e.g., Mouse Genome Informatics; http://www.informatics.jax.org/) databases; Cytoband, cytogenetic location of candidate gene; Gene size, in nucleotides; Case, subjects from discovery sample (n=454) with schizophrenia; CNV start, hg18 (NCBI Build 36.1, March 2006); CNV size, in base pairs; CN, type of copy number aberration; Exonic, CNV overlaps exon(s) of candidate gene (�). aGene implicated in a previous CNV study of schizophrenia: JAK2 (Xu et al., 2008), SOX5 (Walsh et al., 2008), RBFOX1 (Melhem et al., 2011), PRKCA (Xu et al., 2008), NRXN1 (International Schizophrenia Consortium, 2008; Levinson et al., 2011; Melhem et al., 2011; Rujescu et al., 2009; Walsh et al., 2008), GRM7 (Walsh et al., 2008), DOK7 (Xu et al., 2008), CALN1(Stewart et al., 2011), SLC1A1 (Melhem et al., 2011; Rees et al., 2013; Stewart et al., 2011), PTPRM (Walsh et al., 2008), and AP3D1 (Melhem et al., 2011). bGenes implicated in previous exome sequencing studies of schizophrenia: JAK2 (Need et al., 2012), PTPRT (Need et al., 2012), PTGER3 (Need et al., 2012), USH2A (Need et al., 2012), NRXN1 (Need et al., 2012), GRM7 (Need et al., 2012), CACNA2D1 (Need et al., 2012), SLC1A1 (Need et al., 2012), GRIN2A (Need et al., 2012), PTPRM (Xu et al., 2011), PIK3C3(Need et al., 2012). cGenes overlapped by the same CNV.

Page 109: Copy Number Variation in Neurodevelopmental Disorders

96

Figure 5.2. Functional map of schizophrenia from pathway enrichment analysis

Network of 70 pathways (circles) related by mutual overlap (lines) is depicted using the Enrichment Map Cytoscape plugin (Merico et al., 2010). Circle size is proportional to the total number of ‘support’ genes in each pathway and line thickness represents the number of genes in common between two pathways. Each support gene harbors a rare exonic loss CNV in one or more schizophrenia subjects but in no OPGP controls; only CNVs overlapping 15 or fewer genes were tested for pathway enrichment analysis. Groups of functionally related pathways are represented by the filled circles and respective labels; grey circles indicate pathways not assigned to these 13 clusters. Three major functional groups are further highlighted by darker filled circles (blue, Synapse; pink, Nervous system development; red, Axonogenesis) with the respective 19 support genes for schizophrenia shown in color. We also list (in black) other genes in these three functional clusters that were overlapped by very rare loss or gain CNVs in subjects with schizophrenia (i.e., not in 2773 controls), and that have additional supportive data.

Page 110: Copy Number Variation in Neurodevelopmental Disorders

97

Figure 5.3. Overlap of schizophrenia pathway analysis with known ASD genes

Functional map of schizophrenia from Figure 5.2 is presented with highlighting of the clusters of the pathways significantly associated to schizophrenia for which at least half the pathways were also significantly over-represented in known ASD genes (Betancur, 2011).Over-representation was tested using Fisher's Exact Test (FET) and Benjamini-Hochberg FDR for multiple test correction, with a significance threshold of 5% FDR.

Page 111: Copy Number Variation in Neurodevelopmental Disorders

98

5.4 Discussion

Derived from a clinically informed and rigorous approach, the results of this study provide new

information on the genomic architecture of schizophrenia and on the potential role of clinical

microarray testing. Adding novel data to those available from the component studies of existing

consortia reports, this study involved (i) unparalleled systematic recruitment efforts in a

community setting (Bassett et al., 2010a), (ii) consistent phenotyping of all cases using

previously published methods (Fung et al., 2008; Silversides et al., 2012), (iii) rigorous CNV

methodology (Lionel et al., 2011), (iv) exclusion of already known effects of 22q11.2 deletions,

(v) blinded adjudication of rare CNVs by two independent clinical cytogenetic laboratory

directors and (vi) clinical confirmation and return of results to all patients with clinically

significant CNVs and their families. The results strengthen etiologic and mechanistic ties of

schizophrenia to developmental disorders like autism, where clinical microarray testing is now

indicated and suggest a yield of reportable findings approaching that seen in these disorders

(Miller et al., 2010a; Shen et al., 2010). Rare, recurrent structural rearrangements at 2q13 were

implicated for the first time as significant risk factors for schizophrenia. We provide strong

evidence supporting a quantitatively greater, and qualitatively different, genome-wide burden of

rare CNVs in schizophrenia compared with controls. New candidate genes and networks relevant

to schizophrenia identified in this study will facilitate the future clinical interpretation of smaller

CNVs.

The results of this study indicate that ~1 in 13 unrelated patients with schizophrenia may harbor

a clinically significant large rare structural variant that would be detectable using methods

available in most clinical laboratories. This yield is similar to that seen in a 2010 clinically

driven study of genetic testing in ASD (Shen et al., 2010), a related neurodevelopmental

condition in which clinical microarray testing is now considered the first-tier diagnostic test

(Miller et al., 2010a). None of the patients in the schizophrenia sample studied had autism or

multiple congenital anomalies and few would meet the existing criteria for clinical microarray

testing, e.g., on the basis of mild intellectual disability. Our data also suggest that most

individuals with schizophrenia who have large rare CNVs would not have obvious

developmental (syndromal) features, with the exception of some with large exonic losses where

phenotypes tend to be more complex, as in 22q11.2 deletions. Thus, to discover most of the

Page 112: Copy Number Variation in Neurodevelopmental Disorders

99

clinically significant variants identified in this study a genome-wide microarray would be

necessary. Introduction of such genetic testing into clinical practice for an adult neuropsychiatric

disease would have implications for psychiatry and for the still pediatric-focused field of medical

genetics. As for ASD and developmental delay/intellectual disability, the availability of clinical

CNV data for schizophrenia, coupled with an improved ability to interpret these data for the

benefit of patients and families, could have scientific implications and clinical benefits as

discussed below. Further studies are needed to demonstrate the worth of such clinical genetic

testing in schizophrenia.

What are the potential benefits of identifying a ‘clinically significant’ CNV? The impetus to

offer clinical microarray testing to individuals with schizophrenia in the hope of arriving at a

partial etiologic explanation is similar to that described for other neurodevelopmental conditions

(Miller et al., 2010b). Uncertainty about penetrance and expressivity of rare variants is a

common reality of practice in medical genetics, and remains compatible with the provision of

helpful genetic counseling. With the exception of 22q11.2 deletions, most clinically significant

variants in individuals with schizophrenia today would only be discovered on a research basis.

More needs to be known about lifetime, including adult, expression of these CNVs to better

inform medical and/or psychiatric management. Nevertheless, all of these large, rare, clinically

significant variants already have implications for genetic counseling that include, but are not

limited to, reproductive implications for siblings and patients themselves (Costain et al., 2011),

and are of potential benefit to patients and families (Costain et al., 2014a; Costain et al., 2014b).

Our own efforts with this translation to clinical genetic counseling are ongoing, with an overall

positive response to date, comparable with our experience with 22q11.2 deletion syndrome

(Costain et al., 2012) and schizophrenia in general (Costain et al., 2014a; Costain et al., 2014b).

There may be, therefore, a growing obligation to offer the return of selected, clinically relevant

and clinically confirmed individual research results to patients and their families in large-scale

genome-wide studies of schizophrenia such as this (Costain and Bassett, 2013). To our

knowledge, this is currently a rare practice, even for 22q11.2 deletions.

The new candidate genes and network highlighted in this report will facilitate the clinical

interpretation of genome-wide CNVs in schizophrenia, including CNVs that are neither large nor

recurrent. We took a conservative approach to pathway enrichment analysis, using only exonic

losses in the analyses and excluding major multi-genic CNVs. The results showed that the genes

Page 113: Copy Number Variation in Neurodevelopmental Disorders

100

implicated by rare CNVs in schizophrenia are especially overrepresented by those with synaptic

and neurodevelopmental functions, consistent with other studies (Kirov et al., 2012; Walsh et al.,

2008). Moreover, the findings extended attention from such well-recognized candidates to less

obvious genes. Importantly, these were not restricted to genes with multiple functions (e.g.,

DISC1 and RELN). The pathways implicated in this study may assist with the interpretation of

smaller genic CNVs in schizophrenia. This more complex gene network for schizophrenia was

similar to that seen in autism (Pinto et al., 2010). For the first time, a formal analysis has

demonstrated that in schizophrenia there is significant over-representation of known genes for

ASD (Figure 5.3). This finding provides additional evidence in support of using data from ASD

to assist in the clinical interpretation of CNVs in schizophrenia, and vice versa. The fact that

none of the study participants had a history of autism or moderate to severe intellectual disability

supports the likelihood that CNVs shared with these pediatric conditions exhibit true pleiotropy.

Many of the implicated genes in this study show regional gene expression differences in mid-

fetal development of the brain (e.g., CHRNA7 and LIN7A in the thalamus) (Johnson et al., 2009),

a critical time for the pathogenesis of schizophrenia (Insel, 2010). We note, however, that the

prolonged maturation of the human brain provides multiple opportunities for perturbation of

gene expression through gene dosage changes.

Complementary to these results, we demonstrated enrichment in schizophrenia of not only

individuals with large rare exonic CNVs, but also those with two or more rare CNVs >10 kb that

overlapped exons. These results represent some of the first concrete evidence supporting a

‘multiple hit’ hypothesis for schizophrenia, outside of multi-genic large CNVs (Bassett et al.,

2010a; Kirov et al., 2012; Walsh et al., 2008). There is also support for regulatory (e.g., large

intronic CNVs) and epigenetic mechanisms (e.g., chromatin modification cluster in the pathway

enrichment analysis) as important contributory factors to the etiology of schizophrenia. There

remained such evidence after excluding the potentially overpowering influence of multi-genic

CNVs. The new candidate genes and network for schizophrenia identified in this report may also

help guide the future interpretation of results from whole exome and whole genome sequencing

studies (Need et al., 2012; Xu et al., 2011). Indeed, there was a notable degree of overlap

between the promising candidate genes overlapped by rare CNVs in our sample (Tables 5.2 and

5.4) and those genes containing rare SNVs in schizophrenia cases in recent next-generation

sequencing studies (Need et al., 2012; Xu et al., 2011).

Page 114: Copy Number Variation in Neurodevelopmental Disorders

101

This is the first study of genome-wide CNVs in schizophrenia to include clinical adjudication of

identified rare CNVs and to use a systematic clinically informed genetic approach to investigate

a well-characterized cohort of adults assessed a priori for certain ‘syndromic’ features (Bassett et

al., 2010a; Fung et al., 2008; Silversides et al., 2012). To minimize the potential impact of false

positives, we applied identical molecular and conservative analytic methods to unrelated cases

with schizophrenia and to a similar-sized Canadian sample of epidemiologic controls, an

unprecedented approach in neuropsychiatric disease recently validated in the study of congenital

cardiac disease (Silversides et al., 2012). The strategy emphasized CNVs likely to have an

enhanced effect size because we used a more restricted level to define rarity (< 0.1%) than that

usually employed (< 1%) (International Schizophrenia Consortium, 2008; Levinson et al., 2011;

Melhem et al., 2011). This could mean missing more common variants with a lower effect size.

Using an independent population-based control set ensured equal effects for both cases and

OPGP controls, however, and thus all comparative analyses (including the burden and gene

enrichment mapping) are robust to the potential effects of systematic Type 1 or 2 errors.

Published guidelines for CNV analysis support our multi-algorithm approach (Lionel et al.,

2011), and we report a very high positive validation rate by independent methods (100% of 58

CNVs tested, including 17 smaller than 50 kb in size) that is in keeping with our previous

experience (Lionel et al., 2011; Marshall et al., 2008; Pinto et al., 2010; Silversides et al., 2012).

All studies will be affected by the cases and controls studied, and the controls used to adjudicate

rarity of CNVs. We acknowledge a focus on individuals of European ancestry. No genome-wide

CNV results of our cases have been previously published, so there will be no sample overlap

with other datasets.

Adjudicating findings for individually rare variants is challenging—indeed, this was a major

rationale for our study and was performed through blinded assessment of large rare CNVs by

clinical laboratory directors. We also employed conservative qualitative and quantitative

methods in order to identify putative candidate genes for schizophrenia overlapped by rare CNVs

whose expression may be affected by the associated gene dosage changes. We report the details

of a gene network derived from rare CNVs overlapping exons that differentiated schizophrenia

cases from controls (Figure 5.2) to facilitate interpretation of findings of future studies. The

network was further validated by demonstrating significant overrepresentation with known genes

for ASD (Figure 5.3). Nonetheless, all clinical and research interpretations are constrained by

Page 115: Copy Number Variation in Neurodevelopmental Disorders

102

knowledge available at the time. We expect that understanding of coding regions, and non-genic

and regulatory sequences, will continue to evolve. Although the number of cases in our study

appears to be small in comparison to the combined numbers reported by large consortia, we were

sufficiently powered to (i) arrive at a confidence interval (CI) for the prevalence of clinically

significant CNVs in schizophrenia with a lower bound still >5%, and (ii) demonstrate within-

sample recurrence implicating individual genes (Table 5.4) and specific loci (Tables 5.1 and 5.2),

including the novel discovery of enrichment of CNVs at 2q13 in schizophrenia.

Our results provide clinical adjudication of previously reported CNVs for schizophrenia (Bassett

et al., 2010b; Kirov et al., 2012; Levinson et al., 2011; Melhem et al., 2011; Walsh et al., 2008).

Family studies, rare in schizophrenia, will be essential to delineate the inheritance status,

penetrance and variable neuropsychiatric expression, segregation pattern and extended

phenotype associated with those CNVs believed to underlie emerging genetic subtypes of the

disorder. The exception is 22q11.2 deletion syndrome, where these are well established (Costain

and Bassett, 2012). De novo status is neither a necessary nor a sufficient criterion for

pathogenicity, and many of the CNVs highlighted in this study may have been inherited. Many

well-established schizophrenia risk variants of major effect are found in the majority of cases to

be inherited (Sahoo et al., 2011), usually from parents with mild or no overt neuropsychiatric

phenotype (Cooper et al., 2011; Costain et al., 2011). We elected to use previously validated

criteria (Bassett et al., 2010a; Fung et al., 2008; Silversides et al., 2012) for assessing syndromic

features, shown to (i) have a good discriminant ability to identify at least one genomic disorder

(22q11.2 deletion syndrome) and (ii) be feasible for administration in a brief clinical encounter—

a key prerequisite for translation into the psychiatric clinic. There may exist more complex

clinical screening criteria with sufficient discriminant ability to identify patients with large, rare,

clinically significant CNVs. The collectively high prevalence of such variants, and the non-

significant findings for clinical variables in this study, however, support the potential universal

use of clinical microarray testing in schizophrenia. Definitively proving causality of specific

genetic variants for schizophrenia is beyond the scope of our and other genome-wide CNV

studies. Ideally, this would comprise direct functional evidence that can be related to gene

dosage effects in the developing and adult human brain. For many candidate genes implicated by

these CNV results, the functional significance has already been validated in model organisms

and/or human cellular models, as outlined in Tables 5.1 and 5.2. Such models are useful in a

Page 116: Copy Number Variation in Neurodevelopmental Disorders

103

disease where access to the target organ at precise developmental time points is challenging.

Why does the prevalence of large, rare, recurrent CNVs in schizophrenia appear to be higher in

this study than estimates reported in previous studies (Levinson et al., 2011)? First, this is a

prospective study using a consistent ascertainment, recruitment and assessment strategy for

patients with chronic schizophrenia at community clinics. Although demographic features and

age at onset were similar to those in other studies, the chronicity of the illness may have meant

that a relatively few individuals with mild forms of schizophrenia and/or evolving psychiatric

diagnoses were included. Second, although all studies may be expected to have volunteer bias,

we took an inclusive approach to recruitment. We did not exclude individuals, e.g., with learning

difficulties or residual symptoms, and obtained a surrogate consent for individuals where

necessary. Also, recruitment has persisted over several years, and particularly in our catchment

based sample we have made every effort to include all patients (e.g. waiting until their clinical

state improved, making home visits). We did not recruit from university health clinics, use

groups of patients already enrolled in and/or volunteering for other studies or preferentially

enroll patients from intact families with both parents available for study or those with familial

schizophrenia. Such factors can affect the observed prevalence of 22q11.2 deletions (Bassett et

al., 2010b) and likely that of other rare CNVs with high penetrance for severe neuropsychiatric

phenotypes. Nevertheless, given evidence for premature death in some CNV-related conditions

(Bassett et al., 2009), we would have missed individuals who died before enrollment and those

too ill to participate. Replication of our findings in an independent cohort should be a key goal of

future research in this area, and our results may encourage others to adopt our clinically driven

approach, including more epidemiologically-based strategies for patient ascertainment.

In conclusion, for the first time, we have a reliable estimate of the collective prevalence of

clinically significant CNVs in a community-based sample of schizophrenia. Most individuals

with large rare CNVs, including many of those with clinically significant CNVs, will not be

readily distinguishable from the rest of the general schizophrenia population based on

developmental (syndromic) features. The prevalence of clinically significant CNVs discovered in

community-based schizophrenia brings us a step closer to recommending clinical microarray and

other methods of genetic testing (including sequencing) for patients with schizophrenia, and may

initiate a debate among clinicians, scientists, families and policy makers. As for children with

developmental disorders found to have clinically significant CNVs, clinical genetic counseling

Page 117: Copy Number Variation in Neurodevelopmental Disorders

104

could be provided to patients with schizophrenia and their families (Costain and Bassett, 2012),

regardless of the acknowledged need for more information on key genetic parameters, such as

penetrance and variable expression across the lifespan. The complex network of genes being

discovered through studies of rare variants mirrors the complexity of the neural networks

involved. Further research efforts promise to bring order to these for schizophrenia and related

NDDs, as they have for cancer.

5.5 Materials and Methods

5.5.1 Schizophrenia sample ascertainment and assessment

We prospectively recruited 459 unrelated Canadian patients meeting DSM-IV criteria for chronic

schizophrenia or schizoaffective disorder from four community mental health clinics. The

majority (n = 248) comprised a ‘catchment sample’ that originated from the only community

mental health clinic in a catchment area of ~150 000 people (Bassett et al., 2010a). For the

current study, we excluded the five subjects with 22q11.2 deletions with genome-wide CNV data

published elsewhere (Bassett et al., 2008), two of whom were from the catchment area sample

(2). None of the remaining 454 subjects [317 (69.8%) male; mean age 44.8 (SD 12.1) years] had

previously published genome-wide CNV data, though four had previously published abnormal

karyotype results (Bassett et al., 2010a; Caluseriu et al., 2006) and two with typical 1q21.1 gains

were recently reported by us in a case series (Dolcetti et al., 2013). After description of the study

to the subjects, a written informed consent was obtained. The study was approved by local

hospital and university institutional review boards.

All subjects underwent direct clinical screening assessments for potential syndromic features

using a standardized protocol that included a review of available lifetime medical records and

assessment of physical features (Bassett et al., 2010a; Fung et al., 2008; Silversides et al., 2012).

Using the criteria previously validated for identifying adults with 22q11.2 deletion syndrome

[two or more of: global dysmorphic facial features, history of learning difficulties, abnormal

(hypernasal) voice (17)], syndromic [n = 78 (17.2%)] and non-syndromic subgroups were

designated (Silversides et al., 2012). No individual had a pediatric diagnosis of moderate or

severe intellectual disability, multiple major congenital anomalies or autism. All phenotyping

was done blind to genotype. The age at onset was considered the age at first treatment for

Page 118: Copy Number Variation in Neurodevelopmental Disorders

105

psychosis [mean 22.9 (SD 6.8) years], and a positive family history defined as having one or

more first degree relatives with a psychotic disorder [n = 70 (16.8%) of 416].

5.5.2 Control sample for formal analyses

To optimize our analyses, we used an independent Canadian control sample from the OPGP

genetic epidemiologic project comprising 416 unrelated adults [208 (50.0%) male; mean age

45.0 (SD 12.1) years] of European ancestry (Silversides et al., 2012). The OPGP controls are

independent of the 2357 population-based CNV adjudication controls described below and in

Chapter 2. The OPGP was conceived as an opportunity to establish a collection of fully

consented, Ontario based population control DNA samples, as a resource for researchers in

Ontario and elsewhere. The OPGP was constructed in two phases. The first phase involved

collaborations with the Ontario Familial Breast Cancer Registry (OFBCR) (Andrulis et al., 1997)

and the Ontario Familial Colorectal Cancer Registry (OFCCR) (Cotterchio et al., 2000).

Population controls from the OFBCR and OFCCR that had completed the study questionnaires

and provided a blood sample were eligible to participate as controls in the OPGP. The second

phase involved collaboration with the Institute of Social Research at York University (Toronto,

ON, Canada). A database of information on individuals from the population of Ontario (n = 14

000) was obtained, based on a random sample of households across Ontario identified initially

on the basis of telephone directory listings. These households were contacted by means of a

mailed letter and follow-up telephone calls were made by a trained interviewer who, after

establishing eligibility within the household, randomly chose one person as the eligible control.

Individuals were eligible if they were 20–79 years of age and resided in Ontario. In both the

phases, subjects were re-consented for OPGP, and surveyed for comprehensive demographic and

health information. Blood samples were collected and sent via courier to The Centre for Applied

Genomics (TCAG) Biobanking Facility for transformation, DNA preparation and distribution.

The total OPGP collection consists of samples from 1134 males (42%) and 1556 females (58%),

ranging in age from 20 to 82 years. This study uses data from 416 of these OPGP samples of

European ancestry that were genotyped on the Affymetrix 6.0 platform. To minimize laboratory-

related artefacts, identical experimental and analytical protocols for array genotyping and CNV

analyses were followed at the same facility for the OPGP control and schizophrenia case samples

using the methodology described in Chapter 2. Each CNV identified in the schizophrenia and

OPGP cohorts was then adjudicated for rarity by comparison to those identified, using an

Page 119: Copy Number Variation in Neurodevelopmental Disorders

106

identical microarray platform and CNV analysis strategy, in two large population-based control

cohorts comprising 2357 individuals of European ancestry from Ontario and Germany

(Krawczak et al., 2006; Stewart et al., 2009). We used a conservative definition of ‘rare’ CNVs:

those present in < 0.1% of the population controls (Silversides et al., 2012).

5.5.3 CNV prioritization, assessment of pathogenicity and prevalence

We prioritized two main groups (Figure 5.1) of rare (present in < 0.1% of population-based

controls) CNVs in the schizophrenia cases for qualitative analysis: (i) large CNVs (i.e., those

defined as > 500 kb in size) and (ii) CNVs < 500 kb in size that were ‘very rare’ (i.e., not present

in the population-based control cohorts, using a 50% reciprocal overlap criterion) (Lionel et al.,

2011), and that overlapped the same CNS-related gene(s) in two or more unrelated schizophrenia

cases (Silversides et al., 2012). We also identified other very rare CNVs that overlapped

promising candidate genes for schizophrenia (as determined by known biological function and/or

previous reports), but that were found only in a single schizophrenia case.

Each large (> 500 kb) rare CNV in schizophrenia case and OPGP control individuals was

assessed independently by two experienced clinical cytogenetic laboratory directors blind to

case–control status. According to ACMG guidelines for interpretation of postnatal constitutional

CNVs (Kearney et al., 2011), each CNV event was assigned to one of the five main categories:

‘Pathogenic’, ‘Uncertain clinical significance; likely pathogenic’, ‘Uncertain clinical significance

(no sub-classification)’, ‘Uncertain clinical significance; likely benign’ and ‘Benign’.

Disagreements between the two laboratory directors in initial interpretations were subsequently

resolved through discussion to arrive at a consensus determination (while still blind to case-

control status). In the text, variants deemed to be ‘Pathogenic’ or of ‘Uncertain clinical

significance; likely pathogenic’ were collectively termed ‘clinically significant’, and those of

‘Uncertain clinical significance (no sub-classification)’ were termed ‘variants of unknown

significance’ (VUS); the remainder were termed ‘benign’.

We used our schizophrenia community catchment sample to estimate the minimum prevalence

of large, rare, clinically significant structural variants that would be detectable on a clinical

microarray. We estimated that there were 370 unrelated adults with schizophrenia in this area,

including the 248 consecutive subjects for whom genome-wide CNV data were available

(67.0%), as well as individuals who refused or were unable to consent to participation in the

Page 120: Copy Number Variation in Neurodevelopmental Disorders

107

study, and others known to but not attending the clinic or hospital system; all non-participants

were assumed to be unrelated to probands in this study or to each other.

5.5.4 Experimental validation of CNVs

Confirmatory studies of possible schizophrenia-associated CNVs used Stratagene SYBR Green

based qPCR. Each qPCR assay was performed simultaneously in triplicate for both the test

region and a control region of known copy number 2 on chromosome 7. The ratio of the average

value for the test region to that for the control region had to be > 1.3 or < 0.7 in order for the

CNV to be confirmed as a duplication or deletion, respectively. In addition, the standard error of

the ratio had to be < 1.0 on the same scale in order for the assay to be considered reliable. We

validated 58 (100.0%) of 58 CNVs tested, i.e., all 58 assays met both of these criteria for CNV

confirmation. The validated CNVs spanned a range of sizes, and included 17 that were < 50 kb.

Post-study molecular genetic results from clinical laboratories confirmed a further 11 CNVs, and

also the parental origin of the 15q11-q13 duplications. All clinically significant variants were

confirmed in a CLIA-approved laboratory and returned to patients and their families.

5.5.5 CNV burden analyses

For the CNV burden analyses, we compared the proportion of subjects of European ancestry

with one or more rare CNVs in the 420 [n = 291 (69.3%) male] unrelated schizophrenia cases

with that in the 416 [n = 208 (50.0%) male] unrelated OPGP controls (Silversides et al.,

2012).We used only autosomal CNVs because of the significant sex differences between these

groups (χ2 = 32.31, df = 1, P < 0.001). As an additional conservative step, variants that might

reasonably be discovered without use of a genome-wide microarray, i.e., rare CNVs > 6.5 Mb in

size potentially detectable by routine G-band analysis (cases of European ancestry: n = 6,

controls: n = 0; Table 5.1) and 22q11.2 deletions (cases of European ancestry: n = 4, controls: n

= 0), were not used in these formal burden calculations. Statistical analyses were performed

using SAS software. Comparisons between schizophrenia cases and OPGP controls, and between

syndromic and non-syndromic schizophrenia cases, were performed using standard two-sided

statistical tests. Odds ratios and 95% confidence intervals were calculated using standard

methods. A P-value of < 0.05 was considered to be significant. As described above, identical

experimental and analytical protocols for array genotyping and CNV analyses were followed at

the same facility for the schizophrenia case and OPGP control samples, and each CNV was then

Page 121: Copy Number Variation in Neurodevelopmental Disorders

108

adjudicated for rarity by comparison to those identified, using an identical microarray platform

and CNV analysis strategy, in two large population-based control cohorts distinct from the

OPGP controls. Thus, the rate of Type 1 and 2 errors would be expected to be the same in

schizophrenia cases and OPGP controls, with no resulting impact on the relative inter-group

CNV burden.

5.5.6 Functional network from gene association analysis

To identify biologically meaningful groups of genes (i.e., those sharing a common function)

affected by CNVs using the same samples as for the main burden analyses, we used a pathway

enrichment analysis similar to that previously published for ASD (Merico et al., 2010; Pinto et

al., 2010) and congenital cardiac disease (Silversides et al., 2012). Briefly, we first compiled

comprehensive collections of pathways (Silversides et al., 2012) and for each functional

pathway, we tested whether the prevalence of subjects with very rare CNVs was higher in

schizophrenia cases than in OPGP controls using a one-tailed Fisher’s exact test (FET) (Pinto et

al., 2010; Silversides et al., 2012). Nominal P-values were corrected for multiple testing using a

case– control class permutation procedure to estimate an empirical false discovery rate (FDR)

(Pinto et al., 2010; Silversides et al., 2012). Case–control labels were randomly permuted 5000

times and, for each P-value, the empirical FDR was computed as the average number of

pathways with equal or smaller P-value over permutations. We selected 10% as the empirical

FDR significance threshold for final results. As above, the rate of Type 1 and 2 errors would be

expected to be the same in schizophrenia cases and OPGP controls, with no resulting impact on

the pathway enrichment analysis. Finally, we highlighted the clusters of the pathways that were

significantly associated with schizophrenia where for at least half of the pathways there was also

significant overrepresentation of known ASD genes (Betancur, 2011). Over-representation was

tested using FET and Benjamini–Hochberg FDR for multiple test correction, with a significance

threshold of 5% FDR.

Page 122: Copy Number Variation in Neurodevelopmental Disorders

109

Chapter 6

Gephyrin Deletions in Cross-Disorder Risk for Autism,

Schizophrenia and Seizures

Parts of this chapter are adapted, with permission for use from Oxford Press, from the following

published journal article:

Lionel AC, Vaags AK, Sato D, Gazzellone MJ, Mitchell EB, Chen HY, Costain G, Walker S,

Egger G, Thiruvahindrapuram B, Merico D, Prasad A, Anagnostou E, Fombonne E,

Zwaigenbaum L, Roberts W, Szatmari P, Fernandez BA, Georgieva L, Brzustowicz LM, Roetzer

K, Kaschnitz W, Vincent JB, Windpassinger C, Marshall CR, Trifiletti RR, Kirmani S, Kirov G,

Petek E, Hodge JC, Bassett AS, Scherer SW. Rare exonic deletions implicate the synaptic

organizer Gephyrin (GPHN) in risk for autism, schizophrenia and seizures. Human Molecular

Genetics; 2013; 22(10):2055-66.

I performed CNV analysis of the ASD and schizophrenia cases genotyped at TCAG and

discovered the GPHN deletions, performed CNV analysis of controls, coordinated collection of

genetic data from the other clinical sites and drafted and revised the manuscript. Dr. A.K. Vaags

coordinated collection of clinical phenotype data from the individuals with GPHN deletions.

