GENETICS OF OBSESSIVE-COMPULSIVE DISORDER: FROM PHENOTYPES TO PHARMACOGENETICS

by

Gwyneth Ching Mung Zai

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Institute of Medical Science University of Toronto

© Copyright by Gwyneth Ching Mung Zai 2016

Thesis Title: Genetics of Obsessive-Compulsive Disorder: From Phenotypes to Pharmacogenetics

Name: Gwyneth Ching Mung Zai

Degree: Doctor of Philosophy

Department: Institute of Medical Science

Institution: University of Toronto

Year of Convocation: 2016

Overall Abstract

Background: Obsessive-compulsive disorder (OCD) is a debilitating neuropsychiatric disorder that is characterized by a diverse clinical presentation. Evidence suggests genetic involvement in the etiology of OCD; however, genetic association studies have yielded mixed results partly due to clinical heterogeneity.

Aims: We therefore investigated the genetics of OCD subphenotypes: age at onset (AAO), Yale-

Brown Obsessive-Compulsive Scale (Y-BOCS) severity score and symptom dimensions, family history of obsessive-compulsive and related disorders (OCRDs), psychiatric comorbidities, and drug response. We first examined these subphenotypes to ascertain clinically homogeneous dimensions of OCD. We then analyzed these subphenotypes and response for genetic association to identify marker(s) for each subphenotype.

Methods: The sample consists of 560 OCD individuals. For the subphenotypic analyses, admixture analysis (STATA) was performed to analyze AAO and factor analysis (SPSS) was applied to reduce the Y-BOCS symptom checklist. Family history and psychiatric comorbidity

ii were obtained using the modified Family History Index and SCID-IV interviews respectively.

The candidate gene study investigated markers, mostly in the remote regulatory regions, of 17 candidates for association with subphenotypes in 497 OCD participants. Genome-wide association study was conducted in a subset of this sample to examine AAO and Y-BOCS severity. Retrospective antidepressant response data was available in 222 OCD individuals.

Statistical analyses were performed using SPSS, PLINK, and R programs, comparing genotype frequencies between different subphenotype groups.

Results: We identified early (≤8), intermediate (9-17), and late onset (≥18 years) OCD groups, and a 5- or 6-factor model by weighing Y-BOCS symptoms. Our sample revealed the following clinical differences: females had a higher percentage of comorbid lifetime OCRDs and the early onset group was associated with greater symmetry/order and contamination/cleaning symptoms.

However, our genetic studies revealed no significant findings across all OCD subphenotypes and antidepressant response after correction for multiple comparisons.

Conclusions: Interesting results, although negative, suggest the role of and glutamatergic system genes in addition to a newly identified candidate in influencing OCD subphenotypes and antidepressant response. Although replication studies are warranted, these results contribute to the future development of genetic tests to assess for OCD risk and antidepressant response, leading to better outcome for OCD patients.

Word Count: 350

iii

Acknowledgments

I would like to give my special and sincere thanks to my family, especially my parents, Mr. Henry and Mrs. Elizabeth Zai, my brothers, Dr. Clement Zai and Mr. Felix Zai, and my fiancé, Mr. Sunny Chan, my friends and colleagues from medical school and residency, my clinical and research supervisors and mentors for their tremendous help and support throughout my Ph.D. study. I dedicate this Ph.D. thesis to them.

Regarding research, I would like to especially thank my supervisors and mentors, Dr. James L. Kennedy and Dr. Margaret (Peggy) A. Richter, for their guidance, teaching, and insightful comments on my projects as well as advice on my future career path. I would like to express my gratitude to my international mentors (Dr. Katharine Phillips, Dr. Matthew Rudorfer, and Dr. Mark Vawter) and research supervisors (Dr. David Pauls, Professor Barbara Sahakian, and Professor Trevor W. Robbins) for their time, guidance, and support. I would also like to thank the members of my Ph.D. advisory committee, Dr. Joanne Knight and Dr. Paul Arnold, final PAC external advisor, Dr. Paul Sandor, and other members of my Ph.D. examination committee, Dr. Mary Seeman, and Dr. Daniel Geller. For the peer-reviewed process of the original article chapters of this thesis, I would like to thank Dr. Blair Simpson, Dr. Don Black, Dr. Lea Davis, and Dr. Dan Rujescu for their time and constructive feedback. Within the Institute of Medical Science, I would like to acknowledge the director, graduate coordinators, and administrative staff for their hard work and support. For the research team, many thanks to all members of the Psychiatric Neurogenetics Laboratory (Dr. Kennedy’s lab) at the Centre for Addiction and Mental Health (lab manager – Ms. Natalie Freeman; IMPACT study manager – Ms. Nicole Braganza; IMPACT study research analyst – Mr. Sheraz Cheema; lab technicians – Mr. Sajid Shaikh, Mr. David Sibony, and Ms. Maria Tampakeras; visiting postdoctoral fellows – Dr. Vanessa Gonçalves and Dr. Arun Tiwari; colleagues – Dr. Vincenzo de Luca and Dr. Daniel Mueller), the CAMH scientific computing team (manager – Mr. David Rotenberg), and the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre (research assistant – Ms. Marissa Williams; database manager – Ms. Karen Wigg; therapist – Ms. Eliza Burroughs; administrative staff – Ms. Carmen Costa and Ms. Amanda Calzolaio), for making my doctoral experience stimulating and exciting. I would further like to acknowledge iv my international collaborators (Brazil: Dr. Roseli Shavitt, Dr. Euripedes Miguel, Dr. Carolina Cappi, and Dr. Maria Conceição do Rosario; USA: Dr. Carol Mathews, Dr. Gregory Hanna, Dr. Thomas Fernandez, Dr. Lea Davis, Dr. James Knowles, and Dr. Dongmei Yu; Spain: Dr. Pino Alonso and Dr. Xavier Estivill; Netherlands: Dr. Danielle Cath and Dr. Eske Derks; Italy: Dr. Cristina Cavallini; Switzerland: Dr. Edna Grünblatt and Dr. Susanne Walitza; South Africa: Dr. Dan Stein and Dr. Christine Lochner; Mexico: Dr. Humberto Nicolini; Canada: Dr. Evelyn Stewart; UK: Professor Naomi Fineberg, Dr. Sam Chamberlain, Dr. Becky Inkster, Dr. Annette Brühl, Dr. Annemieke Apergis-Schoute, Dr. Sharon Morein-Zamir, and Ms. Matilde Vaghi) and their administrative staff for their contributions in my career development and Ph.D. thesis. Finally, I would like to give my sincere thanks and gratitude to all research participants for their time and effort.

Regarding funding, I would like to acknowledge the Ontario Ministry of Health and Long-Term Care (PGY-5 residency salary from 2012-2014), the Canadian Institutes of Health Research (CIHR) Postdoctoral Fellowship (salary from 2014-2016), the W. Garfield Weston Doctoral Fellowship (fellowship travel expense and salary from 2014-2015), the Younger Foundation (conference allowance in 2014) and the Frederick W. Thompson Anxiety Disorders Centre (travel allowance in 2014-2015) for supporting my salary, research allowance, and/or conference expenses. Research funding was supported by a grant from the Ontario Mental Health Foundation (Drs. Richter and Kennedy) and a private donation from the Frederick W. Thompson family (Dr. Richter).

v

Contributions

Chapter 1: Dr. Gwyneth Zai conducted a thorough literature search on the genetics of obsessive- compulsive disorder (OCD) phenotypes and pharmacogenetics, and she completed the entire chapter. A large part of this chapter was taken from a recent review on the pharmacogenetics of OCD, for which Dr. Gwyneth Zai is the first author, and she performed a thorough literature search and review, wrote, submitted, and revised the manuscript for publication. This review article has been published in the journal of Pharmacogenomics 2014; 15(8):1147-1157.

Chapter 2: Dr. Gwyneth Zai designed the aims, objectives, and hypothese (with guidance from faculty) for this PhD thesis.

Chapter 3: Dr. Gwyneth Zai designed the experiment (with guidance from faculty), performed data collection, corresponded with co-authors, performed all statistical analyses, wrote and revised the manuscript, which is ready for publication.

Chapter 4: Dr. Gwyneth Zai designed the experiment (with guidance from faculty), performed part of the data collection and all genotyping, corresponded with co-authors, performed all the statistical analyses, wrote and revised the manuscript, which is ready for publication after the completion of replication study.

Chapter 5: Dr. Gwyneth Zai designed the experiment (with guidance from faculty), performed part of the data collection, corresponded with co-authors, performed all the statistical analyses, wrote and revised the manuscript, which is ready for publication after the completion of replication study.

Chapter 6: Dr. Gwyneth Zai designed the experiment (with guidance from faculty), performed data collection and most of the genotyping, corresponded with co-authors, performed all the statistical analyses, wrote and revised the manuscript, which is ready for publication after replication study.

Chapter 7: Dr. Gwyneth Zai wrote the entire final chapter, summarizing the results, and providing limitations and future directions to extend the findings from the PhD thesis.

vi

Table of Contents

Acknowledgments ...... iv

Contributions ...... vi

Table of Contents ...... vii

List of Abbreviations ...... xii

List of Tables ...... xix

List of Figures ...... xxii

List of Appendices ...... xxv

Chapter 1 GENERAL INTRODUCTION: LITERATURE REVIEW ...... 1

1 Obsessive-Compulsive Disorder (OCD) ...... 2

1.1 Historical Development and Treatment of Clinical OCD ...... 5

1.2 Treatment of OCD ...... 5

1.3 Gender Differences in OCD ...... 8

1.4 Genetics of OCD ...... 9

1.4.1 Genetics of Gender Differences in OCD ...... 15

1.4.2 Genetics of OCD Subphenotypes ...... 17

1.4.3 Pharmacogenetics of OCD ...... 47

1.5 Genes of Interest ...... 70

1.5.1 Significance of Remote Regulatory Regions ...... 85

1.5.2 Genetic Methodology ...... 87

Chapter 2 AIMS, OBJECTIVES, HYPOTHESES, AND IMPLICATIONS ...... 89

2 Thesis Aims, Hypotheses, and Implications ...... 89

2.1 Aims and Objectives ...... 89

2.1.1 Aims ...... 90

2.1.2 Objectives ...... 90

vii

2.2 Hypotheses ...... 91

2.2.1 Clinical Phenomenology of OCD ...... 91

2.2.2 Genetics of OCD Phenotypes ...... 91

2.2.3 Genetics of SRI Response in OCD ...... 92

2.3 Originality ...... 92

2.4 Implications ...... 92

Chapter 3 CLINICAL PHENOMENOLOGY OF OCD ...... 95

3 An Update on the Clinical Phenomenology of Obsessive-Compulsive Disorder ...... 95

3.1 Abstract ...... 96

3.2 Introduction ...... 97

3.3 Methods ...... 101

3.3.1 Diagnostic Criteria and Clinical Sample ...... 101

3.3.2 Familiality Data to Determine the Presence of Absence of Family History of OCRDs ...... 103

3.3.3 Antidepressant Response Data Description ...... 104

3.3.4 Data Analysis ...... 105

3.4 Results ...... 109

3.4.1 Does Gender Contribute to Clinical Diversity in OCD? ...... 109

3.4.2 Does AAO Present with Multiple Normal Distributions? ...... 111

3.4.3 Y-BOCS Symptom Dimension – Factor Model and Clusters ...... 111

3.4.4 Rate of Psychiatric Comorbidities ...... 112

3.4.5 Rate of Significant Family History of OCRDs ...... 112

3.4.6 Can Clinical Characteristics Predict Antidepressant Response? ...... 113

3.5 Discussion ...... 128

3.5.1 Gender Does Not Contribute to Difference in Clinical Presentations ...... 128

3.5.2 AAO has a Tri-modal Distributions ...... 129

viii

3.5.3 Factor and Cluster Analyses Confirmed a 5- and 6-Factor Model of Y-BOCS Symptom Dimensions ...... 130

3.5.4 High Rates of Psychiatric Comorbidities ...... 132

3.5.5 High Rate of Family History of OCRDs ...... 132

3.5.6 Antidepressant Response Rate is Higher than Published Rates and Differences were Observed between SRI Responders and Non-Responders ...... 133

3.5.7 Limitations of the Clinical Study ...... 135

3.5.8 Conclusion ...... 136

Chapter 4 CANDIDATE GENE STUDY OF OCD PHENOTYPES ...... 137

4 Candidate Gene Study of Obsessive-Compulsive Disorder Phenotypes ...... 137

4.1 Abstract ...... 138

4.2 Introduction ...... 138

4.3 Methods ...... 149

4.3.1 Diagnostic Criteria and Sample for Candidate Gene Analysis ...... 149

4.3.2 Choosing of SNPs of Candidate Genes and Genotyping ...... 153

4.3.3 Quality Control and Statistical Analyses ...... 154

4.4 Results ...... 155

4.5 Discussion ...... 179

Chapter 5 GWAS OF OCD PHENOTYPES ...... 183

5 Genome-Wide Association Study of Obsessive-Compulsive Disorder Phenotypes ...... 183

5.1 Abstract ...... 184

5.2 Introduction ...... 184

5.3 Methods ...... 187

5.3.1 Diagnostic Criteria and Sample ...... 187

5.3.2 Genotyping and Statistical Analyses ...... 187

5.4 Results ...... 189

5.5 Discussion ...... 209 ix

Chapter 6 PHARMACOGENETICS OF OCD ...... 212

6 Multi-Gene Pharmacogenetic Study of Antidepresant Response in Obsessive-Compulsive Disorder ...... 212

6.1 Abstract ...... 213

6.2 Introduction ...... 213

6.3 Methods ...... 217

6.3.1 Diagnostic Criteria and Sample ...... 217

6.3.2 Response Data ...... 218

6.3.3 Choosing of SNPs and Genotyping ...... 219

6.3.4 Statistical Analyses ...... 221

6.4 Results ...... 221

6.5 Discussion ...... 232

Chapter 7 GENERAL DISCUSSION ...... 235

7 General Discussion ...... 235

7.1 Summary of Findings and Implications ...... 236

7.1.1 Clinical OCD (Chapter 3) ...... 237

7.1.2 Genetics of OCD Phenotypes (Chapter 4 and Chapter 5) ...... 238

7.1.3 Pharmacogenetics of OCD (Chapter 6) ...... 240

7.2 Limitations and Considerations ...... 243

7.2.1 Sample Size and Genomic Coverage ...... 243

7.2.2 Corrections for Multiple Comparisons ...... 245

7.2.3 Diagnostic Uncertainty ...... 247

7.2.4 Use of Factor Analysis to Identify OCD Subtypes or Symptom Dimensions .... 248

7.3 Future Directions ...... 248

7.3.1 Trait and Subphenotype Analysis ...... 249

7.3.2 Future Genetic Studies ...... 250

x

7.4 Concluding Remarks ...... 256

References ...... 257

Appendices ...... 334

Appendix I ...... 335

Appendix II ...... 336

Appendix III ...... 349

Appendix IV ...... 350

Copyright Acknowledgements ...... 374

xi

List of Abbreviations

χ2 Chi-squared

5-HT Serotonin

AAO Age at onset

ABLIM1 Actin-binding LIM family, member 1 gene

ACE Angiotensin-converting enzyme gene

Akt Protein kinase B

AKT1 V-Akt murine thymoma viral oncogene homolog 1 gene

ANKFN1 Ankyrin-repeat and fibronectin type III domain containing 1 gene

ARRDC4 Arrestin domain containing 4

BDD Body dysmorphic disorder

BDKRB2 Bradykinin receptor B2 gene

BDNF Brain-derived neurotrophic factor gene

BTBD3 BTB (POZ) domain containing 3 gene

BTBD9 BTB (POZ) domain containing 9 gene

C9orf68 Chromosome 9 open reading frame 68 gene

CBT Cognitive behavioural therapy

CDH9 Cadherin 9, type 2 (T1-cadherin) gene

CGI-I Clinical Global Impression – Improvement

ChIP Chromatin immuno-precipitation

CI Confidence interval

CIHR Canadian Institutes of Health Research

CNV Copy-number variation

COMT Catechol-O-methyltransferase gene

CRHR1 Corticotropin releasing hormone receptor 1

xii

CSTC Cortico-striato-thalamo-cortical

CYP Cytochrome P450 genes

DAT1 Dopamine transporter gene, also known as SLC6A3

DDR1 Discoidin domain receptor tyrosine kinase 1 gene

DHS DNase hypersensitive site

DISP1 Dispatched homolog 1 (drosophila) gene

DLGAP1 Discs, large (drosophila) homolog-associated protein 1 gene

DLGAP2 Discs, large (drosophila) homolog-associated protein 2 gene

DLGAP3 Discs, large (drosophila) homolog-associated protein 3 gene

DNA Deoxyribonucleic acid

DNase Deoxyribonuclease

DRD1 Dopamine D1 receptor gene

DRD2 Dopamine D2 receptor gene

DRD3 Dopamine D3 receptor gene

DRD4 Dopamine D4 receptor gene

DRD5 Dopamine D5 receptor gene

DSM-5 Diagnostic and Statistical Manual of Mental Disorders, 5th Edition

DSM-IV Diagnostic and Statistical Manual of Mental Disorders, 4th Edition

EAAC1 Excitatory amino acid carrier 1

EFNA5 Ephrin-A5 gene

ERBB4 Erb-b2 receptor tyrosine kinase 4

EurMAP European minor allele frequencies

ENCODE ENCyclopedia Of DNA Elements

ERBB4 V-erb-b2 avian erythroblastic leukemia viral oncogene homolog 4 gene

ERE6 Estrogen response element 6 gene

ERK Extracellular signal-regulated kinase xiii

ESRα Estrogen receptor alpha gene

EurMAF European minor allele frequencies

FAIM2 Fas apoptotic inhibitory molecule 2 gene

FBAT Family-based association test

FHI Family History Interview

FKBP5 FK506 binding protein 5 gene

FMM Finite Mixture Models

FS Functional significance

FUT2 Fucosyltransferase 2 gene

GABHS Group A β-hemolytic streptococcal infection

GAD1 Glutamate decarboxylase 1 gene

GAD2 Glutamate decarboxylase 2 gene

GENDEP Genome-based Therapeutic Drugs for Depression study

GI Gastrointestinal

GNB3 Guanine nucleotide binding protein (G protein), beta polypeptide 3 gene

GPC6 Glypican 6 gene

GPCR G-protein-coupled receptor

GRIA1 Glutamate receptor, ionotropic, AMPA 1 gene

GRIA3 Glutamate receptor, ionotropic, AMPA 3 gene

GRIK2 Glutamate receptor, ionotropic, kainate 2 gene

GRIK4 Glutamate receptor, ionotropic, kainate 4 gene

GRIN2B Glutamate receptor, ionotropic, N-methyl D-aspartate 2B gene

GSK3β Glycogen synthase kinase 3 beta gene

GWAS Genome-wide association study

HD Hoarding disorder

HLA Human leukocyte xiv

HOXB8 Homeobox B8 gene

HTR1A Serotonin 1A receptor gene

HTR1B Serotonin 1D beta receptor gene

HTR2A Serotonin 2A receptor gene

HTR2C Serotonin 2C receptor gene

HTR6 Serotonin 6 receptor gene

HTTLPR Serotonin transporter-linked polymorphic region

ICD-10 International Classification of Diseases, 10th Edition

IER3 Immediate early response 3 gene

IL1B Interleukin 1, beta gene

INPP-1 -polyphosphatase-1 gene

Ins/del Insertion/deletion polymorphism

IOCDFGC International OCD Foundation Genetic Collaborative Group

KCNN3 Potassium intermediate/small conductance calcium-activated channel, subfamily N, member 3 gene

LD Linkage disequilibrium

LMX1A LIM homeobox transcription factor 1, alpha gene

MAOA Monoamine oxidase A gene

MAPK Mitogen-activated protein kinase

MARS Munich Antidepressant Response Signature project

MDD Major depressive disorder

MDS Multi-dimensional scaling

MOG Myelin oligodendrocyte glycoprotein gene

MPDZ Multiple PDZ domain protein gene

NaSSA Noradrenergic and specific serotonergic antidepressant

NF-κB Nuclear factor kappa-light-chain-enhancer of activated B cells

xv

NGFR Nerve growth factor receptor gene

NIEHS National Institute of Environmental Health Sciences

NIMH National Institute of Mental Health

NMDA N-methyl-D-aspartate

NPSR1 Neuropeptide S receptor 1 gene

NRG1 Neuregulin 1 gene

NRG3 Neuregulin 3

NSAIDs Non-steroidal anti-inflammatory drugs

NTM Neurotrimin gene

NTRK2 Neurotrophic tyrosine kinase, receptor, type 2 gene

NTRK3 Neurotrophic tyrosine kinase, receptor, type 3 gene

OCRDs Obsessive-compulsive and related disorders

OCD Obsessive-compulsive disorder

OLIG1 Oligodendrocyte lineage transcription factor 1 gene

OLIG2 Oligodendrocyte lineage transcription factor 2 gene

OR Odds ratio

PANDAS Pediatric Autoimmune Neuropsychiatric Disorders Associated with (group A β-hemolytic) Streptococcal infection

PANS Pediatric Acute-onset Neuropsychiatric Syndrome

PCA Principal component analysis

PCDH10 Protocadherin 10 gene

PCDH1D Protocadherin-1D gene

PGC Psychiatric Genomics Consortium

PI3K Phosphoinositide 3 kinase

PKC Protein kinase C

PLC-γ1 Phospholipase C gamma-1 gene

xvi

PLCB1 Phospholipase C, beta 1 (phosphoinositide-specific)

PTPN21 Protein-tyrosine phosphatase, nonreceptor-type, 21 gene

PTPRD Protein-tyrosine phosphatase, receptor-type, delta

RNA Ribonucleic acid

SAPAP3 SAP90/PSD95-associated protein 3 gene

SCID Structured Clinical Interview for the DSM-IV

SD Standard deviation

SIADH Syndrome of inappropriate antidiuretic hormone

SLC1A1 Glutamate transporter, neuronal, gene

SLC18A1 Vesicular monoamine transporter

SLC6A2 Norepinephrine transporter gene

SLC6A4 Serotonin transporter gene, also known as 5HTT

SLITRK5 SLIT and NRTK-like family, member 5 gene

SNRI Serotoin-norepinephrine reuptake inhibitor

SNP Single nucleotide polymorphism

SPD Skin picking disorder

SRI Serotonin reuptake inhibitor

SSRI Selective serotonin reuptake inhibitor

STAR*D Sequenced Treatment Alternatives to Relieve Depression study

TCA

TDT Transmission disequilibrium test

TF Transcription factor

TFBS Transcription factor binding site

TH Tyrosine hydroxylase

TIAM1 T-cell lymphoma invasion and metastasis 1 gene

TNFA Tumor necrosis factor alpha gene xvii

TPH1 Tryptophan hydroxylase 1 gene

TPH2 Tryptophan hydroxylase 2 gene

TSS Transcription start site

TTM Trichotillomania

USCS University of California, Santa Cruz genome browser

VNTR Variable number tandem repeat

WHO World Health Organization

Y-BOCS Yale-Brown Obsessive-Compulsive Scale

xviii

List of Tables

Table 1.1. Common Symptoms of OCD...... 4

Table 1.2. Pharmacological Treatment of OCD...... 7

Table 1.3. Treatment Response Categories in OCD...... 7

Table 1.4. Summary Table of Genetic Studies of Gender in OCD...... 19

Table 1.5. Summary Table of Genetic Studies of AAO in OCD...... 25

Table 1.6. Summary Table of 5-Factor Y-BOCS Symptom Dimensions in OCD...... 32

Table 1.7. Summary Table of 4-Factor Y-BOCS Symptom Dimensions in OCD...... 35

Table 1.8. Summary Table of 3-, 6-, and Mixed-Factor Y-BOCS Symptom Dimensions in OCD...... 37

Table 1.9. Summary Table of Genetic Studies of Y-BOCS Symptom Dimensions in OCD...... 39

Table 1.10. Summary Table of Genetic Studies of Y-BOCS Symptom Severity in OCD...... 41

Table 1.11. Psychiatric Comorbidity of OCD (Percentage of total number of individuals reported with OCD or in the general population)...... 44

Table 1.12. Genetic Studies of Psychiatric Comorbidity in OCD...... 45

Table 1.13. Summary of Pharmacogenetic Studies in OCD...... 49

Table 1.14. Summary of Candidate Gene Studies of OCD...... 71

Table 3.1. Subject Demographics...... 109

Table 3.2. Comparison of Clinical Characteristics between Males and Females...... 110

Table 3.3. Comparison of Clinical Characteristics between Age At Onset (AAO) Groups...... 116

Table 3.4. Y-BOCS Symptom Dimension Factor Analysis (weighing all symptoms as equal). 118

xix

Table 3.5. Y-BOCS Symptom Dimension Factor Analysis (double weighing target symptoms)...... 119

Table 3.6. Comparison of Clinical Characteristics between Familial and Non-Familial (OCD) OCD Individuals...... 125

Table 3.7. Comparison of Clinical Characteristics between Familial and Non-Familial (OCRDs) OCD Individuals...... 126

Table 3.8. Comparison of Clinical Characteristics between SRI Responders and Non- Responders...... 127

Table 4.1. Chosen Genes...... 144

Table 4.2. Chosen SNPs...... 146

Table 4.3. Subject Demographics...... 155

Table 4.4. Subject Demographics (100% Caucasian)...... 156

Table 4.5. Investigated SNPs...... 156

Table 4.6. Genetic Results of Gender Differences in OCD...... 162

Table 4.7. Genetic Results of Age At Onset (AAO) Groups in OCD...... 164

Table 4.8. Genetic Results of Linear Regression of Age At Onset (AAO) in OCD...... 165

Table 4.9. Genetic Results of Linear Regression of Y-BOCS Total Severity Scores in OCD. .. 167

Table 4.10. Genetics Results of Linear Regression of 5-Factor Y-BOCS Symptom Dimensions in OCD...... 169

Table 4.11. Genetic Results of Linear Regression of 6-Factor Y-BOCS Symptom Dimensions in OCD...... 171

Table 4.12. Genetic Results of Psychiatric Comorbidities in OCD...... 173

Table 4.13. Genetic Results of Family History in OCD...... 176 xx

Table 4.14. Genetic Results of Family History (combined OCRDs) in OCD...... 178

Table 5.1. Genome-Wide Association Study (GWAS) Quality Control...... 189

Table 5.2. Demographic Data of the Entire OCD Sample...... 189

Table 5.3. Top Hits of Age At Onset (AAO) in OCD...... 194

Table 5.4. Top Hits of Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score in OCD...... 195

Table 5.5. Population Description for the MDS Plots...... 198

Table 6.1. Brief Summary of Pharmacogenetic Studies of OCD...... 215

Table 6.2. Investigated Genes with Their SNPs for the Single Taqman Assay Protocol...... 216

Table 6.3. Genes and SNPs for the QuantStudio Array Protocol...... 217

Table 6.4. Demographic Data of the Entire OCD Sample...... 222

Table 6.5. Subject Demographics for the Single Taqman Assay Protocol...... 222

Table 6.6. Final Demographic Data for the QuantStudio OCD Pharmacogenetic Sample...... 223

Table 6.7. Pharmacogenetic Results for the Single Taqman Assay Protocol...... 225

Table 6.8. Result of Genetic Variations across Remote Regulatory Regions in 14 OCD Candidate Genes and SRI Response...... 229

xxi

List of Figures

Figure 1.1. Genes Associated with OCD...... 10

Figure 1.2. Annotation of Disease-Associated Variants from ENCyclopedia Of DNA Elements (ENCODE) Data (Kavanagh et al., 2013)...... 88

Figure 2.1. Thesis Hypothesis...... 94

Figure 3.1. Normal Distributions of Age At Onset (AAO)...... 115

Figure 3.2. Y-BOCS Symptom Dimension Cluster Analysis (weighing all symptoms as equal)...... 120

Figure 3.3. Y-BOCS Symptom Dimension Cluster Analysis (double weighing target symptoms)...... 121

Figure 3.4. Current Psychiatric Comorbidities of OCD Sample. Percentage is indicated in brackets...... 122

Figure 3.5. Lifetime Comorbid OCRDs of OCD Sample. Percentage is indicated in brackets. 123

Figure 3.6. Family History of OCD Sample ...... 124

Figure 4.1. Age At Onset (AAO) Distribution with Normal Curve...... 159

Figure 4.2. Log10 Transformed Age At Onset (AAO) Distribution with Normal Curve...... 160

Figure 4.3. Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Distribution with Normal Curve...... 161

Figure 4.4. Linear Regression of Age At Onset (AAO) in OCD...... 166

Figure 4.5. Linear Regression of Y-BOCS Total Severity Scores in OCD...... 168

Figure 4.6. Linear Regression of 5-Factor Y-BOCS Symptom Dimensions in OCD...... 170

Figure 4.7. Linear Regression of 6-Factor Y-BOCS Symptom Dimensions in OCD...... 172

xxii

Figure 5.1. Frequency Distribution of Age At Onset (AAO)...... 190

Figure 5.2. Frequency Distribution of Log10 Transformation of Age At Onset (AAO)...... 191

Figure 5.3. Frequency Distribution of Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score...... 192

Figure 5.4. Quantile-Quantile Plot for Age At Onset (AAO) – Left (Initial QQ Plot) and Right (QQ Plot After Permutation). The genomic inflation value (lambda) is 1.003295...... 196

Figure 5.5. Quantile-Quantile Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score – Left (Initial QQ Plot) and Right (QQ Plot After Permutation). The genomic inflation value (lambda) is 1.048644...... 197

Figure 5.6. Multi-Dimensional Scale (MDS) Plot – Initial...... 199

Figure 5.7. Multi-Dimensional Scale (MDS) Plot – Removal 1...... 200

Figure 5.8. Multi-Dimensional Scale (MDS) Plot – Removal 2...... 201

Figure 5.9. Multi-Dimensional Scale (MDS) Plot – Final...... 202

Figure 5.10. Manhattan Plot for Age At Onset (AAO) – Left (Initial Manhattan Plot) and Right (After Permutation)...... 203

Figure 5.11. Manhattan Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score – Left (Initial Manhattan Plot) and Right (After Permutation)...... 204

Figure 5.12. LocusZoom Plot for Age At Onset (AAO) Result on Chromosome 6...... 205

Figure 5.13. LocusZoom Plot for Age At Onset (AAO) Result on Chromosome 10...... 206

Figure 5.14. LocusZoom Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Result on Chromosome 14...... 207

Figure 5.15. LocusZoom Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Result on Chromosome 9...... 208

xxiii

Figure 6.1. Result of Genetic Variations across Remote Regulatory Regions in 14 OCD Candidate Genes and Antidepressant Response with only Cacausians. Red box consists of SNP within the gene instead of remote regulatory regions...... 231

Figure 7.1. Summary of Genetic Results in OCD Subphenotypes...... 242

xxiv

List of Appendices

Appendix I. Yale-Brown Obsessive Compulsive Scale (Y-BOCS) Symptom Checklist and Severity Score…………………………………………………………………………. 334

Appendix II. Family History Interview (FHI)…………………………………………….. 335

Appendix III. Pharmacogenetics Questionnaire…………………………………………... 348

Appendix IV. Supplementary Tables……………………………………………………....349

xxv 1

Chapter 1 GENERAL INTRODUCTION: LITERATURE REVIEW

Psychiatric disorders are complex, often chronic, and are multi-factorial, with a range of genetic and environmental contributions. Efforts in delineating the genetic etiology of psychiatric disorders have been hampered by the complex genetic transmission patterns, multiple biological mechanisms, gene-gene and gene-environment interactions, and phenotypic heterogeneity challenges that are commonly faced by psychiatric geneticists. Thus, new research paradigms incorporating the use of high-throughput genetic technologies, computational biostatistical methodologies, response and tolerability to different treatment modalities, and brain imaging techniques, have been adopted in an attempt to elucidate the etiologic mechanisms in psychiatric disorders.

Genetics comprises the study of inheritance patterns of variations across genes in living organisms. Genes are regions of deoxyribonucleic acid (DNA) within the human genome that code for ribonucleic acid (RNA) and protein products. Variations across genes are what make each one of us unique in terms of, for example, eye and hair colours, size, shape, and behaviours. The following major milestones in molecular genetics mark the expansion in our understanding of many medical and psychiatric disorders. The concept of heredity came from the ancient Greek physician, Hippocrates, who theorized that “seeds” from different parts of the body were transmitted to newly conceived embryos, also known as pangenesis, which was later supported by Charles Darwin in 1868 (Holterhoff, 2014). The key ingredient of DNA was isolated in 1869 by a Swiss physician, Dr. Johannes Friedrich Miescher (Bernhard, 1970). The first genetic laws, Mendelian laws of inheritance, were reported by Dr. Gregor Mendel in 1866 after he spent a large amount of time crossing pea plants (Castle, 1903). An early geneticist, Dr. Walter Sutton, described and termed the chromosome in 1902 (Crow and Crow, 2002). The link between chromosomes and heredity was confirmed in 1941, revealing that genes code for , by two geneticists, Dr. Edward Tatum and Dr. George Beadle (Raju, 1999; Satzinger, 2008). The discovery of the double-helix structure of DNA in 1953 by Dr. James Watson and Dr. Francis Crick marks the beginning of molecular genetics (Watson and Crick, 1953). With advancing

1

2

molecular technologies, the construction of the complete human genetic library, the Human Genome Project, became a reality in 2001 (Lander et al., 2001). The project identified approximately 23,000 protein-coding genes, which only account for 1.5% of the entire genome, while the remaining so-called “junk DNA” did not appear to have any known functional significance. It was not until 2012 that the ENCyclopedia Of DNA Elements (ENCODE) project announced the importance of these “junk DNA” as having regulatory influence on gene expression (ENCODE Project Consortium, 2012). Thus, the field of molecular genetics has grown exponentially, and large-scale collaborative projects such as the Human Genome Project and ENCODE have provided powerful tools for the investigations of complex traits.

Psychiatric genetics, the study of the role of genetics in psychiatric disorders, has mainly focused on humans, with a limited number of animal models given the complexity of behaviour presented in psychiatric conditions. The founder of psychiatric genetics, Dr. Francis Galton, began his search for hereditary factors of “mental characteristics” in the late nineteenth century (Schulze et al., 2004). With advances in technology and computational capacity, psychiatric genetics has now entered a new era of exploration of the genetic architecture of psychiatric disorders.

Based on the current understanding of psychiatric genetics, the present thesis has investigated genetic contributions that may play a role in the heritability of obsessive-compulsive disorder (OCD) with the main focus being on the identification of different molecular genetic markers that given its heterogeneous nature, may influence different phenotypic expressions of OCD.

1 Obsessive-Compulsive Disorder (OCD)

Obsessive-compulsive disorder (OCD) is a common psychiatric disorder that is characterized by chronic and debilitating symptoms of obsessions and/or compulsions (APA, 2013). Obsessions are described as intrusive, repetitive, and persistent thoughts, urges, or images that cause significant distress and great difficulty in suppression or mental dismissal. Compulsions are excessive and repetitive ritualistic behaviours or mental acts that cannot be

2

3

resisted and are perhaps performed to reduce the severe anxiety caused by the obsessions. These obsessions and/or compulsions are excessive, irrational, and ego-dystonic in nature and patients with OCD devote one hour or more per day to these symptoms, which significantly impair their daily functioning. Common examples of obsessions and their related compulsions are presented in Table 1.1 and include: fear of contamination and cleaning rituals, need for symmetry and rearrangement, aggressive thoughts about one’s safety and checking behaviours. This condition is common, affecting 1 to 3% of the general population (APA, 2000; Ruscio et al., 2010), with subclinical rates of obsessions and/or compulsions ranging from 13% to 49% (Fullana et al., 2009).

According to the World Health Organization (WHO 2001), OCD is one of the top ten leading causes of disability worldwide and it is also the fourth most frequently diagnosed psychiatric disorder. Furthermore, the National Comorbidity Study Replication epidemiological study reported OCD as the anxiety disorder with the highest percentage of “serious” classifications, which was defined using several indicators such as a 12-month suicide attempt rate, substantial work disability, or role impairment (Kessler et al., 2005). Thus, increasing public awareness of this disorder, and optimizing assessment and treatment regimens are necessary to reduce the burden of OCD globally.

Given that OCD is an internalizing disease (individuals who suffer from this type of disorder often keep or internalize their problems/issues/symptoms) with limited public awareness in society, it has not been the focus of research and/or treatment until more recently. Most patients suffering from OCD deem themselves as the only ones who exhibit OCD symptoms, thereby creating an invisible barrier to seeking psychiatric help because of shamefulness and stigma. Understanding the etiology of OCD requires increased public awareness of obsessive- compulsive symptoms, increased screening, and more active patient participation; and these are just some of the reasons why research in OCD has been lagging behind other severe and persistent mental illnesses. Greater public awareness of OCD will encourage research participation, which in turn will improve the statistical power of our research sample in identifying genetic factors of OCD. Improving the understanding of the underlying biological mechanism of OCD will help guide research in targeting treatment for this debilitating mental

3

4

illness. Discoveries from research will bring about innovations in healthcare, which lead to improved outcome for OCD patients.

OCD was grouped as one of the anxiety disorders in the DSM-IV (Diagnostic and Statistical Manual of Mental Disorders, 4th Edition) (APA, 2000), but became a new category, the “obsessive-compulsive and related disorders” (OCRDs), in the recent DSM-5 (APA, 2013). The new section for OCRDs in the DSM-5 includes OCD, body dysmorphic disorder (BDD), hoarding disorder (HD), trichotillomania (TTM) or hair-pulling disorder, and excoriation (skin- picking) disorder (SPD) (APA, 2013). These disorders share clinical phenomenology, patterns of comorbidity, genetic loading, and etiology with OCD (Monzani et al., 2014; Pallanti et al., 2014). It is worth noting that hoarding was previously considered as a subtype or symptom of OCD in the DSM-IV (APA, 2000). Given its relatively distinct clinical profile, it is now widely accepted as a separate disorder (APA, 2013).

Given its complex clinical presentations with diverse symptom characteristics, OCD is known for its heterogeneous nature (Pauls et al., 2014). Therefore, it seems important to identify different homogeneous subgroups or phenotypes of OCD for genetic analysis.

Table 1.1. Common Symptoms of OCD.

Obsession Examples Associated Compulsions

Fear of Concerns with dirt, germs, contaminants, Cleaning, washing contamination household items, animals Concerns with getting self or other ill Symmetry and Recurrent thoughts of needing to do things in a Ordering, arranging, organizing exactness certain way Checking, reassurance seeking, Aggressive Fear of harming self or others monitoring Recurrent violent images Religious Concerns with sacrilege and blasphemy Praying, asking for forgiveness Excessive concerns with morality, Somatic Excessive concerns with illness of disease Physical checking Excessive concerns with body part or aspect of appearance Excessive concerns with discarding waste or Collecting Hoarding junk Recurrent forbidden or perverse sexual Avoiding situations Sexual thoughts, images, or impulses 4

5

Pathologic doubt Recurrent worries of making mistakes or doing Checking, performing rituals and completeness something wrong Superstition- Fear of certain "bad" or "unlucky" numbers or Counting related colours

Common symptoms of OCD with specific examples of obesssions and their associated compulsions are reported in this table.

1.1 Historical Development and Treatment of Clinical OCD

OCD was once known as obsessive compulsive neurosis. Symptoms of OCD were first described as early as the 1400s addressing symptoms from superstitious notions to scrupulosity (Climacus, 1982). The first “self-help” writings came from the late 1600s by Richard Baxter and colleagues (Clifford, 1716). The first attempted treatment for OCD symptoms was described by an astrological healer named Richard Napier with zodiac signs and positioning of the symbols of the Sun, Moon, planets, etc. in the 1600s (MacDonald, 1981). Medical treatment for these symptoms known as “bad thoughts” was described in the 1700s by Ms. Hannah Allen with “bloodletting” (Allen, 1683). Other treatments in history include exorcism (Jenike et al., 1986; Berrios, 1989) as individuals suffering from “bad thoughts” were thought to be possessed by the Devil (Aardema and O’Connor, 2007). Sigmund Freud later developed the diagnostic category of obsessional neurosis in 1894 that captured symptoms of OCD (May-Tolzmann, 1998). He attributed obsessive-compulsive behaviours to unconscious conflicts that manifest as symptoms (Jenike et al., 1986; Berrios, 1989) and used psychoanalytical psychotherapy to treat OCD. A paradigm shift occurred in the 1970s when behavioural psychology (and cognitive psychology to a lesser extent) began to overshadow Freudian theory and became the dominant psychotherapeutic model for the treatment of OCD (Rachman, 2009).

1.2 Treatment of OCD

Returning to the present decade, treatment of OCD comprises two different streams, psychotherapy and pharmacotherapy. For mild to moderate severity of OCD, cognitive behavioural therapy (CBT) with exposure and response prevention strategy is the first-line

5

6

treatment (CPA, 2006). Its response rate is reported to be as high as 86% (Foa et al., 2005). However, psychotherapy requires schedule flexibility, time commitment, psychological mindedness, high motivation, ability to express feelings/affect, ability to tolerate change/anxiety, and absence of severe psychopathology (Sadock et al., 2007). Therefore, some OCD patients are not suitable for psychotherapy and some prefer pharmacotherapy. The first- and second-line pharmacological treatments of OCD are , selective serotonin reuptake inhibitors (SSRIs), serotonin-norepinephrine reuptake inhibitors (SNRIs), and tricyclic antidepressants (TCAs) (Table 1.2) (CPA, 2006; Fineberg et al., 2015; Katzman et al., 2014). The efficacy of antidepressants is lower than that of CBT, ranging between 40% and 60% with reduction of symptoms (Foa et al., 2005; McDougle et al., 1993), which is usually defined as OCD symptom reduction of at least 25% to 35% using the Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score (Tolin et al., 2005) (Table 1.3). Patients who take these antidepressants can experience side effects including gastrointestinal (GI) disturbances (nausea, vomiting, diarrhea, constipation), headaches, insomnia, sedation, fatigue, sexual dysfunction, dry mouth, sweating, anxiety, prolonged QTc intervals, GI bleeding when combined with non-steroidal anti- inflammatory drugs (NSAIDs), serotonin syndrome when combined with other serotonergic agents, in addition to rare adverse events such as osteoporosis and fractures in the elderly, hyponatremia due to syndrome of inappropriate antidiuretic hormone (SIADH), and agranulocytosis. In practice, most patients need to try several antidepressants prior to achieving a satisfactory response. The long therapeutic lag before benefits are seen in OCD pharmacotherapy translates into medication trials requiring 10 to 12 weeks duration each. This factor in combination with the typical pattern of trying a number of different medications before a good response is achieved means that many patients endure periods of many months of ongoing disability even after starting medications. Currently, there are no practice guidelines that differentiate patients according to their response or tolerance to psychotropic medications and/or psychotherapy. Thus, the ability to identify in advance the right medication or therapy that will result in the best response would make a substantial difference to clinical practice. One of the goals of future genetic testing is to utilize genetic materials to determine whether a patient will respond to and/or tolerate a proposed treatment (i.e., medication, psychotherapy).

6

7

Table 1.2. Pharmacological Treatment of OCD.

Lowest Highest FDA Off-Labelled Class Generic Name Brand Name Consensus Guideline Evidence Recommended Approved Highest Dose Dose Dose SSRI Prozac First-line First-line Level 1 40mg/day 80mg/day 80mg/day SSRI Luvox First-line First-line Level 1 150mg/day 300mg/day 300mg/day SSRI Zoloft First-line First-line Level 1 50mg/day 200mg/day 250mg/day SSRI Paxil First-line First-line Level 1 40mg/day 60mg/day 60mg/day SNRI Effexor First-line Second-line Level 2 150mg/day 225mg/day 375mg/day SSRI Celexa First-line Second-line Level 2 20mg/day 60mg/day 60mg/day SSRI Cipralex First-line Third-line Level 4 10mg/day 20mg/day 50mg/day TCA Anafranil Second-line Second-line Level 1 150mg/day 250mg/day 250mg/day NaSSA Mirtazapine Remeron Second-line Second-line Level 2 15mg/day 45mg/day 45mg/day SSRI: selective serotonin reuptake inhibitor SNRI: serotonin-norepinephrine reuptake inhibitor TCA: tricyclic antidepressant NaSSA: noradrenergic and specific serotonergic antidepressant

Table 1.3. Treatment Response Categories in OCD.

Treatment Response Categories in OCD (Pallanti et al., 2002)

I Recovery with Y-BOCS scores ≤8 II Remission with Y-BOCS score ≤16 III Full response with ≥35% reduction in Y-BOCS score IV Partial response with 25-35% reduction in Y-BOCS score V Non-response (resistant) with <25% reduction in Y-BOCS score VI Relapase with symptoms return (25% increase in Y-BOCS score) after 3 months of "adequate" treatment or remission VII Refractory with no change (improvement or worsening) with any conventional therapy Treatment response categories using the Y-BOCS score was defined by Pallanti et al. in 2002.

7

8

1.3 Gender Differences in OCD

OCD affects both genders equally (APA, 2000), although a higher rate in male children and adolescents has been reported (Fineberg et al., 2013). Evidence suggests that gender may play an important role in the neurobiological aspects, psychosocial factors, and behavioural patterns of several psychiatric disorders. Recent clinical studies of OCD postulated that gender may moderate the genetic expression of this complex disorder, resulting in gender-specific phenotypes (Cherian et al., 2014; de Mathis et al., 2011; Torresan et al., 2013).

Gender-related differences exist in the clinical characteristics of this illness but results have been inconsistent across studies. The following gender differences have been reported in OCD studies (Cherian et al., 2014; de Mathis et al., 2011): clinical phenomenology (Cherian et al., 2014; de Mathis et al., 2011; Jaisoorya et al., 2009; Tukel et al., 2013), age at onset (AAO) (Cherian et al., 2014; de Mathis et al., 2011; Weissman et al., 1994), and co-morbidities (Cherian et al., 2014; Tukel et al., 2013). Men appear to have a higher frequency of aggressive (Tukel et al., 2013), sexual, religious, symmetry/exactness obsessions, in addition to checking, ordering/arranging, and hoarding compulsions, while obsessions related to contamination and washing/cleaning compulsions, as well as somatic obsessions and compulsions, appear to be more common among women (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013). OCD males, especially those with early onset OCD, were reported to have a higher rate of comorbid and attention deficit hyperactivity disorder (Leckman et al., 1994). However, one study did not find gender-related phenomenological differences in OCD individuals (Matsunaga et al., 2000).

Additional clinical studies have also found that OCD females tend to have an abrupt onset of illness with episodic presentation following stressful life events, whereas males with OCD generally have an insidious onset with a more chronic course (de Mathis et al., 2011; Nestadt et al., 1998; Rosso et al., 2012) and higher symptom severity (Nestadt et al., 1998; Rosso et al., 2012).

8

9

1.4 Genetics of OCD

The biological and genetic underpinning of OCD was first reported in 1936, describing the heredity of “obsessional neurosis” (Lewis, 1936). Up to 2015, there have been over 500 peer-reviewed publications on the human genetics of OCD. Evidence supporting the genetic basis of OCD originates from family aggregation and twin studies (Pauls, 2010; Pauls et al., 2014), which have provided heritability estimates ranging from 27% to 47% in adults and 45% and 65% in children (van Grootheest et al., 2008). Although knowledge is very limited, there appear to be genetic differences between OCD and the other OCRDs. Twin and family studies have estimated a heritability ranging from 51.1% to 71.4% for hoarding (Monzani et al., 2013; Mathews et al., 2007) and 31.6% to 78% for trichotillomania (Monzani et al., 2014; Novak et al., 2009). Two studies have estimated that 40% and 44% of the vulnerability for skin picking and body dysmorphic concerns respectively was due to genetic factors (Monzani et al., 2012). However, notwithstanding the high genetic loading shared by these disorders, there may also be important genetic distinctions, which may influence medication response differentially across these conditions. For example, Monzani et al. (Monzani et al., 2014) identified a genetic vulnerability factor for trichotillomania and skin picking disorder that is distinct from the others; and hoarding has long been described as showing a pattern of relatively poor response to pharmacotherapy (Saxena et al., 2011).

Evidence from five genetic linkage analyses and over 100 genetic candidate gene association studies supports the relevance of several candidate regions and genes to OCD (Figure 1.1), with the most consistent reports being in the glutamatergic and serotonergic systems (Mathews et al., 2012; Pauls, 2010). The following sections will highlight the major findings from candidate gene studies of OCD.

9

10

region 9p24.2 /(5HT1.) glutamate transporter (SLC1A1) region 12q13.1 glutamate receptor, region 11q14.1 ionotropic, N-methyl brain-derived neurotrophic D-aspartate 2. factor (.DNC) (GRIN2.)

region 3q13.3 dopamine receptor D3 (DRD3) region 13q14.2 region 21q22.11 serotonin receptor 2A (5HT2A) region 17q11.2 oligodendrocyte lineage serotonin transporter transcription factor 2 (hLIG2) (SLC6A4)

Adapted from: Pato et al. (2002) J Clin Psychiatry 63(Suppl. 6):30-3

Figure 1.1. Genes Associated with OCD. This figure illustrates genes located in their associated chromosomal regions with at least one significant association with OCD. Copright permission has been given by Lundbeck Institute to use this figure. 10

11

Glutamatergic System:

Evidence implicates abnormality in the glutamatergic system as part of the etiology of OCD with the most robust genetic results coming from genes that are involved in this system. Support of its involvement comes from the most consistently implicated neuropathophysiological model of OCD, which postulates a disturbance in the cortico-striato-thalamo-cortical (CSTC) circuit (Pauls et al., 2014). Convergent evidence from animal and human research proposes an abnormal threshold of activation in this circuit, leading to hyperactivation of the orbitofrontal- subcortical pathway (Pauls et al., 2014). Research studies have postulated an alteration of glutamatergic neurotransmission in the CSTC circuit that involves the , which mediates exaggerated concerns related to danger, harm, symmetry, or hygiene as seen in OCD patients with obsessions, and this in turn is thought to generate compulsions in order to neutralize these obsessions (Pauls et al., 2014). Temporary relief of anxiety reinforces the behaviour in a positive feedback loop, which leads to eventual debilitating OCD symptoms. Further evidence implicating the glutamatergic system in the etiology of OCD comes from magnetic resonance spectroscopy studies, knock-out mouse models, genetic association studies, and clinical trials of glutamatergic agents (Wu et al., 2012).

The first published genome-wide association study (GWAS) in OCD by the International OCD Foundation Genetics Collaborative group did not detect genome-wide significant association between any tested markers and OCD diagnosis (Stewart et al., 2013). However, interesting trends were observed in several glutamatergic system genes including the ‘discs large (drosophila) homolog-associated protein 1’ (DLGAP1) and glutamate receptor, ionotropic, kainate 2 (GRIK2) genes. Although a second GWAS by the OCD Collaborative Genetics Association Study also did not find any genome-wide significant results (Mattheisen et al., 2015), the authors detected a trend in the top-hit marker on chromosome 9 near the protein- tyrosine phosphatase, receptor-type, delta (PTPRD) gene (P=4.13×10-7), which promotes glutamatergic synaptic differentiation. Despite a non-significant trend in a glutamatergic system gene, the neuronal glutamate transporter gene (SLC1A1) from a meta-analysis (Taylor, 2013), another recent but more in-depth meta-analysis by Stewart et al. (Stewart et al., 2013) of association between nine previously examined single nucleotide polymorphisms (SNPs) across

11

12

the 3’ region of the SLC1A1 gene [first reported by Arnold et al. (Arnold et al., 2006)] and OCD illness revealed a consistent moderately significant finding in one of the SNPs, rs301443 (P=0.046), with another SNP showing modest association when controlled for gender (rs12682897; P=0.012). Furthermore, another glutamate system gene, glutamate receptor, ionotropic N-methyl D-aspartate 2B (GRIN2B), has also been reported to have a significant association with the diagnosis of OCD (Arnold et al., 2004).

Serotonergic System:

The serotonergic system has been extensively studied in OCD given that SSRIs are the first-line pharmacological treatment of OCD based on clinical trials (CPA, 2006; Koran et al., 2007; Soomro et al., 2008). SSRIs increase the level of neurotransmitter serotonin in the synaptic cleft that binds to postsynaptic receptors by inhibiting its reuptake into presynaptic cells. This is a widely accepted mechanism of action of SSRIs that reduces depression and anxiety in a complex interaction with other neurotransmitter systems.

Genes related to the serotonergic neurotransmitter system have been examined extensively in relation to the genetic risk of developing OCD given its functional implication, the role of serotonin in the proposed neuro-circuitry, and the significance of selective serotonin reuptake inhibitors as the first-line pharmacological treatment of OCD. A recent meta-analysis of OCD genetic association studies (Taylor, 2013) showed nominally significant results for a serotonin transporter promoter (SLC6A4) polymorphism (HTTLPR; OR=1.251, 95% confidence interval [CI] 1.048-1.492, P=0.003) and the serotonin 2A receptor (HTR2A) rs6311 marker (OR=1.219, 95% CI 1.037-1.433, P=0.003).

Other Systems:

Interest in the dopaminergic system has come from the treatment of OCD using antipsychotics as augmenting agents to SSRI antidepressants in addition to dopamine interaction with the serotonergic system (Pauls et al., 2014). Further evidence from animal, neuroimaging,

12

13

and neurochemical studies implicates dopaminergic dysfunction in the pathoetiology of OCD (Koo et al., 2010). Although Taylor (2013) in his recent meta-analysis of OCD genetic association studies, showed a nominally significant finding for the catechol-O-methyltransferase (COMT) rs4680 marker (OR=1.200, 95% CI 1.001-1.438, P=0.010), and non-significant trends for two additional dopaminergic system genes, the dopamine transporter (DAT1) and dopamine D3 receptor (DRD3), no further genetic replication studies support the dopamine system in OCD. Genes involved in the gabaergic system have been investigated with inconsistent and non- replicated results (Taylor, 2013). Similar lack of support was reported from the brain-derived neurotrophic factor (BDNF) gene including the most studied Val66Met (rs6265) variant (Zai et al., 2015), as well as genes related to neuroplasticity (Taylor, 2013).

Candidate Genes Identified through Animal Studies: Although human behaviours are highly complex, important confounds are difficult to manipulate for ethical limitations when using invasive techniques in human subjects. Animal models provide the ideal tool to directly modify neural circuitry, in the expectation that it may lead to the manifestation of behaviours similar to those of human beings with OCD. Abnormalities within CSTC circuitry, which involves the orbitofrontal cortex, anterior , and , have been consistently replicated in OCD (Monteiro and Feng, 2015; Saxena and Rauch, 2000; Ting and Feng, 2011). Based on these solid findings, development of novel genetic animal models of OCD that can manipulate this circuit have been used to further our understanding of the etiology of OCD. Four genetic animal models have been reported: SAP90/PSD95-associated protein 3 (SAPAP3) null mice, SLIT and NRTK-like family, member 5 gene (SLITRK5) null mice, homeobox B8 (HOXB8) null mice, and SLC1A1/EAAC1 null mice. SAPAP3 encodes a protein, which is highly expressed in the striatum. It is involved in postsynaptic scaffolding at excitatory (glutamatergic) synapses. Evidence suggests that disruption of the glutamatergic neurotransmitter system in the CSTC circuitry plays a vital role in the etiology of OCD (Wu et al., 2012). Abnormal behaviours consistent with an increase in anxiety and compulsive self-grooming were observed in mice that had a deletion of this gene, SAPAP knock-out or null mice (Welch et al., 2007). More interestingly, the same study showed that these behaviours were alleviated with repeated administration of fluoxetine. Following this

13

14

discovery, three genetic association studies have examined the role of this gene in OCD. Bienvenu et al. (2009) tested the association of SAPAP3 in 383 families with grooming disorders and reported nominally significant association with at least one grooming disorders but not with OCD itself. Another study sequenced SAPAP3 in patients with TTM and OCD (77 TTM, 44 OCD+TTM, 44 OCD) and detected significantly higher percentage of novel nonsynonymous heterozygous variants than healthy controls (4.2% versus 1.1%) (Züchner et al., 2009). The last study investigated seven variants within SAPAP3 in 172 OCD patients, 45 TTM patients, and 153 healthy controls (Boardman et al., 2011) and observed an initial positive association between rs11583978 and TTM, which did not survive correction for multiple testing. Additionally, they detected significant association between earlier AAO and the A-T-A-T (rs11583978-rs7541937- rs6662980-rs4652867) haplotype when compared with the C-G-G-G haplotype, implicating this gene in the development of early onset OCD. SLITRK5 encodes for a protein that regulates neurite outgrowth and is important in neuronal survival. Shmelkov et al. (2010) knocked out SLITRK5 in mice, which then exhibited obsessive-compulsive-like behaviours including excessive self-grooming and increased anxiety in the open maze test. This study showed that these behaviours were once again reversed using chronic fluoxetine treatment, demonstrating similarities with the SAPAP3 null mice. Although SLITRK5 is widely expressed in the CNS, preferential increase of neuronal activity in the orbitofrontal cortex of the SLITRK5 null mice was observed, which is consistent with human OCD functional imaging findings (Grünblatt et al., 2014). This mouse model also showed a reduction in striatal volume in addition to a decrease in dendritic complexity (Shmelkov et al., 2010). Striatal dysfunction has been implicated in OCD (Grünblatt et al., 2014).

HOXB8 encodes a nuclear protein that functions as a sequence-specific transcription factor with an important role in establishing body patterning during embryonic development. HOXB8 knockout mice developed compulsive self-grooming behaviour and fur loss in addition to excessive grooming of wildtype (normal) caged-mice (Greer and Capecchi, 2002).

The last animal model has shown the greatest genetic contribution in OCD. The neuronal excitatory amino acid carrier 1 (EAAC1), which is encoded by SLC1A1, plays a vital function in transporting glutamate and regulating postsynaptic action of glutamate, which is essential in maintaining extracellular glutamate concentrations below oxidative stress and toxic levels

14

15

(Aoyama et al., 2006; Scimemi et al., 2009). Thirty percentage of EAAC1 null mice exhibited an increase in aggression in addition to excessive self-grooming and fur loss (Aoyama et al., 2006), which was reduced with N-acetyl-cysteine (NAC) treatment, a prodrug of glutathione that modulates the glutamatergic system (Aoyama et al., 2006; Berman et al., 2010). This finding further suggests the role of glutamate in the etiology of OCD and provides additional treatment avenues for this debilitating disorder.

Although these animal models have demonstrated compulsive grooming behaviours, further genetic studies in human are necessary to translate these intriguing findings that may help to understand the underlying mechanism and improve the treatment of OCD.

Summary:

In summary, these genetic findings in OCD are exciting but premature in resolving the heterogeneity of OCD. This is because of small sample sizes with limited statistical power, and difficulty in replication. Furthermore, many genes across different neurotransmitter systems confer only a small contribution to the risk of developing OCD, suggesting a polygenic model of liability (Browne et al., 2014).

The difficulty in generating replicable results in the genetics of OCD risk is partly due to the relatively high clinical heterogeneity of OCD; thus, researchers have undertaken strategies, aiming to reduce the phenotypic variability of OCD.

1.4.1 Genetics of Gender Differences in OCD

Evidence suggests important gender differences in human behaviour (i.e., emotion recognition, mentalizing, social behaviours) (Christov-Moore et al., 2014) and in psychiatric disorders (Arnold, 2003; Brady and Randall, 1999; Figueira and Ouakinin, 2010; Kornstein, 1997; Rucklidge, 2008). Although the rate of OCD is equal in males and females as stated earlier, studies have reported OCD males having an earlier onset (Cherian et al., 2014; de Mathis et al., 2011; Weissman et al., 1994), greater severity (Nestadt et al., 1998; Rosso et al., 2012), 15

16

and a more chronic course of OCD (de Mathis et al., 2011; Nestadt et al., 1998; Rosso et al., 2012) than females. Thus, gender may moderate the genetic expression of this complex illness, resulting in specific phenotypes with OCD males perhaps having a higher genetic loading than females. The most consistent findings for the genetics of a gender effect in OCD have been reported in SLC1A1 and COMT (Table 1.4; for further details, please refer to Appendix IV Table 1).

Although only four studies have examined SLC1A1, all showed significant results. Arnold et al. (2006) first reported a preferential transmission of rs301434 (P=0.0007-0.002), rs301435 (P=0.001-0.009), rs3087879 (P=0.002-0.006), and the rs301434-rs301435 haplotype (P=0.0006-0.001) to males and the total sample of 157 OCD families. Within the same year, Dickel et al. (2006) detected another SNP, rs3780412, was associated with early-onset males (P=0.002) in a sample of 71 OCD trios, which was replicated in another 57 OCD trios, over- transmission of rs3780412 C allele to males but not to females (permuted P=0.045) (Stewart et al., 2007). Shugart et al. (2009) later reported that another SNP, rs301443, was associated with OCD males (P=0.00027) in 378 OCD families. However, a recent meta-analysis of 3’ SNPs within this gene in males only reported no significant results (rs12682807; uncorrected P=0.012 but non-significant after correction) (Stewart et al., 2013).

Five studies investigated BDNF for gender effect in OCD and only one presented negative results. Hemmings et al. (2006) investigated multiple candidate genes in OCD phenotypes and specifically for gender, the authors reported significant findings for the rs6265 marker in OCD males (P=0.036 – total sample of 132 OCD and 218 controls). This polymorphism was further examined and results were as follows: the Met allele conferring OCD risk to males (P=0.047) (Hemmings et al., 2008), Met/Met genotype associating with later onset OCD in females (P=0.031) (Katerberg et al., 2009), and Val/Val genotype predicting more severe OCD symptoms in OCD females (P=0.031) (Hemmings et al., 2008); however, one study failed to report positive findings for this marker (Márquez et al., 2013). Nonetheless, the latter study examined additional SNPs and found that the rs7124442 C allele was associated with OCD females (P=0.018). Alonso et al. (2008) tested eight SNPs within BDNF including rs6265 in addition to 46 SNPs within the neurotrophic tyrosine kinase, receptor, type 2 (NTRK2) gene and

16

17

only reported association between the NTRK2 rs2378672 and OCD females (P<0.0001) but not with BDNF in a case-control study comprising of 215 OCD patients and 342 controls.

The most investigated gene is COMT with seven positive and three negative studies in addition to a meta-analysis. Karayiorgou et al. (1997) was the first group to report an increased frequency of the rs4680 Met allele and Met/Met genotype in OCD males (P=0.0002), which was replicated in three case-control studies (Denys et al., 2006; Pooley et al., 2007; Poyurovsky et al., 2005) in addition to a meta-analysis (P<0.0001) (Pooley et al., 2007); however, two later studies detected significant association of the same Met allele with OCD females (P=0.048 with 56 trios) (Alsobrook et al., 2002) (P=0.039) (Katerberg et al., 2010). Two studies detected positive results in the total OCD sample but did not observed differences in gender (Kinnear et al., 2001; Schindler et al., 2000) and Hemmings et al. (2006) only reported an association between rs362204 and AAO in OCD males (P=0.040). Liu et al. (2011) only published data for the rs2075507 SNP and demonstrated that the G/G genotype was more frequent in OCD males (P<0.001) and OCD females (P=0.026) than in healthy controls; however, there was no difference between the gender groups.

1.4.2 Genetics of OCD Subphenotypes

Clinical, epidemiological, treatment, interventional, and genetic studies have consistently observed the heterogeneous nature of OCD (Pauls et al., 2014). Previous research has attempted to define a more phenotypically homogeneous subtype or subgroup of OCD (de Mathis et al., 2006) in order to identify specific genetic and environmental factors that influence the risk of developing OCD.

Evidence suggests that OCD is a clinically and genetically heterogeneous disorder and researchers have focused on the identification of valid subgroups of OCD (Leckman et al., 2010). The examined subgroups include early onset, gender effects (male versus female), symptom dimensions (different OCD symptoms based on Y-BOCS symptom checklist), neuropsychological performance (i.e., cognitive flexibility), neuroimaging data (i.e., caudate, orbitofrontal cortex, anterior cingulate cortex), and treatment response (mainly SRIs). To date,

17

18

only two such subtypes have been accepted into the DSM-5 and these specifiers are the degree of insight and the presence of a lifetime diagnosis of disorder (APA, 2013). The following sub- sections describe the different OCD phenotypes that were examined in genetic association studies within this thesis: age at onset (AAO), symptom dimensions and severity, psychiatric comorbidities, familiality, and SRI treatment response.

18

19

Table 1.4. Summary Table of Genetic Studies of Gender in OCD.

Gene Results Reference GRIK2 - Mas et al., 2014 GRIN2B -- Hemmings et al., 2006; Mas et al., 2014 GRIA1 - Mas et al., 2014 GRIA3 - Mas et al., 2014 SLC1A1 ++++/- Arnold et al., 2006; Dickel et al., 2006; Mas et al., 2014; Shugart et al., 2009; Stewart et al., 2007 DLGAP3 - Mas et al., 2014 GAD1 - Mas et al., 2014 GAD2 + Mas et al., 2014 Baca-Garćia et al., 2005; Dickel et al., 2007; Liu et al., 2011; Lochner et al., 2004; Mas et al., 2014; Voyiaziaki SLC6A4 ++/--- et al., 2011 HTR1B +++/- Dickel et al., 2007; Kim et al., 2009; Lochner et al., 2004; Mas et al., 2014 HTR2A ++/--- Dickel et al., 2007; Enoch et al., 2001; Hemmings et al., 2006; Mas et al., 2014; Walitza et al., 2002 HTR2C +/-- Cavallini et al., 1998; Hemmings et al., 2006; Mas et al., 2014 HTR6 - Hemmings et al., 2006 SLC18A1 - Mas et al., 2014 TPH2 - Mas et al., 2014 DAT/SLC6A3 +/---- Billett et al., 1998; Hemmings et al., 2006; Lochner et al., 2004; Mas et al., 2014; Miguita et al., 2007 DRD1 - Hemmings et al., 2006 DRD2 +/--- Billett et al., 1998; Denys et al., 2006; Hemmings et al., 2006; Mas et al., 2014 DRD3 --- Billett et al., 1998; Hemmings et al., 2006; Mas et al., 2014 DRD4 +++/- Billett et al., 1998; Camarena et al., 2007; Hemmings et al., 2006; Taj et al., 2013 Alsobrook et al., 2002; Denys et al., 2006; Hemmings et al., 2006; Karayiorgou et al., 1997; Katerberg et al., +++++++/---- (+ COMT 2010; Kinnear et al., 2001; Liu et al., 2011; Lochner et al., 2004; Mas et al., 2014; Pooley et al., 2007; meta-analysis) Poyurovsk et al., 2005; Schindler et al., 2000 MAOA ++/- Camarena et al., 2001; Lochner et al., 2004; Mas et al., 2014 Alonso et al., 2008; Dickel et al., 2007; Hemmings et al., 2006; Hemmings et al., 2008; Katerberg et al., 2009; BDNF ++++/--- Márquez et al., 2013; Mas et al., 2014 NTRK2 +/- Alonso et al., 2008; Mas et al., 2014 NTRK3 - Mas et al., 2014 AKT1 - Mas et al., 2014 GSK3β - Mas et al., 2014 NGFR - Mas et al., 2014

19

20

ERBB4 - Mas et al., 2014 NRG1 - Mas et al., 2014 OLIG1 - Mas et al., 2014 OLIG2 - Mas et al., 2014 LMX1A - Mas et al., 2014 BDKRB2 - Mas et al., 2014 CDH9 - Mas et al., 2014 KCNN3 - Mas et al., 2014 EFNA5 - Mas et al., 2014 TH - Lochner et al., 2004 ACE + Hemmings et al., 2006 ESRα - Hemmings et al., 2006 INPP-1 - Hemmings et al., 2006 PLC-γ1 - Hemmings et al., 2006 HOXB8 - Hemmings et al., 2006 + indicates significant study - indicates negative study

20

21

1.4.2.1 The Significance of Age at Onset (AAO) in OCD

The consensus of age at onset (AAO) in OCD is usually between the ages of 10 and 37 years although two thirds of OCD cases had emerged by the age of 22 years (Fineberg et al., 2013).

Most research studies to date have arbitrarily and subjectively defined early onset as less than 10 to 18 years depending on the research groups (Chabane et al., 2005; Janowitz et al., 2009; Maina et al., 2008) and late onset as on or after age 17 to 30 years (Grant et al., 2007; Maina et al., 2008); however, many did not examine AAO closely to calculate the early versus late onset cutoff using analytical approaches. Furthermore, some studies have defined AAO as age at symptom onset (Butwicka and Gmitrowicz, 2010; de Mathis et al., 2009), while some set AAO as age when first diagnosed (Maina et al., 2008; Taylor, 2011; Tukel et al., 2005). The arbitrary and different definitions of AAO used in previous studies will give rise to further clinical heterogeneity and pose significant challenges in comparing different datasets and in combining them in order to increase sample size.

Several studies have since performed the admixture analysis to identify various normal distribution curves within the dataset. This provides an objective method to calculate the exact point where the normal distribution curves meet. Only four such studies have used admixture analysis to determine the best-fitting model for AAO in the OCD population. The first study was conducted by Delorme et al. (2005) in 161 OCD subjects and the admixture analysis detected two distributions with mean ages of 11.1 years for early-onset and 23.5 years as late-onset. Delorme and colleagues (2005) noticed significantly greater Tourette’s syndrome and family history of OCD in early-onset group and higher prevalence of generalized anxiety disorder and MDD in the late-onset group. This study supported the notion that early- and late-onset groups may exhibit different phenotypic characteristics. The second study used mixture analysis in a subset of the current OCD sample (N=196), which was conducted by our group (De Luca et al., 2011). De Luca et al. (2011) detected two normal distributions with a cut-off at 15 years and found that the early-onset OCD patients were assessed at an earlier age and have more frequent panic attacks than the late-onset subjects. The third study performed admixture analysis in 377 OCD individuals and also reported a bimodal distribution for AAO with the best-fitting cut-off

21

22

AAO for early-onset as less than or equal to 19 years (Anholt et al., 2014). Anholt et al. (2014) found that the early-onset group has higher OCD symptom severity across all OCD symptom dimensions in addition to greater ADHD symptoms and higher rates of comorbid bipolar disorder. The authors suggested that the early-onset group may be associated with greater functional impairment and poor clinical prognosis. A very recent paper examined AAO using two different definitions, onset when symptom first appeared and onset when symptom first met diagnostic criteria of OCD (Albert et al., 2015). The authors conducted admixture analysis in 483 OCD subjects with two different sets of AAO. They identified a trimodal distribution for age at symptom onset with mean ages of 6.9 years for early-onset, 14.99 years for the intermediate onset, and 27.7 years for the late-onset group; and reported early- and intermediate- onset as having an insidious onset (rather than acute onset) and a chronic course of illness (rather than episodic) when comparing to the late-onset group. For the age at diagnosis, they confirmed a bimodal distribution with means 18.0 years and 29.5 years. Early-onset group for both AAO definitions was found to have a higher prevalence of younger male subjects and greater severity and rate of symptoms related to religion, symmetry or exactness, contamination and cleaning/washing, in addition to repeating and counting.

A review and meta-analysis (Taylor, 2011) of clinical characteristics of early- versus late- onset OCD indicated that early-onset OCD: a) occurs more often in males (de Mathis et al., 2011; Jaisoorya et al., 2009; Narayanaswamy et al., 2012; Nestadt et al., 1998; Rasmussen and Eisen, 1992); b) associates with higher level of symptom severity and greater prevalence of types of obsessive-compulsive symptoms; c) has more comorbidities including tics and other OCRDs; and associates with a higher prevalence of OCD in first-degree relatives. A recent multi-center study indicated that OCD patients with earlier AAO had significantly longer duration of illness and higher disability in work, social, and family life (P=0.017-0.035) (Dell’Osso et al., 2013). Nakamae (2011) reported individuals with early-onset OCD appear to be more SRI treatment resistant when compared to later onset.

The rationale for separating out the early onset group is that the heritability of childhood OCD is higher than the adult OCD population, ranging from 27% to 47% in adults and 45% and 65% in children (van Grootheest et al., 2008). The concept of early-onset disease as having a more familial pattern of genetic influence over later onset of the same illness is well-documented

22

23

[i.e., bipolar disorder (Rende et al., 2007; Strober et al., 1988), major depressive disorder (Kendler et al., 2007; Levinson et al., 2003; Sullivan et al., 2000; Weissman et al., 1984), mood disorders (Orvaschel, 1990)]. Furthermore, specifically in OCD, earlier onset was postulated to have a higher familial and genetic loading than late onset OCD (Chabane et al., 2005; do Rosario-Campos et al., 2005; Hemmings et al., 2004). Thus, by examining the genetics of different AAO groups as identified using the admixture methodology, we may have a better chance of identifying genetic markers, which may predispose to early rather than late onset OCD.

Mixed and inconsistent genetic results have been reported across different neurotransmitter system candidate genes (Table 1.5; for further details, please refer to Appendix IV Table 1). Mathews et al. (2012) published the only genome-wide linkage study in 33 childhood-onset OCD families and reported a LOD score linkage of greater than two on chromosome 1p36, 2p14, 5q13, 6p25, and 10p13. Although 47 genes have been explored in relation to AAO in OCD, most were single study or with very inconsistent and conflicting findings (Table 1.5). Genes described below are considered to have a more robust result. Two significant meta-analyses examined SLC6A4 5HTTLPR (Bloch et al., 2008) and reported an over-transmission of the La allele to early-onset OCD (P=0.00021) (Walitza et al., 2014) when almost all the individual samples were too small to detect any positive results. In addition to SLC6A4, HTR2A was investigated in nine published articles with mixed results (Table 1.5 and for additional details of each study, please refer to Appendix IV Table 1). Only one study examined linkage and reported a LOD score of two on chromosome 9q where SLC1A1 is located (Hanna et al., 2002) and three out of five studies reported significant findings of different SNPs that were associated with AAO [rs3780412 and rs301430 with early-onset (P=0.04) (Dickel et al., 2006); rs10491734 A allele (P=0.023) and A/A genotype (P=0.009) with early onset (Wu et al., 2013); rs301430 C/C genotype with later onset (P=0.017) (Dallaspezia et al., 2014)]. Regarding BDNF, the first significant study with several SNPs was reported by Hall et al. (2003) and only two of eight subsequent studies supported the role of the rs6265 marker in AAO but with gender bias (Hemmings et al., 2006; Katerberg et al., 2009). Thus, with such scattered genetic findings, it is important to replicate some of these findings in a much larger sample of OCD, which can be stratified into AAO groups.

23

24

24

25

Table 1.5. Summary Table of Genetic Studies of AAO in OCD.

Gene Results Reference GRIK2 - Mas et al., 2014 GRIN2B +/- Hemmings et al., 2006; Mas et al., 2014 GRIA1 - Mas et al., 2014 GRIA3 - Mas et al., 2014 SLC1A1 +++/-- Dallaspezia et al., 2014; Dickel et al., 2006; Mas et al., 2014; Veenstra-VanderWeele et al., 2001; Wu et al., 2013 DLGAP3 - Mas et al., 2014 GAD1 - Mas et al., 2014 GAD2 + Mas et al., 2014 SLC6A4 (++ meta-analyses)/-- Bloch et al., 2008 (meta-analysis); Denys et al., 2006; Dickel et al., 2007; Liu et al., 2011; Lochner et al., ------2004; Lochner et al., 2008; Mas et al., 2014; Walitza et al., 2004; Walitza et al., 2014 HTR1B +/------Denys et al., 2006; Dickel et al., 2007; Hemmings et al., 2004; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Wendland et al., 2007 HTR2A ++++/----- Denys et al., 2006; Dickel et al., 2007; Hemmings et al., 2004; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Walitza et al., 2002; Walitza et al., 2012 HTR2C -- Hemmings et al., 2006; Mas et al., 2014 HTR6 - Hemmings et al., 2006 SLC18A1 - Mas et al., 2014 TPH1 - Walitza et al., 2004 TPH2 -- Mossner et al., 2006; Mas et al., 2014 DAT/SLC6A3 +/------Hemmings et al., 2004; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Miguita et al., 2007; Walitza et al., 2008 DRD1 - Hemmings et al., 2006 DRD2 - Mas et al., 2014 DRD3 +/- Hemmings et al., 2006; Mas et al., 2014 DRD4 ++/---- Hemmings et al., 2004; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Walitza et al., 2008 COMT ++/----- Hemmings et al., 2004; Hemmings et al., 2006; Liu et al., 2011; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Walitza et al., 2008 MAOA ----- Camarena et al., 2001; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014 BDNF +++/------Dickel et al., 2007; Hall et al., 2003; Hemmings et al., 2006; Hemmings et al., 2008; Katerberg et al., 2009; Márquez et al., 2013; Mas et al., 2014; Mossner et al., 2005; Wendland et al., 2007 NTRK2 - Mas et al., 2014

25

26

NTRK3 - Mas et al., 2014 AKT1 - Mas et al., 2014 GSK3β - Mas et al., 2014 NGFR - Mas et al., 2014 ERBB4 - Mas et al., 2014 NRG1 - Mas et al., 2014 OLIG1 - Mas et al., 2014 OLIG2 - Mas et al., 2014 LMX1A - Mas et al., 2014 BDKRB2 - Mas et al., 2014 CDH9 - Mas et al., 2014 KCNN3 - Mas et al., 2014 EFNA5 - Mas et al., 2014 SAPAP3 - Mas et al., 2014 TNFα + Lüleyap et al., 2012 NPSR1 + Lennertz et al., 2013 TH -- Lochner et al., 2004; Lochner et al., 2008 ACE - Hemmings et al., 2006 ESRα - Hemmings et al., 2006 INPP-1 - Hemmings et al., 2006 PLC-γ1 + Hemmings et al., 2006 HOXB8 + Hemmings et al., 2006 + indicates significant study - indicates negative study

26

27

1.4.2.2 Clinical Heterogeneity of OCD: Symptom Dimensions and Severity

One of the main challenges in OCD is the large number of symptom presentations and the varying degree of symptom severity in patients who suffer from OCD. The following section will provide rationale for exploring these subphenotypes in genetics.

1.4.2.2.1 OCD Symptom Dimensions

The majority of OCD individuals endorse both obsessions and compulsions whereas approximately 25% and 5% have obsessions and compulsions only respectively (APA, 2010).

The National Comorbidity Survey Replication epidemiological study reported distinct symptoms of OCD including checking (79.3%), hoarding (62.3%), ordering (57.0%), moral concerns (43.0%), sexual/religious concerns (30.2%), contamination (25.7%), harming (24.2%), concerns related to illness (14.3%), and others (19.0%), with a high 81% of individuals endorsing multiple symptoms (Ruscio et al., 2010).

Given the vast number of possible OCD symptoms for the diagnosis of OCD, each individual with OCD may present very differently. In order to capture all of the OCD symptoms and to avoid over-simplification, factor analysis has been conducted to reduce the number of OCD symptoms. A factor analysis is used to group variables, in this case, different OCD symptoms, according to shared variance. OCD symptoms that are explained by the same variance are grouped together into each factor, symptom dimension. A different approach is the Ward’s method in the hierarchical cluster analysis to join variables (i.e., OCD symptoms) into clusters such that the variance within a cluster is minimized. Each variable begins as its own cluster and clusters are then merged in such a way as to reduce the variability within a cluster if the merging process results in the minimum increase in the error sum of squares. By reducing multiple OCD symptoms into several symptom dimensions, it aids clinician to interpret clinical outcome easier and helps genetic research to increase statistical power by limiting symptom stratification.

27

28

Several validated scales have been developed to capture OCD symptom dimensions. The Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score and symptom checklist is considered to be the gold standard to assess for OCD symptom severity and it is widely used to generate symptom dimensions (Goodman et al., 1989). Previous studies have consistently identified a four- or five-factor of symptom dimensions using the factor analysis based on principal component analysis (PCA) of the Y-BOCS symptom checklist at a category- or item- level including symmetry/ordering, hoarding, contamination/cleaning, aggressive/checking+/- sexual/religious concerns (Table 1.1).

The first study to use statistical approach to reduce the number of Y-BOCS symptom category was conducted in 1994 (Baer et al., 1994) and the 3-factor model explained 48% of the variance for symmetry/hoarding, contamination/cleaning, and pure obsessions. Since 1994, over 30 studies (Table 1.6, Table 1.7, Table 1.8; for full details, please refer to Appendix IV Table 2) have utilized the data reduction technique to calculate the minimal OCD symptom categories for comparing clinical features and treatment outcome. Eighteen studies have identified a 5-factor model for the Y-BOCS symptom checklist and the most common factors include contamination/cleaning, hoarding, aggressive/checking, symmetry/ordering, and sexual/religious (Table 1.6). Foa et al. (2002) used the Obsessive-Compulsive Inventory (Foa et al., 2002), which is a self-report measure of obsessive-compulsive traits with 18 items and six subscales, and the same five factors were reported. The second most common is the 4-factor model with aggressive/checking/sexual, hoarding, symmetry/ordering/counting, and contamination/cleaning as the most consistently identified factors (Table 1.7). The highest percentage of variance that explained the factor model was reported by Cullen et al. (2007) at 76.7% with the following factors: pure obsessions, contamination, symmetry/order, and hoarding. A meta-analysis, which included the above study, consisting of 5,124 OCD individuals, confirmed a four-factor model that explained 79% of the variance (Bloch et al., 2008). Four studies detected three factors and two reported six factors; moreover, the number of identified factors changes with the use of individual item of OCD symptoms versus using category of OCD symptoms (Table 1.8).

Other validated scales such as the Dimensional Y-BOCS (Rosario-Campos et al., 2006), the self-report Dimensional Obsessive-Compulsive Scale (Abramowitz et al., 2010), the self- report Obsessive-Compulsive Inventory (Foa et al., 2002), the self-report Clark-Beck Obsessive-

28

29

Compulsive Inventory (Clark et al., 2005), the self-report Maudsley Obsessional Compulsive Inventory (Hodgson et al., 1977), the self-report Padua Inventory (Sanavio et al., 1988), the self- report Padua Inventory-Revised (Van Oppen et al., 1995), the self-report Vancouver Obsessive Compulsive Inventory (Thordarson et al., 2004), and the self-report Obsessional Intrusive Thoughts Inventory (García-Soriano et al., 2011) have been used but either due to the length or the over-generalization of these inventories, Y-BOCS remains the most extensively utilized in the clinical OCD population.

Symptom dimensions are of clinical utility given that they respond differently to various treatments. Bloch et al. (Bloch et al., 2014) recently conducted a meta-analysis reporting poor treatment outcome in OCD patients with hoarding symptoms. Treatment implication for different symptom subtypes have previously been reported (Nakamae, 2011). OCD individuals with sexual, religious, and harm-related (aggressive) obsessions and checking compulsions appeared to have improved SRI response (>60% with very much improved) (Landeros-Weisenberger et al., 2010), whereas individuals with hoarding symptoms tend to be SRI resistant (Nakamae, 2011).

Studies have also reported clinical differences among different types of symptom dimension. Several studies indicated that somatic and contamination obsessions and washing/cleaning compulsions occur more frequently in female OCD patients (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013), whereas males appear to have a higher frequency of aggressive (Tukel et al., 2013), sexual, religious, symmetry/exactness obsessions, in addition to checking, ordering/arranging, and hoarding compulsions (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013).

A twin study of non-clinical obsessive-compulsive dimensional traits found that the hoarding dimension had the lowest common genetic factor loading and was 54.5% influenced by specific genetic effects (Iervolino et al., 2011); this confirms the separation of hoarding away from the OCD diagnostic criteria in DSM-5 (APA, 2013). Another study identified two statistically separate OCD subgroups according to OCD symptoms (Schooler et al., 2008). Schooler et al. (2008) reported that one group of OCD individuals had significantly higher rates of OCD in the families, earlier onset and more severe OCD symptoms, in addition to more

29

30

psychiatric comorbidities. Additional family studies reported different heritability for specific OCD symptom dimensions including significant correlations between relatives in the dimensions of hoarding (Hasler et al., 2005; Katerberg et al., 2010; Pinto et al., 2007), unacceptable/taboo thoughts (Hasler et al., 2005; Katerberg et al., 2010; Pinto et al., 2007), contamination and symmetry (Katerberg et al., 2010), in addition to checking compulsions (Hasler et al., 2005).

Several studies have examined the genetics of OCD symptom dimensions (Table 1.9; for further details, please refer to Appendix IV Table 1) and OCD symptom severity (Table 1.10; ; for further details, please refer to Appendix IV Table 1) but these studies were mostly under- powered and reported inconsistent results with no replication. Thirteen genetic studies have examined Y-BOCS symptom checklist dimensions in our search for published articles. The most consistent results were reported in the COMT rs4680 SNP, which has been found to be associated with hoarding (Hemmings et al., 2006), obsessional/checking dimension (Lochner et al., 2008), and somatic and (Katerberg et al., 2010); nonetheless, Liu et al. (2011) failed to replicate these findings. Although two studies detected significant associations between SLC6A4 5HTTLPR and similar factor, repeating/counting and symmetry/repeating/counting/ordering, their results implicated opposite genotypes, L/L genotype (P=0.0013) from the study conducted by Cavallini et al. (2002) and S/S genotype by Hasler et al. (2006). Hemmings et al. (2006) demonstrated that the DRD4 rs1800955 and ESRα rs9340799 were associated with hoarding but two studies have failed to replicate the DRD4 results (Lochner et al., 2008; Taj et al., 2013) and Alonso et al. (2012) reported a different marker, rs34535804, predicting lower risk of having contamination/cleaning factor symptoms (P=0.0001).

We aimed to investigate the effect of OCD candidate genes in Y-BOCS symptom checklist generated symptom dimensions because: 1) Limited number of published genetic studies (13 in total) have explored Y- BOCS symptom dimensions 2) Previous studies investigated commonly examined markers within the exons of the candidate genes or SNPs with no known functional role 3) Studies all have relatively small sample size 4) Results were inconsistent and have not been replicable.

30

31

5) Reducing clinical heterogeneity may identify risk genetic variations with strong effect size. 6) Biological mechanisms for symptom dimensions may be different (i.e., as in the case of hoarding symptom, which is now considered a distinct disorder) 7) Diagnostic criteria and/or definition of OCD may change over time.

31

32

Table 1.6. Summary Table of 5-Factor Y-BOCS Symptom Dimensions in OCD.

Variance Study Scoring * Analysis Technique Number 5 Factors Explained Contamination/cleaning/repeating, Tek and symmetry/ordering/somatic, Range=0-1 PCA, current symptoms 45 OCD 65.5% Ulug, 2001 aggressive/counting/hoarding, sexual/religious, checking/hoarding 180 OCD, Contamination/cleaning, hoarding, Cavallini et PCA, (category-level), lifetime Range=0-1 112 sexual/religious/aggressive/checking, 59.87% al., 2002 symptoms controls symmetry/ordering, repeating/counting Symmetry/ordering, hoarding, Mataix-Cols Range=0-2 PCA, category-level 117 OCD contamination/cleaning, aggressive/checking, N/A et al., 2002 sexual/religious Mataix-Cols PCA, category-level, current Aggressive/checking, contamination/washing, Range=0-1 153 OCD 63.7% et al., 2002 symptoms symmetry/ordering, hoarding, sexual/somatic Range=0-4 Symmetry/ordering, hoarding, Foa et al., (Obsessive- Confirmatory PCA, current 215 OCD contamination/cleaning, aggressive/checking, N/A 2002 Compulsive symptoms sexual/religious obsessions Inventory) Aggressive/sexual/religious obsessions, Denys et PCA, item-level, current Range=0-2 150 OCD contamination/cleaning, somatic/checking, 42.5% al., 2004 symptoms symmetry/ordering, high risk aggressive/checking Contamination/cleaning, Denys et PCA, item-level, current aggressive/sexual/religious/somatic, somatic Range=0-2 335 OCD 41.7% al., 2004 symptoms obsessions and checking, symmetry and prefectionism, high risk aggressive and checking Symmetry/ordering, hoarding, Pinto et al., Range=0-1 and Exploratory PCA, category- 293 OCD doubt/somatic/checking, contamination/cleaning, 65.5% 2007 0-2 level taboo thoughts (aggressive/sexual/religious) Contamination/cleaning, harm/checking, Stein et al., Exploratory PCA, item-level, Range=0-2 434 OCD aggressive/sexual/religious, hoarding/symmetry, N/A 2007 current symptoms somatic/hypochondriacal Pinto et al., PCA, item-level, lifetime Taboo thoughts, symmetry/ordering, hoarding, Range=0-1 485 OCD N/A 2008 symptoms contamination/cleaning, doubt/checking

32

33

PCA, categorical-level, lifetime Symmetry/ordering, taboo thoughts, hoarding, 64.1% symptoms doubt/checking, contamination/cleaning

Contamination/cleaning, harm/checking, Stein et al., Range=0-2 Exploratory PCA, item-level 466 OCD hoarding/symmetry, religious/sexual, N/A 2008 somatic/hypochondriacal

Y-BOCS and Thoughts and Behaviour Inventory (TBI; Schooler et Fear of hurting self/others, repetition/symmetry, Range=0-1 Slattery et al., 2004), 398 OCD N/A al., 2008 contamination, checking/counting, hoarding confirmatory PCA, item-level, lifetime symptoms

Contamination/cleaning, hoarding, Jang et al., symmetry/ordering, Range=0-2 PCA, current symptoms 144 OCD 65.41% 2010 aggressive/sexual/religious/checking, repeating/counting 200 OCD, Hoarding, contamination/cleaning, Liu et al., PCA, category-level, current Range=0-1 403 symmetry/ordering, aggressive/checking, 71.8% 2011 symptoms controls somatic/repeating Doubts/checking/repeating, Cherian et contamination/washing/cleaning, hoarding/collecting, N/A PCA, category-level 545 OCD 62% al., 2012 symmetry/ordering, forbidden/sexual/religious/aggressive Brakoulias Hoarding, contamination/cleaning, doubt/checking, Range=0-2 PCA, category-level 154 OCD 64.9% et al., 2013 symmetry/ordering, unacceptable/taboo thoughts 173 OCD, Taj et al., Hoarding, doubts/checking, symmetry/ordering, N/A PCA, category-level 201 65% 2013 contamination, forbidden thoughts controls Hoarding, contamination/cleaning, Brakoulias N/A PCA, category-level 154 OCD symmetry/ordering, unacceptable/taboo thoughts, 64.9% et al., 2014 doubt/checking Doubts about harm/checking, unacceptable/taboo Williams et Exploratory PCA, category- Range=0-2 238 OCD thoughts, contamination/cleaning, hoarding, 68.5% al., 2014 level, current symptoms symmetry/ordering

33

34

Multidimensional item 269 OCD, Martoni et response theory model, Hoarding, washing, symmetry, rituals, forbidden Range=0-1 120 N/A al., 2015 category-level, current thoughts controls symptoms

Scoring: investigators assigned a score of 1 if a symptom category was present, 0 if it was absent, and 2 was assigned in some studies for the most prominent symptom category PCA: principal component analysis N/A: not applicable Highlighted in grey indicates the use of other OCD symptom scales than Y-BOCS symptom checklist * Scoring of 0 indicates the absence of symptom; scoring of 1 indicates the presence of symptom; scoring of 2 indicates the presence of target or principal symptom

34

35

Table 1.7. Summary Table of 4-Factor Y-BOCS Symptom Dimensions in OCD.

Analysis Variance Study Scoring * Number 4 Factors Technique Explained PCA, Number Leckman et category-level, 2 OCD samples Obsessions/checking, symmetry/ordering, cleanliness/washing, of 63.5% al., 1997 lifetime (208 and 98) hoarding symptoms symptoms PCA, current Aggressive/sexual/religious/somatic/checking, Summerfeldt Range=0- symptoms, 203 OCD symmetry/repeating/ordering/counting, contamination and N/A et al., 1997 1 including cleaning/washing, hoarding miscellaneous PCA as per Number 100 OCD, 466 Alsobrook II Leckman et al. Aggressive/checking/sexual, hoarding, of first-degree ~70% et al., 1999 (1997), symmetry/ordering/counting/ritual, contamination symptoms relatives category-level Confirmatory PCA, Aggressive/sexual/religious/somatic/checking, Summerfeldt Range=0- category-level, 203 OCD symmetry/repeating/ordering/counting, contamination and N/A et al., 1999 1 current cleaning/washing, hoarding symptoms PCA, item- Symmetry/exactness/ordering/arranging/repeating/counting/touching, Feinstein et Range=0- level, current 160 OCD contamination/cleaning/aggressive/checking, hoarding, 54.1% al., 2003 1 baseline sexual/religious symptoms PCA, Aggressive/sexual/religious/checking, Hasler et al., Range=0- category-level, 169 OCD symmetry/repeating/counting/ordering/arranging, 64.7% 2005 1 lifetime contamination/cleaning, hoarding symptoms Kim et al., Range=0- PCA, 124 OCD, 171 Hoarding/repeating, contamination/cleaning, aggressive/sexual, 62.8% 2005 2 category-level controls religious/somatic PCA, Hasler et al., Range=0- category-level, Aggressive/religious/checking, symmetry/repeating/counting/ordering, 153 OCD 65.0% 2006 1 lifetime contamination/cleaning, hoarding symptoms PCA, 221 Cullen et al., Range=0- category-level, definite/probable Pure obsessions, contamination, symmetry/order, hoarding 76.7% 2007 1 lifetime OCD 35

36

symptoms

Wendland et Range=0- PCA, 347 OCD, 749 Obsessions/checking, symmetry/ordering, contamination/cleaning, N/A al., 2007 1 category-level controls hoarding PCA and cluster Aggressive/sexual/religious/somatic/checking, Hasler et al., Range=0- analysis, 418 OCD symmetry/repeating/counting/ordering/arranging, 64.0% 2007 1 category-level, contamination/cleaning, hoarding lifetime symptoms PCA, Matsunaga Range=0- category-level, Contamination/washing, hoarding, symmetry/repeating/ordering, 343 OCD 57.7% et al., 2008 2 current aggressive/checking symptoms Albert et al., Range=0- PCA, lifetime Symmetry/repeating/ordering/counting/checking, 329 OCD 58% 2010 1 symptoms aggressive/religious/sexual/somatic, contamination/cleaning, hoarding

58 OCD trios, Lennertz et Range=0- PCA, Symmetry/ordering/repeating/counting, contamination/washing, 236 OCD, 310 53.9% al., 2014 1 category-level hoarding, aggressive/sexual/religious controls Scoring: investigators assigned a score of 1 if a symptom category was present, 0 if it was absent, and 2 was assigned in some studies for the most prominent symptom category PCA: principal component analysis N/A: not applicable * Scoring of 0 indicates the absence of symptom; scoring of 1 indicates the presence of symptom; scoring of 2 indicates the presence of target or principal symptom

36

37

Table 1.8. Summary Table of 3-, 6-, and Mixed-Factor Y-BOCS Symptom Dimensions in OCD.

Analysis Variance Reference Scoring Number 3, 6, or Mixed Factors Technique Explained PCA, category- Baer et al., Range=0- Symmetry and hoarding, contamination and cleaning, pure level, current 107 OCD 48% 1994 2 obsessions symptoms Hantouche PCA, category- and N/A level, current 615 OCD Predominantly compulsive, predominately obsessive, mixed 32.5% Lancrenon, symptoms 1996 53 OCD, 96 non- Confirmatory Contamination/cleaning/repeating/checking/somatic, Wu et al., Range=0- OCD psychiatric factor analysis, aggressive/sexual/religious, N/A 2007 1 patients, 419 category-level hoarding/ordering/counting/symmetry students Cluster analysis, Lochner et Range=0- Contamination/washing, hoarding/symmetry/ordering, item-level, current 261 OCD N/A al., 2008 1 obsessional/checking symptoms Taboo, contamination/cleaning, rituals and superstition, fear Katerberg et Range=0- PCA, item-level, 373 OCD, 462 of harm, intolerance of uncertainty, somatic and sensory 66% al., 2010 1 lifetime symptoms controls phenomena PCA, category- Wiliams et Range=0- 2 OCD samples (74 Contamination/washing, hoarding, sexuality/reassurance, level, current 59.1% al., 2012 2 and 54) aggression/mental, symmetry/perfectionism, doubt/checking symptoms 5 factors (item-level): taboo, contamination/cleaning, doubts, 59% 1224 OCD, 52 OCD- superstitions/rituals, symmetry/hoarding Confirmatory PCA, Katerberg et Range=0- affected item-level, lifetime 4 factors (categorical-level): al., 2010 1 multigenerational symmetry/ordering/arranging/counting/repeating, symptoms 65.5% families aggressive/sexual/religious/checking, contamination/cleaning, hoarding Exploratory factor 5 factors (category-level): Zhang et al., Range=0- analysis (PCA), symmetry/arranging/repeating/counting, 512 OCD 67.1% 2013 1 category- and contamination/cleaning, hoarding, aggressive/checking, item-level, lifetime somatic/religious/sexual;

37

38

symptoms 6 factors (item-level): contamination/cleaning, symmetry/arranging/repeating/counting/superstition, hoarding, 66.1% doubt/checking, somatic, religious/sexual/mental

Scoring: investigators assigned a score of 1 if a symptom category was present, 0 if it was absent, and 2 was assigned in some studies for the most prominent symptom category PCA: principal component analysis N/A: not applicable * Scoring of 0 indicates the absence of symptom; scoring of 1 indicates the presence of symptom; scoring of 2 indicates the presence of target or principal symptom

38

39

Table 1.9. Summary Table of Genetic Studies of Y-BOCS Symptom Dimensions in OCD.

Gene Dimension Results Reference GRIN2B Contamination/cleaning +/- Alonso et al., 2012; Hemmings et al., 2006 SLC6A4 Repeating/counting, ++/--- Cavallini et al., 2002; Hasler et al., 2006; Kim et al., 2005; symmetry/repeating/counting/ordering Lochner et al., 2008; Wendland et al., 2007 HTR1B --- Hemmings et al., 2006; Kim et al., 2009; Lochner et al., 2008 HTR2A -- Hemmings et al., 2006; Lochner et al., 2008 HTR2C - Hemmings et al., 2006 HTR3A - Lennertz et al., 2014 HTR3B - Lennertz et al., 2014 HTR3C - Lennertz et al., 2014 HTR3D - Lennertz et al., 2014 HTR3E Contamination/washing + Lennertz et al., 2014 HTR6 - Hemmings et al., 2006 DAT/SLC6A3 -- Hemmings et al., 2006; Lochner et al., 2008 DRD1 - Hemmings et al., 2006 DRD2 - Hemmings et al., 2006 DRD3 - Hemmings et al., 2006 DRD4 Hoarding +/-- Hemmings et al., 2006; Lochner et al., 2008; Taj et al., 2013 COMT Hoarding, obsessional/checking, somatic and sensory +++/- Hemmings et al., 2006; Katerberg et al., 2010; Liu et al., 2011; phenomena Lochner et al., 2008 MAOA - Lochner et al., 2008 BDNF -- Hemmings et al., 2006; Wendland et al., 2007 TH - Lochner et al., 2008 ACE - Hemmings et al., 2006 ESRα Hoarding, contamination/cleaning ++ Alonso et al., 2011; Hemmings et al., 2006 ESRβ - Alonso et al., 2011 INPP-1 - Hemmings et al., 2006 PLC-γ1 - Hemmings et al., 2006 HOXB8 - Hemmings et al., 2006 + indicates significant study - indicates negative study

39

40

1.4.2.2.2 OCD Symptom Severity

As mentioned in the previous section, Y-BOCS is a clinician-rated, gold standard measure to assess for OCD symptom severity (Goodman et al., 1989). This scale consists of 10 items with each item scoring between 0 (no symptoms) to 4 (extreme), up to a total score ranging from 0 to 40. The scale is divided into two sections, one for obsessions and the other for compulsions, and includes questions related to the time spent on obsessions/compulsions, the interference that the obsessions/compulsions have caused, the amount of distress the patient experienced, the level of resistance, and control over the obsessions/compulsions. The scoring system is as follows: subclinical for 0-7, mild for 8-15, moderate for 16-23, severe for 24-31, and 32-40 for extreme. The severity score allows clinicians to assess treatment outcome.

Higher severity of symptoms has been associated with a greater genetic loading in psychiatric disorders (OCD – Schooler et al., 2008; ADHD – Stergiakouli). Thus, five in total, have looked into the genetics of OCD severity (Table 1.10). Two positive studies demonstrated that the BDNF rs6265 Val/Val genotype was associated with higher severity in OCD females (P=0.013-0.045; Hemmings et al., 2006) (P=0.031; Hemmings et al., 2008); but the samples likely overlapped given that they were conducted by the same group. The study by Márquez et al. (2013) that examined the same SNP was however negative. Three genes, GRIN2B, HTR2A, and DRD1 were found to be associated with OCD severity in one study (Hemmings et al., 2006) but they have not been replicated yet.

Given the scarcity of genetic studies exploring Y-BOCS severity in OCD sample in addition to the potential higher genetic loading in more severe cases of OCD, we aimed to investigate the effect of OCD candidate genes based on OCD severity.

40

41

Table 1.10. Summary Table of Genetic Studies of Y-BOCS Symptom Severity in OCD.

Gene Results Reference GRIN2B + Hemmings et al., 2006 HTR1B - Hemmings et al., 2006 HTR2A +/- Hemmings et al., 2006; Walitza et al., 2012 HTR2C - Hemmings et al., 2006 HTR6 - Hemmings et al., 2006 DAT/SLC6A3 - Hemmings et al., 2006 DRD1 + Hemmings et al., 2006 DRD2 - Hemmings et al., 2006 DRD3 - Hemmings et al., 2006 DRD4 - Hemmings et al., 2006 COMT +/- Hemmings et al., 2006; Katerberg et al., 2010 BDNF ++/- Hemmings et al., 2006; Hemmings et al., 2008; Márquez et al., 2013 ACE - Hemmings et al., 2006 ESRα - Hemmings et al., 2006 INPP-1 - Hemmings et al., 2006 PLC-γ1 - Hemmings et al., 2006 HOXB8 - Hemmings et al., 2006 + indicates significant study - indicates negative study

41

42

1.4.2.3 Psychiatric Comorbidities in OCD

Individuals suffering from OCD often present with additional comorbid psychiatric conditions. According to a recent study (Cross-Disorder Group of the Psychiatric Genomics Consortium, 2013), a set of genetic markers contribute to the risk of developing several psychiatric disorders (major depressive disorder, bipolar disorder, schizophrenia, autism spectrum disorder, and attention-deficit hyperactivity disorder), implicating a common genetic pathway with genetic markers having pleiotropic effect that may lead to different psychiatric presentations. Since genetic studies require large dataset to identify multiple common small gene effect, cross disorder studies are valuable to identify common genetic variants that predispose an individual to a group of psychiatric disorders.

OCD is highly comorbid with other psychiatric disorders that may have a negative impact on OCD outcome and increase therapeutic challenge (de Mathis et al., 2013; Ruscio et al., 2010). Studies have shown that OCD individuals with comorbid psychiatric disorders predict a poorer short- and long-term treatment outcome in addition to worse quality of life (Jakubowski et al., 2013). Large-scale OCD epidemiological studies reported a lifetime comorbidity rate ranging from 76% to 91% and a current comorbidity rate between 42% and 55% with anxiety and mood disorders as the most frequent conditions in individuals who suffer from OCD (Ruscio et al., 2010; Hofmeijer-Savink et al., 2013). Table 1.11 reported frequencies of psychiatric comorbidity across several large OCD studies (Murphy et al., 2013).

Higher rates of comorbid depression (de Mathis et al., 2011) and eating disorders (Torresan et al., 2009) were observed in OCD females (de Mathis et al., 2011) whereas a higher frequency of comorbid social phobia (Jaisoorya et al., 2009) and (de Mathis et al., 2011) was noted in OCD males. In a blinded family study comprising of 80 cases and 73 control probands in addition to 343 OCD cases and 300 control first-degree relatives, body dysmorphic disorder, , eating disorders, and pathologic “grooming” conditions including nail biting, skin picking, and trichotillomania occurred more frequently in OCD cases than controls (Bienvenu et al., 2000). In the Johns Hopkins OCD family study, the authors reported higher frequency of all anxiety and affective disorders except for bipolar disorder in OCD cases than controls (Nestadt et al., 2001), which was supported by other family studies (Carter et al., 2004).

42

43

A Brazilian family study showed higher age-corrected recurrence risks of OCD (22.7% versus 0.9%, odds ratio [OR]=32.5, 95% confidence interval [CI]=4.5-230.8, P=0.0005) and chronic tics (11.6% versus 1.7%, OR=7.9, 95% CI=1.9-33.1, P=0.005) in OCD relatives than control relatives; furthermore, the best predictor of a diagnosis of OCD was found to be a comorbid diagnosis of tics in the relatives (OR=7.35, 95% CI=3.79-14.25, P<0.0001) (do Rosario-Campos et al., 2005).

One of the largest OCD studies from the OCD Collaborative Genetics Study (OCGS) (Nestadt et al., 2009), consisting of 706 OCD individuals, attempted to sub-classify OCD based on comorbidity and the authors reported three subgroups: (1) an OCD simplex class in which only major depressive disorder (MDD) is the most frequent comorbid disorder; (2) an OCD comorbid tic-related class in which tics are highly represented and affective disorders are rarer; and (3) an OCD comorbid affective and anxiety-related class in which panic disorder, affective disorders, and obsessive-compulsive personality disorder are more frequent.

Regarding treatment implication, treatment studies reported that OCD individuals with comorbid tics (Bloch et al., 2006; Masi et al., 2013; McDougle et al., 1994; Nakamae, 2011) and family history of tics (McDougle, 1997) have improved response to antipsychotic (including typical [haloperidol, pimozide], atypical [risperidone, olanzapine, ], and partial agonist []) augmentation.

To date, seven studies have examined the genetic effect of a total of 19 genes in OCD all with comorbid major depressive disorder and tics except for one study (Table 1.12; for further details, please refer to Appendix IV Table 1). Wendland et al. (2007) used another approach and examined three markers within SLC6A4 and the BDNF Val66Met (rs6265) polymorphism in OCD with the total number of comorbid mood and anxiety disorders in addition to suicidality and experience of trauma. The authors detected significant association between the total number of comorbid anxiety disorders and 5HTTLPR (P=0.037) and SLC6A4 rs25531 (P=0.003) respectively. However, most were negative except for four genes, serotonin receptor 6 (HTR6), DRD4, COMT, and BDNF with no replications.

OCD patients with comorbid psychiatric conditions often require different treatment approaches and therefore, psychiatric comorbidities are of clinical importance in the 43

44

management of illness progression and relapse prevention. By determining the genetic basis of OCD with additional comorbid conditions, we may identify common pathways across different psychiatric disorders.

Table 1.11. Psychiatric Comorbidity of OCD (Percentage of total number of individuals reported with OCD or in the general population).

Lifetime Current General Population Comorbid Disorder Rates (%) Rates (%) Prevalence (%)

Obsessive-Compulsive and Related - - - Disorders Trichotillomania 7-36 - 1-2 Body Dysmorphic Disorder 6-12 - 1.7-1.8 Skin Picking 18 - 1.4 Hoarding 2-6 4-8 - 0.3-0.8 Any Mood Disorder 64-74 16-37 8-32 MDD 38-70 15-18 6-17 Bipolar Disorder 1-23 0.7-1 1-2 Dysthymia 6-24 0-6 3-13 Any Anxiety Disorder 46-70 37-38 15-38 Panic Disorder 6-23 7-9 4 Agoraphobia 1-18 1 5 PTSD 5-19 3 9 GAD 8-46 8-9 5-9 SAD 23-44 19 13 Specific Phobia 11-43 8-15 8-14 Any Substance Use Disorder 10-71 5-6 17-35 Any Psychotic Disorder 3-5 2-3 1.5 Any 6-19 2-5 0.4-2 Any Control Disorder 15-39 11 0.3-11 Any Somatoform Disorder 6-8 6 0.1-10 Autism Spectrum Disorder 3 - 1 ADHD 10-13 3 2.5-5 Information adapted from Hasler et al. (2007), Hofmeijer-Savink et al. (2013), LaSalle et al. (2004), Miguel et al. (2008), Nestadt et al. (2001 and 2008), Pallanti and Grassi (2014), Pinto et al. (2006), Ruscio et al. (2010), and Torresan et al. (2013) for the OCD population and Kessler et al. (1994 and 2005) and APA (2013) for the general population.

44

45

Table 1.12. Genetic Studies of Psychiatric Comorbidity in OCD. Gene Disorder Results Reference GRIN2B MDD, tics - Hemmings et al., 2006* SLC6A4 Tics, mood and anxiety disorders --- Dickel et al., 2007^; Liu et al., 2011^; Wendland et al., 2007+ HTR1B Tics, MDD -- Dickel et al., 2007^; Hemmings et al., 2006* HTR2A Tics, MDD -- Dickel et al., 2007^; Hemmings et al., 2006* HTR2C Tics -- Cavallini et al., 1998^; Hemmings et al., 2006* HTR6 MDD, tics + Hemmings et al., 2006* DAT/SLC6A3 MDD, tics - Hemmings et al., 2006* DRD1 MDD, tics - Hemmings et al., 2006* DRD2 MDD, tics - Hemmings et al., 2006* DRD3 MDD, tics - Hemmings et al., 2006* DRD4 Tics, MDD +/- Camarena et al., 2007^; Hemmings et al., 2006* COMT MDD, tics +/- Hemmings et al., 2006*; Liu et al., 2011^ MAOA MDD - Camarena et al., 2001# BDNF Tics, MDD, mood and anxiety disorders +/-- Dickel et al., 2007^; Hemmings et al., 2006*; Wendland et al., 2007+ ACE MDD, tics - Hemmings et al., 2006* ESRα MDD, tics - Hemmings et al., 2006* INPP-1 MDD, tics - Hemmings et al., 2006* PLC-g1 MDD, tics - Hemmings et al., 2006* HOXB8 MDD, tics - Hemmings et al., 2006* + indicates significant result - indicates negative result Bold style indicates significant findings with the bolded disorder(s) * Hemmings et al. (2006) investigated genetic variations between OCD with and without comorbid MDD and tics ^ Camarena et al. (2007), Cavallini et al. (1998), Dickel et al. (2007), and Liu et al. (2011) examined genetic markers between OCD with and without comorbid tics # Camarena et al. (2001) tested genetic variants between OCD with and without MDD + Wendland et al. (2007) studied genetic polymorphisms between OCD according to the total number of comorbid mood and anxiety disorders in addition to suicidality and the experience of trauma.

45

46

1.4.2.4 Importance of Familiality in OCD

Individuals with a significant family history of mental disorders have a higher risk of developing psychiatric disorders. The rationale for OCD being a genetic brain disorder initially came from family studies because OCD runs in the family. It is widely accepted that the risk of OCD is higher when there is a greater number of relatives with OCD in the family. The unadjusted aggregate risk of OCD was found to be ranging from 8.2% to 13.8% as compared to 1-3% in the general population (Cavallaro et al., 2002; Hettema et al., 2010) and several family studies consisting of 343 OCD cases and 300 control relatives reported significantly higher lifetime prevalence of OCD when compared to control relatives (11.7% versus 2.7%; Nestadt et al., 2000). The rate of OCD among relatives of adults with OCD was approximately two times that of controls but the rate dramatically increased to 10-fold among relatives of children and/or adolescents with OCD (Pauls, 2010). Family studies reported prevalence rates of OCD ranging from 7% to 15% in first-degree relatives of child and adolescent probands with OCD, which supported an increased familial loading in OCD probands with earlier AAO (Nestadt et al., 2010).

Several clinical OCD studies have reported significant differences between familial and sporadic cases of OCD. Hanna et al. (2005) indicated that familial OCD individuals presented with ordering compulsions and aberrant grooming behaviours with skin picking more frequently than the sporadic type. Furthermore, a more recent study reported that familial OCD was associated with earlier onset and greater duration of untreated illness with more compulsive symptoms such as ordering and cognitive types, higher comorbidity including depression and anxiety disorders, in addition to non-response to treatment (Viswanath et al., 2011).

To date, only three studies have examined the genetic effect of familiality or family history in OCD (Denys et al., 2006; Katerberg et al., 2009; Katerberg et al., 2010). Denys et al. (2006) reported a significant association between positive family history of OCD and the HTR2A rs6311 G allele (P=0.039-0.043) in 156 OCD patients. On the contrary, Katerberg et al. (2009; 2010) examined the BDNF rs6265 and COMT rs4680 markers respectively and did not detect any differences in allele and genotype frequencies between presence and absence of family history of obsessive-compulsive symptoms. However, no study to date has explored family

46

47

history of other psychiatric disorders in the genetics of OCD. Therefore, it is noteworthy to target OCD individuals with significant family history of OCRDs for genetic study given the higher probability of genetic loading in the familial group.

1.4.3 Pharmacogenetics of OCD

A large part of this section has been taken directly from a review paper (Zai et al., 2014; Pharmacogenomics 2014; 15(8):1147-1157) with granted copyright permission.

The role of pharmacogenetics has become increasingly important in the treatment of mental illness. Evidence suggests that inter-individual genetic variation may largely determine drug response and adverse effects. Pharmacogenetic research in major depressive disorder (MDD) and schizophrenia has provided solid background to utilize genetic data in the treatment of these disorders. Genetic factors influencing antidepressant response have been established and well-studied in MDD but not in anxiety disorders (Tiwari et al., 2009) or the newly categorized obsessive-compulsive and related disorders including obsessive-compulsive disorder (OCD).

Antidepressants are the first-line treatment of MDD, anxiety disorders, and obsessive- compulsive and related disorders. More than 40% of patients do not typically respond to the initial treatment (McDougle et al., 1993) and the duration of the initial trial may take up to six weeks for MDD and up to 12 weeks for anxiety, and obsessive-compulsive and related disorders prior to switching to another medication. Both patients and clinicians are making treatment decisions for psychiatric disorders in a trial-and-error fashion, without any particular guidance. Pharmacogenetics may significantly contribute to a better prediction of drug response and side effect tolerance by personalizing the choice of the best antidepressant for a particular patient, taking into account one’s genetic information that influences drug metabolism, drug transport and drug mechanism. This should ultimately decrease the duration of patient suffering, reduce the number of medication trials, lower health care costs, and improve patients’ quality of life.

Compared to the genetics of OCD, pharmacogenetics concept is a more recent approach in psychiatry. There are only approximately 20 published articles or abstracts reporting on the 47

48

effect of genetic variations in antidepressant response in OCD patients (Qin et al., 2015; Zai et al., 2014).

A summary and review of previous studies on the pharmacogenetics of OCD can be found in (Table 1.13). A total of 18 published articles were found in literature search using these specific search terms: genetics, pharmacogenetics, obsessive-compulsive gene response, treatment response, drug response, antidepressant response, antidepressant, OCD, obsessive- compulsive disorder, side effects, and adverse effects.

Pharmacogenetics can be separated into two categories, pharmacokinetic factors including the cytochrome P450 liver enzyme system, and pharmacodynamic factors including the brain candidate gene systems. Table 1.13 presents all pharmacogenetic studies in OCD to- date.

48

49

Table 1.13. Summary of Pharmacogenetic Studies in OCD.

Gene Polymorphism Study design Medication Sample Results Main finding Reference CYP2D6 *4 (rs3892097) 12 weeks Venlafaxine up Initially 91 patients and - No association of Van *6 (rs5030655) randomized to 300 mg or final analysis with 74 polymorphisms Nieuwerburgh *10 (rs1065852) double blind paroxetine up OCD* patients, m/f: with treatment et al., 2009 *41 (rs28371725) prospective to 60 mg 39.2%/60.8%, mean age response; plasma Response 36.3±11.6 levels of drugs definition: >25% significantly higher reduction of Y- in patients with at BOCS scores least one allele with reduced activity, plasma levels of metabolites significantly lower in cases with two alleles with reduced activity CYP2D6 AmpliChip CYP450 >10 weeks Fluoxetine, 39 patients; m/f: - No overall Müller et al., test, which detects up unless patient sertraline, 41%/59%; mean age significant 2012 to 33 alleles including reported robust fluvoxamine, 46.0±10.2; ethnicity: association 7 duplications in clinical response paroxetine, 87.5% European between CYP2D6 and 3 Response citalopram ancestry metabolizer status alleles in CYP2C19 definition: based and treatment CYP2C19 - on clinical global response but poor impression – metabolizer was improvement associated with (CGI-I) scale lack of response to antidepressants

49

50

CYP2D6 *3 (rs35742686) >10 weeks Fluoxetine, 184 patients; m/f: 71/113; - No significant Brandl et al., *4 (rs3892097) unless patient sertraline, mean age 38±11.4; association but 2014 *10 (rs1065852) reported robust fluvoxamine, ethnicity CYP2D6 non- *17 (rs28371706) clinical response paroxetine, European/others: 166/18; extensive *41 (rs28371725) Response citalopram, mean age at onset metabolism (48% definition: based venlafaxine 14±8.6; mean Y-BOCS versus 22% with on clinical global 29.8±5.9 ≥4 trials) when impression – compared with improvement extensive (CGI-I) scale metabolizers Side effect (P=0.007) and definition: without CYP2D6 non- CYP2C19 *2 (rs4244285) - significant side extensive *3 (rs4986893) effects as “well- metabolism was *17 (rs12248560) tolerated” or “mild associated with side effects” and greater venlafaxine significant side side effects effects as (P=0.022) “marked” or “severe side effects”

SLC6A4 44-bp ins/del 12 weeks Clomipramine 34 family trios with 34 - Transmission McDougle et (5-HTTLPR) prospective 150-250 mg, patients diagnosed with disequilibrium test al., 1998 (rs25531) Response fluvoxamine OCD*; mean age indicated L-allele definition: ≥35% 150-300 mg, 33.0±8.5 years (gender transmitted more improvement on fluoxetine 40- information not provided) frequently to non- the Y-BOCS and 80 mg, responders than S- final Y-BOCS sertraline 100- allele (P=0.052) scores of <16 200 mg or and a final CGI paroxetine 40- score of ‘much 60 mg improved’ or ‘very much improved’; 10 weeks Fluoxetine >60 72 OCD patients*/ 72 - No association with Billett et al., retrospective mg matched controls; mean treatment response 1997 Response or age 36.3±8.4 years definition: 25% clomipramine 50

51

reduction in >150 mg or symptom severity equivalent dose of fluvoxamine, paroxetine or sertraline 12 weeks Fluvoxamine 181 OCD patients*; m/f: - No association with Di Bella et al., prospective 200-300 mg 93/88; 92 patients treatment 2002 Response treated with response; definition: Y- fluvoxamine/191 significant time per BOCS reduction controls; mean age genotype >35% 33.4±11.5 interaction for Y- BOCS compulsions score (P=0.028) and significant time per genotype interaction in non- tic disorder patients for total Y- BOCS score (P=0.041) and Y- BOCS compulsions score (P=0.007) 12 weeks Venlafaxine up 88 OCD patients*; + S/L genotype Denys et al., prospective to 300 mg or m/f:34/54 (paroxetine associated with 2007 randomized paroxetine up N=40); mean age of better response in double-blind to 60 mg responders/non- venlafaxine-treated Response responders patients (P=0.008) definition: >25% 34.1±11.3/31.7±12.0 reduction of Y- BOCS scores 8 weeks Different SRIs 113 OCD patients and - No association Zhang et al., prospective families (information 2004 response regarding age and assessment with gender not provided) Y-BOCS scale, (further details not available) 14 weeks Clomipramine 41 OCD patients*; m/f: - No association Miguita et al., 51

52

STin2 prospective 23/18; age distribution: - No association 2011 Response age 20-30: N=22, age definition: >40% 31-40: N=13, age>41: reduction in Y- N=6 BOCS score HTR1B 861G/C 12 weeks Venlafaxine up 88 OCD patients*; - No association Denys et al., (rs6296) prospective to 300 mg or m/f:34/54 (paroxetine 2007 randomized paroxetine up N=40); mean age of double-blind to 60 mg responders/non- Response responders definition: >25% 34.1±11.3/31.7±12.0 reduction of Y- BOCS score 14 weeks Clomipramine, 41 OCD patients*; - No association Miguita et al., prospective mean dose m/f:23/18; age 2011 Response 235.5 mg distribution: age 20-30: definition: >40% N=22, age 31-40: N=13, reduction in Y- age>41: N=6 BOCS score Response Clomipramine, 239 OCD patients (60 - No association Corregiari et definition: fluoxetine, with response data); al., 2012 responders with fluvoxamine, m/f:36/24 Y-BOCS score>9 citalopram, and Sheehan sertraline, Disability paroxetine Scale>10 for ≥2 months and non- responders with <25% reduction in Y-BOCS score for ≥2 SSRIs and clomipramine for ≥8 weeks HTR2A -1438G/A 12 weeks Venlafaxine up 88 OCD patients*; m/f: - G/G genotype Denys et al., (rs6311) prospective to 300 mg or 34/54 (paroxetine associated with 2007 randomized paroxetine up N=40); mean age of better response in double-blind to 60 mg responders/non- paroxetine-treated Response responders patients (P=0.013) definition: >25% 34.1±11.3/31.7±12.0 reduction of Y- 52

53

BOCS score 8 weeks Different SRIs 113 OCD patients and + Homozygosity Zhang et al., prospective families (information associated with 2004 response regarding age and treatment response assessment with gender not provided) (P=0.030) Y-BOCS scale (further details not available) -1438G/A (rs6311) 12 weeks Fluvoxamine 58 OCD patients*; m/f: - No association Tot et al., and prospective 100-300 mg, 20/38; mean age 27±5 2003 102T/C Response fluoxetine 20- and 83 healthy controls (rs6313) definition: CGI 80 mg, rating “very sertraline 100- much” or “much 200 mg improved 102T/C (rs6313) Response Clomipramine, 239 OCD patients (60 + Higher CC Corregiari et definition: fluoxetine, with response data); genotype in non- al., 2012 responders with fluvoxamine, m/f:36/24 responders Y-BOCS score>9 citalopram, (P<0.01) and Sheehan sertraline, Disability paroxetine Scale>10 for ≥2 516C/T (rs6305) months and non- - No association responders with <25% reduction in Y-BOCS score for ≥2 SSRIs and clomipramine for ≥8 weeks 102T/C (rs6313) 14 weeks Clomipramine, 41 OCD patients*; m/f: - No association Miguita et al., 516C/T (rs6305) prospective mean dose 23/18; age distribution: - 2011 Response 235.5 mg age 20-30: N=22, age definition: >40% 31-40: N=13, age>41: reduction in Y- N=6 BOCS score BDNF rs11030096 12 weeks Fluoxetine 60- 131 OCD patients*; m/f: - No association Real et al., prospective 80 mg, 61/62; mean age after permutation 2009 Response fluvoxamine 33.3±10.5 correction rs925946 definition: >35% 200-300 mg, or - No association 53

54

rs10501087 reduction in Y- clomipramine - No association BOCS score 225-300 mg after permutation correction Val66Met - No association (rs6265) after permutation correction rs12273363 - No association rs908867 - No association of SNP, haplotype containing G-allele associated with poorer response (OR=2.27; P=0.027) rs1491850 + C-allele associated with better response (OR=2.40, P=0.005); haplotype containing T-allele associated with poorer response (OR=2.27; P=0.027) rs1491851 - No association SLC1A1 rs301434, rs301435, + Association of A- Real et al., rs3087879 allele of rs301434 2010 with pharmacological resistance in patients without SLE, association of A-G-C haplotype with pharmacological resistance in patients without SLE

54

55

COMT Details not shown 8 weeks Different SRIs 113 OCD patients and - No association Zhang et al., prospective families (information 2004 response regarding age and assessment with gender not provided) Y-BOCS scale (further details not available) Val158Met (rs4680) 14 weeks Clomipramine, 41 OCD patients*; m/f: - No association Miguita et al., prospective; mean dose 23/18; age distribution: 2011 response 235.5mg age 20-30: N=22, age definition: >40% 31-40: N=13, age>41: reduction in Y- N=6 BOCS score 10 weeks Citalopram up 64 patients; m/f: 31/33; + Association of Vulink et al., prospective to 60mg/day mean age of Met/Met with 2012 Response plus quetiapine responders/non- responders to definition: 450mg/day responders: citalopram as responder as ≥ versus 33.0±11.7/38.2±12.7; compared to non- 25% reduction in citalopram mean age at onset of responders Y-BOCS score versus placebo responders/non- (P=0.007) responders: 15.9±8.9/15.0±4.8; mean Y-BOCS of responders/non- responders: 26.3±4.3/27.8±4.7; mean HAM-A of responders/non- responders: 12.5±7.5/14.6±6.3; mean HAM-D of responders/non- responders: 8.7±4.8/9.8±3.1 36-month Different SSRIs 171 OCD (91 with drug - No association Umehara et prospective (Details not response data) patients; al., 2015 treatment study; shown) m/f: 81/90; mean age of: Response 33.8±11.0; median definition: “A” duration of illness: 6; group responders mean Y-BOCS: 25.8±6.0 55

56

as >35% reduction in Y- BOCS score with only SSRI treatment, “B” group responders as >35% reduction in Y- BOCS score with antipsychotics and SSRI treatment, and “C” as non- responders GRIN2B rs7972211 Details not SSRIs (details 1598 OCD cases from a + P=2.71×10-5 Qin et al., provided but not provided) family-based GWAS with (OR=0.65 [95% 2015 according to the ~800 cases as confidence interval primary author, responders according to 0.49-0.87]) response details the primary author were provided by patient with “yes” or “no” response to a SSRI SLC6A3 40 bp VNTR UTR 14 weeks Clomipramine, 41 OCD patients*; m/f: - No association Miguita et al., region prospective; mean dose 23/18; age distribution: 2011 30 bp VNTR intron 8 response 235.5mg age 20-30: N=22, age - VNTR intron 14 definition: >40% 31-40: N=13, age>41: - SLC6A2 1287G/A reduction in Y- N=6 - (rs5569) BOCS score DRD2 Details not shown 8 weeks Different SRIs 113 OCD patients and - No association Zhang et al., prospective families (information 2004 response regarding age and assessment with gender not provided) Y-BOCS scale (further details not available) Taq1A (rs1800497) 10 weeks Citalopram up 64 patients; m/f: 31/33; - No association Vulink et al., prospective to 60mg/day mean age of 2012 Response plus quetiapine responders/non-

56

57

definition: 450mg/day responders: responder as ≥ versus 33.0±11.7/38.2±12.7; 25% reduction in citalopram mean age at onset of Y-BOCS score versus placebo responders/non- responders: 15.9±8.9/15.0±4.8; mean Y-BOCS of responders/non- responders: 26.3±4.3/27.8±4.7; mean HAM-A of responders/non- responders: 12.5±7.5/14.6±6.3; mean HAM-D of responders/non- responders: 8.7±4.8/9.8±3.1 DRD4 Details not shown 8 weeks Different SRIs 113 OCD patients and - No association Zhang et al., prospective families (information 2004 response regarding age and assessment with gender not provided) Y-BOCS scale (further details not available) VNTR Naturalistic Different SRIs 146 patients; m/f: 98/48; - No association Viswanath et follow-up; mean age of: 28.6±9.1; al., 2013 Response median duration of definition: illness: 6; mean Y- responders as BOCS: 25.8±6.0 >35% reduction in Y-BOCS score and CGI-I of 1 or 2, partial responders as 25-35% reduction in Y-BOCS score, and non- responders as <25% reduction 57

58

in Y-BOCS score MAOA Details not shown 8 weeks Different SRIs 113 OCD patients and - No association Zhang et al., prospective families (information 2004 response regarding age and assessment with gender not provided) Y-BOCS scale (further details not available) GWAS Illumina Treatment SRIs including 804 OCD patients + • DISP1 Qin et al., HumanOmniExpress response using a clomipramine, rs17162912 2015 – 12v1 with 730,525 five-point scale of citalopram, (P=1.76×10-8; SNPs “no response”, escitalopram, OR=0.39, 95% “could not fluoxetine, CI=0.26-0.58) tolerate”, fluvoxamine, • PCDH10 “minimal”, paroxetine, rs7675822 “moderate sertraline, (P=2.86×10-6; improvement”, venlafaxine, OR=0.65, 95% and “total and CI=0.51-0.83) remission” with and rs1911877 “moderate (P=8.41×10-6; improvement” or OR=0.66, 95% “total remission” CI=0.52-0.84) as responders • GRIN2B and “no rs7972211 response” or -5 (P=2.71×10 ; “minimal OR=0.65, 95% response” as CI=0.49-0.87) non-responders • GPC6 rs17253738 (P=2.13×10-5; OR=0.59, 95% CI=0.43-0.82), rs95116369 (P=4.38×10-5; OR=0.61, 95% CI=0.44-0.84), and rs3891616 (P=8.39×10-5; OR=0.63, 95% 58

59

CI=0.46-0.87) • PLCB1 rs722665 (P=8.47×10-5; OR=1.61, 95% CI=1.25-2.08) and rs2423366 (P=0.0001) • PKC rs11158347 (P=5.18×10-5; OR=1.83, 95% CI=1.39-2.41) + indicates significant study - indicates negative study

59

60

1.4.3.1 Pharmacokinetic Factors

Numerous previous studies have examined the effects of cytochrome P450 liver enzymes in drug metabolism (Caley, 2011). CYP2D6 metabolizes most antidepressants including all tricyclic antidepressants, the serotonin norepinephrine reuptake inhibitor (SNRI), venlafaxine, and most selective serotonin reuptake inhibitors (SSRIs), predominantly fluoxetine and paroxetine. CYP2C19 metabolizes several tricyclic antidepressants and SSRIs including escitalopram and citalopram. Furthermore, genetic variations in CYP2D6 and CYP2C19 alter the metabolism of their substrates (i.e., medications), thereby affecting drug efficiency and tolerability. Although other cytochrome P450 enzymes such as CYP1A2, CYP2B6, CYP2C9, and CYP3A4 may also be important to the metabolism of antidepressants including tricyclic antidepressants, bupropion, trazodone, and some SSRIs (of variable degree), very few pharmacogenetic studies of these enzymes and antidepressant response have been published in the literature (Altar et al., 2013) and none has been investigated in OCD. Therefore, we have limited this review to CYP2D6 and CYP2C19.

The genotypes of CYP2D6 allow for a prediction of its enzymatic activity levels in the absence of confounding factors such as co-medication. It is grouped into four different metabolism classifications including “poor” with no CYP2D6 activity, “intermediate” with metabolism between poor and extensive, “extensive” with normal activity, and “ultra-rapid” with an increased metabolism (Jann et al., 2000). Having a poor metabolism status indicates that the individual will metabolize the substrate (i.e., SSRIs, venlafaxine) at a very low rate. This can in turn increase side effects given the higher cumulative drug concentration that remains in the individual’s body. On the other hand, an individual with an ultra-rapid metabolism status will break down the substrate (i.e., SSRIs, tricyclic antidepressants) at a fast rate, resulting in low cumulative serum concentration of the medication and likely a sub-therapeutic level. For metabolism status, on average, 10% of the general population are poor metabolizers, 35% are intermediate metabolizers, and 7% are ultra-rapid metabolizers. However, ethnicity plays an important factor in the CYP2D6 metabolism variability (Neafsey et al., 2009). The prevalence of CYP2D6 poor metabolizers ranges from approximately 2-10% in Caucasians, 2-7% in Hispanics, and 1-2% in Asians, Africans, and African Americans (Bernard et al., 2006).

60

61

Regarding CYP2C19, the same classifications have been used based on the level of enzyme activity. Similarly, inter-ethnic differences have been observed in CYP2C19 metabolism, with Asians having a much higher prevalence (ranging from 12-23%) of poor metabolizers when compared to Caucasian (1-6%) and Black African (1-7.5%) populations (Desta et al., 2002).

Recently, two studies from our group reported influence of genetic variations across these two liver enzymes in SSRI response and OCD (Brandl et al., 2014; Mueller et al., 2012). The first study by Müller and colleagues (Mueller et al., 2012) included 39 patients with OCD who were assessed for SSRI antidepressant response (including fluoxetine, sertraline, fluvoxamine, paroxetine, and citalopram) based on the Clinical Global Impression – Improvement (CGI-I) scale. Although no overall significant association was detected between CYP2D6 or CYP2C19 metabolizer status and treatment response, impairment in CYP2D6 function (poor metabolizer status) was associated with lack of response to SSRIs. The second study (Brandl et al., 2014) investigated the association of CYP2D6 and CYP2C19 status with treatment response and side effects in 184 OCD subjects. Although this study did not show significant association of metabolism status with response to specific medications, the authors noted that patients with CYP2D6 non-extensive metabolism had a significantly increased number of failed SSRI trials when compared to extensive metabolizers (48% of non-extensive metabolizers had required ≥4 trials versus 22% of extensive metabolizers with ≥4 trials; P=0.007). Furthermore, the use of venlafaxine in individuals with CYP2D6 non-extensive metabolism was significantly associated with the occurrence of side effects (P=0.022).

To summarize, only two cytochrome P450 liver enzyme genes, CYP2D6 and CYP2C19 have been studied in antidepressant response and OCD. The genetic status of CYP2D6 appears to modify antidepressant response in OCD patients with non-responders being more common among non-extensive metabolizers. However, the sample size of these studies is relatively small and further validation of these findings is required.

61

62

1.4.3.2 Pharmacodynamic Factors

Pharmacodynamic factors affect drug actions at a molecular level within our brain. Psychotropic medications have various mechanism of actions for which numerous neurotransmitter systems are involved. The most well-known and most studied systems consist of the serotonergic, glutamatergic, and dopaminergic systems, which interact with one another and are associated with drug response and adverse effects of medications (Gvozdic et al., 2012). In this section, we present a general overview of each of the above mentioned systems in addition to other systems of interest. With regard to brain candidate genes across these systems and pharmacogenetics of OCD, there has been little development in this area recently.

1.4.3.2.1 Serotonergic system genes

The serotonergic system has been postulated to be involved in the pathogenesis of OCD given that SSRIs are the first-line treatment of OCD based on extensive clinical trials (Koran et al., 2007). SSRIs increase the level of serotonin in the synaptic cleft by inhibiting its reuptake into the presynaptic cells, thereby allowing more serotonin to bind to the postsynaptic receptors. This is thought to be the mechanism by which these medications reduce depression and anxiety, and engage in complex interaction with other neurotransmitter systems. Many genes that are involved in serotonin metabolism, transport, storage, and signalling cascade, have been studied extensively over the past decade in relation to OCD (Pauls, 2010). Candidate genes in the serotonergic system implicated in OCD include: serotonin transporter (SLC6A4) and its promoter (HTTLPR), serotonin 2A receptor (HTR2A), serotonin 2C receptor (HTR2C), serotonin 1B (1D beta) receptor (HTR1B), and tryptophan hydroxylase (TPH).