Page 123: Copy Number Variation in Neurodevelopmental Disorders

110

6.1 Abstract

The GPHN gene codes for gephyrin, a key scaffolding protein in the neuronal postsynaptic

membrane, responsible for the clustering and localization of glycine and GABA receptors at

inhibitory synapses. Gephyrin has well-established functional links with several synaptic

proteins that have been implicated in genetic risk for NDDs such as Autism Spectrum Disorder

(ASD), schizophrenia and epilepsy including the neuroligins (NLGN2, NLGN4), the neurexins

(NRXN1, NRXN2, NRXN3) and collybistin (ARHGEF9). Moreover, temporal lobe epilepsy

arising due to cellular stress such as temperature and alkalosis associated with epileptogenesis

has been linked to abnormally spliced GPHN mRNA lacking exons encoding the G-domain of

the gephyrin protein. This chapter describes the clinical and genomic characterization of six

unrelated subjects, with a range of neurodevelopmental diagnoses including ASD, schizophrenia

or seizures, who possess rare de novo or inherited hemizygous microdeletions overlapping exons

of GPHN at chromosome 14q23.3. The region of common overlap across the deletions

encompasses exons 3-5, corresponding to the G-domain of the gephyrin protein. These findings,

together with previous reports of homozygous GPHN mutations in connection with autosomal

recessive molybdenum cofactor deficiency, will aid in clinical genetic interpretation of the

GPHN mutation spectrum. This data also add to the accumulating evidence implicating neuronal

synaptic gene products as key molecular factors underlying the etiologies of a diverse range of

neurodevelopmental conditions.

6.2 Introduction

GPHN encodes the protein gephyrin, a key scaffolding molecule in the neuronal postsynaptic

membrane at inhibitory synapses (Fritschy et al., 2008; Tretter et al., 2012). After its initial

discovery by affinity chromatography experiments as co-purifying with the postsynaptic

inhibitory glycine receptor, the protein was named “gephyrin” from the Greek term for “bridge”,

in line with its proposed activity as an intermediary molecule connecting neurotransmitter

receptors to the postsynaptic microtubule cytoskeleton (Prior et al., 1992). Follow-up

experiments demonstrated that the clustering and postsynaptic localization of both major types of

inhibitory receptors (glycinergic and GABAergic) depend on direct protein-protein interactions

of gephyrin with subunits of the glycine receptor (Kirsch et al., 1993) and the GABAA receptor

Page 124: Copy Number Variation in Neurodevelopmental Disorders

111

(Essrich et al., 1998). Gephyrin clustering is essential for GABAergic synapse stability and

aberrant GPHN expression has been observed in the hippocampus and adjacent neocortex of

patients with temporal lobe epilepsy (TLE) and in a rat model of TLE (Fang et al., 2011). These

findings are in line with known deficits in GABAergic synaptic transmission in TLE (Kumar and

Buckmaster, 2006). One potential molecular explanation for gephyrin’s involvement in the

epileptic hippocampus could be abnormally spliced forms of RNA produced by cellular stress-

induced exon skipping, which can lead to altered or inactive protein (Forstera et al., 2010).

However, there have been no reports of genomic GPHN mutations in patients with epilepsy.

In addition to its potential involvement in epilepsy, gephyrin has functional links with several

synaptic proteins, mutations of which have been reported in a range of NDDs. Together with

collybistin, gephyrin forms complexes with the postsynaptic neuroligin proteins NLGN2

(Poulopoulos et al., 2009) and NLGN4 (Hoon et al., 2011). Gephyrin’s specific postsynapatic

localization and clustering are mediated by members of the presynaptic neurexin protein family

(Kang et al., 2008). This complex interplay of gephyrin with the neurexins and neuroligins not

only attests to its importance in the proper formation and function of synapses, but is also of

particular interest given the substantial evidence implicating these trans-synaptic adhesion

molecules in genetic risk for a range of neurodevelopmental conditions (Reichelt et al., 2012;

Sudhof, 2008) including Autism Spectrum Disorder (ASD), schizophrenia and epilepsy as

detailed in Chapter 1. Of specific relevance to gephyrin, rare point mutations in NLGN2 (Sun et

al., 2011) and NLGN4 (Jamain et al., 2003) have been identified in patients with schizophrenia

and ASD, respectively, while mutations in the collybistin gene (ARHGEF9) have been reported

in cases with intellectual disability (ID) and epilepsy (Kalscheuer et al., 2009; Lesca et al., 2011;

Shimojima et al., 2011). Frameshift mutations truncating NRXN1 and NRXN2 in patients with

schizophrenia and ASD, respectively, were also observed to abolish the ability of these proteins

to induce gephyrin clustering at dendrite contact sites when synaptogenic activity was tested in

neuron co-culture assays (Gauthier et al., 2011). The SHANK proteins, which act as postsynaptic

neuronal scaffolding proteins in excitatory synapses (a role analogous to that performed by

gephyrin in inhibitory synapses) have been implicated by rare mutations in genetic risk for ASD

(Berkel et al., 2010; Durand et al., 2007; Sato et al., 2012), schizophrenia (Gauthier et al., 2010)

and epilepsy (Lesca et al., 2012). Given the strong evidence supporting GPHN as a candidate

Page 125: Copy Number Variation in Neurodevelopmental Disorders

112

molecule for involvement in NDDs, the gene was screened for copy number variations (CNVs)

and point mutations in patients from ASD, schizophrenia and seizure disorder case cohorts.

6.3 Results

6.3.1 Detection and inheritance testing of rare microdeletions at the GPHN locus

The GPHN locus was screened for CNVs using high resolution microarray data from 5,384

individuals from ASD, schizophrenia and seizure disorder patient cohorts. In five of these

patients, hemizygous deletions were identified at 14q23.3 interrupting multiple exons of GPHN

(Figure 6.1 & Table 6.1). There was also an earlier report of a 183 kb deletion affecting six

exons of GPHN in a single individual from a cohort of 3,391 schizophrenia patients studied by

the International Schizophrenia Consortium (International Schizophrenia Consortium, 2008).

Additional clinical details were obtained for this patient and his family (Family 6 in Figures 6.1

and 6.2). Independent experimental validation of the array calls was performed using quantitative

PCR (qPCR) or fluorescence in situ hybridization (FISH). None of the CNVs detected at other

loci in probands from the six families were predicted to be of clinical significance when

classified based on American College of Medical Genetics guidelines for interpretation of CNVs

(Kearney et al., 2011). Parental testing for CNV status at the GPHN locus revealed that the

deletions arose de novo in three of the probands (Families 1, 3 and 5 in Figure 6.2). The deletion

in the proband from Family 2 with ASD was inherited from his father, who was reported to have

sub-clinical socialization deficits, but no formal diagnosis of ASD. The deletion was not present

in the three sisters of the proband. DNA samples were not available from the parents in Family 4

for inheritance testing. The deletion in the proband from Family 6 was inherited from his mother,

who was apparently unaffected. DNA samples were not available from the maternal grandmother

of the proband, who was reported to have schizophrenia.

Page 126: Copy Number Variation in Neurodevelopmental Disorders

113

Figure 6.1. Genomic locations of exonic GPHN deletions in the NDD probands

Red bars represent the hemizygous exonic GPHN deletions at 14q23.3 detected in the index probands of the six families presented in this chapter. Information about transcript isoforms and genomic co-ordinates corresponds to human genome Build 36 (hg18). The common region of overlap of the six deletions encompasses exons 3–5, which correspond to the G-domain of the gephyrin protein. Coordinates (hg18) of the deletions observed in control individuals are chr14:66,055,080–66,882,227, chr14:66,153,552–66,267,833 and chr14:66,287,336–66,355,274.

Page 127: Copy Number Variation in Neurodevelopmental Disorders

114

Figure 6.2. Pedigrees of individuals with GPHN deletions

Squares and circles denote males and females, respectively. Arrow indicates the index proband in each family. ‘N/A’ represents individuals for whom DNA was not available for testing. All other individuals were assayed for deletions at the GPHN locus using microarray, FISH or qPCR.

Page 128: Copy Number Variation in Neurodevelopmental Disorders

115

Table 6.1. Genetic and phenotypic details of probands with GPHN deletions Family

(proband sex) Genotype Phenotype

GPHN deletion size Exons

affected Inheritance Diagnosis

Dysmorphic features

Speech Delay

Motor Delay

Other

Family 1 (female)

357 kb 3-12 De novo ASD N Y Y -

Family 2 (male)

319 kb 2-5 Paternal ASD N N N Coarctation of aorta

Family 3 (male)

273 kb 2-5 De novo ASD & seizures

N/A Y N Intellectual disability

Family 4 (male)

134 kb 3-5 Unknown Seizures N/A N/A N/A -

Family 5 (male)

338 kb 3-11 De novo Schizophrenia N N N -

Family 6 (male)

183 kb 3-8 Maternal Schizophrenia N/A N N Learning difficulties

Abbreviations: ASD, Autism Spectrum Disorder; N/A, Not assessed

Page 129: Copy Number Variation in Neurodevelopmental Disorders

116

6.3.2 Frequency comparison of GPHN deletions in cases and controls

Deletions affecting exons of GPHN are extremely rare in the general population. Of the 27,019

control individuals examined for CNVs at this locus (Lionel et al., 2013), only three possessed

exonic GPHN deletions. CNVs present in control individuals might be indicative of variable

expression of GPHN deletions, lack of rigorous assessment of neuropsychiatric phenotype in the

controls, or false-positive calls since DNA was not available from the two SAGE control

individuals for experimental validation. Taken together, the frequency of experimentally

validated exonic deletions at the GPHN locus in cases is significantly greater than that in

controls (6 / 8,775 cases vs. 3 / 27,019 controls, respectively; Fisher’s exact test 2 sided p =

0.009). There were no exonic deletions reported at the GPHN locus in the Database of Genomic

Variants (DGV).

6.3.3 Clinical features of individuals with exonic GPHN deletions

Family 1 (Proband with ASD diagnosis; 357 kb de novo deletion):

The female proband of European ancestry was conceived naturally to a non-consanguineous 35

year old (yro) mother and 33 yro father. The pregnancy was uncomplicated and the child was

delivered at 38 weeks of gestation with a birth weight of 2,810 grams and length of 47 cm. There

are no reported medical or neuropsychiatric conditions in her parents and two older siblings.

During early development she was noted to be a “calm” child, with limited movement when laid

down. She began walking at 12 months but otherwise had slow motor development and gait

issues, which prompted ergotherapy. Upon examination by a pediatrician at two yro and again at

four yro, no dysmorphology was noted. She had language delay with no speech at 2 yro,

followed by gradual speech development by age four. She had episodes of echolalia. She

attended kindergarten for two years with an early intervention program teacher, followed by

regular elementary and secondary school with the aid of an education assistance worker. Up to

six years of age, she continued to desire swaddling and had hypersensitivity to light and sound

associated with self-injury (head banging, tearing hair). Assessment by a psychologist at 13 yro,

which included the Autism Diagnostic Observation Schedule (ADOS) and Autism Diagnostic

Interview – Revised (ADI-R), resulted in a diagnosis of high-functioning ASD. She completed a

secondary school degree with the exception of mathematics. Chromosomal microarray

Page 130: Copy Number Variation in Neurodevelopmental Disorders

117

(Affymetrix 6.0) indicated that she had a deletion at cytoband 14q23.3 (chr14:66,274,499-

66,631,750 [hg18]), which was not present in the parents (Figures 6.1 and 6.2, Family 1).

Family 2 (Proband with ASD diagnosis; 319 kb paternally inherited deletion):

The male proband was conceived naturally to a 28 yro mother and 30 yro father. The proband

was born vaginally at 35 weeks gestation with a birth weight of 3,133 grams and length of 50.8

cm. He was treated for respiratory distress in hospital for nine days following delivery.

Coarctation of the aorta and bicuspid aortic valve were noted on echocardiography performed

upon re-admittance to the hospital at 18 days of life, secondary to continued respiratory distress.

He had an end-to-end anastomosis procedure at one month followed by a coarctation repair with

patch annuloplasty at four months. Post-treatment echocardiograms have been normal and he

remains asymptomatic. He exhibited mild global developmental delay early in life, with walking

at 16 months and speech emergence at 15-16 months. By the age of three years he could speak in

two word sentences. At age six he was diagnosed with ASD by a child psychologist. In

particular, he had difficulties with socialization and repetitive behaviors (specifically throat

clearing and tics). Upon examination by a clinical geneticist at 6 years 8 months, no significant

dysmorphology was noted. The proband experienced academic difficulties and required a

modified program in preschool and kindergarten. By 9 years of age, he had graduated from this

program and received no additional special education. Chromosomal microarray (Agilent 44K)

indicated he had a deletion at cytoband 14q23.3 (chr14:66,102,556-66,421,440 [hg18]), which

was paternally inherited, but not present in his siblings (Figures 6.1 and 6.2, Family 2). The

proband is the third of four children of non-consanguineous parents (Figure 6.2, Family 2). He

has one sister with anxiety, another sister with speech delay and hyperactivity, a third sister with

head-banging behaviors, and a maternal half-brother with bipolar disorder. His mother has been

diagnosed with depression and anxiety and was noted to have speech difficulties in school. In the

maternal extended family there are individuals with additional psychological concerns. The

father is apparently healthy with normal intelligence, but has some challenges with socialization.

There is significant psychiatric illness in the paternal extended family.

Family 3 (Proband with diagnosis of ASD and seizures; 273 kb de novo deletion):

The male proband was conceived naturally to a 36 yro mother and 40 yro father. The pregnancy

was uncomplicated and the proband was delivered naturally with a birth weight of 3,714 grams

Page 131: Copy Number Variation in Neurodevelopmental Disorders

118

and length of 48 cm. He crawled at 7 months and walked at 15 months. Developmental delay

was first noticed by parents at age two, two months after a serious encephalitic illness that

required hospitalization for three days. Cyclical seizures began at the time of this illness and

recurred until age six and subsequently he remained seizure free without anti-convulsant

treatment. He exhibited speech and language delay and did not talk until approximately age four.

He was diagnosed with ASD (Pervasive Developmental Disorder (PDD)) at age four by a child

psychologist on the basis of behavioral assessments provided by parents, teachers and therapists.

He was diagnosed with intellectual disability a few years later by a school psychologist. He

continues to have behavioral issues including anxiety, OCD, tics, and impulsive, sometimes

aggressive behaviors. He talks to himself frequently, and has trouble concentrating and sitting

still. The proband attended early intervention therapy from age three to six. He has received

special education and/or tutoring throughout school as well as speech therapy and occupational

therapy. He is currently in tenth grade. He has had physical therapy in the past for toe walking.

The proband tested negative for fragile X and had normal results from magnetic resonance

imaging (MRI) and computed tomography (CT) brain scans and electroencephalography (EEG).

At 15 years of age, he was referred for further genetic testing. Chromosomal microarray (Agilent

180K) indicated he had a deletion at cytoband 14q23.3 (chr14:66,148,602-66,421,440 [hg18])

(Figure 6.1, Family 3). Neither parent had the deletion when tested by FISH (Figure 6.2, Family

3). After detection of the de novo hemizygous exonic deletion in the proband, clinical metabolic

testing for Molybdenum Cofactor (MoCo) deficiency was performed and the results were

negative. The testing measured urine levels of uric acid, xanthine, hypoxanthine and uracil and

serum levels of molybdenum and uric acid.

Family 4 (Proband with diagnosis of seizures; 134 kb deletion):

The male proband was examined at five years of age for seizures of unknown type or duration.

Clinical chromosomal microarray (Agilent 44K) indicated he had a deletion at cytoband 14q23.3

(chr14:66,287,215-66,421,440 [hg18]) (Figure 6.1, Family 4). This individual was lost to follow-

up and detailed clinical information was not obtained. Samples from the parents were not

available for testing (Figure 6.2, Family 4).

Family 5 (Proband with schizophrenia diagnosis; 338 kb de novo deletion):

Page 132: Copy Number Variation in Neurodevelopmental Disorders

119

The male proband was conceived naturally to non-consanguineous parents, a 29 yro mother and

36 yro father of European ancestry. There was no evidence of developmental delay or autistic

features. The proband completed tenth grade with no reported difficulties, leaving school at age

17 years to work. At 23 years, he began to have increasing anxiety, preoccupation, and paranoia

that responded well to a few months of treatment with chlorpromazine. At age 24 years, he was

admitted to a hospital for schizophrenia. Neuropsychological testing using the Wechsler Adult

Intelligence Scale (WAIS) revealed a Full Scale IQ of 92. He improved with electroconvulsive

therapy, and was discharged after two months. Following two additional hospitalizations, he has

been relatively stable on a standard antipsychotic medication regimen with an adjuvant

antidepressant for many years. A detailed psychiatric assessment, including use of a modified

version of Structured Clinical Interviews for DSM-III-R for Axis I disorders, confirmed a

research diagnosis of chronic schizophrenia. He has no history of seizures and there is no known

history of neuropsychiatric illness or seizures in his parents or siblings. There is a significant

history of schizophrenia in the paternal extended family. There was no significant

dysmorphology noted on examination as an adult by a psychiatric geneticist. Research

chromosomal microarray (Affymetrix 6.0) indicated he had a deletion at cytoband 14q23.3

(chr14:66,267,488-66,605,185 [hg18]) (Figure 6.1, Family 5). Neither parent had the deletion

when tested by Affymetrix 6.0 microarray and qPCR (Figure 6.2, Family 5).

Family 6 (Proband with schizophrenia diagnosis; 183 kb maternally inherited deletion):

The male proband of European ancestry was recruited into the International Schizophrenia

Consortium (ISC) study (30) at age 26 yro. He had normal early milestones. During his

schooling, he had poor marks and left early after 8 years of education. The first onset of

psychiatric symptoms started at 21 years with changes in his behavior, three months before he

attended the National Service. After finishing the service he appeared withdrawn, talked little,

and claimed that “things were different” and that his genitals were changing. At age 22 he was

seen by a psychiatrist after reporting he heard voices laughing at him and people were looking at

him. He believed that he was being poisoned. He was subsequently treated with long-acting

injectable antipsychotic medication for the three years prior to recruitment into the ISC study. He

remained in a chronic condition, with similar delusions, anhedonia (loss of interest in pleasurable

activities) and severe lethargy. Chromosomal microarray testing (Affymetrix 6.0) as part of the

ISC study indicated that he had a deletion at cytoband 14q23.3 (chr14:66,287,324-66,470,381

Page 133: Copy Number Variation in Neurodevelopmental Disorders

120

[hg18]) (Figure 6.1, Family 6). Microarray testing showed the deletion to be present in the

mother, who was reported to be unaffected (Figure 6.2, Family 6). DNA was not available for

testing from the proband’s maternal grandmother, who reportedly also had schizophrenia.

6.4 Discussion

Here, we present clinical and molecular characterization of six unrelated index cases possessing

rare de novo or inherited microdeletions overlapping the GPHN gene at 14q23.3. Perhaps the

most compelling evidence for the pathogenicity of the deletions is the fact that three of the five

in families where DNA from both parents was available for inheritance testing arose as de novo

events in the probands. The deletions in the probands in Families 2 and 6 were inherited from a

parent with sub-clinical socialization deficits and an unaffected parent respectively, suggesting

variable expression of structural mutations at this locus.

To our knowledge, this is the first report of hemizygous deletions at the GPHN locus in

connection with ASD or seizure phenotypes. The deletions in probands from Families 5 and 6

add to previous evidence supporting the involvement of GPHN in genetic risk for schizophrenia.

The GPHN locus was one of the top ranking genetic loci at which runs of homozygosity were

significantly over-represented in cases versus controls in a study examining highly penetrant risk

loci in schizophrenia (Lencz et al., 2007). In addition, the statistical method Gene Relationships

Among Implicated Loci (GRAIL) employed by Raychaudhri et al. pinpointed GPHN as a key

gene highly likely to play an etiological role in schizophrenia based on its functional relatedness

to other genes implicated in risk for the disorder (Raychaudhuri et al., 2009).

The common region of overlap across the six deletions encompasses exons 3 to 5 (Figure 6.1),

corresponding to the coding segment of the G domain of the gephyrin protein. G-domain

trimerization is vital for the formation of the hexagonal gephyrin oligomer scaffolds required for

stable GABA receptor clustering in postsynaptic inhibitory neurons (Fritschy et al., 2008).

Interestingly, abnormally spliced gephyrin mRNA (with no corresponding genomic mutation),

lacking several exons corresponding to the G domain, has been isolated in the hippocampus of

patients with temporal lobe epilepsy (Forstera et al., 2010). Due to the missing G-domain exons,

these aberrant protein variants were no longer able to form trimers and were observed to act in a

Page 134: Copy Number Variation in Neurodevelopmental Disorders

121

dominant-negative manner by interacting with normally spliced gephyrin and impairing its

function, as observed by depleted GABA receptor cluster density and reduced GABAergic

postsynaptic current amplitudes (Forstera et al., 2010). The authors of this study speculated that

cellular stressors such as elevated temperature or alkalosis, which can cause or arise from seizure

activity, respectively, might induce exon-skipping in GPHN mRNA leading to a dominant-

negative effect (Forstera et al., 2010). This type of ‘environmentally’ induced mRNA mutation

might have an equivalent functional impact to dosage effects of a hemizygous genomic deletion

of the underlying GPHN locus (as is seen in our six families). However the situation in vivo may

be more complex than that observed in vitro due to the multiple isoforms and splice variants of

the gene, differences in gephyrin expression across various synapse types and factors affecting

the expression levels of the non-deleted allele. Given the high prevalence of epilepsy in ASD

(>21% in ASD individuals with an intellectual disability (ID) and 8% in those without ID)

(Amiet et al., 2008), our findings call attention to considering models (Rubenstein and

Merzenich, 2003) whereby the effects of seizures may contribute to expression of ASD, instead

of being interpreted as associated medical comorbidities.

In addition to its synaptic function in the central nervous system, gephyrin has a more wide-

ranging role in peripheral tissues in the synthesis of molybdenum cofactor (MoCo) (Feng et al.,

1998; Stallmeyer et al., 1999), the catalytically active center of molybdoenzymes (class of

molybdenum containing enzymes), whose function is essential for the survival of almost all life-

forms. Gephyrin exhibits extensive sequence conservation across different species ranging from

bacteria to humans, attesting to the strong evolutionary pressure preserving MoCo synthesis

(Stallmeyer et al., 1999). To date, there have been two reports of homozygous GPHN mutations

in humans, both in connection with MoCo deficiency, an extremely rare and sometimes lethal

autosomal recessive metabolic disease characterized by untreatable neonatal seizures with

opisthotonos, hypotonia, feeding difficulties, facial dysmorphism and intellectual disability. One

report (Reiss et al., 2001) identified homozygous deletions, which eliminated exons 2 and 3

resulting in a frameshift after only 21 codons of normal coding sequence and complete loss of

protein function, in the last of three affected infants born to Danish first-cousin parents. All three

infants exhibited symptoms of severe MoCo deficiency and died within a month after birth.

While the vast majority of cases of MoCo deficiency are caused by point mutations in one of two

other related genes involved in MoCo biosynthesis, MOCS1 and MOCS2 (Reiss and Hahnewald,

Page 135: Copy Number Variation in Neurodevelopmental Disorders

122

2011), there has been one other case linked to a homozygous point mutation in GPHN (Reiss et

al., 2011). This D580A missense change was recently reported in a female child (of first cousin

parents) of Algerian origin who exhibited a milder MoCo deficiency phenotype when compared

with the patients carrying the GPHN homozygous deletion. The parents, who were carriers of

hemizygous GPHN mutations, did not exhibit any MoCo deficiency symptoms in either of the

two families. We have attempted to see if these parents might have a neuropsychiatric

phenotype, but were unable to obtain information.

The severe symptoms of GPHN null mutations in humans are consistent with phenotypes

observed in geph–/– knockout mice embryos that were born in expected numbers without

apparent physical abnormalities but died within one day of birth (Feng et al., 1998). Knockout

neonates had breathing difficulties and a heightened startle response, becoming rigid and

exhibiting a hyperextended posture when lightly touched. They also had other phenotypic traits

similar to human MoCo deficiency patients including myoclonus and feeding difficulties. The

authors demonstrated that gephyrin was essential for both glycine receptor clustering and MoCo

synthesis. While heterozygous mutant (geph+/–) mice were apparently phenotypically normal

(Feng et al., 1998), there is some evidence from other studies for subtle gene dosage effects. The

concentration of gephyrin protein was reduced by 50% in geph+/– mice (Kneussel et al., 1999),

and reductions in GABAA receptor clustering have been observed in vitro (but not in vivo) in

neuronal cultures derived from geph+/– animals (Fischer et al., 2000). RNAi experiments that

reduced gephyrin expression by ~50% in mouse hippocampal neurons observed significant

reduction of GABA receptor cluster density (Jacob et al., 2005). These findings hint at potential

disturbances in synaptic homeostasis in hemizygous GPHN mutation carriers. Thus while

homozygous GPHN mutations abolishing both the synaptic and molybdenum cofactor

biosynthetic activity of the protein result in the severe metabolic phenotype of MoCo deficiency,

hemizygous deletions reported in this study presumably affect only the synaptic functions of the

protein, as can be inferred from the absence of MoCo deficiency symptoms in our patients and in

previously reported individuals with such mutations (Reiss et al., 2001).

Our data add to the accumulating evidence highlighting the role of dysfunctional inhibitory

signaling with ASD etiology (Coghlan et al., 2012). GPHN joins the growing list of genetic risk

loci evidently shared between ASD and schizophrenia and other NDDs, examples of which are

listed in Chapter 1. Recent research has also highlighted the involvement of CNVs in epilepsy at

Page 136: Copy Number Variation in Neurodevelopmental Disorders

123

several of the loci first implicated in ASD and schizophrenia, including 15q13.3, 16p13.11 and

15q11.2 as well as at individual genes such as CNTNAP2, NRXN1 and SHANK3 (Duong et al.,

2012; Lesca et al., 2012; Mulley and Mefford, 2011; Poot et al., 2011). There is also emerging

evidence of comorbidity and functional links between epilepsy and ASD (Amiet et al., 2008;

Brooks-Kayal, 2010), and between epilepsy and schizophrenia (Chang et al., 2011).

Our finding of different disease states arising from deletions of the same gene is in line with

previous reports of several other genetic loci being involved in genetic risk for a range of NDDs

including autism, schizophrenia and epilepsy. For example, these observations of pleiotropic

effects of mutations at the same gene are also seen in connection with mutations in pre-synaptic

neurexin and in the post-synaptic SHANK scaffolding proteins. Duong et al. recently reported

variable penetrance and pleiotropy of NRXN1 mutations within a single family – individuals with

hemizygous deletions and point mutations had differing neurodevelopmental phenotypes ranging

from ASD and epilepsy to schizophrenia and psychotic disorder (Duong et al., 2012). As with

these and other CNVs, the variable expression and penetrance of the NDD phenotypes associated

with structural mutations at the GPHN locus could be modulated by other risk alleles, protective

genetic modifiers and/or environmental factors. Differences among the translational products,

expression levels and stabilities of the pathogenic gephyrin forms, together with the multiple

types of inhibitory receptors potentially affected downstream, could also contribute to the range

of observed disorder phenotypes.

Our results enhance understanding of the GPHN mutation spectrum and will aid in the

interpretation of clinical genetic testing of NDDs. The molecular findings and clinical case

descriptions provided here will serve as a foundation for future studies elucidating a more

comprehensive correlation study of DNA, mRNA (and protein) variants at this locus with the

resulting neurodevelopmental clinical outcomes. Building on previous work (Forstera et al.,

2010), our findings potentially link genomic CNV dosage effects and epilepsy-induced mRNA

dysregulation to GPHN, possibly providing a sought after model to explore gene-environment

interactions in neuropsychiatric disease.

Page 137: Copy Number Variation in Neurodevelopmental Disorders

124

6.5 Materials and Methods

6.5.1 Study subjects

The patient cohort inspected for GPHN deletions in this study was composed of 1,158 Canadian

patients with ASD, 72 Austrian patients with ASD, 450 Canadian patients with schizophrenia

and a clinical dataset of 3,704 individuals with primary diagnoses of ASD and/or seizure

disorders that were referred for clinical microarray testing at the Mayo Clinic cytogenetics

laboratory. The Canadian ASD and schizophrenia cohorts were recruited as previously described

in Chapters 3 and 5 respectively. All samples were approved for inclusion in the study by the

appropriate local research ethics boards. The GPHN region was inspected for CNVs in

microarray data analyzed by our group (Lionel et al., 2013) from 21,345 population-based adult

controls genotyped with high-resolution arrays and published data from 2,493 controls

genotyped at the University of Washington (Itsara et al., 2009) and from 3,181 controls analyzed

by the International Schizophrenia Consortium (International Schizophrenia Consortium, 2008).

6.5.2 CNV validation and inheritance testing

Validation of deletions in probands from Families 2, 3 and 4 and in their family members from

whom DNA was available, was conducted by FISH. CNV validation in probands from Families

1, 5 and 6, and in their family members from whom DNA was available, was performed using

SYBR-Green-based real-time quantitative PCR (qPCR), for which two independent primer pairs

were designed at the GPHN locus and at the FOXP2 locus on chromosome 7 as a diploid control.

The parentage of the proband in Family 5 with the de novo GPHN deletion, was confirmed by

calculating the Mendelian error rate of Affymetrix SNP 6.0 array genotypes using the PLINK

tool set (Purcell et al., 2007). Parentage of the proband with a de novo deletion in Family 1 was

confirmed using the AmpFlSTR Identifiler PCR Amplification Kit (Life Technologies) using the

manufacturer’s instructions. Briefly, the kit employs a short tandem repeat (STR) multiplex

assay that amplifies 15 tetranucleotide repeat loci and the Amelogenin gender determining

marker in a single PCR reaction. Samples were run on an AB 3100 Genetic Analyzer using

POP4, dye set G5 and analyzed in GeneMapper v3.7.