Most pharmacogenetic studies of antidepressant response came from MDD and findings for HTR2A appear to be most robust (Fabbri et al., 2013; Singh et al., 2013). The study by Corregiari et al. (Corregiari et al., 2012) consisted of 239 OCD patients and of those, 60 had antidepressant response data. Corregiari and colleagues (Corregiari et al., 2012) investigated one SNP across the HTR1B gene, 861G/C (rs6296), and two SNPs within the HTR2A gene, 102T/C (rs6313) and 516C/T (rs6305). Six different antidepressants were examined including clomipramine, fluoxetine, fluvoxamine, citalopram, sertraline, and paroxetine. The authors 62

63

defined responders as having a Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score of less than 9 and a Sheehan Disability Scale score of less than 10 after treatment with an SRI and being stable for at least two months. Patients with less than 25% reduction in the Y- BOCS severity score after at least two different SSRIs and clomipramine for at least 8 weeks were classified as non-responders. This study reported significantly higher CC genotype and C allele of the HTR2A 516C/T marker in non-responders (P<0.01 and P=0.01 respectively). The authors also investigated the relationship between these polymorphisms, study groups (responders, non-responders, and healthy controls), and endocrine and serotonin measures (serotonin, prolactin, growth hormone, and cortisol basal levels); they observed significantly higher cortisol and prolactin levels in response to citalopram treatment in individuals with the CC genotype of the HTR1B 861G/C marker (P<0.01 and P<0.001 respectively). This was the first study examining genetic association of endocrine response to SRI treatment and the authors concluded that serotonergic system genes may play an important role in influencing SRI response and regulating endocrine and serotonin level.

Overall, there were a limited number of OCD pharmacogenetic studies with relatively small sample size, providing consistent but not yet conclusive results implicating HTR2A and HTR1B in SRI antidepressant response in OCD.

1.4.3.2.2 Glutamatergic system genes

There is increasing evidence of the influence of the glutamatergic system on the etiology of many neuropsychiatric disorders including OCD (Kariuki-Nyuthe et al., 2014). Support for involvement of this system in the pathogenesis of OCD comes from research implicating the complex neurotransmitter network, the CSTC circuit, in the underlying neuro-anatomical basis of OCD; altered glutamatergic function is seen as one of the major drivers of observed over- activity in this brain circuit in OCD (Milad et al., 2012). Glutamate has also been implicated in this disorder based on evidence from magnetic resonance spectroscopy studies, knock-out mouse models, clinical studies of glutamatergic agents, and candidate gene association studies (Wu et al., 2012). We briefly summarize the evidence supporting glutamate involvement in the

63

64

neurobiology and pathoetiology of OCD below, but also refer interested readers to a recent extensive review article (Wu et al., 2012).

Glutamatergic medications such as phenobarbital, lamotrigine, topiramate, , N- acetylcysteine (NAC), , ketamine, and D-cycloserine, have been used as an augmenting agent in the management of OCD (Grados et al., 2013). In particular, adjunctive topiramate and riluzole are considered third-line pharmacotherapy for OCD according to the Canadian Psychiatric Association anxiety disorders clinical practice guideline in 2006 (CPA, 2006), and use of glutamate antagonists such as riluzole and topiramate was also recommended by the American Psychiatric Association (Koran et al., 2007). Of interest, new glutamatergic modulating drugs have been developed by pharmaceutical companies, namely Eli Lilly (mGluR2/3 agonist) and Roche (glycine transporter inhibitor). These medications have had mixed results in clinical trials in schizophrenia, and are currently being investigated in OCD. A pilot clinical trial of the glycine transporter inhibitor (Sarcosine) has provided benefits in drug- naïve OCD patients but not in patients who were treatment-resistant or had prior SSRI trial(s) (Wu et al., 2011).

A very recent GWAS of drug response was conducted by Qin et al. (Qin et al., 2015) with 804 OCD cases. GWAS was performed using the Illumina HumanOmniExpress – 12v1 with 730,525 SNPs (Illumina Inc. ®, San Diego, CA, USA). The following is a list of the SRI antidepressants that were included in this GWAS: clomipramine, citalopram, escitalopram, fluoxetine, fluvoxamine, paroxetine, sertraline, venlafaxine, and duloxetine. Drug response status was determined using a five-point scale of “no response”, “could not tolerate”, “minimal response”, “moderate improvement”, and “total remission”. Responders were defined as those who have scored “moderate improvement” or “total remission” and non-responders refer to those who reported “no response” or “minimal response”. The authors concluded that one of the top SNPs was the rs7972211 marker near GRIN2B, which showed modest significant association with SSRI response (OR=0.65 [95% confidence interval 0.49-0.87], P=2.71×10-5). Moreover, three SNPs (rs17253738, rs9516369, and rs3891616) across the glypicans 6 (GPC6) gene detected strong signals of association with SSRI response. GPC6 functions to promote glutamate receptor clustering and receptivity in addition to inducing post-synaptic activity (Allen et al., 2012).

64

65

Only one candidate gene study examined a glutamatergic system gene, SLC1A1, in SRI response in OCD and Real et al. (2010) reported a significant association between the rs301434 A allele and non-response to fluoxetine, fluvoxamine, or clomipramine response in a 12-week prospective study with 131 OCD patients.

With clinical support from utilizing glutamatergic agents in the treatment of OCD and research evidence of glutamatergic system genes from genetic association studies, GWAS, and pharmacogenetics studies, further investigation of glutamatergic pathway genes is crucial in determining the etiology and management of OCD. As previously stated, only one study with positive finding has implicated a role of SLC1A1 in OCD treatment response (Real et al., 2010) but a recent GWAS implicated an enrichment of glutamate-related genes in SRI response (Qin et al., 2015). However, no published pharmacogenetic studies have investigated glutamate medication response in OCD to date.

1.4.3.2.3 Dopaminergic system genes

Disruption of the complex interactions across the CSTC circuit has been implicated in the pathogenesis of OCD, and dopaminergic neurotransmission is an important player in this complicated network (Pauls, 2010). Further evidence of dopamine influence is provided by neuroimaging, neurochemical, and candidate gene studies (Koo et al., 2010). Importantly, antipsychotic medications have established utility as augmentors to SRI antidepressants in the treatment of OCD (CPA, 2006).

Dopaminergic system genes have thus received some interest in genetic association and pharmacogenetic studies of OCD. Only three pharmacogenetic studies, examining the dopaminergic genes in OCD, dopamine D2 receptor (DRD2) and dopamine D4 receptor (DRD4), have been conducted and no associations were detected (Viswanath et al., 2013; Vulink et al., 2012); however, a COMT functional SNP was found to be associated with citalopram response (Vulink et al., 2012). However, a previous published study did not detect any significant association between SRI response and DRD2, DRD4, COMT, and MAOA in OCD (Zhang et al.,

65

66

2004). Two additional studies failed to demonstrate positive results between COMT and SRI response in OCD (Miguita et al., 2011; Umehara et al., 2015).

Vulink et al. (Vulink et al., 2012) (Table 1) examined the DRD2 Taq1A (rs1800497) and COMT Val158Met (rs4680) polymorphisms, with interesting results. This study recruited 64 OCD patients who were randomly assigned to citalopram plus quetiapine versus citalopram plus placebo. Although no significant differences were observed in genotype distributions or allele frequencies of DRD2 or COMT between responders and non-responders to adjunctive quetiapine treatment, 48% of the responders to citalopram had COMT Met/Met genotype as compared to 0% of the non-responders (χ2=10.06, P=0.007). The authors concluded that COMT Val158Met Met/Met genotype may predict better response to citalopram treatment.

A recent naturalistic study by Viswanath et al. (Viswanath et al., 2013) examined 146 OCD patients and the influence of the DRD4 VNTR polymorphism in SRI response. However, the authors did not detect any significant association.

Once again, inconsistent reports of the involvement of dopaminergic system genes suggest that further investigations are needed with significantly larger sample size in antidepressant response in OCD.

1.4.3.2.4 Other candidate genes that have not been examined in pharmacogenetics

Numerous additional candidate systems or genes have emerged in the recent years in the genetics of OCD, including the neurotrophic system genes, the neuronal-related genes, and additional top hits from the GWAS (Stewart et al., 2013) that were not previously known to be involved in neuropsychiatric disorders.

Neurotrophic factor genes such as the brain-derived neurotrophic factor (BDNF) and neurotrophin-3 receptor (NTRK3) have frequently been hypothesized in the etiology of many neuropsychiatric disorders in addition to OCD (Pauls, 2010). BDNF has been studied extensively and although a recent meta-analysis detected no significant association of its functional Val66Met polymorphism and OCD (Taylor, 2013), it is still considered an interesting 66

67

gene given its function in neurodevelopment, neurogenesis, and neuroplasticity (Autry and Monteggia, 2012). Very limited pharmacogenetic studies to-date, examining BDNF and other related genes, have revealed inconsistent and limited findings (Zai et al., 2014).

Neuronal-related genes such as the myelin-oligodendrocyte glycoprotein (MOG) (Zai et al., 2004) and the oligodendrocyte lineage transcription factor 2 (OLIG2) (Stewart et al., 2007) genes were examined in genetic association studies of OCD with promising findings. However, the MOG tetra-nucleotide repeat finding has not been studied for replication purpose and we (Zai et al., 2012) could not replicate the significant results in OLIG2 from Stewart et al. (Stewart et al., 2007).

Many new loci were identified as top hits in the recent GWAS of OCD (Stewart et al., 2013); however, further research is needed to clarify their functional status and their potential role in the pathoetiology of OCD and antidepressant response in OCD.

The GWAS (Qin et al., 2015) (previously described in the glutamatergic system genes section) detected significant top-hits with antidepressant response including: dispatched homolog 1 (drosophila) (DISP1) rs17162912 (P=1.76×10-8), ankyrin-repeat and fibronectin type III domain containing 1 (ANKFN1) rs9303380 (P=1.03×10-6), arrestin domain containing 4 (ARRDC4) rs12437601 (P=1.66×10-6), T-cell lymphoma invasion and metastasis 1 (TIAM1) rs16988159 (P=2.48×10-6), protocadherin 1D (PCDH1D) rs7676822 (P=2.86×10-6), uncharacterized LOC730101 (LOCT30101) rs723815 (P=3.50×10-6), protocadherin 10 (PCDH10) rs1911877 (P=8.41×10-6). Several genes are involved in neuronal cell-cell adhesion including the DISP1 gene (Roessler et al., 2009), which is located on a previously known micro- deletion region for mental retardation on chromosome 1q41-q42 (Shaffer et al., 2007), PCDH10, which was previously implicated in autistic spectrum disorder (Redies et al., 2012), and TIAM1 (Mack et al., 2012), and neurotrimin (NTM) (Liu et al., 2004).

67

68

1.4.3.2.5 Pharmacogenetics summary

No definitive results support a single genetic variation or gene that determines antidepressant response in OCD; however, several SNPs across candidate genes within the cytochrome P450, glutamatergic, and serotonergic systems appear interesting (Zai et al., 2014). These markers include: cytochrome P450 2D6 (CYP2D6), SLC6A4, HTR1B, HTR2A, SLC1A1, and GRIN2B. A recent GWAS of antidepressant response in OCD revealed one genome-wide significant hit for the marker near the dispatched homolog 1 (Drosophila) (DISP1) gene, which was previously implicated in neurological development (Qin et al., 2015). Future directions should focus on expanding the sample size and fine mapping these regions in order to refine our understanding of the functional details of the genetic variation and their relation to the pathophysiology of OCD.

Interest in the pharmacogenetics of OCD has increased over the past decade with a number of international research groups examining genetic factors that determine antidepressant response in OCD. Given the high prevalence and morbidity of OCD, however, there remains a relative paucity of research in pharmacogenetics of OCD and in OCD in general. Although there is no definitive data supportive a single genetic variation or gene that determines antidepressant response in OCD, many SNPs across candidate genes within the cytochrome P450, glutamatergic, and serotonergic systems appear interesting and potentially involved (Table 1.13). Future directions should focus on expanding the sample size and fine mapping these regions in order to refine our understanding of the functional details of the genetic variation and their relation to the pathophysiology of OCD. Neuropsychiatric disorders are likely the result of subtle changes in highly complex neurotransmitter systems. As such, far more extensive functional and pharmacological studies interrogating these systems more deeply are warranted to delineate the effects of genetic variations in antidepressant response.

There are several general limitations in these published pharmacogenetic studies including small sample size, retrospective study design, and a lack of details regarding subjects’ treatment history. Additional larger and well-characterized OCD samples in a prospective study design will increase statistical power to detect true genetic variant(s) that may predict antidepressant response in OCD and its related disorders. Encouraging international groups to join together in collaboration provides a framework to increase sample size, to assess for more 68

69

detailed treatment characteristics (i.e., drug-naïve, drug-free, or treatment resistant), to refine studied phenotypes, to delineate more homogeneous subgroups, and for sharing expertise. Recent advances in technologies such as GWAS, DNA/RNA/whole exome sequencing, and fine mapping allow researchers to conduct more powerful investigations of genetic variations across the entire genome. The recent ENCyclopedia Of DNA Elements (ENCODE) project has explored large regions between genes, and elucidated the importance of a majority of these stretches of non-expressed DNA sequence – previously known as “junk” DNA. These regions contain functional elements and regulators of a large number of known genes, thereby providing a new frame-work for researchers to examine DNA sequence outside of recognized coding regions (Kavanagh et al., 2013). It is easy to imagine that multiple genes will be involved in any given patient for their medication response and side effects, thus development of prediction algorithms that include markers within genes and markers at more remote locations that regulate gene activity will no doubt be necessary. Computational and bioinformatics advancements to address the complex facets of each patient will be required, since it is unlikely that a single clinician will be able to assemble and interpret all the pharmacokinetic, pharmacodynamic, and environmental factors as well as their interactions.

In addition to the strategy of narrowing to more homogeneous subgroups within OCD, there is merit in examining pharmacogenetic associations across disorders. For example, the genes involved in antidepressant response in depression or anxiety may well have similar effects in OCD. Furthermore, studies combining all psychiatric disorders, especially across obsessive- compulsive and related disorders that undergo antidepressant treatment may provide a platform to examine rare variant(s) for the prediction of treatment response and adverse effects by increasing the sample size and power to detect genes of small effect. Identification of genetic variations that predict response and tolerability of medications will likely lead to personalized medicine in the relatively near future, with attendant improvement in clinical outcome and reduction in morbidity and mortality.

69

70

1.5 Genes of Interest

This thesis utilized two different genetic approaches to examine OCD phenotypes. GWAS examines common genetic variants across the entire genome. The GWAS utilized genetic data generated from the genome-wide Illumina Human610-Quadv1_B SNP BeadChip array (Illumina Inc. ®, San Diego, CA, USA). The array uses tag SNPs to the genomic coverage of 924,000 randomly selected SNPs, which offer 80% of genomic coverge at a maximum r2 of 0.9 (Illumina Inc. ®). The required sample size in a case-control study for 80% power at an odds ratio risk of 1.3 with a minor allele frequency of 5% is 2,606 for this array (Illumina Inc. ®). It also offers an additional ~60,000 CNV-specific markers.

The second genetic approach that was conducted in this thesis is using candidate genes with specific rationale, which support the examination of the chosen genes. The rationales for choosing the following 17 genes of interest are described below and briefly summarized in Table 1.14.

As mentioned previously in the genetics and pharmacogenetics sections of this chapter, several important neurotransmitter systems have been consistently implicated in the genetic etiology of OCD. These systems include: glutamatergic, serotonergic, dopaminergic, and others such as neurotrophic factors and additional newly identified genes of unknown connection to OCD.

70

71

Table 1.14. Summary of Candidate Gene Studies of OCD.

Candidate Meta- SRI Gene Chromosome GWAS Gender AAO Y-BOCS Familiality Comorbidity Reference Gene Studies Analysis Response MOG 6p22.1 + + Zai et al., 2004 HTR1B 6q14.1 * - + + --- Corregiari et al., 2012; Denys et al., 2007; Mas et al., 2014; Miguita et al., 2011; Taylor, 2013 DLGAP2 8p23.3 GRIK2 6q16.3 ++ # Delorme et al., 2004; Sampaio et al., 2011; Stewart et al., 2013 SLC1A1 9p24.2 * +/- + + Stewart et al., 2013; Real et al., 2010; Taylor, 2013 BDNF 11p14 * -- ++ + +++ + + Hall et al., 2003; Katerberg et al., 2009; Márquez et al., 2013; Real et al., 2009; Rocha et al., 2010; Taylor, 2013; Tukel et al., 2014; Zai et al., 2015 GRIN2B 12p13.1 * - + + # Alonso et al., 2012; Qin et al., 2015; 71

72

Taylor, 2013 FAIM2 12q13.12 # Stewart et al., 2013 HTR2A 13q14.2 * + + +++-- Corregiari et al., 2012; Denys et al., 2007; Dickel et al., 2007; Miguita et al., 2011; Taylor, 2013; Tot et al., 2003; Zhang et al., 2004 SLITRK5 13q31.2 SLC6A4 17q11.2 *+ + Cengiz et al., 2015; Taylor, 2013 DLGAP1 18p11.31 - # - - + - Li et al., 2015; Stewart et al., 2013 FUT2 19q13.33 # Stewart et al., 2013 BTBD3 20p12.2 # Stewart et al., 2013 OLIG2 21q22.11 ++/- + Stewart et al., 2007; Zai et al., 2012; Zhang et al., 2015 COMT 22q11.21 * + + + Alsobrook et al., 2002; Taylor, 2013; Vulink et al., 2012 MAOA Xp11.3 * - Taylor, 2013 + indicates significant study - indicates negative study * indicates previous meta-analysis # indicates trend association with OCD 72

73

Glutamatergic System Genes:

Five glutamate-related candidate genes were selected for the genetics of OCD phenotypes and pharmacogenetics given their previous significant association with OCD diagnosis.

Glutamate receptor ionotropic kainite 2 (GRIK2) [MIM 113505] is the tenth top hit gene reported in the recent GWAS of OCD (Stewart et al., 2013). Functional studies have implicated its role in RNA editing of glutamate receptor 2 and 6 (Paschen et al., 1994) in addition to induction of long-term potentiation in the hippocampus (Contractor et al., 2001). It is mapped on chromosome 6q16.3 (Paschen et al., 1994) and two studies have implicated its involvement in the genetic etiology of OCD. Delorme et al. (2004) conducted a case-control study in 156 OCD patients and 141 controls in addition to 124 trios and detected a significant under-transmission of the I867 allele of the rs2238076 polymorphism in OCD (P<0.03). Sampaio et al. (2011) investigated the association of OCD and GRIK2 in 47 OCD small nuclear families and reported significant finding at the rs1556995 polymorphism (permutated P=0.03) and the two-marker rs1556995-rs1417182 haplotype (permutated P=0.01) with OCD diagnosis. Furthermore, significant associations of this gene have been reported in clozapine-induced obsessive- compulsive symptoms (Cai et al., 2013), autism (Jamain et al., 2002; Kim et al., 2007; Shuang et al., 2004), schizophrenia (Bah et al., 2004; Shibata et al., 2002), Huntington disease (Rubensztein et al., 1997) and mental retardation (Cordoba et al., 2015; Motazacker et al., 2007). Given the convergence of previous significant candidate gene studies and trend for association in the first GWAS of OCD in addition to the lack of study examining OCD phenotypes, further genetic analysis of this gene in OCD phenotypes was explored.

Neuronal glutamate transporter (SLC1A1) [MIM 133550] has been most consistently implicated in OCD and a meta-analysis investigating nine single nucleotide polymorphisms (SNPs) across the 3’ region of this gene supported its involvement in 306 OCD cases and 634 healthy controls (Stewart et al., 2013) although an earlier meta-analysis was negative (Taylor, 2013). A weak association was reported for the rs301443 variant (P=0.046; non-significant corrected P) with OCD and a slightly stronger association for the rs12682807 marker with OCD males only (P=0.012; non-significant corrected P) (Stewart et al., 2013). This gene has been mapped to chromosome 9p24.2 (Smith et al., 1994), which has been implicated as a 73

74

susceptibility region in OCD from previous linkage studies (Walitza et al., 2010). It encodes a high-affinity neuronal glutamate transporter excitatory amino acid carrier 1 (EAAC1) and is highly expressed in brained regions within the CSTC circuit (Kanai et al., 2004; Rothstein et al., 1994), which has been consistently implicated in the pathoetiology of OCD (Paul et al., 2014). In addition to its association with OCD diagnosis, Wu et al. (2013) detected a significant association of the rs10491734 polymorphism and two four-marker (rs10491734-rs2228622- rs301430-rs301443 A-A-C-C and A-G-C-C) haplotypes with early-onset OCD. The rs3056 polymorphism of this gene was reported to be associated with increased total (P=0.01), left (P=0.02), and right (P=0.02) thalamic volume in OCD patients (Arnold et al., 2009). Furthermore, one pharmacogenetic study reported a significant association between the rs301434 A allele and non-response to SRI response in 131 OCD patients (Real et al., 2010). Thus, additional genetic investigation of this gene is warranted to identify its role in the susceptibility risk of developing OCD and drug response in OCD.

Glutamate receptor, ionotropic N-methyl D-aspartate (NMDA) 2B (GRIN2B) [MIM 138252] is located on chromosome 12p13.1 (Endele et al., 2010) and encodes its receptor, which can be found at excitatory synapses throughout the central nervous system. It has important function in modulating synaptic plasticity, learning, and memory (Chen et al., 1999), maintaining long-term potentiation (Thomas et al., 1996), regulating neuronal morphogenesis (Setou et al., 2000) and cell death (Hardingham et al., 2002). Significant association of this gene was first reported by Arnold et al. (2004) for the rs890 marker (P=0.014) and the 5072G-5988T (rs890- rs1805502) haplotype (P=0.002). The same research team later reported a significant association of this gene with left orbitofrontal cortex (P=0.04) and right anterior cingulate cortex (P=0.02) volume in OCD patients (Arnold et al., 2009). In addition to OCD diagnosis and brain volume, Alonso et al. (2012) examined this gene and OCD subphenotypes and reported the following significant associations between: rs1805476 and OCD males (P=0.002), 4-SNP (rs1805476- rs1805501-rs1805502-rs1805477) haplotype and presence of contamination obsessions and cleaning compulsions (permutated P=0.023). Moreover, a recent GWAS detected association between the rs7972211 marker and SRI response (P=2.71×10-5). Although this gene has been implicated in OCD, only few studies have investigated its involvement in the pathogenesis of OCD.

74

75

SLIT and NTRK-like family member 5 (SLITRK5) [MIM 609680] is one of the top hits from the recent OCD GWAS (Stewart et al., 2013) and it is mapped on chromosome 13q31.2 (Aruga et al., 2003; Nagase et al., 1998). It was first implicated in OCD because the SLITRK5 knockout (deficient) mice were found to have impaired corticostriatal circuitry and develop obsessive- compulsive-like behaviours including excessive self-grooming and increased anxiety in the open maze test (Shmelkov et al., 2010). A member of the SLITRK family, SLITRK1 has been linked to Tourette’s syndrome (Abelson et al., 2005). This SLITRK gene family encodes integral transmembrane proteins with two N-terminal leucine-rich repeat domains, which are similar to those of SLIT proteins. Most of its members including SLITRK5 also have C-terminal regions that share homolog with neurotrophin receptors and are expressed predominantly in neuronal tissues. Aruga and Mikoshiba (2003) reported that SLITRK5 may have neurite-modulating activity but its exact function remains unknown. Given its functional implication and lack of genetic association studies in OCD, this gene has been studied in this thesis.

Discs large (drosophila) homolog-associated protein 1 (DLGAP1) [MIM 605445] is the fifth top hit gene and the strongest glutamatergic gene in the first published GWAS of OCD (Stewart et al., 2013). It maps on chromosome 18p11.31 and its encoded protein interacts with other members of synaptic ion channel clustering proteins, which can mediate the clustering of NMDA receptors and Shaker-type potassium channels (Kim et al., 1997; Satoh et al., 1997). DLGAP1 plays an important in the post-synaptic scaffolding of neuronal cells and synaptic transmission (Satoh et al., 1997). Only one association study has examined this gene more closely in OCD. Li et al. (2015) recently investigated this gene for OCD susceptibility and subphenotype risk in 320 Chinese Han OCD patients and 431 healthy controls. The author did not detect significant differences in allele or genotype frequencies of the rs11081062 polymorphism or when classifying by sex, AAO, and comorbidity; nonetheless, this marker was associated with the presence of contamination obsessions (P=0.021) and cleaning compulsions (P<0.001). In addition to OCD, studies have found positive association of this gene with schizophrenia (Li et al., 2013). Due to the limited number of genetic studies investigating DLGAP1 with potential functional implication, we attempted to explore this gene further in OCD phenotypes.

Discs large (drosophila) homolog-associated protein 2 (DLGAP2) [MIM 605438] is another glutamate-related gene, which is from the same family as DLGAP1 with similar functions (Ranta

75

76

et al., 2000). This gene is located on chromosome 8p23.3, which has been implicated in autism spectrum disorders (Chen et al., 2013; Pinto et al., 2010). A previous imaging genetic study reported the strongest trend between two SNPs in DLGAP2 (rs6558484 and rs7014992) and orbitofrontal cortex white matter volume (Wu et al., 2013). No study to date has examined this gene in OCD risk.

Serotonergic System Genes:

Three serotonin-related genes were chosen for the genetic study of OCD phenotypes. For the above list of genes, SLC6A4 appears to be most consistently implicated in antidepressant response in MDD, followed by HTR2A and HTR1A (Fabbri et al., 2013).

Serotonin 1B receptor (HTR1B) [MIM 182131] maps to chromosome 6q14.1 (Jin et al., 1992) and encodes a terminal auto-receptor that is involved in the regulation of serotonin release with wide physiological functions. The acute administration of non-selective (i.e., mCPP) or selective (i.e., sumatriptan) ligands of this receptor induces a transient worsening of OCD symptoms (Slassi, 2002). Evidence for its role in OCD came from significant genetic association studies with the first reported by Mundo et al. (2000). The authors observed a preferential transmission of the G allele of the rs6296 (G861C) polymorphism to OCD probands in 67 OCD families, which was confirmed in a subsequent family-based association study by the same research team (P=0.023) in 157 OCD families. Although a recent meta-analysis of the same marker in 11 datasets did not yield significant result (P=0.232) (Taylor, 2013), Mas et al. (2014) conducted a transmission disequilibrium study in 84 OCD trios and found that another marker, rs2000292, within this gene was associated with early-onset OCD males (P=0.0006) but not in females. Only three pharmacogenetics studies have been published, examining the rs6296 marker in SRI response in OCD and although they were all negative (Corregiari et al., 2012; Denys et al., 2007; Miguita et al., 2011), this gene has been implicated to predict antidepressant response based on an animal behavioural model using the forced swim test (Redrobe et al., 1996). Therefore, it is vital to examine other genetic variations within this gene in OCD phenotypes and SRI response.

76

77

Serotonin 2A receptor (HTR2A) [MIM 182135] encodes another abundant serotonin receptor with important role in many physiological processes including sleep, appetite, thermoregulation, pain perception, hormone secretion, and sexual behaviour. Abnormalities in the serotonergic neurotransmitter system have consistently been implicated in the etiology of OCD in addition to other psychiatric disorders. It is mapped to chromosome 13q14.2 (Sparkes et al., 1991) and has a specific role in modulating conflict anxiety, which is consistent with the cortical or “top-down” model of cognitive influence (Weisstaub et al., 2006). The initial genetic support of a genetic association was reported by Enoch et al. (1998) who found an association between the A allele of the rs6311 (G-1438A) promoter polymorphism and OCD. The significant results were later replicated in 55 OCD children/adolescents and 223 healthy controls by another group (Walitza et al., 2002). A family-based transmission disequilibrium study examined the same SNP in 54 OCD trios and detected nominal significant association in OCD subjects with comorbid tic disorder (P=0.05) (Dickel et al., 2007). Furthermore, this association was also confirmed with the meta-analysis in 19 datasets (mean OR 1.219, 99th percentile CI 1.037-1.433, P=0.002) (Taylor, 2013). Thus, markers across this gene should be further examined in OCD.

The serotonergic system genes have been widely studied in the pharmacogenetics of major depressive disorder (MDD) and the most robust findings supported a role of HTR2A in antidepressant response of MDD (Fabbri et al., 2013; Singh et al., 2013). Regarding OCD pharmacogenetics, five studies have examined HTR2A. Zhang et al. (2004) first reported that the homozygous genotypes of rs6311 (G-1438A) was associated with SRI response (P=0.030), which was supported by another study (Denys et al., 2007) observing a significant association between the G/G genotype and better paroxetine response (P=0.013); however, an earlier study detected no association (Tot et al., 2003). Conflicting results for other SNPs have been published. Miguita et al. (2011) found no significant association between clomipramine response and rs6313 or rs6305 in 41 OCD patients but Corregiari et al. (2012) detected significant association between the C/C genotype of rs6313 and non-responders (P<0.01) in 239 OCD patients. Therefore, there is a need to clarify the role of this gene in OCD SRI response.

Serotonin transporter (SLC6A4) [MIM 182138] encodes a high-affinity, sodium (Na+) and chloride (Cl-) dependent transporter that localizes in the presynaptic neuronal membranes, and is highly expressed in the central nervous system. SLC6A4 actively clears serotonin from the

77

78

synaptic cleft, which prevents serotonin from attaching to other serotonin receptor for down- stream actions and recycles it to the neurotransmitter pool (Ramamoorthy et al., 1993). It is the main target of SSRI antidepressants. This gene is localized to chromosome 17q11.2 (Ramamoorthy et al., 1993) and Lesch et al. (1996) reported functional significance between the variant of the 5HTTLPR polymorphism – the long allele associated with twice the basal activity than the short allele. The long allele was reported to be twice as common in OCD patients (N=169) than healthy controls (N=253), which was replicated to have a 2-fold over-transmission to OCD patients in a family association study with 175 OCD trios (Hu et al., 2006). Many studies have examined its effect on OCD risk and a meta-analysis of this long-short marker (5HTTLPR) was positive when combining 8 OCD datasets (mean OR 1.251, 99th percentile CI 1.048-1.492, P=0.001) (Taylor, 2013). Furthermore, another very recent small case-control study consisting of 80 OCD patients and 100 healthy controls reported that the G allele of another SNP, rs16965628, was associated with risk of OCD (P=0.02) (Cengiz et al., 2015).

Dopaminergic System Genes:

Catechol-O-methyltransferase (COMT) [MIM 116790] is localized to chromosome 22q11.21 (Brahe et al., 1996; Grossman et al., 1992). Its encoded protein, COMT, is an enzyme that catalyzes the transfer of methyl group from S-adenosylmethionine to catecholamines including the common neurotransmitters of dopamine, epinephrine, and norepinephrine. Karayiorgou et al. (1997; 1999) reported significant associations between the homozygous low activity genotype of the rs4680 (Val158Met) polymorphism and OCD males in two different studies, a case-control (73 OCD subjects and 148 healthy controls) and a family (110 OCD families) study; and another study detected nominal association of this low activity allele in OCD females (P=0.049) but not in males (Alsobrook et al., 2002). Significant finding was reported for this polymorphism in the recent meta-analysis using 25 OCD datasets (mean OR 1.200, 99th percentile CI 1.001-1.438, P=0.010). Although many studies have investigated COMT in relation to OCD risk, very few have examined other SNPs within this gene. Regarding pharmacogenetics of antidepressant response, Vulink et al. (2012) detected significantly more responders to citalopram with the rs4680 Met/Met genotype (P=0.007).

78

79

Monoamine oxidase A (MAOA) [MIM 309850] is the only examined gene that is localized to chromosome X, more specifically Xp11.3 (Levy et al., 1989). It encodes MAOA, which are expressed in the outer mitochondrial membrane. It functions to regulate amines by oxidizing neurotransmitters such as serotonin, norepinephrine, and dopamine, in addition to dietary amines, which is vital in maintaining mental states. Low levels of MAO activity and mutations in this gene have been associated with violent, criminal, and/or impulsive behaviour (Chen et al., 2004). MAOA inhibitors including and are third-line pharmacological treatment of OCD (CPA, 2006); however, given its side effect profile and dietary restriction, this drug class has not been widely prescribed. Very few studies have examined MAOA in OCD and Karayiorgou et al. (1999) reported a sexual dimorphic association between the allele with high MAOA enzymatic activity and OCD in 110 OCD families. However, a meta-analysis of six OCD datasets did not yield significant finding (P=0.483) (Taylor, 2013).

Neurotrophic and Neuroplasticity Gene:

Brain-derived neurotrophic factor (BDNF) [MIM 113505] is located at chromosome 11p14 (Hanson et al., 1992). BDNF is a member of the neurotrophin superfamily of survival- promoting molecules such as neurotrophin-3/4/5, and nerve growth factor (NGF), which play an important role in cell survival, differentiation, and cell death (Jones and Reichardt, 1990; Numakawa et al., 2010). Members of this family play a vital role in promoting neuronal growth and maintenance during normal development and differentiation of the vertebrate central nervous system (Hofer and Barde, 1988). BDNF specifically has trophic effects on neuronal plasticity and learning (Cowansage et al., 2010; Tyler et al., 2002). BDNF is therefore an attractive candidate for molecular analysis in OCD, based on potential functional relevance, and reported interaction between this gene and other neurotransmitters implicated in this disorder. BDNF is involved in both early brain development and adult synaptic formation and decay. There is a reciprocal relationship between BDNF and the serotonergic system: BDNF has important modulatory effects on serotonergic transmission (Celada et al., 1996; Martinowich and Lu, 2008; Mössner et al., 2000), and serotonergic inputs influence BDNF expression (Garcia et al., 2003;

79

80

Molteni et al., 2010; Nibuya et al., 1995). The interaction of serotonin and BDNF is particularly relevant given the large extant literature implicating involvement of serotonergic neurotransmission in OCD and the well described efficacy of serotonin reuptake inhibitors (SRIs) in the treatment of this disorder (Koran et al., 2007; Soomro et al., 2008; Zohar et al., 2000). Moreover, BDNF is also functionally interconnected with many neurotransmitters involved in OCD, including serotonin, glutamate, and dopamine. Several studies have indicated a possible effect of BDNF on the differentiation of serotonergic and dopaminergic neurons (Eaton et al., 1995; Studer et al., 1995; Zhou et al., 1994). BDNF enhances phosphorylation of the postsynaptic N-methyl D-aspartate (NMDA) receptor subunit 1 (Suen et al., 1997), an interaction that may be particularly relevant given the apparent role of glutamate in OCD based on genetic studies (Arnold et al., 2006; Rotge et al., 2010) and promising early reports of anti-obsessional efficacy of glutamate modulating agents (Wu et al., 2012). Further, two published reports describe lower peripheral BDNF serum (Maina et al., 2010) and plasma levels (Wang et al., 2011) in individuals with OCD, which was recently confirmed by another group (Fontenelle et al., 2012) and a meta-regression analysis by Suliman et al. (2013).

BDNF polymorphisms have been associated with a variety of psychiatric phenotypes such as mood disorders (Neves-Pereira et al., 2002; Sen et al., 2003; Sklar et al., 2002) and late age at onset in schizophrenia (Krebs et al., 2000; Muglia et al., 2003). Hall and colleagues first reported association for a number of BDNF polymorphisms in OCD, including the Met allele of the Val-66-Met polymorphism, a finding that was most significant in their early onset cases (Hall et al., 2003). Several other studies have also found significant association between OCD and BDNF including a five-marker protective haplotype containing the Val allele of the Val-66-Met polymorphism (P = 0.006) (Alonso et al., 2008) and Val as the protective allele in a case-control study (P < 0.01) (Rocha et al., 2010). Interestingly, another case-control study reported no significant differences in allele or genotype frequency between patients with OCD and healthy controls, however the Met/Met genotype was associated with later illness onset (P = 0.008) and a trend observed towards association of the Val/Val genotype and lower Y-BOCS scores (P = 0.051) in the female subgroup (Katerberg et al., 2009). Furthermore, Hemmings et al. (2013) recently detected that the Met allele increased the risk of OCD in individuals with a history of childhood emotional abuse. By contrast, the Val allele has also been associated with OCD diagnosis in the most recent study by Márquez et al. (2013) (P = 0.0001) and more severe OCD 80

81

in female OCD cases (Hemmings et al., 2008). The Val/Val genotype was significantly associated with OCD (Márquez et al., 2013; P = 0.0001) and hoarding symptoms (Timpano et al., 2011). However, there have also been six negative studies published in this population, including Dickel et al. (2007), Klaffke et al. (2006), Mossner et al. (2005), Tükel et al. (2012), Wendland and colleagues (2007), as well as our team in a previous study of both the Val-66-Met or BDNF-(GT)n polymorphisms with OCD (Zai et al., 2005). In addition to the Val-66-Met polymorphism, a recent study reported that in OCD patients the C allele of rs2883187 was associated with more severe symptoms and higher familial loading of OCD (Tükel et al., 2014). Despite these mixed findings and a recent negative meta-analysis of the Val-66-Met polymorphism (Zai et al., 2015), it appears possible that variable BDNF expression or function, due to DNA sequence variation, may contribute to the neuropathogenesis of OCD phenotypes. Furthermore, only one pharmacogenetic study has been published and Real et al. (2009) reported that the rs1491851 C allele was associated with better SRI response (P=0.005) whereas other SNPs within this gene was negative. Thus, further investigation is necessary to fully explore its role in OCD.

Myelin-Related Genes:

Brain imaging studies have identified potential neurobiological factors involved in the development of OCD (Kwon et al., 2009; MacMaster, 2010; Saxena and Rauch, 2000). Most studies have implicated abnormal link between the CSTC circuit, particularly the pathway that connects the orbito-frontal cerebral cortex to the (Graybiel and Rauch, 2000; Saxena and Rauch, 2000). Several early neuroimaging studies reported decreased white matter volume in OCD when compared with healthy controls (Breiter et al., 1994; Jenike et al., 1996; MacMaster et al., 1999; Rosenberg et al., 1997), which has led to questions regarding possible white matter dysfunction in OCD. This was recently confirmed by a recent meta-analysis of neuroimaging in OCD using multimodal voxel-based methodology (Radua et al., 2014), which reported widespread white matter abnormalities with the most robust finding of an increase in white matter volume and reduction in fractional anisotropy (FA) in the anterior midline tracts between the anterior parts of the cingulum bundle and body of corpus callosum. Diffusion tensor

81

82

imaging (DTI) studies comparing white matter of patients with OCD to healthy volunteers have reported lower FA bilaterally in the anterior cingulate gyrus (Szeszko et al., 2005), the rostrum of the corpus callosum, as well as the cingulum bundle (Saito et al., 2008). FA is reflective of the integrity of myelination and axonal function, with lower values reflecting disturbance in the unidirectional diffusion expected in normal axonal bundles. Cannistraro et al. (2007) reported greater left than right FA in the cingulum bundle and anterior limb of the internal capsule regions. Menzies et al. (2008) reported significantly reduced FA in a large region of right inferior parietal white matter and increased FA in the right medial frontal region in patients with OCD in addition to abnormal FA findings in their relatives. A recent article reviewed all DTI studies in OCD and confirmed that the majority of studies showed decreased FA in OCD patients when compared to healthy controls, with the most consistent white matter connectivity abnormalities found in the cingulate bundle, corpus callosum, and anterior limb of internal capsule (Koch et al., 2014). Thus, based on this literature, researchers have hypothesized that impairment in the development or maintenance of myelination might play a role in the etiology of OCD in addition to providing rationale for white matter as an endophenotype of OCD. This hypothesis has received some support from preliminary genetic studies implicating two myelination genes, the myelin oligodendrocyte glycoprotein (MOG; Zai et al., 2004) and the oligodendrocyte lineage transcription factor 2 (OLIG2; Stewart et al., 2007), in OCD.

Myelin oligodendrocyte glycoprotein (MOG) [MIM 159465] is mapped to chromosome 6p22.1 (Pham-Dinh et al., 1993), just distal to the human leukocyte antigen region, which contains genes that contribute to immune response. Its encoded protein is a minor component of the myelin sheath, which is expressed preferentially on the extracellular surface of oligodendrocytes (Pham- Dinh et al., 1993). It has been postulated that MOG functions as a cellular adhesion molecule, a regulator of oligodendrocyte microtubule stability, and/or a mediator of interactions between myelin and the immune system, particularly as an activator of the complement cascade (Johns and Bernard, 1999). It contains an immunoglobulin-like variable domain characteristic of members of the immunoglobulin super-family (Pham-Dinh et al., 1993) and variants of this gene may therefore be a target for immune-related demyelination and contribute to the development or progression of autoimmune disorders including multiple sclerosis (Lee and Linker, 2012; Reindl et al., 2013). An immune-related proposed mechanism of the development of OCD came from the Pediatric Autoimmune Neuropsychiatric Disorders Associated with (group A β-hemolytic) 82

83

Streptococcal (GABHS) infection (PANDAS), which has been characterized with an abrupt onset of OCD symptoms following GABHS infection (da Rocha et al., 2008; Snider and Swedo, 2003). Another term, Pediatric Acute-onset Neuropsychiatric Syndrome (PANS), has been proposed in 2012 to expand the clinical entity of the existing PANDAS by removing the cause (GABHS) (Swedo et al., 2012). Only one study has investigated genetic variants of this gene in OCD. Zai et al. (2004) a significant association between OCD and allele 2 at the tetranucleotide repeat (TAAA)n located in the 3’-untranslated region (P=0.022) in 160 small nuclear families with an OCD proband. The same allele 2 was found to be associated with highly Y-BOCS severity score (P=0.020). Additionally, the MOG haplotype C-C-13-2 [rs2251711-rs2071653- (CA)n-(TAAA)n] was significantly over-transmitted from parents to their OCD-affected offspring (P=0.011). Atmaca et al. (2010) provided further support for this gene in the etiology of OCD, reporting significantly larger total white matter volumes in OCD patients carrying the (Val142Leu) Val/Val genotype (P<0.01).

Oligodendrocyte lineage transcription factor 2 (OLIG2) [MIM 606386] encodes for a protein that is an essential regulator in the development of human white matter (myelin)-producing cells (Takebayashi et al., 2002) and it is highly expressed in the amygdala, caudate nucleus, and , brain regions that have been associated with OCD (Rauch and Savage, 1997; Saxena et al., 1998). It may act as a transcription factor required for oligodendrocyte progenitor cell differentiation into different lineages during re-myelination throughout life (Ligon et al., 2006). It is mapped on chromosome 21q22.11 (Georgieva et al., 2006; Wang et al., 2000). A family- based association study of this gene first reported an over-transmission of the rs9653711 C allele (P=0.0035), over-transmission of the G-A-T-T-C (rs762178-rs1059004-rs9653711-rs6517137- rs13046814) haplotype (P=0.020), and an under-transmission of the A-C-T-T-G haplotype (P=0.0035) in 66 families with 160 OCD probands (P=0.0035) (Stewart et al., 2007). Two studies have been published since with one negative and one positive result. Zai et al. (2012) failed to replicate these findings in 160 OCD small nuclear families using the family-based association approach. However, a recent study supported the initial findings and demonstrated that in addition to the rs9653711 polymorphism, rs762178 and rs1059004 were also associated with OCD diagnosis and early-onset OCD in a case-control study comprising of 400 OCD subjects and 459 healthy controls (Zhang et al., 2015). Thus, this gene may be an interesting candidate to examine OCD subphenotypes in OCD. 83

84

Other Genes:

One large study (Chapters 4 and 6) mainly focused on the remote regulatory regions (Zai et al., 2014), testing markers with potential functional impact on several interesting top-hit OCD candidate genes from the first GWAS (Stewart et al., 2013): fas apoptotic inhibitory molecule 2 (FAIM2), fucosyl-transferase 2 (FUT2), and BTB (POZ) domain containing 3 (BTBD3).

Fas apoptotic inhibitory molecule 2 (FAIM2) [MIM 604306] is the eighth top hit gene in the OCD GWAS (Stewart et al., 2013). It is mapped to chromosome 12q13.12 (Somia et al., 1999) and has been implicated in obesity given its role in apoptosis (Williams et al., 2012). Its encoded protein is a 35.1 kDa membrane protein, which is highly expressed in the hippocampus and it regulates Fas ligand-mediated apoptosis in neurons (Fernandez et al., 2007). Specifically, FAIM2 is a neuron-specific inhibitor of Fas/CD95-mediated apoptosis with potential neuroprotective effect (Reich et al., 2011). Evidence has shown that this gene is regulated by nutritional state and methylation levels of the FAIM2 promoter region have been found to be associated with obesity (Wu et al., 2015). However, the molecular mechanism by which this gene affects obesity has yet to be clarified. Moreover, its role in psychiatric disorders or drug response has not yet been determined.

Fucosyl-transferase 2 (FUT2) [MIM 182100] is one of the top hit genes from the GWAS of OCD (Stewart et al., 2013) and is localized onto chromosome 19q13.33 (Rouquier et al., 1995). It has been found to be positively predict plasma vitamin B12 levels in two GWASs (Hazra et al., 2008; Tanaka et al., 2009) and associated with Crohn’s disease (McGovern et al., 2010). Interestingly, vitamin B12 serum level was reported to be abnormally low in OCD patients when compared to healthy controls (Hermesh et al., 1988). A case report of vitamin B12 deficiency was associated with OCD symptoms (Sharma and Biswas, 2012). A recent study also implicated a low vitamin B12 level in OCD patients (Türksoy et al., 2014). Thus, this gene may be associated with OCD in the context of low vitamin B12 serum level. Exploration of this gene in psychiatric disorders or drug response has not been done.

84

85

BTB (POZ) domain containing 3 (BTBD3) [MIM 615566], localized on chromosome 20p12.2, is the strongest associated gene in the trio sample of the first GWAS of OCD with a P value of 3.84E-08 (Stewart et al., 2013). BTBD3 has multiple cellular functions including the regulation of dendritic orientation (Matsui et al., 2013), cytoskeleton dynamics, transcriptional regulation, ion channel assembly and gating, protein ubiquitination and degradation (Perez-Torrado et al., 2006). Its close family of transcription factors, BTBD9, has been found to be associated with Tourette syndrome, which is frequently comorbid with OCD (Riviere et al., 2009); however, its role in other psychiatric disorders or drug response remains currently unknown.

These candidate genes described above have been selected based on their previously significant genetic association with OCD and several were recent top hits from the first published GWAS of OCD (Stewart et al., 2013). The rationale for SNP selection was mainly based on the ENCyclopedia Of DNA Elements (ENCODE) project, which reported that 80.4% of the human genome displays some functionality in gene regulation, gene-gene, and gene-environment interaction (Kavanagh et al., 2013). This provides strong support for the examination of genetic variations across these remote regulatory regions in terms of OCD phenotypes and drug response in OCD.

1.5.1 Significance of Remote Regulatory Regions

The human genome provides the fundamental script for human biology. However, our current understanding of the genome is still underway given that less than 3% of the genome is devoted to protein coding. The rest of the genome was once thought to be made up of “junk” DNA with unknown function; nonetheless, recent advances in human genetics identified over 80% of the genome as containing sequences of non-coding ribonucleic acids (RNAs), alternatively spliced transcripts, and regulatory regions with important regulatory functions (ENCODE Project Consortium, 2012; Kavanagh et al., 2013; Nair and Howard, 2013) (Figure 1.2). The identification of these newly discovered regulatory elements has provided researchers new insights into the mechanisms of gene regulation.

85

86

Previous OCD genetic studies have mainly examined known genetic variations within or surrounding genes of interest (Pauls et al., 2014). Almost none has investigated remote regulatory regions of these candidate genes. As mentioned above, the notion that our genome is made up of mostly “junk” DNA was disproven after the ENCyclopedia Of DNA Elements (ENCODE) project reported 80.4% of human genome displaying some functionality into gene regulation, gene-gene, and gene-environment interactions (Kavanagh et al., 2013; Nair and Howard, 2013). This provides a strong evidence of support to examine genetic variations across these remote regulatory regions in psychiatric disorders.

Defining the regulatory regions that control gene activity has become a key biomedical issue. Recent technological advances have provided an increasing repertoire of methods to identify regulatory elements. Regulation of gene expression is controlled by the sequence- specific binding of transcription factors (TFs) that may act as activators, repressors, or both, activating or preventing transcription. As illustrated recently by the ENCODE project (Figure 1.2), chromatin immuno-precipitation (ChIP) is a powerful assay that is used to detect TFs that attach to TF-binding sites (regulatory elements), chromatin-associated markers, and transcription-associated protein complexes. This invaluable tool helps to catalog sequences with characteristics of regulatory elements (ENCODE Project Consortium, 2012). Examples of functional regulatory elements include RNA transcribed regions, TF-binding sites for cis- regulatory modules, repressors, activators, enhancers, and silencers, chromatic structure, and DNA methylation sites. The ongoing discovery of these functional regulatory elements will allow researchers to address specific underlying mechanism to further our understanding of psychiatric disorders. This will in turn accelerate gene discovery efforts, leading to novel therapeutics.

One of the major studies in this thesis examines genetic variations that have functional implications on their respective genes. There are many different publicly available search engines including functional significance (FS) score, European minor allele frequencies (EurMAF), BrainCloud (Colantuoni et al., 2011), UCSC Genome Browser (Rosenbloom et al., 2013), HaploReg (Ward et al., 2012), RegulomeDB (Boyle et al., 2012), and National Institute of Environmental Health Sciences (NIEHS) (Xu et al., 2009), which facilitate the identification of SNP with higher allele frequency for meaningful genetic analysis and known functions.

86

87

Thus, one of the goals of this PhD thesis is to utilize these publicly available databases to identify SNPs in the remote regulatory regions of OCD candidate genes and to examine the impact of these SNPs in OCD subphenotypes and drug response.

1.5.2 Genetic Methodology

Traditional genetic studies in psychiatric disorders have focused on the hypothesis-driven candidate gene approach. Candidates were chosen based on the putative biological mechanisms and drug targets of existing psychotropic medications. With the advances in genomic technology, genome-wide association studies (GWASs) have the ability to screen for genetic variations across the entire genome with varying degree of coverage. By using the latter approach and in contrast to the candidate gene studies, a hypothesis-free screening for genetic involvement can be made possible. Given the large amount of generated data, the significance level threshold has to be corrected to 5E-08 for multiple comparisons. Therefore, the studied sample size will have to be substantially enlarged in the genetics of psychiatric disorders given the relatively smaller effect size of genetic findings to date.

The goal of GWAS is to detect signals across the entire genome but it is unable to fine map the region of interest. Additional algorithms have been developed to test for specific hypotheses in the context of GWAS. Candidate gene studies have the advantage of tailoring the SNP selection to cover all common variations across each gene, taking into the consideration of functional significance of each SNP.

The increasing use of GWAS approach has shifted stand-alone institutional contributions to large-scale international collaboration because the sample sizes required for GWAS discovery and replication are beyond the reach of a single research group. This shift has led to the establishment of the Psychiatric Genomics Consortium (PGC) since the early 2007. PGC currently consists of over 800 investigators from 38 countries and over 900,000 samples have been collected for analysis. To date, there are only two GWASs of OCD diagnosis, which did not detect any genome-wide significant findings. Hence, in this thesis, we have explored OCD phenotypes in order to better understand the complexity of this illness.

87

88

Figure 1.2. Annotation of Disease-Associated Variants from ENCyclopedia Of DNA Elements (ENCODE) Data (Kavanagh et al., 2013). LD: linkage disequilibrium; TSS: transcription start site; TFBS: transcription factor binding site

88

89

Chapter 2 AIMS, OBJECTIVES, HYPOTHESES, AND IMPLICATIONS

2 Thesis Aims, Hypotheses, and Implications

The main purpose of this thesis is to aid in the discovery of genetic influence in the etiology of OCD by examining its subphenotypes. There are many approaches in genetic studies and this thesis will only focus on common genetic variations that affect risk of developing specific OCD subphenotypes in addition to predicting response to serotonin reuptake inhibitor (SRI) antidepressant treatment in OCD.

2.1 Aims and Objectives

Prior to investigating the genetic basis of OCD subphenotypes, a comprehensive examination of the clinical phenomenology of OCD was necessary as a precursor to genetic studies. Several OCD subphenotypes were explored including age at onset (AAO), Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) symptom severity, symptom dimensions, comorbid psychiatric conditions, family history, and SRI response.

In determining the genetic architecture of OCD subphenotypes and drug response in OCD, serotonin system genes are obvious candidates given the extensive literature supporting serotonergic mechanisms as etiologically important in this illness. Moreover as the first-line agents for this condition, all SRIs exert their effects by modifying serotonergic neurotransmission, and this provides further support for the importance of this pathway. Additionally, evidence suggests that the neuropathology of OCD lies in the complex neurotransmitter network of the cortico-striato-thalamo-cortical (CSTC) circuit, where dopamine and glutamate dysfunctions also have been implicated in this disorder (Pauls, 2010; Pauls et al., 2014). Genotyping data derived from a genome wide association study (GWAS) of OCD (Stewart et al., 2013), of which our group was a contributor, provides an interesting opportunity to explore the genetic basis of OCD subphenotypes and drug response in OCD. 89

90

2.1.1 Aims

The following is a list of objectives for this PhD thesis:

1) To identify and refine clinically useful subphenotypes in OCD for genetic studies;

2) To develop further understanding of the complex molecular and genetic basis of OCD;

3) To conduct statistical analyses of complex traits in OCD using different genetic approaches including candidate gene study and non-hypothesis driven GWAS.

2.1.2 Objectives

There are two main sections of this thesis. The initial objective is to examine the complex clinical characteristics of OCD and to analyze them in an approach that reduces the dimensions of the data, therefore making it more amenable to genetic analysis.

The second objective of this thesis is to utilize genetic methodology to examine these OCD subphenotypes in order to identify genetic variants that contribute to:

1) The susceptibility risk of developing a particular subgroup of OCD (distinguishing OCD subphenotypes), and;

2) SRI response in the treatment of OCD.

This thesis will attempt to dissect the various shared and distinct genetic variations that influence the risk of developing certain subtypes of OCD, earlier onset, severity, familial form, additional comorbid psychiatric conditions, and predict response to SRIs.

90

91

2.2 Hypotheses 2.2.1 Clinical Phenomenology of OCD

The null hypotheses of the clinical phenomenology of OCD are that:

1. There are no differences within gender and AAO groups;

2. The Y-BOCS symptom checklist will generate the similar symptom dimensional factors as previous published studies;

3. The rates of significant family history and comorbid psychiatric condition(s) are the same as published reports.

The alternative hypotheses of the clinical phenomenology of OCD are that:

1. There are significant differences within gender and AAO groups;

2. The Y-BOCS symptom checklist will generate different symptom dimensional factors from previous published studies;

3. The rates of significant family history and comorbid psychiatric condition(s) are different from published reports.

2.2.2 Genetics of OCD Phenotypes

The null hypothesis of the genetics of OCD phenotypes is that genetic variations within the serotonergic system (i.e., serotonin transporter [5HTT/SLC6A4], serotonin 2A receptor [HTR2A]) and the glutamatergic system (i.e., glutamate receptor, ionotropic, N-methyl D- aspartate 2B [GRIN2B], glutamate transporter [SLC1A1]) candidate genes are not associated with any phenotypes (i.e., AAO, Y-BOCS severity, OCD symptom dimensions, psychiatric comorbidity, family history) in OCD.

91

92

The alternative hypothesis is that genetic variations within the serotonergic system (i.e., 5HTT/SLC6A4, HTR2A) and the glutamatergic system (i.e., GRIN2B, SLC1A1) candidate genes are associated with specific subphenotypes (i.e., AAO, Y-BOCS severity, OCD symptom dimensions, psychiatric comorbidity, family history) in OCD (Figure 2.1).

2.2.3 Genetics of SRI Response in OCD

The null hypothesis of the genetics of SRI response in OCD is that genetic variations within the serotonergic system (i.e., 5HTT/SLC6A4, HTR2A) and the glutamatergic system (i.e., GRIN2B, SLC1A1) candidate genes are not associated with SRI response in OCD.

The alternative hypothesis is that genetic variations within the serotonergic system (i.e., 5HTT/SLC6A4, HTR2A) and the glutamatergic system (i.e., GRIN2B, SLC1A1) candidate genes are associated with SRI response in OCD (Figure 2.1).

2.3 Originality

This is the first attempt to utilize a GWAS approach to examine genetic susceptibility of OCD subphenotypes. Furthermore, given the recent interest in regulatory elements from the ENCyclopedia Of DNA Elements (ENCODE) project (ENCODE Project Consortium, 2012), it is important to examine regulatory regions in the vicinity of top OCD candidate genes, regions that can potentially alter gene function and protein activity.

2.4 Implications

This thesis is designed to investigate the genetic susceptibility and pharmacogenetics of OCD. Thus, we may better understand genetic factors for risk of developing various forms of OCD. The identification of genetic factors that contribute to SRI response is vitally important in identifying those who are responders. The clinicians will be assisted in choosing the best tolerated medications in order to prevent the significant morbidity and mortality in patients with 92

93

OCD. Furthermore, by developing an algorithm to determine the dimensions of OCD symptoms, it can help researchers and clinicians to identify a more homogenous OCD group with potentially different response to a given SRI.

93

94

Figure 2.1. Thesis Hypothesis. This figure illustrates the hypothesis of this Ph.D. thesis – different genetic variations may predispose to different phenotypes of OCD and there may be shared genetic factors influencing the risk of OCD phenotypes.

94

95

Chapter 3 CLINICAL PHENOMENOLOGY OF OCD

3 An Update on the Clinical Phenomenology of Obsessive-Compulsive Disorder

Gwyneth Zai 1,2,4, David Pauls 3, Vincenzo de Luca 1,2, Marissa Williams 4, Karen Wigg 4, James L. Kennedy 1,2, Margaret A. Richter 1,2,4

1 Neurogenetics Section, Centre for Addiction and Mental Health, Clarke Division, Toronto, ON, Canada

2 Department of Psychiatry, University of Toronto, Toronto, ON, Canada

3 Psychiatric and Neurodevelopmental Genetics Unit, Center for Human Genetic Research, Massachusetts General Hospital, Boston, MA, USA

4 Frederick W. Thompson Anxiety Disorders Centre, Department of Psychiatry, Sunnybrook Health Sciences Centre, Toronto, ON, Canada

* Corresponding author: Dr. James L. Kennedy, Director of the Neuroscience Department and

Head of Neurogenetics Section at the Centre for Addiction and Mental Health (CAMH),

Professor of the Department of Psychiatry and the Institute of Medical Science at University of

Toronto, Toronto, ON M5T 1R8, Canada (Phone: 1-416-979-4987; Fax: 1-416-979-4666; E- mail: [email protected])

95

96

3.1 Abstract

Background: Obsessive-compulsive disorder (OCD) is a complex and debilitating disorder with high rate of comorbid psychiatric illness. Evidence suggests that OCD is a heterogeneous condition. Existing genetic studies in psychiatric disorders have been limited by poor reproducibility and the use of current diagnostic systems, which lacks biological justification. The main goal of this chapter is to identify homogeneous subphenotypes of OCD that are clinically relevant and suitable for future genetic approaches with greater effect size in search for the underlying mechanism in the etiology of OCD.

Methods: We explored demographics, clinical presentations, and treatment outcome in a large Canadian OCD sample consisting of 560 OCD individuals and 273 family members. Examined variables included: gender, age at onset (AAO), OCD symptoms and severity as measured by the Yale-Brown Obsessive-Compulsive Scale (Y-BOCS), DSM-IV psychiatric comorbidity, family history of obsessive-compulsive and related disorders (OCRDs), and serotonin reuptake inhibitor (SRI) response. The distribution of AAO was explored using admixture analysis in the STATA program in order to identify statistically calculated subgroups of AAO, and symptom dimensional factors were derived using the factor and cluster analyses in the SPSS software on the Y-BOCS symptom checklist. Frequency of comorbid psychiatric disorders including lifetime OCRDs and current psychiatric disorders was calculated using the diagnoses that were determined from the SCID-IV interview. Family history of OCRDs including OCD, hoarding disorder (HD), trichotillomania (TTM), body dysmorphic disorder (BDD), and skin picking disorder (SPD) was examined using a modified version of the Family History Interview (FHI). Serotonin reuptake inhibitor (SRI) response was determined retrospectively using the Clinical Global Impression – Improvement (CGI-I) scale. Clinical measures were compared between gender, statistically derived AAO subgroups (early-, intermediate-, and late-onset), familial status, and SRI response status (responders and non-responders).

Results: Female OCD subjects in our sample had a higher percentage of comorbid lifetime OCRDs when compared with OCD males but no differences in other clinical characteristics were detected. AAO analysis showed a tri-modal distribution, separating early (≤7 years), 96

97

intermediate (8-17 years), and late (≥18 years) onset groups. The early onset group showed significantly more symmetry/order and contamination/cleaning symptoms. Factor analysis for Y-BOCS confirmed a 5-factor model (factor 1: symmetry/order; factor 2: aggression/sexual/religious; factor 3: hoarding; factor 4: contamination/cleaning; factor 5: somatic) when treating all symptoms equally, and detected a 6-factor model (factor 1: symmetry/order; factor 2: hoarding; factor 3: contamination/cleaning; factor 4: aggression/checking; factor 5: somatic; factor 6: sexual/religious) when double-weighting the principal/target symptoms (where aggression/sexual/religious dimension in the 5-factor model was divided), which were both confirmed with the cluster analysis. Fifty-nine percent and 24% of the OCD participants had one or more current comorbid psychiatric diagnoses and lifetime OCRDs respectively. Over 21% of the OCD individuals in our sample had a significant family history of OCRDs and only one individual had an overlapping comorbid and family history of the same condition (TTM). Familial OCD cases had less comorbidity. The SSRI/SRI antidepressant response rate in our OCD sample was 70.9%/74.7%. SRI responders had more symmetry/order symptoms, higher rates of comorbid mood disorders and family history of OCD and hoarding when compared to non-responders.

Conclusions: This is one of the largest North American OCD sample with an extensive investigation of clinical phenomenology. We derived clinically relevant and easily interpretable subphenotypes in OCD, which also detected significantly different clinical characteristics between the identified groups. The identification of more homogeneous subphenotypes will aid in genetic studies of OCD and should enable more definitive detection of genes contributing to different aspects of this complex disorder.

3.2 Introduction

Obsessive-compulsive disorder (OCD) is a chronic and debilitating psychiatric disorder that affects approximately 2% to 3% of the general population (APA, 2000; Ruscio et al., 2010). This disorder is characterized by obsessions that are recurrent intrusive thoughts, images, or urges, and compulsions, which are repetitive rituals or behaviours performed in order to counteract the anxiety-provoking obsessions (APA, 2013). These symptoms are typically 97

98

egodystonic and are frequently thought of as irrational and excessive by patients. The DSM-5 classifies OCD as one of the obsessive-compulsive and related disorders (OCRDs), which are a new category of disorders that have been separated from the DSM-IV anxiety disorders (APA, 2013). In addition to OCD, this new diagnostic category of OCRDs includes body dysmorphic disorder (BDD), hoarding disorder (HD), trichotillomania (TTM) or hair-pulling disorder, and excoriation (skin-picking) disorder (SPD) (APA, 2013). Hoarding used to be considered a symptom of OCD (APA, 2000) but given its distinct clinical presentation, treatment outcome, and genetic influence, it is now considered a separate disorder within OCRDs (APA, 2013).

The clinical presentation of DSM-IV OCD is highly heterogeneous given that there are many different symptom characteristics such as fear of contamination and cleaning rituals, obsessions and compulsions with symmetry and ordering, checking, hoarding, aggressive, religious, sexual, and somatic nature, as well as hoarding (APA, 2000) prior to the introduction of DSM-5. Furthermore, phenotypic heterogeneity has been supported by many epidemiological and clinical studies of OCD in terms of symptom subtype or dimension, age at onset (AAO), and treatment outcome.

To better understand the complexity of OCD, it may be important to delineate the diagnosis into different phenotypically distinct subtypes. Furthermore, unique and common pathophysiological mechanisms may be easier to identify by studying more homogeneous diagnostic subgroups of this disorder (Robins et al., 1970).

Researchers have postulated different schemes to ascertain different subgroups of OCD. OCD can be subdivided according to gender (Torresan et al., 2013), AAO (de Mathis et al., 2008; Janowitz et al., 2009; Taylor et al., 2011), comorbidities such as the presence or absence of schizotypal personality traits (Brakoulias et al., 2014; Jin Lee et al., 2006), obsessive-compulsive personality traits (Coles et al., 2008; Wetterneck et al., 2011), tics (Gomes de Alvarenga et al., 2012; Miguel et al., 1997), hoarding (Samuels et al., 2007; Torres et al., 2012), and other disorders (Murphy et al., 2010), in addition to treatment outcome (Nakamae et al., 2011; Sumitani et al., 2006), and symptom dimensions (Bloch et al., 2008; Fontenelle et al., 2005; Hasler et al., 2007). However, no one study has comprehensively examined a large OCD sample with regard to individuals’ demographics, clinical characteristics, and other OCD subphenotypes.

98

99

We aimed to explore the relationship of gender, age at onset (AAO), Yale-Brown Obsessive- Compulsive Scale (Y-BOCS) symptom checklist dimensions, psychiatric comorbid conditions, family history of obsessive-compulsive and related disorders (OCRDs), and serotonin reuptake inhibitor (SRI) response in a large well-characterized OCD sample. The rationale supporting our objective to examine gender and these OCD subphenotypes is explained below.

Although the gender ratio of OCD is 1:1 (APA, 2000), male children and adolescents have a slightly higher rate of OCD than females (Fineberg et al., 2013). Differences in clinical presentation between OCD males and females have previously been documented; nonetheless, there is one study that did not detect any gender differences (Matsunaga et al., 2000). Men suffering from OCD tend to have a chronic course of illness (de Mathis et al., 2011; Nestadt et al., 1998; Rosso et al., 2012) with greater symptom severity (Nestadt et al, 1998; Rosso et al., 2012) and they endorse more symptoms related to aggression (Tukel et al., 2013), sexual themes, religious themes, symmetry or exactness, and hoarding behaviours (de Mathis et al., 2008; Jaissorya et la., 2009; Rosso et al., 2012; Tukel et al., 2013). On the other hand, females with OCD have been reported to have an acute onset of symptoms and an episodic course of illness (de Mathis et al., 2011; Nestadt et al., 1998; Rosso et al., 2012); they also appear to develop symptoms associated with contamination and somatization more often than OCD males (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013).

As mentioned previously, males appear to have an earlier onset of OCD and AAO for OCD has been reported to be in the range of 10 and 37 years although most cases have emerged by the age of 22 years (Fineberg et al., 2013). Previous studies have reported differences in the clinical characteristics of OCD depending on early- or late-onset; however, most have arbitrarily set the cut-off age for early- and late-onset without any validation (Chabane et al., 2005; Grant et al., 2007; Janowitz et al., 2009; Maina et al., 2008). Only four studies to date have examined AAO with a more in-depth approach using admixture analysis to identify subgroup of OCD (Albert et al., 2015; Anholt et al., 2014; Delorme et al., 2005; De Luca et al., 2011) and all confirmed a bimodal distribution in which differences in the clinical characteristics were observed between early- and late-onset groups. These independent studies implicated the potential utility of AAO to ascertain clinically informative subgroups of OCD. However,

99

100

variations in the differing clinical features for the AAO groups between the studies were substantial, suggesting the need for further research.

Variable clinical symptoms of OCD have been described and no one patient presents with the exact same severity and clusters of symptoms. The Y-BOCS symptom checklist has 74 items related to a distinct OCD symptom. Thus, this measure enables ascertainment of the breadth of symptoms for clinical outcome comparison, treatment guidance, and genetic studies. Multiple studies have utilized factor and cluster analyses to reduce the number of OCD symptoms identified with the Y-BOCS into meaningful and informative symptom dimensions. Although the majority of studies have identified a 5-factor model with factors including contamination/cleaning, hoarding, aggressive/checking, symmetry/ordering, and sexual/religious (Table 1.6), differences in clinical characteristics amongst the groups were reported (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013). The identification of dimensional structure of OCD symptoms using this quantitative approach has the potential to advance our understanding of OCD. This is an attempt to derive OCD symptom dimensions in our large and well-characterized sample of OCD because these dimensional phenotypes may be useful in our future efforts to understand the natural history, genetics, neurobiology, treatment response, and outcomes of OCD (Leckman et al., 2007; Mataix-Cols et al., 2005).