Page 138: Copy Number Variation in Neurodevelopmental Disorders

125

Chapter 7

Summary and Future Research Directions

The following study provides an excellent illustration of several key themes of my thesis

research and I use this to illustrate possible extensions of my work. Parts of this chapter are

adapted, with permission from Oxford Press, from the following published journal article:

Lionel AC*, Tammimies K*, Vaags AK*, Rosenfeld JA, Ahn JW, Merico D, Noor A, Runke

CK, Pillalamarri VK, Carter MT, Gazzellone MJ, Thiruvahindrapuram B, Fagerberg C, Laulund

LW, Pellecchia G, Lamoureux S, Deshpande C, Clayton-Smith J, White AC, Leather S, Trounce

J, Melanie Bedford H, Hatchwell E, Eis PS, Yuen RK, Walker S, Uddin M, Geraghty MT,

Nikkel SM, Tomiak EM, Fernandez BA, Soreni N, Crosbie J, Arnold PD, Schachar RJ, Roberts

W, Paterson AD, So J, Szatmari P, Chrysler C, Woodbury-Smith M, Brian Lowry R,

Zwaigenbaum L, Mandyam D, Wei J, Macdonald JR, Howe JL, Nalpathamkalam T, Wang Z,

Tolson D, Cobb DS, Wilks TM, Sorensen MJ, Bader PI, An Y, Wu BL, Musumeci SA, Romano

C, Postorivo D, Nardone AM, Monica MD, Scarano G, Zoccante L, Novara F, Zuffardi O,

Ciccone R, Antona V, Carella M, Zelante L, Cavalli P, Poggiani C, Cavallari U, Argiropoulos B,

Chernos J, Brasch-Andersen C, Speevak M, Fichera M, Ogilvie CM, Shen Y, Hodge JC,

Talkowski ME, Stavropoulos DJ, Marshall CR, Scherer SW. Disruption of the ASTN2/TRIM32

locus at 9q33.1 is a risk factor in males for autism spectrum disorders, ADHD and other

neurodevelopmental phenotypes. Human Molecular Genetics; 2014; doi:10.1093/hmg/ddt669.

*Joint first authors

I coordinated collection of genetic data from the different molecular diagnostic sites involved in

this study, performed CNV analysis of controls and the TCAG cases, conducted the statistical

analysis of the CNV results presented in Section 7.2.1 and drafted and revised the manuscript.

Dr. K. Tammimies performed the functional characterization experiments, prepared the figures

presented in Section 7.2.2 and drafted and revised the manuscript. Dr A.K. Vaags coordinated

collection of clinical phenotype information from individuals with ASTN2/TRIM32 deletions.

Page 139: Copy Number Variation in Neurodevelopmental Disorders

126

7.1 Summary and Implications

My thesis research had two broad aims. First, in order to discover novel genes involved in NDD

risk, I performed the genome-wide CNV analysis of three newly characterized Canadian patient

cohorts of individuals with one of 3 different disorders: ASD, ADHD and schizophrenia.

Following high-confidence CNV calling using multiple computational algorithms, rare CNVs

were identified by filtering against CNV results generated from controls from the general

population using identical methodology as for the cases. Within each NDD dataset, the strategy

of rare CNV prioritization focused on de novo variants (in those ASD and ADHD probands for

whom parental DNA was available), loci affected by rare CNVs across unrelated probands and

loci previously reported in literature. In addition to replicating pathogenic CNVs previously

reported in literature, these analyses uncovered several novel risk genes for each disorder. For

ASD, discovery of de novo and rare inherited exonic deletions affecting the NRXN3 gene in

multiple unrelated probands provided the final piece of evidence for the involvement of all three

members of the synaptic neurexin protein family in risk for the disorder. Follow-up of these

findings in additional clinical datasets and phenotype characterization of individuals with such

variants revealed clinical heterogeneity and variance in phenotype severity potentially correlated

with the specific gene isoforms affected by the deletions. The rate of de novo CNVs in ADHD

was assessed for the first time and was found to be relatively lower than the established rate of

such mutations in ASD and schizophrenia. Several rare inherited variants were identified in the

ADHD dataset in unrelated patients that affected loci previously implicated in NDD risk

including ASTN2/TRIM32 and DDX53/PTCHD1.The clinical diagnostic yield of microarray

testing of schizophrenia individuals was investigated and was found to be in line with findings

from ASD, providing the first evidence for the benefits of such testing for schizophrenia.

Another novel finding in the schizophrenia dataset involved rare CNVs at the 2q13 region, thus

expanding the phenotype spectrum of this genomic locus from its previous known role in ASD

and ID. The second aim of my thesis research was to probe genetic overlap between the three

NDDs by a systematic comparison of the genome-wide rare CNV results from each disorder

cohort. In addition to confirming the phenotypic heterogeneity of established genomic

syndromes such as the 16p11.2 microduplication and 15q11.2 microduplication regions, this

investigation revealed several novel candidate genes involved in cross-disorder risk including

ASTN2/TRIM32 in ADHD and ASD and GPHN for ASD and schizophrenia. In addition to the

Page 140: Copy Number Variation in Neurodevelopmental Disorders

127

results presented here, the data from my thesis project has also facilitated publications with

collaborators describing other novel NDD risk loci such as 16q24.2 deletions (Handrigan et al.,

2013), 1q21.1 duplications (Dolcetti et al., 2013), 15q13.3 duplications (Williams et al., 2012),

2q23.1 duplications (Chung et al., 2012), 3q29 duplications (Goobie et al., 2008), AUTS2

(Beunders et al., 2013), SHANK1 (Sato et al., 2012), PTCHD1 (Noor et al., 2010), TMLHE

(Celestino-Soper et al., 2011) and DPYD (Carter et al., 2011). The findings of my thesis research

have important implications for different facets of molecular medicine including diagnostic

testing, genetic counseling, clinical screening and intervention, as well as for the development of

novel therapeutic agents.

7.1.1 Implications for clinical genetic testing and molecular diagnostics

The rare CNV findings of my thesis research are of immediate relevance to guide clinical genetic

testing. In Canada and many other countries, genome-wide microarray testing in diagnostic

laboratories is now a routine part of the clinical work-up of individuals presenting with NDDs

such as ASD, ID or developmental delay or with congenital anomalies (Anagnostou et al., 2014;

Miller et al., 2010a). Before the results from such testing can be reported to the referring

clinicians, clinical cytogeneticists interpret the CNVs detected and classify them into different

categories ranging from known pathogenic mutations to known benign variants with

intermediate categories of unknown clinical significance (Kearney et al., 2011). A key step in

this categorization process involves literature searches for publications describing the clinical

relevance of specific CNVs and the biological functions of the genes encompassed by them

(Kearney et al., 2011). Thus, research articles such as those arising from my thesis work are

instrumental in the interpretation of test results. Indeed, we have often received requests for

additional information about loci such as GPHN, NRXN3 and ASTN2/TRIM32 from

cytogeneticists who have encountered novel cases with mutations at these genes.

Based on the results of my thesis research, I envision an expansion of the scope of clinical

genetic testing. In particular, the findings from the CNV analysis of the schizophrenia cohort

presented in Chapter 5 make a case for routine clinical microarray testing in schizophrenia,

similar to what is currently standard practice for patients with ASD, ID and developmental delay.

As described in Chapter 5, the percentage of schizophrenia patients with a clinically significant

CNV (as defined by clinical cytogeneticists using consensus parameters for CNV categorization)

Page 141: Copy Number Variation in Neurodevelopmental Disorders

128

is in line with yields observed for ASD genetic testing (Miller et al., 2010a; Shen et al., 2010).

Routine employment of clinical microarray testing for schizophrenia would greatly facilitate the

discovery of additional risk genes for the disorder. Such an expansion of clinical genetic testing

into what is largely an adult onset disorder would also represent a paradigm shift in medical

genetics and will greatly broaden its current predominant focus on pediatric conditions.

7.1.2 Implications for genetic counseling of patients and their families

CNV findings highlighting specific genes in NDD risk are especially informative for genetic

counseling (De Wolf et al., 2013). Return of these genetic results, concurrent with explanations

of their implications by genetic counselors and clinical geneticists, can be beneficial to patients

and their families. For instance, the availability of these rare CNV results aids in the calculation

of recurrence risk for NDD in the family and is informative for family planning (Costain and

Bassett, 2014). If the pathogenic rare CNV is found to be of de novo origin in the proband, this

implies that the recurrence risk for siblings is equal to the population prevalence of the disorder

(~1% for schizophrenia and ASD and ~5% for ADHD). On the other hand if the mutation is

found to be inherited from a parent, the siblings will be at a much higher risk for recurrence than

the general population risk. For example, presence of 22q11.23 deletion in a parent would imply

a risk for schizophrenia of approximately 20%–25% in each child (Costain and Bassett, 2014).

The X-linked recessive inheritance of risk CNVs such as PTCHD1 deletions has specific

implications for the counseling of female carriers.

In addition to better informed reproductive decision making, genetic counseling of patients and

their families also has other important therapeutic benefits. Of particular relevance to this thesis

are the comprehensive investigations of the efficacy of genetic counseling of individuals with

schizophrenia (Costain et al., 2014b) and their family members (Costain et al., 2014a) (by my

clinical collaborators, Dr. Anne Bassett and Dr. Gregory Costain). These longitudinal studies

followed up on the rare CNV findings described in the schizophrenia cohort in Chapter 5 of this

thesis and assessed individuals for certain knowledge-based and psychological factors before and

after the receipt of their CNV results via genetic counseling. Individuals with schizophrenia

reported significant improvements after counseling in understanding of the disorder and its

recurrence risk as well as significant reductions in internalized stigma and self-blame (Costain et

al., 2014b). Similar post-counseling increases in knowledge about the condition and decreases

Page 142: Copy Number Variation in Neurodevelopmental Disorders

129

sense of stigma and fear of recurrence were also observed in family members of individuals with

the disorder (Costain et al., 2014a). Participants in both studies reported high satisfaction with

genetic counseling and strongly endorsed its necessity.

7.1.3 Implications for screening, diagnosis and treatment of NDDs

The findings from my research will enhance the detection yield in clinical diagnostic laboratories

of rare CNVs implicating specific NDD risk genes in patients. The rapid reporting of these

results to the referring clinicians will enable much faster diagnosis of NDD conditions than was

possible earlier in the absence of genetic testing. This is turn will allow for earlier and more

effective clinical intervention to benefit the patients. For example, in the case of ASD, there is

clear evidence that children diagnosed at earlier ages benefit from timely introduction to

behavioral therapy regimens (Carbone, 2013; Myers and Johnson, 2007). One particularly

effective and widely prescribed regimen is Applied Behavioral Analysis (ABA), which increases

the functional abilities of children with ASD by teaching them behavior skills using repeated

trials (Vismara and Rogers, 2010). Early administration of intensive ABA or other behavioral

interventions such as the Early Start Denver Model have been shown by several studies to

comprehensively alter the developmental trajectory of ASD and yield significant improvements

in long-term outcomes (Dawson et al., 2012; Dawson et al., 2010; Eldevik et al., 2009; Vismara

and Rogers, 2010). Although maximum benefits are obtained when these therapeutic approaches

are started at age 3 or younger, the current average age of diagnosis is between 48 months

(classic autism) and 75 months (Asperger syndrome), and ASD diagnosis is not made for half of

school-aged children before the age of 5 (Autism and Developmental Disabilities Monitoring

Network, 2012; Pringle et al., 2012). The results from genetic screening can bridge this gap

between optimal age for initiating clinical interventions and current age of diagnosis. Earlier

diagnosis and treatment are also of substantial benefit for long-term outcomes in adult-onset

NDDs such as schizophrenia (Emsley et al., 2008).

In addition to benefiting NDD screening, the rare CNV findings have important implications for

the clinical management of patients with such genetic changes. For some of the well-established

microdeletion syndromes featuring recurrent changes of relatively high prevalence such as

22q11.2 DS, consensus clinical practice guidelines have been formulated (Bassett et al., 2011).

These feature anticipatory care over the lifetime of the patient focused on the optimal medical

Page 143: Copy Number Variation in Neurodevelopmental Disorders

130

management of the clinical features that are typically known to arise in connection with the

deletion at different time points (Costain and Bassett, 2012). For other rarer CNVs that were

discovered more recently, longitudinal genotype-phenotype studies are essential to define the

phenotypes resulting from the clinical manifestation of the CNV and establish appropriate

guidelines for their clinical management. Knowledge of the biological function of key genes

affected by the CNV can inform treatment strategies. One recent study reported

pharmacogenetically guided treatment of a patient with a large 15q13.3 deletion who presented

with epilepsy, ID, schizophrenia and episodes of aggressive rage (Cubells et al., 2011). This

recurrent microdeletion has been reported in several other individuals with epilepsy, ID,

aggressive behavior, schizophrenia, and autism (Ben-Shachar et al., 2009; Miller et al., 2009).

Since one of the genes affected by the deletion, CHRNA7, encodes a7 nicotinic cholinergic

receptor (NCHR) protein, Cubells et al. hypothesized that deficits in NCHR related

neurotransmission was involved in their patient’s phenotype. Indeed, substantial improvement

was observed upon treatment with galantamine, a U.S. Food and Drug Administration (FDA)

approved modulator of NCHR neurotransmission. The presence of comorbid symptoms of

different NDDs in individuals with CNVs involved in cross-disorder risk also has relevance for

the selection of optimal pharmacological treatment. For instance, individuals with

ASTN2/TRIM32 deletions often present with both ASD and ADHD traits and specific drugs such

as methylphenidate and atomoxetine, together with behavioral therapy, have been recommended

for individuals with these comorbid symptoms (Mahajan et al., 2012; Murray, 2010; Taurines et

al., 2012). Other examples of pharmaceuticals with potential therapeutic benefit across different

NDDs include antipsychotic drugs usually prescribed for schizophrenia showing efficacy when

used for different conditions such as ASD (Zuddas et al., 2011). The use of oxytocin has shown

promise for the resolution of deficits in social cognition and mood problems in both autism and

schizophrenia (Andari et al., 2010; Feifel et al., 2010; Hollander et al., 2007).

7.1.4 Implications for development of novel therapeutics

There is an urgent need for new therapeutic agents for different NDDs in general and in

particular for ASD, for which there is a particular dearth relative to ADHD and schizophrenia

(Delorme et al., 2013; Ghosh et al., 2013; Hampson et al., 2012). Currently, only two drugs are

approved by the FDA specifically for the treatment of ASD: risperidone and aripiprazole

Page 144: Copy Number Variation in Neurodevelopmental Disorders

131

(Delorme et al., 2013; McPheeters et al., 2011). These drugs are aimed at reducing symptoms of

irritability associated with ASD and as such are ineffective at resolving core disorder symptoms

including communication deficits and repetitive behaviors (McPheeters et al., 2011; Politte and

McDougle, 2013). Their therapeutical applications are also restricted due to side effects such as

weight gain and sedation. The recent progress in ASD genetic research, including findings

presented in this thesis, could potentially facilitate the development of novel knowledge-based

therapeutics targeting the core molecular etiology of the disorder (Muglia, 2011). Rare CNV and

SNV findings have implicated hundreds of risk genes in ASD (Betancur, 2011; Devlin and

Scherer, 2012). At first glance, this extensive genetic heterogeneity would seem to pose

considerable challenges for the design and targeting of new pharmaceutical therapies. However,

many of these ASD risk genes often converge onto specific pathways. The most striking in terms

of their consistency across findings from multiple independent CNV and SNV studies are gene

networks involved in the formation, maturation and function of neuronal synapses. Examples of

such synaptic ASD risk genes first discovered in my thesis research are NRXN3 (Vaags et al.,

2012) and GPHN (Lionel et al., 2013) among others. It is now evident that the disruption of

synaptic homeostasis is an overarching etiological feature of a significant subset of cases with

ASD and other NDDs (Bourgeron, 2009; Grant, 2012; Guilmatre et al., 2009; Heck and Lu,

2012; Penzes et al., 2013; Ramocki and Zoghbi, 2008; Toro et al., 2010; Waltereit et al., 2013)

and resolution of these synaptic deficits could be the gateway to experimental therapeutics

(Canitano, 2013; Delorme et al., 2013). Two particularly powerful methodologies for modeling

synaptic defects in NDDs are animal models of specific mutations seen in human patients

(Nomura and Takumi, 2012) and patient derived induced pluripotent stem cells (iPSCs)

(Brennand et al., 2011; Mackay-Sim, 2013). In addition to yielding valuable insights into the

precise molecular mechanisms involved in NDD etiology, these technologies can also be used to

test novel candidate pharmaceutical agents with the goal of reversing NDD symptoms at the

organismal and cellular levels. For instance, exciting new research has shown that mice with

overexpression of the synaptic SHANK3 gene show dramatic resolution of their seizure and

manic symptoms upon treatment with valproate, a finding of direct relevance to similar

symptoms in human patients with SHANK3 duplications (Han et al., 2013). Independent work on

functional neurons produced by the iPSC differentiation of skin fibroblasts of ASD patients with

SHANK3 deletions has revealed that the administration of insulin-like growth factor (IGF1) can

Page 145: Copy Number Variation in Neurodevelopmental Disorders

132

compensate for the deficits in excitatory synaptic transmission observed in such neurons

(Shcheglovitov et al., 2013).

7.2 Future Research Directions

7.2.1 Follow-up of rare CNV findings in clinical genetic cohorts

The rare CNV findings presented in this thesis and the specific genes highlighted by them

represent a resource that can be used to guide future genetic studies. Replication of these findings

in much larger patient cohorts would provide further evidence that these are bona fide risk loci.

Discovery of additional patients with mutations at the regions reported in this thesis would

enable better understanding of the genotype-phenotype correlation and allow more precise

definition of the clinical features of individuals with such mutations. Particularly valuable

sources of genomic data for future replication and follow-up studies are the genome-wide CNV

results generated by clinical diagnostic laboratories during the genetic testing of patients referred

to them. Since microarray testing is now routinely performed as part of the clinical workup of

cases with ASD, ID, developmental delay and congenital anomalies, clinical diagnostic labs have

access to genome wide CNV data from thousands of such patients, which can be mined for

mutations at specific genes. As an illustration of this strategy, I present here the results of a

follow-up study (Lionel et al., 2014) of the ASTN2/TRIM32 locus, which was first implicated in

ASD and ADHD by rare deletions in the work presented in Chapter 4 (Lionel et al., 2011).

This follow-up screening study utilized microarray data from 89,985 individuals referred for

clinical microarray testing to 10 different molecular diagnostic centers in Canada, Denmark,

Italy, the United Kingdom and the United States (Table 7.1). The reasons for referral for clinical

microarray testing were systematically assessed at each of the study sites and 64,114 individuals

were determined to be NDD cases based on the presence of one or more of the following

phenotypes: ADHD, ASD, behavioral disorders, cognitive impairment, developmental delay, ID,

learning disability, macrocephaly, microcephaly, neurological disorders, OCD, psychoses,

seizures and speech/language disorders. The remaining 25,871 patients in the clinical cohorts,

whose reasons for referral for genetic testing did not contain any of the NDD terms listed above,

were counted as non-NDD cases. The gender composition of each clinical cohort was also

tabulated in order to test for sex-biased effects. For the purpose of statistical testing of findings in

Page 146: Copy Number Variation in Neurodevelopmental Disorders

133

the case cohorts, high resolution microarray data from 44,085 control individuals were inspected

for exonic deletions at the ASTN2/TRIM32 regions (Lionel et al., 2014). Fisher’s one-sided exact

test was used to test for enrichment of CNVs in cases versus non-NDD cases and versus controls

with a significance threshold of p < 0.05. A challenge in combining data across multiple case and

control cohorts is the heterogeneity of microarray platforms and the resulting differences in

probe coverage. The array platforms used for the control dataset had much higher probe densities

on average, both genome-wide and in the ASTN2/TRIM32 region of interest, than those

constituting the clinical dataset. The higher resolution of the control CNV dataset relative to the

cases decreases the likelihood of spurious enrichment findings driven by false negatives in the

controls and provides a conservative estimate of the significance and effect size of the findings.

Deletions impacting exons of ASTN2 were identified in 46 individuals (Figure 7.1). In the 20

individuals with parental DNA available for inheritance testing, the deletions were maternally

inherited in 10 (50%), paternally inherited in 8 (40%) and arose de novo in two individuals

(10%). This observed rate of de novo deletions in the cases (2/20) was significantly higher (one-

sided binomial test p = 0.017) than the expected genome-wide background rate of 1% for de

novo deletions in the general population. The latter rate was derived from findings in control

individuals by previous work (Levy et al., 2011; Sanders et al., 2011; Xu et al., 2008), which

used microarrays of similar resolution to those in this study. De novo ASTN2 deletions have

recently been reported in individuals with ID (Bernardini et al., 2010; Vulto-van Silfhout et al.,

2013). The majority of the individuals with exonic ASTN2 deletions (40/46) belonged to the case

subset of 64,114 with NDD phenotypes (Table 7.2). Significant enrichment of ASTN2 exonic

deletions (p = 0.01) was observed in the NDD cases compared with the 25,871 non-NDD cases

in the clinical cohort. Inspection of microarray data from 44,085 control individuals revealed 18

exonic ASTN2 deletions and one deletion exonic solely to TRIM32. Deletions around the 3’ end

of ASTN2 (Figure 7.1) that disrupted multiple transcript isoforms of ASTN2 were significantly

enriched in NDD cases versus controls (p = 0.002) but not in non-NDD cases versus controls

(Table 7.3). A striking difference was observed in sex-specific frequencies for ASTN2 exonic

deletions among the NDD cases, with an excess of such events in male cases compared with

female cases (p = 0.003). Deletion frequencies among the controls did not exhibit a sex-specific

difference. The sex-specific difference was also observed for the deletions at the 3’ end of

ASTN2 that disrupted multiple transcript isoforms of the gene. Such deletions were enriched in

Page 147: Copy Number Variation in Neurodevelopmental Disorders

134

the male NDD cases compared with male controls (p = 0.005) but not in female NDD cases

relative to female controls.

Page 148: Copy Number Variation in Neurodevelopmental Disorders

135

Table 7.1. Clinical diagnostic cohorts used for screening of rare ASTN2/TRIM32 CNVs

Cohort1 Total # of cases

Total # exonic ASTN2 losses2

# NDD individuals (males/females)

# exonic ASTN2 losses in NDD individuals

Alberta Children’s Hospital 1,619 1 1,170 (675/495) 1

BBGRE 14,847 3 9,650 (6,486/3,164) 1

Boston Children's Hospital 7,320 6 6,623 (4,152/2,471) 5

Credit Valley Hospital 3,552 1 3,098 (2,055/1,043) 1

Hospital for Sick Children 7,411 5 4,863 (3,267/1,596) 5

Italian diagnostic labs3 6,626 3 5,568 (3,272/2,296) 3

Mayo Clinic 19,131 6 11,208 (7,282/3,926) 5

Odense University Hospital 551 2 289 (182/107) 2

Signature Genomics 26,973 13 19,690 (11,617/8,073)4 11

The Centre for Applied Genomics5 1,955 6 1,955 (1,450/505) 6

Total 89,985 46 losses 64,114 (40,438/23,676) 40 losses

Abbreviations: BBGRE, Brain and Body Genetic Resource Exchange database; NDD, Neurodevelopmental disorders.

1 Ten different molecular diagnostic sites that contributed clinical microarray data for this study. Further descriptions are available in the following references: (Ahn et al., 2013). for BBGRE data, (Chen et al., 2013) for Boston Children’s Hospital data, (Hodge et al., 2013) for Mayo Clinic data and (Rosenfeld et al., 2013)for Signature Genomics data.

2 All deletions in the clinical cohorts smaller than 6 Mb that overlapped one or more exons of ASTN2 were included in the counts above.

3.Italian cohort includes data from individuals tested at five different molecular diagnostic sites: Cremona, Pavia, San Giovanni Rotondo, Tor Vergata and Troina.

4 Sex distribution of the Signature cohort was extrapolated from that found in a sampling cross-section of the data by (Ernst et al., 2012)

5 The Centre for Applied Genomics cohort includes 415 Canadian individuals with ADHD (Lionel et al., 2011) genotyped on the Affymetrix 6.0 (n = 248) and the Affymetrix CytoScan HD (n = 167), 174 individuals with OCD genotyped on the Illumina Omni2.5M-quad, and 1,366 Canadian individuals with ASD (Sato et al., 2012) genotyped on one of the following microarray platforms: Affymetrix 6.0, Agilent 1M, Illumina 1M or Affymetrix CytoScan HD.

Page 149: Copy Number Variation in Neurodevelopmental Disorders

136

These findings suggest greater penetrance in males relative to females. Similar sex-specific

effects have been recently reported for CNVs at other autosomal loci including SHANK1

deletions in ASD risk (Sato et al., 2012) and 16p13.11 CNVs in NDD risk (Tropeano et al.,

2013). Such autosomal loci with male-biased penetrance, together with X-linked risk genes,

could help explain the skewed sex ratios observed in the prevalence of NDDs such as ADHD,

ASD and ID. Given the enrichment of exonic ASTN2 deletions in cases relative to controls, the

clinical features of individuals with such CNVs were examined for phenotypic trends. The

reasons for referral for genetic testing, together with more detailed clinical phenotypes when

available, were obtained. The major commonality among these individuals was some form of a

NDD phenotype, which was observed in 87% of ASTN2 deletion cases (n = 41). The most

common NDD diagnoses were speech/language delay (n=18), ASD (n=12), ADHD (n=9),

generalized developmental delay (DD) (n=9), anxiety (n=9), obsessive compulsive disorder

(OCD) (n=6) and learning disability (n=8) (Table 7.2). These findings are in line with the initial

observations in Chapter 4, where ASD and ADHD were the primary phenotypes observed in

connected with exonic ASTN2 deletions. Gross motor delay was present in 12 cases and fine

motor delay in 6 cases. In addition, a wide range of dysmorphic features were present in ASTN2

deletion carriers, although macrocephaly was the only feature found to be common in more than

10% of the cases (n=7). Phenotypic information was available for 12 (8 mothers and 4 fathers) of

the 17 parents with exonic ASTN2 deletions. Seventy-five percent of the paternal carriers and

50% of the maternal carriers reported some form of neurodevelopmental trait (primarily anxiety,

depression, learning disabilities, dyslexia and in some cases formal adult diagnoses of ASD or

ADHD), though usually milder than those seen in the probands.

In summary, this follow-up study replicated the initial enrichment observed for ASTN2 deletions

in Chapter 4 and confirmed ASD and ADHD as the most frequent NDD diagnoses observed in

connection with such mutations. In addition, the massive size of the new clinical datasets

provided sufficient statistical power to test additional trends such as the sex-specific gender

differences and the differing penetrance of the deletions based on location within the gene. These

data emphasize the need to characterize rare CNVs (and other genetic variants) in the context of

large case and control cohorts in order to extract meaningful genotype and phenotype data

necessary for proper clinical genetic interpretation.

Page 150: Copy Number Variation in Neurodevelopmental Disorders

137

Figure 7.1. Exonic ASTN2/TRIM32 deletions in clinical and control cohorts

Filled red bars represent deletions detected in individuals with NDD phenotypes. Empty red bars denote deletions in cases without known NDD phenotypes (from available clinical information) and in controls. Shaded gray region denotes the critical region defined by deletions that disrupt multiple isoforms of ASTN2. Numbers adjacent to the bars are the randomized sample identifiers of individuals with the deletions and correlate with information in Table 7.2. Gender information was not available for the two control individuals marked with ∗ at the bottom of the figure. Dashed purple lines intersect deletions that overlap exons shared by multiple ASTN2 isoforms and dashed green lines intersect those affecting only the long isoform. Dashed vertical black line intersects deletions that overlap an exon of TRIM32. Genomic locations and coordinates are based on hg18 (NCBI36). The three transcript isoforms of ASTN2 possessing different numbers of exons are depicted including the long isoform (NM_014010) and two shorter isoforms (NM_198186 and NM_001184735). The three other shorter isoforms of the gene (NM_198187, NM_198188 and NM_001184734) have the same number and location of exons as NM_198186 but differ slightly in the length of their first and terminal exons and UTRs.

Page 151: Copy Number Variation in Neurodevelopmental Disorders

138

Table 7.2. Genetic and clinical information for individuals with rare ASTN2 deletions

Case# Sex Site1 CNV coordinates (hg18) CNV size CNV NDD phenotypes2

1 M TCAG chr9:117,954,428-118,356,243 401,816 loss ADHD, Anx, LD, SD

2 M OUH chr9:118,055,333-118,646,904 591,572 loss DD, MD

3 M ACH chr9:118,069,649-119,679,670 1,610,022 loss Chiari I malformation

4 M MC chr9:118,130,121-119,029,857 899,737 loss Hydrocephalus, Mac

5 M HSC chr9:118,164,272-118,358,705 194,434 loss ADHD, ASD, DD, LD, Mic, MD, OCD,

SD

6 M SG chr9:118,291,060-118,661,674 370,615 loss Anx, ASD, LD, Mac, MD, OCD, SD

7 M SG chr9:118,327,395-118,595,433 268,039 loss ASD, DD

8 M CVH chr9:118,342,936-118,685,436 342,501 loss DD, seizures

9 M BCH chr9:118,358,646-118,459,563 100,918 loss Anx, DD, Mac

10 M ITA chr9:118,358,837-118,728,270 369,434 loss ASD, SD

11 M HSC chr9:118,390,436-118,524,432 133,997 loss DD, ID, LD, MD, SD

12 M SG chr9:118,395,767-118,520,501 124,735 loss ADHD, SD, CNS disorder

13 M SG chr9:118,395,767-118,595,433 199,667 loss Hydrocephalus

14 M TCAG chr9:118,407,129-118,523,510 116,382 loss Anx, ASD, SD

15 M SG chr9:118,421,170-118,683,092 261,923 loss Behavioral problems

16 M SG chr9:118,430,585-118,569,556 138,972 loss DD, MD, seizures, SD

17 M SG chr9:118,430,585-118,610,907 180,323 loss DD, MD, SD

18 M DEC chr9:118,440,935-118,584,415 143,481 loss ID, Mac

19 M MC chr9:118,459,294-118,616,407 157,114 loss Chronic static encephalopathy

20 M BCH chr9:118,469,713-118,524,312 54,600 loss DD, LD, tics

21 M TCAG chr9:118,479,893-118,627,637 147,745 loss ADHD, Anx, OCD, tics

22 M TCAG chr9:118,480,042-118,570,447 90,406 loss ADHD, ASD, Mac, OCD, SD, seizures

23 M TCAG chr9:118,481,308-118,654,031 172,724 loss Anx, OCD

24 M BCH chr9:118,488,204-118,558,274 70,071 loss ASD, seizures

25 M BCH chr9:118,488,204-118,700,657 212,454 loss -

Page 152: Copy Number Variation in Neurodevelopmental Disorders

139

26 M TCAG chr9:118,493,276-118,670,608 177,333 Loss ADHD, Anx, LD

27 M BBG chr9:118,497,759-118,661,673 163,915 Loss -

28 M MC chr9:118,502,294-118,616,407 114,114 Loss ADHD, Anx, ASD, Mac, SD

29 M SG chr9:118,530,143-118,569,556 39,414 Loss Mic

30 M MC chr9:118,541,180-118,685,465 144,286 Loss Dizziness, dysgraphia, migraines

31 M BCH chr9:118,572,937-118,637,250 64,314 Loss DCD, DD, MD, SD

32 M ITA chr9:118,580,317-118,814,591 234,275 Loss ASD, structural brain anomaly

33 M SG chr9:118,580,891-118,630,518 49,628 Loss DD

34 M HSC chr9:118,616,347-118,907,058 290,712 Loss Anx, ASD, hydrocephalus, LD, Mac,

MD, structural brain anomaly

35 M SG chr9:118,620,063-118,781,101 161,039 Loss ASD

36 M BCH chr9:118,743,266-118,991,977 248,712 Loss ASD, DD

37 M ITA chr9:118,840,027-118,935,027 95,001 Loss Behavioral problems, ID, OCD, SD

38 M SG chr9:118,874,947-119,109,618 234,672 Loss -

39 F SG chr9:118,199,243-118,248,950 49,708 Loss ADHD, DD, Mic, SD

40 F BBG chr9:118,202,805-118,227,359 24,555 Loss MD, plagiocephaly, SD

41 F HSC chr9:118,202,811-118,459,563 256,753 Loss Mac

42 F HSC chr9:118,390,436-118,524,432 133,997 Loss ID, MD, SD, seizures

43 F SG chr9:118,481,080-118,610,907 129,828 Loss -

44 F BBG chr9:118,497,759-118,661,673 163,915 Loss -

45 F OUH chr9:118,608,198-118,669,889 61,692 Loss ADHD, MD, SD

46 F MC chr9:118,728,240-118,992,036 263,797 Loss -

47 F MC chr9:118,829,818-118,890,615 60,798 Loss Septo-optic dysplasia

1. Molecular diagnostic testing site of origin of the case: ACD, Alberta Children’s Hospital; BBG, Brain and Body Genetic Resource Exchange (BBGRE); BCH, Boston Children’s Hospital; CVH, Credit Valley Hospital; DEC DECIPHER database; HSC, The Hospital for Sick Children; ITA, Italian diagnostic labs; MC, Mayo Clinic; OUH, Odense University Hospital; SG, Signature Genomics; TCAG, The Centre for Applied Genomics.