Large-scale epidemiological studies have reported a high lifetime and current rate of comorbid psychiatric conditions in OCD patients, ranging from 76% to 91% and 42% and 55% respectively (Ruscio et al., 2010; Hofmeijer-Savink et al., 2013). Anxiety and mood disorders are amongst the most frequent comorbid conditions in individuals who suffer from OCD (Ruscio et al., 2010; Hofmeijer-Savink et al., 2013). Although these psychiatric disorders are distinct, they do share similar features and may form a basis for investigating the cross-disorder etiological overlap in the near future. However, most clinical studies only examine the difference in OCD presentation between the presence and absence of tics (do Rosario-Campos et al., 2005; Nestadt et al., 2009). The putative tic-related subtype may account for 10% to 40% of the childhood-onset OCD cases (Leckman et al., 1994). OCD individuals with a personal history of tics have been reported to show a male predominance with earlier onset and greater symmetry and hoarding symptoms (Baer, 1994; Leckman et al., 1997).

100

101

A higher rate of OCD in relatives of OCD patients, two times in OCD adults and 10-fold in OCD children and adolescents, when compared to the general population has been consistently reported (Mataix-Cols et al., 2013; Pauls, 2010; Pauls et al., 2014). When comparing clinical characteristics between familial and sporadic cases of OCD, familial OCD was associated with earlier onset of diagnosis and longer duration of untreated illness in addition to reporting different types of symptoms such as and cognitive compulsions (Viswanath et al., 2011), ordering and arranging compulsions (Hanna et al., 2005) and higher comorbid OCRDs especially SPD and anxiety disorders (Hanna et al., 2005) in addition to major depression (Viswanath et al., 2011). Furthermore, familial OCD was found to be associated with treatment resistance (Pallanti and Grassi, 2014). Although multiple studies have supported the familial relationship between OCD and other OCRDs including BDD, TTM, SPD, and pathologic nail biting, no studies to date have examined the difference in clinical characteristics between the OCD individuals with and without a family history of OCRDs. These reported relationships of OCRDs may support the potential overlapping genetic risk between OCD and other OCRDs.

Hence, there is a critical need for deconstructing the current OCD diagnosis into different clinically relevant subgroups to determine the underlying mechanisms of OCD, to improve response to current treatment, and to provide basis for new therapies. Thus, the goal of this study is to examine the clinical phenomenology in a large OCD sample regarding gender, AAO, Y- BOCS symptom dimensions, psychiatric comorbidities, familiality, and SRI antidepressant treatment response.

3.3 Methods 3.3.1 Diagnostic Criteria and Clinical Sample

Five hundred and sixty OCD patients and 273 family members were recruited from consecutive referrals to the Anxiety Disorders Clinic at the Centre for Addiction and Mental Health and the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre in Toronto, Canada. All research participants were assessed using the Structured Clinical Interview for the DSM-IV (SCID) (First et al., 1996). Age at onset (AAO) 101

102

was defined as when the OCD subject first met diagnostic criteria of OCD (retrospective recall of when OCD symptoms first caused significant impairment or distress). The Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) (Goodman et al., 1989) was completed for each participant to determine the severity of OCD symptoms. Lifetime severity of OCD symptoms was estimated by re-administering the Y-BOCS scale whiles individuals focused on the time period when the most severe OCD symptoms were experienced.

All assessments and interviews were performed by trained research assistants who were supervised by an experienced psychiatrist. All participants met DSM-IV criteria for the diagnosis of OCD as determined by the SCID and available medical records. Psychiatric comorbid conditions were also determined by the SCID interview and OCRDs were defined using the DSM-5 diagnostic criteria. The presence of tics was determined by DSM-IV diagnostic criteria for tic disorder and Tourette’s syndrome during the SCID-IV interview. We did not screen participant for any Axis II psychiatric disorders in this study protocol.

All subjects provided their written informed consent to participate in this study, and ethics approval was obtained from the local Research Ethics Board. Inclusion criteria included adults aged between 18 and 65 years who met DSM-IV criteria for OCD diagnosis with more than one year symptom duration or more than one year since onset if waxing/waning course. Exclusion criteria included any metabolic or chronic neurological disease (other than tic disorder), active substance dependence or abuse (OCD in the context of active substance use, i.e., include OCD participant if OCD predated substance use), schizophrenia, or schizoaffective disorder. Psychiatric comorbidities were obtained from the SCID assessment.

3.3.1.1 Description of Y-BOCS Symptom Checklist and Severity Score

The Y-BOCS (Goodman et al., 1989) is the gold-standard, validated and clinician- administered instrumental measure that consists of a symptom checklist and severity scale (Appendix I). The symptom checklist has 74 specific symptom items that are clustered into 15 different groups of obsessions or compulsions including two miscellaneous categories, each for obsessions and compulsions. Each checklist item is rated according to whether the symptom

102

103

occurs currently in the present (past week), in the past, or never occurs. The five most interfering, severe, or distressing obsessions and compulsions to a total of 10 are identified and ranked as the “target” or “principal” symptoms. The Y-BOCS 10-item index of severity measures the degree of control and resistance that the patient has over his/her obsessions and compulsions, time spent on these obsessions and compulsions, in addition to the interference and distress caused by them. This severity scale is divided into two subscales for obsessions and compulsions (five questions each). Each question is scored between 0 and 4 with 0 as none and 4 as most severe. The Y-BOCS severity score can range from 0 (non-clinical) to 40 (most severe). A score less than 7 is considered to be non-clinical and a score great than or equal to 16 indicates clinically significant OCD symptoms.

The Y-BOCS has two versions, a clinician-rated and self-report questionnaire. The clinician-rated version has been reported to have high inter-rater reliability and internal consistency (α=0.89) (Goodman et al., 1989; Steketee et al., 1996); therefore, we have chosen to use the clinician-reported version of Y-BOCS.

3.3.2 Familiality Data to Determine the Presence of Absence of Family History of OCRDs

Family history data for all available first- and second-degree relatives were obtained using the OCD Family History Interview (OCD-FHI), which is a semi-structured interview developed by the investigators (Drs. Richter and Kennedy; see Appendix II for details). It was modeled on the Family History Research Diagnostic Criteria (Andreasen et al., 1986), yielding diagnoses of OCD and relevant spectrum conditions (OCRDs). Reliability of information was also collected due to recall bias of the participants and their family members. The criteria of defining a significant (positive) family history of OCRDs included either of the following:

1) A clinical diagnosis of OCRDs, or

2) Reliability measure with an informant providing accurate information or close relationship between the informant and subject who provided reasonable details on their mental health that met criteria for an OCRD, or

103

104

3) If the informant reported having no or very little information on or regular indirect contact with the subject (not meeting reliability cut-off of good or excellent), then at least two members from the same family (including proband and first-degree relatives) reported on the same symptoms that met criteria for an OCRD separately.

The actual step-by-step algorithm for the modified OCD-FHI is listed in Appendix II.

3.3.3 Antidepressant Response Data Description

Two hundred and twenty-two of the 560 OCD patients had adequate antidepressant response data. Of the 222 sample, 217 Caucasian patients had drug response data for analysis. Six antidepressants including five selective serotonin reuptake inhibitors (SSRIs: fluoxetine, fluvoxamine, sertraline, paroxetine, citalopram) and clomipramine (serotonin reuptake inhibitor [SRI]), were examined for response or non-response.

The drug response data was collected by a questionnaire (Appendix III), which was developed by Drs. Kennedy and Richter, in addition to chart review retrospectively. Information including adequacy of trials (based on widely accepted criteria regarding optimal dose and duration), self-rated estimates of reliability of recollection, and response was obtained. Response to antidepressant medications was scored according to the Clinical Global Impression – Improvement Scale (CGI-I). Individuals deemed to be drug “responders” rated themselves as “much improved” or “very much improved” on at least one antidepressant; to be considered as “non-responders” had to give ratings of “minimal improvement”, “no change”, or “worse” in response to at least two different antidepressants.

The specific criteria for a complete and analyzable antidepressant treatment trial included the following:

1) Adequate duration of trial of at least 10 weeks

2) Adequate dosage of SRI (maximum dose achieved of: fluoxetine ≥20mg, fluvoxamine ≥150mg, paroxetine ≥20mg, sertraline ≥100mg, citalopram ≥20mg, or clomipramine ≥150mg) 104

105

3) Participant rated their recall of the drug trial as good or excellent

4) Individuals with intolerable side effects were excluded from the antidepressant response analysis given tolerability and response are different constructs and may have different genetic pathway

5) At least two failed trials of SRIs for coding as non-responder

3.3.4 Data Analysis

3.3.4.1 Gender Differences in Clinical Presentation of OCD

For each gender group, we investigated the relationship between gender and clinical characteristics including AAO, Y-BOCS severity, Y-BOCS symptom dimension factor, psychiatric comorbidities, family history of OCRDs, and SRI response using SPSS (version 20.0, Armonk, NY, USA). Levene’s test of homogeneity of variances was performed on continuous variables (age, AAO, Y-BOCS severity and symptom dimensions).

3.3.4.2 Identifying AAO subgroups using Admixture Analysis

We performed admixture analysis, a method for identifying the model that best fits the observed distribution of a continuous variable, AAO, as described in Delorme et al. (Delorme et al., 2005) using the DENORMIX (Kolenikov, 2001) module from the STATA program (version 11.0, Texas, USA). Admixture analysis has been used to reduce clinical heterogeneity. The DENORMIX analysis performs a decomposition of the AAO distribution into a mixture of normal components and estimates the number of components of the mixture. The maximum likelihood estimation of the finite normal mixture was applied to determine a theoretical model that best fits the observed distribution of AAO. The χ2 goodness-of-fit test was performed and the P value indicates the degree to which each model approximated the empirical distribution function of AAO in this sample. The model with the highest P value was selected to be the best- fitting model. The FMM procedure is then used to confirm the model, which fits finite mixture regression model to univariate outcomes using both maximum likelihood and Bayesian 105

106

techniques. Point of intersection between the normal distributions was calculated using the means and standard deviations of the identified components using the uniroot function in R program (version 3.2.1).

For each identified AAO group, we investigated the relationship between AAO and clinical characteristics including gender, age, Y-BOCS severity, Y-BOCS symptom dimension factor, psychiatric comorbidities, family history of OCRDs, and SRI response using SPSS (version 20.0, Armonk, NY, USA). Levene’s test of homogeneity of variances was performed on continuous variables (age, Y-BOCS severity and symptom dimensions).

3.3.4.3 Identifying Symptom Dimensions using Factor and Cluster Analyses on Y-BOCS Symptom Checklist

To evaluate whether this large OCD sample would generate a Y-BOCS severity score factor model similar to those previously reported, an exploratory factor analysis for data reduction was performed on seven obsession categories and eight compulsion categories with a total of 15 a priori categories according to the Y-BOCS Symptom Checklist (Appendix I) using SPSS (version 20.0, Armonk, NY, USA). Lifetime Y-BOCS symptoms (past or present/current) were used for this analysis as per Leckman et al. (Leckman et al., 1997). The miscellaneous obsessive and compulsive symptoms were excluded from the analysis since previous studies have shown that these symptoms are not clearly associated with previously described OCD symptom dimensions (Baer et al., 1994) and an exploratory factor analysis including these latter miscellaneous symptoms did not yield any meaningful results.

We utilized two different coding systems. First, symptoms present at the time of assessment, in the past, or both (i.e., lifetime symptoms) were coded as “1”, and symptoms that were never present were coded as “0”. Second, principal or target symptoms were coded as “2” to give a double weighting whereas symptoms present but not described as target or principal by subjects were coded as “1” and absence of symptoms were coded as “0”. Symptoms were summed for each category. Initial factors were extracted using the principal component analysis (PCA) method, and rotations based on whether the symptoms are related to one another were then performed by the PROMAX method using SPSS (version 20.0, Armonk, NY, USA) 106

107

because the oblique rotation provided by this method is most appropriate if the factors are expected to be correlated with each other (De Geus et al., 2004). Orthogonal rotation such as VARIMAX imposes restriction so that the factors cannot be correlated. Nonetheless, reanalysis using the VARIMAX method yielded similar results. Factor scores of the generated factors (symptom dimensions) for each subject were automatically calculated using SPSS (version 20.0, Armonk, NY, USA).

A cluster analysis using the Ward’s method in SPSS (version 20.0, Armonk, NY, USA) was also performed to compare with the factor analysis. Ward’s method is an agglomerative algorithm, applied in hierarchical cluster analysis for choosing the pair of clusters with the least incremental increase sum of squares to merge at each step. Symptom scores were generated based on the above described criteria (previous two paragraphs) for each of the 15 symptom categories in the Y-BOCS Symptom Checklist. Ward’s method for partitioning the data was used to minimize within-groups variance and to maximize between-groups variance (Abramowitz et al., 2003). Squared Euclidian distance was calculated as the similarity measure. Scores for each symptom category were standardized to correct for range differences within each category.

3.3.4.4 Determining the Presence or Absence of Comorbid Psychiatric Conditions

The frequency of current comorbid psychiatric conditions and lifetime/current OCRDs was determined by the SCID-IV DSM-IV diagnosis. Individuals with primary psychotic disorders and active substance use disorders that pre-dated the onset of OCD symptoms were excluded. Axis II diagnoses were not assessed.

3.3.4.5 Determining the Presence or Absence of Family History of OCRDs

To determine familiality, family history of OCD, OCD with hoarding, HD, TTM, and BDD was obtained for the proband and all identified first-degree relatives through interview of

107

108

the proband and any participating first-degree relatives were collected using the Family History Interview (FHI) questionnaire (Appendix II). Family history was obtained for OCD, OCD with hoarding, HD, TTM, and BDD. Reliability measures were also collected and data obtained were considered reliable if the informant was the subject who could recall his/her own symptoms or if the informant and subject: a) were reasonably close, b) had regular and frequent direct contact, and c) could provide details of subject’s symptoms. If the informant failed the reliability cut-off (no contact, not close to the subject, or indirect contact only with the subject), then two independent sources (including proband or any participating family members) would have to provide the same information or the participants indicated that the information provided had been confirmed. Familiality for each patient is coded either as a “0”, which indicates the absence of family history, or “1”, which refers to the presence of family history.

Familial form of OCD was first defined as OCD individual with a positive family history of OCD. The second familial analysis was defined as OCD individuals with a significant family history of OCRDs. We then examined the relationship between familial status and clinical characteristics including gender, Y-BOCS severity, Y-BOCS symptom dimension factor, family history of OCRDs, and current psychiatric comorbidities using SPSS (version 20.0, Armonk, NY, USA).

3.3.4.6 Determining Responder or Non-responder Status of Antidepressant Response

The response rate of the OCD sample was explored as a categorical dichotomous variable using frequency.

We investigated the relationship between SRI response status and clinical characteristics including gender, Y-BOCS severity, Y-BOCS symptom dimension factor, family history of OCRDs, and current psychiatric comorbidities using SPSS (version 20.0, Armonk, NY, USA).

108

109

3.4 Results

Subject demographics for the OCD sample have been reported in Table 3.1. In brief, the final number of OCD sample that had enough data for analysis is 519 with a mean age of 44 years, a mean AAO of 13.6 years, and a mean lifetime Y-BOCS severity score of 26.0. Almost 58% of the subjects were female and approximately 90% of the sample identified themselves as Caucasian.

Table 3.1. Subject Demographics.

Descriptors Mean (S.D.) N (final data set) 519 (probands) Age 44 (12.8) Gender 57.6% female Ethnicity 89.6% Caucasian 3.3% Asian 2.5% East Indian 0.7% African American 3.9% Others Age at Onset 13.6 (8.0) Yale-Brown Obsessive Compulsive Scale (Y-BOCS) severity score 26.0 (6.9) OCD subject demographics are shown in this table.

3.4.1 Does Gender Contribute to Clinical Diversity in OCD?

Subjects were divided according to their gender (Table 3.2). We observed significant greater number of comorbid lifetime history of SPD and TTM (P<0.001) in female OCD subjects than males (Table 3.2). Otherwise, there were no additional significant differences in clinical characteristics between males and females (Table 3.2) in terms of AAO, Y-BOCS severity and symptom dimensions as well as comorbid current psychiatric conditions, family history of OCRDs, and SRI response.

109

110

Table 3.2. Comparison of Clinical Characteristics between Males and Females.

Male Female Gender Group Mean±SD (N) Mean±SD (N) t P N 211 296 Age 42.7±13.1 (211) 44.9±12.6 (296) -1.967 0.063 AAO 13.2±7.3 (80) 13.6±8.5 (95) -0.266 0.791 Y-BOCS Severity 26.7±7.6 (89) 25.6±6.1 (112) 1.111 0.268 5-Factor 1 (Symmetry/Order) 0.22±1.0 (92) 0.00±1.1 (115) 1.553 0.122 5-Factor 2 0.13±1.0 (92) -0.03±1.0 (115) 1.128 0.261 (Aggression/Sexual/Religious) 5-Factor 3 (Hoarding) -0.06±1.0 (92) 0.12±1.0 (115) -1.276 0.203 5-Factor 4 (Contamination/Cleaning) 0.07±1.0 (92) -0.02±1.0 (115) 0.640 0.523 5-Factor 5 (Somatic) -0.06±1.1 (92) 0.04±1.0 (115) -0.728 0.467 6-Factor 1 (Symmetry/Order) 0.24±1.0 (92) 0.01±1.0 (115) 1.597 0.112 6-Factor 2 (Hoarding) -0.04±1.0 (92) 0.09±1.0 (115) -0.954 0.341 6-Factor 3 (Contamination/Cleaning) 0.05±1.0 (92) -0.01±1.0 (115) 0.366 0.715 6-Factor 4 (Aggression/Checking) 0.11±1.0 (92) -0.01±1.0 (115) 0.852 0.395 6-Factor 5 (Somatic) -0.04±1.1 (92) 0.01±1.0 (115) -0.329 0.743 6-Factor 6 (Sexual/Religious) 0.04±1.1 (92) 0.01±1.0 (115) 0.220 0.826 Disorders Percentage (N) Percentage (N) z/χ2 P Comorbid Lifetime OCRDs 15.5 (30) 42.6 (116) -0.308 0.757 - SPD 5.7 (11) 19.5 (53) -5.431 <0.001 - BDD 5.7 (11) 9.9 (27) -2.107 0.035 - TTM 3.6 (7) 12.5 (34) -3.967 <0.001 - HD 0.5 (1) 0.7 (2) -0.373 0.711 * Comorbid Current Psychiatric (194) (343) 0.699 0.403 Disorders (pure vs comorbid) - Anxiety disorders 55.4 (107) 60.7 (165) 0.450 0.646 - Mood disorders 38.5 (55) 33.1 (90) -0.142 0.889 - Tics/Tourette 8.3 (16) 4.8 (13) 1.911 0.056 - Others 8.3 (16) 15.8 (43) -1.898 0.057 - Pure OCD 9.3 (18) 11.8 (32) -0.836 0.401 Family History of OCRDs 19.7 (15/76) 21.6 (21/97) -0.308 0.757 - OCD 18.4 (13) 17.5 (13) 0.010 0.915 - HD 1.3 (1) 1.0 (1) 0.020 0.889 - TTM 0 (0) 1.0 (1) 0.670 0.414 - BDD 1.3 (1) 2.1 (2) 0.100 0.757 - SPD 0 (0) 0 (0) 0.000 1.000 SRI Response N N χ2 P Responders 33 57 0.423 0.515 Non-Responders 13 17 HD: hoarding disorder; TTM: trichotillomania; BDD: body dysmorphic disorder; SPD: skin picking disorder; SD: standard deviation Highlighted in yellow indicates significant P value * χ2 test was performed comparing the number of pure OCD and OCD with comorbid conditions Levene’s test of homogeneity of variances was not significant for age, Y-BOCS severity and symptom dimensions

110

111

3.4.2 Does AAO Present with Multiple Normal Distributions?

AAO in our sample ranged from 2 to 62 years with a mean of 13.6 and a standard deviation of 8.0. The overall sample showed a non-normal distribution of AAO in a histogram with kernel density and normal curve (Figure 3.1). Using the DENORMIX module in the STATA program (version 11.0, Texas, USA), the admixture analysis was performed and yielded a combination of one, two, three, or four normal theoretical distributions. The χ2 goodness-of-fit test for one-component was 81.824 (P=0.000), two-component was 8.488 (P=0.037), three- component was 6.116 (P=0.047), and four-component was 6.682 (P=0.035). The best-fit model with the highest P value (P=0.047) was the three-component model with means of 5.379, 11.583, and 22.291, standard deviations of 1.594, 3.426, and 8.726, and proportions of 0.196, 0.413, and 0.390, for the three normal distributions (Figure 3.2). The point of intersection between the first and second normal distributions is 8.265 and between the second and third normal distributions is 17.499. Therefore, subjects were divided into three separate groups according to AAO: early (≤8 years), intermediate (between 9 to 17 years), and late (≥18 years). Early onset group had significantly younger age, higher symmetry/order and contamination/cleaning symptom dimensional scores than intermediate or late onset group (Table 3.3). Otherwise, there were no significant differences in other clinical characteristics including Y-BOCS severity, psychiatric comorbidities, family history of OCRDs, and SRI response between the early, intermediate, and late AAO groups (Table 3.3).

3.4.3 Y-BOCS Symptom Dimension – Factor Model and Clusters

Table 3.4 and Table 3.5 provide the PROMAX rotated factor structure that created five distinct factors after treating all symptoms equally and six different factors after double- weighting target symptoms. Approximately 65% of the variance was explained by the 5-factor model whereas 70% of the variance for the 6-factor model. The cluster analyses confirmed both factor models that were generated from the PCA (Figure 3.2 and Figure 3.3). For the 5-factor model: factor 1 includes symmetry obsessions, ordering and arranging, counting, repeating, and checking compulsions; factor 2 comprises of sexual and religious obsessions, and aggressive obsessions and compulsions; factor 3 consists of hoarding and saving obsessions, hoarding and

111

112

collecting compulsions; factor 4 contains contamination obsessions and cleaning compulsions; and factor 5 includes somatic obsessions and compulsions. For the 6-factor model: factor 1 consists of symmetry obsession, ordering and arranging, repeating, and counting compulsions; factor 2 contains hoarding and saving obsessions, hoarding and collecting compulsions; factor 3 comprises of contamination obsessions and cleaning compulsions; factor 4 includes aggressive obsessions and compulsions, and checking compulsions; factor 5 includes somatic obsessions and compulsions; factor 6 consists of religious and sexual obsessions.

3.4.4 Rate of Psychiatric Comorbidities

Of 560 OCD participants, 520 had full SCID-IV data. Current comorbid psychiatric conditions were observed in 58.8% of the OCD participants and 23.7% had lifetime comorbid OCRDs (Figure 3.5 for current psychiatric diagnoses and Figure 3.6 for lifetime OCRDs). The top three current psychiatric comorbid conditions were , major depressive disorder, and hoarding disorder.

3.4.5 Rate of Significant Family History of OCRDs

We had 373 OCD subjects with one or more participating family members who provided FHI data. Approximately 86% of the OCD individuals had one informant, 10% had two, 4% had three, and 1% had four informants. Over 21% of the interview data was discarded due to unreliable recall. Within our OCD sample, 21.7% has significant family history of obsessive- compulsive and related disorders (OCRDs) including OCD, HD, TTM, and BDD (Figure 3.7). Positive family history was noted for the following disorders: OCD (15.5%; N=77), OCD with hoarding (1.4%; N=7), HD (2.0%; N=10), TTM (1.2%; N=6), and BDD (1.6%; N=8). Of the OCD subjects who reported positive family history of OCRDs, only one also had a diagnosis of TTM. Significant family history of OCRDs (except for OCD) did not appear to correlate with OCD subjects’ comorbidity status.

When comparing the clinical characteristics between familial and non-familial OCD individuals, first based on a positive family history of OCD only (Table 3.6), we detected trend 112

113

differences with more aggressive/sexual/religious symptoms, higher total comorbid psychiatric disorders and lower pure OCD cases in the familial OCD group. When combining the family history of other OCRDs (Table 3.7), we detected significantly lower number of pure OCD cases (participants who only had a diagnosis of OCD without any comorbid psychiatric conditions) in the familial group when compared to the non-familial group (P=0.035). There were also trend associations of higher comorbid anxiety disorders and tics in the familial group.

3.4.6 Can Clinical Characteristics Predict Antidepressant Response?

For the drug response data, of the 310 OCD subjects who completed the pharmacogenetic questionnaire, only 66% had an adequate trial of SRI (N=204) (also excluded those who could not tolerate the medication and therefore did not receive an adequate trial), which significantly reduce our sample. Of these 66% of OCD participants, over 92% had good or excellent recall (N=188).

After grouping all five SSRIs or all six SRIs (SSRIs and clomipramine) together, the response rate in our OCD sample is 70.9% and 74.7% respectively. When comparing clinical characteristics between responders and non-responders to SRI(s) (Table 3.8), we detected significantly more symmetry/order symptoms (P=0.008), higher rates of comorbid current psychiatric disorders (P=0.007) especially mood disorders (P=0.004), and higher rates of positive family history of OCD (P=0.042) and hoarding (P=0.002) in the SRI responder group. Responders had significantly higher rate of pure OCD (without any comorbid psychiatric disorders) (P=0.028).

113

114

Figure 3.1. Frequency Distribution of Age At Onset (AAO). This graph illustrates the frequency distribution of AAO in the OCD sample.

114

115

Age At Onset (AAO) Three Normal Distributions

0.5

0.4

0.3

0.2 Probability Density Probability

0.1

0 5.379058 11.58332 22.29146 62 Age at Onset (years)

Figure 3.1. Normal Distributions of Age At Onset (AAO). This graph illustrates three normal distributions within the AAO distribution in the OCD sample.

115

116

Table 3.3. Comparison of Clinical Characteristics between Age At Onset (AAO) Groups.

Early (≤8 years) Intermediate (between 9-17 years) Late (≥18 years) F P AAO Group Mean±SD (N) Mean±SD (N) Mean±SD (N) N 58 70 47 Gender 41.4% Male 50% Male 44.7% Male 0.291 0.864 Age 33.2±10.4 35.5±12.5 39.5±9.4 4.403 0.014 (4.371) (0.014) AAO 5.8±1.7 12.4±2.2 24.4±5.4 N/A N/A Y-BOCS Severity 26.4±7.1 (58) 26.5±6.9 (69) 25.3±6.6 (47) 0.500 0.608 5-Factor 1 (Symmetry/Order) 0.40±0.9 (58) 0.16±1.0 (70) -0.43±0.9 (47) 10.206 <0.001 (8.700) (<0.001) 5-Factor 2 0.25±1.0 (58) 0.16±1.0 (70) -0.14±0.8 (47) 2.305 0.103 (Aggression/Sexual/Religious) 5-Factor 3 (Hoarding) 0.09±1.0 (58) 0.14±1.0 (70) -0.11±1.0 (47) 0.928 0.397 5-Factor 4 (Contamination/Cleaning) 0.37±1.1 (58) -0.03±1.0 (70) -0.28±1.0 (47) 5.472 0.005 (5.724) (0.004) 5-Factor 5 (Somatic) 0.07±1.1 (58) 0.00±1.1 (70) -0.05±1.0 (47) 0.188 0.828 6-Factor 1 (Symmetry/Order) 0.37±1.0 (58) 0.16±1.1 (70) -0.29±0.9 (47) 6.004 0.003 (4.354) (0.014) 6-Factor 2 (Hoarding) 0.04±0.9 (58) 0.08±1.0 (70) -0.02±1.0 (47) 0.134 0.875 6-Factor 3 (Contamination/Cleaning) 0.38±1.1 (58) -0.09±1.0 (70) -0.21±1.0 (47) 5.219 0.006 (5.171) (0.007) 6-Factor 4 (Aggression/Checking) 0.31±1.0 (58) 0.01±0.9 (70) 0.05±0.9 (47) 1.615 0.202 6-Factor 5 (Somatic) -0.03±0.9 (58) 0.07±1.1 (70) -0.15±0.9 (47) 0.672 0.512 6-Factor 6 (Sexual/Religious) 0.04±1.1 (92) 0.19±1.0 (70) -0.20±0.9 (47) 2.110 0.124 2 Disorders Percentage (N) Percentage (N) Percentage (N) z/χ P Comorbid Lifetime OCRDs 28.6 (14/49) 22.6 (14/62) 15.6 (5/32) 2.422 0.298 - SPD 12.2 (6) 12.9 (8) 9.4 (3) - BDD 16.3 (8) 4.8 (3) 3.1 (1) - TTM 0 (0) 4.8 (3) 3.1 (1) - HD 0 (0) 0 (0) 0 (0) Comorbid Current Psychiatric Disorders 100 (61) 81.4 (57/70) 59.6 (28/47) 3.663 0.160 - Anxiety disorders 60.3 (35) 52.9 (37) 40.4 (19) 0.679 0.712

116

117

- Mood disorders 31.0 (18) 21.4 (15) 14.9 (7) 2.268 0.322 - Tics/Tourette 6.9 (4) 5.7 (4) 0 (0) - Others 6.9 (4) 1.4 (1) 4.3 (2) Family History of OCRDs 25.5 (13/51) 24.2 (15/62) 12.5 (4/32) 2.422 0.298 - OCD 21.6 (9) 17.7 (11) 9.4 (3) 2.046 0.360 - HD 2.0 (2) 4.8 (3) 3.1 (1) - TTM 1.9 (1) 0 (0) 0 (0) - BDD 1.9 (1) 1.6 (1) 0 (0) - SPD 0 (0) 0 (0) 0 (0) 2 SRI Response N N Percentage (N) χ P Responders 13 10 9 0.605 0.739 Non-Responders 5 6 3 HD: hoarding disorder; TTM: trichotillomania; BDD: body dysmorphic disorder; SPD: skin picking disorder Highlighted in yellow indicates significant P value and results remained significant after including gender and age as covariates in a linear regression model (in bracket) Levene’s test of homogeneity of variances was not significant for age, Y-BOCS severity and symptom dimensions

117

118

Table 3.4. Y-BOCS Symptom Dimension Factor Analysis (weighing all symptoms as equal).

118

119

Table 3.5. Y-BOCS Symptom Dimension Factor Analysis (double weighing target symptoms).

119

120

Y-BOCS Symptom Dimensions Cluster Analysis

Hoarding obsession Hoarding compulsion

Ordering/Arranging

Counting compulsion

Somatic compulsion

All symptoms treated equally as 1

Figure 3.2. Y-BOCS Symptom Dimension Cluster Analysis (weighing all symptoms as equal). This figure illustrates the grouped symptom dimensions using cluster analysis when weighing all symptoms equally.

120

121

Y-BOCS Symptom Dimensions Cluster Analysis

Hoarding obsession Hoarding compulsion

Ordering/Arranging

Counting compulsion

Somatic compulsion

Target Symptoms = 2 Identified Symptoms = 1 No Symptoms = 0 Figure 3.3. Y-BOCS Symptom Dimension Cluster Analysis (double weighing target symptoms). This figure illustrates the grouped symptom dimensions using cluster analysis when weighing target and other symptoms differently.

121

122

Figure 3.4. Current Psychiatric Comorbidities of OCD Sample. Percentage is indicated in brackets. This graph illustrates the count and percentage of current psychiatric comorbid conditions in the OCD sample.

122

123

Figure 3.5. Lifetime Comorbid OCRDs of OCD Sample. Percentage is indicated in brackets. This graph illustrates the count and percentage of lifetime obsessive-compulsive and related disorders in the OCD sample.

123

124

21.7% + Family History

1.4% 15.5% 2.0% 1.6% 1.2% OCD with OCD HD BDD TTM Hoarding

Figure 3.6. Family History of OCD Sample This figure illustrates the percentage of family history of obsessive-compulsive and related disorders in the OCD sample.

124

125

Table 3.6. Comparison of Clinical Characteristics between Familial and Non-Familial (OCD) OCD Individuals.

Familial Non-Familial Familial Group Mean±SD (N) Mean±SD (N) t P N 85 288 37.8% Male 39.9% Male -0.348 0.726 Gender (31/82) (115/288) Age 35.5±12.2 (80) 36.0±12.9 (274) 0.324 0.746 AAO 12.0±5.5 (20) 13.4±7.9 (112) 0.771 0.425 Y-BOCS Severity 27.3±4.6 (23) 26.9±6.7 (127) -0.221 0.826 5-Factor 1 (Symmetry/Order) -0.13±1.0 (23) 0.16±1.1 (130) 1.230 0.221 5-Factor 2 0.52±1.2 (23) 0.08±1.0 (130) -1.925 0.056 (Aggression/Sexual/Religious) 5-Factor 3 (Hoarding) -0.03±1.0 (23) 0.11±1.0 (130) 0.615 0.539 5-Factor 4 (Contamination/Cleaning) -0.02±1.1 (23) 0.13±1.1 (130) 0.584 0.560 5-Factor 5 (Somatic) 0.00±1.1 (23) 0.10±1.1 (130) 0.398 0.691 6-Factor 1 (Symmetry/Order) -0.01±1.1 (23) 0.14±1.0 (130) 0.683 0.496 6-Factor 2 (Hoarding) -0.13±0.9 (23) 0.09±1.0 (130) 1.017 0.311 6-Factor 3 (Contamination/Cleaning) -0.12±1.1 (23) 0.11±1.1 (130) 0.925 0.356 6-Factor 4 (Aggression/Checking) 0.23±1.1 (23) 0.14±0.9 (130) -0.432 0.666 6-Factor 5 (Somatic) 0.10±1.1 (23) 0.09±1.1 (130) -0.047 0.963 6-Factor 6 (Sexual/Religious) 0.55±1.2 (23) 0.01±1.0 (130) -2.271 0.025 Disorders Percentage (N) Percentage (N) z/χ2 P * Comorbid Current Psychiatric 62.4 (53/85) 52.1 (150/288) 2.791 0.095 Disorders (pure vs comorbid) - OCRDs 18.8 (16) 19.1 (55) -0.057 0.952 - Anxiety disorders 40.0 (34) 30.6 (88) 1.631 0.103 - Mood disorders 17.6 (15) 16.0 (46) 0.367 0.711 - Tics/Tourette 5.9 (5) 4.9 (14) 0.376 0.704 - Others 12.9 (11) 6.6 (19) 1.890 0.059 - Pure OCD 37.6 (32) 47.9 (138) -1.671 0.095 SRI Response N N χ2 P Responders 21 49 0.060 0.806 Non-Responders 6 16 HD: hoarding disorder; TTM: trichotillomania; BDD: body dysmorphic disorder; SPD: skin picking disorder Highlighted in yellow indicates significant P value Highlighted in grey indicates trend association * χ2 test was performed comparing the number of pure OCD and OCD with comorbid conditions

125

126

Table 3.7. Comparison of Clinical Characteristics between Familial and Non-Familial (OCRDs) OCD Individuals.

Familial Non-Familial Familial Group Mean±SD (N) Mean±SD (N) t P N 108 265 36.2% Male 40.8% Male -0.810 0.418 Gender (38/105) (108/265) Age 36.0±12.2 (102) 35.8±12.9 (252) -0.126 0.900 AAO 11.9±5.7 (26) 13.5±8.0 (106) 0.953 0.342 Y-BOCS Severity 28.3±5.8 (30) 26.7±6.6 (120) -1.249 0.214 5-Factor 1 (Symmetry/Order) 0.14±1.1 (30) 0.11±1.0 (123) -0.140 0.889 5-Factor 2 0.39±1.2 (30) 0.09±0.9 (123) -1.488 0.139 (Aggression/Sexual/Religious) 5-Factor 3 (Hoarding) 0.11±1.0 (30) 0.08±1.0 (123) -0.128 0.898 5-Factor 4 (Contamination/Cleaning) -0.03±1.1 (30) 0.14±1.1 (123) 0.737 0.462 5-Factor 5 (Somatic) 0.11±1.1 (30) 0.08±1.1 (123) -0.171 0.865 6-Factor 1 (Symmetry/Order) 0.18±1.1 (30) 0.11±1.0 (123) -0.326 0.745 6-Factor 2 (Hoarding) -0.01±0.9 (30) 0.08±1.0 (123) 0.423 0.673 6-Factor 3 (Contamination/Cleaning) -0.12±1.1 (30) 0.12±1.1 (123) 1.132 0.259 6-Factor 4 (Aggression/Checking) 0.15±1.1 (30) 0.15±1.0 (123) 0.041 0.968 6-Factor 5 (Somatic) 0.24±1.1 (30) 0.05±1.0 (123) -0.906 0.366 6-Factor 6 (Sexual/Religious) 0.39±1.1 (30) 0.02±1.0 (123) -1.751 0.082 2 Disorders Percentage (N) Percentage (N) z/χ P Comorbid Current Psychiatric 63.0 (68/108) 50.1 (135/265) 1.679 0.195 Disorders (pure vs comorbid) - OCRDs 22.2 (24) 17.7 (47) 1.001 0.317 - Anxiety disorders 39.8 (43) 29.8 (79) 1.868 0.061 - Mood disorders 15.7 (17) 16.6 (44) -0.204 0.841 - Tics/Tourette 8.3 (9) 3.8 (10) 1.817 0.069 - Others 11.1 (12) 6.8 (18) 1.391 0.165 - Pure OCD 37.0 (40) 49.1 (130) -2.114 0.035 2 SRI Response N N χ P Responders 26 44 0.206 0.650 Non-Responders 7 15 HD: hoarding disorder; TTM: trichotillomania; BDD: body dysmorphic disorder; SPD: skin picking disorder Highlighted in yellow indicates significant P value Highlighted in grey indicates trend association * χ2 test was performed comparing the number of pure OCD and OCD with comorbid conditions

126

127

Table 3.8. Comparison of Clinical Characteristics between SRI Responders and Non- Responders.

Responders Non-Responders SRI Response Mean±SD (N) Mean±SD (N) t P N 98 30 Gender 36.7% Male (33/90) 43.3% Male (13/30) -0.650 0.516 Age 38.1±11.6 (84) 38.4±11.5 (30) 0.090 0.928 AAO 13.3±8.2 (35) 13.6±7.9 (14) -0.180 0.857 Y-BOCS Severity 27.9±7.3 (91) 29.9±4.7 (30) -1.768 0.080 5-Factor 1 (Symmetry/Order) -0.16±1.1 (42) 0.43±1.0 (16) -2.760 0.008 5-Factor 2 (Aggression/Sexual/Religious) 0.22±0.9 (42) 0.19±1.1 (16) 0.136 0.892 5-Factor 3 (Hoarding) -0.07±1.0 (42) 0.25±1.1 (16) -1.423 0.165 5-Factor 4 (Contamination/Cleaning) 0.07±1.0 (42) 0.19±1.2 (16) -0.497 0.622 5-Factor 5 (Somatic) 0.16±1.1 (42) 0.40±1.4 (16) -0.861 0.398 6-Factor 1 (Symmetry/Order) -0.16±0.9 (42) 0.41±1.1 (16) -2.586 0.013 6-Factor 2 (Hoarding) -0.02±1.0 (42) 0.07±0.9 (16) -0.467 0.642 6-Factor 3 (Contamination/Cleaning) 0.05±1.0 (42) 0.10±1.1 (16) -0.222 0.826 6-Factor 4 (Aggression/Checking) 0.17±1.0 (42) 0.07±1.0 (16) 0.479 0.634 6-Factor 5 (Somatic) 0.11±1.0 (42) 0.60±1.6 (16) -1.585 0.123 6-Factor 6 (Sexual/Religious) 0.22±1.1 (42) 0.25±1.4 (16) -0.108 0.915 SRI Response Percentage (N) Percentage (N) z/χ2 P Pure OCD 45.9 (45/98) 23.3 (7/30) 2.204 0.028 * Comorbid Current Psychiatric Disorders (80) (35) 7.276 0.007 - OCRDs 16.3 (16) 16.7 (5) -0.044 0.968 - Anxiety disorders 35.7 (35) 46.7 (14) -1.080 0.280 - Mood disorders 15.3 (15) 40.0 (12) -2.901 0.004 - Tics/Tourette 6.1 (6) 3.3 (1) 0.588 0.555 - Others 8.2 (8) 10.0 (3) -0.314 0.757 Family History of OCRDs 40.0 (28/70) 45.5 (10/22) -0.453 0.653 - OCD 10.0 (7) 27.3 (6) -2.029 0.042 - HD 0 (0) 13.6 (3) -3.141 0.002 - TTM 0 (0) 4.5 (1) -1.079 0.073 - BDD 0 (0) 0 (0) 0.000 1.000 - SPD 0 (0) 0 (0) 0.000 1.000 HD: hoarding disorder; TTM: trichotillomania; BDD: body dysmorphic disorder; SPD: skin picking disorder Highlighted in yellow indicates significant P value Highlighted in grey indicates trend association * χ2 test was performed comparing the number of pure OCD and OCD with comorbid conditions

127

128

3.5 Discussion

To our knowledge, this is the single largest North American and most comprehensive clinical OCD study to date that explored demographical data and phenomenological characteristics of OCD including gender, AAO, Y-BOCS symptom dimensions and severity scores, psychiatric comorbidities, family history of OCRDs, and SRI antidepressant response. Our results were comparable to other large clinical OCD studies (Hofmeijer-Sevink et al., 2013; Torresan et al., 2013). The female OCD subjects in this study have a higher prevalence of psychiatric comorbidities especially OCRDs and the earlier AAO group appears to have higher symmetry/order and contamination/cleaning symptoms. Identified symptom dimensions from the 5- or 6-factor model were consistent with previous published studies (please refer to Chapter 1 Table 1.5a for details).

3.5.1 Gender Does Not Contribute to Difference in Clinical Presentations

Our current OCD sample did not detected any significant differences in AAO, symptom severity, symptom dimensions, and family history or OCRDs based on proband gender; however, we observed significantly higher rate of lifetime comorbid TTM and SPD, in females. Although we detected a significant difference of the rate of TTM between male and female individuals with OCD, the published gender ratio of TTM was approximately 10:1 (O’Sullivan et al., 1997; Stein et al., 2010), which was almost three times the rate of our current sample (3.5:1). The difference in the rate of SPD between males and females (3.4:1) was comparable to previous epidemiological studies with ¾ of cases were reported in female (Grant and Odlaug, 2009; Keuthen et al., 2001). However, OCRDs have recently been established as a diagnostic group and besides OCD, TTM and BDD, SPD and HD did not exist as a distinct disorder prior to the DSM-5. Changes made in DSM-5 may potentially affect future research study and there is a need to further explore the rates of OCRDs in the general population with the new diagnostic criteria.

Our findings for the remaining clinical characteristics appeared to contradict previous OCD studies except for one clinical study by Matsunaga et al. (2000), which detected no gender-

128

129

related phenomenological differences in OCD patients. Most studies reported notable gender differences. For example, OCD males had earlier AAO (de Mathis et al., 2011; Jaisoorya et al., 2009; Narayanaswamy et al., 2012; Nestadt et al., 1998; Rasmussen and Eisen, 1992), higher symptom severity (Nestadt et al., 1998; Rosso et al., 2012), more frequent aggressive (Tukel et al., 2013), sexual, religious, symmetry/exactness obsessions, checking, ordering/arranging, and hoarding compulsions (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013). On the other hand, OCD females were observed to endorse higher frequency of contamination obsessions, washing/cleaning compulsions, and somatic obsessions and compulsions (de Mathis et al., 2008; Jaisoorya et al., 2009; Rosso et al., 2012; Tukel et al., 2013), and significantly more comorbid depression (Cherian et al., 2014). One reason for the general lack of differences between males and females in our current OCD sample is likely the sampling bias (which will be expanded in the limitation section below).

3.5.2 AAO has a Tri-modal Distributions

This current study demonstrated that the early-onset OCD group had younger age, higher frequency of symmetry/order and contamination/cleaning symptoms but did not differ in symptom severity, rate of comorbid psychiatric conditions, or rate of family history between the AAO groups. From a clinical perspective, individuals with early onset OCD often seek psychiatric help earlier given the longer duration of suffering when compared to later onset OCD patients and this clinical observation was supported by our finding. To date, only four clinical studies have examined AAO using admixture analysis in the OCD population. Delorme et al. (2005) identified two AAO groups, early-onset as defined by AAO less than or equal to 20 and late-onset as AAO greater than or equal to 21 years. The authors detected a higher rate of comorbid Tourette’s syndrome and greater percentage of positive family history of OCD in the early-onset group, and conversely a higher prevalence of generalized anxiety disorder and major depressive disorder in the late-onset group. The second study used mixture analysis in a subset of the current OCD sample (N=196), which was conducted by our group previously (De Luca et al., 2011). De Luca et al. (2011) detected two normal distributions with a cut-off at 15 years and found that the early-onset OCD patients were assessed at an earlier age and have more frequent panic attacks than the late-onset subjects. Similarly to the current larger sample, we also 129

130

observed an earlier age of assessment for the early-onset group when compared to the intermediate-onset group and for the intermediate-onset group when compared to the late-onset group. Although we did not detect any significant difference in psychiatric comorbidity, our results implicated a higher rate of anxiety disorders in the early-onset than the late-onset group. The third study was conducted by Anholt et al. (2014), who reported a bimodal distribution for AAO with the best-fitting cut-off AAO for early-onset as less than or equal to 19 years. The early-onset group was reported to have greater OCD symptom severity across all OCD symptom dimensions in addition to greater number of ADHD symptoms and higher rate of comorbid bipolar disorder. The last and most recent study utilized two sets of AAO as defined by age at which symptom first appeared and age at which symptom first met diagnostic criteria of OCD (Albert et al., 2015). Albert et al. (2015) also detected a bimodal distribution for AAO at which diagnosis was first met. However, they identified a trimodal distribution for AAO at symptom onset with means similar to our current study and a cut-off of 9 and 24 years (8 and 18 years in this study). The late-onset cut-off appeared to be comparable with these studies, ranging from 18 to 21 years except for Albert et al. (2015) who detected a cut-off at age 26 years. Nonetheless, we defined AAO as onset when OCD was first diagnosed and all three of the above mentioned studies identified a bimodal distribution for AAO using the same definition as our current study. The inconsistency between the present study and the others could be explained by sampling and recruitment bias. Our OCD sample had a higher mean of OCD Y-BOCS severity score (26 in this current study versus 23.6-25.4 for the other studies) and a younger AAO (13.6 years in this current study versus 17.1-22.1 years). These discrepancies suggested that different studies might have recruited different clinical OCD populations depending on the clinical settings and the city or country where the studies were conducted. Similar to the lack of gender differences, the general lack of differences between the AAO groups in the present study is also likely due to sampling and recruitment bias in a tertiary and highly specialized OCD clinical setting.

3.5.3 Factor and Cluster Analyses Confirmed a 5- and 6-Factor Model of Y-BOCS Symptom Dimensions

In this study, we demonstrated that OCD symptom dimensions could be reliably generated from the Y-BOCS symptom checklist using either the factor or cluster analysis. We 130

131

used two different coding methods, either weighing all OCD symptoms equally or double- weighing the target or principal symptoms, and identified five or six factors with the addition of somatic obsessions and compulsions to the consistently reported symptom factors or dimensions – symmetry/ordering, hoarding, contamination/cleaning, aggressive/checking+/-sexual/religious concerns from previous studies (Table 1.6, Table 1.7, Table 1.8). We detected an extra symptom dimension in both our 5- and 6-factor models, somatic subgroup, which is in concert with two previous studies (Stein et al., 2007; Stein et al., 2008). Somatic symptoms have been frequently grouped with aggressive/sexual/religious/checking obsessions and compulsions (Denys et al., 2004; Hasler et al., 2007; Lochner et al., 2008; Pinto et al., 2007). Liu et al. (2011), who weighed all symptoms equally, reported somatic and repeating as one symptom dimension, with hoarding, contamination/cleaning, symmetry/ordering, and aggressive/checking as the remaining four factors. Zhang et al. (2013) detected somatic symptoms as a separate dimension out of six different identified factors only when examining Y-BOCS symptom checklist at an item level rather than the commonly used categorical level. Most studies did not report somatic symptoms as the one of the prominent symptom dimensions within the 4 or 5 factors (Table 1.6 and Table 1.7).

Although the current study derived a 5-factor model that was consistent with most published studies, we did not replicate the identified symptom dimensions and the generated factors appeared to be consistent with only the minority of the studies. Furthermore, from a clinical perspective, somatic symptoms are very closely related to hypochondriasis. Recurrent health-related concerns with frequent visits to the family doctor’s office might not be perceived as a symptom of OCD but rather a somatic symptom within the DSM-IV somatization disorder category; thus, some of the previous studies might have under-detected this symptom domain. Another possible reason for a more common presentation of somatic symptom is the location of the recruitment sites. Toronto, the recruiting site for this study, is considered to be one of the most multi-cultural cities in the world with a large percentage, 49.1%, of Toronto’s population as Asians according to the 2011 National Household Survey (Statistics Canada, 2011). Many Asian individuals who suffer from anxiety or depression very often present with somatic complaints (Hinton et al., 2009; Lin and Cheung, 1999; Mumford, 1992). Due to stigma and misconception of mental illness, physical discomforts are more widely acceptable than emotional or behavioural issues in the Asian culture (Lin and Cheung, 1999). Our OCD sample consists of 3.3% of Asian 131

132

individuals, which might potentially affect the results of the factor analysis. Almost all of the previous studies included Caucasian OCD participants although many did not specify the ethnicity of their samples. Only one study has investigated the African American population and detected a 6-factor model that accounted for 59.1% of the variance: contamination and washing, hoarding, sexual obsessions and reassurance, aggression and mental compulsions, symmetry and perfectionism, and doubt and checking (Williams et al., 2012). Although the factors were similar to the previous studies, the authors reported significantly more contamination symptoms and excessive concerns with animals in African Americans than Europeans. These findings further implicate cultural differences in OCD symptom presentation.

3.5.4 High Rates of Psychiatric Comorbidities

Over 50% of our sample’s OCD participants have concurrent psychiatric disorders and over 20% have a lifetime history of OCRD(s). These numbers are consistent with previous published studies. Comorbidity rates in epidemiological studies vary widely, ranging from 19% to 90% depending on the methodological and semantic (i.e., lifetime versus current rates) differences (Overbeek et al., 2002; Pallanti et al., 2011). However, one key concept is that OCD individuals often present with comorbid psychiatric conditions, which can often influence treatment plan and outcome. Patients who are non-responders to pharmacological and psychological treatments are more likely to meet criteria for comorbid Axis I or Axis II disorders (Pallanti et al., 2011).

3.5.5 High Rate of Family History of OCRDs

Given that OCRDs formed a new and separate section in the recently published DSM-5 (APA, 2013), many previous studies have not investigated the presence or absence family history of OCRDs in OCD samples. Approximately 22% of our sample’s OCD probands reported having one or more first-degree relative who suffers from at least one OCRD. This is not surprising because of the similarities these disorders share. Many studies have reported a prevalence rate ranging from 7% to 37% of first-degree relatives of OCD individuals who also

132

133

have a diagnosis of OCD (Nestadt et al., 2010) and the present study detected 15.5% of OCD individuals having a positive family history of OCD. However, none has ever investigated the prevalence of OCRDs in first-degree relatives of OCD patients. This is likely because most research studies still utilize the DSM-IV diagnostic criteria, which do not account for SPD and HD and with slight differences in the diagnostic criteria of BDD and TTM. Hence, this is the first study, which reported 6.2% of OCD individuals with a positive family history of OCRDs other than OCD. Interestingly, after further exploration of the probands with positive family history of OCRDs, we only detected one individual with comorbid TTM who had a family history of TTM. The rest of the probands did not have congruent personal and family history of the same disorders (SPD, BDD, and HD). This is intriguing given that this study is the first to show a general liability of OCRDs in OCD patients without a comorbid diagnosis of other OCRDs.

Although the number of the comparison groups was small, we observed clinical differences between familial and non-familial form of OCD. The results for incorporating OCD only and other OCRDs together did not differ dramatically. Higher comorbidity was detected in the familial group in general when compared with the non-familial group, suggesting that family history may allow researchers to identify more homogeneous subgroups of OCD.

3.5.6 Antidepressant Response Rate is Higher than Published Rates and Differences were Observed between SRI Responders and Non-Responders

Although our OCD subjects were recruited from a tertiary and highly specialized OCD clinic, the SRI response rate is relatively higher, at over 70%, than other published rates of 40- 60% (Foa et al., 2005; Greist et al., 1995; McDougle et al., 1993). This could be explained by several factors: the definition of drug response and study design. Firstly, these three cited studies above (Foa et al., 2005; Greist et al., 1995; McDougle et al., 1993) have defined response as OCD symptom reduction of at least 25% to 35% according to the Y-BOCS severity score (Tolin et al., 2005), whereas our current study has employed the CGI-I scale to report response rate. Documented Y-BOCS severity score may provide a more objective SRI response that is specific

133

134

for OCD whereas the CGI-I scale offers a readily understood and practical measurement tool that can easily be administrated by any physicians but it lacks specificity to OCD and objectivity of a change in SRI response over time. Secondly, Y-BOCS severity scores are often documented prospectively by a physician who monitors the progress of an SRI treatment trial; however, in our current study, most of the response data have been generated retrospectively or by chart review; thus, recall bias could explain the small difference in SRI response. We attempted to address the limitation of recall bias by introducing a reliability measure. Thirdly, one additional possible reason for a higher response rate in this sample could be the increasing awareness of OCD in general, leading to referral of less severe cases to the tertiary clinic, which may be more representative of the general OCD population.

This study revealed clear clinical prognostic predictors of SRI response. Poor SRI response was associated with symmetry/order symptoms, higher number of comorbid current psychiatric disorders especially mood disorders, and family history of OCD and HD. Previous treatment studies have examined clinical factors that predict antidepressant response in OCD. Denys et al. (2003) reported the absence of previous treatment, moderate symptom severity (Y- BOCS severity score < 23), and low depressive symptoms (Hamilton Depression Rating Scale score between 6 and 15) as good prognostic factors of antidepressant response. Although we did not examine previous treatment history, we observed a trend for lower severity and a significantly lower number of comorbid current mood disorders in responders when compared with non-responders. Another study examined additional clinical predictors of SSRI response and detected higher level of baseline symptom severity and disability in non-responders (Tükel et al., 2006). Marazziti and Consoli (2010) provided a comprehensive review on treatment strategies for OCD and reported several robust positive clinical predictors of SSRI response, which included later onset of illness, less symptom severity, absence of previous treatment, and moderate severity of depression; however, the authors concluded that the strongest symptom predictor of poor treatment response was hoarding. Although inconsistent findings have been reported for comorbid tics as influencing SRI response (Husted et al., 2007; Skarphedinsson et al., 2015), the presence of tics (Bloch et al., 2006; Masi et al., 2013; McDougle et al., 1994; Nakamae, 2011) and family history of tics (McDougle, 1997) both predict a better response to antipsychotic augmentation strategy. Thus, in addition to examining SRI response, determining clinical predictors of antipsychotic response and response to other treatment modalities such as 134

135

psychotherapy and transmagnetic stimulation will allow us to move beyond standard treatment protocol to individualized medicine. Further investigation on other clinical and demographic variables including duration of illness and number of previous treatment is warranted to provide a comprehensive treatment guide that is tailored to each individual patient.

3.5.7 Limitations of the Clinical Study

One limitation of our study is regarding data collection. There are missing data within the current sample, which is often unavoidable in large research dataset. Many reasons could contribute to missing data and they are generally categorized into two main areas, participant factors and researcher factors. Participants may have lost interest given the lengthy assessment and complicated interviewing questions. The duration of the interview necessitated more than one visit for some subjects, and some probands did not return to complete the research assessment. Missing observations might occur due to human error or the lack of man-power to transfer written documents into electronic databases. Additionally, missing data could mislead our current findings and further reduce the power of our sample. Another limitation related to data collection is that the clinical data from this study including AAO, lifetime Y-BOCS severity, presence or absence of Y-BOCS OCD symptoms, family history, and comorbid psychiatric conditions, in addition to SRI antidepressant treatment response data were ascertained in a retrospective manner and chart review and there is a heightened chance of recall bias and unreliable assessment. We attempted to minimize the effect of recall bias by incorporating a reliability measure. Although using the reliability measure drastically reduces the number of positive family history and complete drug response data in our sample, this is a more stringent way to reduce recall bias but with the risk of increased missing data. For the family history interview, 21.5% of the interview data was discarded due to unreliable recall. For the drug response data, 33% had inadequate trial of SRI and an additional 8% of these individuals reported poor recall, which significantly reduce our sample. Prospective study will ultimately eliminate the risk of prematurely discontinuing SRI at a suboptimal dose (adequacy of trial) and recall bias by utilizing objective measures to quantify treatment response and to follow participants in addition to their relatives longitudinally. Therefore, prospective clinical and pharmacogenetic studies are warranted in order to clarify causality in defining clinical outcome 135

136

and treatment response. The third limitation of this study is the tertiary-centre sample, which may not fully represent the general OCD population but is rather a relatively more severe OCD sample with higher comorbidity; however, the rate of treatment responders is higher than the published rate, which may suggest either the contrary or methodological differences between the published studies. Nonetheless, our results suggested that this OCD sample is comparative to other large published OCD clinical studies with similar rates of major comorbidity.

3.5.8 Conclusion

The results of this study demonstrated that there may be phenotypically different and homogeneous subgroups of OCD, which may be scientifically relevant to be examined in genetic studies of OCD. Gender differences, early versus late AAO, symptom severity, symptom dimensional subtypes, positive versus negative family history, different psychiatric comorbidities, and antidepressant responders versus non-responders are clinically informative factors that may be associated with specific vulnerability pattern, illness chronicity, and illness severity. By combining associated genetic variations from ongoing OCD genetic studies with clinical and demographic risk factors, we may be able to develop a risk model for OCD. Once this model has been validated, it will be a useful tool for clinical application. A better understanding of clinically homogeneous OCD subgroups or subphenotypes has the potential to substantially improve treatment outcomes and to promote the development of new therapies. Further research should focus on the identification of homogeneous subgroups for genetic research and for treatment outcome study.

136

137

Chapter 4 CANDIDATE GENE STUDY OF OCD PHENOTYPES

4 Candidate Gene Study of Obsessive-Compulsive Disorder Phenotypes

Gwyneth Zai 1,2,4, Clement Zai 1,2, Vanessa Gonçalves 1,2, David Pauls 3, Karen Wigg 4, Margaret A. Richter 4, James L. Kennedy 1,2

1 Neurogenetics Section, Centre for Addiction and Mental Health, Clarke Division, Toronto, ON, Canada

2 Department of Psychiatry, University of Toronto, Toronto, ON, Canada

3 Psychiatric and Neurodevelopmental Genetics Unit, Center for Human Genetic Research, Massachusetts General Hospital, Boston, MA, USA

4 Frederick W. Thompson Anxiety Disorders Centre, Department of Psychiatry, Sunnybrook Health Sciences Centre, Toronto, ON, Canada

* Corresponding author: Dr. James L. Kennedy, Director of the Neuroscience Department and

Head of Neurogenetics Section at the Centre for Addiction and Mental Health (CAMH),

Professor of the Department of Psychiatry and the Institute of Medical Science at University of

Toronto, Toronto, ON M5T 1R8, Canada (Phone: 1-416-979-4987; Fax: 1-416-979-4666; E- mail: [email protected])

137

138

4.1 Abstract

Background: Obsessive-compulsive disorder (OCD) is a heterogeneous illness with a strong genetic etiology. However, genetic studies in obsessive-compulsive disorder (OCD) have yielded inconsistent results and two recent genome-wide association studies (GWASs) did not report any genome-wide significant results. Using homogeneous phenotypes in OCD may aid in the identification of risk genes that may contribute to the susceptibility to OCD. We aimed to examine the genetics of various phenotypes in OCD including age at onset (AAO), Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score and symptom dimensions, psychiatric comorbidities, and the presence or absence of family history.

Method: A candidate gene study of 497 OCD individuals with 336 first-degree relatives was conducted using the 32-SNP QuantStudio array. Statistical analyses were performed using PLINK and R programs.

Results: While some markers showed nominal significance, none of the findings remained significant for any phenotypes after correcting for multiple comparisons.

Conclusions: The use of more refined OCD phenotypes in genetic study may help to identify common genetic variants that contribute to risk of developing OCD but larger and well- characterized samples are required to provide more definitive conclusion.

4.2 Introduction

Obsessive-compulsive disorder is a serious and debilitating psychiatric disorder, which is characterized by recurrent, persistent, and unwanted obsessions and/or repetitive compulsions (APA, 2013). It affects approximately 1-3% of the general population and causes significant distress and impairment in those affected with this condition (Ruscio et al., 2010). The World Health Organization ranked OCD within the top ten most disabling illnesses in terms of lost earnings and reduced quality of life (WHO, 2009). However, current first-line pharmacological treatment of OCD offers up to 60% of response (McDougle et al., 1993). Thus, advances in

138

139

identifying the underlying mechanisms of OCD are vital to the development of new treatment therapies.

Support for a genetic etiology in obsessive-compulsive disorder (OCD) has come from family, twins, aggregation, candidate gene, and genome-wide association studies (GWASs) (Pauls et al., 2014). The estimated heritability for OCD ranges from 27% to 65% (van Grootheest et al., 2008). Family studies have reported rates of OCD among relatives of OCD individuals as two- to ten-fold higher when compared to control relatives (Pauls, 2010). Segregation studies demonstrated strong familial genetic rather than sporadic and environmental risks as an etiology of OCD (Alsobrook et al., 1999; Cavallini et al., 1999; Nestadt et al., 2010; Pauls et al., 2014). Although preliminary findings have suggested a major role of the glutamatergic system in the neurobiological basis of OCD, results from the existing GWASs and candidate gene studies have not yielded replicable results in any one single glutamate-related gene linking to the diagnosis of OCD (Pauls et al., 2014). Researchers have implemented a strategy to develop new ways of classifying psychiatric disorders based on behavioural and neurobiological measures (Etkin and Cuthbert, 2014). This is an attempt to move away from diagnosis-based category to specific subphenotypes of the illness. Individuals suffering from OCD often have complex and diverse clinical presentations given its heterogeneous phenomenology and high rate of psychiatric comorbidities. In order to dissect the genetics of homogeneous OCD subphenotypes, large sample is required to stratify these different OCD homogeneous subgroups for genetic analysis.

Many different studies with relatively small sample size have examined the genetic effect of OCD subphenotypes including age at onset (AAO), Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity and symptom dimensions, in addition to psychiatric comorbidities (Table 1.5, Table 1.9, Table 1.10, and Table 1.12). Only one study has investigated the genetics of familiality (significant family history of OCD) in OCD (Denys et al., 2006) and two studies have investigated the brain-derived neurotrophic factor (BDNF) and the catechol-O-methyltransferase (COMT) genes respectively with the presence of obsessive-compulsive symptoms (subclinical) in the first-degree relatives (Katerberg et al., 2009; Katerberg et al., 2010).

139

140

Evidence supports a stronger genetic loading for early-onset OCD, which was consistently found to be phenotypically different from (Dell’ Osso et al., 2013; de Mathis et al., 2011; Jaisoorya et al., 2009; Narayanaswamy et al., 2012; Nestadt et al., 1998; Rasmussen and Eisen, 1992; Taylor, 2011) and has a higher heritability (van Grootheest et al., 2008) than late- onset OCD. Forty-seven genes have been examined in relation to AAO in OCD but mixed and inconsistent genetic results have been reported across these different neurotransmitter system candidate genes (Table 1.5). Two significant meta-analyses investigated the serotonin transporter (SLC6A4) 5HTTLPR polymorphism (Bloch et al., 2008) and reported an over- transmission of the La allele to early-onset OCD (P=0.00021) (Walitza et al., 2014). Another serotonin gene, the serotonin 2A receptor (HTR2A), was investigated in nine published studies with mixed results (Table 1.5). One linkage study reported a LOD score of two on chromosome 9q where the neuronal glutamate transporter gene (SLC1A1) is located (Hanna et al., 2002) and three of five studies reported significant findings but each with different SNPs, which were associated with early-onset OCD [rs3780412 and rs301430 (P=0.04) (Dickel et al., 2006); rs10491734 A allele (P=0.023) and A/A genotype (P=0.009) (Wu et al., 2013); rs301430 C/C genotype with later onset (P=0.017) (Dallaspezia et al., 2014)]. The brain-derived neurotrophic factor gene (BDNF) was also previously implicated in AAO in OCD. The first significant study was reported by Hall et al. (2003) and only two of eight subsequent studies supported the role of the BDNF rs6265 marker in AAO and the findings were confounded with gender bias (Hemmings et al., 2006; Katerberg et al., 2009). Although genetics of OCD AAO phenotype have previously been explored, the cut-off between early- and late-onset was subjectively defined. We therefore used two statistical methods to investigate AAO, first as a continuous variable and second by separating AAO into statistically identified cut-off point for early- and late-onset groups.

Thirteen candidate gene studies have previously examined Y-BOCS symptom checklist- generated symptom dimensions in OCD (Table 1.9) but they were mostly under-powered and reported inconsistent results. The most robust results were reported in the catechol-O- methyltransferase (COMT) rs4680 polymorphism, which has been found to be associated with hoarding (Hemmings et al., 2006), obsessional/checking dimension (Lochner et al., 2008), and somatic and sensory phenomena (Katerberg et al., 2010); nonetheless, Liu et al. (2011) failed to replicate these findings. Two studies detected significant associations between 140

141

repeating/counting and symmetry/repeating/counting/ordering dimensions and the SLC6A4 5HTTLPR L/L genotype (P=0.0013) (Cavallini et al., 2002) and S/S genotype (Hasler et al., 2006). Hemmings et al. (2006) demonstrated that the dopamine D4 receptor (DRD4) rs1800955 and the estrogen receptor alpha (ESRα) rs9340799 markers were associated with hoarding but two independent studies failed to replicate these results (Lochner et al., 2008; Taj et al., 2013). Alonso et al. (2012) reported a different marker within DRD4, rs34535804, which predicted lower risk of having contamination/cleaning factor symptoms (P=0.0001). We therefore aimed to investigate the effect of OCD candidate genes in Y-BOCS symptom checklist-generated symptom dimensions because: a) limited number of published genetic studies (13 in total) have explored Y-BOCS symptom dimensions; b) previous studies investigated commonly examined markers within the exons of the candidate genes or single nucleotide polymorphisms with no known functional role; c) published studies all had relatively small sample size; d) results were inconsistent and have not been replicable.

Higher severity of psychiatric symptoms has been found to have higher genetic loading (i.e., OCD – Schooler et al., 2008; attention deficit hyperactivity disorder – Stergiakouli et al., 2015). Thus, five studies in total, have investigated into the genetics of OCD severity. Two significant studies from the same research groups demonstrated that the BDNF rs6265 Val/Val genotype was associated with higher severity in OCD females (P=0.013-0.045; Hemmings et al., 2006) (P=0.031; Hemmings et al., 2008). Márquez et al. (2013) examined the same SNP but was a negative study. One study reported significant associations between OCD symptom severity and three genes, glutamate receptor, ionotropic, N-methyl D-aspartate 2B (GRIN2B), HTR2A, and dopamine D1 receptor (DRD1) but they have not been replicated yet. Given the potential higher genetic loading in more severe cases of OCD and limited number of genetic studies exploring Y-BOCS severity in OCD, we aimed to investigate the effect of OCD candidate genes on OCD severity.