2. NDD trait abbreviations: ADHD, attention deficit hyperactivity disorder; Anx, anxiety; ASD, autism spectrum disorder; CNS, central nervous system; DCD, developmental coordination disorder; DD, developmental delay; ID, intellectual disability; LD, learning disability; Mac, macrocephaly; MD, motor delay; Mic, microcephaly; NDD; neurodevelopmental disorder; OCD, obsessive compulsive disorder; SD, speech delay. “-“indicates non-NDD cases (no NDD terms were present in their reasons for referral for clinical microarray testing)

Page 153: Copy Number Variation in Neurodevelopmental Disorders

140

Table 7.3. Results of CNV enrichment analysis of ASTN2/TRIM32 locus

Total NDD dataset Male NDD Female NDD

CNV type NDD cases

Controls p-

value1 NDD males

Male controls2

p-value1

NDD females

Female controls2

p-value1

Exonic ASTN2 losses

40 18 0.084 34 10 0.328 6 6 0.773

Exonic losses affecting only long

ASTN2 isoform 13 13 0.876 11 9 0.976 2 3 0.883

Exonic losses affecting multiple ASTN2 isoforms

27 5 0.002 23 1 0.005 4 3 0.64

Exonic losses affecting TRIM323

23 6 0.019 22 2 0.024 1 3 0.964

Abbreviations: NDD, Neurodevelopmental disorders

1. P-values are from one-sided Fisher’s exact test. Values in bold are significant at threshold of p < 0.05.

2. Gender information was available for 33,171 control individuals, and this subset was used for the sex-specific enrichment analysis.

3. All deletions which affected TRIM32 in the cases also overlapped exon(s) of ASTN2. One of the exonic TRIM32 deletions in the controls did not overlap an ASTN2 exon.

Page 154: Copy Number Variation in Neurodevelopmental Disorders

141

7.2.2 Functional characterization of rare CNVs and risk genes

Following the discovery of a rare CNV implicating a specific gene in NDD risk, experimental

and computational investigations can be performed in order to characterize the gene’s biological

role, explore causal links between the CNV and gene function and better understand the

molecular mechanisms by which the CNV influences phenotype. Examples of such

investigations include assessing the CNV’s effect on gene expression, quantifying expression of

the gene in different brain tissue types across different neurodevelopmental time points,

measuring relative expression of different transcript isoforms of the gene, and performing

comparative genetics analyses to identify evolutionarily conserved regions within the gene and

functionally important domains of its protein. As an illustration of some of these approaches, I

present here data from the functional characterization of ASTN2 following its implication in

NDD risk by exonic deletions as described in the previous section and in Chapter 4.

Since exonic deletions near the 3’ terminal end of ASTN2, which affected multiple isoforms of

the gene, were strongly enriched in NDD cases compared to controls, the functional impact of

these deletions on the gene’s expression was investigated. The expression testing was performed

in lymphoblast cell lines from six individuals with such deletions including two patients from the

follow-up study described in the previous section and their families (Figure 7.2A). Expression

was significantly lower in ASTN2 deletion carriers (Figure 7.2B) compared with expression

levels in nine individuals with two copies of ASTN2. These results suggest that the deletions

could disrupt the expression of all transcript isoforms of ASTN2 and potentially lead to complete

haploinsufficiency of the protein. This may have more severe phenotypic consequences than

those deletions affecting only the long transcript isoform. Similar trends of greater penetrance of

deletions impacting multiple isoforms of a gene, relative to those overlapping a single isoform,

were also observed for the NRXN3 deletions presented in Chapter 3 (Vaags et al., 2012) and at

other loci such as NRXN1 (Schaaf et al., 2012) and AUTS2 (Beunders et al., 2013) in connection

with risk for NDDs.

Page 155: Copy Number Variation in Neurodevelopmental Disorders

142

Figure 7.2. Impact of ASTN2 deletions on gene expression in lymphoblast cell lines

A) Pedigrees of the two families used for ASTN2 expression analysis. Arrows indicate the index probands of the pedigrees (corresponding to patients 14 and 22 in Figure 7.1 and Table 7.2). Individuals with black shading have an ASD diagnosis while the person represented by gray shading reported depression and anxiety issues. B) Relative ASTN2 expression in lymphoblast cell lines from the six individuals (3 males and 3 females) with ASTN2 deletions in the two pedigrees were compared to lymphoblast cell-lines from 9 individuals (5 males and 4 females) with two copies of ASTN2. The latter group included the three individuals without deletions in the two pedigrees as well as six additional samples with two copies of ASTN2. The lymphoblast cell lines were cultured as previously described (Seno et al., 2011). Total RNA was extracted using Qiagen RNeasy mini kit with DNase I treatment (QIAGEN, Valencia, CA, USA). cDNA was synthesized using Superscript III First strand Synthesis Supermix (Invitrogen, Carlsbad, CA, USA) with 1 µg of poly (A+) or total DNase I treated RNA as a template. The expression was measured using two different primer pairs which detected all ASTN2 isoforms. ASTN2 expression was normalized using GAPDH (dCt) and the fold change was calculated using the ∆∆Ct method. Student’s T-test was used to assess the expression difference between the two sample groups for statistical significance.

Page 156: Copy Number Variation in Neurodevelopmental Disorders

143

Given the location-dependent penetrance observed for the ASTN2 exonic deletions (Table 7.3),

the average nucleotide conservation of all exons in ASTN2 (Figure 7.3A) and TRIM32 was

assessed across vertebrates, placental mammals and primates (Figure 7.3B). Exons with coding

sequence were well-conserved across all vertebrate species. Interestingly, the conservation was

much lower between humans and other vertebrates for the exons unique to the shorter isoforms

(exons 1B and 5C in Figure 7.3A). Additionally, inspection of EST databases such as Aceview,

Ensembl and Fantom did not reveal any evidence for the presence of shorter 3’-terminal Astn2

isoforms in mouse. These observations suggest that the unique exons of the shorter isoforms (and

their alternative transcription site in exon 1B) are of more recent evolutionary origin and are

potentially derived in or just prior to the primate lineage. In support of this claim, both exons 1B

and 5C contain transposable elements, which have recently been shown to contribute novel

regulatory elements and thus give rise to new transcript isoforms (Jacques et al., 2013).

Although there was no evident difference in the average nucleotide conservation between ASTN2

coding exons, the amino acid alignment of eight ASTN2 protein orthologs revealed that the C-

terminal end exhibits a greater degree of conservation compared to the rest of the protein (Figure

7.4). This section of the protein corresponds to the 3’-terminal region of the gene exhibiting the

strongest enrichment of exonic deletions in male NDD cases (Figure 7.1) and encodes the

fibronectin type III domain (Figure 7.4). Interestingly, this domain is also found in other genes

implicated in NDDs such as the contactins (Burbach and van der Zwaag, 2009). In addition, the

data suggest that the shorter transcript isoforms are of functional importance especially during

early brain development. This claim is supported by mRNA quantification assays in which the

shorter isoforms account for forty percent of the total expression of the gene in fetal brain and

were detected in all the different brain regions tested (Figure 7.5). The evolutionary conservation

analyses suggest that the first exon shared by the shorter ASTN2 isoforms is of more recent

evolutionary origin and could have a function specific to primates. There is no indication from

previous functional work (Wilson et al., 2010) or from mouse EST databases for existence of

shorter 3’-terminal Astn2 transcript isoforms in the mouse. Taken together, the recent emergence

of the shorter isoforms comprising a functionally important domain and their widespread

expression in the human brain could indicate primate-specific involvement in neurodevelopment.

Page 157: Copy Number Variation in Neurodevelopmental Disorders

144

Figure 7.3. Exon conservation analysis of ASTN2 and TRIM32

(A) Schematic presentation of the ASTN2 transcript isoforms. The three transcript isoforms of ASTN2 possessing different numbers of exons are depicted including the long isoform (NM_014010) and two shorter isoforms (NM_198186 and NM_001184735). The three other shorter isoforms of the gene (NM_198187, NM_198188 and NM_001184734) have the same number and location of exons as NM_198186 but differ slightly in the length of their first and terminal exons and UTRs. The ‘∗’ symbol denotes exons with variable length in different isoforms. (B) Conservation profile of ASTN2 and TRIM32 exons across vertebrates, placental mammals and primates. The nucleotide conservation scores for each base present in all ASTN2 and TRIM32 exons were computed using the PhyloP program based on alignment between 46 vertebrate species including 23 placental mammals and eight primates. The average score was calculated for each exon and for the unique UTRs.

Page 158: Copy Number Variation in Neurodevelopmental Disorders

145

Figure 7.4. Amino acid conservation analysis of ASTN2 protein

The amino acid sequences of 1:1 ASTN2 orthologs from the following eight species were downloaded from Ensembl: human ASTN2 (ENSP00000354504): macaque (ENSMMUP00000039777), chimpanzee (ENSPTRP00000036394), horse (ENSECAP00000003457), mouse (ENSMUSP00000065786), dog (ENSCAFP00000005232), chicken (ENSGALP00000011381) and clawed frog (ENSXETP00000047008). Following multiple sequence alignment using ClustalW2, amino acid conservation was quantified using Scorecons server. The known protein domains and features of ASTN2 are shown in the schematic illustration above the chart. The signal peptide (SP) is marked red, trans-membrane (TM) domains are marked light blue, EGF/laminin superfamily (EGF) domains are marked light green, the MACPF domain is marked blue and the fibronectin domain type 3 (FN III) is marked dark green. The critical region with enrichment of exonic deletions affecting multiple ASTN2 isoforms in NDD cases from Figure 7.1 is shown in grey. Vertical red dashed lines correspond to the exon boundaries.

Page 159: Copy Number Variation in Neurodevelopmental Disorders

146

Figure 7.5. Relative expression levels of ASTN2 transcript isoforms

Figure on the left shows the expression of ASTN2 transcript isoforms (long, shorter and all) in eight different human brain regions. ACTB was used as a control gene. Results are from RT-PCR analysis using primers with locations as depicted by red arrows in Figure 7.3A. Abundant expression was detected for primers amplifying exons present in all isoforms (ASTN2 all) and for exons specific to the long ASTN2 isoform NM_014010 (ASTN2 long isoform) in all the brain regions. The shorter transcripts (NM_198187, NM_198188, NM_001184734 and NM_001184735) were expressed at a lower level (ASTN2 shorter isoforms).

Figure on the right shows the results from quantification of ASTN2 isoform expression levels in triplicate from adult brain and fetal brain RNA samples by qRT–PCR (standard curve method). The expression was normalized using ACTB as a housekeeping gene, and the expression ratio is relative to the expression from all ASTN2 isoforms (mean ratio+standard deviation). The results were replicated using normalization against GAPDH, a housekeeping gene. A change was observed in the expression ratios between the developmental stages. In the fetal brain, the shorter isoforms accounted for approximately 40% of the total ASTN2 expression, which decreased to 20% in the adult brain. This suggests a functional role for the shorter isoforms during early development.

Page 160: Copy Number Variation in Neurodevelopmental Disorders

147

Figure 7.6. Expression profile of ASTN2 across human brain development

The gene level expression profiles of ASTN2 across developmental time points in nine regions of the human brain; amygdala (AMY), cerebellar cortex (CBC), diencephalon (DIE), frontal cortex (FC), hippocampus (HIP), occipital cortex (OC), parietal cortex (PC), temporal cortex (TC) and ventral forebrain (VF). This analysis utilized expression data from the BrainSpan database (www.brainspan.org) (Kang et al., 2011), which contains extensive transcriptome profiles for 16 brain regions from 41 individuals. The age range of the subjects spans from 8 post-conception weeks to 40 years. The data was quantile normalized and the gene level expression was averaged across donors for each time point in the nine regions. The expression levels of ASTN2 in the different brain regions were plotted across age range using the R package ggplot2. Smooth curve lines were computed by loess using a span of 0.4.

Page 161: Copy Number Variation in Neurodevelopmental Disorders

148

The spatiotemporal expression pattern of ASTN2 in different brain regions during human

development was analyzed by utilizing comprehensive gene expression data from the BrainSpan

database (Kang et al., 2011). Overall, ASTN2 is expressed at a moderate level throughout

development and exhibits an increase in the late prenatal period and during postnatal

development in many of the brain regions (Figure 7.6). The highest level of ASTN2 expression

was observed in the cerebellar cortex (CBC), where the expression increase during infancy and

early childhood is in concordance with the expression pattern previously reported in mice

(Wilson et al., 2010). Other genes from the BrainSpan database exhibiting similar expression

patterns in the CBC were involved in synaptic transmission and plasticity. Interestingly, during

the prenatal period, ASTN2 has a dynamic spike in the expression around 12-13 gestational

weeks in the neocortical regions (frontal cortex, parietal cortex and occipital cortex) (Figure 7.6).

This expression pattern is enriched in neuronal development genes involved in axonogenesis and

neuron differentiation. The dynamic expression pattern of ASTN2 together with the biological

annotation of genes with similar expression patterns suggests that ASTN2 could have an

important role in both prenatal and postnatal brain development.

Comparison of the human brain expression profile of ASTN2 with previous work in mice

revealed both intriguing differences and similarities. The striking prenatal spike in ASTN2

expression in the neocortical regions, towards the end of the first trimester, has not been reported

in the embryonic mouse brain (Wilson et al., 2010). This finding might indicate transcriptional

regulation patterns or additional functions of the protein and/or transcript isoforms specific to

primates. Interestingly, this period in the developmental timeline of the human cerebral cortex is

marked by extensive neuronal migration, increasing axonal outgrowth and the formation of the

early synapses (Kang et al., 2011; Pescosolido et al., 2012). Several genes previously implicated

in risk for ASD (and also other NDDs) such as DOCK4, and NRCAM also exhibit increasing

cortical expression around this time-point (Pagnamenta et al., 2010; Sakurai, 2012). In

concordance with the experimental work from mouse (Wilson et al., 2010), the highest ASTN2

expression is expressed in the cerebellar cortex shortly after birth. Several studies investigating

the neuropathology of ASD (Fatemi et al., 2012) and ADHD (O'Halloran et al., 2012) have

consistently highlighted the cerebellum as a major region of interest. Reported neuroanatomical

abnormalities include disrupted neuronal migration in the cerebellar cortex (Fatemi et al., 2012)

as well as reduction in volume of the cerebellar vermis (Anagnostou and Taylor, 2011), which

Page 162: Copy Number Variation in Neurodevelopmental Disorders

149

has been linked to repetitive and stereotyped behavior in ASD (Pierce and Courchesne, 2001).

Furthermore, the number of Purkinje cells, one of the main cell types in the cerebellum, has been

found to be decreased by up to 50% in individuals with ASD (Fatemi et al., 2012). Astn2 has also

been shown to be very highly expressed in the cerebellum in general, and in the Purkinje cells in

particular, during both embryonic and postnatal development (Wilson et al., 2010).

7.2.3 Higher resolution genomic studies of NDDs

Although substantial initial progress has been made over the past decade in the discovery of the

specific genes involved in NDD risk, this monumental task is far from complete. For instance, in

ASD genetics research, it is estimated that the etiological mutations (in the form of rare CNVs,

SNVs and chromosomal abnormalities) identified to date account for around 20% of cases

(Devlin and Scherer, 2012). It is clear that the majority of NDD cases are still of idiopathic or

unknown causation. The continued elucidation of the genetic architectures of these conditions is

critical for further advances in molecular diagnostics, clinical screening and therapeutic

development. Other intriguing sources of genetic risk for NDDs remain relatively unexplored

including random monoallelic inactivation (Chess, 2012) and specific somatic mutations in

affected parts of the brain that are absent in other tissues (Insel, 2014). Additionally, the roles of

“modifier genes” and gene-gene interactions have to be systematically investigated for their

potential to explain the complex and variable phenotypes observed in the clinical presentation of

different NDDs.

While genome-wide microarrays such as those used in this thesis research have been effective at

detecting CNVs of relatively large sizes, they suffer from certain limitations. These include their

inability to detect either smaller genetic changes such as SNVs and indels or balanced events

such as inversions and translocations. The scope of CNV detection by microarrays is also limited

to those regions with sufficient probe coverage, which is often preferentially biased towards

regions with known genes at the expense of intergenic regions. As a result of these deficiencies,

broad swathes of the genetic variant spectrum are missed by microarrays, thus necessitating

higher-resolution genomics technology for NDD gene discovery. Arguably the most promising

of these new approaches is whole genome sequencing (WGS) technology. By presenting a

comprehensive and unbiased view of the total spectrum of genetic variation, WGS represents a

considerable advance from older genetic technology. WGS studies of individuals with NDD

Page 163: Copy Number Variation in Neurodevelopmental Disorders

150

have the potential to substantially increase the detection yield of genetic testing by capturing all

classes of genetic mutations within a patient in a single experiment. The first such applications of

WGS in ASD genetic research have indeed found high yields of clinically relevant mutations (in

~50% of families) and identified specific mutational hotspots missed by previous microarray and

exome sequencing studies (Jiang et al., 2013; Michaelson et al., 2012). The exponentially

increasing use of WGS in psychiatric genetic research and its ongoing transition into clinical

diagnostic laboratories will catalyze the next significant advances in uncovering the genomic

architectures of NDDs.

Page 164: Copy Number Variation in Neurodevelopmental Disorders

151

References

Abrahams, B.S., Arking, D.E., Campbell, D.B., Mefford, H.C., Morrow, E.M., Weiss, L.A., Menashe, I., Wadkins, T., Banerjee-Basu, S., and Packer, A. (2013). SFARI Gene 2.0: a community-driven knowledgebase for the autism spectrum disorders (ASDs). Mol Autism 4, 36.

Abrahams, B.S., and Geschwind, D.H. (2008). Advances in autism genetics: on the threshold of a new neurobiology. Nat Rev Genet 9, 341-355.

Adams, N.C., Tomoda, T., Cooper, M., Dietz, G., and Hatten, M.E. (2002). Mice that lack astrotactin have slowed neuronal migration. Development 129, 965-972.

Ahn, J.W., Dixit, A., Johnston, C., Ogilvie, C.M., Collier, D.A., Curran, S., and Dobson, R.J. (2013). BBGRE: brain and body genetic resource exchange. Database (Oxford) 2013, bat067.

Ahuja, A., Martin, J., Langley, K., and Thapar, A. (2013). Intellectual disability in children with attention deficit hyperactivity disorder. J Pediatr 163, 890-895 e891.

Ambrosini, P.J. (2000). Historical development and present status of the schedule for affective disorders and schizophrenia for school-age children (K-SADS). J Am Acad Child Adolesc Psychiatry 39, 49-58.

American Psychiatric Association (2013). Diagnostic and statistical manual of mental disorders (5th ed.) (Arlington, VA, American Psychiatric Publishing).

Amiet, C., Gourfinkel-An, I., Bouzamondo, A., Tordjman, S., Baulac, M., Lechat, P., Mottron, L., and Cohen, D. (2008). Epilepsy in autism is associated with intellectual disability and gender: evidence from a meta-analysis. Biol Psychiatry 64, 577-582.

Anagnostou, E., and Taylor, M.J. (2011). Review of neuroimaging in autism spectrum disorders: what have we learned and where we go from here. Mol Autism 2, 4.

Anagnostou, E., Zwaigenbaum, L., Szatmari, P., Fombonne, E., Fernandez, B.A., Woodbury-Smith, M., Brian, J., Bryson, S., Smith, I.M., Drmic, I., et al. (2014). Autism spectrum disorder: advances in evidence-based practice. CMAJ.

Anckarsater, H., Stahlberg, O., Larson, T., Hakansson, C., Jutblad, S.B., Niklasson, L., Nyden, A., Wentz, E., Westergren, S., Cloninger, C.R., et al. (2006). The impact of ADHD and autism spectrum disorders on temperament, character, and personality development. Am J Psychiatry 163, 1239-1244.

Andari, E., Duhamel, J.R., Zalla, T., Herbrecht, E., Leboyer, M., and Sirigu, A. (2010). Promoting social behavior with oxytocin in high-functioning autism spectrum disorders. Proc Natl Acad Sci USA 107, 4389-4394.

Page 165: Copy Number Variation in Neurodevelopmental Disorders

152

Andrulis, I.L., Boyd, N.F., and Sutherland, H. (1997). New Ontario familial breast cancer registry to facilitate genetic and epidemiologic studies. Can Fam Physician 43, 949-950.

Anney, R., Klei, L., Pinto, D., Almeida, J., Bacchelli, E., Baird, G., Bolshakova, N., Bolte, S., Bolton, P.F., Bourgeron, T., et al. (2012). Individual common variants exert weak effects on the risk for autism spectrum disorderspi. Hum Mol Genet 21, 4781-4792.

Anney, R., Klei, L., Pinto, D., Regan, R., Conroy, J., Magalhaes, T.R., Correia, C., Abrahams, B.S., Sykes, N., Pagnamenta, A.T., et al. (2010). A genome-wide scan for common alleles affecting risk for autism. Hum Mol Genet 19, 4072-4082.

Arcos-Burgos, M., Castellanos, F.X., Pineda, D., Lopera, F., Palacio, J.D., Palacio, L.G., Rapoport, J.L., Berg, K., Bailey-Wilson, J.E., and Muenke, M. (2004). Attention-deficit/hyperactivity disorder in a population isolate: linkage to loci at 4q13.2, 5q33.3, 11q22, and 17p11. Am J Hum Genet 75, 998-1014.

Arnold, P.D., Siegel-Bartelt, J., Cytrynbaum, C., Teshima, I., and Schachar, R. (2001). Velo-cardio-facial syndrome: Implications of microdeletion 22q11 for schizophrenia and mood disorders. Am J Med Genet 105, 354-362.

Autism and Developmental Disabilities Monitoring Network (2012). Prevalence of autism spectrum disorders--Autism and Developmental Disabilities Monitoring Network, 14 sites, United States, 2008. MMWR Surveill Summ 61, 1-19.

Awadalla, P., Gauthier, J., Myers, R.A., Casals, F., Hamdan, F.F., Griffing, A.R., Cote, M., Henrion, E., Spiegelman, D., Tarabeux, J., et al. (2010). Direct measure of the de novo mutation rate in autism and schizophrenia cohorts. Am J Hum Genet 87, 316-324.

Bailey, A., Le Couteur, A., Gottesman, I., Bolton, P., Simonoff, E., Yuzda, E., and Rutter, M. (1995). Autism as a strongly genetic disorder: evidence from a British twin study. Psychol Med 25, 63-77.

Baker, P., Piven, J., Schwartz, S., and Patil, S. (1994). Brief report: duplication of chromosome 15q11-13 in two individuals with autistic disorder. J Autism Dev Disord 24, 529-535.

Bakker, S.C., van der Meulen, E.M., Buitelaar, J.K., Sandkuijl, L.A., Pauls, D.L., Monsuur, A.J., van 't Slot, R., Minderaa, R.B., Gunning, W.B., Pearson, P.L., et al. (2003). A whole-genome scan in 164 Dutch sib pairs with attention-deficit/hyperactivity disorder: suggestive evidence for linkage on chromosomes 7p and 15q. Am J Hum Genet 72, 1251-1260.

Baribeau, D.A., and Anagnostou, E. (2013). A Comparison of Neuroimaging Findings in Childhood Onset Schizophrenia and Autism Spectrum Disorder: A Review of the Literature. Front Psychiatry 4, 175.

Barneveld, P.S., Pieterse, J., de Sonneville, L., van Rijn, S., Lahuis, B., van Engeland, H., and Swaab, H. (2011). Overlap of autistic and schizotypal traits in adolescents with Autism Spectrum Disorders. Schizophr Res 126, 231-236.

Page 166: Copy Number Variation in Neurodevelopmental Disorders

153

Barr, C.L., Feng, Y., Wigg, K., Bloom, S., Roberts, W., Malone, M., Schachar, R., Tannock, R., and Kennedy, J.L. (2000a). Identification of DNA variants in the SNAP-25 gene and linkage study of these polymorphisms and attention-deficit hyperactivity disorder. Mol Psychiatry 5, 405-409.

Barr, C.L., Wigg, K.G., Feng, Y., Zai, G., Malone, M., Roberts, W., Schachar, R., Tannock, R., and Kennedy, J.L. (2000b). Attention-deficit hyperactivity disorder and the gene for the dopamine D5 receptor. Mol Psychiatry 5, 548-551.

Bassett, A.S. (2011). Parental origin, DNA structure, and the schizophrenia spectrum. Am J Psychiatry 168, 350-353.

Bassett, A.S., and Chow, E.W. (2008). Schizophrenia and 22q11.2 deletion syndrome. Curr Psychiatry Rep 10, 148-157.

Bassett, A.S., Chow, E.W., Husted, J., Hodgkinson, K.A., Oechslin, E., Harris, L., and Silversides, C. (2009). Premature death in adults with 22q11.2 deletion syndrome. J Med Genet 46, 324-330.

Bassett, A.S., Costain, G., Fung, W.L., Russell, K.J., Pierce, L., Kapadia, R., Carter, R.F., Chow, E.W., and Forsythe, P.J. (2010a). Clinically detectable copy number variations in a Canadian catchment population of schizophrenia. J Psychiatr Res 44, 1005-1009.

Bassett, A.S., Marshall, C.R., Lionel, A.C., Chow, E.W., and Scherer, S.W. (2008). Copy number variations and risk for schizophrenia in 22q11.2 deletion syndrome. Hum Mol Genet 17, 4045-4053.

Bassett, A.S., McDonald-McGinn, D.M., Devriendt, K., Digilio, M.C., Goldenberg, P., Habel, A., Marino, B., Oskarsdottir, S., Philip, N., Sullivan, K., et al. (2011). Practical guidelines for managing patients with 22q11.2 deletion syndrome. J Pediatr 159, 332-339 e331.

Bassett, A.S., Scherer, S.W., and Brzustowicz, L.M. (2010b). Copy number variations in schizophrenia: critical review and new perspectives on concepts of genetics and disease. Am J Psychiatry 167, 899-914.

Baum, A.E., Akula, N., Cabanero, M., Cardona, I., Corona, W., Klemens, B., Schulze, T.G., Cichon, S., Rietschel, M., Nothen, M.M., et al. (2008). A genome-wide association study implicates diacylglycerol kinase eta (DGKH) and several other genes in the etiology of bipolar disorder. Mol Psychiatry 13, 197-207.

Ben-Shachar, S., Lanpher, B., German, J.R., Qasaymeh, M., Potocki, L., Nagamani, S.C., Franco, L.M., Malphrus, A., Bottenfield, G.W., Spence, J.E., et al. (2009). Microdeletion 15q13.3: a locus with incomplete penetrance for autism, mental retardation, and psychiatric disorders. J Med Genet 46, 382-388.

Berkel, S., Marshall, C.R., Weiss, B., Howe, J., Roeth, R., Moog, U., Endris, V., Roberts, W., Szatmari, P., Pinto, D., et al. (2010). Mutations in the SHANK2 synaptic scaffolding gene in autism spectrum disorder and mental retardation. Nat Genet 42, 489-491.

Page 167: Copy Number Variation in Neurodevelopmental Disorders

154

Bernardini, L., Alesi, V., Loddo, S., Novelli, A., Bottillo, I., Battaglia, A., Digilio, M.C., Zampino, G., Ertel, A., Fortina, P., et al. (2010). High-resolution SNP arrays in mental retardation diagnostics: how much do we gain? Eur J Hum Genet 18, 178-185.

Betancur, C. (2011). Etiological heterogeneity in autism spectrum disorders: more than 100 genetic and genomic disorders and still counting. Brain Res 1380, 42-77.