Regarding the genetic study of comorbidity in OCD, although the shared genetic susceptibility to OCD and tics has previously been established, the relationship between OCD and other psychiatric disorders has not been examined except for one study by Wendland et al. (2007) who investigated the total number of anxiety and mood disorders and detected significant results for the serotonin transporter 5HTTLPR and rs25531 with the total number of comorbid

141

142

anxiety disorders (P=0.007-0.037). Only three candidate gene studies to date have examined the genetic effect of familiality in OCD symptoms (Denys et al., 2006; Katerberg et al., 2009; Katerberg et al., 2010). Denys et al. (2006) reported a significant association between positive family history of OCD and the HTR2A rs6311 G allele (P=0.039-0.043) in 156 OCD patients. Katerberg et al. (2009; 2010) examined the BDNF rs6265 and COMT rs4680 markers respectively and did not detect any differences in allele and genotype frequencies between presence and absence of family history of obsessive-compulsive symptoms. However, no study to date has explored family history of other psychiatric disorders in the genetics of OCD. The rationale behind the investigation of familial OCRDs in OCD came from recent heritability literature suggesting a common genetic pattern between the OCRDs. Higher rates of psychiatric comorbidity and family history of OCRDs may arise from the shared genetic vulnerability with OCD. Therefore, it is noteworthy to target OCD individuals with different comorbid conditions or significant family history of OCRDs for genetic study given the higher probability of genetic loading in individuals with higher number of comorbid psychiatric conditions or with a family history of OCRDs.

One novelty of this study is the investigation of single nucleotide polymorphisms (SNPs) in the remote regulatory regions of the OCD genes of interest. Less than 2% of the entire human genome accounts for protein-coding regions (Lander et al., 2001) and the rest of the non-coding sequences were once thought to be “junk” DNA with unknown function. With the recent breakthrough of identifying these “junk” as remote regulatory elements from the ENCyclopedia Of DNA Elements (ENCODE) project (ENCODE Project Consortium, 2012), we now have a better understanding of the importance of non-coding regions in gene regulation.

OCD candidate genes have been selected (Table 4.1 and Table 4.2) from previously significant association studies, meta-analyses, and the first genome-wide association study (GWAS) of OCD (Stewart et al., 2013). SNPs were chosen based on the functionality using different publicly available databases that utilize information from the ENCODE project (ENCODE Project Consortium, 2012).

Therefore, the present study aimed to examine the genetics of OCD subphenotypes in a large OCD sample, in order to quantify AAO, severity, symptom dimension, psychiatric disorder

142

143

burden, and family history burden of the highest risk of developing OCD. This is the first comprehensive genetic study that investigates multiple different subphenotypes of OCD across multiple genetic variations with functional basis within OCD candidate genes.

143

144

Table 4.1. Chosen Genes.

Candidate Meta- Y- SRI Gene Chromosome Gene GWAS Gender AAO Familiality Comorbidity Reference Analysis BOCS Response Studies HTR1B 6q14.1 * - + + --- Corregiari et al., 2012; Denys et al., 2007; Mas et al., 2014; Miguita et al., 2011; Taylor, 2013 GRIK2 6q16.3 ++ # Delorme et al., 2004; Sampaio et al., 2011; Stewart et al., 2013 SLC1A1 9p24.2 * +/- + + Stewart et al., 2013; Real et al., 2010; Taylor, 2013 BDNF 11p14 * -- ++ + +++ + + Hall et al., 2003; Katerberg et al., 2009; Márquez et al., 2013; Real et al., 2009; Rocha et al., 2010; Taylor, 2013; Tukel et al., 2014; Zai et al., 2015 GRIN2B 12p13.1 * - + + # Alonso et al., 2012; Qin et al., 2015; Taylor, 2013

144

145

FAIM2 12q13.12 # Stewart et al., 2013 HTR2A 13q14.2 * + + +++-- Corregiari et al., 2012; Denys et al., 2007; Dickel et al., 2007; Miguita et al., 2011; Taylor, 2013; Tot et al., 2003; Zhang et al., 2004 SLITRK5 13q31.2 SLC6A4 17q11.2 *+ + Cengiz et al., 2015; Taylor, 2013 DLGAP1 18p11.31 - # - - + - Li et al., 2015; Stewart et al., 2013 FUT2 19q13.33 # Stewart et al., 2013 BTBD3 20p12.2 # Stewart et al., 2013 COMT 22q11.21 * + + + Alsobrook et al., 2002; Taylor, 2013; Vulink et al., 2012 MAOA Xp11.3 * - Taylor, 2013 + indicates significant study - indicates negative study * indicates previous meta-analysis # indicates trend association with OCD

145

146

Table 4.2. Chosen SNPs.

Gene Chromosome SNP Location FS_score BrainCloud UCSC* HaploReg^ EurMAF RegulomeDB_NIEHS Comments and References HTR1B 6 rs1778258 5' dt o 0.13 1_0.354537 GRIK2 6 rs1556995 o 0.21 OCD; Sampaio et al., 2011; Stewart et al., 2013 SLC1A1 9 rs3933331 5' 0.16 OCD; Wendland et al., 2009 SLC1A1 9 rs7031998 5', histone dt o 0.25 1_0.297757 mark SLC1A1 9 rs7022369 Intron o 0.49 Linked to increase expression of SLC1A1 in schizophrenia; Horiuchi et al., 2012 BDNF 11 rs3763965 5' 0.44 1_0.272426 Tagged by rs10835187 BDNF 11 rs7124442 3'UTR (also o 0.29 5_0.092802 Associated with high BDNF serum in level in anorexia and bulimia; BDNF_AS1) Mercader et al., 2007 BDNF 11 rs61888800 CpG dth rs56133711 0.23 BDNF 11 rs11030119 Exon, CpG dth o 0.28 Tagged by rs7944119 BDNF 11 rs7944119 Exon, CpG dth o 0.28 Tagged by rs11030119; Mercader et al., 2007 BDNF 11 rs2883187 Exon 1, dt o 0.48 histone mark, CpG GRIN2B 12 rs1805482 Exon o 0.33 3_0.357405 GRIN2B 12 rs7301328 Exon o 0.37 3_0.467752 FAIM2 12 rs1044677 o 0.35 5_0.298166 FAIM2 12 rs7132908 3'-UTR dt o 0.36 5_0.370372 Associated with obesity and height; Li et al., 2010 FAIM2 12 rs706795 5'-UTR, d o 0.41 13_0.440516 CpG HTR2A 13 rs7997012 Intron 0.44 Associated with citalopram response and thalamic 5-HTT binding; Laje et al., 2010; McMahon et al., 2006 HTR2A 13 rs9534510 Intron 2 dt o 0.46 SLITRK5 13 rs10450811 125kb 3' of 0.18 See QTL: LINC00438 cis_eQTL_hapmap3_LCL_meta: Match Score = 0.738636452456 SLITRK5 13 rs9557425 5', CpG 0.37 1_0.359264 SLC6A4 17 rs3813034 3'UTR rs1042173 0.45 Polyadenylation polymorphism associated with spontaneous fear recovery after extinction; Hartley et al., 2012 DLGAP1 18 rs2240899 Intron o 0.42 1_0.280006 Tagged by rs2240898 146

147

DLGAP1 18 rs8096794 Intron, dth rs8096794 0.47 histone mark, enhancer FUT2 19 rs681343 CpG rs601338 0.43 BTBD3 20 rs1996132 20kb 3' of 0.26 See QTL: C20orf61 cis_eQTL_Zeller_Monocyte: Match Score = 3.03374677682 COMT 22 rs9617850 o 0.21 See QTL: cis_eQTL_Zeller_Monocyte: Match Score = 4.64984078336 COMT 22 rs737865 Intron 1, dth o 0.28 May affect mRNA, protein, and histone enzyme activity in postmortem mark, human brain; Chen et al., 2004 enhancer COMT 22 rs933271 Intron 1, dth o 0.31 histone mark, enhancer COMT 22 rs4818 Exon, 0.365 ~0.40~ histone mark, enhancer MAOA X rs3788862 Intron o 0.36 Associated with thrombocyte-MAO activity; Jansson et al., 2005 MAOA X rs1465107 Intron o 0.36 cis-acting interactions with VNTR and other SNPs associated with MDD and with higher MAOA mRNA; Zhang et al., 2010 MAOA X rs979605 Intron o 0.39 Associated with thrombocyte-MAO activity; Jansson et al., 2005 MAOA X rs1137070 Associated with thrombocyte-MAO activity; Jansson et al., 2005; cis- acting interactions with VNTR and other SNPs associated with MDD rs6323; 0.27 3_0.139747 and with higher MAOA mRNA; =rs1801291 Zhang et al., 2010; MAOA enzyme activity associated with hyperuricemia and gout; Tu et al., 2010 FS: functional significance UCSC: University of California, Santa Cruz genome browser EurMAF: European minor allele frequency was obtained from HaploReg database NIEHS: National Institute of Environmental Health Sciences * d=DNase, t=TF ChIP, h=H3K27ac

147

148

DNase: deoxyribonuclease, which is an enzyme that breaks down DNA TF: transcription factor ChIP: chromatin immunoprecipitation H3K27Ac: histone H3 (acetyl K27), which is a histone protein that can be chemically modified to regulate gene transcription ^ o indicates SNP that tags other SNP(s) Highlighted in light blue indicates that the SNP was not examined but tagged with another SNP on the list

148

149

4.3 Methods 4.3.1 Diagnostic Criteria and Sample for Candidate Gene Analysis

Four hundred and ninety-seven OCD patients and 336 family members (total N=833) were recruited from consecutive referrals to the Anxiety Disorders Clinic at the Centre for Addiction and Mental Health, Toronto and the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre, Toronto. All research participants were assessed using the Structured Clinical Interview for the DSM-IV (SCID) (First et al., 1996). All assessments and interviews were performed by trained research assistants who were supervised by an experienced psychiatrist. All participants met DSM-IV criteria for the diagnosis of OCD as determined by the SCID and available medical records. All subjects provided their written informed consent to participate in this study, and research ethics approval was obtained from the local Research Ethics Board. Exclusion criteria included any metabolic or chronic neurological disease (other than tic disorder), active substance abuse/dependence, schizophrenia, or schizoaffective disorder.

4.3.1.1 Using AAO as Continuous and Categorical Variable

Age at onset (AAO) was defined when the subject first met DSM-IV diagnostic criteria of OCD. We performed an admixture analysis to identify the best-fitting model with the observed distribution of the continuous variable, AAO, according to the protocol described by Delorme et al. (2005) using the DENORMIX module (Kolenikov, 2001) in the STATA program (version 11.0, Texas, USA). The DENORMIX module performs a decomposition of the AAO distribution into a mixture of normal distributions and estimates the number of distributions within the mixture. We used the highest P value to which the model best approximated the empirical distribution function of AAO from the χ2 goodness-of-fit test to select the best-fitting model. The point of intersection between two normal distributions was calculated using the means and standard deviations of the identified components using the uniroot function in the R program (version 3.2.1).

149

150

This analysis of AAO yielded three normal distributions in the best-fitting model and AAO was divided into three groups, early (≤8 years), intermediate (between 9 to 17 years), and late (≥18 years) onset. We investigated the relationship between AAO and genotypes of each tested SNP using SPSS (version 20.0, Armonk, NY, USA) with two different analytical approaches. We identified three AAO groups and performed chi-squared (χ2) test to determine whether there is a different genotype frequency in each AAO group. We also used AAO as a continuous variable and conducted linear regression with the genotyping.

4.3.1.2 Using Y-BOCS Severity Score and Symptom Dimensions for Genetic Analysis

The Yale-Brown Obsessive-Compulsive Scale (Y-BOCS), a gold-standard, validated and clinician-administered instrumental measure was completed for each participant to determine the severity of OCD symptoms (Goodman et al., 1989). Lifetime OCD symptom severity was retrospectively estimated by re-administering the Y-BOCS scale while each OCD participant was requested to focus on the time when the most severe OCD symptoms were endorsed. The presence or absence of OCD symptoms was documented using the Y-BOCS symptom checklist (Goodman et al., 1989; Appendix I).

An exploratory factor analysis for data reduction of the Y-BOCS symptom checklist was conducted on seven obsession categories and eight compulsion categories with a total of 15 a priori categories using SPSS (version 20.0, Armonk, NY, USA). Lifetime Y-BOCS symptoms (past or present/current) were chosen for this analysis as described by Leckman et al. (1997). The miscellaneous obsessive and compulsive symptom categories were excluded from the analysis because previous studies have shown that these symptoms are not clearly associated with previously known OCD symptom dimensions (Baer et al., 1994).

We utilized two different coding approaches. The first approach coded the presence of symptoms as “1” and the absence of symptoms as “0”. The second approach coded principal or target symptoms as “2” by double-weighing these severe symptoms and coded the presence of minor symptoms that were not described as target or principal as “1” as well as the absence of

150

151

symptoms as “0”. Sum across the symptoms for each category was calculated and used in the analysis. Initial factors were extracted using the principal component analysis (PCA) method and the PROMAX rotations for relatedness of symptoms (De Geus et al., 2004) were performed using SPSS (version 20.0, Armonk, NY, USA). Factor scores for each generated factor (symptom dimensions) for each subject were automatically calculated using SPSS (version 20.0, Armonk, NY, USA) and used as continuous variables for genetic analysis.

The factor analysis generated five distinct factors when treating all lifetime symptoms equally whereas a 6-factor model was detected after double-weighing the target symptoms. Approximately 65% of the variance was explained by the 5-factor model and 70% of the variance for the 6-factor model. The five symptom dimensions include: factor 1 includes symmetry obsessions, ordering and arranging, counting, repeating, and checking compulsions; factor 2 comprises of sexual and religious obsessions, and aggressive obsessions and compulsions; factor 3 consists of hoarding and saving obsessions, hoarding and collecting compulsions; factor 4 contains contamination obsessions and cleaning compulsions; and factor 5 includes somatic obsessions and compulsions. The six symptom dimensions consist of: factor 1 consists of symmetry obsession, ordering and arranging, repeating, and counting compulsions; factor 2 contains hoarding and saving obsessions, hoarding and collecting compulsions; factor 3 comprises of contamination obsessions and cleaning compulsions; factor 4 includes aggressive obsessions and compulsions, and checking compulsions; factor 5 includes somatic obsessions and compulsions; factor 6 consists of religious and sexual obsessions.

The factor automatically generated factor scores were then used as a continuous variable for genetic analysis using linear regression in SPSS (version 20.0, Armonk, NY, USA).

4.3.1.3 Presence or Absence of Comorbid Psychiatric Disorder(s)

The frequency of current comorbid psychiatric conditions and lifetime/current OCRDs was determined by the SCID-IV interview. The presence or absence of psychiatric comorbidity was explored as a categorical dichotomous variable and compared using a Pearson χ2 test or

151

152

Fisher’s exact test when required regardless of the diagnosis status of the proband in SPSS (version 20.0, Armonk, NY, USA).

4.3.1.4 The Presence or Absence of Family History of OCRDs

To determine familiality, family psychiatric history from OCD probands and available first-degree relatives was collected using the modified OCD Family History Interview (FHI) questionnaire (Appendix II). FHI is a semi-structured interview that was developed by Drs. Richter and Kennedy and it has been modified based on the Family History Diagnostic Criteria (Andreasen et al., 1986) for OCD, OCD with hoarding, hoarding disorder (HD), trichotillomania (TTM), and body dysmorphic disorder (BDD). Reliability measures were also collected and data obtained were considered reliable if the informant was the subject who could recall his/her own symptoms or if the informant and subject: a) were reasonably close, b) had regular and frequent direct contact, and c) could provide details of subject’s symptoms. If the informant failed the reliability cut-off (no contact or indirect contact only with the subject, or not close to the subject), then two independent sources (including proband or any participating family members) would have to provide the same information or the participants indicated that the information provided had been confirmed. The criteria for a significant (positive) family history of OCRDs should have either of the following:

1) A clinical diagnosis of OCRDs; or

2) Reliability measure with an informant providing accurate information or close relationship between the informant and subject who provided reasonable details on the subject’s mental health that met criteria for an OCRD; or

3) If the informant reported having no or very little information on or only indirect contact with the subject (not meeting the reliability cut-off), then at least two members from the same family (including proband and first-degree relatives) reported on the same symptoms that met criteria for an OCRD separately/independently.

152

153

Familiality for each patient is coded either as a “0”, which indicates absence of family history, or “1”, which refers to the presence of family history. The presence or absence of family history was explored as a categorical dichotomous variable and compared using a Pearson χ2 test or Fisher’s exact test when required regardless of the diagnosis status of the proband in SPSS (version 20.0, Armonk, NY, USA).

4.3.2 Choosing of SNPs of Candidate Genes and Genotyping

Venous blood was obtained from the participants in two 10cc EDTA tubes, and genomic DNA was extracted from blood lymphocytes using a high salt method (Lahiri and Nurnberger, 1991).

Thirty-two SNPs, which mostly lie within the remote regulatory regions across 14 candidate genes, were investigated. A 32-SNP QuantStudio Chip (Life Technologies ®, Burlington, ON) was designed. The chip consisted of regulatory SNPs across 14 different candidate genes including genes from previous significant genetic association studies and recent top hits from the first and recently published GWAS (Stewart et al., 2013). SNPs were chosen based on potential functional status (Table 4.2) using different publicly available search engines including functional significance (FS) score, European minor allele frequencies (EurMAF), BrainCloud (Colantuoni et al., 2011), University of California, Santa Cruz genome browser (UCSC) (Rosenbloom et al., 2013), HaploReg (Ward et al., 2012), RegulomeDB (Boyle et al., 2012), and National Institute of Environmental Health Sciences (NIEHS) (Xu et al., 2009).

The 32 SNPs in the following 14 candidate genes are shown in Table 4.1 and Table 4.2: serotonin 1B receptor (HTR1B), glutamate receptor ionotropic kainite 2 (GRIK2), neuronal glutamate transporter (SLC1A1), brain-derived neurotrophic factor (BDNF), glutamate receptor ionotropic N-methyl D-aspartate 2B (GRIN2B), fas apoptotic inhibitory molecule 2 (FAIM2), serotonin 2A receptor (HTR2A), SLIT and NTRK-like family member 5 (SLITRK5), serotonin transporter (SLC6A4), discs large (drosophila) homolog-associated protein 1 (DLGAP1), fucosyl- transferase 2 (FUT2), BTB (POZ) domain containing 3 (BTBD3), catechol-O-methyltransferase (COMT), and monoamine oxidase A (MAOA).

153

154

Genotyping was performed using the QuantStudio technology (Life Technologies ®, Burlington, ON). All genotypes were determined with the Life Technology ® allelic discrimination software and confirmed by two experienced researchers. Approximately 10% of the genotypes were confirmed by repeating the experiment and using positive and negative controls.

4.3.3 Quality Control and Statistical Analyses

Quality control was performed using the whole genome association analysis toolset, PLINK software (version 1.07) (Purcell et al., 2007). Samples with <98% completion, SNP assays with call rates <90%, and SNPs with significant Hardy-Weinberg equilibrium test were excluded. Only two SNPs, GRIN2B rs1805482 and COMT rs737865, were excluded from the final analysis. We tested for the association between genotypes of 30 SNPs and different phenotypes using the linear regression model in the PLINK software (version 1.07) (Purcell et al., 2007). For the quantitative analyses, linear regression was performed on the continuous variables including AAO, Y-BOCS severity scores, and Y-BOCS symptom dimension factor scores (please refer to Chapter 2 for details) using the PLINK program (version 1.07) (Purcell et al., 2007).

Permutation to one million was conducted for all linear regression analyses using the PLINK software (version 1.07) (Purcell et al., 2007). Pearson χ2 test was conducted for AAO groups, the presence or absence of family history, with genotype and allele frequencies using SPSS (version 20.0, Armonk, NY, USA). Hardy Weinberg Equilibrium test for each SNP and linkage disequilibrium between the markers within the same gene were determined using the Haploview program (version 3.32) (Barrett et al., 2005). Haplotype analysis was conducted using the sliding window method in the Unphased software (version 3.1.7) (Dudbridge, 2008).

Results were assumed to be significant initially if P<0.05 in all cases and P<0.002 (26 independent tests due to 3 for the 5 markers across BDNF and 2 for 4 markers across MAOA) in all cases after corrected for multiple comparisons using the Nyholt method (Nyholt, 2004; Li et

154

155

al., 2005).

Our sample had enough power (80.5%) to detect 7.4% of the variance of AAO and Y- BOCS severity that are accounted for by the SNP (α=0.002, additive model) using the QUANTO program (version 1.2.4) (Gauderman and Morrison, 2006).

4.4 Results

Demographic data of the OCD sample is presented in Table 4.3 and Table 4.4. Given that 3% of our sample consists of Asians and African Americans, we excluded them for the final analysis. Thirty SNPs remained after quality control for genetic analyses (Table 4.5).

Table 4.3. Subject Demographics.

Descriptors Mean (S.D.) N (mixed data set) 497 (probands) Age 35.2 (12.5) Gender 57.3% female Ethnicity 97.0% Caucasian 2.6% Asian 0.4% African American Age at Onset (N=328) 14.5 (8.9) Yale-Brown Obsessive Compulsive Scale (Y-BOCS) severity score 27.1 (6.6) (N=362) This table presents the demographics of all OCD subjects.

155

156

Table 4.4. Subject Demographics (100% Caucasian).

Descriptors Mean (S.D.) N (final data set) 421 (probands) Age 35.3 (12.4) Gender 54.6% female Ethnicity 100% Caucasian Age at Onset (N=163) 13.5 (7.9) Yale-Brown Obsessive Compulsive Scale (Y-BOCS) severity score (N=189) 26.0 (6.7) This table presents the demographics of Caucasian OCD subjects only.

Table 4.5. Investigated SNPs.

Chromosome Gene SNP MAF 6 HTR1B rs1778258 0.092 6 GRIK2 rs1556995 0.162 9 SLC1A1 rs3933331 0.166 9 SLC1A1 rs7031998 0.259 9 SLC1A1 rs7022369 0.473 11 BDNF rs3763965 0.433 11 BDNF rs7124442 0.252 11 BDNF rs61888800 0.187 11 BDNF rs11030119 0.234 11 BDNF rs2883187 0.482 12 GRIN2B rs1805482 * 12 GRIN2B rs7301328 0.382 12 FAIM2 rs1044677 0.296 12 FAIM2 rs7132908 0.377 12 FAIM2 rs706795 0.376 13 HTR2A rs7997012 0.390 13 HTR2A rs9534510 0.398 13 SLITRK5 rs10450811 0.178 13 SLITRK5 rs9557425 0.379 17 SLC6A4 rs3813034 0.439 18 DLGAP1 rs2240899 0.383 18 DLGAP1 rs8096794 0.499 19 FUT2 rs681343 0.496 20 BTBD3 rs1996132 0.242 22 COMT rs9617850 0.173 22 COMT rs737865 * 22 COMT rs933271 0.269 22 COMT rs4818 0.412 X MAOA rs3788862 0.303 X MAOA rs1465107 0.300 X MAOA rs979605 0.302 X MAOA rs1137070 0.303 * Highlighted in grey indicates failed SNP in the QuantStudio 32-SNP Chip MAF = minor allele frequency

156

157

Genotypic frequencies did not deviate from Hardy Weinberg Equilibrium with a P value greater than 0.090 (Nyholt correction P>0.002). Significant linkage disequilibrium and haplotype blocks were observed for the following SNPs: BDNF rs7124442 and rs61888800 (D’=0.96, LOD=58.34, r2=0.685), BDNF rs11030119 and rs2883187 (D’=1.0, LOD=33.36, r2=0.349), FAIM2 rs1044677 and rs7132908 (D’=0.98, LOD=28.96, r2=0.284), COMT rs9617850 and rs933271 (D’=0.986, LOD=54.81, r2=0.594), MAOA rs3788862 and rs1465107 (D’=0.993, LOD=140.26, r2=0.978), and MAOA rs979605 and rs1137070 (D’=0.985, LOD=136.95, r2=0.971). The skewness of the AAO distribution was 1.210, which significantly deviated from normal distribution (Figure 4.1) and thus it was transformed into Log10 of AAO with a resulting skewness of -0.333 (Figure 4.2). Y-BOCS severity score did not deviate greatly from the normal distribution with a skewness of -0.345 (Figure 4.3).

Trends in gender differences were observed (Table 4.6) for the DLGAP1 rs2240899 (P=0.009), SLC6A4 rs3813034 (P=0.028), and COMT rs4818 (P=0.042). The strongest trend for genetic differences between the AAO groups (early-, intermediate-, and late-onset) were detected (Table 4.7) for the FUT2 rs681343 SNP (P=0.017).

Linear regression of AAO (Table 4.8 and Figure 4.4), Y-BOCS severity score (Table 4.9 and Figure 4.5), and Y-BOCS symptom dimensions for the 5-factor (Table 4.10 and Figure 4.6) and 6-factor (Table 4.11 and Figure 4.7) models were also negative. However, trends were observed for: earlier AAO with COMT rs4818 A-G haplotype (P=0.021) and later onset with FUT2 rs681343 T allele (P=0.007); however, these results did not survive correction for multiple comparisons. The following haplotypes were nominally implicated in AAO groups: COMT rs9617850-rs933271 A-T with early onset (P=0.043) and rs9617850-rs4818 G-C with early onset (P=0.031). Trend associations were noted for the 5- and 6-factor models of Y-BOCS symptom dimension but these findings were not significant after correcting for multiple testing (Tables 4.10 and 4.11, Figures 4.6 and 4.7).

The following presents the percentage for each comorbid psychiatric condition in the current sample of OCD individuals with genetic data: 5.6% with tics, 6.4% with any mood disorders, 9.1% with any anxiety disorders, and 5.3% with any OCRDs in addition to OCD. Only trend associations were detected in the genetic analyses of comorbid psychiatric conditions

157

158

(Table 4.12). Several markers were found to be associated with OCD and comorbid OCRDs including HTR1B rs1778258 (P=0.011), BDNF rs7124442 (P=0.040) and rs11030119 (P=0.028), and FUT2 rs681343 (P=0.014); however, these findings did not survive correction for multiple comparisons. BDNF appear to confer risk to developing OCD and comorbid mood, anxiety, and OCRDs but further research is needed to expand on these results.

Approximately 22% of the OCD participants had positive family history of obsessive- compulsive and related disorders (15.5% OCD, 1.4% OCD with hoarding disorder, 2.0% hoarding disorder, 1.2% trichotillomania, and 1.6% body dysmorphic disorder). Significant associations (Table 4.13) were detected for positive family history of: OCD and HTR2A rs7997012 AA genotype (χ2=9.016, P=0.011) and rs9534510 TT genotype (χ2=7.353, P=0.025), trichotillomania and SLC1A1 rs7022369 CC genotype (χ2=16.007, P=0.0003), trichotillomania and DLGAP1 rs8096794 AG genotype (χ2=6.025, P=0.045), body dysmorphic disorder and SLC1A1 rs3933331 CC genotype (χ2=9.149, P=0.010), and body dysmorphic disorder and BDNF rs2883187 AA genotype (χ2=6.170, P=0.046). However, these results are preliminary given the relatively small sample size for each positive family history group. When examining the family history of all OCRDs together (Table 4.14), interesting findings were detected in FAIM2 rs1044677 (χ2=4.739, P=0.094) and rs7132908 (χ2=5.049, P=0.080) in addition to HTR2A rs9534510 (χ2=5.572, P=0.062).

These results did not change significantly when incorporating covariates including age and sex into the analytical model (i.e., linear regression for continuous AAO and Y-BOCS severity, logistic regression for categorical AAO groups, family history, and psychiatric comorbidity).

158

159

Figure 4.1. Age At Onset (AAO) Distribution with Normal Curve. This graph illustrates the AAO distribution of the OCD sample.

159

160

Figure 4.2. Log10 Transformed Age At Onset (AAO) Distribution with Normal Curve. This graph illustrates the log10 transformation of AAO distribution of the OCD sample.

160

161

Figure 4.3. Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Distribution with Normal Curve. This graph illustrates the Y-BOCS distribution of the OCD sample.

161

162

Table 4.6. Genetic Results of Gender Differences in OCD.

Gender Gene SNP Allele/Genotype Male Female χ2 (Genotype/Allele) P (Genotype/Allele) HTR1B rs1778258 CC/CT/TT/C/T 158/32/3/348/38 228/51/2/507/55 0.938/0.000 0.625/1.000 GRIK2 rs1556995 CC/CT/TT/C/T 11/60/121/82/302 8/99/172/115/443 2.947/0.040 0.229/0.846 SLC1A1 rs3933331 CC/CG/GG/C/G 10/52/125/72/302 19/81/178/119/437 0.601/0.510 0.740/0.475 rs7031998 GG/GT/TT/G/T 94/84/11/272/106 132/121/20/385/161 0.422/0.160 0.810/0.687 rs7022369 CC/CG/GG/C/G 55/86/52/196/190 89/130/58/308/246 2.381/1.930 0.304/0.164 BDNF rs3763965 AA/AT/TT/A/T 33/99/58/165/215 54/128/94/236/316 1.479/0.020 0.477/0.893 rs7124442 CC/CT/TT/C/T 13/76/103/102/282 21/97/158/139/413 0.976/0.160 0.614/0.690 rs61888800 GG/GT/TT/G/T 123/62/6/308/74 182/81/12/445/105 0.823/0.000 0.663/0.982 rs11030119 AA/AG/GG/A/G 11/75/105/97/285 21/87/168/129/423 3.187/0.400 0.203/0.527 rs2883187 AA/AG/GG/A/G 48/95/46/191/187 76/136/65/288/266 0.243/0.140 0.886/0.711 GRIN2B rs7301328 CC/CG/GG/C/G 26/95/69/147/233 43/138/98/224 0.281/0.140 0.869/0.703 FAIM2 rs1044677 CC/CT/TT/C/T 96/75/20/267/115 127/123/25/377 1.405/0.130 0.495/0.714 rs7132908 AA/AG/GG/A/G 36/89/68/161/225 35/139/107/209/353 3.452/1.780 0.178/0.182 rs706795 CC/CT/TT/C/T 70/97/26/237/149 105/122/51/332/224 2.720/0.210 0.257/0.651 HTR2A rs7997012 AA/AG/GG/A/G 31/86/74/148/234 35/136/106/206/248 1.438/0.170 0.487/0.678 rs9534510 GG/GT/TT/G/T 59/100/32/218/164 84/146/49/314/244 0.066/0.030 0.968/0.861 SLITRK5 rs10450811 AA/AG/GG/A/G 4/55/131/63/317 10/76/189/96/454 0.944/0.070 0.624/0.795 rs9557425 CC/CG/GG/C/G 65/89/37/219/163 112/117/51/341/219 1.759/1.050 0.415/0.305 SLC6A4 rs3813034 AA/AC/CC/A/C 60/73/54/193/181 90/134/52/314/238 7.171/2.300 0.028/0.129 DLGAP1 rs2240899 CC/CG/GG/C/G 51/103/37/205/177 113/122/44/348/210 9.505/6.730 0.009/0.009 rs8096794 AA/AG/GG/A/G 45/98/49/188/196 73/131/75/277/281 0.809/0.020 0.667/0.889 FUT2 rs681343 CC/CT/TT/C/T 58/88/47/204/182 66/150/63/282/276 3.434/0.400 0.179/0.527 BTBD3 rs1996132 CC/CG/GG/C/G 12/69/111/93/291 28/103/148/159/399 2.419/1.910 0.298/0.167 COMT rs9617850 AA/AG/GG/A/G 4/61/127/69/315 15/87/178/117/443 3.167/1.050 0.205/0.305 rs933271 CC/CT/TT/C/T 13/71/109/97/289 27/112/143/166/398 2.070/1.910 0.355/0.167 rs4818 CC/CG/GG/C/G 54/93/44/201/181 105/121/52/331/225 4.712/4.130 0.095/0.042 MAOA rs3788862 A/G 138/248 191/371 0.240 0.623 rs1465107 A/G 139/247 184/372 0.740 0.391 rs979605 A/G 133/247 194/362 <0.001 1.000

162

163

rs1137070 CC/CT/TT/C/T 123/1/69/247/139 127/108/44/362/196 98.223/0.040 <0.001/0.834 Highlighted in yellow indicates significant P value prior to correction for multiple comparisons Highlighted in grey indicates trend association prior to correction for multiple comparisons Highlighted in light blue indicates X-linked transmission

163

164

Table 4.7. Genetic Results of Age At Onset (AAO) Groups in OCD.

AAO Group Gene SNP Genotype ≤8 9-17 ≥18 χ2 P HTR1B rs1778258 CC/CT/TT 39/14/1 51/12/0 40/6/0 4.831 0.305 GRIK2 rs1556995 CC/CT/TT 3/12/38 2/23/37 2/13/31 3.095 0.542 SLC1A1 rs3933331 CC/CG/GG 1/15/38 3/20/39 2/14/29 1.240 0.871 rs7031998 GG/GT/TT 27/22/3 35/25/3 22/22/2 0.830 0.934 rs7022369 CC/CG/GG 17/25/12 21/26/16 12/26/8 2.569 0.632 BDNF rs3763965 AA/AT/TT 11/28/15 11/28/24 6/26/14 2.627 0.622 rs7124442 CC/CT/TT 5/23/26 2/23/37 2/19/25 3.006 0.557 rs61888800 GG/GT/TT 34/17/3 43/19/1 28/16/2 1.756 0.781 rs11030119 AA/AG/GG 5/22/27 1/22/40 3/18/25 4.482 0.345 rs2883187 AA/AG/GG 12/28/14 20/29/13 10/28/8 3.418 0.490 GRIN2B rs7301328 CC/CG/GG 5/30/19 7/29/26 9/22/15 3.453 0.485 FAIM2 rs1044677 CC/CT/TT 19/28/7 29/25/8 22/22/2 4.388 0.356 rs7132908 AA/AG/GG 6/25/23 11/25/27 5/25/16 2.957 0.565 rs706795 CC/CT/TT 16/28/9 22/36/5 16/17/13 9.121 0.058 HTR2A rs7997012 AA/AG/GG 10/20/24 9/30/23 7/22/17 1.818 0.769 rs9534510 GG/GT/TT 18/30/6 18/32/13 18/24/4 4.226 0.376 SLITRK5 rs10450811 AA/AG/GG 2/19/33 1/15/46 1/18/27 3.810 0.432 rs9557425 CC/CG/GG 21/25/8 26/20/16 16/24/6 5.990 0.200 SLC6A4 rs3813034 AA/AC/CC 17/20/15 21/24/18 13/25/8 3.876 0.423 DLGAP1 rs2240899 CC/CG/GG 15/30/9 19/29/15 18/21/7 3.059 0.548 rs8096794 AA/AG/GG 11/26/17 16/33/14 15/21/10 2.997 0.558 FUT2 rs681343 CC/CT/TT 4/34/16 17/25/21 14/22/10 12.061 0.017 BTBD3 rs1996132 CC/CG/GG 3/19/32 8/25/30 3/18/25 2.970 0.563 COMT rs9617850 AA/AG/GG 2/23/29 2/20/41 1/9/36 6.626 0.157 rs933271 CC/CT/TT 6/22/26 8/19/36 4/15/27 2.018 0.732 rs4818 CC/CG/GG 17/28/9 18/33/12 20/19/7 2.904 0.574 MAOA rs3788862 AA/AG/GG 12/11/31 10/7/45 11/13/22 7.763 0.101 rs1465107 AA/AG/GG 12/11/31 10/7/46 10/13/23 7.432 0.115 rs979605 AA/AG/GG 13/11/30 11/7/45 9/15/22 9.414 0.052 rs1137070 CC/CT/TT 30/11/13 44/8/11 22/14/9 7.130 0.129

Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

164

165

Table 4.8. Genetic Results of Linear Regression of Age At Onset (AAO) in OCD.

Chromosome Gene SNP Allele 1 Model BETA STAT P 6 HTR1B rs1778258 T ADD -0.054 -1.439 0.151 6 GRIK2 rs1556995 C ADD 0.029 0.997 0.320 9 SLC1A1 rs3933331 C ADD -0.015 -0.552 0.582 9 SLC1A1 rs7031998 T ADD -0.004 -0.165 0.869 9 SLC1A1 rs7022369 G ADD 0.020 0.909 0.364 11 BDNF rs3763965 A ADD -0.023 -1.025 0.306 11 BDNF rs7124442 C ADD -0.038 -1.518 0.130 11 BDNF rs61888800 T ADD -0.035 -1.275 0.203 11 BDNF rs11030119 A ADD -0.042 -1.652 0.100 11 BDNF rs2883187 G ADD -0.020 -0.906 0.366 12 GRIN2B rs7301328 C ADD 0.030 1.291 0.198 12 FAIM2 rs1044677 T ADD -0.041 -1.664 0.097 12 FAIM2 rs7132908 A ADD 0.022 0.974 0.331 12 FAIM2 rs706795 T ADD -0.023 -0.993 0.321 13 HTR2A rs7997012 A ADD 0.034 1.498 0.135 13 HTR2A rs9534510 T ADD -0.027 -1.189 0.235 13 SLITRK5 rs10450811 A ADD 0.024 0.830 0.407 13 SLITRK5 rs9557425 G ADD 0.004 0.164 0.870 17 SLC6A4 rs3813034 C ADD -0.030 -1.375 0.170 18 DLGAP1 rs2240899 G ADD -0.022 -0.967 0.334 18 DLGAP1 rs8096794 A ADD 0.037 1.672 0.096 19 FUT2 rs681343 T ADD 0.061 2.727 0.007 20 BTBD3 rs1996132 C ADD -0.003 -0.102 0.919 22 COMT rs9617850 A ADD -0.051 -1.775 0.077 22 COMT rs933271 C ADD -0.010 -0.422 0.674 22 COMT rs4818 G ADD -0.043 -1.942 0.053 X MAOA rs3788862 A ADD 0.022 0.813 0.417 SEX -0.005 -0.154 0.878 X MAOA rs1465107 A ADD 0.017 0.623 0.534 SEX -0.009 -0.276 0.783 X MAOA rs979605 A ADD 0.022 0.826 0.409 SEX -0.006 -0.186 0.853 X MAOA rs1137070 T ADD 0.021 0.794 0.428 SEX -0.004 -0.107 0.915 ADD: additive model SEX: X-linked model Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

165

166

P=0.002 Linear Regression of AAO in OCD N=313 3

FUT2 2.5 P=0.009

2 COMT 1.5 P=0.038 P=0.05

- Log10 P value - Log10 1

0.5

0

rs4818 rs706795 rs681343 rs933271 rs979605 rs1778258rs1556995rs3933331rs7031998rs7022369rs3763965rs7124442 rs2883187rs7301328rs1044677rs7132908rs7997012rs9534510rs9557425rs3813034rs2240899rs8096794rs1996132rs9617850 rs3788862rs1465107rs1137070Within rs61888800rs11030119 rs10450811 Exon

Nyholt & FUT2 BDNF COMT GRIK2 MAOA FAIM2 Bonferroni: HTR1B HTR2A BTBD3 GRIN2B SLC6A4 SLC1A1 DLGAP1

SLITRK5 P<0.002

Figure 4.4. Linear Regression of Age At Onset (AAO) in OCD. This graph illustrates the linear regression results of AAO in OCD.

166

167

Table 4.9. Genetic Results of Linear Regression of Y-BOCS Total Severity Scores in OCD.

Chromosome Gene SNP Allele 1 Model BETA STAT P 6 HTR1B rs1778258 T ADD 0.008 0.460 0.646 6 GRIK2 rs1556995 C ADD 0.004 0.299 0.765 9 SLC1A1 rs3933331 C ADD -0.002 -0.121 0.904 9 SLC1A1 rs7031998 T ADD 0.015 1.256 0.210 9 SLC1A1 rs7022369 G ADD 0.000 -0.024 0.981 11 BDNF rs3763965 A ADD 0.018 1.667 0.097 11 BDNF rs7124442 C ADD 0.004 0.336 0.737 11 BDNF rs61888800 T ADD 0.017 1.285 0.200 11 BDNF rs11030119 A ADD -0.001 -0.042 0.966 11 BDNF rs2883187 G ADD -0.003 -0.269 0.789 12 GRIN2B rs7301328 C ADD -0.012 -1.050 0.294 12 FAIM2 rs1044677 T ADD 0.008 0.671 0.503 12 FAIM2 rs7132908 A ADD 0.004 0.355 0.723 12 FAIM2 rs706795 T ADD -0.004 -0.331 0.741 13 HTR2A rs7997012 A ADD 0.012 1.147 0.252 13 HTR2A rs9534510 T ADD 0.008 0.746 0.457 13 SLITRK5 rs10450811 A ADD 0.006 0.420 0.675 13 SLITRK5 rs9557425 G ADD 0.000 -0.011 0.992 17 SLC6A4 rs3813034 C ADD -0.009 -0.884 0.377 18 DLGAP1 rs2240899 G ADD -0.012 -1.167 0.244 18 DLGAP1 rs8096794 A ADD 0.012 1.178 0.240 19 FUT2 rs681343 T ADD -0.014 -1.283 0.200 20 BTBD3 rs1996132 C ADD 0.016 1.382 0.168 22 COMT rs9617850 A ADD 0.010 0.787 0.432 22 COMT rs933271 C ADD -0.003 -0.270 0.788 22 COMT rs4818 G ADD 0.011 0.989 0.324 X MAOA rs3788862 A ADD 0.023 1.852 0.065 SEX 0.001 0.062 0.950 X MAOA rs1465107 A ADD 0.022 1.719 0.087 SEX 0.002 0.159 0.874 X MAOA rs979605 A ADD 0.020 1.579 0.115 SEX 0.002 0.161 0.872 X MAOA rs1137070 T ADD 0.022 1.777 0.077 SEX 0.000 0.014 0.989 ADD: additive model SEX: X-linked model Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

167

168

Figure 4.5. Linear Regression of Y-BOCS Total Severity Scores in OCD. This graph illustrates the linear regression results of Y-BOCS severity scores in OCD.

168

169

Table 4.10. Genetics Results of Linear Regression of 5-Factor Y-BOCS Symptom Dimensions in OCD.

Y-BOCS 5-Factor Model Factor 1 Factor 2 Factor 3 Factor 4 Factor 5 Symmetry/Ordering/Repeating/Checking Aggression/Sexual/Religious Hoarding Contamination/Cleaning Somatic Chromosome Gene SNP A1 BETA STAT P BETA STAT P BETA STAT P BETA STAT P BETA STAT P 6 HTR1B rs1778258 T 0.053 0.394 0.694 -0.006 -0.042 0.967 -0.129 -0.955 0.340 -0.223 -1.667 0.096 0.062 0.457 0.648 6 GRIK2 rs1556995 C 0.013 0.128 0.898 0.085 0.848 0.397 0.112 1.129 0.260 0.089 0.908 0.365 0.045 0.453 0.651 9 SLC1A1 rs3933331 C 0.089 0.944 0.346 -0.085 -0.892 0.373 -0.025 -0.260 0.795 -0.008 -0.084 0.933 0.050 0.529 0.597 9 SLC1A1 rs7031998 T -0.091 -1.014 0.311 0.021 0.233 0.816 -0.012 -0.130 0.897 -0.125 -1.399 0.163 -0.104 -1.153 0.250 9 SLC1A1 rs7022369 G -0.083 -1.067 0.287 0.041 0.518 0.605 0.017 0.221 0.826 -0.011 -0.142 0.888 0.005 0.058 0.954 11 BDNF rs3763965 A 0.193 2.441 0.015 -0.010 -0.124 0.901 -0.019 -0.233 0.816 0.200 2.547 0.011 0.176 2.202 0.028 11 BDNF rs7124442 C 0.001 0.008 0.994 -0.021 -0.248 0.805 -0.099 -1.143 0.254 0.049 0.569 0.570 0.083 0.959 0.338 11 BDNF rs61888800 T -0.028 -0.298 0.766 0.013 0.138 0.890 -0.082 -0.866 0.387 0.033 0.354 0.723 0.125 1.316 0.189 11 BDNF rs11030119 A -0.059 -0.688 0.492 0.011 0.121 0.904 -0.122 -1.405 0.161 0.044 0.504 0.615 0.052 0.588 0.557 11 BDNF rs2883187 G 0.061 0.782 0.435 -0.028 -0.357 0.721 -0.037 -0.467 0.641 0.149 1.919 0.056 0.053 0.665 0.507 12 GRIN2B rs7301328 C -0.169 -2.055 0.041 -0.070 -0.839 0.402 0.115 1.388 0.166 0.042 0.516 0.607 0.216 2.609 0.009 12 FAIM2 rs1044677 T -0.034 -0.397 0.691 -0.122 -1.457 0.146 0.091 1.070 0.285 -0.007 -0.078 0.938 0.135 1.582 0.115 12 FAIM2 rs7132908 A -0.019 -0.246 0.806 0.015 0.185 0.854 -0.042 -0.534 0.594 -0.067 -0.849 0.397 -0.070 -0.880 0.380 12 FAIM2 rs706795 T 0.054 0.680 0.497 0.087 1.100 0.272 0.010 0.126 0.900 0.084 1.063 0.289 0.157 1.974 0.049 13 HTR2A rs7997012 A 0.007 0.089 0.929 0.121 1.509 0.132 -0.031 -0.383 0.702 0.057 0.717 0.474 0.004 0.051 0.960 13 HTR2A rs9534510 T 0.130 1.629 0.104 0.114 1.417 0.158 -0.075 -0.933 0.352 -0.049 -0.620 0.536 -0.149 -1.854 0.065 13 SLITRK5 rs10450811 A 0.060 0.606 0.545 -0.032 -0.322 0.747 -0.126 -1.279 0.202 -0.060 -0.610 0.542 0.005 0.054 0.957 13 SLITRK5 rs9557425 G -0.017 -0.214 0.831 -0.095 -1.209 0.228 -0.134 -1.703 0.089 -0.109 -1.404 0.161 -0.075 -0.946 0.345 17 SLC6A4 rs3813034 C 0.022 0.288 0.774 -0.125 -1.664 0.097 0.012 0.153 0.878 -0.163 -2.180 0.030 -0.023 -0.304 0.761 18 DLGAP1 rs2240899 G 0.141 1.795 0.074 -0.122 -1.544 0.124 0.029 0.363 0.717 -0.063 -0.806 0.421 0.084 1.047 0.296 18 DLGAP1 rs8096794 A 0.145 1.892 0.059 0.176 2.286 0.023 0.120 1.553 0.121 0.098 1.280 0.201 0.158 2.047 0.041 19 FUT2 rs681343 C -0.104 -1.331 0.184 -0.076 -0.963 0.337 -0.107 -1.366 0.173 0.000 0.000 1.000 -0.128 -1.621 0.106 20 BTBD3 rs1996132 C 0.084 0.970 0.333 0.127 1.461 0.145 0.049 0.555 0.579 0.051 0.587 0.557 0.008 0.096 0.924 22 COMT rs9617850 A 0.055 0.563 0.574 0.045 0.456 0.649 0.078 0.792 0.429 0.167 1.726 0.085 -0.020 -0.205 0.838 22 COMT rs933271 C -0.046 -0.554 0.580 0.042 0.501 0.617 0.043 0.511 0.610 0.154 1.871 0.062 0.005 0.065 0.948 22 COMT rs4818 G 0.062 0.784 0.434 0.103 1.289 0.198 -0.005 -0.061 0.951 -0.047 -0.604 0.547 -0.186 -2.352 0.019 ADD: additive model Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

169

170

Nyholt & Bonferroni: P=0.002 Factor 1 (Symmetry, Ordering, OCD Symptom Dimension and Genotype using PLINK Linear Regression Repeating and Checking) 2.5 Factor 2 (Aggression, Sexual, and Religious) Factor 3 (Hoarding)

Factor 4 (Contamination and 2 Cleaning) Factor 5 (Somatic)

1.5 P=0.05

- Log10 P Value P - Log10 1

0.5

0

COMT rs4818 FAIM2 rs706795 FUT2 rs681343 COMT rs933271 BDNF rs3763965BDNF rs7124442 BDNF rs2883187 COMT rs9617850 MAOA rs979605 HTR1B rs1778258 BDNF rs61888800BDNF rs11030119 FAIM2 rs1044677FAIM2 rs7132908HTR2A rs7997012HTR2A rs9534510 BTBD3 rs1996132 MAOA rs3788862MAOA rs1465107 MAOA rs1137070 GRINK2 SLC1A1rs1556995 rs3933331SLC1A1 rs7031998SLC1A1 rs7022369 GRIN2B rs7301328 SLITRK5 SLC6A4rs9557425 rs3813034DLGAP1 rs2240899DLGAP1 rs8096794 SLITRK5 rs10450811 Figure 4.6. Linear Regression of 5-Factor Y-BOCS Symptom Dimensions in OCD. This graph illustrates the linear regression results of the 5-factor model of Y-BOCS derived symptom dimensions in OCD.

170

171

Table 4.11. Genetic Results of Linear Regression of 6-Factor Y-BOCS Symptom Dimensions in OCD.

Y-BOCS 6-Factor Model Factor 1 Factor 2 Factor 3 Factor 4 Factor 5 Factor 6 Symmetry/Order Hoarding Contamination/Cleaning Aggression/Checking Somatic Sexual/Religious Chromosome Gene SNP A1 BETA STAT P BETA STAT P BETA STAT P BETA STAT P BETA STAT P BETA STAT P 6 HTR1B rs1778258 T 0.114 0.869 0.385 -0.129 -0.982 0.327 -0.185 -1.436 0.152 0.012 0.090 0.928 -0.025 -0.189 0.851 -0.093 -0.725 0.469 6 GRIK2 rs1556995 C -0.058 -0.592 0.554 0.065 0.651 0.515 0.114 1.174 0.241 0.088 0.880 0.380 0.032 0.322 0.748 0.034 0.346 0.729 9 SLC1A1 rs3933331 C 0.086 0.911 0.363 0.038 0.399 0.690 -0.003 -0.036 0.971 -0.124 -1.295 0.196 0.038 0.410 0.682 -0.018 -0.193 0.847 9 SLC1A1 rs7031998 T -0.012 -0.129 0.898 -0.004 -0.046 0.963 -0.064 -0.721 0.472 -0.050 -0.550 0.583 -0.060 -0.678 0.498 0.082 0.925 0.356 9 SLC1A1 rs7022369 G -0.074 -0.944 0.346 -0.007 -0.090 0.928 -0.008 -0.110 0.913 0.085 1.082 0.280 -0.010 -0.126 0.900 -0.067 -0.875 0.382 11 BDNF rs3763965 A 0.203 2.562 0.011 -0.015 -0.186 0.852 0.155 1.998 0.047 0.038 0.478 0.633 0.172 2.189 0.029 -0.077 -0.994 0.321 11 BDNF rs7124442 C 0.035 0.412 0.681 -0.126 -1.468 0.143 0.036 0.422 0.674 -0.078 -0.897 0.371 0.118 1.375 0.170 0.036 0.428 0.669 11 BDNF rs61888800 T 0.000 -0.004 0.997 -0.090 -0.957 0.339 -0.027 -0.286 0.775 0.016 0.171 0.865 0.122 1.303 0.193 0.052 0.565 0.572 11 BDNF rs11030119 A -0.028 -0.326 0.745 -0.110 -1.273 0.204 0.035 0.404 0.687 -0.042 -0.479 0.632 0.076 0.881 0.379 0.083 0.980 0.328 11 BDNF rs2883187 G 0.070 0.900 0.369 -0.001 -0.011 0.991 0.139 1.810 0.071 0.006 0.075 0.940 0.043 0.552 0.582 -0.022 -0.286 0.775 12 GRIN2B rs7301328 C -0.163 -1.965 0.050 0.122 1.473 0.142 0.050 0.614 0.539 0.008 0.090 0.928 0.205 2.511 0.013 -0.096 -1.184 0.237 12 FAIM2 rs1044677 T -0.004 -0.050 0.960 0.076 0.895 0.371 -0.034 -0.399 0.690 -0.106 -1.253 0.211 0.185 2.211 0.028 -0.058 -0.701 0.484 12 FAIM2 rs7132908 A -0.057 -0.721 0.472 -0.047 -0.599 0.550 -0.036 -0.462 0.645 -0.002 -0.019 0.985 -0.080 -1.013 0.312 0.019 0.241 0.810 12 FAIM2 rs706795 T 0.008 0.098 0.922 0.008 0.103 0.918 0.060 0.765 0.445 0.081 1.031 0.303 0.059 0.750 0.454 0.000 -0.006 0.996 13 HTR2A rs7997012 A -0.005 -0.062 0.951 0.059 0.739 0.461 0.037 0.465 0.642 0.070 0.870 0.385 -0.023 -0.293 0.770 0.126 1.616 0.107 13 HTR2A rs9534510 T 0.203 2.556 0.011 -0.082 -1.018 0.310 -0.095 -1.207 0.228 0.050 0.624 0.533 -0.092 -1.154 0.249 0.120 1.537 0.125 13 SLITRK5 rs10450811 A 0.042 0.423 0.673 -0.113 -1.159 0.247 -0.044 -0.452 0.651 0.060 0.606 0.545 -0.039 -0.395 0.693 -0.092 -0.960 0.338 13 SLITRK5 rs9557425 G -0.036 -0.463 0.644 -0.102 -1.302 0.194 -0.050 -0.656 0.512 -0.126 -1.608 0.109 -0.017 -0.218 0.827 -0.024 -0.308 0.758 17 SLC6A4 rs3813034 C 0.056 0.734 0.463 0.009 0.123 0.902 -0.134 -1.815 0.070 -0.069 -0.912 0.363 0.017 0.221 0.825 -0.143 -1.968 0.050 18 DLGAP1 rs2240899 G 0.127 1.602 0.110 0.058 0.731 0.465 -0.067 -0.863 0.389 -0.093 -1.171 0.243 0.196 2.515 0.012 -0.124 -1.605 0.110 18 DLGAP1 rs8096794 A 0.114 1.476 0.141 0.062 0.799 0.425 0.053 0.705 0.481 0.151 1.963 0.051 0.096 1.259 0.209 0.100 1.322 0.187 19 FUT2 rs681343 C -0.100 -1.281 0.201 -0.063 -0.811 0.418 0.036 0.472 0.637 -0.068 -0.869 0.385 -0.125 -1.618 0.107 -0.075 -0.975 0.330 20 BTBD3 rs1996132 C 0.070 0.794 0.428 0.034 0.385 0.701 0.022 0.251 0.802 0.037 0.424 0.672 -0.030 -0.347 0.729 0.137 1.609 0.109 22 COMT rs9617850 A -0.020 -0.206 0.837 0.041 0.418 0.676 0.146 1.525 0.128 -0.006 -0.060 0.952 -0.022 -0.229 0.819 0.080 0.834 0.405 22 COMT rs933271 C -0.078 -0.942 0.347 0.001 0.017 0.986 0.141 1.734 0.084 -0.032 -0.380 0.704 -0.041 -0.494 0.622 0.071 0.881 0.379 22 COMT rs4818 G 0.013 0.168 0.867 -0.018 -0.223 0.823 -0.050 -0.642 0.521 -0.065 -0.812 0.418 -0.096 -1.220 0.224 0.228 2.986 0.003 ADD: additive model Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

171

172

Nyholt & Bonferroni: P=0.002

P=0.05

Figure 4.7. Linear Regression of 6-Factor Y-BOCS Symptom Dimensions in OCD. This graph illustrates the linear regression results of the 5-factor model of Y-BOCS derived symptom dimensions in OCD.

172

173

Table 4.12. Genetic Results of Psychiatric Comorbidities in OCD.

Comorbidity Mood Disorders Anxiety Disorders Yes No Yes No Gene SNP Genotype χ2 P Genotype χ2 P HTR1B rs1778258 16/3/0 247/64/2 0.377 0.828 55/12/0 208/55/2 0.807 0.668 GRIK2 rs1556995 2/8/8 11/99/202 4.436 0.109 3/23/41 10/84/169 0.235 0.889 SLC1A1 rs3933331 3/5/11 15/84/210 4.174 0.124 4/19/44 14/70/177 0.119 0.942 rs7031998 11/7/1 154/137/20 0.503 0.778 29/31/6 136/113/15 1.780 0.411 rs7022369 6/10/3 88/150/75 0.667 0.716 19/33/15 75/127/63 0.063 0.969 BDNF rs3763965 5/12/2 63/156/92 3.196 0.202 17/36/14 51/132/80 2.759 0.252 rs7124442 4/8/7 22/131/159 5.182 0.075 5/25/36 20/114/130 0.782 0.676 rs61888800 8/10/1 192/105/15 2.984 0.225 40/22/4 160/93/12 0.304 0.859 rs11030119 4/8/7 20/126/167 6.275 0.043 6/24/37 18/110/137 0.902 0.637 rs2883187 1/11/7 71/158/82 3.437 0.179 9/32/25 63/137/64 6.292 0.043 GRIN2B rs7301328 3/8/8 42/167/102 0.988 0.610 9/37/21 36/138/89 0.179 0.915 FAIM2 rs1044677 12/6/1 137/140/31 2.563 0.278 33/26/5 116/120/27 1.233 0.540 rs7132908 6/7/6 44/148/121 4.305 0.116 11/31/25 39/124/102 0.125 0.939 rs706795 9/9/1 112/149/50 1.986 0.371 23/33/11 98/125/40 0.209 0.901 HTR2A rs7997012 5/7/7 45/149/117 2.125 0.346 8/35/24 42/121/100 1.080 0.583 rs9534510 5/12/2 104/150/57 1.697 0.428 19/38/10 90/124/49 1.962 0.375 SLITRK5 rs10450811 1/6/11 13/99/197 0.100 0.951 2/17/46 12/88/162 1.825 0.401 rs9557425 8/7/4 115/145/51 0.731 0.694 20/34/12 103/118/43 1.722 0.423 SLC6A4 rs3813034 6/9/3 92/144/73 0.468 0.791 21/28/17 77/125/59 0.656 0.720 DLGAP1 rs2240899 7/8/4 107/149/55 0.271 0.873 19/31/16 95/126/43 2.667 0.264 rs8096794 3/13/3 82/151/80 2.918 0.232 15/29/23 70/135/60 3.896 0.143 FUT2 rs681343 2/10/7 78/158/77 2.617 0.270 14/34/19 66/134/65 0.667 0.716 BTBD3 rs1996132 1/8/10 25/111/176 0.430 0.806 9/23/35 17/96/151 3.623 0.163 COMT rs9617850 0/6/13 13/102/196 0.885 0.643 2/23/41 11/85/168 0.306 0.858 rs933271 2/6/11 31/119/163 0.320 0.852 7/23/37 26/102/137 0.395 0.821 rs4818 7/9/3 107/150/56 0.083 0.960 15/36/16 99/123/43 5.851 0.054 MAOA rs3788862 7/2/10 63/63/186 3.347 0.188 16/9/42 54/56/154 2.114 0.347 rs1465107 6/3/10 62/62/187 1.503 0.472 16/9/42 52/56/155 2.238 0.327

173

174

rs979605 7/4/8 63/59/187 3.327 0.189 16/7/43 54/56/152 3.960 0.138 rs1137070 8/4/7 186/60/66 2.966 0.227 43/6/17 151/58/56 5.589 0.061 OCRDs Tics Yes No Yes No Gene SNP Genotype χ2 P Genotype χ2 P HTR1B rs1778258 14/2/1 249/65/1 8.943 0.011 19/0/0 244/67/2 5.287 0.071 GRIK2 rs1556995 1/4/12 12/103/198 0.745 0.689 1/5/13 12/102/197 0.394 0.821 SLC1A1 rs3933331 1/3/13 17/86/208 0.820 0.664 1/4/14 17/85/207 0.396 0.820 rs7031998 9/7/1 156/137/20 0.062 0.969 9/10/0 156/134/21 1.685 0.431 rs7022369 4/9/4 90/151/74 0.229 0.892 8/4/7 86/156/71 5.963 0.051 BDNF rs3763965 3/8/5 65/160/89 0.076 0.963 6/11/2 62/157/92 3.654 0.161 rs7124442 0/12/5 26/127/161 6.434 0.040 3/7/9 23/132/157 1.774 0.412 rs61888800 9/8/0 191/107/16 1.815 0.404 11/8/0 189/107/16 1.310 0.519 rs11030119 0/12/5 24/122/169 7.159 0.028 1/8/10 23/126/164 0.123 0.940 rs2883187 2/12/3 70/157/86 2.722 0.256 1/12/6 71/157/83 3.253 0.197 GRIN2B rs7301328 2/12/3 43/163/107 2.411 0.300 6/7/6 39/168/104 5.775 0.056 FAIM2 rs1044677 4/11/2 145/135/30 3.599 0.165 8/9/2 141/137/30 0.098 0.952 rs7132908 1/11/5 49/144/122 2.614 0.271 4/10/5 46/145/122 1.384 0.501 rs706795 7/7/3 114/151/48 0.323 0.851 8/9/2 113/149/49 0.481 0.786 HTR2A rs7997012 2/12/3 48/144/121 4.091 0.129 5/6/8 45/150/116 2.819 0.244 rs9534510 7/8/2 102/154/57 0.750 0.687 5/10/4 104/152/55 0.440 0.802 SLITRK5 rs10450811 1/6/9 13/99/199 0.448 0.799 1/5/12 13/100/196 0.212 0.900 rs9557425 8/7/2 115/145/53 0.812 0.666 8/8/2 115/144/53 0.626 0.731 SLC6A4 rs3813034 3/7/7 95/146/69 3.511 0.173 5/8/6 93/145/70 0.788 0.674 DLGAP1 rs2240899 6/7/4 108/150/55 0.477 0.788 6/10/3 108/147/56 0.209 0.901 rs8096794 6/8/3 79/156/80 1.065 0.587 7/8/4 78/156/79 1.337 0.512 FUT2 rs681343 8/9/0 72/159/84 8.473 0.014 4/10/5 76/158/79 0.102 0.950 BTBD3 rs1996132 1/5/11 25/114/175 0.533 0.766 3/4/12 23/115/174 3.034 0.219 COMT rs9617850 0/6/11 13/102/198 0.747 0.688 0/9/10 13/99/199 2.479 0.289 rs933271 1/8/8 32/117/166 0.816 0.665 1/9/9 32/116/165 1.047 0.593 rs4818 8/6/3 106/153/56 1.438 0.487 5/10/4 109/149/55 0.594 0.743 MAOA rs3788862 2/5/10 68/60/186 1.618 0.445 2/2/15 68/63/181 3.253 0.197 rs1465107 2/5/10 66/60/187 1.541 0.463 2/2/15 66/63/182 3.106 0.212

174

175

rs979605 2/5/10 68/58/185 1.744 0.418 2/0/17 68/63/178 8.037 0.018 rs1137070 9/5/3 185/59/70 1.197 0.550 16/0/3 178/64/70 6.510 0.039 Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

175

176

Table 4.13. Genetic Results of Family History in OCD.

FHI OCD TTM Gene SNP Genotype Positive Negative χ2 P Positive Negative χ2 P HTR1B rs1778258 CC/TC/TT 25/10/1 122/37/1 1.782 0.410 5/1/0 226/65/3 0.172 0.917 GRIK2 rs1556995 CC/CT/TT 2/15/19 7/57/96 0.642 0.725 0/1/5 13/102/177 1.331 0.514 SLC1A1 rs3933331 CC/CG/GG 1/9/26 11/38/109 0.885 0.642 0/2/4 17/78/196 0.444 0.801 rs7031998 GG/GT/TT 11/21/4 82/67/10 5.375 0.068 2/4/0 144/126/22 1.504 0.471 rs7022369 CC/CG/GG 14/12/10 44/83/33 4.058 0.131 6/0/0 77/145/72 16.007 <0.001 BDNF rs3763965 AA/AT/TT 7/20/9 36/80/44 0.371 0.831 0/5/1 63/137/92 3.346 0.188 rs7124442 CC/CT/TT 2/15/19 11/72/76 0.317 0.853 1/2/3 21/124/148 0.834 0.659 rs61888800 GG/GT/TT 21/15/0 95/57/7 1.864 0.394 4/2/0 178/103/12 0.285 0.867 rs11030119 AA/AG/GG 2/13/21 11/68/81 0.703 0.704 1/1/4 21/115/158 1.676 0.433 rs2883187 AA/AG/GG 7/18/10 30/75/52 0.274 0.872 1/3/2 68/136/85 0.160 0.923 GRIN2B rs7301328 CC/CG/GG 6/15/15 22/84/51 1.664 0.435 1/3/2 39/149/103 0.055 0.973 FAIM2 rs1044677 CC/CT/TT 17/17/1 66/72/20 2.894 0.235 3/3/0 132/126/31 0.723 0.697 rs7132908 AA/AG/GG 6/21/9 25/71/64 3.015 0.221 1/4/1 49/138/107 1.119 0.571 rs706795 CC/CT/TT 12/15/9 49/85/25 2.304 0.316 2/4/0 96/150/47 1.236 0.539 HTR2A rs7997012 AA/AG/GG 5/25/6 30/68/62 9.016 0.011 1/5/0 48/137/109 3.901 0.142 rs9534510 GG/GT/TT 10/17/9 61/75/23 2.889 0.236 2/3/1 101/139/53 0.017 0.992 SLITRK5 rs10450811 AA/AG/GG 1/9/26 8/54/97 1.638 0.441 0/1/5 15/92/184 1.102 0.576 rs9557425 CC/CG/GG 12/20/4 59/75/25 0.958 0.619 1/3/2 104/145/44 1.876 0.391 SLC6A4 rs3813034 AA/AC/CC 9/17/8 54/71/33 0.754 0.686 0/4/2 95/135/59 2.943 0.230 DLGAP1 rs2240899 CC/CG/GG 15/12/8 50/85/25 4.097 0.129 2/3/1 100/141/51 0.007 0.996 rs8096794 AA/AG/GG 5/24/7 38/75/47 4.634 0.099 0/6/0 71/143/80 6.205 0.045 FUT2 rs681343 CC/CT/TT 5/19/12 39/84/37 2.665 0.264 0/5/1 72/151/71 2.763 0.251 BTBD3 rs1996132 CC/CG/GG 4/14/18 14/64/80 0.181 0.913 0/2/4 26/106/160 0.699 0.705 COMT rs9617850 AA/AG/GG 0/13/23 10/51/99 2.431 0.297 1/1/4 12/98/182 2.648 0.266 rs933271 CC/CT/TT 1/15/20 19/59/82 2.667 0.264 1/2/3 29/115/150 0.324 0.851 rs4818 CC/CG/GG 10/22/4 51/82/27 1.314 0.518 3/3/0 95/149/50 1.584 0.453 MAOA rs3788862 AA/AG/GG 8/9/19 33/33/93 0.442 0.802 2/2/2 62/65/166 1.302 0.521 rs1465107 AA/AG/GG 8/9/19 32/33/94 0.512 0.774 2/2/2 60/66/167 1.353 0.508 rs979605 AA/AG/GG 6/7/21 34/33/93 0.240 0.887 2/1/3 60/66/165 0.596 0.742 rs1137070 CC/CT/TT 22/7/7 92/33/34 0.130 0.937 3/1/2 165/67/61 0.578 0.749

176

177

FHI BDD Hoarding Gene SNP Genotype Positive Negative χ2 P Positive Negative χ2 P HTR1B rs1778258 CC/CT/TT 5/3/0 226/63/3 1.204 0.548 10/0/0 122/37/1 3.059 0.217 GRIK2 rs1556995 CC/CT/TT 0/3/5 13/100/177 0.382 0.826 0/2/8 7/57/96 1.715 0.424 SLC1A1 rs3933331 CC/CG/GG 2/4/2 15/76/198 9.149 0.010 1/3/6 11/38/109 0.369 0.832 SLC1A1 rs7031998 GG/GT/TT 4/4/0 142/126/22 0.685 0.710 6/4/0 82/67/10 0.768 0.681 SLC1A1 rs7022369 CC/CG/GG 3/4/1 80/141/71 0.745 0.689 3/4/3 44/83/33 0.671 0.715 BDNF rs3763965 AA/AT/TT 2/4/2 61/138/91 0.169 0.919 1/5/4 36/80/44 1.197 0.550 rs7124442 CC/CT/TT 1/3/4 21/123/147 0.338 0.845 0/6/4 11/72/76 1.254 0.534 rs61888800 GG/GT/TT 6/2/0 176/103/12 0.839 0.657 4/6/0 95/57/7 2.539 0.281 rs11030119 AA/AG/GG 1/3/4 21/113/158 0.326 0.849 0/5/5 11/68/81 0.812 0.666 rs2883187 AA/AG/GG 1/1/5 68/138/82 6.170 0.046 3/6/0 30/75/52 4.470 0.107 GRIN2B rs7301328 CC/CG/GG 0/3/5 40/149/100 3.122 0.210 1/7/2 22/84/51 1.041 0.594 FAIM2 rs1044677 CC/CT/TT 4/4/0 131/125/31 0.970 0.616 4/5/1 66/72/20 0.101 0.951 rs7132908 AA/AG/GG 2/5/1 48/137/107 2.003 0.367 2/5/3 25/71/64 0.419 0.811 rs706795 CC/CT/TT 2/6/0 96/148/47 2.325 0.313 4/6/0 49/85/25 1.900 0.387 HTR2A rs7997012 AA/AG/GG 2/4/2 47/138/107 0.681 0.711 1/5/4 30/69/62 0.522 0.770 rs9534510 GG/GT/TT 2/4/2 101/138/52 0.443 0.801 5/2/3 61/75/23 3.326 0.190 SLITRK5 rs10450811 AA/AG/GG 0/2/6 15/91/183 0.687 0.709 0/3/6 8/54/97 0.498 0.779 rs9557425 CC/CG/GG 2/4/2 103/144/44 0.734 0.693 5/3/2 59/75/25 1.122 0.571 SLC6A4 rs3813034 AA/AC/CC 2/3/3 93/136/58 1.419 0.492 2/7/1 54/71/33 2.390 0.303 DLGAP1 rs2240899 CC/CG/GG 1/5/1 101/139/51 1.666 0.435 4/5/1 50/85/25 0.439 0.803 rs8096794 AA/AG/GG 3/5/0 68/144/80 3.129 0.209 1/7/2 38/75/47 2.106 0.349 FUT2 rs681343 CC/CT/TT 2/4/2 70/152/70 0.013 0.993 3/4/3 39/84/37 0.594 0.743 BTBD3 rs1996132 CC/CG/GG 0/2/6 26/106/158 1.599 0.449 1/3/6 14/64/80 0.435 0.804 COMT rs9617850 AA/AG/GG 0/4/4 13/95/182 1.258 0.533 0/2/8 10/51/99 1.542 0.463 rs933271 CC/CT/TT 0/5/3 30/112/150 2.279 0.320 0/3/7 19/59/82 1.941 0.379 rs4818 CC/CG/GG 3/5/0 95/147/50 1.656 0.437 3/5/2 51/82/27 0.067 0.967 MAOA rs3788862 AA/AG/GG 3/1/4 61/66/164 1.411 0.494 2/2/6 33/33/93 0.009 0.996 MAOA rs1465107 AA/AG/GG 3/1/4 59/67/165 1.555 0.460 2/2/6 32/33/94 0.004 0.998 MAOA rs979605 AA/AG/GG 4/1/3 58/66/165 4.238 0.120 1/2/7 34/33/93 0.806 0.668 MAOA rs1137070 CC/CT/TT 3/1/4 165/67/59 4.155 0.125 7/2/1 92/33/34 0.828 0.661 Highlighted in yellow indicates significant P value prior to correction for multiple comparisons

177

178

Table 4.14. Genetic Results of Family History (combined OCRDs) in OCD.

FHI OCRD(s) Gene SNP Genotype Positive Negative χ2 P HTR1B rs1778258 CC/TC/TT 63/20/1 168/46/2 0.277 0.871 GRIK2 rs1556995 CC/CT/TT 2/28/54 11/75/128 1.302 0.522 SLC1A1 rs3933331 CC/CG/GG 3/23/57 14/57/143 0.952 0.621 rs7031998 GG/GT/TT 35/40/8 111/90/14 2.437 0.296 rs7022369 CC/CG/GG 25/33/26 58/112/46 4.511 0.105 BDNF rs3763965 AA/AT/TT 48/100/66 15/42/27 0.765 0.682 rs7124442 CC/CT/TT 6/30/48 16/96/103 2.172 0.388 rs61888800 GG/GT/TT 53/28/3 129/77/9 0.258 0.879 rs11030119 AA/AG/GG 6/26/52 16/90/110 3.151 0.207 rs2883187 AA/AG/GG 22/39/22 47/100/65 0.829 0.661 GRIN2B rs7301328 CC/CG/GG 12/38/34 28/114/71 1.735 0.420 FAIM2 rs1044677 CC/CT/TT 45/32/5 90/97/26 4.739 0.094 rs7132908 AA/AG/GG 15/47/22 35/95/86 5.049 0.080 rs706795 CC/CT/TT 25/43/16 73/111/31 1.149 0.563 HTR2A rs7997012 AA/AG/GG 13/44/27 36/98/82 1.244 0.537 rs9534510 GG/GT/TT 24/38/22 79/104/32 5.572 0.062 SLITRK5 rs10450811 AA/AG/GG 4/22/57 11/71/132 1.322 0.516 rs9557425 CC/CG/GG 22/48/14 83/100/32 4.155 0.125 SLC6A4 rs3813034 AA/AC/CC 23/40/19 72/99/42 1.017 0.602 DLGAP1 rs2240899 CC/CG/GG 32/37/13 70/107/39 1.165 0.558 rs8096794 AA/AG/GG 19/48/17 52/101/63 3.175 0.204 FUT2 rs681343 CC/CT/TT 16/44/24 56/112/48 2.211 0.331 BTBD3 rs1996132 CC/CG/GG 9/26/49 17/82/115 1.665 0.435 COMT rs9617850 AA/AG/GG 2/31/51 11/68/135 1.585 0.453 rs933271 CC/CT/TT 6/37/41 24/80/112 1.824 0.402 rs4818 CC/CG/GG 27/46/11 71/106/39 1.289 0.525 MAOA rs3788862 AA/AG/GG 20/17/47 44/50/121 0.562 0.755 rs1465107 AA/AG/GG 19/18/47 43/50/122 0.295 0.863 rs979605 AA/AG/GG 19/16/47 43/51/121 0.763 0.683 rs1137070 CC/CT/TT 48/16/20 120/52/43 1.136 0.567 This table illustrates the genetic results of presence versus absence of family history of OCRDs in OCD.

178

179

4.5 Discussion This is the first comprehensive genetic study of OCD subphenotypes in a large well- characterized OCD sample using SNP with implicated remote regulatory function. Furthermore, this is the first study that investigated the significance of family history of OCRDs in the genetics of OCD. Two studies have investigated positive family history of OCD in influencing the risk of OCD. Deny et al. (2006) previously reported that the familial OCD group (with positive family history of OCD) had significantly higher GG genotype of HTR2A rs6313 (P=0.015). However, the study did not disclose any details regarding the assessment of family history. Two studies by the same research group used the presence of obsessive-compulsive symptoms in the first-degree relatives of OCD subjects as significant family history of OCD (Katerberg et al., 2009; Katerberg et al., 2010). The studies examined COMT rs4680 (Katerberg et al., 2010) and BDNF rs6265 (Katerberg et al., 2009), which were both negative for family history of obsessive-compulsive symptoms. Therefore, this is the first genetic study that used a comprehensive interview to determine the family history status of the OCD individual. In this study, glutamatergic system genes (SLC1A1 and DLGAP1) may possibly be implicated in OCD individuals who have a significant family history of TTM and BDD while serotonergic system gene (HTR2A) may otherwise be involved in the familial cases of OCD.

Although this is a relatively large OCD sample with consistent recruitment criteria and assessment protocol, stratifying into different subgroups of OCD would require an even larger sample in order to detect any true positive findings and to reduce the likelihood of false negative results. The main reason for the negative study is likely due to the lack of power, especially for the family history and psychiatric comorbidity analyses.

Putting the limited power of this study aside, interesting results in the newly identified gene, FUT2, may potentially predict later onset of OCD in addition to confer risk to both OCD and other OCRDs as shown in the comorbidity analysis. However, the role of FUT2 has never been studied or associated with any psychiatric disorders except for the first GWAS of OCD (Stewart et al., 2013), which provided the initial rationale for investigating this gene in OCD subphenotypes. Two GWASs have reported that genetic variations within FUT2 significantly predicted plasma vitamin B12 levels (Hazra et al., 2008; Tanaka et al., 2009) and FUT2 has been

179

180

associated with Crohn’s disease (McGovern et al., 2010). Interestingly, OCD individuals were found to have significantly lower vitamin B12 serum level than healthy controls (Hermesh et al., 1988; Türksoy et al., 2014). Furthermore, vitamin B12 deficiency has been reported to be associated with OCD symptoms (Sharma and Biswas, 2012). Thus, this gene may be associated with OCD in the context of low vitamin B12 serum level. Thus, exploration of this gene with vitamin B12 serum level in OCD patients may lead to the identification of a potential biomarker for OCD subphenotype and/or diagnosis of OCD.

Another highly implicated gene, BDNF, was shown to potentially predict greater OCD severity and confer risk to both mood and anxiety disorders in addition to OCD. This suggests an overlapping and common genetic pathway, which may predispose an individual to multiple psychiatric disorders. Evidence suggested that BDNF has pleiotropic effects on neuroplasticity and has been implicated in the susceptibility to many psychiatric disorders including OCD (Alonso et al., 2008; Hall et al., 2003; Hemmings et al., 2013; Márquez et al., 2013; Katerberg et al., 2009; Rocha et al., 2010), mood disorders (Neves-Pereira et al., 2002; Sen et al., 2003; Sklar et al., 2002), and schizophrenia (Krebs et al., 2000; Muglia et al., 2003), supporting a genetic overlap between these disorders.

Caution should be taken when interpreting our findings for the comorbidity and family history analyses because of the stratification strategies that we performed, which significantly lower the power of each stratified group to detect meaningful differences in each SNP. Results of OCD symptom dimensions did not yield conclusive evidence of genetic differences especially between hoarding and the other OCD symptoms although there was a high probability of a false negative study due to the sample size. Hoarding symptoms have been removed from the diagnosis of OCD and are now considered to be one of the diagnostic criteria of hoarding disorder (APA, 2013).

Another limitation of this study is the complexity and consistency of the provided clinical information. Since many of the examined subphenotypes (clinical variables) were collected retrospectively, recall bias may impact on the reliability of the clinical data. This study attempted to minimize recall bias by incorporating a reliability measure to detail whether the participants had good recall of the given clinical information, especially in the family history

180

181

data. All unreliable data were excluded from the analysis (approximately 25% of the family history data were deemed unreliable) and although this further reduced the power of this study, this is a more stringent way to analyze cleaner dataset. Thus, an even larger OCD sample with additional clinical information, for example, a consistent definition of OCD onset (i.e., first onset of symptom versus first diagnosis), symptom severity, duration of illness, symptom dimensional subtype (i.e., target/principal symptoms), illness progression, and other treatment modalities in addition to antidepressants (i.e., cognitive behavioural therapy, transmagnetic stimulation, deep brain stimulation) is required to further explore the neurobiological basis of OCD.

OCD is a complex, heterogeneous, and polygenic disorder, which is challenging to study with traditional genetic approach of using disease status as the phenotypic trait. The identification of genetic predictors depends on various factors including the genetic architecture of the trait, in particular, the number of markers that affect the trait and the distribution of their effects. The relative risk of one single gene conferring susceptibility to a psychiatric disorder is small (Schizophrenia Working Group of the Psychiatric Genomics Consortium, 2014). Thus, in addition to increasing the sample size, gene-gene interaction and epigenetics (or gene- environment interactions) or epigenetics, may prove to amplify the risk of OCD traits and to identify the missing heritability of this disorder while using complex computational model to handle large dataset. Recent advances in human genomics including the ENCODE project have provided us with enormous potentials to further our breakthroughs in the discovery of new genetic basis of OCD. Moreover, current OCD diagnosis is a dichotomous distinction between healthy and disease status; thus, by examining obsessive-compulsive trait in a bidirectional continuum in healthy, subclinical, and clinical population may allow researchers to collect larger sample and to identify specific targets for each trait or symptom.

Recent advances in technology and international collaboration provide important insights into the genetic dissection of OCD. Recent GWASs and many candidate gene studies have identified genetic variations that may contribute to a small effect of the susceptibility risk to OCD. However, it is currently still immature to conclude effects of a single gene in any psychiatric disorders and a vast majority of the genetic architecture and risk profile of psychiatric disorders remains unknown. Future research should continue to expand our current knowledge

181

182

by improving clinical and biological assessment of individualized risk, which may aid in the development of novel therapeutics.

182

183

Chapter 5 GWAS OF OCD PHENOTYPES

5 Genome-Wide Association Study of Obsessive- Compulsive Disorder Phenotypes

Gwyneth Zai 1,2,4, Clement Zai 1,2, Vanessa Gonçalves 1,2, Joanne Knight 1,2, David Pauls 3, Karen Wigg 4, Margaret A. Richter 4, James L. Kennedy 1,2

1 Neurogenetics Section, Centre for Addiction and Mental Health, Clarke Division, Toronto, ON, Canada

2 Department of Psychiatry, University of Toronto, Toronto, ON, Canada

3 Psychiatric and Neurodevelopmental Genetics Unit, Center for Human Genetic Research, Massachusetts General Hospital, Boston, MA, USA

4 Frederick W. Thompson Anxiety Disorders Centre, Department of Psychiatry, Sunnybrook Health Sciences Centre, Toronto, ON, Canada

* Corresponding author: Dr. James L. Kennedy, Director of the Neuroscience Department and

Head of Neurogenetics Section at the Centre for Addiction and Mental Health (CAMH),

Professor of the Department of Psychiatry and the Institute of Medical Science at University of

Toronto, Toronto, ON M5T 1R8, Canada (Phone: 1-416-979-4987; Fax: 1-416-979-4666; E- mail: [email protected])

183

184

5.1 Abstract

Background: Genetic studies in obsessive-compulsive disorder (OCD) have yielded inconsistent results and two recent genome-wide association studies (GWASs) did not report any genome- wide significant results. OCD is a heterogeneous illness with a strong genetic etiology. Identification of a more homogeneous phenotype in OCD may provide a better understanding of the diversity in OCD. We aimed to examine the genetics of two common phenotypes in OCD including age at onset (AAO) and Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score.

Method: A genome-wide association study of 269 OCD individuals with 137 first-degree relatives and the initial quality control were conducted according to the International OCD Foundation Genetic Collaborative (IOCDFGC) Group. After further quality control, 594,332 SNPs, 203 OCD subjects and 94 relatives remained for quantitative analysis using PLINK and R programs.

Results: No genome-wide significance was detected in AAO or Y-BOCS severity score in OCD.

Conclusions: GWAS requires larger sample size to identify common genetic variants that may contribute to OCD phenotypes.

5.2 Introduction

Obsessive-compulsive disorder is one of the major disorders in the newly defined category of psychiatric disorders according to the DSM-5 (APA, 2013). Individuals suffering from OCD present with recurrent, intrusive, and persistent obsessional thoughts, images, urges, or impulses that cause marked anxiety or distress, in addition to repetitive behaviours with an attempt to alleviate anxiety or distress from the obsession (APA, 2013). These symptoms are persistent, taking at least one hour per day, which cause significant impairment in social, occupational, or other important areas of functioning (APA, 2013). OCD patients often recognized these symptoms as excessive and irrational but they have difficulty to ignore them.

184

185

Evidence supports a strong genetic etiology in the underlying mechanism of OCD from previous family, twins, segregation, candidate gene, and genome-wide association studies (GWASs) (Pauls et al., 2014). Family studies repeatedly detected higher rates of OCD in the relatives of OCD patients (Pauls, 2010). Twins studies have reported heritability between 27% and 65% (van Grootheest et al., 2008). Segregation studies have consistently detected genetic transmission within families (Pauls, 2010). Over 500 candidate gene studies have implicated the involvement of several important neurotransmitter systems in OCD (Pauls et al., 2014). More recently, two GWASs of OCD have been published (Mattheisen et al., 2015; Stewart et al., 2013). However, most candidate gene studies and both GWASs mainly explored the genetics of OCD diagnosis and did not examine subphenotypes of OCD. Furthermore, results have not been robustly replicated. Given the complexity of OCD in terms of clinical presentation with multiple common psychiatric comorbidities, it is important to also examine OCD subphenotypes in addition to disease status in GWAS of OCD.

Two GWASs of OCD have been published recently. The first GWAS did not detect genome-wide significant association between any tested markers and OCD diagnosis (Stewart et al., 2013); however, interesting trend associations were observed in several glutamatergic system genes including the ‘discs large (drosophila) homolog-associated protein 1’ (DLGAP1) and glutamate receptor, ionotropic, kainate 2 (GRIK2) genes. The second GWAS also did not find any genome-wide significant results (Mattheisen et al., 2015) but detected a trend in a top-hit marker on chromosome 9 near the protein-tyrosine phosphatase, receptor-type, delta (PTPRD) gene (P=4.13×10-7), which promotes glutamatergic synaptic differentiation. Nonetheless, none of the GWASs investigated the role of age at onset (AAO) and severity of OCD symptoms in contributing to the genetic etiology of OCD.

A meta-analysis comparing clinical characteristics between early- and late-onset OCD reported that males with greater severity, higher rates of comorbid tics, and higher prevalence of OCD in their first-degree relatives are associated with earlier onset (Taylor, 2011). Thus, earlier onset of OCD has been postulated to have a higher genetic loading than late-onset OCD (Chabane et al., 2005; do Rosario-Campos et al., 2005; Hemmings et al., 2004). Although a total of 47 genes have been explored in relation to AAO in OCD, most were either single study with no published replication or studies with inconsistent and conflicting results (Table 1.5). The

185

186

most robust findings were two significant meta-analyses that examined a polymorphism, 5HTTLPR, within the serotonin transporter gene, SLC6A4 (Bloch et al., 2008) and reported an over-transmission of the La allele to early-onset OCD (P=0.00021) (Walitza et al., 2014). In addition to SLC6A4, the serotonin 2A receptor gene, HTR2A, has also been investigated in nine studies with mixed results (Denys et al., 2006; Dickel et al., 2007; Hemmings et al., 2004; Hemmings et al., 2006; Lochner et al., 2004; Lochner et al., 2008; Mas et al., 2014; Walitza et al., 2002; Walitza et al., 2012). Only one study examined linkage and reported a LOD score of two on chromosome 9q where the glutamate transporter gene, SLC1A1, is located (Hanna et al., 2002) and three (Dallaspezia et al., 2014; Dickel et al., 2006; Wu et al., 2013) out of five studies [negative: (Mas et al., 2014; Veenstra-VanderWeele et al., 2001)] reported significant findings, each with different single nucleotide polymorphisms (SNPs), which were associated with AAO. In addition to the serotonin and glutamate system genes, the brain-derived neurotrophic factor gene (BDNF) was first reported to be significantly associated with AAO by Hall et al. (2003) and only two of eight subsequent studies supported the role of the rs6265 marker in AAO with gender bias (Hemmings et al., 2006; Katerberg et al., 2009). Thus, with such discrepant genetic findings, it is important to replicate some of these findings in a much larger sample of OCD, which can take into account the influence of AAO in the genetic architecture of OCD.

OCD severity as defined by the gold standard Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score has previously been investigated in genetic studies. Higher severity of OCD has been found to be associated with greater genetic loading (i.e., higher familiality) (Schooler et al., 2008). However, only five studies have explored the effect of severity in the basis of OCD. Both studies by Hemmings et al. (2006; 2008) reported significant association between the BDNF rs6265 Val/Val genotype and higher severity in OCD females (P=0.013- 0.045) but was not replicated in a recent study (Márquez et al., 2013). Three additional genes, the glutamate receptor, ionotropic, N-methyl D-aspartate 2B (GRIN2B), HTR2A, and dopamine D1 receptor (DRD1) from the earlier study by Hemmings et al. (2006) were found to be associated with OCD severity but they have yet been replicated. Hence, given the limited studies exploring OCD severity, additional research is needed.

This is the first attempt to investigate the role of genetics in two different OCD subphenotypes including AAO and Y-BOCS severity in a non-hypothesis-driven GWAS. By

186

187

examining the genetics of AAO and severity in OCD, we may have a better chance of identifying genetic markers that may predict prognosis and help guide treatment decisions in addition to further improve our understanding in the genetic architecture of OCD.

5.3 Methods 5.3.1 Diagnostic Criteria and Sample

Two hundred and sixty-nine OCD subjects and 137 first-degree relatives were recruited from the Centre for Addiction and Mental Health and the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre. DSM-IV OCD diagnosis was confirmed using the Structured Clinical Interview for DSM-IV (SCID) (First et al., 1996) and available medical records. Age at onset (AAO) was defined as the age when the OCD subject first met diagnostic criteria. Symptom severity was determined using the Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) (Goodman et al., 1989). Trained research assistants who were supervised by an experienced psychiatrist performed all assessments and interviews. All participants provided their written informed consent to take part in this study, and research ethics approval was obtained from the local Research Ethics Boards. Inclusion criteria consisted of adult who met DSM-IV criteria of OCD diagnosis with more than one year symptom duration or more than one year since onset if waxing/waning course, and ages between 18 and 65 years. Exclusion criteria comprised of any metabolic or chronic neurological disease (other than tic disorder), schizophrenia, schizoaffective disorder, or bipolar disorder.

5.3.2 Genotyping and Statistical Analyses

Genotyping for this genome-wide association study (GWAS) was conducted using the genome-wide Illumina Human610-Quadv1_B SNP BeadChip array (Illumina Inc. ®, San Diego, CA, USA) and initial quality control was performed by the International OCD Foundation Genetic Collaborative (IOCDFGC) Group (Stewart et al., 2013). Additional quality control and statistical analyses were performed using PLINK (version 1.07) (Purcell et al., 2007), R (version 3.0.2) (R Core Team, 2013), and LocusZoom (version 1.1) (Pruim et al., 2010) programs. 187

188

For further quality control (see Table 5.1 for full details), all 406 individuals had greater than 95% of the markers genotyped, and all 594,332 single nucleotide polymorphisms (SNPs) were greater than 95% genotyped or had a minor allele frequency greater than 5%. All SNPs did not deviate significantly from Hardy-Weinberg Equilibrium (P>0.0001). Thirty-six individuals were removed because of sex discrepancy from the genetic data (N=14), call rate cut-off of greater than 2% (N=17), and outliers from the mean heterozygosity (N=5). Cryptic relatedness was assessed and one individual of each pair of related individuals (defined as pairs with PI^HAT>0.185) was excluded if there is more missing phenotype or genotype information for that individual, and a total of three individuals were removed. We performed a multi- dimensional scaling (MDS) analysis of the genotypes to ascertain the ethnicity of the samples and 88 outliers were removed. After sample refinement and updating map position, 594,332 SNPs, 203 OCD subjects and 94 family members remained for statistical analyses.

Our sample has enough power (81.5%) to detect 18% of the variance of AAO and Y- BOCS severity that are accounted for by the SNP (α=5E-08, additive model) using the QUANTO program (version 1.2.4) (Gauderman and Morrison, 2006).

We performed quantitative analyses on the AAO (using Log10 of AAO) and Y-BOCS severity variables (qfam algorithm) in the PLINK software (version 1.07) (Purcell et al., 2007). The qfam algorithm adopts the within and between model (Abecasis et al., 2000; Fulker et al., 1999) and uses permutation to account for related individuals (e.g., Perroud et al., 2011). We included age, sex, and the first ten components from the multi-dimensional scaling (MDS) analysis as covariates. The quantile-quantile plots, MDS plots, and Manhattan plots were drawn in R program.

188

189

Table 5.1. Genome-Wide Association Study (GWAS) Quality Control.

Quality Control Sample (remaining total) SNP Raw Total 406 (269 OCD+137 relatives) 594332 Missingness 0 0 Minor allele frequency - 0 Sex discrepancy 14 (258 OCD+134 relatives=392) - Call rate cut-off >2% 17 (248 OCD+127 relatives=375) - Heterozygosity rates 5 (245 OCD+125 relatives=370) - Identity-By-Descent (IBD) >0.185 3 (243 OCD+124 relatives=367) - Ancestry outliers 88 (203 OCD+94 relatives=297) - This table illustrates the quality control steps for this GWAS in OCD.

5.4 Results

The OCD sample demographics are described in Table 5.2. The skewness of the AAO distribution was 1.209, which significantly deviates from normal distribution (Figure 5.1) and thus it was transformed into log base 10 of AAO with a resulting skewness of -0.347 (Figure 5.2), which was used for the analysis. Y-BOCS severity score did not deviate greatly from the normal distribution with a skewness of -0.571 (Figure 5.3). The genomic inflation value (lambda) for AAO is 1.003295 and for Y-BOCS is 1.048644.

Table 5.2. Demographic Data of the Entire OCD Sample.

Descriptors Mean (S.D.) N (final data set) 203 (probands) & 94 (relatives) Age 34.4 (12.0) Gender 52.3% female Ethnicity 100% Caucasian (CEU & TSI) Age at Onset 13.3 (7.9) Yale-Brown Obsessive Compulsive Scale (Y-BOCS) severity score 24.6 (8.2) This table illustrates the OCD subject demographics of this GWAS.

189

190

Figure 5.1. Frequency Distribution of Age At Onset (AAO). This graph illustrates the frequency distribution of AAO in the OCD sample.

190

191

Figure 5.2. Frequency Distribution of Log10 Transformation of Age At Onset (AAO). This graph illustrates the log10 transformation of the frequency distribution of AAO in the OCD sample.

191

192

Figure 5.3. Frequency Distribution of Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score. This graph illustrates the frequency distribution of Y-BOCS severity score in the OCD sample.

192

193

The results for AAO and Y-BOCS severity score are summarized in Table 5.3 and Table 5.4, Figure 5.4 and Figure 5.5 (quantile-quantile plots), Figures 5.5-5.9 (MDS plots – Table 5.5 for ancestry label), Figure 5.10 and Figure 5.11 (Manhattan plots) respectively. No genome- wide significant results were detected for AAO or Y-BOCS severity. We found markers in two chromosomal regions to have suggestive associations with earlier AAO (11 markers on chromosome 6 with permutated P<3.80E-06 and four markers on chromosome 10 with permutated P<4.92E-06; Table 5.3) and detected a top-hit marker, rs366476, on chromosome 14 for Y-BOCS severity score (permutated P=4.04E-06; Table 5.4). The first region of interest for AAO resides in an intergenic region on chromosome 6p21.3 (Figure 5.12). The closest genes are the immediate early response 3 (IER3) and the discoidin domain receptor tyrosine kinase 1 (DDR1) genes (Figure 5.12). The second region for AAO is mapped to chromosome 10q25.3 where the SNPs reside within the actin-binding LIM protein family, member 1 (ABLIM1) gene (Figure 5.13). The top-hit marker for increased Y-BOCS severity score, rs366476, resides within the protein-tyrosine phosphatase, nonreceptor-type, 21 (PTPN21) gene on chromosome 14q31.3 (Figure 5.14). The second identified SNP rs1328906 for association with OCD severity is located on chromosome 9, which lies close to the multiple PDZ domain protein gene (MPDZ) (Figure 5.15).

193

194

Table 5.3. Top Hits of Age At Onset (AAO) in OCD.

SNP CHR EMP1 NP BP A1 TEST NIND BETA STAT RAW_P rs4665011 2 4.70E-06 10000000 159377004 A TOT 158 0.1349 4.76 4.38E-06 rs2671762 3 4.40E-06 10000000 5110111 T TOT 159 -0.2282 -4.695 5.76E-06 rs12190167 6 4.10E-06 10000000 30874173 A TOT 156 -0.2138 -4.947 1.95E-06 rs12197154 6 9.00E-06 10000000 30874223 C TOT 158 -0.2071 -4.793 3.80E-06 rs4713376 6 7.40E-06 10000000 30881293 C TOT 158 -0.2007 -4.811 3.52E-06 rs4327730 6 7.20E-06 10000000 30888915 T TOT 158 -0.2155 -4.834 3.18E-06 rs12210318 6 7.20E-06 10000000 30889883 A TOT 158 -0.2155 -4.834 3.18E-06 rs12192760 6 7.20E-06 10000000 30893364 C TOT 158 -0.2155 -4.834 3.18E-06 rs16897921 6 5.90E-06 10000000 30897505 A TOT 157 -0.2024 -4.847 3.02E-06 rs12195964 6 7.40E-06 10000000 30898460 G TOT 158 -0.2007 -4.811 3.52E-06 rs4713391 6 7.40E-06 10000000 30900214 C TOT 158 -0.2007 -4.811 3.52E-06 rs12192704 6 7.30E-06 10000000 30900249 A TOT 158 -0.2007 -4.811 3.52E-06 rs12198723 6 7.20E-06 10000000 30904713 G TOT 158 -0.2155 -4.834 3.18E-06 rs2183918 9 8.60E-06 10000000 28010170 G TOT 158 -0.1529 -4.587 9.20E-06 rs7913199 10 2.60E-06 10000000 116284487 T TOT 159 -0.1606 -4.81 3.51E-06 rs2483578 10 3.10E-06 10000000 116336214 C TOT 159 -0.1641 -4.839 3.09E-06 rs2483583 10 3.70E-06 10000000 116344186 G TOT 159 -0.1595 -4.732 4.92E-06 rs2497663 10 3.70E-06 10000000 116346418 A TOT 159 -0.1595 -4.732 4.92E-06 rs1262781 13 1.65E-05 10000000 49964172 T TOT 158 0.1352 4.649 7.07E-06 rs1753637 13 1.65E-05 10000000 49982174 T TOT 158 0.1352 4.649 7.07E-06 rs10502895 18 6.50E-06 10000000 44485436 A TOT 159 -0.1989 -4.698 5.70E-06 This table illustrates the top hit genetic markers for the results of AAO in OCD.

194

195

Table 5.4. Top Hits of Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score in OCD.

SNP CHR EMP1 NP BP A1 TEST NIND BETA STAT RAW_P rs1947571 2 3.23E-05 10000000 1.94E+08 A TOT 200 3.712 4.337 2.30E-05 rs10932669 2 9.53E-05 10000000 2.17E+08 T TOT 200 -5.708 -4.347 2.21E-05 rs9341145 2 7.17E-05 10000000 2.17E+08 T TOT 200 -7.165 -4.379 1.93E-05 rs1516338 3 4.18E-05 10000000 211759 C TOT 199 4.166 4.397 1.79E-05 rs11706917 3 3.31E-05 10000000 72437744 A TOT 200 -4.81 -4.535 9.97E-06 rs2210968 6 2.82E-05 10000000 1.02E+08 A TOT 200 4.228 4.405 1.73E-05 rs1328906 9 1.29E-05 10000000 13091241 T TOT 199 4.665 4.687 5.16E-06 rs10830004 10 4.09E-05 10000000 1.33E+08 A TOT 200 -3.796 -4.37 2.00E-05 rs10734664 12 0.00015 10000000 17159827 T TOT 200 -15.84 -4.399 1.77E-05 rs33219 12 2.07E-05 10000000 30960265 T TOT 200 -7.024 -4.443 1.47E-05 rs2145811 13 5.04E-05 10000000 40008295 T TOT 200 -3.708 -4.414 1.67E-05 rs366476 14 7.1E-06 10000000 88004404 A TOT 200 4.165 4.742 4.04E-06 rs1008712 16 3.53E-05 10000000 83888530 T TOT 200 -5.525 -4.422 1.61E-05 rs9890602 17 1.93E-05 10000000 27904495 T TOT 199 -5.88 -4.339 2.28E-05 rs235360 21 8.68E-05 10000000 45019671 A TOT 200 -4.194 -4.348 2.20E+00 This table illustrates the top hit genetic markers for the results of Y-BOCS severity score in OCD.

195

196

Figure 5.4. Quantile-Quantile Plot for Age At Onset (AAO) – Left (Initial QQ Plot) and Right (QQ Plot After Permutation). The genomic inflation value (lambda) is 1.003295.

196

197

Figure 5.5. Quantile-Quantile Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score – Left (Initial QQ Plot) and Right (QQ Plot After Permutation). The genomic inflation value (lambda) is 1.048644.

197

198

Table 5.5. Population Description for the MDS Plots.

Population Code Description CEU Utah Residents (CEPH) with Northern and Western European Ancestry JPT Japanese in Tokyo, Japan YRI Yoruba in Ibadan, Nigeria ASW Americans of African Ancestry in SW USA CHB Han Chinese in Bejing, China CHD Chinese in Metropolitan Denver, Colorado GIH Gujarati Indians from Houston, Texas LWK Luhya in Webuye, Kenya MEX Mexican Ancestry MKK Maasai in Kinyawa, Kenya TSI Toscani in Italia This table illustrates population description for the MDS plots.

198

199

Population:  Sample  CEU  JPT  YRI  ASW  CHB  CHD  GIH  LWK  MEX  MKK  TSI

Figure 5.6. Multi-Dimensional Scale (MDS) Plot – Initial. This graph illustrates the MDS plot of the OCD sample for population stratification.

199

200

Removed

Population:  Sample  CEU  JPT  YRI  ASW  CHB  CHD  GIH  LWK  MEX  MKK  TSI

Figure 5.7. Multi-Dimensional Scale (MDS) Plot – Removal 1. This graph illustrates the MDS plot after the first removal of population outliers of the OCD sample.

200

201

Removed

Removed

Figure 5.8. Multi-Dimensional Scale (MDS) Plot – Removal 2. This graph illustrates the MDS plot after the second removal of population outliers of the OCD sample.

201

202

Figure 5.9. Multi-Dimensional Scale (MDS) Plot – Final. This graph illustrates the final MDS plot after removal of population outliers of the OCD sample.

202

203

Figure 5.10. Manhattan Plot for Age At Onset (AAO) – Left (Initial Manhattan Plot) and Right (After Permutation).

203

204

Figure 5.11. Manhattan Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score – Left (Initial Manhattan Plot) and Right (After Permutation).

204

205

Figure 5.12. LocusZoom Plot for Age At Onset (AAO) Result on Chromosome 6. This graph illustrates the LocusZoom plot for the AAO results of OCD on chromosome 6.

205

206

Figure 5.13. LocusZoom Plot for Age At Onset (AAO) Result on Chromosome 10. This graph illustrates the LocusZoom plot for the AAO results of OCD on chromosome 10.

206

207

Figure 5.14. LocusZoom Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Result on Chromosome 14. This graph illustrates the LocusZoom plot for the Y-BOCS results of OCD on chromosome 14.

207

208

Figure 5.15. LocusZoom Plot for Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score Result on Chromosome 9. This graph illustrates the LocusZoom plot for the Y-BOCS results of OCD on chromosome 9.

208

209

5.5 Discussion

This is the first GWAS to examine the role of genetics in OCD in the context of AAO and Y-BOCS severity. Although no genome-wide significant results were detected for AAO or Y-BOCS severity, interesting findings were noted.

The strongest associations for earlier AAO came from SNPs on chromosome 6p21.3, which is close to the human leukocyte antigen (HLA) region. The closest upstream gene to the SNP rs12190167, IER3, has previously been implicated in human diseases such as cancer, autoimmune diseases, inflammatory diseases, and hypertension (Wu et al., 2013) and has potential roles in protecting cells from Fas- or tumor necrosis factor type alpha-induced apoptosis (Wu et al., 1998). It is also involved in several signaling pathways, in particular the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK), and phosphoinositide 3 kinase (PI3K)/ protein kinase B (also known as Akt) (Arlt and Schäfer, 2011). DDR1, which is the closest downstream gene to rs12190167, encodes for receptor tyrosine kinase that play a key role in the regulation of cell growth, differentiation, and metabolism, in addition to its interaction with the G-protein-coupled receptor (GPCR) and Akt signaling pathways (Borza and Pozzi, 2014). This gene has previously been postulated in the development and progression of various diseases including cancer, inflammatory process in atherosclerosis, lung and liver fibrosis, renal injury, and osteoarthritis (Borza and Pozzi, 2014). Furthermore, this gene has been implicated as a novel susceptibility gene for schizophrenia (Roig et al., 2007). One common theme across these genes is their inflammatory and immunological properties, which have been hypothesized in the etiology of at least a subgroup of OCD, the pediatric autoimmune neuropsychiatric disorders associated with [(group A β-hemolytic) streptococcal (GABHS)] infection (PANDAS) (Snider and Swedo, 2003). Individuals with PANDAS-related subtype of OCD have earlier AAO (between three years of age and the onset of puberty) (Snider and Swedo, 2003) and OCD patients have been found to have changes in immune parameters such as interleukin-1β (IL-1β) (Gray and Bloch, 2012). Furthermore, significant genetic associations have been reported in genes near the HLA region including the myelin oligodendrocyte glycoprotein (MOG) gene (Zai et al., 2004) and the tumor necrosis factor-alpha (TNF-α) gene (Cappi et al., 2012). Therefore, it

209

210

may be important to identify this subgroup of OCD for a more in-depth examination of immunological genes as genetic risk interacting with an insult (i.e., GABHS infection) for the development of OCD.

The second region on chromosome 10q25.3 lies within ABLIM1, and its encoded LIM domain proteins play important roles in the regulation of numerous biological processes such as embryonic development, cell lineage determination, and cancer differentiation (Krupp et al., 2006). Moreover, deletion of this genomic region has been shown to be associated with various types of cancer (Krupp et al., 2006). However, this gene has not been shown to have any connection with psychiatric disorders to date. Thus, future study is warranted to explore its functional significance in OCD.

The strongest association with Y-BOCS severity score came from a marker located within PTPN21, which has previously been associated with schizophrenia in a GWAS (Chen et al., 2011). PTPN21 encodes for a protein that influences neuronal growth and survival in addition to its role in interacting with erb-b2 receptor tyrosine kinase 4 (ERBB4) and neuregulin 3 (NRG3) signaling pathways, which are key players in neuronal plasticity (Plani-Lam et al., 2015). Neuroplasticity plays a key role in neurodevelopment, repair, learning and memory. Disturbance of this process has been repeated postulated to contribute to psychiatric disorders including OCD (Pittenger, 2013). The second top SNP for Y-BOCS severity is a marker that lies close to MPDZ, which has previously been reported to be associated with hydrocephalus (Al- Dosari et al., 2013) and studies have shown that MPDZ interacts with HTR2A (Jones et al., 2009), HTR2C (Ullmer et al., 1998), and more importantly NMDA receptor, which plays a vital role in facilitating and stabilizing signaling (Elias and Nicoll, 2007; Krapivinsky et al., 2004). This again points to the potential role of the consistently implicated glutamatergic and serotonergic systems in influencing severity of OCD.

All of the above mentioned genes appear to have pleiotropic effects on multiple systems, especially in neuroplasticity and cell development. It is therefore vital to examine these genes and to identify common pathways, leading to the development of an earlier onset and/or a more severe form of OCD. This will hopefully promote a more aggressive approach in the treatment of OCD in order to reduce patients’ suffering.

210

211

Small sample size is one of the limitations of this study. Using the linear regression model on AAO and Y-BOCS symptom severity allow for optimal power in this small sample of OCD participants. However, this sample was only able to capture significant results at a relatively high variance of 18% and while considering that OCD is a multifactorial disease, lower variance will increase the chance to detect additional small gene effects. Larger sample size will allow for further stratification of AAO, separating into early, intermediate, and late onset in addition to using the Y-BOCS symptom checklist to examine different symptom dimensions. However, given the current study’s small sample size, there is not enough power to detect a significant difference between the stratified groups according to AAO and Y-BOCS symptom dimensions. The second limitation is the lower SNP coverage of the Illumina Human610-Quadv1_B SNP BeadChip array (Illumina Inc. ®, San Diego, CA, USA) when compared to the recently developed PsychChip array; thus, genotyping using an array with higher quality and finer mapping of the genome will enable us to detect additional SNPs, which are not covered in this study’s array.

In summary, there are many factors and complexities associated with the diagnosis of OCD given its heterogeneous nature. Future studies will require larger, well-phenotyped samples, and more sophisticated analytical strategies (i.e., incorporating GABHS infection status for gene-environment interaction) in order to quantify and characterize the genetic susceptibility for earlier onset and more severe subgroups of OCD.

211

212

Chapter 6 PHARMACOGENETICS OF OCD

6 Multi-Gene Pharmacogenetic Study of Antidepresant Response in Obsessive-Compulsive Disorder

Gwyneth Zai 1,2,3,4, Clement C. Zai 1,2, Vanessa F. Gonçalves 1,2, Karen Wigg 4, James L. Kennedy 1,2,3*, Margaret A. Richter 2,3,4*

1 Neurogenetics Section, Centre for Addiction and Mental Health, Toronto, ON M5T 1R8, Canada

2 Department of Psychiatry, University of Toronto, Toronto, ON M5T 1R8, Canada

3 Institute of Medical Science, University of Toronto, Toronto M5S 1A8, Canada

4 The Frederick W. Thompson Anxiety Disorders Centre, Department of Psychiatry, Sunnybrook Health Sciences Centre, Toronto, ON M4N 3M5, Canada

* These authors are co-senior authors on this manuscript.

* Corresponding author: Dr. James L. Kennedy, Director of the Neuroscience Department and

Head of Neurogenetics Section at the Centre for Addiction and Mental Health (CAMH),

Professor of the Department of Psychiatry and the Institute of Medical Science at University of

Toronto, Toronto, ON M5T 1R8, Canada (Phone: 1-416-979-4987; Fax: 1-416-979-4666; E- mail: [email protected])

212

213

6.1 Abstract

Limited number of pharmacogenetic studies in search of genetic variations to predict antidepressant response in obsessive-compulsive disorder (OCD) has been published. Most of these published studies have examined known single nucleotide polymorphisms (SNPs) within OCD candidate genes. Recent advances from the ENCyclopedia Of DNA Elements (ENCODE) project has expanded our knowledge in the human genome, which consists of at least 80% functional regulatory elements. Therefore, in this study, we investigated SNPs in the remote regulatory regions of 14 OCD candidate genes in 217 Caucasian OCD patients in addition to examining candidate SNPs within 10 genes in 117 OCD patients with antidepressant response data. For the remote regulatory region SNPs, we observed significant association between the serotonin 2A receptor rs7997012 (P=0.047), serotonin 1B receptor rs1778258 (P=0.005-0.010), fas apoptotic inhibitory molecule 2 rs706795 (P=0.017-0.045) with antidepressant response. Additional robust significant associations were detected between: SLC1A1 and citalopram, sertraline, and fluvoxamine response, C9orf68 and clomipramine and fluvoxamine response, GRIN2B and clomipramine, citalopram, fluoxetine, paroxetine, and fluvoxamine response, and COMT and citalopram, sertraline, and paroxetine response. However, given the large number of examined SNPs, none of these results survived testing for multiple comparisons. This study provides further support for a role of genetic influence in predicting antidepressant response in OCD.

6.2 Introduction

Pharmacogenetics in psychiatry has received increasing attention over the past several years given that only approximately 50% of patients respond to psychotropic medications. Urgent advances to explore factors that predict medication response are required to decrease health care cost and duration of suffering from failed medication trials.

OCD is chronic and severe psychiatric disorder that is characterized by obsessions and compulsions (APA, 2013). This illness affects approximately 2.5% of the general population (Ruscio et al., 2010). The first-line pharmacological treatment of this disease is the

213

214

antidepressant class, selective serotonin reuptake inhibitors (SSRIs) (CPA, 2006; Fineberg et al., 2015; Katzman et al., 2014), but between 40% and 60% of patients with OCD do not respond to these medications (McDougle et al., 1993), and only 32% to 65% of SSRI-treated OCD patients showed clinical improvement (Fineberg et al., 2005). The duration of these medication trials may take up to 12 weeks prior to switching to another antidepressant. Thus, both patients and physicians are choosing psychotropic medications to treat psychiatric disorders in a trial-and- error matter without any guidance. Emerging research in pharmacogenetics may guide better prediction of medication response and tolerability, which may in turn enable physicians to choose the best antidepressant for a particular patient. Ultimately, this may decrease the patients’ duration of suffering and lower health care cost by reducing the number of medication trials.

Evidence confirmed a strong genetic etiology in this illness with a heritability ranging from 26% to 65% (van Grootheest et al., 2005}. The identification of genetic susceptibility risk to OCD has been difficult due to the complexity and heterogeneity of this disorder. The pathogenesis of this illness is likely a combination of epistasis, epigenetic mechanisms, and environmental insults. Therefore, the research focus has been targeted to pharmacogenetics in determining biomarkers to predict psychotropic medication response and tolerability. Many factors affect drug response and genetic variations have been shown to influence drug metabolism (Caley, 2011) and mechanisms of drug action (Gvozdic et al., 2012). Given that there has been limited number of pharmacogenetic studies in OCD (Zai et al., 2014), researchers predict that there will be overlapping genetic association between OCD diagnosis and antidepressant response.

Antidepressant pharmacogenetics has been well-established in major depressive disorder (MDD) but not in anxiety disorders or obsessive-compulsive and related disorders (OCRDs) (Tiwari et al., 2009). No pharmacogenetic studies have been published in obsessive-compulsive and related disorders except for obsessive-compulsive disorder (OCD) (Table 6.1) (Zai et al., 2014).

214

215

Table 6.1. Brief Summary of Pharmacogenetic Studies of OCD.

Gene Chromosome PharmacogeneticsReferences Brandl et al., 2013; Müller et al., 2012; Van Nieuwerburgh et CYP2D6 22q13.1 +/- al., 2009 CYP2C19 10q24 +/- Müller et al., 2012 Corregiari et al., 2012; Denys et al., 2007; Miguita et al., HTR1B 6p14.1 - 2011 SLC1A1 9p24.2 + Real et al., 2010 BDNF 11p14.1 + Real et al., 2009 - Miguita et al., 2011; Tot et al., 2003 HTR2A 13q14.2 + Corregiari et al., 2013; Denys et al., 2007; Zhang et al., 2004 + McDougle et al., 1998; Denys et al., 2007 SLC6A4 17q11.2 Billett et al., 1997; Di Bella et al., 2002; Miguita et al., 2011; - Zhang et al., 2004 + Liu et al., 2011; Michaelovsky et al., 2008 COMT 22p11.21 - Miguita et al., 2011; Zhang et al., 2004 MAO-A Xp11.3 - Zhang et al., 2004 + indicates significant study - indicates negative study +/- indicates mixed results

Previous pharmacogenetic study in OCD reported several modest findings in known genetic variants across the cytochrome P450 liver enzymes, glutamatergic and serotonergic system genes (Table 6.1) (Zai et al., 2014). The ENCyclopedia Of DNA Elements (ENCODE) project reported 80.4% of human genome displaying some functionality into gene regulation, gene-gene, and gene-environment interaction (Kavanagh et al., 2013; Nair and Howard, 2013), which provided a strong evidence of support to examine genetic variations across these remote regulatory regions. The identification of these novel regulatory elements has provided researchers new insights into the functional implications of known genes.

Genetic associations between OCD and several candidate genes including the glutamatergic system genes [glutamate transporter (SLC1A1), glutamate NMDA receptor 2B (GRIN2B)], serotonergic system genes [serotonin 1B receptor (5HT1B), serotonin 2A receptor (HTR2A), serotonin transporter (SLC6A4)], dopaminergic system genes [catecholamine-O- methyl-transferase (COMT), monoamine oxidase (MAOA)], myelin-related genes [myelin oligodendrocyte glycoprotein (MOG), oligodendrocyte lineage transcription factor 2 (OLIG2)] and brain-derived neurotrophic factor (BDNF), have been reported with promising but inconsistent results (Zai et al., 2015). We also included additional candidates in the glutamatergic neurotransmitter system, the disks large (drosophila) homolog-associated protein 1

215

216

(DLGAP1) and DLGAP2 genes. The following is a list of the GWAS top hits that were examined in this study: glutamate receptor ionotropic kainite 2 (GRIK2), fas apoptotic inhibitory molecule 2 (FAIM2), SLIT and NTRK-like family member 5 (SLITRK5), discs large (drosophila) homolog-associated protein 1 (DLGAP1), fucosyl-transferase 2 (FUT2), and BTB (POZ) domain containing 3 (BTBD3).

The purpose of this study is to examine genetic variations within these 17 candidate genes (Table 1.12, Table 6.2, and Table 6.3) and antidepressant response in OCD patients. Candidate genes were chosen based on their previously reported significant genetic association with OCD diagnosis including the top hits genes from the first and recent genome-wide association study (GWAS) of OCD (Stewart et al., 2013).

Table 6.2. Investigated Genes with Their SNPs for the Single Taqman Assay Protocol.

Gene Chromosome SNP MOG 6p22.1 rs2252711, rs2071653 HTR1B 6q14.1 rs2000292, rs6297, rs6296, rs130058, rs11568817, rs1213371 DLGAP2 8p23.3 rs14103, rs4653107, rs4653108, rs4653109, rs11583978, rs1001616, rs12727066, rs11587343, rs4653112, rs7541937, rs11264126, rs12141243, rs35688758, rs7525948, rs11264155, rs6662980, rs2184187, rs4417025, rs4259608, rs1931059, rs4652867, rs11264172, rs6686484, rs11264173, rs10493064, rs12120523, rs7555884, rs16837122, rs4652869, rs6699355 SLC1A1 9p24.2 rs1360329, rs7031998, rs7045401, rs9918970, rs10814995, rs1980943, rs7022369, rs6476875, rs2026828, rs3780415, rs10815020, rs10974625, rs2228622, rs3780413, rs3780412, rs12682807 GRIN2B 12p13.1 rs1805502, rs1805476, rs890, rs1805246, rs1805247, rs1806191, rs1806201, rs1120905, rs7301328, rs1019385 HTR2A 13q14.2 rs3125, rs6314, rs1923882, rs7997012, rs12584920, rs1328684, rs2296973, rs2070037, rs2070039, rs2070040, rs9534511, rs1033847, rs6312, rs6306, rs1328685, rs731245 SLC6A4 17q11.2 rs1042173, rs140701, rs140700, rs6354, rs2066713, rs8071667, rs16965628, rs2020934, rs2020933, rs25533 OLIG2 21q22.11 rs1122807, rs762178, rs9982080, rs1059004, rs6517136, rs6517137, rs13046814, rs7278343, rs9653711, rs881666 COMT 22q11.21 rs737866, rs933271, rs1544325, rs5993883, rs740603, rs2239393, rs4680, rs4646316, rs165774, rs9332377 MAOA Xp11.3 rs1465107, rs1465108, rs2179098, rs6323, rs979606, rs979605 Highlighted in yellow indicates replicated SNP from the QuantStudio array protocol

216

217

Table 6.3. Genes and SNPs for the QuantStudio Array Protocol.

Gene Chromosome SNP HTR1B 6q14.1 rs1778258 GRIK2 6q16.3 rs1556995 SLC1A1 9p24.2 rs3933331, rs7031998, rs7022369 BDNF 11p14 rs3763965, rs7124442, rs61888800, rs11030119, rs2883187 GRIN2B 12p13.1 rs1805482*, rs7301328 FAIM2 12q13.12 rs1044677, rs7132908, rs706795 HTR2A 13q14.2 rs7997012, rs9534510 SLITRK5 13q31.2 rs10450811, rs9557425 SLC6A4 17q11.2 rs3813034 DLGAP1 18p11.31 rs2240899, rs8096794 FUT2 19q13.33 rs681343 BTBD3 20p12.2 rs1996132 COMT 22q11.21 rs9617850, rs737865*, rs933271, rs4818 MAOA Xp11.3 rs3788862, rs1465107, rs979605, rs1137070 * Highlighted in grey indicates failed SNP in the QuantStudio 32-SNP Chip Highlighted in yellow indicates replicated SNP from the Single Taqman Assay Protocol Highlighted in blue indicates SNPs within exon and not within remote regulatory regions

6.3 Methods 6.3.1 Diagnostic Criteria and Sample

Five hundred and sixty OCD patients and 273 family members were recruited from consecutive referrals to the Anxiety Disorders Clinic at the Centre for Addiction and Mental Health in Toronto, Canada and the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre in Toronto, Canada. All research participants were assessed using the Structured Clinical Interview for the DSM-IV (SCID) (First et al., 1996). The Yale- Brown Obsessive-Compulsive Scale (Y-BOCS) was completed for each participant to determine the severity of OCD symptoms (Goodman et al., 1989). Lifetime severity of OCD symptoms was estimated by re-administering the Y-BOCS scale while individuals were asked to focus retrospectively on the time period when the most severe OCD symptoms were experienced. All assessments and interviews were performed by trained research assistants who were supervised by an experienced psychiatrist. All participants met DSM-IV criteria for the diagnosis of OCD as determined by the SCID interview and available medical records. All subjects provided their written informed consent to participate in this study, and ethics approval was obtained from the

217

218

local Research Ethics Board. Exclusion criteria included any metabolic or chronic neurological disease (other than tic disorder), active substance use disorder, schizophrenia, or schizoaffective disorder.

6.3.2 Response Data

Two hundred and twenty-two of the 560 OCD patients have antidepressant response data. Of the entire sample, 222 OCD participants have complete drug response data for analysis. Six SRI antidepressants including five SSRIs (fluoxetine, fluvoxamine, sertraline, paroxetine, citalopram) and clomipramine were examined for an association of response or non-response. The drug response data was collected retrospectively by using a questionnaire (Appendix III), which was developed by Drs. Kennedy and Richter, in addition to chart review. Response to antidepressant medications was scored according to the Clinical Global Impression – Improvement (CGI-I) scale. Individuals were coded as “responders” if they rated “much improved” or “very much improved” for at least one antidepressant and individuals were defined as “non-responders” if they rated “minimal improvement”, “no change”, or “worse” for at least two different antidepressants. Complete response data must include the following: adequacy of trials (based on widely accepted criteria regarding optimal dose and duration), self-rated estimates of reliability of recollection, and response according to CGI-I scoring system.

The specific criteria for a complete and analyzable antidepressant treatment trial included the following:

1) Adequate duration of trial of at least 10 weeks

2) Adequate dosage of SRI (maximum dose achieved of: fluoxetine ≥20mg, fluvoxamine ≥150mg, paroxetine ≥20mg, sertraline ≥100mg, citalopram ≥20mg, or clomipramine ≥150mg)

3) Participants rated their recall of the drug trail as good or excellent

218

219

4) Individuals with intolerable side effects were excluded from the antidepressant response analysis given tolerability and response are different measures and may have different genetic pathway

5) At least two failed trials of SRIs for coding as non-responder

6.3.3 Choosing of SNPs and Genotyping

Venous blood was obtained from the participants in two 10cc EDTA tubes, and genomic DNA was extracted from blood lymphocytes using a high salt method (Lahiri and Nurnberger, 1991). Genotyping was completed with two different methodologies, Taqman assay for each SNP and QuantStudio array (Life Technologies ®, Burlington, ON).

6.3.3.1 Single Taqman Assay

One hundred and sixteen SNPs across 10 candidate genes were investigated. The examined genes (Table 6.2) include: 2 SNPs across myelin oligodendrocyte glycoprotein (MOG), 6 SNPs across serotonin 1D beta receptor (HTR1B), 30 SNPs across discs large (drosophila) homolog-associated protein 2 (DLGAP2), 16 SNPs across neuronal glutamate transporter (SLC1A1), 10 SNPs across glutamate receptor ionotropic N-methyl D-aspartate 2B (GRIN2B), 16 SNPs across serotonin 2A receptor (HTR2A), 10 SNPs across serotonin transporter (SLC6A4), 10 SNPs across oligodendrocyte lineage transcription factor 2 (OLIG2), 10 SNPs across catechol-O-methyltransferase (COMT), and 6 SNPs across monoamine oxidase A (MAOA). These candidate genes were chosen given their evidence of genetic association with OCD susceptibility.

Part of the genotyping was conducted by the International Obsessive-Compulsive Disorder Foundation Genetic Collaborative Group and the genotyping protocol has been described elsewhere (Stewart et al., 2013). Additional genotyping was performed using the Taqman assay for allelic discrimination with the Avant ABI Prism ® 7500 Genetic Analyzer (Applied Biosystems, Foster City) for MOG (Zai et al., 2004), HTR1B (Mundo et al., 2002),

219

220

SLC1A1 (Arnold et al., 2006), GRIN2B (Arnold et al., 2004), and OLIG2 (Zai et al., 2012) and the genotyping protocol was previously reported in other studies.

6.3.3.2 QuantStudio Array

Thirty-two SNPs, of which 27 reside in the remote regulatory regions, across 14 candidate genes, were investigated. A 32-SNP QuantStudio Chip (Life Technologies ®, Burlington, ON) was designed. The chip consisted of regulatory SNPs across 14 different candidate genes including genes from previous significant genetic association studies and recent top hits from the first and recently published GWAS (Stewart et al., 2013). SNPs were chosen based on potential functional status using different publicly available search engines including functional significance (FS) score, European minor allele frequencies (EurMAF), BrainCloud (Colantuoni et al., 2011), UCSC (Rosenbloom et al., 2013), HaploReg (Ward et al., 2012), RegulomeDB (Boyle et al., 2012), and National Institute of Environmental Health Sciences (NIEHS) (Xu et al., 2009).

Thirty-two SNPs across the remote regulatory regions in 14 candidate genes, serotonin 1B receptor (HTR1B), glutamate receptor ionotropic kainite 2 (GRIK2), neuronal glutamate transporter (SLC1A1), brain-derived neurotrophic factor (BDNF), glutamate receptor ionotropic N-methyl D-aspartate 2B (GRIN2B), fas apoptotic inhibitory molecule 2 (FAIM2), serotonin 2A receptor (HTR2A), SLIT and NTRK-like family member 5 (SLITRK5), serotonin transporter (SLC6A4), discs large (drosophila) homolog-associated protein 1 (DLGAP1), fucosyl-transferase 2 (FUT2), BTB (POZ) domain containing 3 (BTBD3), catechol-O-methyltransferase (COMT), and monoamine oxidase A (MAOA), are shown in Table 6.3.

Genotyping was performed using the QuantStudio technology (Life Technologies ®, Burlington, ON). All genotypes were determined with the Life Technology ® allelic discrimination software and confirmed by two experienced researchers. Approximately 10% of the genotypes were confirmed by repeating the experiment and using positive and negative controls.

Quality control was performed using PLINK (Purcell et al., 2007). Samples with <98%

220

221

completion and SNP assays with call rates <90% were excluded. Only two SNPs, GRIN2B rs1805482 and COMT rs737865, were excluded from the final analysis.

6.3.4 Statistical Analyses Genotype frequency of each of the tested SNP did not deviate from Hardy-Weinberg Equilibrium test (P>0.1). We tested for the association between the genotypes of the examined SNPs and antidepressant response status (responders versus non-responders) using Pearson chi- squared test (SPSS version 20.0, Armonk, NY, USA). Analysis was completed for each of the six antidepressants for the single Taqman assay protocol in addition to combining them into either SSRI(s) or SRI(s) in order to improve power for the QuantStudio protocol.

Results were assumed to be significant initially if P<0.05 in all cases and P<6.33E-4 for 79 independent markers for the single Taqman assay protocol and P<0.002 (26 independent tests due to 3 for the 5 markers across BDNF and 2 for 4 markers across MAOA) for the QuantStudio protocol using the Nyholt method (Nyholt, 2004; Li et al., 2005). Therefore, the results were deemed to be significant if P value is less than 5.05E-4 in all cases after corrected for multiple comparisons using the Nyholt method for 99 SNPs (Single Taqman Assay 79 + QuantStudio Array 26 – 6 replicated SNPs).

Our sample has enough power (80.5%) to detect 8.3% of the variance of SSRI/SRI response that is accounted for by the SNP (α=5.05E-4, additive model) using the QUANTO program (version 1.2.4) (Gauderman and Morrison, 2006).

6.4 Results

Demographics data of the OCD sample is presented in Table 6.4, Table 6.5, and Table 6.6.

221

222

Table 6.4. Demographic Data of the Entire OCD Sample.

Descriptors N/Mean (S.D.)

N 497 Age 35 (12.5) Gender 57% female 97% Caucasian Ethnicity 3% Asian & African American Age at Onset (N=328) 14.5 (8.9) Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score - Lifetime 27.0 (7.0) (N=388) This table illustrates the demographics of the entire OCD sample.

Table 6.5. Subject Demographics for the Single Taqman Assay Protocol.

Descriptors Mean (S.D.)

N 108 Age 37.2 (12.01) Gender 56.4% female Ethnicity 100% Caucasian Age at Onset 13.5 (7.9) [N=79] YBOCS Severity Score (Lifetime) 27.0 (6.8) [N=91] Fluoxetine (number of subjects) 27 Fluvoxamine (number of subjects) 18 Sertraline (number of subjects) 23 Paroxetine (number of subjects) 23 Citalopram (number of subjects) 17 Clomipramine (number of subjects) 26 This table illustrates the demographics of the OCD subjects in the single Taqman Assay protocol.

222

223

Table 6.6. Final Demographic Data for the QuantStudio OCD Pharmacogenetic Sample.

Descriptors N/Mean (S.D.) N 217 Age (N=191) 37.6 (11.9) Gender (N=197) 57.9% female Ethnicity 100% Caucasian Age of onset (N=179) 14.6 (9.2) Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score (N=193) 28.1 (6.5) Fluoxetine (number of subjects) 52 Fluvoxamine (number of subjects) 26 Sertraline (number of subjects) 52 Paroxetine (number of subjects) 49 Citalopram (number of subjects) 41 Clomipramine (number of subjects) 23 This table illustrates the demographics of the OCD subjects in the QuantStudio pharmacogenetics sample.

For the single Taqman assay study, multiple markers were significantly associated with response to different SRI(s). Significant results (Table 6.7) include: HTR1B rs6297 with fluvoxamine (P=0.047) and citalopram (P=0.018) and any SSRI(s) response (P=0.048), HTR1B rs130058 with clomipramine (P=0.029) and any SRI(s) response (P=0.007), four markers (rs14103 and P=0.045; rs12727066 and P=0.034; rs1931059 and P=0.025; rs6686484 and P=0.010) within DLGAP2 with sertraline response, DLGAP2 rs11264155 with clomipramine response (P=0.031), DLGAP2 rs11264173 with fluoxetine (P=0.040) and any SSRI(s) response (P=0.045), SLC1A1 rs7031998 with paroxetine response (P=0.008), SLC1A1 rs7045401 with sertraline response (P=0.010), SLC1A1 rs10814995 with fluvoxamine response (P=0.018), GRIN2B rs1805247 with any SRI(s) response (P=0.036), GRIN2B rs1806191 with citalopram response (P=0.007), GRIN2B rs1806201 with fluvoxamine response (P=0.024), HTR2A rs6314 with citalopram response (P=0.032), SLC6A4 rs1042173 with paroxetine response (P=0.036), SLC6A4 rs16965628 with paroxetine (P=0.044) and clomipramine response (P=0.036), OLIG2 rs762178 with any SSRI(s) (P=0.044) and any SRI(s) response (P=0.038), OLIG2 rs13046814 with sertraline response (P=0.014), OLIG2 rs9653711 with any SRI(s) response (P=0.048),

223

224

COMT rs1544325 with sertraline (P=0.032) and citalopram response (P=0.032), COMT rs5993883 with paroxetine response (P=0.036), COMT rs740603 with paroxetine (P=0.015) and any SRI(s) response (P=0.027), and four markers within MAOA (rs1465107, rs1465108, rs6323, rs979606, P=0.032) with citalopram response. The most interesting results were shown in Table 6.7 for HTR1B, GRIN2B, and OLIG2. However, none of these findings survived Nyholt correction for multiple comparisons.

For the QuantStudio study, we found initial significant results (Table 6.8 and Figure 6.1) for the: HTR1B rs1778258 with any SSRI(s) (P=0.005), FAIM2 rs706795 with any SSRI(s) (P=0.017) and any SRI(s) (P=0.045), HTR2A rs7997012 and any SRI(s) (P=0.047). For the analysis by each SRI, initial significant findings were detected for: HTR1B rs1778258 (P=0.035) and FAIM2 rs1044677 (P=0.046) with citalopram response, SLC1A1 rs7031998 (P=0.031) and FAIM2 rs706795 (P=0.020) with paroxetine response, in addition to FAIM2 rs706795 (P=0.015) and SLITRK5 rs10450811 (P=0.007) with sertraline response. However, none of the results remained significant after Nyholt correction.

Six SNPs overlapped between the single Taqman and the QuantStudio studies. Sample size was increased for these markers when examined in the QuantStudio sample. Of these overlapping SNPs, the HTR2A rs7997012 shared initial interesting results, the first study with citalopram response (P=0.077) and the latter with any SRI(s) response (P=0.047). Otherwise, none of the results from the Taqman study persisted with the large QuantStudio sample.

224

225

Table 6.7. Pharmacogenetic Results for the Single Taqman Assay Protocol.