Beunders, G., Voorhoeve, E., Golzio, C., Pardo, L.M., Rosenfeld, J.A., Talkowski, M.E., Simonic, I., Lionel, A.C., Vergult, S., Pyatt, R.E., et al. (2013). Exonic deletions in AUTS2 cause a syndromic form of intellectual disability and suggest a critical role for the C terminus. Am J Hum Genet 92, 210-220.

Bevan Jones, R., Thapar, A., Lewis, G., and Zammit, S. (2012). The association between early autistic traits and psychotic experiences in adolescence. Schizophr Res 135, 164-169.

Biederman, J., Faraone, S.V., Mick, E., Spencer, T., Wilens, T., Kiely, K., Guite, J., Ablon, J.S., Reed, E., and Warburton, R. (1995). High risk for attention deficit hyperactivity disorder among children of parents with childhood onset of the disorder: a pilot study. Am J Psychiatry 152, 431-435.

Biederman, J., Mick, E., Faraone, S.V., Braaten, E., Doyle, A., Spencer, T., Wilens, T.E., Frazier, E., and Johnson, M.A. (2002). Influence of gender on attention deficit hyperactivity disorder in children referred to a psychiatric clinic. Am J Psychiatry 159, 36-42.

Biederman, J., Monuteaux, M.C., Mick, E., Spencer, T., Wilens, T.E., Silva, J.M., Snyder, L.E., and Faraone, S.V. (2006). Young adult outcome of attention deficit hyperactivity disorder: a controlled 10-year follow-up study. Psychol Med 36, 167-179.

Biederman, J., Petty, C.R., Fried, R., Kaiser, R., Dolan, C.R., Schoenfeld, S., Doyle, A.E., Seidman, L.J., and Faraone, S.V. (2008). Educational and occupational underattainment in adults with attention-deficit/hyperactivity disorder: a controlled study. J Clin Psychiatry 69, 1217-1222.

Binder, M.D., and Kilpatrick, T.J. (2009). TAM receptor signalling and demyelination. Neurosignals 17, 277-287.

Bochukova, E.G., Huang, N., Keogh, J., Henning, E., Purmann, C., Blaszczyk, K., Saeed, S., Hamilton-Shield, J., Clayton-Smith, J., O'Rahilly, S., et al. (2010). Large, rare chromosomal deletions associated with severe early-onset obesity. Nature 463, 666-670.

Bourgeron, T. (2009). A synaptic trek to autism. Curr Opin Neurobiol 19, 231-234.

Bradley, W.E., Raelson, J.V., Dubois, D.Y., Godin, E., Fournier, H., Prive, C., Allard, R., Pinchuk, V., Lapalme, M., Paulussen, R.J., et al. (2010). Hotspots of large rare deletions in the human genome. PLoS One 5, e9401.

Bray, N.J., Leweke, F.M., Kapur, S., and Meyer-Lindenberg, A. (2010). The neurobiology of schizophrenia: new leads and avenues for treatment. Curr Opin Neurobiol 20, 810-815.

Page 168: Copy Number Variation in Neurodevelopmental Disorders

155

Brennand, K.J., Simone, A., Jou, J., Gelboin-Burkhart, C., Tran, N., Sangar, S., Li, Y., Mu, Y., Chen, G., Yu, D., et al. (2011). Modelling schizophrenia using human induced pluripotent stem cells. Nature 473, 221-225.

Brooks-Kayal, A. (2010). Epilepsy and autism spectrum disorders: are there common developmental mechanisms? Brain Dev 32, 731-738.

Bruce, B., Thernlund, G., and Nettelbladt, U. (2006). ADHD and language impairment: A study of the parent questionnaire FTF (Five to Fifteen). Eur Child Adolesc Psychiatry 15, 52-60.

Brunetti-Pierri, N., Berg, J.S., Scaglia, F., Belmont, J., Bacino, C.A., Sahoo, T., Lalani, S.R., Graham, B., Lee, B., Shinawi, M., et al. (2008). Recurrent reciprocal 1q21.1 deletions and duplications associated with microcephaly or macrocephaly and developmental and behavioral abnormalities. Nat Genet 40, 1466-1471.

Bucan, M., Abrahams, B.S., Wang, K., Glessner, J.T., Herman, E.I., Sonnenblick, L.I., Alvarez Retuerto, A.I., Imielinski, M., Hadley, D., Bradfield, J.P., et al. (2009). Genome-wide analyses of exonic copy number variants in a family-based study point to novel autism susceptibility genes. PLoS Genet 5, e1000536.

Buizer-Voskamp, J.E., Muntjewerff, J.W., Strengman, E., Sabatti, C., Stefansson, H., Vorstman, J.A., and Ophoff, R.A. (2011). Genome-wide analysis shows increased frequency of copy number variation deletions in Dutch schizophrenia patients. Biol Psychiatry 70, 655-662.

Burbach, J.P., and van der Zwaag, B. (2009). Contact in the genetics of autism and schizophrenia. Trends Neurosci 32, 69-72.

Buxbaum, J.D., Baron-Cohen, S., and Devlin, B. (2010). Genetics in psychiatry: common variant association studies. Mol Autism 1, 6.

Caluseriu, O., Mirza, G., Ragoussis, J., Chow, E.W., MacCrimmon, D., and Bassett, A.S. (2006). Schizophrenia in an adult with 6p25 deletion syndrome. Am J Med Genet A 140, 1208-1213.

Canitano, R. (2013). Novel treatments in autism spectrum disorders: from synaptic dysfunction to experimental therapeutics. Behav Brain Res 251, 125-132.

Cannon, T.D., Kaprio, J., Lonnqvist, J., Huttunen, M., and Koskenvuo, M. (1998). The genetic epidemiology of schizophrenia in a Finnish twin cohort. A population-based modeling study. Arch Gen Psychiatry 55, 67-74.

Carbone, P.S. (2013). Moving from research to practice in the primary care of children with autism spectrum disorders. Acad Pediatr 13, 390-399.

Cardno, A.G., Marshall, E.J., Coid, B., Macdonald, A.M., Ribchester, T.R., Davies, N.J., Venturi, P., Jones, L.A., Lewis, S.W., Sham, P.C., et al. (1999). Heritability estimates for psychotic disorders: the Maudsley twin psychosis series. Arch Gen Psychiatry 56, 162-168.

Carroll, L.S., and Owen, M.J. (2009). Genetic overlap between autism, schizophrenia and bipolar disorder. Genome Med 1, 102.

Page 169: Copy Number Variation in Neurodevelopmental Disorders

156

Carter, M.T., Nikkel, S.M., Fernandez, B.A., Marshall, C.R., Noor, A., Lionel, A.C., Prasad, A., Pinto, D., Joseph-George, A.M., Noakes, C., et al. (2011). Hemizygous deletions on chromosome 1p21.3 involving the DPYD gene in individuals with autism spectrum disorder. Clin Genet 80, 435-443.

Celestino-Soper, P.B., Shaw, C.A., Sanders, S.J., Li, J., Murtha, M.T., Ercan-Sencicek, A.G., Davis, L., Thomson, S., Gambin, T., Chinault, A.C., et al. (2011). Use of array CGH to detect exonic copy number variants throughout the genome in autism families detects a novel deletion in TMLHE. Hum Mol Genet 20, 4360-4370.

Chang, Y.T., Chen, P.C., Tsai, I.J., Sung, F.C., Chin, Z.N., Kuo, H.T., Tsai, C.H., and Chou, I.C. (2011). Bidirectional relation between schizophrenia and epilepsy: a population-based retrospective cohort study. Epilepsia 52, 2036-2042.

Charach, A., Chen, S., Hogg-Johnson, S., and Schachar, R.J. (2009). Using the Conners' Teacher Rating Scale-Revised in school children referred for assessment. Can J Psychiatry 54, 232-241.

Chen, X., Shen, Y., Zhang, F., Chiang, C., Pillalamarri, V., Blumenthal, I., Talkowski, M., Wu, B.L., and Gusella, J.F. (2013). Molecular analysis of a deletion hotspot in the NRXN1 region reveals the involvement of short inverted repeats in deletion CNVs. Am J Hum Genet 92, 375-386.

Chess, A. (2012). Mechanisms and consequences of widespread random monoallelic expression. Nat Rev Genet 13, 421-428.

Chiang, A.P., Beck, J.S., Yen, H.J., Tayeh, M.K., Scheetz, T.E., Swiderski, R.E., Nishimura, D.Y., Braun, T.A., Kim, K.Y., Huang, J., et al. (2006). Homozygosity mapping with SNP arrays identifies TRIM32, an E3 ubiquitin ligase, as a Bardet-Biedl syndrome gene (BBS11). Proc Natl Acad Sci USA 103, 6287-6292.

Ching, M.S., Shen, Y., Tan, W.H., Jeste, S.S., Morrow, E.M., Chen, X., Mukaddes, N.M., Yoo, S.Y., Hanson, E., Hundley, R., et al. (2010). Deletions of NRXN1 (neurexin-1) predispose to a wide spectrum of developmental disorders. Am J Med Genet B Neuropsychiatr Genet 153B, 937-947.

Christian, S.L., Brune, C.W., Sudi, J., Kumar, R.A., Liu, S., Karamohamed, S., Badner, J.A., Matsui, S., Conroy, J., McQuaid, D., et al. (2008). Novel submicroscopic chromosomal abnormalities detected in autism spectrum disorder. Biol Psychiatry 63, 1111-1117.

Chung, B.H., Mullegama, S., Marshall, C.R., Lionel, A.C., Weksberg, R., Dupuis, L., Brick, L., Li, C., Scherer, S.W., Aradhya, S., et al. (2012). Severe intellectual disability and autistic features associated with microduplication 2q23.1. Eur J Hum Genet 20, 398-403.

Clarke, A.R., Barry, R.J., Irving, A.M., McCarthy, R., and Selikowitz, M. (2011). Children with attention-deficit/hyperactivity disorder and autistic features: EEG evidence for comorbid disorders. Psychiatry Res 185, 225-231.

Clemmensen, L., Vernal, D.L., and Steinhausen, H.C. (2012). A systematic review of the long-term outcome of early onset schizophrenia. BMC Psychiatry 12, 150.

Page 170: Copy Number Variation in Neurodevelopmental Disorders

157

Coghlan, S., Horder, J., Inkster, B., Mendez, M.A., Murphy, D.G., and Nutt, D.J. (2012). GABA system dysfunction in autism and related disorders: From synapse to symptoms. Neurosci Biobehav Rev 36, 2044-2055.

Collins, P.Y., Patel, V., Joestl, S.S., March, D., Insel, T.R., Daar, A.S., Anderson, W., Dhansay, M.A., Phillips, A., Shurin, S., et al. (2011). Grand challenges in global mental health. Nature 475, 27-30.

Constantino, J.N., Todorov, A., Hilton, C., Law, P., Zhang, Y., Molloy, E., Fitzgerald, R., and Geschwind, D. (2013). Autism recurrence in half siblings: strong support for genetic mechanisms of transmission in ASD. Mol Psychiatry 18, 137-138.

Constantino, J.N., Zhang, Y., Frazier, T., Abbacchi, A.M., and Law, P. (2010). Sibling recurrence and the genetic epidemiology of autism. Am J Psychiatry 167, 1349-1356.

Cook, E.H., Jr., and Scherer, S.W. (2008). Copy-number variations associated with neuropsychiatric conditions. Nature 455, 919-923.

Cooper, G.M., Coe, B.P., Girirajan, S., Rosenfeld, J.A., Vu, T.H., Baker, C., Williams, C., Stalker, H., Hamid, R., Hannig, V., et al. (2011). A copy number variation morbidity map of developmental delay. Nat Genet 43, 838-846.

Cooper, M., Martin, J., Langley, K., Hamshere, M., and Thapar, A. (2014). Autistic traits in children with ADHD index clinical and cognitive problems. Eur Child Adolesc Psychiatry 23, 23-34.

Costain, G., and Bassett, A.S. (2012). Clinical applications of schizophrenia genetics: genetic diagnosis, risk, and counseling in the molecular era. Appl Clin Genet 5, 1-18.

Costain, G., and Bassett, A.S. (2013). Incomplete knowledge of the clinical context as a barrier to interpreting incidental genetic research findings. Am J Bioeth 13, 58-60.

Costain, G., and Bassett, A.S. (2014). Individualizing recurrence risks for severe mental illness: epidemiologic and molecular genetic approaches. Schizophr Bull 40, 21-23.

Costain, G., Chow, E.W., Ray, P.N., and Bassett, A.S. (2012). Caregiver and adult patient perspectives on the importance of a diagnosis of 22q11.2 deletion syndrome. J Intellect Disabil Res 56, 641-651.

Costain, G., Chow, E.W., Silversides, C.K., and Bassett, A.S. (2011). Sex differences in reproductive fitness contribute to preferential maternal transmission of 22q11.2 deletions. J Med Genet 48, 819-824.

Costain, G., Esplen, M.J., Toner, B., Hodgkinson, K.A., and Bassett, A.S. (2014a). Evaluating genetic counseling for family members of individuals with schizophrenia in the molecular age. Schizophr Bull 40, 88-99.

Page 171: Copy Number Variation in Neurodevelopmental Disorders

158

Costain, G., Esplen, M.J., Toner, B., Scherer, S.W., Meschino, W.S., Hodgkinson, K.A., and Bassett, A.S. (2014b). Evaluating genetic counseling for individuals with schizophrenia in the molecular age. Schizophr Bull 40, 78-87.

Costain, G., Lionel, A.C., Merico, D., Forsythe, P., Russell, K., Lowther, C., Yuen, T., Husted, J., Stavropoulos, D.J., Speevak, M., et al. (2013). Pathogenic rare copy number variants in community-based schizophrenia suggest a potential role for clinical microarrays. Hum Mol Genet 22, 4485-4501.

Cotterchio, M., McKeown-Eyssen, G., Sutherland, H., Buchan, G., Aronson, M., Easson, A.M., Macey, J., Holowaty, E., and Gallinger, S. (2000). Ontario familial colon cancer registry: methods and first-year response rates. Chronic Dis Can 21, 81-86.

Crespi, B., Stead, P., and Elliot, M. (2010). Evolution in health and medicine Sackler colloquium: Comparative genomics of autism and schizophrenia. Proc Natl Acad Sci USA 107 Suppl 1, 1736-1741.

Cristino, A.S., Williams, S.M., Hawi, Z., An, J.Y., Bellgrove, M.A., Schwartz, C.E., Costa, L.D., and Claudianos, C. (2013). Neurodevelopmental and neuropsychiatric disorders represent an interconnected molecular system. Mol Psychiatry.

Cross-Disorder Group of the Psychiatric Genomics Consortium (2013). Identification of risk loci with shared effects on five major psychiatric disorders: a genome-wide analysis. Lancet 381, 1371-1379.

Cubells, J.F., Deoreo, E.H., Harvey, P.D., Garlow, S.J., Garber, K., Adam, M.P., and Martin, C.L. (2011). Pharmaco-genetically guided treatment of recurrent rage outbursts in an adult male with 15q13.3 deletion syndrome. Am J Med Genet A 155A, 805-810.

Cuffe, S.P., Moore, C.G., and McKeown, R.E. (2005). Prevalence and correlates of ADHD symptoms in the national health interview survey. J Atten Disord 9, 392-401.

Curran, S., Bolton, P., Rozsnyai, K., Chiocchetti, A., Klauck, S.M., Duketis, E., Poustka, F., Schlitt, S., Freitag, C.M., Lee, I., et al. (2011). No association between a common single nucleotide polymorphism, rs4141463, in the MACROD2 gene and autism spectrum disorder. Am J Med Genet B Neuropsychiatr Genet 156B, 633-639.

Dalsgaard, S. (2013). Attention-deficit/hyperactivity disorder (ADHD). Eur Child Adolesc Psychiatry 22 Suppl 1, S43-48.

Dalsgaard, S., Mortensen, P.B., Frydenberg, M., Maibing, C.M., Nordentoft, M., and Thomsen, P.H. (2013a). Association between Attention-Deficit Hyperactivity Disorder in childhood and schizophrenia later in adulthood. Eur Psychiatry.

Dalsgaard, S., Nielsen, H.S., and Simonsen, M. (2013b). Five-fold increase in national prevalence rates of attention-deficit/hyperactivity disorder medications for children and adolescents with autism spectrum disorder, attention-deficit/hyperactivity disorder, and other psychiatric disorders: a Danish register-based study. J Child Adolesc Psychopharmacol 23, 432-439.

Page 172: Copy Number Variation in Neurodevelopmental Disorders

159

Daniels, J.L., Forssen, U., Hultman, C.M., Cnattingius, S., Savitz, D.A., Feychting, M., and Sparen, P. (2008). Parental psychiatric disorders associated with autism spectrum disorders in the offspring. Pediatrics 121, e1357-1362.

Dawson, G., Jones, E.J., Merkle, K., Venema, K., Lowy, R., Faja, S., Kamara, D., Murias, M., Greenson, J., Winter, J., et al. (2012). Early behavioral intervention is associated with normalized brain activity in young children with autism. J Am Acad Child Adolesc Psychiatry 51, 1150-1159.

Dawson, G., Rogers, S., Munson, J., Smith, M., Winter, J., Greenson, J., Donaldson, A., and Varley, J. (2010). Randomized, controlled trial of an intervention for toddlers with autism: the Early Start Denver Model. Pediatrics 125, e17-23.

de Bruin, E.I., Ferdinand, R.F., Meester, S., de Nijs, P.F., and Verheij, F. (2007). High rates of psychiatric co-morbidity in PDD-NOS. J Autism Dev Disord 37, 877-886.

de la Serna, E., Baeza, I., Andres, S., Puig, O., Sanchez-Guistau, V., Romero, S., Bernardo, M., Moreno, D., Noguera, A., and Castro-Fornieles, J. (2011). Comparison between young siblings and offspring of subjects with schizophrenia: clinical and neuropsychological characteristics. Schizophr Res 131, 35-42.

de la Serna, E., Baeza, I., Toro, J., Andres, S., Puig, O., Sanchez-Guistau, V., Romero, S., Bernardo, M., and Castro-Fornieles, J. (2010). Relationship between clinical and neuropsychological characteristics in child and adolescent first degree relatives of subjects with schizophrenia. Schizophr Res 116, 159-167.

de Lacy, N., and King, B.H. (2013). Revisiting the relationship between autism and schizophrenia: toward an integrated neurobiology. Annu Rev Clin Psychol 9, 555-587.

De Wolf, V., Brison, N., Devriendt, K., and Peeters, H. (2013). Genetic counseling for susceptibility loci and neurodevelopmental disorders: the del15q11.2 as an example. Am J Med Genet A 161A, 2846-2854.

Delorme, R., Ey, E., Toro, R., Leboyer, M., Gillberg, C., and Bourgeron, T. (2013). Progress toward treatments for synaptic defects in autism. Nat Med 19, 685-694.

Devlin, B., Melhem, N., and Roeder, K. (2011). Do common variants play a role in risk for autism? Evidence and theoretical musings. Brain Res 1380, 78-84.

Devlin, B., and Scherer, S.W. (2012). Genetic architecture in autism spectrum disorder. Curr Opin Genet Dev 22, 229-237.

Di Martino, A., Zuo, X.N., Kelly, C., Grzadzinski, R., Mennes, M., Schvarcz, A., Rodman, J., Lord, C., Castellanos, F.X., and Milham, M.P. (2013). Shared and distinct intrinsic functional network centrality in autism and attention-deficit/hyperactivity disorder. Biol Psychiatry 74, 623-632.

Dolcetti, A., Silversides, C.K., Marshall, C.R., Lionel, A.C., Stavropoulos, D.J., Scherer, S.W., and Bassett, A.S. (2013). 1q21.1 Microduplication expression in adults. Genet Med 15, 282-289.

Page 173: Copy Number Variation in Neurodevelopmental Disorders

160

Donev, R., Gantert, D., Alawam, K., Edworthy, A., Hassler, F., Meyer-Lindenberg, A., Dressing, H., and Thome, J. (2011). Comorbidity of schizophrenia and adult attention-deficit hyperactivity disorder. World J Biol Psychiatry 12 Suppl 1, 52-56.

Duong, L., Klitten, L.L., Moller, R.S., Ingason, A., Jakobsen, K.D., Skjodt, C., Didriksen, M., Hjalgrim, H., Werge, T., and Tommerup, N. (2012). Mutations in NRXN1 in a family multiply affected with brain disorders: NRXN1 mutations and brain disorders. Am J Med Genet B Neuropsychiatr Genet 159B, 354-358.

Durand, C.M., Betancur, C., Boeckers, T.M., Bockmann, J., Chaste, P., Fauchereau, F., Nygren, G., Rastam, M., Gillberg, I.C., Anckarsater, H., et al. (2007). Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat Genet 39, 25-27.

Eack, S.M., Bahorik, A.L., McKnight, S.A., Hogarty, S.S., Greenwald, D.P., Newhill, C.E., Phillips, M.L., Keshavan, M.S., and Minshew, N.J. (2013). Commonalities in social and non-social cognitive impairments in adults with autism spectrum disorder and schizophrenia. Schizophr Res 148, 24-28.

El Zein, F., Solis, M., Vaughn, S., and McCulley, L. (2013). Reading Comprehension Interventions for Students with Autism Spectrum Disorders: A Synthesis of Research. J Autism Dev Disord.

Eldevik, S., Hastings, R.P., Hughes, J.C., Jahr, E., Eikeseth, S., and Cross, S. (2009). Meta-analysis of Early Intensive Behavioral Intervention for children with autism. J Clin Child Adolesc Psychol 38, 439-450.

Elia, J., Gai, X., Hakonarson, H., and White, P.S. (2011). Structural variations in attention-deficit hyperactivity disorder. Lancet 377, 377-378; author reply 378.

Elia, J., Gai, X., Xie, H.M., Perin, J.C., Geiger, E., Glessner, J.T., D'Arcy, M., deBerardinis, R., Frackelton, E., Kim, C., et al. (2010). Rare structural variants found in attention-deficit hyperactivity disorder are preferentially associated with neurodevelopmental genes. Mol Psychiatry 15, 637-646.

Elia, J., Glessner, J.T., Wang, K., Takahashi, N., Shtir, C.J., Hadley, D., Sleiman, P.M., Zhang, H., Kim, C.E., Robison, R., et al. (2012a). Genome-wide copy number variation study associates metabotropic glutamate receptor gene networks with attention deficit hyperactivity disorder. Nat Genet 44, 78-84.

Elia, J., Sackett, J., Turner, T., Schardt, M., Tang, S.C., Kurtz, N., Dunfey, M., McFarlane, N.A., Susi, A., Danish, D., et al. (2012b). Attention-deficit/hyperactivity disorder genomics: update for clinicians. Curr Psychiatry Rep 14, 579-589.

Emsley, R., Chiliza, B., and Schoeman, R. (2008). Predictors of long-term outcome in schizophrenia. Curr Opin Psychiatry 21, 173-177.

Ernst, C., Marshall, C.R., Shen, Y., Metcalfe, K., Rosenfeld, J., Hodge, J.C., Torres, A., Blumenthal, I., Chiang, C., Pillalamarri, V., et al. (2012). Highly penetrant alterations of a

Page 174: Copy Number Variation in Neurodevelopmental Disorders

161

critical region including BDNF in human psychopathology and obesity. Arch Gen Psychiatry 69, 1238-1246.

Essrich, C., Lorez, M., Benson, J.A., Fritschy, J.M., and Luscher, B. (1998). Postsynaptic clustering of major GABAA receptor subtypes requires the gamma 2 subunit and gephyrin. Nat Neurosci 1, 563-571.

Esterberg, M.L., Trotman, H.D., Brasfield, J.L., Compton, M.T., and Walker, E.F. (2008). Childhood and current autistic features in adolescents with schizotypal personality disorder. Schizophr Res 104, 265-273.

Etherton, M.R., Blaiss, C.A., Powell, C.M., and Sudhof, T.C. (2009). Mouse neurexin-1alpha deletion causes correlated electrophysiological and behavioral changes consistent with cognitive impairments. Proc Natl Acad Sci USA 106, 17998-18003.

Fang, M., Shen, L., Yin, H., Pan, Y.M., Wang, L., Chen, D., Xi, Z.Q., Xiao, Z., Wang, X.F., and Zhou, S.N. (2011). Downregulation of gephyrin in temporal lobe epilepsy neurons in humans and a rat model. Synapse 65, 1006-1014.

Faraone, S.V., Biederman, J., and Monuteaux, M.C. (2000). Toward guidelines for pedigree selection in genetic studies of attention deficit hyperactivity disorder. Genet Epidemiol 18, 1-16.

Fatemi, S.H., Aldinger, K.A., Ashwood, P., Bauman, M.L., Blaha, C.D., Blatt, G.J., Chauhan, A., Chauhan, V., Dager, S.R., Dickson, P.E., et al. (2012). Consensus paper: pathological role of the cerebellum in autism. Cerebellum 11, 777-807.

Feifel, D., Macdonald, K., Nguyen, A., Cobb, P., Warlan, H., Galangue, B., Minassian, A., Becker, O., Cooper, J., Perry, W., et al. (2010). Adjunctive intranasal oxytocin reduces symptoms in schizophrenia patients. Biol Psychiatry 68, 678-680.

Feng, G., Tintrup, H., Kirsch, J., Nichol, M.C., Kuhse, J., Betz, H., and Sanes, J.R. (1998). Dual requirement for gephyrin in glycine receptor clustering and molybdoenzyme activity. Science 282, 1321-1324.

Fernandez, B.A., Roberts, W., Chung, B., Weksberg, R., Meyn, S., Szatmari, P., Joseph-George, A.M., Mackay, S., Whitten, K., Noble, B., et al. (2010). Phenotypic spectrum associated with de novo and inherited deletions and duplications at 16p11.2 in individuals ascertained for diagnosis of autism spectrum disorder. J Med Genet 47, 195-203.

Fernandez, T.V., Sanders, S.J., Yurkiewicz, I.R., Ercan-Sencicek, A.G., Kim, Y.S., Fishman, D.O., Raubeson, M.J., Song, Y., Yasuno, K., Ho, W.S., et al. (2012). Rare copy number variants in tourette syndrome disrupt genes in histaminergic pathways and overlap with autism. Biol Psychiatry 71, 392-402.

Feuk, L., Kalervo, A., Lipsanen-Nyman, M., Skaug, J., Nakabayashi, K., Finucane, B., Hartung, D., Innes, M., Kerem, B., Nowaczyk, M.J., et al. (2006). Absence of a paternally inherited FOXP2 gene in developmental verbal dyspraxia. Am J Hum Genet 79, 965-972.

Page 175: Copy Number Variation in Neurodevelopmental Disorders

162

Filges, I., Rothlisberger, B., Blattner, A., Boesch, N., Demougin, P., Wenzel, F., Huber, A.R., Heinimann, K., Weber, P., and Miny, P. (2011). Deletion in Xp22.11: PTCHD1 is a candidate gene for X-linked intellectual disability with or without autism. Clin Genet 79, 79-85.

Fischer, F., Kneussel, M., Tintrup, H., Haverkamp, S., Rauen, T., Betz, H., and Wassle, H. (2000). Reduced synaptic clustering of GABA and glycine receptors in the retina of the gephyrin null mutant mouse. J Comp Neurol 427, 634-648.

Fishell, G., and Hatten, M.E. (1991). Astrotactin provides a receptor system for CNS neuronal migration. Development 113, 755-765.

Fombonne, E. (2009). Epidemiology of pervasive developmental disorders. Pediatr Res 65, 591-598.

Forstera, B., Belaidi, A.A., Juttner, R., Bernert, C., Tsokos, M., Lehmann, T.N., Horn, P., Dehnicke, C., Schwarz, G., and Meier, J.C. (2010). Irregular RNA splicing curtails postsynaptic gephyrin in the cornu ammonis of patients with epilepsy. Brain 133, 3778-3794.

Fournier, K.A., Hass, C.J., Naik, S.K., Lodha, N., and Cauraugh, J.H. (2010). Motor coordination in autism spectrum disorders: a synthesis and meta-analysis. J Autism Dev Disord 40, 1227-1240.

Franke, B., Faraone, S.V., Asherson, P., Buitelaar, J., Bau, C.H., Ramos-Quiroga, J.A., Mick, E., Grevet, E.H., Johansson, S., Haavik, J., et al. (2012). The genetics of attention deficit/hyperactivity disorder in adults, a review. Mol Psychiatry 17, 960-987.

Franzek, E., and Beckmann, H. (1998). Different genetic background of schizophrenia spectrum psychoses: a twin study. Am J Psychiatry 155, 76-83.

Friedman, J.M., Baross, A., Delaney, A.D., Ally, A., Arbour, L., Armstrong, L., Asano, J., Bailey, D.K., Barber, S., Birch, P., et al. (2006). Oligonucleotide microarray analysis of genomic imbalance in children with mental retardation. Am J Hum Genet 79, 500-513.

Fritschy, J.M., Harvey, R.J., and Schwarz, G. (2008). Gephyrin: where do we stand, where do we go? Trends Neurosci 31, 257-264.

Fung, W.L., Chow, E.W., Webb, G.D., Gatzoulis, M.A., and Bassett, A.S. (2008). Extracardiac features predicting 22q11.2 deletion syndrome in adult congenital heart disease. Int J Cardiol 131, 51-58.

Gai, X., Xie, H.M., Perin, J.C., Takahashi, N., Murphy, K., Wenocur, A.S., D'Arcy, M., O'Hara, R.J., Goldmuntz, E., Grice, D.E., et al. (2012). Rare structural variation of synapse and neurotransmission genes in autism. Mol Psychiatry 17, 402-411.

Gargaro, B.A., Rinehart, N.J., Bradshaw, J.L., Tonge, B.J., and Sheppard, D.M. (2011). Autism and ADHD: how far have we come in the comorbidity debate? Neurosci Biobehav Rev 35, 1081-1088.

Page 176: Copy Number Variation in Neurodevelopmental Disorders

163

Gauthier, J., Champagne, N., Lafreniere, R.G., Xiong, L., Spiegelman, D., Brustein, E., Lapointe, M., Peng, H., Cote, M., Noreau, A., et al. (2010). De novo mutations in the gene encoding the synaptic scaffolding protein SHANK3 in patients ascertained for schizophrenia. Proc Natl Acad Sci USA 107, 7863-7868.

Gauthier, J., Siddiqui, T., Huashan, P., Yokomaku, D., Hamdan, F., Champagne, N., Lapointe, M., Spiegelman, D., Noreau, A., Lafrenière, R., et al. (2011). Truncating mutations in NRXN2 and NRXN1 in autism spectrum disorders and schizophrenia. Hum Genet 130, 563-573.

Geissler, J., and Lesch, K.P. (2011). A lifetime of attention-deficit/hyperactivity disorder: diagnostic challenges, treatment and neurobiological mechanisms. Expert Rev Neurother 11, 1467-1484.

Gejman, P.V., Sanders, A.R., and Kendler, K.S. (2011). Genetics of schizophrenia: new findings and challenges. Annu Rev Genomics Hum Genet 12, 121-144.

Ghosh, A., Michalon, A., Lindemann, L., Fontoura, P., and Santarelli, L. (2013). Drug discovery for autism spectrum disorder: challenges and opportunities. Nat Rev Drug Discov 12, 777-790.

Gillberg, C., Cederlund, M., Lamberg, K., and Zeijlon, L. (2006). Brief report: "the autism epidemic". The registered prevalence of autism in a Swedish urban area. J Autism Dev Disord 36, 429-435.

Gilman, S.R., Chang, J., Xu, B., Bawa, T.S., Gogos, J.A., Karayiorgou, M., and Vitkup, D. (2012). Diverse types of genetic variation converge on functional gene networks involved in schizophrenia. Nat Neurosci 15, 1723-1728.

Gilman, S.R., Iossifov, I., Levy, D., Ronemus, M., Wigler, M., and Vitkup, D. (2011). Rare de novo variants associated with autism implicate a large functional network of genes involved in formation and function of synapses. Neuron 70, 898-907.

Girard, S.L., Dion, P.A., and Rouleau, G.A. (2012). Schizophrenia genetics: putting all the pieces together. Curr Neurol Neurosci Rep 12, 261-266.

Girard, S.L., Gauthier, J., Noreau, A., Xiong, L., Zhou, S., Jouan, L., Dionne-Laporte, A., Spiegelman, D., Henrion, E., Diallo, O., et al. (2011). Increased exonic de novo mutation rate in individuals with schizophrenia. Nat Genet 43, 860-863.

Girirajan, S., Campbell, C.D., and Eichler, E.E. (2011). Human copy number variation and complex genetic disease. Annu Rev Genet 45, 203-226.

Gizer, I.R., Ficks, C., and Waldman, I.D. (2009). Candidate gene studies of ADHD: a meta-analytic review. Hum Genet 126, 51-90.

Glessner, J.T., Reilly, M.P., Kim, C.E., Takahashi, N., Albano, A., Hou, C., Bradfield, J.P., Zhang, H., Sleiman, P.M., Flory, J.H., et al. (2010). Strong synaptic transmission impact by copy number variations in schizophrenia. Proc Natl Acad Sci USA 107, 10584-10589.

Page 177: Copy Number Variation in Neurodevelopmental Disorders

164

Glessner, J.T., Wang, K., Cai, G., Korvatska, O., Kim, C.E., Wood, S., Zhang, H., Estes, A., Brune, C.W., Bradfield, J.P., et al. (2009). Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 459, 569-573.

Glynn, D., Bortnick, R.A., and Morton, A.J. (2003). Complexin II is essential for normal neurological function in mice. Hum Mol Genet 12, 2431-2448.

Gomez, R.L., Janowsky, D., Zetin, M., Huey, L., and Clopton, P.L. (1981). Adult psychiatric diagnosis and symptoms compatible with the hyperactive child syndrome: a retrospective study. J Clin Psychiatry 42, 389-394.

Goobie, S., Knijnenburg, J., Fitzpatrick, D., Sharkey, F.H., Lionel, A.C., Marshall, C.R., Azam, T., Shago, M., Chong, K., Mendoza-Londono, R., et al. (2008). Molecular and clinical characterization of de novo and familial cases with microduplication 3q29: guidelines for copy number variation case reporting. Cytogenet Genome Res 123, 65-78.

Grant, S.G. (2012). Synaptopathies: diseases of the synaptome. Curr Opin Neurobiol 22, 522-529.

Grozeva, D., Kirov, G., Ivanov, D., Jones, I.R., Jones, L., Green, E.K., St Clair, D.M., Young, A.H., Ferrier, N., Farmer, A.E., et al. (2010). Rare copy number variants: a point of rarity in genetic risk for bipolar disorder and schizophrenia. Arch Gen Psychiatry 67, 318-327.

Grzadzinski, R., Di Martino, A., Brady, E., Mairena, M.A., O'Neale, M., Petkova, E., Lord, C., and Castellanos, F.X. (2011). Examining autistic traits in children with ADHD: does the autism spectrum extend to ADHD? J Autism Dev Disord 41, 1178-1191.

Guilmatre, A., Dubourg, C., Mosca, A.L., Legallic, S., Goldenberg, A., Drouin-Garraud, V., Layet, V., Rosier, A., Briault, S., Bonnet-Brilhault, F., et al. (2009). Recurrent rearrangements in synaptic and neurodevelopmental genes and shared biologic pathways in schizophrenia, autism, and mental retardation. Arch Gen Psychiatry 66, 947-956.

Hallmayer, J., Cleveland, S., Torres, A., Phillips, J., Cohen, B., Torigoe, T., Miller, J., Fedele, A., Collins, J., Smith, K., et al. (2011). Genetic heritability and shared environmental factors among twin pairs with autism. Arch Gen Psychiatry 68, 1095-1102.

Hampson, D.R., Gholizadeh, S., and Pacey, L.K. (2012). Pathways to drug development for autism spectrum disorders. Clin Pharmacol Ther 91, 189-200.

Hamshere, M.L., Stergiakouli, E., Langley, K., Martin, J., Holmans, P., Kent, L., Owen, M.J., Gill, M., Thapar, A., O'Donovan, M., et al. (2013). Shared polygenic contribution between childhood attention-deficit hyperactivity disorder and adult schizophrenia. Br J Psychiatry 203, 107-111.

Han, K., Holder, J.L., Jr., Schaaf, C.P., Lu, H., Chen, H., Kang, H., Tang, J., Wu, Z., Hao, S., Cheung, S.W., et al. (2013). SHANK3 overexpression causes manic-like behaviour with unique pharmacogenetic properties. Nature 503, 72-77.

Page 178: Copy Number Variation in Neurodevelopmental Disorders

165

Handrigan, G.R., Chitayat, D., Lionel, A.C., Pinsk, M., Vaags, A.K., Marshall, C.R., Dyack, S., Escobar, L.F., Fernandez, B.A., Stegman, J.C., et al. (2013). Deletions in 16q24.2 are associated with autism spectrum disorder, intellectual disability and congenital renal malformation. J Med Genet 50, 163-173.

Heck, D.H., and Lu, L. (2012). The social life of neurons: synaptic communication deficits as a common denominator of autism, schizophrenia, and other cognitive disorders. Biol Psychiatry 72, 173-174.

Hehir-Kwa, J.Y., Rodriguez-Santiago, B., Vissers, L.E., de Leeuw, N., Pfundt, R., Buitelaar, J.K., Perez-Jurado, L.A., and Veltman, J.A. (2011). De novo copy number variants associated with intellectual disability have a paternal origin and age bias. J Med Genet 48, 776-778.

Hermey, G., Plath, N., Hubner, C.A., Kuhl, D., Schaller, H.C., and Hermans-Borgmeyer, I. (2004). The three sorCS genes are differentially expressed and regulated by synaptic activity. J Neurochem 88, 1470-1476.

Hodge, J.C., Mitchell, E., Pillalamarri, V., Toler, T.L., Bartel, F., Kearney, H.M., Zou, Y.S., Tan, W.H., Hanscom, C., Kirmani, S., et al. (2013). Disruption of MBD5 contributes to a spectrum of psychopathology and neurodevelopmental abnormalities. Mol Psychiatry.

Hofvander, B., Delorme, R., Chaste, P., Nyden, A., Wentz, E., Stahlberg, O., Herbrecht, E., Stopin, A., Anckarsater, H., Gillberg, C., et al. (2009). Psychiatric and psychosocial problems in adults with normal-intelligence autism spectrum disorders. BMC Psychiatry 9, 35.

Hollander, E., Bartz, J., Chaplin, W., Phillips, A., Sumner, J., Soorya, L., Anagnostou, E., and Wasserman, S. (2007). Oxytocin increases retention of social cognition in autism. Biol Psychiatry 61, 498-503.

Holtmann, M., Bolte, S., and Poustka, F. (2007). Attention deficit hyperactivity disorder symptoms in pervasive developmental disorders: association with autistic behavior domains and coexisting psychopathology. Psychopathology 40, 172-177.

Hoon, M., Soykan, T., Falkenburger, B., Hammer, M., Patrizi, A., Schmidt, K.F., Sassoe-Pognetto, M., Lowel, S., Moser, T., Taschenberger, H., et al. (2011). Neuroligin-4 is localized to glycinergic postsynapses and regulates inhibition in the retina. Proc Natl Acad Sci USA 108, 3053-3058.

Huguet, G., Ey, E., and Bourgeron, T. (2013). The genetic landscapes of autism spectrum disorders. Annu Rev Genomics Hum Genet 14, 191-213.

Hultman, C.M., Sandin, S., Levine, S.Z., Lichtenstein, P., and Reichenberg, A. (2011). Advancing paternal age and risk of autism: new evidence from a population-based study and a meta-analysis of epidemiological studies. Mol Psychiatry 16, 1203-1212.

Ickowicz, A., Schachar, R.J., Sugarman, R., Chen, S.X., Millette, C., and Cook, L. (2006). The parent interview for child symptoms: a situation-specific clinical research interview for attention-deficit hyperactivity and related disorders. Can J Psychiatry 51, 325-328.

Page 179: Copy Number Variation in Neurodevelopmental Disorders

166

Ingraham, L.J., and Kety, S.S. (2000). Adoption studies of schizophrenia. Am J Med Genet 97, 18-22.

Insel, T.R. (2010). Rethinking schizophrenia. Nature 468, 187-193.

Insel, T.R. (2014). Brain somatic mutations: the dark matter of psychiatric genetics? Mol Psychiatry 19, 156-158.

International Schizophrenia Consortium (2008). Rare chromosomal deletions and duplications increase risk of schizophrenia. Nature 455, 237-241.

Iossifov, I., Ronemus, M., Levy, D., Wang, Z., Hakker, I., Rosenbaum, J., Yamrom, B., Lee, Y.H., Narzisi, G., Leotta, A., et al. (2012). De novo gene disruptions in children on the autistic spectrum. Neuron 74, 285-299.

Itsara, A., Cooper, G.M., Baker, C., Girirajan, S., Li, J., Absher, D., Krauss, R.M., Myers, R.M., Ridker, P.M., Chasman, D.I., et al. (2009). Population analysis of large copy number variants and hotspots of human genetic disease. Am J Hum Genet 84, 148-161.

Jablensky, A. (2010). The diagnostic concept of schizophrenia: its history, evolution, and future prospects. Dialogues Clin Neurosci 12, 271-287.

Jacob, T.C., Bogdanov, Y.D., Magnus, C., Saliba, R.S., Kittler, J.T., Haydon, P.G., and Moss, S.J. (2005). Gephyrin regulates the cell surface dynamics of synaptic GABAA receptors. J Neurosci 25, 10469-10478.

Jacquemont, M.L., Sanlaville, D., Redon, R., Raoul, O., Cormier-Daire, V., Lyonnet, S., Amiel, J., Le Merrer, M., Heron, D., de Blois, M.C., et al. (2006). Array-based comparative genomic hybridisation identifies high frequency of cryptic chromosomal rearrangements in patients with syndromic autism spectrum disorders. J Med Genet 43, 843-849.

Jacquemont, S., Reymond, A., Zufferey, F., Harewood, L., Walters, R.G., Kutalik, Z., Martinet, D., Shen, Y., Valsesia, A., Beckmann, N.D., et al. (2011). Mirror extreme BMI phenotypes associated with gene dosage at the chromosome 16p11.2 locus. Nature 478, 97-102.

Jacques, P.E., Jeyakani, J., and Bourque, G. (2013). The majority of primate-specific regulatory sequences are derived from transposable elements. PLoS Genet 9, e1003504.

Jamain, S., Quach, H., Betancur, C., Rastam, M., Colineaux, C., Gillberg, I.C., Soderstrom, H., Giros, B., Leboyer, M., Gillberg, C., et al. (2003). Mutations of the X-linked genes encoding neuroligins NLGN3 and NLGN4 are associated with autism. Nat Genet 34, 27-29.

Jarick, I., Volckmar, A.L., Putter, C., Pechlivanis, S., Nguyen, T.T., Dauvermann, M.R., Beck, S., Albayrak, O., Scherag, S., Gilsbach, S., et al. (2014). Genome-wide analysis of rare copy number variations reveals PARK2 as a candidate gene for attention-deficit/hyperactivity disorder. Mol Psychiatry 19, 115-121.

Page 180: Copy Number Variation in Neurodevelopmental Disorders

167

Jensen, P.S., Garcia, J.A., Glied, S., Crowe, M., Foster, M., Schlander, M., Hinshaw, S., Vitiello, B., Arnold, L.E., Elliott, G., et al. (2005). Cost-effectiveness of ADHD treatments: findings from the multimodal treatment study of children with ADHD. Am J Psychiatry 162, 1628-1636.

Jiang, Y.-h., Yuen, Ryan K.C., Jin, X., Wang, M., Chen, N., Wu, X., Ju, J., Mei, J., Shi, Y., He, M., et al. (2013). Detection of Clinically Relevant Genetic Variants in Autism Spectrum Disorder by Whole-Genome Sequencing. Am J Hum Genet 93, 249-263.

Johnson, M.B., Kawasawa, Y.I., Mason, C.E., Krsnik, Z., Coppola, G., Bogdanovic, D., Geschwind, D.H., Mane, S.M., State, M.W., and Sestan, N. (2009). Functional and evolutionary insights into human brain development through global transcriptome analysis. Neuron 62, 494-509.

Jokiranta, E., Brown, A.S., Heinimaa, M., Cheslack-Postava, K., Suominen, A., and Sourander, A. (2013). Parental psychiatric disorders and autism spectrum disorders. Psychiatry Res 207, 203-211.

Joshi, G., Wozniak, J., Petty, C., Martelon, M.K., Fried, R., Bolfek, A., Kotte, A., Stevens, J., Furtak, S.L., Bourgeois, M., et al. (2013). Psychiatric comorbidity and functioning in a clinically referred population of adults with autism spectrum disorders: a comparative study. J Autism Dev Disord 43, 1314-1325.

Kalscheuer, V.M., Musante, L., Fang, C., Hoffmann, K., Fuchs, C., Carta, E., Deas, E., Venkateswarlu, K., Menzel, C., Ullmann, R., et al. (2009). A balanced chromosomal translocation disrupting ARHGEF9 is associated with epilepsy, anxiety, aggression, and mental retardation. Hum Mutat 30, 61-68.

Kaminsky, E.B., Kaul, V., Paschall, J., Church, D.M., Bunke, B., Kunig, D., Moreno-De-Luca, D., Moreno-De-Luca, A., Mulle, J.G., Warren, S.T., et al. (2011). An evidence-based approach to establish the functional and clinical significance of copy number variants in intellectual and developmental disabilities. Genet Med 13, 777-784.

Kan, K.J., Dolan, C.V., Nivard, M.G., Middeldorp, C.M., van Beijsterveldt, C.E., Willemsen, G., and Boomsma, D.I. (2013). Genetic and environmental stability in attention problems across the lifespan: evidence from the Netherlands twin register. J Am Acad Child Adolesc Psychiatry 52, 12-25.

Kang, H.J., Kawasawa, Y.I., Cheng, F., Zhu, Y., Xu, X., Li, M., Sousa, A.M., Pletikos, M., Meyer, K.A., Sedmak, G., et al. (2011). Spatio-temporal transcriptome of the human brain. Nature 478, 483-489.

Kang, Y., Zhang, X., Dobie, F., Wu, H., and Craig, A.M. (2008). Induction of GABAergic postsynaptic differentiation by alpha-neurexins. J Biol Chem 283, 2323-2334.

Kearney, H.M., Thorland, E.C., Brown, K.K., Quintero-Rivera, F., and South, S.T. (2011). American College of Medical Genetics standards and guidelines for interpretation and reporting of postnatal constitutional copy number variants. Genet Med 13, 680-685.

Page 181: Copy Number Variation in Neurodevelopmental Disorders

168

Kebir, O., Tabbane, K., Sengupta, S., and Joober, R. (2009). Candidate genes and neuropsychological phenotypes in children with ADHD: review of association studies. J Psychiatry Neurosci 34, 88-101.

Kerjan, G., Koizumi, H., Han, E.B., Dube, C.M., Djakovic, S.N., Patrick, G.N., Baram, T.Z., Heinemann, S.F., and Gleeson, J.G. (2009). Mice lacking doublecortin and doublecortin-like kinase 2 display altered hippocampal neuronal maturation and spontaneous seizures. Proc Natl Acad Sci USA 106, 6766-6771.

Kessler, R.C., Lane, M., Stang, P.E., and Van Brunt, D.L. (2009). The prevalence and workplace costs of adult attention deficit hyperactivity disorder in a large manufacturing firm. Psychol Med 39, 137-147.

Kety, S.S., Wender, P.H., Jacobsen, B., Ingraham, L.J., Jansson, L., Faber, B., and Kinney, D.K. (1994). Mental illness in the biological and adoptive relatives of schizophrenic adoptees. Replication of the Copenhagen Study in the rest of Denmark. Arch Gen Psychiatry 51, 442-455.

Kim-Cohen, J., Caspi, A., Moffitt, T.E., Harrington, H., Milne, B.J., and Poulton, R. (2003). Prior juvenile diagnoses in adults with mental disorder: developmental follow-back of a prospective-longitudinal cohort. Arch Gen Psychiatry 60, 709-717.

Kim, H.G., Kishikawa, S., Higgins, A.W., Seong, I.S., Donovan, D.J., Shen, Y., Lally, E., Weiss, L.A., Najm, J., Kutsche, K., et al. (2008). Disruption of neurexin 1 associated with autism spectrum disorder. Am J Hum Genet 82, 199-207.

King, B.H., and Lord, C. (2011). Is schizophrenia on the autism spectrum? Brain Res 1380, 34-41.

Kirov, G., Grozeva, D., Norton, N., Ivanov, D., Mantripragada, K.K., Holmans, P., Craddock, N., Owen, M.J., and O'Donovan, M.C. (2009). Support for the involvement of large copy number variants in the pathogenesis of schizophrenia. Hum Mol Genet 18, 1497-1503.

Kirov, G., Gumus, D., Chen, W., Norton, N., Georgieva, L., Sari, M., O'Donovan, M.C., Erdogan, F., Owen, M.J., Ropers, H.H., et al. (2008). Comparative genome hybridization suggests a role for NRXN1 and APBA2 in schizophrenia. Hum Mol Genet 17, 458-465.

Kirov, G., Pocklington, A.J., Holmans, P., Ivanov, D., Ikeda, M., Ruderfer, D., Moran, J., Chambert, K., Toncheva, D., Georgieva, L., et al. (2012). De novo CNV analysis implicates specific abnormalities of postsynaptic signalling complexes in the pathogenesis of schizophrenia. Mol Psychiatry 17, 142-153.

Kirsch, J., Wolters, I., Triller, A., and Betz, H. (1993). Gephyrin antisense oligonucleotides prevent glycine receptor clustering in spinal neurons. Nature 366, 745-748.

Klaning, U., Mortensen, P.B., and Kyvik, K.O. (1996). Increased occurrence of schizophrenia and other psychiatric illnesses among twins. Br J Psychiatry 168, 688-692.

Page 182: Copy Number Variation in Neurodevelopmental Disorders

169

Klei, L., Sanders, S.J., Murtha, M.T., Hus, V., Lowe, J.K., Willsey, A.J., Moreno-De-Luca, D., Yu, T.W., Fombonne, E., Geschwind, D., et al. (2012). Common genetic variants, acting additively, are a major source of risk for autism. Mol Autism 3, 9.

Kneussel, M., Brandstatter, J.H., Laube, B., Stahl, S., Muller, U., and Betz, H. (1999). Loss of postsynaptic GABA(A) receptor clustering in gephyrin-deficient mice. J Neurosci 19, 9289-9297.

Kochhar, P., Batty, M.J., Liddle, E.B., Groom, M.J., Scerif, G., Liddle, P.F., and Hollis, C.P. (2011). Autistic spectrum disorder traits in children with attention deficit hyperactivity disorder. Child Care Health Dev 37, 103-110.

Kong, A., Frigge, M.L., Masson, G., Besenbacher, S., Sulem, P., Magnusson, G., Gudjonsson, S.A., Sigurdsson, A., Jonasdottir, A., Wong, W.S., et al. (2012). Rate of de novo mutations and the importance of father's age to disease risk. Nature 488, 471-475.

Korn, J.M., Kuruvilla, F.G., McCarroll, S.A., Wysoker, A., Nemesh, J., Cawley, S., Hubbell, E., Veitch, J., Collins, P.J., Darvishi, K., et al. (2008). Integrated genotype calling and association analysis of SNPs, common copy number polymorphisms and rare CNVs. Nat Genet 40, 1253-1260.

Kotte, A., Joshi, G., Fried, R., Uchida, M., Spencer, A., Woodworth, K.Y., Kenworthy, T., Faraone, S.V., and Biederman, J. (2013). Autistic traits in children with and without ADHD. Pediatrics 132, e612-622.

Krawczak, M., Nikolaus, S., von Eberstein, H., Croucher, P.J., El Mokhtari, N.E., and Schreiber, S. (2006). PopGen: population-based recruitment of patients and controls for the analysis of complex genotype-phenotype relationships. Community Genet 9, 55-61.

Krumm, N., O'Roak, B.J., Karakoc, E., Mohajeri, K., Nelson, B., Vives, L., Jacquemont, S., Munson, J., Bernier, R., and Eichler, E.E. (2013). Transmission disequilibrium of small CNVs in simplex autism. Am J Hum Genet 93, 595-606.

Krumm, N., O'Roak, B.J., Shendure, J., and Eichler, E.E. (2014). A de novo convergence of autism genetics and molecular neuroscience. Trends Neurosci 37, 95-105.

Ku, C.S., Polychronakos, C., Tan, E.K., Naidoo, N., Pawitan, Y., Roukos, D.H., Mort, M., and Cooper, D.N. (2013). A new paradigm emerges from the study of de novo mutations in the context of neurodevelopmental disease. Mol Psychiatry 18, 141-153.

Kumar, R.A., KaraMohamed, S., Sudi, J., Conrad, D.F., Brune, C., Badner, J.A., Gilliam, T.C., Nowak, N.J., Cook, E.H., Jr., Dobyns, W.B., et al. (2008). Recurrent 16p11.2 microdeletions in autism. Hum Mol Genet 17, 628-638.

Kumar, S.S., and Buckmaster, P.S. (2006). Hyperexcitability, interneurons, and loss of GABAergic synapses in entorhinal cortex in a model of temporal lobe epilepsy. J Neurosci 26, 4613-4623.

Page 183: Copy Number Variation in Neurodevelopmental Disorders

170

Kuntsi, J., Neale, B.M., Chen, W., Faraone, S.V., and Asherson, P. (2006). The IMAGE project: methodological issues for the molecular genetic analysis of ADHD. Behav Brain Funct 2, 27.

Kustanovich, V., Merriman, B., McGough, J., McCracken, J.T., Smalley, S.L., and Nelson, S.F. (2003). Biased paternal transmission of SNAP-25 risk alleles in attention-deficit hyperactivity disorder. Mol Psychiatry 8, 309-315.

Lai, M.C., Lombardo, M.V., Chakrabarti, B., and Baron-Cohen, S. (2013). Subgrouping the autism "spectrum": reflections on DSM-5. PLoS Biol 11, e1001544.

Langley, K., Martin, J., Agha, S.S., Davies, C., Stergiakouli, E., Holmans, P., Williams, N., Owen, M., O'Donovan, M., and Thapar, A. (2011). Clinical and cognitive characteristics of children with attention-deficit hyperactivity disorder, with and without copy number variants. Br J Psychiatry 199, 398-403.

Larsson, H., Asherson, P., Chang, Z., Ljung, T., Friedrichs, B., Larsson, J.O., and Lichtenstein, P. (2013a). Genetic and environmental influences on adult attention deficit hyperactivity disorder symptoms: a large Swedish population-based study of twins. Psychol Med 43, 197-207.

Larsson, H., Dilshad, R., Lichtenstein, P., and Barker, E.D. (2011). Developmental trajectories of DSM-IV symptoms of attention-deficit/hyperactivity disorder: genetic effects, family risk and associated psychopathology. J Child Psychol Psychiatry 52, 954-963.

Larsson, H., Ryden, E., Boman, M., Langstrom, N., Lichtenstein, P., and Landen, M. (2013b). Risk of bipolar disorder and schizophrenia in relatives of people with attention-deficit hyperactivity disorder. Br J Psychiatry 203, 103-106.

Larsson, H.J., Eaton, W.W., Madsen, K.M., Vestergaard, M., Olesen, A.V., Agerbo, E., Schendel, D., Thorsen, P., and Mortensen, P.B. (2005). Risk factors for autism: perinatal factors, parental psychiatric history, and socioeconomic status. Am J Epidemiol 161, 916-925; discussion 926-918.

Lasky-Su, J., Neale, B.M., Franke, B., Anney, R.J., Zhou, K., Maller, J.B., Vasquez, A.A., Chen, W., Asherson, P., Buitelaar, J., et al. (2008). Genome-wide association scan of quantitative traits for attention deficit hyperactivity disorder identifies novel associations and confirms candidate gene associations. Am J Med Genet B Neuropsychiatr Genet 147B, 1345-1354.

Laumonnier, F., Bonnet-Brilhault, F., Gomot, M., Blanc, R., David, A., Moizard, M.P., Raynaud, M., Ronce, N., Lemonnier, E., Calvas, P., et al. (2004). X-linked mental retardation and autism are associated with a mutation in the NLGN4 gene, a member of the neuroligin family. Am J Hum Genet 74, 552-557.

Le Couteur, A., Bailey, A., Goode, S., Pickles, A., Robertson, S., Gottesman, I., and Rutter, M. (1996). A broader phenotype of autism: the clinical spectrum in twins. J Child Psychol Psychiatry 37, 785-801.

Lee, C., and Scherer, S.W. (2010). The clinical context of copy number variation in the human genome. Expert Rev Mol Med 12, e8.

Page 184: Copy Number Variation in Neurodevelopmental Disorders

171

Lencz, T., Lambert, C., DeRosse, P., Burdick, K.E., Morgan, T.V., Kane, J.M., Kucherlapati, R., and Malhotra, A.K. (2007). Runs of homozygosity reveal highly penetrant recessive loci in schizophrenia. Proc Natl Acad Sci USA 104, 19942-19947.

Lesca, G., Rudolf, G., Labalme, A., Hirsch, E., Arzimanoglou, A., Genton, P., Motte, J., de Saint Martin, A., Valenti, M.P., Boulay, C., et al. (2012). Epileptic encephalopathies of the Landau-Kleffner and continuous spike and waves during slow-wave sleep types: Genomic dissection makes the link with autism. Epilepsia 53, 1526-1538.

Lesca, G., Till, M., Labalme, A., Vallee, D., Hugonenq, C., Philip, N., Edery, P., and Sanlaville, D. (2011). De novo Xq11.11 microdeletion including ARHGEF9 in a boy with mental retardation, epilepsy, macrosomia, and dysmorphic features. Am J Med Genet A 155A, 1706-1711.

Lesch, K.P., Selch, S., Renner, T.J., Jacob, C., Nguyen, T.T., Hahn, T., Romanos, M., Walitza, S., Shoichet, S., Dempfle, A., et al. (2011). Genome-wide copy number variation analysis in attention-deficit/hyperactivity disorder: association with neuropeptide Y gene dosage in an extended pedigree. Mol Psychiatry 16, 491-503.

Lesch, K.P., Timmesfeld, N., Renner, T.J., Halperin, R., Roser, C., Nguyen, T.T., Craig, D.W., Romanos, J., Heine, M., Meyer, J., et al. (2008). Molecular genetics of adult ADHD: converging evidence from genome-wide association and extended pedigree linkage studies. J Neural Transm 115, 1573-1585.

Levinson, D.F., Duan, J., Oh, S., Wang, K., Sanders, A.R., Shi, J., Zhang, N., Mowry, B.J., Olincy, A., Amin, F., et al. (2011). Copy number variants in schizophrenia: confirmation of five previous findings and new evidence for 3q29 microdeletions and VIPR2 duplications. Am J Psychiatry 168, 302-316.

Levy, D., Ronemus, M., Yamrom, B., Lee, Y.H., Leotta, A., Kendall, J., Marks, S., Lakshmi, B., Pai, D., Ye, K., et al. (2011). Rare de novo and transmitted copy-number variation in autistic spectrum disorders. Neuron 70, 886-897.

Leyfer, O.T., Folstein, S.E., Bacalman, S., Davis, N.O., Dinh, E., Morgan, J., Tager-Flusberg, H., and Lainhart, J.E. (2006). Comorbid psychiatric disorders in children with autism: interview development and rates of disorders. J Autism Dev Disord 36, 849-861.

Lichtenstein, P., Carlstrom, E., Rastam, M., Gillberg, C., and Anckarsater, H. (2010). The genetics of autism spectrum disorders and related neuropsychiatric disorders in childhood. Am J Psychiatry 167, 1357-1363.

Lionel, A.C., Crosbie, J., Barbosa, N., Goodale, T., Thiruvahindrapuram, B., Rickaby, J., Gazzellone, M., Carson, A.R., Howe, J.L., Wang, Z., et al. (2011). Rare copy number variation discovery and cross-disorder comparisons identify risk genes for ADHD. Sci Transl Med 3, 95ra75.

Lionel, A.C., Tammimies, K., Vaags, A.K., Rosenfeld, J.A., Ahn, J.W., Merico, D., Noor, A., Runke, C.K., Pillalamarri, V.K., Carter, M.T., et al. (2014). Disruption of the ASTN2/TRIM32

Page 185: Copy Number Variation in Neurodevelopmental Disorders

172

locus at 9q33.1 is a risk factor in males for autism spectrum disorders, ADHD and other neurodevelopmental phenotypes. Hum Mol Genet.