Gene/SRI Fluoxetine Paroxetine Clomipramine Sertraline Fluvoxamine Citalopram SSRIs SRIs MOG χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs2252711 3.250 0.197 0.738 0.390 0.650 0.420 0.857 0.355 0.135 0.714 3.010 0.222 1.352 0.509 1.737 0.420 rs2071653 3.175 0.204 3.877 0.144 0.438 0.803 0.057 0.811 0.950 0.622 1.497 0.221 1.338 0.512 0.200 0.905 HTR1B χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs2000292 1.715 0.424 2.489 0.288 0.730 0.694 1.251 0.535 2.443 0.295 0.413 0.814 5.734 0.057 3.384 0.184 rs6297 0.122 0.727 0.011 0.916 N/A N/A 0.748 0.387 3.958 0.047 5.615 0.018 3.923 0.048 3.411 0.065 rs6296 1.110 0.574 3.292 0.193 1.377 0.502 0.480 0.787 2.222 0.329 0.413 0.814 3.605 0.165 3.331 0.189 rs130058 2.550 0.279 1.551 0.461 7.071 0.029 0.281 0.869 2.727 0.256 2.974 0.226 3.642 0.162 9.942 0.007 rs11568817 1.589 0.452 0.482 0.786 3.442 0.179 1.034 0.596 1.061 0.588 0.932 0.628 1.517 0.468 3.042 0.218 rs1213371 0.862 0.650 0.482 0.786 3.923 0.141 1.034 0.596 1.760 0.415 2.273 0.321 0.934 0.627 2.618 0.270 DLGAP2 χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs14103 0.002 0.960 0.553 0.457 1.182 0.277 6.788 0.034 0.669 0.413 0.528 0.467 1.670 0.434 1.588 0.452 rs4653107 0.351 0.839 0.909 0.635 1.143 0.565 1.064 0.588 0.022 0.881 0.905 0.636 0.700 0.705 0.753 0.686 rs4653108 0.318 0.573 1.197 0.550 1.143 0.565 0.194 0.660 2.165 0.339 0.528 0.467 1.186 0.553 0.701 0.704 rs4653109 0.512 0.774 0.015 0.993 0.410 0.815 0.803 0.669 1.837 0.399 0.209 0.901 0.392 0.822 0.446 0.800 rs11583978 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs1001616 <0.001 1.000 0.191 0.909 3.026 0.220 1.247 0.536 0.303 0.859 0.573 0.751 1.031 0.597 2.522 0.283 rs12727066 0.338 0.561 0.553 0.457 2.229 0.135 6.788 0.034 0.669 0.413 0.528 0.467 2.016 0.365 2.282 0.320 rs11587343 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs4653112 2.879 0.237 0.011 0.916 3.467 0.063 4.461 0.107 0.669 0.413 0.528 0.467 1.034 0.596 1.599 0.450 rs7541937 0.938 0.625 1.022 0.600 0.103 0.950 0.569 0.752 0.034 0.983 0.573 0.751 1.655 0.437 0.538 0.764 rs11264126 0.173 0.917 0.753 0.686 0.886 0.642 2.081 0.353 0.900 0.638 0.573 0.751 0.999 0.607 0.410 0.815 rs12141243 1.001 0.317 0.101 0.750 1.963 0.161 3.057 0.080 0.460 0.498 0.873 0.350 0.495 0.481 1.552 0.213 rs35688758 0.830 0.362 N/A N/A N/A N/A 0.351 0.554 0.741 0.389 0.244 0.621 0.344 0.557 0.288 0.591 rs7525948 0.020 0.888 0.899 0.343 1.182 0.277 4.461 0.107 0.861 0.353 1.273 0.259 0.424 0.809 0.440 0.803 rs11264155 3.090 0.213 1.389 0.499 6.970 0.031 1.825 0.401 0.257 0.879 0.745 0.689 2.714 0.257 0.368 0.832 rs6662980 0.455 0.797 0.091 0.955 2.030 0.362 1.853 0.396 1.284 0.526 0.069 0.966 2.405 0.300 1.868 0.393 rs2184187 1.974 0.373 4.745 0.093 1.250 0.535 2.033 0.362 0.389 0.823 0.152 0.927 0.368 0.832 0.857 0.651 rs4417025 0.227 0.893 0.012 0.994 2.030 0.362 1.555 0.460 0.752 0.687 0.069 0.966 2.114 0.348 1.894 0.388 rs4259608 1.675 0.433 0.753 0.686 0.027 0.870 4.411 0.110 0.024 0.876 1.257 0.262 1.940 0.379 1.486 0.476

225

226

rs1931059 0.006 0.940 0.053 0.974 2.500 0.287 7.367 0.025 0.087 0.769 3.661 0.160 1.292 0.524 1.622 0.445 rs4652867 1.234 0.539 1.299 0.522 2.850 0.240 0.585 0.445 0.642 0.423 0.013 0.910 0.022 0.881 0.839 0.657 rs11264172 2.035 0.362 0.715 0.699 0.504 0.777 1.209 0.546 1.730 0.421 0.016 0.992 0.197 0.906 0.537 0.765 rs6686484 0.157 0.924 4.750 0.093 3.923 0.141 9.181 0.010 0.087 0.958 0.749 0.688 3.897 0.142 2.182 0.336 rs11264173 6.416 0.040 0.493 0.781 0.277 0.871 3.463 0.177 3.446 0.179 0.573 0.751 6.217 0.045 3.180 0.204 rs10493064 1.208 0.272 0.048 0.827 1.111 0.574 2.049 0.359 0.051 0.822 1.537 0.464 1.059 0.304 3.749 0.153 rs12120523 0.816 0.366 0.973 0.615 1.182 0.277 0.194 0.660 1.507 0.471 3.010 0.222 0.830 0.660 0.735 0.692 rs7555884 1.706 0.426 0.484 0.785 0.411 0.814 0.283 0.868 3.449 0.178 2.534 0.282 0.339 0.844 0.052 0.974 rs16837122 1.815 0.404 1.840 0.398 0.170 0.680 0.504 0.777 2.043 0.360 2.148 0.342 3.179 0.204 2.830 0.243 rs4652869 0.793 0.673 0.604 0.739 1.963 0.375 0.888 0.641 4.102 0.129 1.959 0.375 1.531 0.465 1.777 0.411 rs6699355 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A SLC1A1 χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs1360329 2.088 0.352 2.157 0.142 0.357 0.836 4.719 0.094 1.184 0.553 2.914 0.233 0.764 0.683 0.130 0.937 rs7031998 3.324 0.190 7.036 0.008 2.631 0.268 5.314 0.070 0.540 0.463 3.034 0.219 2.832 0.243 2.997 0.223 rs7045401 0.199 0.656 3.680 0.159 1.758 0.415 9.127 0.010 3.838 0.147 1.316 0.518 4.409 0.110 3.442 0.179 rs9918970 1.706 0.426 0.953 0.621 3.442 0.179 1.114 0.573 0.482 0.786 0.069 0.966 0.499 0.779 1.401 0.496 rs10814995 4.916 0.086 1.440 0.487 0.726 0.696 2.712 0.258 8.052 0.018 2.974 0.226 2.744 0.254 0.701 0.704 rs1980943 2.413 0.299 2.872 0.238 0.538 0.764 4.825 0.090 0.560 0.756 3.692 0.158 1.222 0.543 0.215 0.898 rs7022369 0.332 0.847 1.601 0.449 3.450 0.178 1.590 0.452 0.816 0.665 4.984 0.083 1.117 0.572 1.460 0.482 rs6476875 0.153 0.696 1.155 0.561 2.933 0.231 3.142 0.208 1.919 0.383 3.514 0.173 2.077 0.354 2.047 0.359 rs2026828 N/A N/A <0.001 1.000 N/A N/A 0.750 0.386 N/A N/A N/A N/A 0.139 0.709 0.375 0.540 rs3780415 3.506 0.173 0.431 0.806 0.786 0.675 1.773 0.412 1.818 0.403 2.730 0.255 4.767 0.092 2.130 0.345 rs10815020 0.615 0.735 1.557 0.459 0.619 0.431 0.011 0.918 0.002 0.964 0.636 0.728 0.397 0.820 0.914 0.633 rs10974625 0.911 0.634 1.204 0.548 0.680 0.409 0.562 0.454 0.059 0.809 0.636 0.728 1.316 0.518 0.351 0.839 rs2228622 2.868 0.238 1.222 0.543 0.538 0.764 0.291 0.865 2.044 0.360 0.513 0.774 0.065 0.968 0.052 0.974 rs3780413 1.273 0.529 3.618 0.164 2.156 0.340 3.882 0.144 4.961 0.084 0.076 0.963 2.301 0.316 3.780 0.151 rs3780412 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs12682807 1.140 0.286 0.451 0.502 1.128 0.288 0.349 0.555 0.014 0.906 0.064 0.800 1.343 0.246 1.360 0.243 GRIN2B χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs1805502 2.120 0.145 5.414 0.067 0.687 0.709 1.118 0.572 0.281 0.869 0.825 0.662 0.871 0.647 1.461 0.482 rs1805476 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs890 2.518 0.284 0.537 0.765 0.202 0.904 1.320 0.517 1.299 0.522 2.444 0.295 1.681 0.432 0.369 0.831 rs1805246 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A

226

227

rs1805247 1.147 0.284 0.862 0.353 1.182 0.277 3.211 0.201 1.684 0.194 0.825 0.364 1.850 0.397 6.654 0.036 rs1806191 1.424 0.491 1.755 0.416 0.380 0.827 2.211 0.331 3.977 0.137 9.900 0.007 0.722 0.697 0.389 0.823 rs1806201 0.020 0.990 1.867 0.393 1.200 0.549 1.247 0.536 7.446 0.024 3.094 0.213 0.519 0.771 1.055 0.590 rs1120905 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs7301328 2.169 0.338 0.887 0.642 3.667 0.160 2.424 0.298 1.684 0.431 1.547 0.461 0.285 0.867 2.721 0.257 rs1019385 0.658 0.720 2.593 0.274 1.333 0.513 1.221 0.543 0.920 0.631 0.393 0.822 1.178 0.555 0.667 0.717 HTR2A χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs3125 0.405 0.524 2.912 0.088 1.102 0.294 0.731 0.393 1.619 0.445 0.064 0.800 1.894 0.388 1.193 0.551 rs6314 0.816 0.366 0.101 0.750 2.248 0.619 0.465 0.495 0.022 0.881 4.618 0.032 0.095 0.757 0.004 0.947 rs1923882 0.208 0.648 2.068 0.150 1.067 0.587 3.205 0.201 1.818 0.403 4.015 0.134 3.065 0.216 2.400 0.301 rs7997012 5.136 0.077 0.929 0.629 2.618 0.270 1.853 0.396 1.102 0.576 0.093 0.955 0.421 0.810 1.651 0.438 rs12584920 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs1328684 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs2296973 1.784 0.410 2.561 0.278 0.154 0.695 1.616 0.446 0.961 0.618 2.394 0.302 0.073 0.964 0.002 0.999 rs2070037 0.004 0.952 2.510 0.285 1.529 0.216 0.970 0.616 2.973 0.226 3.385 0.184 1.291 0.524 1.557 0.459 rs2070039 1.298 0.523 2.068 0.150 1.400 0.497 0.194 0.660 0.067 0.795 0.009 0.926 1.140 0.566 1.185 0.553 rs2070040 0.963 0.618 0.795 0.672 1.377 0.502 1.040 0.595 2.941 0.230 2.826 0.243 4.885 0.087 5.476 0.065 rs9534511 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs1033847 1.271 0.260 0.011 0.916 2.167 0.141 0.349 0.555 1.818 0.178 N/A N/A 0.043 0.835 0.082 0.774 rs6312 0.002 0.960 0.011 0.916 3.391 0.066 1.933 0.164 0.194 0.660 N/A N/A 0.467 0.495 0.837 0.360 rs6306 0.020 0.888 2.959 0.228 1.040 0.308 0.349 0.555 2.828 0.243 3.094 0.213 1.324 0.516 0.725 0.696 rs1328685 0.494 0.482 0.011 0.916 2.229 0.135 1.285 0.257 0.194 0.660 0.393 0.531 0.097 0.756 1.667 0.197 rs731245 0.042 0.979 1.226 0.542 3.923 0.141 0.431 0.806 1.804 0.406 0.233 0.890 1.014 0.602 1.736 0.420 SLC6A4 χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs1042173 1.867 0.393 6.634 0.036 1.054 0.590 0.445 0.801 0.087 0.958 1.247 0.536 0.254 0.881 0.245 0.885 rs140701 1.293 0.524 3.877 0.144 1.099 0.577 0.437 0.804 1.428 0.490 0.153 0.926 1.420 0.492 1.723 0.423 rs140700 3.775 0.151 1.197 0.550 1.043 0.593 0.970 0.616 1.818 0.403 1.303 0.521 0.902 0.637 1.014 0.602 rs6354 0.167 0.920 1.867 0.393 3.161 0.206 2.392 0.302 3.449 0.178 0.476 0.788 2.159 0.340 3.808 0.149 rs2066713 1.253 0.534 0.269 0.874 3.758 0.153 1.011 0.603 0.683 0.711 0.825 0.662 1.188 0.552 4.890 0.087 rs8071667 1.556 0.459 2.116 0.347 2.674 0.263 2.392 0.302 1.886 0.390 0.393 0.822 1.050 0.592 1.956 0.376 rs16965628 0.396 0.529 4.044 0.044 4.396 0.036 2.078 0.149 0.281 0.596 0.013 0.910 2.565 0.109 3.168 0.075 rs2020934 1.992 0.369 0.456 0.796 1.140 0.566 1.117 0.572 1.428 0.490 0.489 0.783 1.872 0.392 1.652 0.438 rs2020933 0.193 0.660 0.719 0.396 3.147 0.076 1.359 0.244 0.051 0.822 0.064 0.800 0.321 0.571 0.608 0.436

227

228

rs25533 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A OLIG2 χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs1122807 N/A N/A N/A N/A 1.140 0.286 N/A N/A N/A N/A N/A N/A N/A N/A 0.357 0.550 rs762178 1.852 0.396 5.245 0.073 1.492 0.474 5.606 0.061 3.433 0.180 1.293 0.524 6.244 0.044 6.527 0.038 rs9982080 4.441 0.109 1.262 0.261 0.103 0.748 0.587 0.746 0.113 0.945 0.328 0.849 0.484 0.785 0.667 0.716 rs1059004 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs6517136 N/A N/A 0.899 0.343 1.040 0.308 N/A N/A 1.287 0.257 N/A N/A 0.407 0.523 0.716 0.398 rs6517137 1.271 0.260 2.021 0.364 1.040 0.308 0.083 0.773 0.861 0.353 0.064 0.800 0.433 0.805 0.346 0.841 rs13046814 4.350 0.114 0.929 0.629 2.250 0.325 8.538 0.014 1.635 0.442 3.514 0.173 1.837 0.399 2.076 0.354 rs7278343 N/A N/A 0.899 0.343 1.128 0.288 N/A N/A 1.173 0.279 N/A N/A 0.382 0.537 0.688 0.407 rs9653711 4.087 0.130 4.362 0.113 2.400 0.301 3.540 0.170 1.626 0.444 0.705 0.703 5.015 0.081 6.083 0.048 rs881666 0.535 0.765 1.197 0.550 0.405 0.817 0.083 0.959 0.090 0.956 3.598 0.165 0.546 0.761 0.456 0.796 COMT χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs737866 0.035 0.851 2.593 0.274 1.295 0.523 1.276 0.528 1.914 0.384 0.069 0.793 0.680 0.712 0.948 0.623 rs933271 2.331 0.312 0.169 0.919 1.188 0.552 2.354 0.308 2.828 0.243 1.329 0.514 2.711 0.258 4.520 0.104 rs1544325 0.043 0.979 2.671 0.263 1.648 0.439 6.865 0.032 0.460 0.795 6.875 0.032 5.240 0.073 4.071 0.131 rs5993883 0.116 0.944 6.643 0.036 2.156 0.340 2.574 0.276 0.434 0.805 3.262 0.196 3.215 0.200 4.931 0.085 rs740603 0.532 0.767 8.430 0.015 3.161 0.206 1.810 0.405 0.148 0.929 2.706 0.258 5.869 0.053 7.242 0.027 rs2239393 0.071 0.965 2.837 0.242 1.983 0.371 0.541 0.763 0.833 0.659 3.839 0.147 0.713 0.700 0.235 0.889 rs4680 0.079 0.961 2.436 0.296 0.933 0.627 0.431 0.806 0.471 0.790 1.761 0.415 0.057 0.972 0.180 0.914 rs4646316 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A rs165774 0.133 0.936 0.570 0.752 0.051 0.975 1.590 0.452 3.442 0.179 1.203 0.548 0.007 0.997 0.280 0.869 rs9332377 1.994 0.369 0.905 0.636 1.111 0.574 0.476 0.788 0.087 0.769 0.009 0.926 2.714 0.257 3.430 0.180 MAOA χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P rs1465107 1.752 0.416 1.867 0.393 0.933 0.627 0.815 0.665 4.983 0.083 6.875 0.032 0.319 0.853 0.492 0.782 rs1465108 0.010 0.995 1.867 0.393 0.400 0.819 0.815 0.665 3.714 0.156 6.875 0.032 0.194 0.907 0.779 0.677 rs2179098 2.214 0.137 N/A N/A N/A N/A 0.469 0.493 N/A N/A N/A N/A 0.418 0.518 0.360 0.549 rs6323 0.346 0.841 0.392 0.822 1.771 0.412 1.247 0.536 1.953 0.377 6.875 0.032 0.699 0.705 0.318 0.853 rs979606 0.227 0.893 0.392 0.822 1.771 0.412 1.247 0.536 1.953 0.377 6.875 0.032 0.516 0.773 0.509 0.775 rs979605 N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A N/A Highlighted in yellow indicates initial significant P value prior to adjusting for multiple comparisons

228

229

Table 6.8. Result of Genetic Variations across Remote Regulatory Regions in 14 OCD Candidate Genes and SRI Response.

SRI Fluoxetine Paroxetine Sertraline Fluvoxamine Citalopram Clomipramine SSRIs SRIs Gene SNP χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P χ2 P HTR1B rs1778258 0.014 0.907 1.201 0.273 2.663 0.264 2.880 0.090 4.425 0.035 1.096 0.295 7.781 0.005 6.692 0.010 GRIK2 rs1556995 1.973 0.373 4.798 0.091 2.088 0.352 1.588 0.452 0.096 0.953 0.423 0.809 5.513 0.064 1.993 0.369 SLC1A1 rs3933331 0.842 0.656 2.595 0.273 3.749 0.153 0.888 0.641 2.669 0.263 5.636 0.060 2.927 0.231 5.825 0.054 rs7031998 4.766 0.092 6.962 0.031 4.421 0.110 2.517 0.113 2.654 0.265 0.750 0.387 3.160 0.206 2.360 0.307 rs7022369 0.554 0.758 0.415 0.813 0.630 0.730 0.136 0.934 4.778 0.092 0.283 0.868 0.289 0.866 0.035 0.983 BDNF rs3763965 0.846 0.655 1.548 0.461 1.968 0.374 1.143 0.565 1.777 0.411 3.979 0.137 1.125 0.570 0.471 0.790 rs7124442 0.117 0.732 0.983 0.612 1.714 0.424 2.103 0.147 0.972 0.615 1.585 0.453 0.208 0.901 0.399 0.819 rs61888800 0.578 0.447 0.835 0.659 1.436 0.488 0.107 0.744 2.467 0.291 1.296 0.523 0.750 0.687 0.794 0.672 rs11030119 1.755 0.416 0.244 0.885 0.682 0.711 2.517 0.113 1.785 0.410 1.645 0.439 0.201 0.905 0.606 0.739 rs2883187 0.598 0.742 0.465 0.792 2.557 0.278 5.410 0.067 3.854 0.146 1.000 0.607 0.201 0.904 1.044 0.593 GRIN2B rs7301328 1.885 0.390 0.051 0.975 2.158 0.340 0.604 0.739 0.725 0.696 4.587 0.101 0.801 0.670 0.520 0.771 FAIM2 rs1044677 3.141 0.208 0.428 0.807 3.526 0.172 4.561 0.102 6.139 0.046 1.363 0.506 3.802 0.149 5.069 0.079 rs7132908 0.101 0.951 1.083 0.582 0.115 0.944 5.679 0.058 4.786 0.091 0.949 0.622 2.640 0.267 2.155 0.340 rs706795 2.507 0.285 7.800 0.020 8.424 0.015 0.188 0.910 0.299 0.861 4.259 0.119 8.159 0.017 6.190 0.045 HTR2A rs7997012 2.240 0.326 1.436 0.488 0.205 0.902 1.868 0.393 0.826 0.662 3.039 0.219 2.444 0.295 6.112 0.047 rs9534510 2.659 0.265 0.286 0.867 1.117 0.572 0.510 0.775 2.451 0.294 2.420 0.298 0.022 0.989 0.578 0.749 SLITRK5 rs10450811 1.308 0.520 2.490 0.288 9.887 0.007 2.584 0.108 0.007 0.936 0.159 0.924 3.739 0.154 4.079 0.130 rs9557425 1.668 0.434 2.073 0.355 1.506 0.471 1.630 0.443 4.189 0.123 3.356 0.187 0.127 0.938 0.433 0.805 SLC6A4 rs3813034 3.491 0.175 4.841 0.089 4.875 0.087 1.966 0.374 3.111 0.211 5.796 0.055 0.266 0.875 2.672 0.263 DLGAP1 rs2240899 2.997 0.223 4.521 0.104 3.370 0.185 0.986 0.611 0.445 0.801 2.012 0.366 0.479 0.787 0.571 0.752 rs8096794 1.495 0.474 2.105 0.349 0.646 0.724 2.572 0.276 2.459 0.292 0.036 0.982 0.972 0.615 0.832 0.660 FUT2 rs681343 1.094 0.579 1.460 0.482 1.666 0.435 0.330 0.848 0.358 0.836 0.937 0.626 0.061 0.970 2.912 0.233 BTBD3 rs1996132 1.219 0.544 2.854 0.240 1.051 0.591 5.382 0.068 0.025 0.988 0.462 0.794 1.596 0.450 0.772 0.680 COMT rs9617850 2.054 0.152 0.044 0.978 2.265 0.322 1.671 0.434 0.489 0.783 2.000 0.368 0.074 0.964 0.043 0.979 rs933271 0.957 0.620 0.079 0.962 3.328 0.189 2.667 0.264 1.504 0.472 1.845 0.398 0.238 0.888 1.028 0.598 rs4818 0.583 0.747 3.181 0.204 1.543 0.462 1.719 0.423 2.230 0.328 0.767 0.681 0.156 0.925 0.589 0.745 MAOA rs3788862 0.395 0.821 1.091 0.580 0.261 0.878 3.638 0.162 1.131 0.568 0.352 0.839 0.522 0.770 1.014 0.602 rs1465107 0.271 0.873 0.767 0.681 0.577 0.749 3.638 0.162 1.663 0.435 0.352 0.839 0.688 0.709 0.657 0.720 rs979605 0.326 0.849 1.817 0.403 0.878 0.645 4.304 0.116 2.711 0.258 0.980 0.613 1.455 0.483 1.892 0.388

229

230

rs1137070 0.199 0.905 2.717 0.257 1.391 0.499 4.847 0.089 2.495 0.287 0.098 0.952 2.761 0.251 2.536 0.281 Highlighted in yellow indicates initial significant P value prior to adjusting for multiple comparisons

230

231

SRI Response in OCD

Figure 6.1. Result of Genetic Variations across Remote Regulatory Regions in 14 OCD Candidate Genes and Antidepressant Response with only Cacausians. Red box consists of SNP within the gene instead of remote regulatory regions.

231

232

6.5 Discussion

This is the first pharmacogenetic study in OCD that examined multiple system genes and regulatory region variations within OCD candidate genes. We concluded that serotonergic system genes may play a vital role in OCD antidepressant response although our results did not survive correction for multiple comparisons.

However, several glutamate-related genes appeared to potentially influence SRI response in OCD. SLC1A1 rs3933331, the closest to significance level in addition to another SNP within SLC1A1, rs3780413, SLITRK5 rs10450811, and DLGAP1 rs2240899 may alter antidepressant response and this is likely because of the increasing recognition of the role of glutamate neurotransmission in the pathoetiology of OCD, resulting in promising novel targets for the treatment of psychiatric disorders including OCD (Grados et al., 2013), mood disorders (Krystal et al., 2002), and schizophrenia (Kantrowitz and Javitt, 2012). Glutamatergic system genes have been consistently implicated in the etiology of OCD (Pauls et al., 2014). Meta-analysis of SLC1A1 reported significant association with OCD (Stewart et al., 2013); furthermore, additional glutamate-related genes have arisen from the two recent GWASs of OCD (Stewart et al., 2013). It is also interesting to note that a SLC1A1 haplotype, which consists of rs3780413, was associated with antipsychotic-induced obsessive-compulsive symptoms (P=0.04 after permutation) (Kwon et al., 2009).

In addition to the importance of glutamate in OCD, serotonergic system genes including HTR1B have been repeatedly implicated in the genetic basis of OCD. Multiple serotonergic system genes have been postulated in the pathogenesis of OCD (Pauls, 2010; Pauls et al., 2014) because SSRI(s) as the first-line pharmacological treatment of OCD alter serotonin level in the brain. Although a recent meta-analysis of all OCD association studies (Taylor, 2013) showed nominal significant results in HTTLPR (OR=1.251, 99% confidence interval [CI] 1.048-1.492, P=0.003) and HTR2A rs6311 (OR=1.219, 99% CI 1.037-1.433, P=0.003), the result for the HTR1B rs6296 marker was negative (OR=1.139, 99% CI 0.860-1.509, P=0.232). In the present study, several polymorphisms, rs2000292, rs6296, and rs1778258, were observed to have trend associations with SRI response in OCD. Although three previous studies have failed to indicate

232

233

this gene in the pharmacogenetics of OCD (Corregiari et al., 2012; Denys et al., 2007; Miguita et al., 2011), this is likely due to limited power to detect significant findings.

HTR2A appeared to be associated with SRI response and a recent meta-analysis of all OCD association studies (Taylor, 2013) showed nominal significant results in the HTR2A rs6311 marker (OR=1.219, 95% CI 1.037-1.433, P=0.003), supporting its role in the genetic risk for OCD. This SNP has also been shown to be associated with early-onset and severity in OCD (Walitza et al., 2012), which provided further evidence of this marker in the etiology of OCD. Interestingly, it appears that HTR2A rs7997012 has also been shown to predict antidepressant response in major depressive disorder (MDD) (Fabbri et al., 2013; Singh et al., 2014), which overlaps with antidepressant response in our current Caucasian OCD sample.

FAIM2 was one of the top hit genes from the first and recent GWAS of OCD (Stewart et al., 2013). It encodes for an anti-apoptotic protein that protects cells from fas-induced apoptosis and it is thought to play a role in cerebellar development by affecting cerebellar size, internal granular layer thickness, and Purkinje cell development (Fernandez et al., 2007; Somia et al., 1999). It has previously been shown to be associated with obesity (Speliotes et al., 2010); however, further studies will need to be conducted to examine its role in psychiatric disorders and antidepressant response.

BTBD3 was the top hit gene from the first and recent GWAS of OCD with a P value of 3.84E-08 in the trio sample (Stewart et al., 2013). It encodes for a protein, which is involved in multiple cellular functions including the regulation of dendritic orientation (Matsui et al., 2013), cytoskeleton dynamics, transcription, ion channel assembly and gating, protein ubiquitination and degradation (Perez-Torrado et al., 2006). BTBD9, a close family to BTBD3, has previously been shown to be associated with Tourette syndrome (Riviere et al., 2009), which is frequently comorbid with OCD. However, further studies will need to be conducted to examine its role in OCD.

Some of the limitations of this study include small sample size and retrospective medication response data. However, this is currently the largest single site OCD sample for pharmacogenetic study to-date. Prospective study design would ensure better quality of drug response data and would reduce recall bias. Another potential limitation is the intra-ethic 233

234

heterogeneity. Although we attempted to remove non-Caucasians from the final analysis and detected different findings, the best approach is either to perform a GWAS or to identify ethnic groups using ancestry informative markers.

Although our results did not survive multiple testing, these interesting findings provide the first step in further examining the potential influence of genetic variations across regulatory regions of OCD multi-system candidate genes in antidepressant response, suggesting multiple neurotransmission pathway involvement in the genetic determinant of antidepressant response in OCD.

Antidepressant pharmacogenetics studies in OCD to date are still limited as compared to MDD and thus require further research to expand our understanding of genetic factors that influence antidepressant response and tolerability. Several promising candidate genes in the serotonergic and cytochrome P450 systems provide insight into SSRI response in OCD. Future direction will include interaction between serotonergic system genes and cytochrome P450 system genes. Larger samples and replication studies in a prospective study design are warranted to further explore these interesting findings. Pharmacogenetic research in OCD will ultimately provide physician guidance in choosing the most appropriate and well-tolerable antidepressants for each patient in order to reduce clinical symptoms and duration of suffering.

234

235

Chapter 7 GENERAL DISCUSSION

7 General Discussion

Obsessive-compulsive disorder (OCD) is a highly complex and common psychiatric disorder, which is often comorbid with other psychiatric illnesses. Individuals suffering from OCD often present with unique symptoms and characteristics of this illness. Despite the long existence of OCD, this condition is only recently better well-recognized. Although there are relatively useful pharmacotherapy and psychotherapy for the treatment of OCD, approximately 30% of OCD patients do not get relieved from these treatments. Genetics is one of the fast- advancing approaches to identify the underlying mechanism of OCD. This will ultimately give rise to new targets for drug development and new technologies for the management of this chronic and debilitating disease.

Genetics across psychiatric disorders have not yielded consistent results given the need for extremely large sample size in order to detect the impact of multiple genes of small effect in polygenic diseases. Similarly in OCD, no one single gene has been consistently replicated across over 500 human genetic studies (Pauls et al., 2014). Recruitment of over 10,000 OCD participants [as in the case of schizophrenia GWASs: first PGC GWAS with initial 8,832 cases for meta-analysis and 7,413 cases for replication and identified 22 genome-wide significant loci (Ripke et al., 2013); second PGC GWAS with 36,989 cases, which identified 108 significant genetic loci (Schizophrenia Working Group of the Psychiatric Genomics Consortium, 2014)] for genetic research study is a highly challenging task even with international collaborative effort because many individuals with OCD do not seek help due to numerous barriers to mental health service access including societal reason (i.e., stigma, cross cultural diversity), limited resources (i.e., shortage of mental health care professionals), system issues (i.e., regional disparities, lack of integration between mental health and health care services), and individual (patient) factors (i.e., indecision, risk aversion, severity of illness, poor insight into illness). Moreover, OCD is not as well-recognized as other severe and persistent mental illnesses such as depression and

235

236

schizophrenia. Therefore, from a clinical perspective, there is a need to increase public awareness and psychoeducation regarding mental health, especially this debilitating illness, OCD, and its related disorders in addition to promoting research and recruitment of psychiatric participants for clinical and genetic studies.

From the genetic viewpoint, technological advances certainly have progressed into a new era. The seminal publication of the human genome in 2001 identified less than 2% of the three billion nucleotide sequences as functional (Lander et al., 2001). It was since 2012 that researchers have detected most of the genetic sequences to have functional properties. Previously considered “junk” DNAs have been shown to regulate gene expression remotely. These regions act as enhancer or silencer, allowing transcription factors (TFs), repressors, or co- activators to bind. Alteration of gene expression ultimately changes the level of gene product including proteins, which are the fundamental cellular building blocks of life. Therefore, investigation of remote regulatory elements may provide further insights into the mechanism of disease development and progression.

Most previous studies have focused on markers within or immediately surrounding genes of interest. Many examined single nucleotide polymorphisms (SNPs) had no known functional significance at the time of the study. Emerging data from the ENCyclopedia Of DNA Elements (ENCODE) project (ENCODE Project Consortium, 2012) allowed researchers to select SNPs based on function. The aim of our study is to utilize this functional data to choose SNPs to study possible association with OCD subphenotypes and drug response in OCD.

7.1 Summary of Findings and Implications

This thesis provides an overview of the clinical phenomenology of OCD and supports the heterogeneity and complexity of this disorder. Given that genetic studies in OCD have yielded mixed findings, the goal of genetic studies in this thesis was to identify genetic variants that affect various subphenotypes of OCD. Demographical data (gender) and OCD subphenotypes, which included age at onset (AAO), Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) severity score, Y-BOCS symptom dimensions, psychiatric comorbidities, family history of

236

237

obsessive-compulsive and related disorders (OCRDs), and serotonin reuptake inhibitor (SRI) response were examined in this thesis. In addition to investigating these subphenotypes, two genetic approaches were used in this thesis including candidate gene approach and genome-wide association study (GWAS).

A limited number of genetic studies in search of genetic variations to examine various subphenotypes of OCD have been published. Furthermore, previous genetic studies have mainly focused on SNPs within the exons of a candidate gene. With scientific breakthrough and the identification of regulatory elements (ENCODE Project Consortium, 2012), new SNPs with plausible functional significance can be tested.

Although the studies in this thesis did not yield significant results after correcting for multiple comparisons, important trends were detected in genes within the serotonergic and glutamatergic neurotransmitter systems in addition to new genes with currently unknown relationship to psychiatric disorders.

The following sections will highlight the clinical characteristics of the OCD sample and trend associations of the genetics of OCD subphenotypes and pharmacogenetics (Figure 7.1).

7.1.1 Clinical OCD (Chapter 3)

Patients who suffer from OCD often present with diverse clinical characteristics. Although our OCD sample did not show significant clinical differences between males and females, female OCD patients in this sample reported a higher lifetime rate of skin picking disorder (SPD) and trichotillomania (TTM). Using statistical and analytic approaches, we detected subgroups of OCD based on AAO (early, intermediate, and late) and Y-BOCS symptom dimension (5- and 6-factor models). In contrast to gender having no differing clinical presentations, early-onset OCD had greater symmetry/order and contamination/cleaning than the intermediate- or late-onset groups but did not show any differences in the rates of psychiatric comorbidity and family history of OCRDs. Our sample appeared to have similar rates of psychiatric comorbidity when comparing to other published studies (Hofmeijer-Sevink et al., 2013; Torresan et al., 2013). Moreover, this study also confirmed a heightened family history of 237

238

OCRDs in our OCD sample and the familial OCD group appeared to have a higher rate of psychiatric comorbidity. One interesting observation is that this OCD sample had a relatively higher SRI response rate than other published studies (Foa et al., 2005; Greist et al., 1995; McDougle et al., 1993). Furthermore, when comparing the OCD symptom severity between the current study and other published studies, our sample had an overall higher mean of the Y-BOCS severity score (Albert et al., 2015). Higher severity of OCD symptoms was previously found to be associated with poorer SRI response, which did not support the high SRI response rate in our sample. This could be due to the method used to define SRI response (retrospective versus prospective, CGI-I scale versus Y-BOCS severity score, number of different SRIs per participant versus drug trial with one SRI, inclusion criteria of adequate SRI trial) and possible increasing awareness and psychoeducation of OCD in the public. Regarding clinical predictors of SRI response, we found that symmetry/order, higher number of comorbid current psychiatric disorders especially mood disorders, family history of OCD and hoarding disorder (HD) were significantly associated with non-response to SRI and we also detected a trend of greater severity in non-responders. Our results could be interpreted as similar to the reported clinical variables predicting poor SRI response: more depressive symptoms (Denys et al., 2003) (presence of comorbid mood disorders in this study) and hoarding symptoms (Marazziti and Consoli, 2010) (significant family history of hoarding with higher loading of hoarding in the proband).

This study’s OCD participants have been recruited from two tertiary care psychiatric centres and these participants are often more debilitated and severe than those from the primary care centres. This large OCD clinical dataset provided support of its clinical heterogeneity, which poses an important consideration when using these clinical data for genetic analyses.

7.1.2 Genetics of OCD Phenotypes (Chapter 4 and Chapter 5)

The genetic effects of OCD subphenotypes were investigated using two study approaches, candidate gene association study and GWAS.

The first study (Chapter 4) utilized a novel selection methodology to identify SNPs of OCD candidate genes with potential functionality based on the ENCODE project. OCD

238

239

phenotypes, in particular, AAO and Y-BOCS severity score, were examined as continuous variables in order to preserve the study’s statistical power. We then explored the genetics of Y- BOCS symptom dimensions, comorbid psychiatric conditions, and family history of OCRDs, which drastically reduce the statistical power due to sample stratification. In this study, SNP from the newly identified gene from the first GWAS (Stewart et al., 2013), the DLGAP1 rs2240899 (P=0.009) GG genotype occurred more frequently in OCD females than in males. The FUT2 rs681343 (P=0.003) and COMT rs4818 (P=0.042) may potentially predict earlier onset of OCD. Regarding symptom dimensions, given further increase in multiple comparisons with the additional five or six symptoms dimensions of OCD, the results will need larger samples to reflect conclusive evidence for genetic variants associating with different symptom dimensions. In the preliminary findings, BDNF rs3763965 was associated with symmetry/order/repeating/checking, contamination/cleaning, and somatic factors; GRIN2B rs7301328 was associated with symmetry/ordering/repeating/checking and somatic factors; DLGAP1 rs8096794 was associated with aggression/sexual/religious and somatic factors, which was relatively similar to the 6-factor model. Significant results were observed between comorbid mood disorders and GRIK2 rs1556995 (P=0.040), comorbid OCRDs and FUT2 rs681343 (P=0.006), family history of OCD and HTR2A rs7997012 (P=0.011), family history of TTM and SLC1A1 rs7022369 (P<0.001) and DLGAP1 rs8096794 (P=0.045), family history of BDD and SLC1A1 rs3933331 (P=0.010) and BDNF rs2883187 (P=0.046). Although none of the results persisted after Nyholt correction for multiple comparisons, interesting preliminary findings implicated both the null and alternative hypotheses –genetic variants may be associated with specific and several subphenotypes of OCD.

The second study (Chapter 5) was conducted using GWAS approach and this is the first to examine OCD phenotypes to date. Given its relative small sample size (N=203), we performed family-based quantitative trait analysis on the continuous variables, AAO and Y- BOCS severity score, to boost the power of the study by incorporating family members (N=94) into the analysis. No genome-wide significance was detected for AAO or Y-BOCS severity. However, interestingly, the strongest signals for earlier AAO were observed on chromosome 6p21.3, which is located adjacent to the human leukocyte antigen (HLA) region. Immunological factors have been proposed in OCD including the proposed subgroup of OCD, Pediatric Acute- onset Neuropsychiatric Syndrome (PANS). Furthermore, nearby genes detected from this region 239

240

have important roles in signaling cascade that are crucial in the regulation of cell growth, differentiation, and metabolism. Top hit SNPs were detected within genes that have pleiotropic effects on multiple systems, especially in neuroplasticity and cell development. Thus, further investigations into immunological factors such as infection status of Group A β-hemolytic streptococcal infection (GABHS), pathway analysis or hypothesis-driven GWAS will be warranted to help identify genetic risk of an earlier onset or a more severe form of OCD.

7.1.3 Pharmacogenetics of OCD (Chapter 6)

Pharmacogenetics of OCD is considered to be at its early stage. Although there were few published pharmacogenetic studies of OCD, none of the significant results have been replicated solidly. We have approached the genetics of SRI antidepressant response with two different strategies, one examining important OCD candidate SNPs and the other investigating SNPs within the remote regulatory regions in OCD candidate genes.

The first smaller study identified several interesting candidates (HTR1B, DLGAP2, SLC1A1, GRIN2B, HTR2A, SLC6A4, OLIG2, COMT, and MAOA) that may help predict SRI antidepressant response. The second larger study further demonstrated the importance of the serotonergic and glutamatergic system genes (HTR1B, HTR2A, SLC1A1) in SRI antidepressant response. The HTR2A rs7997012 marker has been relatively consistently shown to be associated with SRI antidepressant response in major depressive disorder (MDD) (Fabbri et al., 2013; Singh et al., 2014) and an increasing trend (P=0.077 to P=0.047) was noted from the smaller sample to the enlarged sample (N=108 to N=217). Intriguingly, findings closest to reaching a significance level were mainly SNPs within the serotonin-related genes including HTR1B, HTR2A, and SLC6A4, and glutamate-related genes such as SLC1A1, DLGAP2, and GRIN2B. The serotonergic system has been studied extensively in the pharmacogenetics of MDD (Fabbri et al., 2013; Singh et al., 2014) but limited research has been conducted in OCD. These initial preliminary results suggest further examination into serotonin-related genes given that almost all antidepressants act on the serotonergic system. Genes involved in the glutamatergic neurotransmitter system have been robustly implicated in OCD diagnosis as shown with converging evidence using several different genetic approaches including linkage studies (Hanna

240

241

et al., 2002; Willour et al., 2004), candidate gene studies (Pauls et al., 2014; Stewart et al., 2013), and the recent two GWASs (Mattheisen et al., 2015; Stewart et al., 2013). Additionally, the very recent and first GWAS of OCD pharmacogenetics also revealed the impact of genes in the glutamate pathway in influencing drug response (Qin et al., 2015). Further support of the importance of glutamate in OCD came from studies examining the complex neurocircuitry of OCD (Milad et al., 2012; Pauls et al., 2014) and the increasing use of glutamate agents for the treatment of OCD (Grados et al., 2013). Although none of our pharmacogenetic results survived multiple testing, these preliminary findings may provide the first step to further examine the effect of genetics in SRI antidepressant response in OCD.

A genetic variant within the novel top hit candidate gene from the first OCD GWAS, FAIM2, was shown to have potential predictive value in SRI antidepressant response in OCD (Stewart et al., 2013). This gene has been implicated in cerebellar development (Fernandez et al., 2007; Somia et al., 1999) and was shown to be associated with obesity (Speliotes et al., 2010). Despite our initial interesting finding with FAIM2, the mechanism of this gene in antidepressant response or psychiatric illness remains unknown to date; thus, future directions should involve conducting functional investigations and additional genetic studies to examine its role in OCD, SRI antidepressant response, and other psychiatric illnesses.

241

242

Figure 7.1. Summary of Genetic Results in OCD Subphenotypes. This figure illustrates the results of this Ph.D. thesis, showing shared and different common genetic variations that are associated with OCD phenotypes.

242

243

7.2 Limitations and Considerations

In most genetic studies, small sample size and testing for multiple comparisons are the two main limitations to consider when interpreting results. Other additional limitations include the use of categorical diagnostic criteria to identify individuals who suffer from OCD and the current method of identifying different OCD subtypes. These limitations are general to all psychiatric research to date.

7.2.1 Sample Size and Genomic Coverage

The power of a genetic association study depends on numerous variables, which can be divided into three main categories: statistical model, genetic variation, and sample. The underlying genetic model is assumed to be additive, recessive, or dominant. Several factors regarding genetic variation can affect the power of statistical analysis. The estimated effect size and the allele frequency of the examined genetic variation in addition to the linkage disequilibrium between the examined and risk genetic variations are all important variables to consider. One of the most important factors is the sample size and the examined phenotype.

The number of participants for our Toronto-based clinical OCD sample is considered as one of the largest OCD cohorts from one institution in the world with consistent clinical assessment protocol. Although this OCD sample is large compared to many published studies, it has a statistical power of 80% to detect an effect size of 0.43 for the Student T tests (99% to detect an effect size of 0.5) using the G*Power program (version 3.1.9.2) (Faul et al., 2007; Faul et al., 2009). Therefore, there is still a need to combine and compare between large international OCD samples in order to detect clinically meaningful and statistically significant results.

For GWAS, limited power to detect genome-wide significance should be considered. The GWAS of OCD in the thesis used data generated from the Illumina Human610-Quadv1_B SNP BeadChip array (Illumina Inc. ®, San Diego, CA, USA) and the required case-control sample size to have at least 80% power to detect an odds ratio risk of 1.3 with a minor allele frequency of 5% is 2,606 (Illumina ®). Therefore, even the first published GWAS did not have

243

244

enough power to detect small genetic risk. Furthermore, the genomic coverage for the r2 at which at least 95% of SNPs are tagged is 0.55 (Illumina ®); therefore, newer SNP chip arrays may offer better genomic coverage for future GWASs of OCD. Our sample had enough power (81.5%) to detect 18% of the variance of AAO or Y-BOCS severity that were accounted for by a SNP (α=5E-08, additive model) using the QUANTO program (version 1.2.4) (Gauderman and Morrison, 2006). OCD is a polygenic condition and may well be characterized by the contributions of many genes of small effect as appeared to be the case in schizophrenia or bipolar disorder, for example. Our current sample did not have enough power to detect genetic differences of small impact on OCD subphenotypes.

For the continuous phenotypes in the candidate gene studies, the OCD sample had at least 80% power to detect 7.4% of the variance of AAO and Y-BOCS severity that were accounted for by each SNP (α=0.002, additive model) using the QUANTO program (Gauderman and Morrison, 2006). For a polygenic disease, 7.4% variance is considered to be high. Given the recent explosion of GWASs across different psychiatric disorders, we now understand that multiple genetic variations may confer small risk, likely with a relative risk ranging from 1.1 to 1.5, to these mental illnesses. Thus, larger and better characterized samples are required, in order to extend the interesting results in this thesis, to better delineate subgroups of OCD, and to enable more directed studies regarding the localization and characterization of genetic factors in the etiology of subphenotypes of OCD and its related disorders.

Although our OCD sample with adequate pharmacogenetic data for analysis had enough power (80.5%) to detect an even larger (8.3%) variance of SRI response that was accounted for by each SNP (α=5.05E-4, additive model) using the QUANTO program (version 1.2.4) (Gauderman and Morrison, 2006), drug response phenotype can have a larger effect size when compared to disease status as shown by the recent small-sample (discovery sample N=139) pharmacogenetic GWAS of antipsychotic-induced weight gain (Malhotra et al., 2012). In this small sample study, a GWAS-significant hit was found for the rs489693 near the melanocortin-4 receptor gene.

244

245

7.2.1.1 Trend Associations

Genetic studies in psychiatric disorders require replication to confirm initial significant findings given the high false positive and negative rates; therefore, each experiment within this thesis will need to be replicated in a larger and independent OCD sample.

The present study only represents a small part of the ongoing and larger international OCD genetics collaborative research investigation in which polymorphisms across candidate genes and GWASs of potential candidates in the international OCD population are screened, in order to elucidate the potential important role these candidates may play in the etiology of OCD. Therefore, by utilizing the large international OCD sample, we can extend the current studies’ preliminary trends to determine whether these findings will eventually reach statistical significance. Replication samples have been identified through the International OCD Foundation Genetic Collaborative (IOCDFGC) group and also from our ongoing international collaborators outside of the IOCDFGC group.

Regarding the clinical phenomenology of OCD, two large samples, 1,001 Brazilian (Professor Euripedes C. Miguel and Professor Roseli Shavitt from the University of São Paulo) and 500 Spanish individuals (Professor Pino Alonso from the Bellvitge Hospital in Barcelona) , will be participating in the replication of our current preliminary findings. This is in addition to the new direction to compare OCD clinical features across different cultural backgrounds. For the GWAS and genetics of OCD subphenotypes as well as pharmacogenetics of OCD, replication will also be conducted using the Spanish OCD sample (N=500). A proposed GWAS of SRI antidepressant response in OCD will be investigated in a subset sample of the first published GWAS of OCD (Stewart et al., 2013).

7.2.2 Corrections for Multiple Comparisons

For each single-marker test, we set the critical P-value α at 0.05 for which the false- positive (type I error) rate is 5% for each test. Given that we analyzed multiple markers in each study, the random chances of yielding a false positive result was increased in the overall study.

245

246

Multiple comparisons should always be taken into account when implicating the involvement of a gene variant to the diagnosis or phenotype of a disorder.

Our results have been corrected for multiple comparisons with the Nyholt method (Nyholt, 2004; Li et al., 2005) that accounts for inter-marker linkage disequilibrium (LD). Bonferroni correction is the most stringent test, which offers the most conservative approach to control for false positive results (Bonferroni, 1936). It is generally used for post-hoc pair-wise comparisons among independent comparison groups after a global comparison among all comparison groups yields a significant result. It involves setting the critical P-value α at a level calculated by dividing 0.05 by the number of pair-wise comparisons. It has been adopted by many genetic studies. For example, we tested 30 SNPs across the remote regulatory regions in OCD subphenotypes. The critical P-value threshold, below which we should consider the results significant, would have been 0.05/30, or 0.0017. However, some argue that Bonferroni correction is overly conservative. Another issue with Bonferroni correction is that studied markers are in high LD with one another. When a high degree of LD exists between markers, which are assumed to be completely independent, then the Bonferroni correction would markedly overcorrect for the inflated false-positive rate, resulting in a reduction in power (Perneger, 1998). Nyholt (Nyholt, 2004) offered a solution by accounting for correlation between markers in the calculation of the critical P-value adjustment, thus decreasing the effective number of markers used in the study. Therefore, the correction factor would change with different sets of tested polymorphisms that have different LD patterns among the markers tested (Perneger, 1998; Sham and Purcell, 2014).

With the hypothesis-driven candidate gene approach, the number of multiple comparisons is much lower than for GWAS. For candidate gene approach, we used the Nyholt method (Nyholt, 2004; Li et al., 2005). As for GWAS, we used the significance threshold of 5E- 08 that was established for GWAS based on data collected on the International Haplotype Map Consortium, where up to an estimated one million independent markers across the genome may be tested (Pe’er et al., 2008).

There are other methods to correct for multiple comparisons. The Westfall and Young Permutation is the only correction accounting for genes co-regulation (Westfall and Young,

246

247

1993); nonetheless, computation is very slow and is also very conservative. The least stringent of all corrections is the Benjamini and Hochberg False Discovery Rate (Benjamini and Hochberg, 1995). This correction provides a good balance between discovery of statistically significant genes and limitation of higher rates of false positive findings; however, this does not take into consideration that some examined markers in the study are not independent of one another and is also less stringent.

Given that the current thesis did not yield strong significant results prior to correction for multiple testing, we used the Nyholt method (Nyholt, 2004; Li et al., 2005) for candidate gene studies and the 5E-08 threshold for the GWAS.

For the genetics of OCD subphenotypes, since we investigated multiple subphenotypes, correcting the significance level after testing several traits should be accounted for. However, since our results did not reach significance level after Nyholt correction, there was no need to further correct for subphenotypes (given the negative results). In the case when the results were significant after correction for multiple SNP comparisons, adjustement of P-value would be used to correct for multiple phenotype comparisons using the p_ACT module (version 1.2) in the R program (Conneely and Boehnke, 2007).

7.2.3 Diagnostic Uncertainty

OCD, similar to other psychiatric disorders, is likely characterized by the extreme of the clinical phenomena that are on a normally distributed continuum in the general population. These continuous, bidirectional phenomena are not limited to the traditional diagnostic criteria of OCD. Categorical approach as set by the DSM-IV and the most recent DSM-5 only allows for one subtype or cluster, which is one of the major limitations of the DSM model. Here, we identified a symptom dimensional approach (utilizing the Y-BOCS symptom checklist and conducting factor analysis to generate symptom dimensions), which has been replicated with consistency and can better deal with the issues of comorbidity and/or co-existence of various symptoms in OCD (Mataix-Cols et al., 2005). However, all research participants have been

247

248

recruited using the diagnostic approach as set by the DSM-IV or ICD-10 criteria, which may potentially hinder the growth of genetic research in psychiatric disorders.

7.2.4 Use of Factor Analysis to Identify OCD Subtypes or Symptom Dimensions

It is important to note that the classification of OCD subjects into their respective symptom dimensions is not mutually exclusive given that patients often present with multiple symptoms across different symptom dimensions. Therefore, the complexity of grouping an OCD individual into different symptom dimensional factors requires the identification of rigorous statistical model. In this study, we attempted to use factor scores as generated by the factor analysis of the Y-BOCS symptom checklist in two different approaches for genetic analyses. We initially utilized the simple method of weighing all OCD items (symptoms) equally and later compared the factor analysis by double-weighting 10 target symptoms as identified from OCD participants. Although the simple analysis generated five factors and the more complex analysis produced six factors, they were all consistent with previous literatures. However, the best way to generate symptom dimensions is to use the pre-existing dimensional Y-BOCS (DY-BOCS; Rosario-Campos et al., 2006), which takes double the time of clinical assessment. Therefore, better approaches to identify symptom dimensions should help to reduce the time of assessment and to generate more robust symptom dimensions that are comparable to DY-BOCS.

7.3 Future Directions

Over the last decade, advances in genetic technology and international collaborative efforts have led to a vastly improved understanding of the complex genetics of many psychiatric disorders including OCD. New discoveries certainly have the potential to transform the field of psychiatric genetics, which is moving along the goal of eventual translation into clinical practice. These new advances promise new understanding and novel avenues for prevention and treatment of mental illnesses; but they will also present with significant clinical and ethical challenges.

248

249

The current status of psychiatric genetics is still at an early stage and further investigations are required to move our progress forward into meaningful targets for prevention and therapeutic interventions, which will ultimately lead to decrease of patient suffering and disability globally.

7.3.1 Trait and Subphenotype Analysis

7.3.1.1 Traits rather than Disease Status

Development of new therapies has been halted partly because of the limited clinical validation in psychiatric disorders. Diagnostic systems including the DSM-IV, DSM-5, and ICD-10 define disorders according to symptoms and syndrome rather than their neurobiological substrates. Biological heterogeneity within the diagnostic groups has delayed the identification of underpinning mechanisms.

The National Institute of Mental Health (NIMH) has already initiated a strategic plan, the Research Domain Criteria (RDoC) to “develop, for research purposes, new ways of classifying mental disorders based on behavioral dimensions and neurobiological measures” (http://www.nimh.nih.gov; Etkin and Cuthbert, 2014), as opposed to the current DSM-IV or DSM-5 categorical diagnostic approach. The goal of this initiative is to call for new ways to classify psychopathology based on dimensions of observable behaviour with established biological validity rather than according to the current diagnostic categories. Therefore, trans- diagnostic approaches may be of considerable value for understanding obsessionality and compulsivity as potential new neuropsychiatric domains in order to support innovation in developing evidence-based treatments.

Given the difficulty in recruiting large OCD samples, the evaluation of broader spectrum phenotypes or traits may be helpful to understand the role of candidate genes in the pathoetiology of OCD and its related disorders. Moreover, OCD is characterized by diverse clinical symptomatology. Using the Y-BOCS symptom checklist as an example, there are 74 different items, of which each item corresponds to a separate symptom of OCD. Therefore, using a trans-diagnostic approach to identify obsessionality and compulsivity as dimensional

249

250

phenotypes may improve our understanding of the extent of traits (i.e., the extreme of both spectrums of obsessionality and compulsivity) in psychiatric disorders. Conducting research on polygenic dimensional trait across the general population will increase power to detect multiple genes of small effect.

7.3.1.2 Categorization by OCD Symptom Dimensions

An attempt to overcome the limitation regarding the categorization of OCD symptom dimensions is to determine these dimensions using the DY-BOCS but this takes double the time of assessment and only several international groups have adopted the DY-BOCS. Furthermore, many OCD patients present with symptoms across two or more dimensions. Therefore, there is a clear need for a practical, readily available, and standardized way to assess OCD symptoms for the exploration of genetic vulnerability in OCD. In order to maximize the use of Y-BOCS, our group has initiated a Canadian-Brazilian joint study to use collected Y-BOCS data including severity, symptom checklist, and target symptoms to estimate DY-BOCS dimensional scores in the Canadian sample. This joint collaboration aims to develop a statistical algorithm using complex computational statistical modelling, in collaboration with the Brazilian group who has both Y-BOCS and DY-BOCS in 1,001 OCD patients, in order to use the current available Y- BOCS data to best estimate DY-BOCS data. The intention of this future study is to explore the relationship between dimensional factors and OCD candidate genes in both Canadian and Brazilian samples. Another goal of this project is to examine the role of ethnicity and culture in OCD symptoms between the Canadian and Brazilian samples. This study is currently underway.

7.3.2 Future Genetic Studies

7.3.2.1 Gene Functions

Ever since the ENCODE project has shed light on the previously so-called “junk DNA” as having functional implications on remote genes (Kavanagh et al., 2013), recent genetic studies have targeted variations in these remote regulatory regions that influence disease and/or phenotype risk. Although functionality has been implicated, experiments to examine the 250

251

proposed functions of these regulatory region variations that affect remote or nearby genes are necessary in order to determine disease mechanism and/or etiology. Gene expression studies will likely add to the current genetic bible of the function of our human genome.

7.3.2.2 Gene-gene interaction

Many phenotypes are quantitative and complex in etiology with multiple genetic and environmental causes. High-throughput genomics and biostatistical models have provided an unprecedented view into the genetic architecture of complex traits.

Evidence has implicated the role of gene-gene interaction that influences risk and has suggested that susceptibility genes may act in concert with one another rather than independently such that a certain combination of alleles within specific genes may have a much stronger effect on disease or phenotype vulnerability (Mackay, 2014). Many studies have investigated the role of gene-gene interaction; however, genetic studies have been limited by relatively small sample size. Theoretically, a whole genome interaction experiment would require testing each SNP with every other SNP, leading to correcting for an exponentially larger number of statistical tests. The results are further limited by power loss after sample stratification. Gene-gene interactions are feasible in smaller samples if specific biological interactions are tested (with an a priori hypothesis).

The first recent and the only published negative family study examining gene-gene interaction is the Brazilian collaborative group and they investigated the epistatic effect between COMT and MAOA in 783 OCD trios (Sampaio et al., 2015). Therefore, this is an area where research is lacking in OCD genetics, and with international collaborative efforts, the role of gene-gene interaction can be more effectively explored.

7.3.2.3 Gene-environment interaction

We should always remember that psychiatric disorders are heritable illnesses, which are affected by multiple factors including the environment. Other known factors that can increase the risk of developing OCD can be divided into several categories including biological such as 251

252

genetics and childhood group A streptococcal infection, and environmental such as stressful life events. Stressful life events including trauma or abuse can predispose a person to react strongly and negatively to stress, which can trigger intrusive thoughts, rituals, and emotional distress characteristic of OCD. Thus, with emphasis on the improvement of clinical characterization including environmental stressors (i.e., trauma), gene-environment interaction is likely a productive step to investigate the combinatory effects in determining the risk of OCD.

Genetic studies in OCD have not yet investigated gene-environment interaction, primarily due to the lack of environmental data on the studied sample and the lack of power with small sample sizes and stratification.

Epigenetics refers to the heritable changes in gene expression via DNA methylation or histone modification without changes to the underlying DNA sequence – a change in phenotype without a change in genotype. To date, there are no published epigenetic studies in OCD.

This thesis implicated two interesting candidates for further exploration as examples of gene-environment interaction. FUT2 was found to be an interesting candidate based on a recent GWAS of OCD (Stewart et al., 2013). Furthermore, this gene predicted plasma vitamin B12 levels in two independent GWASs (Hazra et al., 2008; Tanaka et al., 2009). Low vitamin B12 serum level was previously reported in OCD patients when compared to healthy controls (Hermesh et al., 1988; Türksoy et al., 2014) and deficient serum vitamin B12 was associated with OCD symptoms (Sharma and Biswas, 2012). Therefore, this gene may be associated with OCD in the context of low vitamin B12 serum level and using gene-environment interaction to examine the effect of FUT2 and vitamin B12 serum level in OCD may identify a novel pathway of the development of OCD. Another intriguing finding from this thesis is the multiple trend associations of SNPs within the chromosome 6p21.3 region, which has been implicated in OCD. Several immune-related genes were associated with OCD (Cappi et al., 2012; Zai et al., 2004). A subgroup of OCD called the Pediatric Acute-onset Neuropsychiatric Syndrome (PANS) has recently been proposed in order to expand the clinical characteristics of the previously termed Pediatric Autoimmune Neuropsychiatric Disorders Associated with (group A β-hemolytic) Streptococcal (GABHS) infection (PANDAS) (Swedo et al., 2012). Therefore, a previous

252

253

medical history of GABHS infection may help identify yet another subgroup of OCD in conjunction with studying immune-related genes.

7.3.2.4 Other Genetic Variations

This thesis only examined the effects of common genetic variations in OCD phenomenology and SRI antidepressant response. There are other genetic variations to consider including copy-number variations (CNVs) using cytogenetic technologies and rare genetic mutation using whole exome or whole genome sequencing techniques.

Microsatellite markers including SLC6A4 5HTTLPR and STin2 VNTR, BDNF (GT)n, DAT1 VNTR, and DRD4 VNTR have been examined in the genetic basis of OCD and it is worth noting that a meta-analysis comprising eight datasets examined the 5HTTLPR marker and reported an overall significant finding (mean OR=1.251, 99th percentile CI 1.048-1.492, P=0.001) (Taylor, 2013). This thesis only explored biallelic SNPs, however, it will be important to also investigate multi-allelic markers for genetic association with OCD.

To date, only two studies have examined CNVs in OCD and none in whole exome or genome sequencing for rare genetic mutations. Walitza et al. (2012) investigated whether the HTR2A SNP (rs6311) and/or a nearby CNV was associated with OCD severity and age of onset in 136 pediatric OCD individuals and the results supported a very early onset OCD (2.5 years earlier than the typical onset) in carriers of one-copy deletion CNV and the rs6311 A allele. Another recent study conducted by the IOCDFGC group (McGrath et al., 2014) did not detect an increase in global CNV burden in OCD or OCD and Tourette syndrome but the findings implicated a 3.3-fold increased burden of large deletions on chromosome 16p13.11, a region previously associated with other neurodevelopmental disorders.

A recent submitted exome sequencing study (Cappi et al., submitted) of 20 OCD cases and their unaffected parents detected an increase de novo mutation rate relative to the healthy controls from the publicly available 1000 Genomes Project and the mutations identified in the OCD individuals suggested an enrichment of genes involved in immunological systems and the central nervous system. 253

254

Given that these different types of genetic variations have been implicated across psychiatric disorders, further investigation, specifically in OCD, is necessary, to identify other genetic contribution to the susceptibility of developing OCD or its related phenotypes.

7.3.2.5 Genetics of OCD Endophenotypes

The identification of endophenotypes with an attempt to ascertain a more homogeneous phenotype, for genetic studies, is likely important to elucidate the etiology of OCD. The search of endophenotypes in the genetics of OCD is generally guided by their strong association with OCD, high heritability, and observable similar deficits in unaffected relatives (Glahn et al., 2014; Gottesman and Gould, 2003). Endophenotypes may be of considerable utility in the search for underlying genetic diatheses and thus, for informing diagnostic classification and ultimately providing guidance to treatment.

Deficit across several cognitive domains has been reported in OCD patients; however, the nature of the deficit is inconsistent. The observed cognitive deficits in OCD are mainly within the executive function domain (Tukel et al., 2012), namely cognitive inflexibility (Bradbury et al., 2011; Chamberlain et al., 2006) and motor inhibitory control deficits (Chamberlain et al., 2005). However, only two genetic studies to date have examined the effects of genetic vulnerability to cognitive deficits in OCD (Tukel et al., 2012; Tukel et al., 2013). These two studies were performed by the same group and they examined a functional genetic variation within BDNF and COMT in cognitive functions of OCD patients. The authors detected poorer performance of the Trail-Making Test in OCD COMT rs4680 Met carrier, suggesting that the low activity enzyme creating higher level of dopamine in the prefrontal cortex may lead to poorer executive function in OCD (Tukel et al., 2013). Tukel et al. (2012) also reported that the BDNF rs6265 Met carriers had slower performance on Trail-Making Test A and B and poorer performance on verbal fluency when compared with healthy control Met carriers. Given that there are only two such studies investigating genetic variations of cognitive substrates in OCD, replication studies with large samples and the development of a broad battery of systematic and well-standardized cognitive tasks that are reliable, easy to interpret, and comparable based on modern cognitive neuroscience approaches will be required in order to derive more definitive

254

255

conclusions. A collaborative study between the Neurogenetics Section at the Centre for Addiction and Mental Health, the Frederick W. Thompson Anxiety Disorders Centre at the Sunnybrook Health Sciences Centre, and the Behavioural and Clinical Neuroscience Institute and Department of Psychiatry at the University of Cambridge was recently initiated in 2014 to examine the genetics of cognitive substrates in OCD patients. The main purpose of this study is to identify genetic variants that contribute to the risk for developing cognitive deficits in patients with OCD in comparison with their unaffected relative(s) and healthy unrelated controls. A number of computerized psychological tasks from CANTAB® will assess behavioural flexibility and motor inhibition in individuals with OCD, their unaffected relative(s), and healthy unrelated controls. This study is designed to examine the genetic vulnerability of cognitive deficit(s) in OCD. The identification of genetic factors that contribute to cognitive substrates is vitally important in ascertaining endophenotypes, as well as those who may benefit from psychotherapy (i.e., cognitive behavioural therapy) and/or medications. This study is currently underway.

Another potential source of endophenotypes is imaging genetics, which is the use of anatomical or physiological imaging technologies as phenotypes to evaluate genetic associations. In a recent systematic review of imaging genetics in OCD (Grünblatt et al., 2014), the most promising results came from the serotonergic (SLC6A4, HTR2A), glutamatergic (SLC1A1, SAP90/PSD95-associated protein 3 [SAPAP]), and dopaminergic (COMT, dopamine transporter) systems. More specifically, genetic variations of the serotonergic system have been linked to anatomical changes in the orbitofrontal cortex and the raphe nuclei; glutamatergic system to the orbitofrontal cortex, anterior cingulate cortex and the thalamus; and dopamine-related variants to the putamen (Grünblatt et al., 2014). However, the results are based on limited findings and further replication and exploration is needed.

Cognitive substrates and imaging findings are just two examples of endophenotypes, which can be explored in the genetics of OCD.

255

256

7.4 Concluding Remarks

OCD is a psychiatric syndrome with diverse and heterogeneous symptom characteristics, complex genetics, and complicated neurobiological mechanisms. Patients with OCD often present with varied symptoms but treatment tends to be the same, either with the use of SRI antidepressant, cognitive behavioural therapy, or a combination of both. Management of OCD has not provided tremendous success given its chronic and persisting nature. Therefore, better treatment with personalized/individualized medicine may enable clinicians to treat each patient with targeted therapies in the future.

Although this thesis was not able to confirm personalized medicine (precision psychiatry), it provided a framework for future genetic studies to examine homogeneous subgroups in functional genetic variations and remote regulatory regions using GWAS and fine mapping by targeting SNPs across candidate genes.

The present studies yielded helpful datasets and interesting preliminary results that will require further exploration, in a larger and well-characterized OCD sample.

Given the etiological and neurobiological complexity of OCD, it is not surprising that the identification of specific genetic factors has been relatively unproductive thus far. In the next several decades, we hope to use new developments in genetic knowledge to gain fresh new insight into the complexity of OCD, which should prove to be invaluable in establishing the etiology. This will subsequently lead to the introduction of new treatment strategies for this incapacitating psychiatric disorder

256

257

References

(APA), A. P. A. (2000). Diagnostic and statistical Manual of Mental Disorders.

(APA), A. P. A. (2013). Diagnostic and Statistical Manual of Mental Disorders, American Psychiatric Association.

(CPA), C. P. A. (2006). "Clinical practice guidelines. Management of anxiety disorders." Can J Psychiatry 51(8 Suppl 2): 9S-91S.

(WHO), W. H. O. (2009). "Global health risks: mortality and burden of disease attributable to selected major risks." Retrieved October 15, 2013, from http://www.who.int/healthinfo/global_burden_disease/GlobalHealthRisks_report_full.pdf.

Aardema, F. and K. O'Connor (2007). "The menance within: obsessions and the self." Journal of Cognitive Psychotherapy 21(3): 182-197.

Abecasis, G. R., L. R. Cardon and W. O. Cookson (2000). "A general test of association for quantitative traits in nuclear families." Am J Hum Genet 66(1): 279-292.

Abelson, J. F., K. Y. Kwan, B. J. O'Roak, D. Y. Baek, A. A. Stillman, T. M. Morgan, C. A. Mathews, D. L. Pauls, M. R. Rasin, M. Gunel, N. R. Davis, A. G. Ercan-Sencicek, D. H. Guez, J. A. Spertus, J. F. Leckman, L. S. t. Dure, R. Kurlan, H. S. Singer, D. L. Gilbert, A. Farhi, A. Louvi, R. P. Lifton, N. Sestan and M. W. State (2005). "Sequence variants in SLITRK1 are associated with Tourette's syndrome." Science 310(5746): 317-320.

Abramowitz, J. S., B. J. Deacon, B. O. Olatunji, M. G. Wheaton, N. C. Berman, D. Losardo, K. R. Timpano, P. B. McGrath, B. C. Riemann, T. Adams, T. Bjorgvinsson, E. A. Storch and L. R. Hale (2010). "Assessment of obsessive-compulsive symptom dimensions: development and evaluation of the Dimensional Obsessive-Compulsive Scale." Psychol Assess 22(1): 180-198.

Abramowitz, J. S., M. E. Franklin, S. A. Schwartz and J. M. Furr (2003). "Symptom presentation and outcome of cognitive-behavioral therapy for obsessive-compulsive disorder." J Consult Clin Psychol 71(6): 1049-1057.

257

258

Albert, U., F. Bogetto, G. Maina, P. Saracco, C. Brunatto and D. Mataix-Cols (2010). "Family accommodation in obsessive-compulsive disorder: Relation to symptom dimensions, clinical and family characteristics." Psychiatry Res 179(2): 204-211.

Albert, U., M. Manchia, A. Tortorella, U. Volpe, G. Rosso, B. Carpiniello and G. Maina (2015). "Admixture analysis of age at symptom onset and age at disorder onset in a large sample of patients with obsessive-compulsive disorder." J Affect Disord 187: 188-196.

Al-Dosari, M. S., M. Al-Owain, M. Tulbah, W. Kurdi, N. Adly, A. Al-Hemidan, T. A. Masoodi, B. Albash and F. S. Alkuraya (2013). "Mutation in MPDZ causes severe congenital hydrocephalus." J Med Genet 50(1): 54-58.

Allen, H. (1683). A Narrative of God's Gracious Dealings with that Choice Christian Mrs. Hannah Allen... London, John Wallis.

Allen, N. J., M. L. Bennett, L. C. Foo, G. X. Wang, C. Chakraborty, S. J. Smith and B. A. Barres (2012). "Astrocyte glypicans 4 and 6 promote formation of excitatory synapses via GluA1 AMPA receptors." Nature 486(7403): 410-414.

Alonso, P., M. Gratacos, J. M. Menchon, J. Saiz-Ruiz, C. Segalas, E. Baca-Garcia, J. Labad, J. Fernandez-Piqueras, E. Real, C. Vaquero, M. Perez, H. Dolengevich, J. R. Gonzalez, M. Bayes, R. de Cid, J. Vallejo and X. Estivill (2008). "Extensive genotyping of the BDNF and NTRK2 genes define protective haplotypes against obsessive-compulsive disorder." Biol Psychiatry 63(6): 619-628.