Lionel, A.C., Vaags, A.K., Sato, D., Gazzellone, M.J., Mitchell, E.B., Chen, H.Y., Costain, G., Walker, S., Egger, G., Thiruvahindrapuram, B., et al. (2013). Rare exonic deletions implicate the synaptic organizer Gephyrin (GPHN) in risk for autism, schizophrenia and seizures. Hum Mol Genet 22, 2055-2066.

Liu, P., Carvalho, C.M., Hastings, P.J., and Lupski, J.R. (2012). Mechanisms for recurrent and complex human genomic rearrangements. Curr Opin Genet Dev 22, 211-220.

Lo-Castro, A., Galasso, C., Cerminara, C., El-Malhany, N., Benedetti, S., Nardone, A.M., and Curatolo, P. (2009). Association of syndromic mental retardation and autism with 22q11.2 duplication. Neuropediatrics 40, 137-140.

Losh, M., Childress, D., Lam, K., and Piven, J. (2008). Defining key features of the broad autism phenotype: a comparison across parents of multiple- and single-incidence autism families. Am J Med Genet B Neuropsychiatr Genet 147B, 424-433.

Lundstrom, S., Chang, Z., Kerekes, N., Gumpert, C.H., Rastam, M., Gillberg, C., Lichtenstein, P., and Anckarsater, H. (2011). Autistic-like traits and their association with mental health problems in two nationwide twin cohorts of children and adults. Psychol Med 41, 2423-2433.

Mackay-Sim, A. (2013). Patient-derived stem cells: pathways to drug discovery for brain diseases. Front Cell Neurosci 7, 29.

Mahajan, R., Bernal, M.P., Panzer, R., Whitaker, A., Roberts, W., Handen, B., Hardan, A., Anagnostou, E., and Veenstra-VanderWeele, J. (2012). Clinical practice pathways for evaluation and medication choice for attention-deficit/hyperactivity disorder symptoms in autism spectrum disorders. Pediatrics 130 Suppl 2, S125-138.

Malhotra, D., McCarthy, S., Michaelson, J.J., Vacic, V., Burdick, K.E., Yoon, S., Cichon, S., Corvin, A., Gary, S., Gershon, E.S., et al. (2011). High frequencies of de novo CNVs in bipolar disorder and schizophrenia. Neuron 72, 951-963.

Malhotra, D., and Sebat, J. (2012). CNVs: harbingers of a rare variant revolution in psychiatric genetics. Cell 148, 1223-1241.

Manolio, T.A. (2013). Bringing genome-wide association findings into clinical use. Nat Rev Genet 14, 549-558.

Manshadi, M., Lippmann, S., O'Daniel, R.G., and Blackman, A. (1983). Alcohol abuse and attention deficit disorder. J Clin Psychiatry 44, 379-380.

Mao, Y., Ge, X., Frank, C.L., Madison, J.M., Koehler, A.N., Doud, M.K., Tassa, C., Berry, E.M., Soda, T., Singh, K.K., et al. (2009). Disrupted in schizophrenia 1 regulates neuronal progenitor proliferation via modulation of GSK3beta/beta-catenin signaling. Cell 136, 1017-1031.

Page 186: Copy Number Variation in Neurodevelopmental Disorders

173

Marshall, C.R., Noor, A., Vincent, J.B., Lionel, A.C., Feuk, L., Skaug, J., Shago, M., Moessner, R., Pinto, D., Ren, Y., et al. (2008). Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet 82, 477-488.

Marshall, C.R., and Scherer, S.W. (2012). Detection and characterization of copy number variation in autism spectrum disorder. Methods Mol Biol 838, 115-135.

Martin, J., Hamshere, M.L., O'Donovan, M.C., Rutter, M., and Thapar, A. (2014). Factor structure of autistic traits in children with ADHD. J Autism Dev Disord 44, 204-215.

Martin, N.C., Piek, J., Baynam, G., Levy, F., and Hay, D. (2010). An examination of the relationship between movement problems and four common developmental disorders. Hum Mov Sci 29, 799-808.

Matson, J.L., and Shoemaker, M. (2009). Intellectual disability and its relationship to autism spectrum disorders. Res Dev Disabil 30, 1107-1114.

Matsunami, N., Hadley, D., Hensel, C.H., Christensen, G.B., Kim, C., Frackelton, E., Thomas, K., da Silva, R.P., Stevens, J., Baird, L., et al. (2013). Identification of rare recurrent copy number variants in high-risk autism families and their prevalence in a large ASD population. PLoS One 8, e52239.

Mazzoni, P., Kimhy, D., Khan, S., Posner, K., Maayan, L., Eilenberg, M., Messinger, J., Kestenbaum, C., and Corcoran, C. (2009). Childhood onset diagnoses in a case series of teens at clinical high risk for psychosis. J Child Adolesc Psychopharmacol 19, 771-776.

McCarroll, S.A., Kuruvilla, F.G., Korn, J.M., Cawley, S., Nemesh, J., Wysoker, A., Shapero, M.H., de Bakker, P.I., Maller, J.B., Kirby, A., et al. (2008). Integrated detection and population-genetic analysis of SNPs and copy number variation. Nat Genet 40, 1166-1174.

McCarthy, M.I., and Hirschhorn, J.N. (2008). Genome-wide association studies: past, present and future. Hum Mol Genet 17, R100-101.

McCarthy, S.E., Makarov, V., Kirov, G., Addington, A.M., McClellan, J., Yoon, S., Perkins, D.O., Dickel, D.E., Kusenda, M., Krastoshevsky, O., et al. (2009). Microduplications of 16p11.2 are associated with schizophrenia. Nat Genet 41, 1223-1227.

McMahon, H.T., Missler, M., Li, C., and Sudhof, T.C. (1995). Complexins: cytosolic proteins that regulate SNAP receptor function. Cell 83, 111-119.

McPheeters, M.L., Warren, Z., Sathe, N., Bruzek, J.L., Krishnaswami, S., Jerome, R.N., and Veenstra-Vanderweele, J. (2011). A systematic review of medical treatments for children with autism spectrum disorders. Pediatrics 127, e1312-1321.

Melhem, N., Middleton, F., McFadden, K., Klei, L., Faraone, S.V., Vinogradov, S., Tiobech, J., Yano, V., Kuartei, S., Roeder, K., et al. (2011). Copy number variants for schizophrenia and related psychotic disorders in Oceanic Palau: risk and transmission in extended pedigrees. Biol Psychiatry 70, 1115-1121.

Page 187: Copy Number Variation in Neurodevelopmental Disorders

174

Merico, D., Isserlin, R., Stueker, O., Emili, A., and Bader, G.D. (2010). Enrichment map: a network-based method for gene-set enrichment visualization and interpretation. PLoS One 5, e13984.

Merikangas, A.K., Corvin, A.P., and Gallagher, L. (2009). Copy-number variants in neurodevelopmental disorders: promises and challenges. Trends Genet 25, 536-544.

Michaelson, J.J., Shi, Y., Gujral, M., Zheng, H., Malhotra, D., Jin, X., Jian, M., Liu, G., Greer, D., Bhandari, A., et al. (2012). Whole-genome sequencing in autism identifies hot spots for de novo germline mutation. Cell 151, 1431-1442.

Millar, J.K., Wilson-Annan, J.C., Anderson, S., Christie, S., Taylor, M.S., Semple, C.A., Devon, R.S., St Clair, D.M., Muir, W.J., Blackwood, D.H., et al. (2000). Disruption of two novel genes by a translocation co-segregating with schizophrenia. Hum Mol Genet 9, 1415-1423.

Miller, D.T., Adam, M.P., Aradhya, S., Biesecker, L.G., Brothman, A.R., Carter, N.P., Church, D.M., Crolla, J.A., Eichler, E.E., Epstein, C.J., et al. (2010a). Consensus statement: chromosomal microarray is a first-tier clinical diagnostic test for individuals with developmental disabilities or congenital anomalies. Am J Hum Genet 86, 749-764.

Miller, D.T., Shen, Y., Weiss, L.A., Korn, J., Anselm, I., Bridgemohan, C., Cox, G.F., Dickinson, H., Gentile, J., Harris, D.J., et al. (2009). Microdeletion/duplication at 15q13.2q13.3 among individuals with features of autism and other neuropsychiatric disorders. J Med Genet 46, 242-248.

Miller, F.A., Hayeems, R.Z., and Bytautas, J.P. (2010b). What is a meaningful result? Disclosing the results of genomic research in autism to research participants. Eur J Hum Genet 18, 867-871.

Miranda, A., Mercader, J., Fernandez, M.I., and Colomer, C. (2013). Reading Performance of Young Adults With ADHD Diagnosed in Childhood: Relations With Executive Functioning. J Atten Disord.

Missler, M., Zhang, W., Rohlmann, A., Kattenstroth, G., Hammer, R.E., Gottmann, K., and Sudhof, T.C. (2003). Alpha-neurexins couple Ca2+ channels to synaptic vesicle exocytosis. Nature 423, 939-948.

Mody, M., and Belliveau, J.W. (2013). Speech and Language Impairments in Autism: Insights from Behavior and Neuroimaging. N Am J Med Sci (Boston) 5, 157-161.

Moessner, R., Marshall, C.R., Sutcliffe, J.S., Skaug, J., Pinto, D., Vincent, J., Zwaigenbaum, L., Fernandez, B., Roberts, W., Szatmari, P., et al. (2007). Contribution of SHANK3 mutations to autism spectrum disorder. Am J Hum Genet 81, 1289-1297.

Molin, A.M., Andrieux, J., Koolen, D.A., Malan, V., Carella, M., Colleaux, L., Cormier-Daire, V., David, A., de Leeuw, N., Delobel, B., et al. (2012). A novel microdeletion syndrome at 3q13.31 characterised by developmental delay, postnatal overgrowth, hypoplastic male genitals, and characteristic facial features. J Med Genet 49, 104-109.

Page 188: Copy Number Variation in Neurodevelopmental Disorders

175

Moore, S., Kelleher, E., and Corvin, A. (2011). The shock of the new: progress in schizophrenia genomics. Curr Genomics 12, 516-524.

Moreno-De-Luca, A., Myers, S.M., Challman, T.D., Moreno-De-Luca, D., Evans, D.W., and Ledbetter, D.H. (2013). Developmental brain dysfunction: revival and expansion of old concepts based on new genetic evidence. Lancet Neurol 12, 406-414.

Moreno-De-Luca, D., Mulle, J.G., Kaminsky, E.B., Sanders, S.J., Myers, S.M., Adam, M.P., Pakula, A.T., Eisenhauer, N.J., Uhas, K., Weik, L., et al. (2010). Deletion 17q12 is a recurrent copy number variant that confers high risk of autism and schizophrenia. Am J Hum Genet 87, 618-630.

Morrow, E.M., Yoo, S.Y., Flavell, S.W., Kim, T.K., Lin, Y., Hill, R.S., Mukaddes, N.M., Balkhy, S., Gascon, G., Hashmi, A., et al. (2008). Identifying autism loci and genes by tracing recent shared ancestry. Science 321, 218-223.

Mouridsen, S.E., Rich, B., and Isager, T. (2008). Psychiatric disorders in adults diagnosed as children with atypical autism. A case control study. J Neural Transm 115, 135-138.

Muglia, P. (2011). From genes to therapeutic targets for psychiatric disorders - what to expect? Curr Opin Pharmacol 11, 563-571.

Mulle, J.G. (2012). Schizophrenia genetics: progress, at last. Curr Opin Genet Dev 22, 238-244.

Mulle, J.G., Dodd, A.F., McGrath, J.A., Wolyniec, P.S., Mitchell, A.A., Shetty, A.C., Sobreira, N.L., Valle, D., Rudd, M.K., Satten, G., et al. (2010). Microdeletions of 3q29 confer high risk for schizophrenia. Am J Hum Genet 87, 229-236.

Mulley, J.C., and Mefford, H.C. (2011). Epilepsy and the new cytogenetics. Epilepsia 52, 423-432.

Mulligan, A., Anney, R.J., O'Regan, M., Chen, W., Butler, L., Fitzgerald, M., Buitelaar, J., Steinhausen, H.C., Rothenberger, A., Minderaa, R., et al. (2009). Autism symptoms in Attention-Deficit/Hyperactivity Disorder: a familial trait which correlates with conduct, oppositional defiant, language and motor disorders. J Autism Dev Disord 39, 197-209.

Murray, M.J. (2010). Attention-deficit/Hyperactivity Disorder in the context of Autism spectrum disorders. Curr Psychiatry Rep 12, 382-388.

Myers, S.M., and Johnson, C.P. (2007). Management of children with autism spectrum disorders. Pediatrics 120, 1162-1182.

Nagamani, S.C., Erez, A., Ben-Zeev, B., Frydman, M., Winter, S., Zeller, R., El-Khechen, D., Escobar, L., Stankiewicz, P., Patel, A., et al. (2013). Detection of copy-number variation in AUTS2 gene by targeted exonic array CGH in patients with developmental delay and autistic spectrum disorders. Eur J Hum Genet 21, 343-346.

Page 189: Copy Number Variation in Neurodevelopmental Disorders

176

Neale, B.M., Kou, Y., Liu, L., Ma'ayan, A., Samocha, K.E., Sabo, A., Lin, C.F., Stevens, C., Wang, L.S., Makarov, V., et al. (2012). Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485, 242-245.

Neale, B.M., Lasky-Su, J., Anney, R., Franke, B., Zhou, K., Maller, J.B., Vasquez, A.A., Asherson, P., Chen, W., Banaschewski, T., et al. (2008). Genome-wide association scan of attention deficit hyperactivity disorder. Am J Med Genet B Neuropsychiatr Genet 147B, 1337-1344.

Neale, B.M., Medland, S., Ripke, S., Anney, R.J., Asherson, P., Buitelaar, J., Franke, B., Gill, M., Kent, L., Holmans, P., et al. (2010a). Case-control genome-wide association study of attention-deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry 49, 906-920.

Neale, B.M., Medland, S.E., Ripke, S., Asherson, P., Franke, B., Lesch, K.P., Faraone, S.V., Nguyen, T.T., Schafer, H., Holmans, P., et al. (2010b). Meta-analysis of genome-wide association studies of attention-deficit/hyperactivity disorder. J Am Acad Child Adolesc Psychiatry 49, 884-897.

Need, A.C., McEvoy, J.P., Gennarelli, M., Heinzen, E.L., Ge, D., Maia, J.M., Shianna, K.V., He, M., Cirulli, E.T., Gumbs, C.E., et al. (2012). Exome sequencing followed by large-scale genotyping suggests a limited role for moderately rare risk factors of strong effect in schizophrenia. Am J Hum Genet 91, 303-312.

Nijmeijer, J.S., Hoekstra, P.J., Minderaa, R.B., Buitelaar, J.K., Altink, M.E., Buschgens, C.J., Fliers, E.A., Rommelse, N.N., Sergeant, J.A., and Hartman, C.A. (2009). PDD symptoms in ADHD, an independent familial trait? J Abnorm Child Psychol 37, 443-453.

Nishimura, T., Kubosaki, A., Ito, Y., and Notkins, A.L. (2009). Disturbances in the secretion of neurotransmitters in IA-2/IA-2beta null mice: changes in behavior, learning and lifespan. Neuroscience 159, 427-437.

Nomura, J., and Takumi, T. (2012). Animal models of psychiatric disorders that reflect human copy number variation. Neural Plast 2012, 589524.

Noor, A., Whibley, A., Marshall, C.R., Gianakopoulos, P.J., Piton, A., Carson, A.R., Orlic-Milacic, M., Lionel, A.C., Sato, D., Pinto, D., et al. (2010). Disruption at the PTCHD1 Locus on Xp22.11 in Autism spectrum disorder and intellectual disability. Sci Transl Med 2, 49ra68.

O'Donovan, M.C., Craddock, N., Norton, N., Williams, H., Peirce, T., Moskvina, V., Nikolov, I., Hamshere, M., Carroll, L., Georgieva, L., et al. (2008). Identification of loci associated with schizophrenia by genome-wide association and follow-up. Nat Genet 40, 1053-1055.

O'Halloran, C.J., Kinsella, G.J., and Storey, E. (2012). The cerebellum and neuropsychological functioning: a critical review. J Clin Exp Neuropsychol 34, 35-56.

O'Roak, B.J., Deriziotis, P., Lee, C., Vives, L., Schwartz, J.J., Girirajan, S., Karakoc, E., Mackenzie, A.P., Ng, S.B., Baker, C., et al. (2011). Exome sequencing in sporadic autism spectrum disorders identifies severe de novo mutations. Nat Genet 43, 585-589.

Page 190: Copy Number Variation in Neurodevelopmental Disorders

177

O'Roak, B.J., Vives, L., Girirajan, S., Karakoc, E., Krumm, N., Coe, B.P., Levy, R., Ko, A., Lee, C., Smith, J.D., et al. (2012). Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485, 246-250.

Ollila, H.M., Soronen, P., Silander, K., Palo, O.M., Kieseppa, T., Kaunisto, M.A., Lonnqvist, J., Peltonen, L., Partonen, T., and Paunio, T. (2009). Findings from bipolar disorder genome-wide association studies replicate in a Finnish bipolar family-cohort. Mol Psychiatry 14, 351-353.

Oti, M., Huynen, M.A., and Brunner, H.G. (2008). Phenome connections. Trends Genet 24, 103-106.

Owen, M.J., O'Donovan, M.C., Thapar, A., and Craddock, N. (2011). Neurodevelopmental hypothesis of schizophrenia. Br J Psychiatry 198, 173-175.

Ozonoff, S., Young, G.S., Carter, A., Messinger, D., Yirmiya, N., Zwaigenbaum, L., Bryson, S., Carver, L.J., Constantino, J.N., Dobkins, K., et al. (2011). Recurrence risk for autism spectrum disorders: a baby siblings research consortium study. Pediatrics 128, e488-495.

Pagnamenta, A.T., Bacchelli, E., de Jonge, M.V., Mirza, G., Scerri, T.S., Minopoli, F., Chiocchetti, A., Ludwig, K.U., Hoffmann, P., Paracchini, S., et al. (2010). Characterization of a family with rare deletions in CNTNAP5 and DOCK4 suggests novel risk loci for autism and dyslexia. Biol Psychiatry 68, 320-328.

Paridaen, J.T., Danesin, C., Elas, A.T., van de Water, S., Houart, C., and Zivkovic, D. (2009). Apc1 is required for maintenance of local brain organizers and dorsal midbrain survival. Dev Biol 331, 101-112.

Parisi, P., Miano, S., Mei, D., Paolino, M.C., Castaldo, R., and Villa, M.P. (2010). Diffuse subcortical band heterotopia, periodic limb movements during sleep and a novel "de novo" mutation in the DCX gene. Brain Dev 32, 511-515.

Pelletier, A.L., and Mittal, V.A. (2012). An autism dimension for schizophrenia in the next Diagnostic and Statistical Manual? Schizophr Res 137, 269-270.

Penzes, P., Buonanno, A., Passafaro, M., Sala, C., and Sweet, R.A. (2013). Developmental vulnerability of synapses and circuits associated with neuropsychiatric disorders. J Neurochem 126, 165-182.

Peralta, V., de Jalon, E.G., Campos, M.S., Zandio, M., Sanchez-Torres, A., and Cuesta, M.J. (2011). The meaning of childhood attention-deficit hyperactivity symptoms in patients with a first-episode of schizophrenia-spectrum psychosis. Schizophr Res 126, 28-35.

Pescosolido, M.F., Gamsiz, E.D., Nagpal, S., and Morrow, E.M. (2013). Distribution of disease-associated copy number variants across distinct disorders of cognitive development. J Am Acad Child Adolesc Psychiatry 52, 414-430 e414.

Pescosolido, M.F., Yang, U., Sabbagh, M., and Morrow, E.M. (2012). Lighting a path: genetic studies pinpoint neurodevelopmental mechanisms in autism and related disorders. Dialogues Clin Neurosci 14, 239-252.

Page 191: Copy Number Variation in Neurodevelopmental Disorders

178

Pierce, K., and Courchesne, E. (2001). Evidence for a cerebellar role in reduced exploration and stereotyped behavior in autism. Biol Psychiatry 49, 655-664.

Pinto, D., Darvishi, K., Shi, X., Rajan, D., Rigler, D., Fitzgerald, T., Lionel, A.C., Thiruvahindrapuram, B., Macdonald, J.R., Mills, R., et al. (2011). Comprehensive assessment of array-based platforms and calling algorithms for detection of copy number variants. Nat Biotechnol 29, 512-520.

Pinto, D., Pagnamenta, A.T., Klei, L., Anney, R., Merico, D., Regan, R., Conroy, J., Magalhaes, T.R., Correia, C., Abrahams, B.S., et al. (2010). Functional impact of global rare copy number variation in autism spectrum disorders. Nature 466, 368-372.

Pober, B.R. (2010). Williams-Beuren syndrome. N Engl J Med 362, 239-252.

Polanczyk, G., de Lima, M.S., Horta, B.L., Biederman, J., and Rohde, L.A. (2007). The worldwide prevalence of ADHD: a systematic review and metaregression analysis. Am J Psychiatry 164, 942-948.

Politte, L.C., and McDougle, C.J. (2013). Atypical antipsychotics in the treatment of children and adolescents with pervasive developmental disorders. Psychopharmacology (Berl).

Poot, M. (2013). Towards identification of individual etiologies by resolving genomic and biological conundrums in patients with autism spectrum disorders. Mol Syndromol 4, 213-226.

Poot, M., van der Smagt, J.J., Brilstra, E.H., and Bourgeron, T. (2011). Disentangling the myriad genomics of complex disorders, specifically focusing on autism, epilepsy, and schizophrenia. Cytogenet Genome Res 135, 228-240.

Posey, D.J., Wiegand, R.E., Wilkerson, J., Maynard, M., Stigler, K.A., and McDougle, C.J. (2006). Open-label atomoxetine for attention-deficit/ hyperactivity disorder symptoms associated with high-functioning pervasive developmental disorders. J Child Adolesc Psychopharmacol 16, 599-610.

Posthuma, D., and Polderman, T.J. (2013). What have we learned from recent twin studies about the etiology of neurodevelopmental disorders? Curr Opin Neurol 26, 111-121.

Poulopoulos, A., Aramuni, G., Meyer, G., Soykan, T., Hoon, M., Papadopoulos, T., Zhang, M., Paarmann, I., Fuchs, C., Harvey, K., et al. (2009). Neuroligin 2 drives postsynaptic assembly at perisomatic inhibitory synapses through gephyrin and collybistin. Neuron 63, 628-642.

Prasad, A., Merico, D., Thiruvahindrapuram, B., Wei, J., Lionel, A.C., Sato, D., Rickaby, J., Lu, C., Szatmari, P., Roberts, W., et al. (2012). A discovery resource of rare copy number variations in individuals with autism spectrum disorder. G3 (Bethesda) 2, 1665-1685.

Pringle, B., Colpe, L.J., Blumberg, S.J., Avila, R.M., and Kogan, M.D. (2012). Diagnostic history and treatment of school-aged children with autism spectrum disorder and special health care needs. NCHS Data Brief, 1-8.

Page 192: Copy Number Variation in Neurodevelopmental Disorders

179

Prior, P., Schmitt, B., Grenningloh, G., Pribilla, I., Multhaup, G., Beyreuther, K., Maulet, Y., Werner, P., Langosch, D., Kirsch, J., et al. (1992). Primary structure and alternative splice variants of gephyrin, a putative glycine receptor-tubulin linker protein. Neuron 8, 1161-1170.

Pritchard, J.K., Stephens, M., and Donnelly, P. (2000). Inference of population structure using multilocus genotype data. Genetics 155, 945-959.

Purcell, S., Neale, B., Todd-Brown, K., Thomas, L., Ferreira, M.A., Bender, D., Maller, J., Sklar, P., de Bakker, P.I., Daly, M.J., et al. (2007). PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet 81, 559-575.

Purcell, S.M., Wray, N.R., Stone, J.L., Visscher, P.M., O'Donovan, M.C., Sullivan, P.F., and Sklar, P. (2009). Common polygenic variation contributes to risk of schizophrenia and bipolar disorder. Nature 460, 748-752.

Ramocki, M.B., and Zoghbi, H.Y. (2008). Failure of neuronal homeostasis results in common neuropsychiatric phenotypes. Nature 455, 912-918.

Ramos-Quiroga, J.A., Sanchez-Mora, C., Casas, M., Garcia-Martinez, I., Bosch, R., Nogueira, M., Corrales, M., Palomar, G., Vidal, R., Coll-Tane, M., et al. (2014). Genome-wide copy number variation analysis in adult attention-deficit and hyperactivity disorder. J Psychiatr Res 49, 60-67.

Rapoport, J., Chavez, A., Greenstein, D., Addington, A., and Gogtay, N. (2009). Autism spectrum disorders and childhood-onset schizophrenia: clinical and biological contributions to a relation revisited. J Am Acad Child Adolesc Psychiatry 48, 10-18.

Raychaudhuri, S., Plenge, R.M., Rossin, E.J., Ng, A.C., Purcell, S.M., Sklar, P., Scolnick, E.M., Xavier, R.J., Altshuler, D., and Daly, M.J. (2009). Identifying relationships among genomic disease regions: predicting genes at pathogenic SNP associations and rare deletions. PLoS Genet 5, e1000534.

Reaven, J.A., Hepburn, S.L., and Ross, R.G. (2008). Use of the ADOS and ADI-R in children with psychosis: importance of clinical judgment. Clin Child Psychol Psychiatry 13, 81-94.

Rees, E., Walters, J.T., Chambert, K.D., O'Dushlaine, C., Szatkiewicz, J., Richards, A.L., Georgieva, L., Mahoney-Davies, G., Legge, S.E., Moran, J.L., et al. (2013). CNV analysis in a large schizophrenia sample implicates deletions at 16p12.1 and SLC1A1 and duplications at 1p36.33 and CGNL1. Hum Mol Genet.

Reichelt, A.C., Rodgers, R.J., and Clapcote, S.J. (2012). The role of neurexins in schizophrenia and autistic spectrum disorder. Neuropharmacology 62, 1519-1526.

Reiersen, A.M., Constantino, J.N., Grimmer, M., Martin, N.G., and Todd, R.D. (2008). Evidence for shared genetic influences on self-reported ADHD and autistic symptoms in young adult Australian twins. Twin Res Hum Genet 11, 579-585.

Reiersen, A.M., Constantino, J.N., Volk, H.E., and Todd, R.D. (2007). Autistic traits in a population-based ADHD twin sample. J Child Psychol Psychiatry 48, 464-472.

Page 193: Copy Number Variation in Neurodevelopmental Disorders

180

Reiss, J., Gross-Hardt, S., Christensen, E., Schmidt, P., Mendel, R.R., and Schwarz, G. (2001). A mutation in the gene for the neurotransmitter receptor-clustering protein gephyrin causes a novel form of molybdenum cofactor deficiency. Am J Hum Genet 68, 208-213.

Reiss, J., and Hahnewald, R. (2011). Molybdenum cofactor deficiency: Mutations in GPHN, MOCS1, and MOCS2. Hum Mutat 32, 10-18.

Reiss, J., Lenz, U., Aquaviva-Bourdain, C., Joriot-Chekaf, S., Mention-Mulliez, K., and Holder-Espinasse, M. (2011). A GPHN point mutation leading to molybdenum cofactor deficiency. Clin Genet 80, 598-599.

Riley, B., Thiselton, D., Maher, B.S., Bigdeli, T., Wormley, B., McMichael, G.O., Fanous, A.H., Vladimirov, V., O'Neill, F.A., Walsh, D., et al. (2010). Replication of association between schizophrenia and ZNF804A in the Irish Case-Control Study of Schizophrenia sample. Mol Psychiatry 15, 29-37.

Risch, N., Spiker, D., Lotspeich, L., Nouri, N., Hinds, D., Hallmayer, J., Kalaydjieva, L., McCague, P., Dimiceli, S., Pitts, T., et al. (1999). A genomic screen of autism: evidence for a multilocus etiology. Am J Hum Genet 65, 493-507.

Risi, S., Lord, C., Gotham, K., Corsello, C., Chrysler, C., Szatmari, P., Cook, E.H., Jr., Leventhal, B.L., and Pickles, A. (2006). Combining information from multiple sources in the diagnosis of autism spectrum disorders. J Am Acad Child Adolesc Psychiatry 45, 1094-1103.

Romanos, M., Freitag, C., Jacob, C., Craig, D.W., Dempfle, A., Nguyen, T.T., Halperin, R., Walitza, S., Renner, T.J., Seitz, C., et al. (2008). Genome-wide linkage analysis of ADHD using high-density SNP arrays: novel loci at 5q13.1 and 14q12. Mol Psychiatry 13, 522-530.

Rommelse, N.N., Franke, B., Geurts, H.M., Hartman, C.A., and Buitelaar, J.K. (2010). Shared heritability of attention-deficit/hyperactivity disorder and autism spectrum disorder. Eur Child Adolesc Psychiatry 19, 281-295.

Rommelse, N.N., Geurts, H.M., Franke, B., Buitelaar, J.K., and Hartman, C.A. (2011). A review on cognitive and brain endophenotypes that may be common in autism spectrum disorder and attention-deficit/hyperactivity disorder and facilitate the search for pleiotropic genes. Neurosci Biobehav Rev 35, 1363-1396.

Ronald, A., Simonoff, E., Kuntsi, J., Asherson, P., and Plomin, R. (2008). Evidence for overlapping genetic influences on autistic and ADHD behaviours in a community twin sample. J Child Psychol Psychiatry 49, 535-542.

Ronemus, M., Iossifov, I., Levy, D., and Wigler, M. (2014). The role of de novo mutations in the genetics of autism spectrum disorders. Nat Rev Genet 15, 133-141.

Rosenberg, R.E., Law, J.K., Yenokyan, G., McGready, J., Kaufmann, W.E., and Law, P.A. (2009). Characteristics and concordance of autism spectrum disorders among 277 twin pairs. Arch Pediatr Adolesc Med 163, 907-914.

Page 194: Copy Number Variation in Neurodevelopmental Disorders

181

Rosenfeld, J.A., Ballif, B.C., Torchia, B.S., Sahoo, T., Ravnan, J.B., Schultz, R., Lamb, A., Bejjani, B.A., and Shaffer, L.G. (2010). Copy number variations associated with autism spectrum disorders contribute to a spectrum of neurodevelopmental disorders. Genet Med 12, 694-702.

Rosenfeld, J.A., Coe, B.P., Eichler, E.E., Cuckle, H., and Shaffer, L.G. (2013). Estimates of penetrance for recurrent pathogenic copy-number variations. Genet Med 15, 478-481.

Rosenthal, M., Wallace, G.L., Lawson, R., Wills, M.C., Dixon, E., Yerys, B.E., and Kenworthy, L. (2013). Impairments in real-world executive function increase from childhood to adolescence in autism spectrum disorders. Neuropsychology 27, 13-18.

Rowen, L., Young, J., Birditt, B., Kaur, A., Madan, A., Philipps, D.L., Qin, S., Minx, P., Wilson, R.K., Hood, L., et al. (2002). Analysis of the human neurexin genes: alternative splicing and the generation of protein diversity. Genomics 79, 587-597.

Rubenstein, J.L., and Merzenich, M.M. (2003). Model of autism: increased ratio of excitation/inhibition in key neural systems. Genes Brain Behav 2, 255-267.

Rujescu, D., Ingason, A., Cichon, S., Pietilainen, O.P., Barnes, M.R., Toulopoulou, T., Picchioni, M., Vassos, E., Ettinger, U., Bramon, E., et al. (2009). Disruption of the neurexin 1 gene is associated with schizophrenia. Hum Mol Genet 18, 988-996.

Russell, G., and Pavelka, Z. (2013). Co-Occurrence of Developmental Disorders: Children Who Share Symptoms of Autism, Dyslexia and Attention Deficit Hyperactivity Disorder. In Recent Advances in Autism Spectrum Disorders, M. Fitzgerald, ed. (InTech).

Rzhetsky, A., Wajngurt, D., Park, N., and Zheng, T. (2007). Probing genetic overlap among complex human phenotypes. Proc Natl Acad Sci USA 104, 11694-11699.

Sahoo, T., Theisen, A., Rosenfeld, J.A., Lamb, A.N., Ravnan, J.B., Schultz, R.A., Torchia, B.S., Neill, N., Casci, I., Bejjani, B.A., et al. (2011). Copy number variants of schizophrenia susceptibility loci are associated with a spectrum of speech and developmental delays and behavior problems. Genet Med 13, 868-880.

Sakai, Y., Shaw, C.A., Dawson, B.C., Dugas, D.V., Al-Mohtaseb, Z., Hill, D.E., and Zoghbi, H.Y. (2011). Protein interactome reveals converging molecular pathways among autism disorders. Sci Transl Med 3, 86ra49.

Sakurai, T. (2012). The role of NrCAM in neural development and disorders--beyond a simple glue in the brain. Mol Cell Neurosci 49, 351-363.

Salyakina, D., Cukier, H.N., Lee, J.M., Sacharow, S., Nations, L.D., Ma, D., Jaworski, J.M., Konidari, I., Whitehead, P.L., Wright, H.H., et al. (2011). Copy number variants in extended autism spectrum disorder families reveal candidates potentially involved in autism risk. PLoS One 6, e26049.

Sanders, S.J., Ercan-Sencicek, A.G., Hus, V., Luo, R., Murtha, M.T., Moreno-De-Luca, D., Chu, S.H., Moreau, M.P., Gupta, A.R., Thomson, S.A., et al. (2011). Multiple recurrent de novo

Page 195: Copy Number Variation in Neurodevelopmental Disorders

182

CNVs, including duplications of the 7q11.23 Williams syndrome region, are strongly associated with autism. Neuron 70, 863-885.

Sanders, S.J., Murtha, M.T., Gupta, A.R., Murdoch, J.D., Raubeson, M.J., Willsey, A.J., Ercan-Sencicek, A.G., DiLullo, N.M., Parikshak, N.N., Stein, J.L., et al. (2012). De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485, 237-241.

Santosh, P.J., Baird, G., Pityaratstian, N., Tavare, E., and Gringras, P. (2006). Impact of comorbid autism spectrum disorders on stimulant response in children with attention deficit hyperactivity disorder: a retrospective and prospective effectiveness study. Child Care Health Dev 32, 575-583.

Sato, D., Lionel, A.C., Leblond, C.S., Prasad, A., Pinto, D., Walker, S., O'Connor, I., Russell, C., Drmic, I.E., Hamdan, F.F., et al. (2012). SHANK1 Deletions in Males with Autism Spectrum Disorder. Am J Hum Genet 90, 879-887.

Schaaf, C.P., Boone, P.M., Sampath, S., Williams, C., Bader, P.I., Mueller, J.M., Shchelochkov, O.A., Brown, C.W., Crawford, H.P., Phalen, J.A., et al. (2012). Phenotypic spectrum and genotype-phenotype correlations of NRXN1 exon deletions. Eur J Hum Genet 20, 1240-1247.

Schwamborn, J.C., Berezikov, E., and Knoblich, J.A. (2009). The TRIM-NHL protein TRIM32 activates microRNAs and prevents self-renewal in mouse neural progenitors. Cell 136, 913-925.

Sebat, J., Lakshmi, B., Malhotra, D., Troge, J., Lese-Martin, C., Walsh, T., Yamrom, B., Yoon, S., Krasnitz, A., Kendall, J., et al. (2007). Strong association of de novo copy number mutations with autism. Science 316, 445-449.

Seno, M.M., Hu, P., Gwadry, F.G., Pinto, D., Marshall, C.R., Casallo, G., and Scherer, S.W. (2011). Gene and miRNA expression profiles in autism spectrum disorders. Brain Res 1380, 85-97.

Shcheglovitov, A., Shcheglovitova, O., Yazawa, M., Portmann, T., Shu, R., Sebastiano, V., Krawisz, A., Froehlich, W., Bernstein, J.A., Hallmayer, J.F., et al. (2013). SHANK3 and IGF1 restore synaptic deficits in neurons from 22q13 deletion syndrome patients. Nature 503, 267-271.

Sheitman, B.B., Kraus, J.E., Bodfish, J.W., and Carmel, H. (2004). Are the negative symptoms of schizophrenia consistent with an autistic spectrum illness? Schizophr Res 69, 119-120.

Shen, Y., Dies, K.A., Holm, I.A., Bridgemohan, C., Sobeih, M.M., Caronna, E.B., Miller, K.J., Frazier, J.A., Silverstein, I., Picker, J., et al. (2010). Clinical genetic testing for patients with autism spectrum disorders. Pediatrics 125, e727-735.

Shi, J., Levinson, D.F., Duan, J., Sanders, A.R., Zheng, Y., Pe'er, I., Dudbridge, F., Holmans, P.A., Whittemore, A.S., Mowry, B.J., et al. (2009). Common variants on chromosome 6p22.1 are associated with schizophrenia. Nature 460, 753-757.

Shih, R.A., Belmonte, P.L., and Zandi, P.P. (2004). A review of the evidence from family, twin and adoption studies for a genetic contribution to adult psychiatric disorders. Int Rev Psychiatry 16, 260-283.

Page 196: Copy Number Variation in Neurodevelopmental Disorders

183

Shimojima, K., Sugawara, M., Shichiji, M., Mukaida, S., Takayama, R., Imai, K., and Yamamoto, T. (2011). Loss-of-function mutation of collybistin is responsible for X-linked mental retardation associated with epilepsy. J Hum Genet 56, 561-565.

Shinawi, M., Liu, P., Kang, S.H., Shen, J., Belmont, J.W., Scott, D.A., Probst, F.J., Craigen, W.J., Graham, B.H., Pursley, A., et al. (2010). Recurrent reciprocal 16p11.2 rearrangements associated with global developmental delay, behavioural problems, dysmorphism, epilepsy, and abnormal head size. J Med Genet 47, 332-341.

Silversides, C.K., Lionel, A.C., Costain, G., Merico, D., Migita, O., Liu, B., Yuen, T., Rickaby, J., Thiruvahindrapuram, B., Marshall, C.R., et al. (2012). Rare Copy Number Variations in Adults with Tetralogy of Fallot Implicate Novel Risk Gene Pathways. PLoS Genet 8, e1002843.

Simonoff, E., Pickles, A., Charman, T., Chandler, S., Loucas, T., and Baird, G. (2008). Psychiatric disorders in children with autism spectrum disorders: prevalence, comorbidity, and associated factors in a population-derived sample. J Am Acad Child Adolesc Psychiatry 47, 921-929.

Sinzig, J., Walter, D., and Doepfner, M. (2009). Attention deficit/hyperactivity disorder in children and adolescents with autism spectrum disorder: symptom or syndrome? J Atten Disord 13, 117-126.

Smoller, J.W. (2013). Disorders and borders: psychiatric genetics and nosology. Am J Med Genet B Neuropsychiatr Genet 162B, 559-578.

Sporn, A.L., Addington, A.M., Gogtay, N., Ordonez, A.E., Gornick, M., Clasen, L., Greenstein, D., Tossell, J.W., Gochman, P., Lenane, M., et al. (2004). Pervasive developmental disorder and childhood-onset schizophrenia: comorbid disorder or a phenotypic variant of a very early onset illness? Biol Psychiatry 55, 989-994.

Sprenger, L., Buhler, E., Poustka, L., Bach, C., Heinzel-Gutenbrunner, M., Kamp-Becker, I., and Bachmann, C. (2013). Impact of ADHD symptoms on autism spectrum disorder symptom severity. Res Dev Disabil 34, 3545-3552.

Sprich, S., Biederman, J., Crawford, M.H., Mundy, E., and Faraone, S.V. (2000). Adoptive and biological families of children and adolescents with ADHD. J Am Acad Child Adolesc Psychiatry 39, 1432-1437.

Stallmeyer, B., Schwarz, G., Schulze, J., Nerlich, A., Reiss, J., Kirsch, J., and Mendel, R.R. (1999). The neurotransmitter receptor-anchoring protein gephyrin reconstitutes molybdenum cofactor biosynthesis in bacteria, plants, and mammalian cells. Proc Natl Acad Sci USA 96, 1333-1338.

Stankiewicz, P., Kulkarni, S., Dharmadhikari, A.V., Sampath, S., Bhatt, S.S., Shaikh, T.H., Xia, Z., Pursley, A.N., Cooper, M.L., Shinawi, M., et al. (2012). Recurrent deletions and reciprocal duplications of 10q11.21q11.23 including CHAT and SLC18A3 are likely mediated by complex low-copy repeats. Hum Mutat 33, 165-179.

Page 197: Copy Number Variation in Neurodevelopmental Disorders

184

Stefansson, H., Ophoff, R.A., Steinberg, S., Andreassen, O.A., Cichon, S., Rujescu, D., Werge, T., Pietilainen, O.P., Mors, O., Mortensen, P.B., et al. (2009). Common variants conferring risk of schizophrenia. Nature 460, 744-747.

Stefansson, H., Rujescu, D., Cichon, S., Pietilainen, O.P., Ingason, A., Steinberg, S., Fossdal, R., Sigurdsson, E., Sigmundsson, T., Buizer-Voskamp, J.E., et al. (2008). Large recurrent microdeletions associated with schizophrenia. Nature 455, 232-236.

Steffenburg, S., Gillberg, C., Hellgren, L., Andersson, L., Gillberg, I.C., Jakobsson, G., and Bohman, M. (1989). A twin study of autism in Denmark, Finland, Iceland, Norway and Sweden. J Child Psychol Psychiatry 30, 405-416.

Steinberg, S., Mors, O., Borglum, A.D., Gustafsson, O., Werge, T., Mortensen, P.B., Andreassen, O.A., Sigurdsson, E., Thorgeirsson, T.E., Bottcher, Y., et al. (2011). Expanding the range of ZNF804A variants conferring risk of psychosis. Mol Psychiatry 16, 59-66.

Stergiakouli, E., Hamshere, M., Holmans, P., Langley, K., Zaharieva, I., Hawi, Z., Kent, L., Gill, M., Williams, N., Owen, M.J., et al. (2012). Investigating the contribution of common genetic variants to the risk and pathogenesis of ADHD. Am J Psychiatry 169, 186-194.

Stewart, A.F., Dandona, S., Chen, L., Assogba, O., Belanger, M., Ewart, G., LaRose, R., Doelle, H., Williams, K., Wells, G.A., et al. (2009). Kinesin family member 6 variant Trp719Arg does not associate with angiographically defined coronary artery disease in the Ottawa Heart Genomics Study. J Am Coll Cardiol 53, 1471-1472.

Stewart, L.R., Hall, A.L., Kang, S.H., Shaw, C.A., and Beaudet, A.L. (2011). High frequency of known copy number abnormalities and maternal duplication 15q11-q13 in patients with combined schizophrenia and epilepsy. BMC Med Genet 12, 154.

Stone, W.S., and Iguchi, L. (2011). Do Apparent Overlaps between Schizophrenia and Autistic Spectrum Disorders Reflect Superficial Similarities or Etiological Commonalities? N Am J Med Sci (Boston) 4, 124-133.

Sudhof, T.C. (2008). Neuroligins and neurexins link synaptic function to cognitive disease. Nature 455, 903-911.

Sullivan, P.F., Daly, M.J., and O'Donovan, M. (2012a). Genetic architectures of psychiatric disorders: the emerging picture and its implications. Nat Rev Genet 13, 537-551.

Sullivan, P.F., Kendler, K.S., and Neale, M.C. (2003). Schizophrenia as a complex trait: evidence from a meta-analysis of twin studies. Arch Gen Psychiatry 60, 1187-1192.

Sullivan, P.F., Magnusson, C., Reichenberg, A., Boman, M., Dalman, C., Davidson, M., Fruchter, E., Hultman, C.M., Lundberg, M., Langstrom, N., et al. (2012b). Family history of schizophrenia and bipolar disorder as risk factors for autism. Arch Gen Psychiatry 69, 1099-1103.

Page 198: Copy Number Variation in Neurodevelopmental Disorders

185

Sullivan, S., Rai, D., Golding, J., Zammit, S., and Steer, C. (2013). The association between autism spectrum disorder and psychotic experiences in the Avon longitudinal study of parents and children (ALSPAC) birth cohort. J Am Acad Child Adolesc Psychiatry 52, 806-814 e802.

Sun, C., Cheng, M.C., Qin, R., Liao, D.L., Chen, T.T., Koong, F.J., Chen, G., and Chen, C.H. (2011). Identification and functional characterization of rare mutations of the neuroligin-2 gene (NLGN2) associated with schizophrenia. Hum Mol Genet 20, 3042-3051.

Sundaram, S.K., Huq, A.M., Wilson, B.J., and Chugani, H.T. (2010). Tourette syndrome is associated with recurrent exonic copy number variants. Neurology 74, 1583-1590.

Szatmari, P., Jones, M.B., Zwaigenbaum, L., and MacLean, J.E. (1998). Genetics of autism: overview and new directions. J Autism Dev Disord 28, 351-368.

Szatmari, P., Paterson, A.D., Zwaigenbaum, L., Roberts, W., Brian, J., Liu, X.Q., Vincent, J.B., Skaug, J.L., Thompson, A.P., Senman, L., et al. (2007). Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nat Genet 39, 319-328.

Tabuchi, K., and Sudhof, T.C. (2002). Structure and evolution of neurexin genes: insight into the mechanism of alternative splicing. Genomics 79, 849-859.

Talkowski, M.E., Mullegama, S.V., Rosenfeld, J.A., van Bon, B.W., Shen, Y., Repnikova, E.A., Gastier-Foster, J., Thrush, D.L., Kathiresan, S., Ruderfer, D.M., et al. (2011). Assessment of 2q23.1 microdeletion syndrome implicates MBD5 as a single causal locus of intellectual disability, epilepsy, and autism spectrum disorder. Am J Hum Genet 89, 551-563.

Taurines, R., Schwenck, C., Westerwald, E., Sachse, M., Siniatchkin, M., and Freitag, C. (2012). ADHD and autism: differential diagnosis or overlapping traits? A selective review. Atten Defic Hyperact Disord 4, 115-139.

Taylor, M.J., Charman, T., Robinson, E.B., Plomin, R., Happe, F., Asherson, P., and Ronald, A. (2013). Developmental associations between traits of autism spectrum disorder and attention deficit hyperactivity disorder: a genetically informative, longitudinal twin study. Psychol Med 43, 1735-1746.

Thapar, A., Cooper, M., Eyre, O., and Langley, K. (2013). What have we learnt about the causes of ADHD? J Child Psychol Psychiatry 54, 3-16.

Thapar, A., Langley, K., Owen, M.J., and O'Donovan, M.C. (2007). Advances in genetic findings on attention deficit hyperactivity disorder. Psychol Med 37, 1681-1692.

Toro, R., Konyukh, M., Delorme, R., Leblond, C., Chaste, P., Fauchereau, F., Coleman, M., Leboyer, M., Gillberg, C., and Bourgeron, T. (2010). Key role for gene dosage and synaptic homeostasis in autism spectrum disorders. Trends Genet 26, 363-372.

Tretter, V., Mukherjee, J., Maric, H.M., Schindelin, H., Sieghart, W., and Moss, S.J. (2012). Gephyrin, the enigmatic organizer at GABAergic synapses. Front Cell Neurosci 6, 23.

Page 199: Copy Number Variation in Neurodevelopmental Disorders

186

Tropeano, M., Ahn, J.W., Dobson, R.J., Breen, G., Rucker, J., Dixit, A., Pal, D.K., McGuffin, P., Farmer, A., White, P.S., et al. (2013). Male-biased autosomal effect of 16p13.11 copy number variation in neurodevelopmental disorders. PLoS One 8, e61365.

Tseng, W.L., and Gau, S.S. (2013). Executive function as a mediator in the link between attention-deficit/hyperactivity disorder and social problems. J Child Psychol Psychiatry 54, 996-1004.

Tuy, F.P., Saillour, Y., Kappeler, C., Chelly, J., and Francis, F. (2008). Alternative transcripts of Dclk1 and Dclk2 and their expression in doublecortin knockout mice. Dev Neurosci 30, 171-186.

Ullrich, B., Ushkaryov, Y.A., and Sudhof, T.C. (1995). Cartography of neurexins: more than 1000 isoforms generated by alternative splicing and expressed in distinct subsets of neurons. Neuron 14, 497-507.

Unenge Hallerback, M., Lugnegard, T., and Gillberg, C. (2012). Is autism spectrum disorder common in schizophrenia? Psychiatry Res 198, 12-17.

Vaags, A.K., Lionel, A.C., Sato, D., Goodenberger, M., Stein, Q.P., Curran, S., Ogilvie, C., Ahn, J.W., Drmic, I., Senman, L., et al. (2012). Rare deletions at the neurexin 3 locus in autism spectrum disorder. Am J Hum Genet 90, 133-141.

Vacic, V., McCarthy, S., Malhotra, D., Murray, F., Chou, H.H., Peoples, A., Makarov, V., Yoon, S., Bhandari, A., Corominas, R., et al. (2011). Duplications of the neuropeptide receptor gene VIPR2 confer significant risk for schizophrenia. Nature 471, 499-503.

van der Zwaag, B., Franke, L., Poot, M., Hochstenbach, R., Spierenburg, H.A., Vorstman, J.A., van Daalen, E., de Jonge, M.V., Verbeek, N.E., Brilstra, E.H., et al. (2009). Gene-network analysis identifies susceptibility genes related to glycobiology in autism. PLoS One 4, e5324.

Vannucchi, G., Masi, G., Toni, C., Dell'osso, L., Marazziti, D., and Perugi, G. (2013). Clinical features, developmental course, and psychiatric comorbidity of adult autism spectrum disorders. CNS Spectr, 1-8.

Veltman, J.A., and Brunner, H.G. (2012). De novo mutations in human genetic disease. Nat Rev Genet 13, 565-575.

Vincent, J.B., Horike, S.I., Choufani, S., Paterson, A.D., Roberts, W., Szatmari, P., Weksberg, R., Fernandez, B., and Scherer, S.W. (2006). An inversion inv(4)(p12-p15.3) in autistic siblings implicates the 4p GABA receptor gene cluster. J Med Genet 43, 429-434.

Vismara, L.A., and Rogers, S.J. (2010). Behavioral treatments in autism spectrum disorder: what do we know? Annu Rev Clin Psychol 6, 447-468.

Vrijenhoek, T., Buizer-Voskamp, J.E., van der Stelt, I., Strengman, E., Sabatti, C., Geurts van Kessel, A., Brunner, H.G., Ophoff, R.A., and Veltman, J.A. (2008). Recurrent CNVs disrupt three candidate genes in schizophrenia patients. Am J Hum Genet 83, 504-510.

Page 200: Copy Number Variation in Neurodevelopmental Disorders

187

Vulto-van Silfhout, A.T., Hehir-Kwa, J.Y., van Bon, B.W., Schuurs-Hoeijmakers, J.H., Meader, S., Hellebrekers, C.J., Thoonen, I.J., de Brouwer, A.P., Brunner, H.G., Webber, C., et al. (2013). Clinical significance of de novo and inherited copy-number variation. Hum Mutat 34, 1679-1687.

Walsh, T., McClellan, J.M., McCarthy, S.E., Addington, A.M., Pierce, S.B., Cooper, G.M., Nord, A.S., Kusenda, M., Malhotra, D., Bhandari, A., et al. (2008). Rare structural variants disrupt multiple genes in neurodevelopmental pathways in schizophrenia. Science 320, 539-543.

Waltereit, R., Banaschewski, T., Meyer-Lindenberg, A., and Poustka, L. (2013). Interaction of neurodevelopmental pathways and synaptic plasticity in mental retardation, autism spectrum disorder and schizophrenia: Implications for psychiatry. World J Biol Psychiatry.

Walters, R.G., Jacquemont, S., Valsesia, A., de Smith, A.J., Martinet, D., Andersson, J., Falchi, M., Chen, F., Andrieux, J., Lobbens, S., et al. (2010). A new highly penetrant form of obesity due to deletions on chromosome 16p11.2. Nature 463, 671-675.

Wang, K., Zhang, H., Ma, D., Bucan, M., Glessner, J.T., Abrahams, B.S., Salyakina, D., Imielinski, M., Bradfield, J.P., Sleiman, P.M., et al. (2009). Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature 459, 528-533.

Wang, K.S., Liu, X.F., and Aragam, N. (2010). A genome-wide meta-analysis identifies novel loci associated with schizophrenia and bipolar disorder. Schizophr Res 124, 192-199.

Waris, P., Lindberg, N., Kettunen, K., and Tani, P. (2013). The relationship between Asperger's syndrome and schizophrenia in adolescence. Eur Child Adolesc Psychiatry 22, 217-223.

Weischenfeldt, J., Symmons, O., Spitz, F., and Korbel, J.O. (2013). Phenotypic impact of genomic structural variation: insights from and for human disease. Nat Rev Genet 14, 125-138.

Weiss, L.A., Arking, D.E., Daly, M.J., and Chakravarti, A. (2009). A genome-wide linkage and association scan reveals novel loci for autism. Nature 461, 802-808.

Weiss, L.A., Shen, Y., Korn, J.M., Arking, D.E., Miller, D.T., Fossdal, R., Saemundsen, E., Stefansson, H., Ferreira, M.A., Green, T., et al. (2008). Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med 358, 667-675.

Werling, D.M., and Geschwind, D.H. (2013). Sex differences in autism spectrum disorders. Curr Opin Neurol 26, 146-153.

Williams, H.J., Craddock, N., Russo, G., Hamshere, M.L., Moskvina, V., Dwyer, S., Smith, R.L., Green, E., Grozeva, D., Holmans, P., et al. (2011a). Most genome-wide significant susceptibility loci for schizophrenia and bipolar disorder reported to date cross-traditional diagnostic boundaries. Hum Mol Genet 20, 387-391.

Williams, H.J., Norton, N., Dwyer, S., Moskvina, V., Nikolov, I., Carroll, L., Georgieva, L., Williams, N.M., Morris, D.W., Quinn, E.M., et al. (2011b). Fine mapping of ZNF804A and genome-wide significant evidence for its involvement in schizophrenia and bipolar disorder. Mol Psychiatry 16, 429-441.

Page 201: Copy Number Variation in Neurodevelopmental Disorders

188

Williams, N.M., Franke, B., Mick, E., Anney, R.J., Freitag, C.M., Gill, M., Thapar, A., O'Donovan, M.C., Owen, M.J., Holmans, P., et al. (2012). Genome-wide analysis of copy number variants in attention deficit hyperactivity disorder: the role of rare variants and duplications at 15q13.3. Am J Psychiatry 169, 195-204.

Williams, N.M., Zaharieva, I., Martin, A., Langley, K., Mantripragada, K., Fossdal, R., Stefansson, H., Stefansson, K., Magnusson, P., Gudmundsson, O.O., et al. (2010). Rare chromosomal deletions and duplications in attention-deficit hyperactivity disorder: a genome-wide analysis. Lancet 376, 1401-1408.

Wilson, G.M., Flibotte, S., Chopra, V., Melnyk, B.L., Honer, W.G., and Holt, R.A. (2006). DNA copy-number analysis in bipolar disorder and schizophrenia reveals aberrations in genes involved in glutamate signaling. Hum Mol Genet 15, 743-749.

Wilson, P.M., Fryer, R.H., Fang, Y., and Hatten, M.E. (2010). Astn2, a novel member of the astrotactin gene family, regulates the trafficking of ASTN1 during glial-guided neuronal migration. J Neurosci 30, 8529-8540.

Xia, K., Guo, H., Hu, Z., Xun, G., Zuo, L., Peng, Y., Wang, K., He, Y., Xiong, Z., Sun, L., et al. (2013). Common genetic variants on 1p13.2 associate with risk of autism. Mol Psychiatry.

Xu, B., Ionita-Laza, I., Roos, J.L., Boone, B., Woodrick, S., Sun, Y., Levy, S., Gogos, J.A., and Karayiorgou, M. (2012). De novo gene mutations highlight patterns of genetic and neural complexity in schizophrenia. Nat Genet 44, 1365-1369.

Xu, B., Roos, J.L., Dexheimer, P., Boone, B., Plummer, B., Levy, S., Gogos, J.A., and Karayiorgou, M. (2011). Exome sequencing supports a de novo mutational paradigm for schizophrenia. Nat Genet 43, 864-868.

Xu, B., Roos, J.L., Levy, S., van Rensburg, E.J., Gogos, J.A., and Karayiorgou, M. (2008). Strong association of de novo copy number mutations with sporadic schizophrenia. Nat Genet 40, 880-885.

Xu, B., Woodroffe, A., Rodriguez-Murillo, L., Roos, J.L., van Rensburg, E.J., Abecasis, G.R., Gogos, J.A., and Karayiorgou, M. (2009). Elucidating the genetic architecture of familial schizophrenia using rare copy number variant and linkage scans. Proc Natl Acad Sci USA 106, 16746-16751.

Xu, J., Zwaigenbaum, L., Szatmari, P., and Scherer, S.W. (2004). Molecular Cytogenetics of Autism. Curr Genomics 5, 347-364.

Yang, L., Neale, B.M., Liu, L., Lee, S.H., Wray, N.R., Ji, N., Li, H., Qian, Q., Wang, D., Li, J., et al. (2013). Polygenic transmission and complex neuro developmental network for attention deficit hyperactivity disorder: genome-wide association study of both common and rare variants. Am J Med Genet B Neuropsychiatr Genet 162B, 419-430.

Yoshida, Y., and Uchiyama, T. (2004). The clinical necessity for assessing Attention Deficit/Hyperactivity Disorder (AD/HD) symptoms in children with high-functioning Pervasive Developmental Disorder (PDD). Eur Child Adolesc Psychiatry 13, 307-314.

Page 202: Copy Number Variation in Neurodevelopmental Disorders

189

Youle, R.J., and Strasser, A. (2008). The BCL-2 protein family: opposing activities that mediate cell death. Nat Rev Mol Cell Biol 9, 47-59.

Yu, H.E., Hawash, K., Picker, J., Stoler, J., Urion, D., Wu, B.L., and Shen, Y. (2012). A recurrent 1.71 Mb genomic imbalance at 2q13 increases the risk of developmental delay and dysmorphism. Clin Genet 81, 257-264.

Zahir, F.R., Baross, A., Delaney, A.D., Eydoux, P., Fernandes, N.D., Pugh, T., Marra, M.A., and Friedman, J.M. (2008). A patient with vertebral, cognitive and behavioural abnormalities and a de novo deletion of NRXN1alpha. J Med Genet 45, 239-243.

Zhang, D., Cheng, L., Qian, Y., Alliey-Rodriguez, N., Kelsoe, J.R., Greenwood, T., Nievergelt, C., Barrett, T.B., McKinney, R., Schork, N., et al. (2009). Singleton deletions throughout the genome increase risk of bipolar disorder. Mol Psychiatry 14, 376-380.

Zhang, X., Snijders, A., Segraves, R., Niebuhr, A., Albertson, D., Yang, H., Gray, J., Niebuhr, E., Bolund, L., and Pinkel, D. (2005). High-resolution mapping of genotype-phenotype relationships in cri du chat syndrome using array comparative genomic hybridization. Am J Hum Genet 76, 312-326.

Zheng, C., Heintz, N., and Hatten, M.E. (1996). CNS gene encoding astrotactin, which supports neuronal migration along glial fibers. Science 272, 417-419.

Zuddas, A., Zanni, R., and Usala, T. (2011). Second generation antipsychotics (SGAs) for non-psychotic disorders in children and adolescents: a review of the randomized controlled studies. Eur Neuropsychopharmacol 21, 600-620.

Zwaigenbaum, L., Bryson, S., Rogers, T., Roberts, W., Brian, J., and Szatmari, P. (2005). Behavioral manifestations of autism in the first year of life. Int J Dev Neurosci 23, 143-152.

Zweier, C., de Jong, E.K., Zweier, M., Orrico, A., Ousager, L.B., Collins, A.L., Bijlsma, E.K., Oortveld, M.A., Ekici, A.B., Reis, A., et al. (2009). CNTNAP2 and NRXN1 are mutated in autosomal-recessive Pitt-Hopkins-like mental retardation and determine the level of a common synaptic protein in Drosophila. Am J Hum Genet 85, 655-666.


Recommended