Alonso, P., M. Gratacos, C. Segalas, G. Escaramis, E. Real, M. Bayes, J. Labad, C. Lopez-Sola, X. Estivill and J. M. Menchon (2012). "Association between the NMDA glutamate receptor GRIN2B gene and obsessive-compulsive disorder." J Psychiatry Neurosci 37(4): 273-281.

Alonso, P., M. Gratacos, C. Segalas, G. Escaramis, E. Real, M. Bayes, J. Labad, A. Pertusa, J. Vallejo, X. Estivill and J. M. Menchon (2011). "Variants in estrogen receptor alpha gene are associated with phenotypical expression of obsessive-compulsive disorder." Psychoneuroendocrinology 36(4): 473-483.

258

259

Alsobrook, I. J., J. F. Leckman, W. K. Goodman, S. A. Rasmussen and D. L. Pauls (1999). "Segregation analysis of obsessive-compulsive disorder using symptom-based factor scores." Am J Med Genet 88(6): 669-675.

Alsobrook, J. P., 2nd, A. H. Zohar, M. Leboyer, N. Chabane, R. P. Ebstein and D. L. Pauls (2002). "Association between the COMT locus and obsessive-compulsive disorder in females but not males." Am J Med Genet 114(1): 116-120.

Altar, C. A., J. Hornberger, A. Shewade, V. Cruz, J. Garrison and D. Mrazek (2013). "Clinical validity of cytochrome P450 metabolism and serotonin gene variants in psychiatric pharmacotherapy." Int Rev Psychiatry 25(5): 509-533.

Andreasen, N. C., J. Rice, J. Endicott, T. Reich and W. Coryell (1986). "The family history approach to diagnosis. How useful is it?" Arch Gen Psychiatry 43(5): 421-429.

Anholt, G. E., I. M. Aderka, A. J. van Balkom, J. H. Smit, K. Schruers, N. J. van der Wee, M. Eikelenboom, V. De Luca and P. van Oppen (2014). "Age of onset in obsessive-compulsive disorder: admixture analysis with a large sample." Psychol Med 44(1): 185-194.

Aoyama, K., S. W. Suh, A. M. Hamby, J. Liu, W. Y. Chan, Y. Chen and R. A. Swanson (2006). "Neuronal glutathione deficiency and age-dependent neurodegeneration in the EAAC1 deficient mouse." Nat Neurosci 9(1): 119-126.

Arlt, A. and H. Schafer (2011). "Role of the immediate early response 3 (IER3) gene in cellular stress response, inflammation and tumorigenesis." Eur J Cell Biol 90(6-7): 545-552.

Arnold, L. M. (2003). "Gender differences in bipolar disorder." Psychiatr Clin North Am 26(3): 595-620.

Arnold, P. D., F. P. Macmaster, G. L. Hanna, M. A. Richter, T. Sicard, E. Burroughs, Y. Mirza, P. C. Easter, M. Rose, J. L. Kennedy and D. R. Rosenberg (2009). "Glutamate system genes associated with ventral prefrontal and thalamic volume in pediatric obsessive-compulsive disorder." Brain Imaging Behav 3(1): 64-76.

259

260

Arnold, P. D., F. P. Macmaster, M. A. Richter, G. L. Hanna, T. Sicard, E. Burroughs, Y. Mirza, P. C. Easter, M. Rose, J. L. Kennedy and D. R. Rosenberg (2009). "Glutamate receptor gene (GRIN2B) associated with reduced anterior cingulate glutamatergic concentration in pediatric obsessive-compulsive disorder." Psychiatry Res 172(2): 136-139.

Arnold, P. D., D. R. Rosenberg, E. Mundo, S. Tharmalingam, J. L. Kennedy and M. A. Richter (2004). "Association of a glutamate (NMDA) subunit receptor gene (GRIN2B) with obsessive- compulsive disorder: a preliminary study." Psychopharmacology (Berl) 174(4): 530-538.

Arnold, P. D., T. Sicard, E. Burroughs, M. A. Richter and J. L. Kennedy (2006). "Glutamate transporter gene SLC1A1 associated with obsessive-compulsive disorder." Arch Gen Psychiatry 63(7): 769-776.

Aruga, J. and K. Mikoshiba (2003). "Identification and characterization of Slitrk, a novel neuronal transmembrane protein family controlling neurite outgrowth." Mol Cell Neurosci 24(1): 117-129.

Aruga, J., N. Yokota and K. Mikoshiba (2003). "Human SLITRK family genes: genomic organization and expression profiling in normal brain and brain tumor tissue." Gene 315: 87-94.

Atmaca, M., E. Onalan, H. Yildirim, H. Yuce, M. Koc and S. Korkmaz (2010). "The association of myelin oligodendrocyte glycoprotein gene and white matter volume in obsessive-compulsive disorder." J Affect Disord 124(3): 309-313.

Autry, A. E. and L. M. Monteggia (2012). "Brain-derived neurotrophic factor and neuropsychiatric disorders." Pharmacol Rev 64(2): 238-258.

Baca-Garcia, E., B. R. Salgado, H. D. Segal, C. V. Lorenzo, M. N. Acosta, M. A. Romero, M. D. Hernandez, J. Saiz-Ruiz, J. Fernandez Piqueras and J. de Leon (2005). "A pilot genetic study of the continuum between compulsivity and impulsivity in females: the serotonin transporter promoter polymorphism." Prog Neuropsychopharmacol Biol Psychiatry 29(5): 713-717.

Baer, L. (1994). "Factor analysis of symptom subtypes of obsessive compulsive disorder and their relation to personality and tic disorders." J Clin Psychiatry 55 Suppl: 18-23.

260

261

Bah, J., H. Quach, R. P. Ebstein, R. H. Segman, J. Melke, S. Jamain, M. Rietschel, I. Modai, K. Kanas, O. Karni, B. Lerer, D. Gourion, M. O. Krebs, B. Etain, F. Schurhoff, A. Szoke, M. Leboyer and T. Bourgeron (2004). "Maternal transmission disequilibrium of the glutamate receptor GRIK2 in schizophrenia." Neuroreport 15(12): 1987-1991.

Baker, D. W. (2011). Electronic medical records: new electronic medical records research from D.W. Baker and co-researchers described. Computers, Networks & Communications.

Barrett, J. C., B. Fry, J. Maller and M. J. Daly (2005). "Haploview: analysis and visualization of LD and haplotype maps." Bioinformatics 21(2): 263-265.

Benjamini, Y. and Y. Hochberg (1995). "Controlling the false discovery rate: a practical and powerful approach to multiple testing." Journal of the Royal Statistical Society, Series B 57(1): 289-300.

Berman, A. E., W. Y. Chan, A. M. Brennan, R. C. Reyes, B. L. Adler, S. W. Suh, T. M. Kauppinen, Y. Edling and R. A. Swanson (2011). "N-acetylcysteine prevents loss of dopaminergic neurons in the EAAC1-/- mouse." Ann Neurol 69(3): 509-520.

Bernard, S., K. A. Neville, A. T. Nguyen and D. A. Flockhart (2006). "Interethnic differences in genetic polymorphisms of CYP2D6 in the U.S. population: clinical implications." Oncologist 11(2): 126-135.

Bernhard, K. (1970). "[Jonhannes Friedrich Miescher Symposium. 100th anniversary of the discovery of nucleic acids. Welcome]." Bull Schweiz Akad Med Wiss 25(1-2): 32-34.

Berrios, G. E. (1989). "Obsessive-compulsive disorder: its conceptual history in France during the 19th century." Compr Psychiatry 30(4): 283-295.

Bevan, J. L., J. A. Lynch, T. N. Dubriwny, T. M. Harris, P. J. Achter, A. L. Reeder and C. M. Condit (2003). "Informed lay preferences for delivery of racially varied pharmacogenomics." Genet Med 5(5): 393-399.

261

262

Bhattacharyya, S., C. L. Prasanna, S. Khanna, Y. C. Janardhan Reddy and S. Sheshadri (2005). "A family genetic study of clinical subtypes of obsessive-compulsive disorder." Psychiatr Genet 15(3): 175-180.

Bienvenu, O. J., J. F. Samuels, M. A. Riddle, R. Hoehn-Saric, K. Y. Liang, B. A. Cullen, M. A. Grados and G. Nestadt (2000). "The relationship of obsessive-compulsive disorder to possible spectrum disorders: results from a family study." Biol Psychiatry 48(4): 287-293.

Billett, E. A., M. A. Richter, F. Sam, R. P. Swinson, X. Y. Dai, N. King, F. Badri, T. Sasaki, J. A. Buchanan and J. L. Kennedy (1998). "Investigation of dopamine system genes in obsessive- compulsive disorder." Psychiatr Genet 8(3): 163-169.

Bloch, M. H., C. A. Bartley, L. Zipperer, E. Jakubovski, A. Landeros-Weisenberger, C. Pittenger and J. F. Leckman (2014). "Meta-analysis: hoarding symptoms associated with poor treatment outcome in obsessive-compulsive disorder." Mol Psychiatry 19(9): 1025-1030.

Bloch, M. H., A. Landeros-Weisenberger, B. Kelmendi, V. Coric, M. B. Bracken and J. F. Leckman (2006). "A systematic review: antipsychotic augmentation with treatment refractory obsessive-compulsive disorder." Mol Psychiatry 11(7): 622-632.

Bloch, M. H., A. Landeros-Weisenberger, M. C. Rosario, C. Pittenger and J. F. Leckman (2008). "Meta-analysis of the symptom structure of obsessive-compulsive disorder." Am J Psychiatry 165(12): 1532-1542.

Bloch, M. H., A. Landeros-Weisenberger, S. Sen, P. Dombrowski, B. Kelmendi, V. Coric, C. Pittenger and J. F. Leckman (2008). "Association of the serotonin transporter polymorphism and obsessive-compulsive disorder: systematic review." Am J Med Genet B Neuropsychiatr Genet 147B(6): 850-858.

Boardman, L., L. van der Merwe, C. Lochner, C. J. Kinnear, S. Seedat, D. J. Stein, J. C. Moolman-Smook and S. M. Hemmings (2011). "Investigating SAPAP3 variants in the etiology of obsessive-compulsive disorder and trichotillomania in the South African white population." Compr Psychiatry 52(2): 181-187.

262

263

Bonferroni, C. E. (1936). Teoria statistica delle classi e calcolo delle probabilita, Pubblicazioni del R Istituto Superiore di Scienze Economiche e Commerciali di Firenze.

Borza, C. M. and A. Pozzi (2014). "Discoidin domain receptors in disease." Matrix Biol 34: 185- 192.

Botoseneanu, A., J. A. Alexander and J. Banaszak-Holl (2011). "To test or not to test? The role of attitudes, knowledge, and religious involvement among U.S. adults on intent-to-obtain adult genetic testing." Health Educ Behav 38(6): 617-628.

Boyle, A. P., E. L. Hong, M. Hariharan, Y. Cheng, M. A. Schaub, M. Kasowski, K. J. Karczewski, J. Park, B. C. Hitz, S. Weng, J. M. Cherry and M. Snyder (2012). "Annotation of functional variation in personal genomes using RegulomeDB." Genome Res 22(9): 1790-1797.

Bradbury, C., S. E. Cassin and N. A. Rector (2011). "Obsessive beliefs and neurocognitive flexibility in obsessive-compulsive disorder." Psychiatry Res 187(1-2): 160-165.

Brady, K. T. and C. L. Randall (1999). "Gender differences in substance use disorders." Psychiatr Clin North Am 22(2): 241-252.

Brahe, C., P. Bannetta, A. Serra and F. Arwert (1986). "The increased COMT activity in Down syndrome patients is not a consequence of dosage effect owing to location of the gene on chromosome 21: further evidence." Am J Med Genet 24(1): 203-204.

Brakoulias, V., V. Starcevic, D. Berle, D. Milicevic, A. Hannan and A. Martin (2014). "The relationships between obsessive-compulsive symptom dimensions and cognitions in obsessive- compulsive disorder." Psychiatr Q 85(2): 133-142.

Brakoulias, V., V. Starcevic, D. Berle, D. Milicevic, A. Hannan, K. Viswasam and K. Mann (2014). "The clinical characteristics of obsessive compulsive disorder associated with high levels of schizotypy." Aust N Z J Psychiatry.

Brakoulias, V., V. Starcevic, D. Berle, D. Milicevic, K. Moses, A. Hannan, P. Sammut and A. Martin (2013). "The characteristics of unacceptable/taboo thoughts in obsessive-compulsive disorder." Compr Psychiatry 54(7): 750-757.

263

264

Brakoulias, V., V. Starcevic, D. Berle, P. Sammut, D. Milicevic, K. Moses, A. Hannan and A. Martin (2013). "Further support for five dimensions of obsessive-compulsive symptoms." J Nerv Ment Dis 201(6): 452-459.

Brandl, E. J., D. J. Muller and M. A. Richter (2012). "Pharmacogenetics of obsessive- compulsive disorders." Pharmacogenomics 13(1): 71-81.

Brandl, E. J., A. K. Tiwari, X. Zhou, J. Deluce, J. L. Kennedy, D. J. Muller and M. A. Richter (2014). "Influence of CYP2D6 and CYP2C19 gene variants on antidepressant response in obsessive-compulsive disorder." Pharmacogenomics J 14(2): 176-181.

Breckenridge, A., K. Lindpaintner, P. Lipton, H. McLeod, M. Rothstein and H. Wallace (2004). "Pharmacogenetics: ethical problems and solutions." Nat Rev Genet 5(9): 676-680.

Breiter, H. C., P. A. Filipek, D. N. Kennedy, L. Baer, D. A. Pitcher, M. J. Olivares, P. F. Renshaw and V. S. Caviness, Jr. (1994). "Retrocallosal white matter abnormalities in patients with obsessive-compulsive disorder." Arch Gen Psychiatry 51(8): 663-664.

Browne, H. A., S. L. Gair, J. M. Scharf and D. E. Grice (2014). "Genetics of obsessive- compulsive disorder and related disorders." Psychiatr Clin North Am 37(3): 319-335.

Burdick, K. E., D. G. Robinson, A. K. Malhotra and P. R. Szeszko (2008). "Neurocognitive profile analysis in obsessive-compulsive disorder." J Int Neuropsychol Soc 14(4): 640-645.

Cai, J., W. Zhang, Z. Yi, W. Lu, Z. Wu, J. Chen, S. Yu, Y. Fang and C. Zhang (2013). "Influence of polymorphisms in genes SLC1A1, GRIN2B, and GRIK2 on clozapine-induced obsessive-compulsive symptoms." Psychopharmacology (Berl) 230(1): 49-55.

Caley, C. F. (2011). "Interpreting and applying CYP450 genomic test results to psychotropic medications." J Pharm Pract 24(5): 439-446.

Camarena, B., C. Loyzaga, A. Aguilar, K. Weissbecker and H. Nicolini (2007). "Association study between the dopamine receptor D(4) gene and obsessive-compulsive disorder." Eur Neuropsychopharmacol 17(6-7): 406-409.

264

265

Camarena, B., G. Rinetti, C. Cruz, A. Gomez, J. R. de La Fuente and H. Nicolini (2001). "Additional evidence that genetic variation of MAO-A gene supports a gender subtype in obsessive-compulsive disorder." Am J Med Genet 105(3): 279-282.

Canada, S. (2013, September 11, 2013). "National Household Survey (NHS) Profile." 2011 National Household Survey, Statistics Canada Catalogue NO. 99-004-XWE Retrieved August 31, 2015, from http://www12.statcan.gc.ca/nhs-enm/2011/dp-pd/prof/index.cfm?Lang=E.

Cannistraro, P. A., N. Makris, J. D. Howard, M. M. Wedig, S. M. Hodge, S. Wilhelm, D. N. Kennedy and S. L. Rauch (2007). "A diffusion tensor imaging study of white matter in obsessive-compulsive disorder." Depress Anxiety 24(6): 440-446.

Cappi, C., H. Brentani, L. Lima, S. J. Sanders, G. Zai, J. B. Diniz, M. Walker, V. Reis, A. G. Hounie, D. Mariani, G. L. Requena, R. Puga, R. G. Shavitt, D. L. Pauls, E. C. Miguel and T. V. Fernandez (Submitted). "Whole-exome sequencing in obsessive-compulsive disorder identifies rare mutations and immunological pathways." Translational Psychiatry.

Cappi, C., R. K. Muniz, A. S. Sampaio, Q. Cordeiro, H. Brentani, S. A. Palacios, A. H. Marques, H. Vallada, E. C. Miguel, L. Guilherme and A. G. Hounie (2012). "Association study between functional polymorphisms in the TNF-alpha gene and obsessive-compulsive disorder." Arq Neuropsiquiatr 70(2): 87-90.

Carter, A. S., R. A. Pollock, M. K. Suvak and D. L. Pauls (2004). "Anxiety and major depression comorbidity in a family study of obsessive-compulsive disorder." Depress Anxiety 20(4): 165- 174.

Castle, W. E. (1903). "Mendel's Law of Heredity." Science 18(456): 396-406.

Cavallaro, R., G. Galardi, M. C. Cavallini, M. Henin, S. Amodio, L. Bellodi and G. Comi (2002). "Obsessive compulsive disorder among idiopathic focal dystonia patients: an epidemiological and family study." Biol Psychiatry 52(4): 356-361.

Cavallini, M. C., D. Di Bella, L. Pasquale, M. Henin and L. Bellodi (1998). "5HT2C CYS23/SER23 polymorphism is not associated with obsessive-compulsive disorder." Psychiatry Res 77(2): 97-104. 265

266

Cavallini, M. C., D. Di Bella, F. Siliprandi, F. Malchiodi and L. Bellodi (2002). "Exploratory factor analysis of obsessive-compulsive patients and association with 5-HTTLPR polymorphism." Am J Med Genet 114(3): 347-353.

Cavallini, M. C., L. Pasquale, L. Bellodi and E. Smeraldi (1999). "Complex segregation analysis for obsessive compulsive disorder and related disorders." Am J Med Genet 88(1): 38-43.

Celada, P., J. A. Siuciak, T. M. Tran, C. A. Altar and J. M. Tepper (1996). "Local infusion of brain-derived neurotrophic factor modifies the firing pattern of dorsal raphe serotonergic neurons." Brain Res 712(2): 293-298.

Celeux, G. and J. Diebolt (1990). "Une version de type recuit simul e de l'algorithme EM." C. R. Acad. Sci. Paris S er. I Math 310: 119-124.

Cengiz, M., S. N. Okutan, B. Bayoglu, A. Sakalli Kani, R. Bayar and N. Kocabasoglu (2015). "Genetic Polymorphism of the Serotonin Transporter Gene, SLC6A4 rs16965628, Is Associated with Obsessive Compulsive Disorder." Genet Test Mol Biomarkers 19(5): 228-234.

Chabane, N., R. Delorme, B. Millet, M. C. Mouren, M. Leboyer and D. Pauls (2005). "Early- onset obsessive-compulsive disorder: a subgroup with a specific clinical and familial pattern?" J Child Psychol Psychiatry 46(8): 881-887.

Chamberlain, S. R., A. D. Blackwell, N. A. Fineberg, T. W. Robbins and B. J. Sahakian (2005). "The neuropsychology of obsessive compulsive disorder: the importance of failures in cognitive and behavioural inhibition as candidate endophenotypic markers." Neurosci Biobehav Rev 29(3): 399-419.

Chamberlain, S. R., N. A. Fineberg, A. D. Blackwell, T. W. Robbins and B. J. Sahakian (2006). "Motor inhibition and cognitive flexibility in obsessive-compulsive disorder and trichotillomania." Am J Psychiatry 163(7): 1282-1284.

Chamberlain, S. R., N. A. Fineberg, L. A. Menzies, A. D. Blackwell, E. T. Bullmore, T. W. Robbins and B. J. Sahakian (2007). "Impaired cognitive flexibility and motor inhibition in unaffected first-degree relatives of patients with obsessive-compulsive disorder." Am J Psychiatry 164(2): 335-338. 266

267

Chamberlain, S. R. and L. Menzies (2009). "Endophenotypes of obsessive-compulsive disorder: rationale, evidence and future potential." Expert Rev Neurother 9(8): 1133-1146.

Chen, J., G. Lee, A. H. Fanous, Z. Zhao, P. Jia, A. O'Neill, D. Walsh, K. S. Kendler, X. Chen and C. International Schizophrenia (2011). "Two non-synonymous markers in PTPN21, identified by genome-wide association study data-mining and replication, are associated with schizophrenia." Schizophr Res 131(1-3): 43-51.

Chen, J., B. K. Lipska, N. Halim, Q. D. Ma, M. Matsumoto, S. Melhem, B. S. Kolachana, T. M. Hyde, M. M. Herman, J. Apud, M. F. Egan, J. E. Kleinman and D. R. Weinberger (2004). "Functional analysis of genetic variation in catechol-O-methyltransferase (COMT): effects on mRNA, protein, and enzyme activity in postmortem human brain." Am J Hum Genet 75(5): 807- 821.

Chen, K., D. P. Holschneider, W. Wu, I. Rebrin and J. C. Shih (2004). "A spontaneous point mutation produces monoamine oxidase A/B knock-out mice with greatly elevated monoamines and anxiety-like behavior." J Biol Chem 279(38): 39645-39652.

Chen, N., T. Luo and L. A. Raymond (1999). "Subtype-dependence of NMDA receptor channel open probability." J Neurosci 19(16): 6844-6854.

Cherian, A. V., J. C. Narayanaswamy, R. Srinivasaraju, B. Viswanath, S. B. Math, T. Kandavel and Y. C. Reddy (2012). "Does insight have specific correlation with symptom dimensions in OCD?" J Affect Disord 138(3): 352-359.

Chien, W. H., S. S. Gau, H. M. Liao, Y. N. Chiu, Y. Y. Wu, Y. S. Huang, W. C. Tsai, H. M. Tsai and C. H. Chen (2013). "Deep exon resequencing of DLGAP2 as a candidate gene of autism spectrum disorders." Mol Autism 4(1): 26.

Christov-Moore, L., E. A. Simpson, G. Coude, K. Grigaityte, M. Iacoboni and P. F. Ferrari (2014). "Empathy: gender effects in brain and behavior." Neurosci Biobehav Rev 46 Pt 4: 604- 627.

267

268

Clark, D. A., M. M. Antony, A. T. Beck, R. P. Swinson and R. A. Steer (2005). "Screening for obsessive and compulsive symptoms: validation of the Clark-Beck Obsessive-Compulsive Inventory." Psychol Assess 17(2): 132-143.

Clifford, S. (1716). The Signs and Causes of Melancholy with Directions Suited to the Case of Those Who Are Afflicted With It, Collected Out of the Works of Mr. Richard Baxter. London, S. Cruttenden.

Climacus, J. (1982). The Ladder of Divine Ascent. New York, Paulist Press.

Colantuoni, C., B. K. Lipska, T. Ye, T. M. Hyde, R. Tao, J. T. Leek, E. A. Colantuoni, A. G. Elkahloun, M. M. Herman, D. R. Weinberger and J. E. Kleinman (2011). "Temporal dynamics and genetic control of transcription in the human prefrontal cortex." Nature 478(7370): 519-523.

Coles, M. E., A. Pinto, M. C. Mancebo, S. A. Rasmussen and J. L. Eisen (2008). "OCD with comorbid OCPD: a subtype of OCD?" J Psychiatr Res 42(4): 289-296.

Collins, E. F. (1961). The Treatment of Scrupulosity in the Summa Moralis of St. Antoninus: A Historical-Theological Study. Rome, Pontificia Universitas Gregoriana.

Conneely, K. N. and M. Boehnke (2007). "So many correlated tests, so little time! Rapid adjustment of P values for multiple correlated tests." Am J Hum Genet 81(6): 1158-1168.

Consortium, E. P., B. E. Bernstein, E. Birney, I. Dunham, E. D. Green, C. Gunter and M. Snyder (2012). "An integrated encyclopedia of DNA elements in the human genome." Nature 489(7414): 57-74.

Contractor, A., G. Swanson and S. F. Heinemann (2001). "Kainate receptors are involved in short- and long-term plasticity at mossy fiber synapses in the hippocampus." Neuron 29(1): 209- 216.

Cordoba, M., S. Rodriguez, D. Gonzalez Moron, N. Medina and M. A. Kauffman (2015). "Expanding the spectrum of Grik2 mutations: intellectual disability, behavioural disorder, epilepsy and dystonia." Clin Genet 87(3): 293-295.

268

269

Corey-Lisle, P. K., R. Nash, P. Stang and R. Swindle (2004). "Response, partial response, and nonresponse in primary care treatment of depression." Arch Intern Med 164(11): 1197-1204.

Corregiari, F. M., M. Bernik, Q. Cordeiro and H. Vallada (2012). "Endophenotypes and serotonergic polymorphisms associated with treatment response in obsessive-compulsive disorder." Clinics (Sao Paulo) 67(4): 335-340.

Couzin, J. (2008). "Science and commerce. Gene tests for psychiatric risk polarize researchers." Science 319(5861): 274-277.

Cowansage, K. K., J. E. LeDoux and M. H. Monfils (2010). "Brain-derived neurotrophic factor: a dynamic gatekeeper of neural plasticity." Curr Mol Pharmacol 3(1): 12-29.

Cross-Disorder Group of the Psychiatric Genomics, C., S. H. Lee, S. Ripke, B. M. Neale, S. V. Faraone, S. M. Purcell, R. H. Perlis, B. J. Mowry, A. Thapar, M. E. Goddard, J. S. Witte, D. Absher, I. Agartz, H. Akil, F. Amin, O. A. Andreassen, A. Anjorin, R. Anney, V. Anttila, D. E. Arking, P. Asherson, M. H. Azevedo, L. Backlund, J. A. Badner, A. J. Bailey, T. Banaschewski, J. D. Barchas, M. R. Barnes, T. B. Barrett, N. Bass, A. Battaglia, M. Bauer, M. Bayes, F. Bellivier, S. E. Bergen, W. Berrettini, C. Betancur, T. Bettecken, J. Biederman, E. B. Binder, D. W. Black, D. H. Blackwood, C. S. Bloss, M. Boehnke, D. I. Boomsma, G. Breen, R. Breuer, R. Bruggeman, P. Cormican, N. G. Buccola, J. K. Buitelaar, W. E. Bunney, J. D. Buxbaum, W. F. Byerley, E. M. Byrne, S. Caesar, W. Cahn, R. M. Cantor, M. Casas, A. Chakravarti, K. Chambert, K. Choudhury, S. Cichon, C. R. Cloninger, D. A. Collier, E. H. Cook, H. Coon, B. Cormand, A. Corvin, W. H. Coryell, D. W. Craig, I. W. Craig, J. Crosbie, M. L. Cuccaro, D. Curtis, D. Czamara, S. Datta, G. Dawson, R. Day, E. J. De Geus, F. Degenhardt, S. Djurovic, G. J. Donohoe, A. E. Doyle, J. Duan, F. Dudbridge, E. Duketis, R. P. Ebstein, H. J. Edenberg, J. Elia, S. Ennis, B. Etain, A. Fanous, A. E. Farmer, I. N. Ferrier, M. Flickinger, E. Fombonne, T. Foroud, J. Frank, B. Franke, C. Fraser, R. Freedman, N. B. Freimer, C. M. Freitag, M. Friedl, L. Frisen, L. Gallagher, P. V. Gejman, L. Georgieva, E. S. Gershon, D. H. Geschwind, I. Giegling, M. Gill, S. D. Gordon, K. Gordon-Smith, E. K. Green, T. A. Greenwood, D. E. Grice, M. Gross, D. Grozeva, W. Guan, H. Gurling, L. De Haan, J. L. Haines, H. Hakonarson, J. Hallmayer, S. P. Hamilton, M. L. Hamshere, T. F. Hansen, A. M. Hartmann, M. Hautzinger, A. C. Heath, A. K. Henders, S. Herms, I. B. Hickie, M. Hipolito, S. Hoefels, P. A. Holmans, F. Holsboer, W. J. 269

270

Hoogendijk, J. J. Hottenga, C. M. Hultman, V. Hus, A. Ingason, M. Ising, S. Jamain, E. G. Jones, I. Jones, L. Jones, J. Y. Tzeng, A. K. Kahler, R. S. Kahn, R. Kandaswamy, M. C. Keller, J. L. Kennedy, E. Kenny, L. Kent, Y. Kim, G. K. Kirov, S. M. Klauck, L. Klei, J. A. Knowles, M. A. Kohli, D. L. Koller, B. Konte, A. Korszun, L. Krabbendam, R. Krasucki, J. Kuntsi, P. Kwan, M. Landen, N. Langstrom, M. Lathrop, J. Lawrence, W. B. Lawson, M. Leboyer, D. H. Ledbetter, P. H. Lee, T. Lencz, K. P. Lesch, D. F. Levinson, C. M. Lewis, J. Li, P. Lichtenstein, J. A. Lieberman, D. Y. Lin, D. H. Linszen, C. Liu, F. W. Lohoff, S. K. Loo, C. Lord, J. K. Lowe, S. Lucae, D. J. MacIntyre, P. A. Madden, E. Maestrini, P. K. Magnusson, P. B. Mahon, W. Maier, A. K. Malhotra, S. M. Mane, C. L. Martin, N. G. Martin, M. Mattheisen, K. Matthews, M. Mattingsdal, S. A. McCarroll, K. A. McGhee, J. J. McGough, P. J. McGrath, P. McGuffin, M. G. McInnis, A. McIntosh, R. McKinney, A. W. McLean, F. J. McMahon, W. M. McMahon, A. McQuillin, H. Medeiros, S. E. Medland, S. Meier, I. Melle, F. Meng, J. Meyer, C. M. Middeldorp, L. Middleton, V. Milanova, A. Miranda, A. P. Monaco, G. W. Montgomery, J. L. Moran, D. Moreno-De-Luca, G. Morken, D. W. Morris, E. M. Morrow, V. Moskvina, P. Muglia, T. W. Muhleisen, W. J. Muir, B. Muller-Myhsok, M. Murtha, R. M. Myers, I. Myin-Germeys, M. C. Neale, S. F. Nelson, C. M. Nievergelt, I. Nikolov, V. Nimgaonkar, W. A. Nolen, M. M. Nothen, J. I. Nurnberger, E. A. Nwulia, D. R. Nyholt, C. O'Dushlaine, R. D. Oades, A. Olincy, G. Oliveira, L. Olsen, R. A. Ophoff, U. Osby, M. J. Owen, A. Palotie, J. R. Parr, A. D. Paterson, C. N. Pato, M. T. Pato, B. W. Penninx, M. L. Pergadia, M. A. Pericak-Vance, B. S. Pickard, J. Pimm, J. Piven, D. Posthuma, J. B. Potash, F. Poustka, P. Propping, V. Puri, D. J. Quested, E. M. Quinn, J. A. Ramos-Quiroga, H. B. Rasmussen, S. Raychaudhuri, K. Rehnstrom, A. Reif, M. Ribases, J. P. Rice, M. Rietschel, K. Roeder, H. Roeyers, L. Rossin, A. Rothenberger, G. Rouleau, D. Ruderfer, D. Rujescu, A. R. Sanders, S. J. Sanders, S. L. Santangelo, J. A. Sergeant, R. Schachar, M. Schalling, A. F. Schatzberg, W. A. Scheftner, G. D. Schellenberg, S. W. Scherer, N. J. Schork, T. G. Schulze, J. Schumacher, M. Schwarz, E. Scolnick, L. J. Scott, J. Shi, P. D. Shilling, S. I. Shyn, J. M. Silverman, S. L. Slager, S. L. Smalley, J. H. Smit, E. N. Smith, E. J. Sonuga-Barke, D. St Clair, M. State, M. Steffens, H. C. Steinhausen, J. S. Strauss, J. Strohmaier, T. S. Stroup, J. S. Sutcliffe, P. Szatmari, S. Szelinger, S. Thirumalai, R. C. Thompson, A. A. Todorov, F. Tozzi, J. Treutlein, M. Uhr, E. J. van den Oord, G. Van Grootheest, J. Van Os, A. M. Vicente, V. J. Vieland, J. B. Vincent, P. M. Visscher, C. A. Walsh, T. H. Wassink, S. J. Watson, M. M. Weissman, T. Werge, T. F. Wienker, E. M. Wijsman, G.

270

271

Willemsen, N. Williams, A. J. Willsey, S. H. Witt, W. Xu, A. H. Young, T. W. Yu, S. Zammit, P. P. Zandi, P. Zhang, F. G. Zitman, S. Zollner, C. International Inflammatory Bowel Disease Genetics, B. Devlin, J. R. Kelsoe, P. Sklar, M. J. Daly, M. C. O'Donovan, N. Craddock, P. F. Sullivan, J. W. Smoller, K. S. Kendler and N. R. Wray (2013). "Genetic relationship between five psychiatric disorders estimated from genome-wide SNPs." Nat Genet 45(9): 984-994.

Crow, E. W. and J. F. Crow (2002). "100 years ago: Walter Sutton and the chromosome theory of heredity." Genetics 160(1): 1-4. da Rocha, F. F., H. Correa and A. L. Teixeira (2008). "Obsessive-compulsive disorder and immunology: a review." Prog Neuropsychopharmacol Biol Psychiatry 32(5): 1139-1146.

Dallaspezia, S., M. Mazza, C. Lorenzi, F. Benedetti and E. Smeraldi (2014). "A single nucleotide polymorphism in SLC1A1 gene is associated with age of onset of obsessive-compulsive disorder." Eur Psychiatry 29(5): 301-303. de Bartolomeis, A. and G. Fiore (2004). "Postsynaptic density scaffolding proteins at excitatory synapse and disorders of synaptic plasticity: implications for human behavior pathologies." Int Rev Neurobiol 59: 221-254.

De Geus, F. and D. Denys (2004). "Pitfalls in factor analytic techniques." Am J Psychiatry 161(3): 579-580.

De Luca, V., V. Gershenzon, E. Burroughs, N. Javaid and M. A. Richter (2011). "Age at onset in Canadian OCD patients: mixture analysis and systematic comparison with other studies." J Affect Disord 133(1-2): 300-304. de Mathis, M. A., J. B. Diniz, A. G. Hounie, R. G. Shavitt, V. Fossaluza, Y. Ferrao, J. F. Leckman, C. de Braganca Pereira, M. C. do Rosario and E. C. Miguel (2013). "Trajectory in obsessive-compulsive disorder comorbidities." Eur Neuropsychopharmacol 23(7): 594-601. de Mathis, M. A., J. B. Diniz, R. G. Shavitt, A. R. Torres, Y. A. Ferrao, V. Fossaluza, C. Pereira, E. Miguel and M. C. do Rosario (2009). "Early onset obsessive-compulsive disorder with and without tics." CNS Spectr 14(7): 362-370.

271

272

de Mathis, M. A., M. C. do Rosario, J. B. Diniz, A. R. Torres, R. G. Shavitt, Y. A. Ferrao, V. Fossaluza, C. A. de Braganca Pereira and E. C. Miguel (2008). "Obsessive-compulsive disorder: influence of age at onset on comorbidity patterns." Eur Psychiatry 23(3): 187-194.

Deary, I. J. (2012). "Intelligence." Annu Rev Psychol 63: 453-482.

Dell'Osso, B., B. Benatti, M. Buoli, A. C. Altamura, D. Marazziti, E. Hollander, N. Fineberg, D. J. Stein, S. Pallanti, H. Nicolini, M. Van Ameringen, C. Lochner, G. Hranov, O. Karamustafalioglu, L. Hranov, J. M. Menchon, J. Zohar and I. group (2013). "The influence of age at onset and duration of illness on long-term outcome in patients with obsessive-compulsive disorder: a report from the International College of Obsessive Compulsive Spectrum Disorders (ICOCS)." Eur Neuropsychopharmacol 23(8): 865-871.

Delorme, R., A. Bille, C. Betancur, F. Mathieu, N. Chabane, M. C. Mouren-Simeoni and M. Leboyer (2006). "Exploratory analysis of obsessive compulsive symptom dimensions in children and adolescents: a prospective follow-up study." BMC Psychiatry 6: 1.

Delorme, R., J. L. Golmard, N. Chabane, B. Millet, M. O. Krebs, M. C. Mouren-Simeoni and M. Leboyer (2005). "Admixture analysis of age at onset in obsessive-compulsive disorder." Psychol Med 35(2): 237-243.

Delorme, R., M. O. Krebs, N. Chabane, I. Roy, B. Millet, M. C. Mouren-Simeoni, W. Maier, T. Bourgeron and M. Leboyer (2004). "Frequency and transmission of glutamate receptors GRIK2 and GRIK3 polymorphisms in patients with obsessive compulsive disorder." Neuroreport 15(4): 699-702.

Denys, D., H. Burger, H. van Megen, F. de Geus and H. Westenberg (2003). "A score for predicting response to pharmacotherapy in obsessive-compulsive disorder." Int Clin Psychopharmacol 18(6): 315-322.

Denys, D., F. de Geus, H. J. van Megen and H. G. Westenberg (2004). "Symptom dimensions in obsessive-compulsive disorder: factor analysis on a clinician-rated scale and a self-report measure." Psychopathology 37(4): 181-189.

272

273

Denys, D., F. de Geus, H. J. van Megen and H. G. Westenberg (2004). "Use of factor analysis to detect potential phenotypes in obsessive-compulsive disorder." Psychiatry Res 128(3): 273-280.

Denys, D., F. Van Nieuwerburgh, D. Deforce and H. Westenberg (2006). "Association between the dopamine D2 receptor TaqI A2 allele and low activity COMT allele with obsessive- compulsive disorder in males." Eur Neuropsychopharmacol 16(6): 446-450.

Denys, D., F. Van Nieuwerburgh, D. Deforce and H. G. Westenberg (2006). "Association between serotonergic candidate genes and specific phenotypes of obsessive compulsive disorder." J Affect Disord 91(1): 39-44.

Desta, Z., X. Zhao, J. G. Shin and D. A. Flockhart (2002). "Clinical significance of the cytochrome P450 2C19 genetic polymorphism." Clin Pharmacokinet 41(12): 913-958.

Dickel, D. E., J. Veenstra-VanderWeele, N. C. Bivens, X. Wu, D. J. Fischer, M. Van Etten-Lee, J. A. Himle, B. L. Leventhal, E. H. Cook, Jr. and G. L. Hanna (2007). "Association studies of serotonin system candidate genes in early-onset obsessive-compulsive disorder." Biol Psychiatry 61(3): 322-329.

Dickel, D. E., J. Veenstra-VanderWeele, N. J. Cox, X. Wu, D. J. Fischer, M. Van Etten-Lee, J. A. Himle, B. L. Leventhal, E. H. Cook, Jr. and G. L. Hanna (2006). "Association testing of the positional and functional candidate gene SLC1A1/EAAC1 in early-onset obsessive-compulsive disorder." Arch Gen Psychiatry 63(7): 778-785. do Rosario-Campos, M. C., J. F. Leckman, M. Curi, S. Quatrano, L. Katsovitch, E. C. Miguel and D. L. Pauls (2005). "A family study of early-onset obsessive-compulsive disorder." Am J Med Genet B Neuropsychiatr Genet 136B(1): 92-97.

Dold, M., M. Aigner, R. Lanzenberger and S. Kasper (2013). "Antipsychotic augmentation of serotonin reuptake inhibitors in treatment-resistant obsessive-compulsive disorder: a meta- analysis of double-blind, randomized, placebo-controlled trials." Int J Neuropsychopharmacol 16(3): 557-574.

Dudbridge, F. (2008). "Likelihood-based association analysis for nuclear families and unrelated subjects with missing genotype data." Hum Hered 66(2): 87-98. 273

274

Eaton, M. J., J. K. Staley, M. Y. Globus and S. R. Whittemore (1995). "Developmental regulation of early serotonergic neuronal differentiation: the role of brain-derived neurotrophic factor and membrane depolarization." Dev Biol 170(1): 169-182.

Egan, M. F., T. E. Goldberg, B. S. Kolachana, J. H. Callicott, C. M. Mazzanti, R. E. Straub, D. Goldman and D. R. Weinberger (2001). "Effect of COMT Val108/158 Met genotype on frontal lobe function and risk for schizophrenia." Proc Natl Acad Sci U S A 98(12): 6917-6922.

Egan, M. F., M. Kojima, J. H. Callicott, T. E. Goldberg, B. S. Kolachana, A. Bertolino, E. Zaitsev, B. Gold, D. Goldman, M. Dean, B. Lu and D. R. Weinberger (2003). "The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function." Cell 112(2): 257-269.

Elias, G. M. and R. A. Nicoll (2007). "Synaptic trafficking of glutamate receptors by MAGUK scaffolding proteins." Trends Cell Biol 17(7): 343-352.

Endele, S., G. Rosenberger, K. Geider, B. Popp, C. Tamer, I. Stefanova, M. Milh, F. Kortum, A. Fritsch, F. K. Pientka, Y. Hellenbroich, V. M. Kalscheuer, J. Kohlhase, U. Moog, G. Rappold, A. Rauch, H. H. Ropers, S. von Spiczak, H. Tonnies, N. Villeneuve, L. Villard, B. Zabel, M. Zenker, B. Laube, A. Reis, D. Wieczorek, L. Van Maldergem and K. Kutsche (2010). "Mutations in GRIN2A and GRIN2B encoding regulatory subunits of NMDA receptors cause variable neurodevelopmental phenotypes." Nat Genet 42(11): 1021-1026.

Enoch, M. A., B. D. Greenberg, D. L. Murphy and D. Goldman (2001). "Sexually dimorphic relationship of a 5-HT2A promoter polymorphism with obsessive-compulsive disorder." Biol Psychiatry 49(4): 385-388.

Enoch, M. A., W. H. Kaye, A. Rotondo, B. D. Greenberg, D. L. Murphy and D. Goldman (1998). "5-HT2A promoter polymorphism -1438G/A, anorexia nervosa, and obsessive- compulsive disorder." Lancet 351(9118): 1785-1786.

Etkin, A. and B. Cuthbert (2014). "Beyond the DSM: development of a transdiagnostic psychiatric neuroscience course." Acad Psychiatry 38(2): 145-150.

274

275

Fabbri, C., G. Di Girolamo and A. Serretti (2013). "Pharmacogenetics of antidepressant drugs: an update after almost 20 years of research." Am J Med Genet B Neuropsychiatr Genet 162B(6): 487-520.

Faragian, S., A. Pashinian, C. Fuchs and M. Poyurovsky (2009). "Obsessive-compulsive symptom dimensions in schizophrenia patients with comorbid obsessive-compulsive disorder." Prog Neuropsychopharmacol Biol Psychiatry 33(6): 1009-1012.

Fargher, E. A., C. Eddy, W. Newman, F. Qasim, K. Tricker, R. A. Elliott and K. Payne (2007). "Patients' and healthcare professionals' views on pharmacogenetic testing and its future delivery in the NHS." Pharmacogenomics 8(11): 1511-1519.

Faul, F., E. Erdfelder, A. Buchner and A. G. Lang (2009). "Statistical power analyses using G*Power 3.1: tests for correlation and regression analyses." Behav Res Methods 41(4): 1149- 1160.

Faul, F., E. Erdfelder, A. G. Lang and A. Buchner (2007). "G*Power 3: a flexible statistical power analysis program for the social, behavioral, and biomedical sciences." Behav Res Methods 39(2): 175-191.

Feinstein, S. B., B. A. Fallon, E. Petkova and M. R. Liebowitz (2003). "Item-by-item factor analysis of the Yale-Brown Obsessive Compulsive Scale Symptom Checklist." J Neuropsychiatry Clin Neurosci 15(2): 187-193.

Fernandez, M., M. F. Segura, C. Sole, A. Colino, J. X. Comella and V. Cena (2007). "Lifeguard/neuronal membrane protein 35 regulates Fas ligand-mediated apoptosis in neurons via microdomain recruitment." J Neurochem 103(1): 190-203.

Figueira, M. L. and S. Ouakinin (2010). "Gender-related endocrinological dysfunction and mental disorders." Curr Opin Psychiatry 23(4): 369-372.

Fineberg, N. A. and T. M. Gale (2005). "Evidence-based pharmacotherapy of obsessive- compulsive disorder." Int J Neuropsychopharmacol 8(1): 107-129.

275

276

Fineberg, N. A., M. P. Hengartner, C. E. Bergbaum, T. M. Gale, A. Gamma, V. Ajdacic-Gross, W. Rossler and J. Angst (2013). "A prospective population-based cohort study of the prevalence, incidence and impact of obsessive-compulsive symptomatology." Int J Psychiatry Clin Pract 17(3): 170-178.

Fineberg, N. A., S. Reghunandanan, H. B. Simpson, K. A. Phillips, M. A. Richter, K. Matthews, D. J. Stein, J. Sareen, A. Brown, D. Sookman and D. Accreditation Task Force of The Canadian Institute for Obsessive Compulsive (2015). "Obsessive-compulsive disorder (OCD): Practical strategies for pharmacological and somatic treatment in adults." Psychiatry Res 227(1): 114-125.

First, M. B., M. Gibbon, R. L. Spitzer and J. B. W. Williams (1996). Structured Clinical Interview for DSM-IV Axis I disorders - Research Version (SCID-I/P, version 2.0, February, Final Version). Washington, DC, American Psychiatric Press.

Flockhart, D. A. (2007). "Drug interactions: cytochrome P450 drug interaction table." Retrieved January 10, 2014, from http://medicine.iupui.edu/clinpharm/ddis/clinical-table/.

Foa, E. B., J. D. Huppert, S. Leiberg, R. Langner, R. Kichic, G. Hajcak and P. M. Salkovskis (2002). "The Obsessive-Compulsive Inventory: development and validation of a short version." Psychol Assess 14(4): 485-496.

Foa, E. B., M. R. Liebowitz, M. J. Kozak, S. Davies, R. Campeas, M. E. Franklin, J. D. Huppert, K. Kjernisted, V. Rowan, A. B. Schmidt, H. B. Simpson and X. Tu (2005). "Randomized, placebo-controlled trial of exposure and ritual prevention, clomipramine, and their combination in the treatment of obsessive-compulsive disorder." Am J Psychiatry 162(1): 151-161.

Fontenelle, L. F., I. G. Barbosa, J. V. Luna, N. P. Rocha, A. Silva Miranda and A. L. Teixeira (2012). "Neurotrophic factors in obsessive-compulsive disorder." Psychiatry Res 199(3): 195- 200.

Fontenelle, L. F. and G. Hasler (2008). "The analytical epidemiology of obsessive-compulsive disorder: risk factors and correlates." Prog Neuropsychopharmacol Biol Psychiatry 32(1): 1-15.

276

277

Fontenelle, L. F., M. V. Mendlowicz, C. Marques and M. Versiani (2003). "Early- and late-onset obsessive-compulsive disorder in adult patients: an exploratory clinical and therapeutic study." J Psychiatr Res 37(2): 127-133.

Fontenelle, L. F., M. V. Mendlowicz and M. Versiani (2005). "Clinical subtypes of obsessive- compulsive disorder based on the presence of checking and washing compulsions." Rev Bras Psiquiatr 27(3): 201-207.

Fulker, D. W., S. S. Cherny, P. C. Sham and J. K. Hewitt (1999). "Combined linkage and association sib-pair analysis for quantitative traits." Am J Hum Genet 64(1): 259-267.

Fullana, M. A., P. Alonso, M. Gratacos, N. Jaurrieta, S. Jimenez-Murcia, C. Segalas, E. Real, X. Estivill and J. M. Menchon (2012). "Variation in the BDNF Val66Met polymorphism and response to cognitive-behavior therapy in obsessive-compulsive disorder." Eur Psychiatry 27(5): 386-390.

Fullana, M. A., D. Mataix-Cols, A. Caspi, H. Harrington, J. R. Grisham, T. E. Moffitt and R. Poulton (2009). "Obsessions and compulsions in the community: prevalence, interference, help- seeking, developmental stability, and co-occurring psychiatric conditions." Am J Psychiatry 166(3): 329-336.

Fullana, M. A., G. Vilagut, S. Rojas-Farreras, D. Mataix-Cols, R. de Graaf, K. Demyttenaere, J. M. Haro, G. de Girolamo, J. P. Lepine, H. Matschinger, J. Alonso and E. S. M. investigators (2010). "Obsessive-compulsive symptom dimensions in the general population: results from an epidemiological study in six European countries." J Affect Disord 124(3): 291-299.

Garcia, C., M. J. Chen, A. A. Garza, C. W. Cotman and A. Russo-Neustadt (2003). "The influence of specific noradrenergic and serotonergic lesions on the expression of hippocampal brain-derived neurotrophic factor transcripts following voluntary physical activity." Neuroscience 119(3): 721-732.

Garcia-Soriano, G., A. Belloch, C. Morillo and D. A. Clark (2011). "Symptom dimensions in obsessive-compulsive disorder: from normal cognitive intrusions to clinical obsessions." J Anxiety Disord 25(4): 474-482.

277

278

Gauderman, W. J. and J. M. Morrison. (2006). "QUANTO 1.1: A computer program for power and sample size calculations for genetic-epidemiology studies." from http://hydra.usc.edu/gxe.

Georgieva, L., V. Moskvina, T. Peirce, N. Norton, N. J. Bray, L. Jones, P. Holmans, S. Macgregor, S. Zammit, J. Wilkinson, H. Williams, I. Nikolov, N. Williams, D. Ivanov, K. L. Davis, V. Haroutunian, J. D. Buxbaum, N. Craddock, G. Kirov, M. J. Owen and M. C. O'Donovan (2006). "Convergent evidence that oligodendrocyte lineage transcription factor 2 (OLIG2) and interacting genes influence susceptibility to schizophrenia." Proc Natl Acad Sci U S A 103(33): 12469-12474.

Girishchandra, B. G. and S. Khanna (2001). "Phenomenology of obsessive compulsive disorder: a factor analytic approach." Indian J Psychiatry 43(4): 306-316.

Glahn, D. C., E. E. Knowles, D. R. McKay, E. Sprooten, H. Raventos, J. Blangero, I. I. Gottesman and L. Almasy (2014). "Arguments for the sake of endophenotypes: examining common misconceptions about the use of endophenotypes in psychiatric genetics." Am J Med Genet B Neurospsychiatr Genet 165B(2): 122-130.

Gomes de Alvarenga, P., M. A. de Mathis, A. C. Dominguez Alves, M. C. do Rosario, V. Fossaluza, A. G. Hounie, E. C. Miguel and A. Rodrigues Torres (2012). "Clinical features of tic- related obsessive-compulsive disorder: results from a large multicenter study." CNS Spectr 17(2): 87-93.

Goodman, W. K., L. H. Price, S. A. Rasmussen, C. Mazure, P. Delgado, G. R. Heninger and D. S. Charney (1989). "The Yale-Brown Obsessive Compulsive Scale. II. Validity." Arch Gen Psychiatry 46(11): 1012-1016.

Goodman, W. K., L. H. Price, S. A. Rasmussen, C. Mazure, R. L. Fleischmann, C. L. Hill, G. R. Heninger and D. S. Charney (1989). "The Yale-Brown Obsessive Compulsive Scale. I. Development, use, and reliability." Arch Gen Psychiatry 46(11): 1006-1011.

Gottesman, I. I. and T. D. Gould (2003). "The endophenotype concept in psychiatry: etymology and strategic intentions." Am J Psychiatry 160(4): 636-645.

278

279

Grados, M. A., M. W. Specht, H. M. Sung and D. Fortune (2013). "Glutamate drugs and pharmacogenetics of OCD: a pathway-based exploratory approach." Expert Opin Drug Discov 8(12): 1515-1527.

Grant, J. E., M. C. Mancebo, A. Pinto, K. A. Williams, J. L. Eisen and S. A. Rasmussen (2007). "Late-onset obsessive compulsive disorder: clinical characteristics and psychiatric comorbidity." Psychiatry Res 152(1): 21-27.

Grant, J. E. and B. L. Odlaug (2009). "Update on pathological skin picking." Curr Psychiatry Rep 11(4): 283-288.

Gray, S. M. and M. H. Bloch (2012). "Systematic review of proinflammatory cytokines in obsessive-compulsive disorder." Curr Psychiatry Rep 14(3): 220-228.

Graybiel, A. M. and S. L. Rauch (2000). "Toward a neurobiology of obsessive-compulsive disorder." Neuron 28(2): 343-347.

Greer, J. M. and M. R. Capecchi (2002). "Hoxb8 is required for normal grooming behavior in mice." Neuron 33(1): 23-34.

Greist, J. H., J. W. Jefferson, K. A. Kobak, D. J. Katzelnick and R. C. Serlin (1995). "Efficacy and tolerability of serotonin transport inhibitors in obsessive-compulsive disorder. A meta- analysis." Arch Gen Psychiatry 52(1): 53-60.

Grossman, M. H., B. S. Emanuel and M. L. Budarf (1992). "Chromosomal mapping of the human catechol-O-methyltransferase gene to 22q11.1----q11.2." Genomics 12(4): 822-825.

Grunblatt, E., T. U. Hauser and S. Walitza (2014). "Imaging genetics in obsessive-compulsive disorder: linking genetic variations to alterations in neuroimaging." Prog Neurobiol 121: 114- 124.

Guillin, O., J. Diaz, P. Carroll, N. Griffon, J. C. Schwartz and P. Sokoloff (2001). "BDNF controls dopamine D3 receptor expression and triggers behavioural sensitization." Nature 411(6833): 86-89.

279

280

Gvozdic, K., E. J. Brandl, D. L. Taylor and D. J. Muller (2012). "Genetics and personalized medicine in antidepressant treatment." Curr Pharm Des 18(36): 5853-5878.

Haga, S. B. and W. Burke (2008). "Pharmacogenetic testing: not as simple as it seems." Genet Med 10(6): 391-395.

Haga, S. B., G. Tindall and J. M. O'Daniel (2012). "Public perspectives about pharmacogenetic testing and managing ancillary findings." Genet Test Mol Biomarkers 16(3): 193-197.

Hall, D., A. Dhilla, A. Charalambous, J. A. Gogos and M. Karayiorgou (2003). "Sequence variants of the brain-derived neurotrophic factor (BDNF) gene are strongly associated with obsessive-compulsive disorder." Am J Hum Genet 73(2): 370-376.

Hanna, G. L., D. J. Fischer, K. R. Chadha, J. A. Himle and M. Van Etten (2005). "Familial and sporadic subtypes of early-onset Obsessive-Compulsive disorder." Biol Psychiatry 57(8): 895- 900.

Hanna, G. L., J. A. Himle, G. C. Curtis and B. W. Gillespie (2005). "A family study of obsessive-compulsive disorder with pediatric probands." Am J Med Genet B Neuropsychiatr Genet 134B(1): 13-19.

Hanna, G. L., J. Veenstra-VanderWeele, N. J. Cox, M. Boehnke, J. A. Himle, G. C. Curtis, B. L. Leventhal and E. H. Cook, Jr. (2002). "Genome-wide linkage analysis of families with obsessive-compulsive disorder ascertained through pediatric probands." Am J Med Genet 114(5): 541-552.

Hanson, I. M., A. Seawright and V. van Heyningen (1992). "The human BDNF gene maps between FSHB and HVBS1 at the boundary of 11p13-p14." Genomics 13(4): 1331-1333.

Hardingham, G. E., Y. Fukunaga and H. Bading (2002). "Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways." Nat Neurosci 5(5): 405-414.

280

281

Hartley, C. A., M. C. McKenna, R. Salman, A. Holmes, B. J. Casey, E. A. Phelps and C. E. Glatt (2012). "Serotonin transporter polyadenylation polymorphism modulates the retention of fear extinction memory." Proc Natl Acad Sci U S A 109(14): 5493-5498.

Hasler, G., D. Kazuba and D. L. Murphy (2006). "Factor analysis of obsessive-compulsive disorder YBOCS-SC symptoms and association with 5-HTTLPR SERT polymorphism." Am J Med Genet B Neuropsychiatr Genet 141B(4): 403-408.

Hasler, G., V. H. LaSalle-Ricci, J. G. Ronquillo, S. A. Crawley, L. W. Cochran, D. Kazuba, B. D. Greenberg and D. L. Murphy (2005). "Obsessive-compulsive disorder symptom dimensions show specific relationships to psychiatric comorbidity." Psychiatry Res 135(2): 121-132.

Hasler, G., A. Pinto, B. D. Greenberg, J. Samuels, A. J. Fyer, D. Pauls, J. A. Knowles, J. T. McCracken, J. Piacentini, M. A. Riddle, S. L. Rauch, S. A. Rasmussen, V. L. Willour, M. A. Grados, B. Cullen, O. J. Bienvenu, Y. Y. Shugart, K. Y. Liang, R. Hoehn-Saric, Y. Wang, J. Ronquillo, G. Nestadt, D. L. Murphy and O. C. D. C. G. Study (2007). "Familiality of factor analysis-derived YBOCS dimensions in OCD-affected sibling pairs from the OCD Collaborative Genetics Study." Biol Psychiatry 61(5): 617-625.

Hazra, A., P. Kraft, J. Selhub, E. L. Giovannucci, G. Thomas, R. N. Hoover, S. J. Chanock and D. J. Hunter (2008). "Common variants of FUT2 are associated with plasma vitamin B12 levels." Nat Genet 40(10): 1160-1162.

Hemmings, S. M., C. J. Kinnear, C. Lochner, D. J. Niehaus, J. A. Knowles, J. C. Moolman- Smook, V. A. Corfield and D. J. Stein (2004). "Early- versus late-onset obsessive-compulsive disorder: investigating genetic and clinical correlates." Psychiatry Res 128(2): 175-182.

Hemmings, S. M., C. J. Kinnear, L. Van der Merwe, C. Lochner, V. A. Corfield, J. C. Moolman- Smook and D. J. Stein (2008). "Investigating the role of the brain-derived neurotrophic factor (BDNF) val66met variant in obsessive-compulsive disorder (OCD)." World J Biol Psychiatry 9(2): 126-134.

281

282

Hemmings, S. M., C. Lochner, L. van der Merwe, D. C. Cath, S. Seedat and D. J. Stein (2013). "BDNF Val66Met modifies the risk of childhood trauma on obsessive-compulsive disorder." J Psychiatr Res 47(12): 1857-1863.

Hemmings, S. M. J. (2006). Investigating the Molecular Aetiology of Obsessive-Compulsive Disorder (OCD) and Clinically-defined Subsets of OCD, University of Stellenbosch.

Hermesh, H., A. Weizman, A. Shahar and H. Munitz (1988). "Vitamin B12 and folic acid serum levels in obsessive compulsive disorder." Acta Psychiatr Scand 78(1): 8-10.

Hettema, J. M., M. C. Neale and K. S. Kendler (2001). "A review and meta-analysis of the genetic epidemiology of anxiety disorders." Am J Psychiatry 158(10): 1568-1578.

Hicks, J. K., J. J. Swen, C. F. Thorn, K. Sangkuhl, E. D. Kharasch, V. L. Ellingrod, T. C. Skaar, D. J. Muller, A. Gaedigk, J. C. Stingl and C. Clinical Pharmacogenetics Implementation (2013). "Clinical Pharmacogenetics Implementation Consortium guideline for CYP2D6 and CYP2C19 genotypes and dosing of tricyclic antidepressants." Clin Pharmacol Ther 93(5): 402-408.

Hinton, D. E., L. Park, C. Hsia, S. Hofmann and M. H. Pollack (2009). "Anxiety disorder presentations in Asian populations: a review." CNS Neurosci Ther 15(3): 295-303.

Hodgson, R. J. and S. Rachman (1977). "Obsessional-compulsive complaints." Behav Res Ther 15(5): 389-395.

Hofer, M. M. and Y. A. Barde (1988). "Brain-derived neurotrophic factor prevents neuronal death in vivo." Nature 331(6153): 261-262.

Hofmeijer-Sevink, M. K., P. van Oppen, H. J. van Megen, N. M. Batelaan, D. C. Cath, N. J. van der Wee, M. A. van den Hout and A. J. van Balkom (2013). "Clinical relevance of comorbidity in obsessive compulsive disorder: the Netherlands OCD Association study." J Affect Disord 150(3): 847-854.

Holterhoff, K. (2014). "The history and reception of Charles Darwin's hypothesis of pangenesis." J Hist Biol 47(4): 661-695.

282

283

Hoop, J. G., M. I. Lapid, R. M. Paulson and L. W. Roberts (2010). "Clinical and ethical considerations in pharmacogenetic testing: views of physicians in 3 "early adopting" departments of psychiatry." J Clin Psychiatry 71(6): 745-753.

Horiuchi, Y., S. Iida, M. Koga, H. Ishiguro, Y. Iijima, T. Inada, Y. Watanabe, T. Someya, H. Ujike, N. Iwata, N. Ozaki, H. Kunugi, M. Tochigi, M. Itokawa, M. Arai, K. Niizato, S. Iritani, A. Kakita, H. Takahashi, H. Nawa and T. Arinami (2012). "Association of SNPs linked to increased expression of SLC1A1 with schizophrenia." Am J Med Genet B Neuropsychiatr Genet 159B(1): 30-37.

Hu, X. Z., R. H. Lipsky, G. Zhu, L. A. Akhtar, J. Taubman, B. D. Greenberg, K. Xu, P. D. Arnold, M. A. Richter, J. L. Kennedy, D. L. Murphy and D. Goldman (2006). "Serotonin transporter promoter gain-of-function genotypes are linked to obsessive-compulsive disorder." Am J Hum Genet 78(5): 815-826.

Husted, D. S., N. A. Shapira, T. K. Murphy, G. D. Mann, H. E. Ward and W. K. Goodman (2007). "Effect of comorbid tics on a clinically meaningful response to 8-week open-label trial of fluoxetine in obsessive compulsive disorder." J Psychiatr Res 41(3-4): 332-337.

Iervolino, A. C., F. V. Rijsdijk, L. Cherkas, M. A. Fullana and D. Mataix-Cols (2011). "A multivariate twin study of obsessive-compulsive symptom dimensions." Arch Gen Psychiatry 68(6): 637-644.

Informatics, I. I. f. H. (2012). "The use of medicines in the United States: review of 2011." Retrieved October 15, 2013, from http://www.imshealth.com/ims/Global/Content/Insights/IMS%20Institute%20for%20Healthcare %20Informatics/IHII_Medicines_in_U.S_Report_2011.pdf.

International HapMap, C. (2003). "The International HapMap Project." Nature 426(6968): 789- 796.

International HapMap, C., D. M. Altshuler, R. A. Gibbs, L. Peltonen, D. M. Altshuler, R. A. Gibbs, L. Peltonen, E. Dermitzakis, S. F. Schaffner, F. Yu, L. Peltonen, E. Dermitzakis, P. E. Bonnen, D. M. Altshuler, R. A. Gibbs, P. I. de Bakker, P. Deloukas, S. B. Gabriel, R. Gwilliam,

283

284

S. Hunt, M. Inouye, X. Jia, A. Palotie, M. Parkin, P. Whittaker, F. Yu, K. Chang, A. Hawes, L. R. Lewis, Y. Ren, D. Wheeler, R. A. Gibbs, D. M. Muzny, C. Barnes, K. Darvishi, M. Hurles, J. M. Korn, K. Kristiansson, C. Lee, S. A. McCarrol, J. Nemesh, E. Dermitzakis, A. Keinan, S. B. Montgomery, S. Pollack, A. L. Price, N. Soranzo, P. E. Bonnen, R. A. Gibbs, C. Gonzaga- Jauregui, A. Keinan, A. L. Price, F. Yu, V. Anttila, W. Brodeur, M. J. Daly, S. Leslie, G. McVean, L. Moutsianas, H. Nguyen, S. F. Schaffner, Q. Zhang, M. J. Ghori, R. McGinnis, W. McLaren, S. Pollack, A. L. Price, S. F. Schaffner, F. Takeuchi, S. R. Grossman, I. Shlyakhter, E. B. Hostetter, P. C. Sabeti, C. A. Adebamowo, M. W. Foster, D. R. Gordon, J. Licinio, M. C. Manca, P. A. Marshall, I. Matsuda, D. Ngare, V. O. Wang, D. Reddy, C. N. Rotimi, C. D. Royal, R. R. Sharp, C. Zeng, L. D. Brooks and J. E. McEwen (2010). "Integrating common and rare genetic variation in diverse human populations." Nature 467(7311): 52-58.

Investigators, G., M. Investigators and S. D. Investigators (2013). "Common genetic variation and antidepressant efficacy in major depressive disorder: a meta-analysis of three genome-wide pharmacogenetic studies." Am J Psychiatry 170(2): 207-217.

Jaisoorya, T. S., Y. C. Reddy, S. Srinath and K. Thennarasu (2009). "Sex differences in Indian patients with obsessive-compulsive disorder." Compr Psychiatry 50(1): 70-75.

Jakubovski, E., J. B. Diniz, C. Valerio, V. Fossaluza, C. Belotto-Silva, C. Gorenstein, E. Miguel and R. G. Shavitt (2013). "Clinical predictors of long-term outcome in obsessive-compulsive disorder." Depress Anxiety 30(8): 763-772.

Jamain, S., C. Betancur, H. Quach, A. Philippe, M. Fellous, B. Giros, C. Gillberg, M. Leboyer, T. Bourgeron and S. Paris Autism Research International Sibpair (2002). "Linkage and association of the glutamate receptor 6 gene with autism." Mol Psychiatry 7(3): 302-310.

Jang, J. H., H. S. Kim, T. H. Ha, N. Y. Shin, D. H. Kang, J. S. Choi, K. Ha and J. S. Kwon (2010). "Nonverbal memory and organizational dysfunctions are related with distinct symptom dimensions in obsessive-compulsive disorder." Psychiatry Res 180(2-3): 93-98.

Jann, M. W. and L. J. Cohen (2000). "The influence of ethnicity and antidepressant pharmacogenetics in the treatment of depression." Drug Metabol Drug Interact 16(1): 39-67.

284

285

Janowitz, D., H. J. Grabe, S. Ruhrmann, S. Ettelt, F. Buhtz, A. Hochrein, S. Schulze- Rauschenbach, K. Meyer, S. Kraft, C. Ferber, R. Pukrop, H. J. Freyberger, J. Klosterkotter, P. Falkai, U. John, W. Maier and M. Wagner (2009). "Early onset of obsessive-compulsive disorder and associated comorbidity." Depress Anxiety 26(11): 1012-1017.

Jansson, M., S. McCarthy, P. F. Sullivan, P. Dickman, B. Andersson, L. Oreland, M. Schalling and N. L. Pedersen (2005). "MAOA haplotypes associated with thrombocyte-MAO activity." BMC Genet 6: 46.

Jenike, M. A., L. Baer and W. E. Minichiello (1986). Obsessive Compulsive Disorders: Theory and Management. Littleton, MA, PSG Publishing.

Jenike, M. A., H. C. Breiter, L. Baer, D. N. Kennedy, C. R. Savage, M. J. Olivares, R. L. O'Sullivan, D. M. Shera, S. L. Rauch, N. Keuthen, B. R. Rosen, V. S. Caviness and P. A. Filipek (1996). "Cerebral structural abnormalities in obsessive-compulsive disorder. A quantitative morphometric magnetic resonance imaging study." Arch Gen Psychiatry 53(7): 625-632.

Jin, H., D. Oksenberg, A. Ashkenazi, S. J. Peroutka, A. M. Duncan, R. Rozmahel, Y. Yang, G. Mengod, J. M. Palacios and B. F. O'Dowd (1992). "Characterization of the human 5- hydroxytryptamine1B receptor." J Biol Chem 267(9): 5735-5738.

Jin Lee, K., Y. Wook Shin, H. Wee, Y. Youn Kim and J. S. Kwon (2006). "Gray matter volume reduction in obsessive-compulsive disorder with schizotypal personality trait." Prog Neuropsychopharmacol Biol Psychiatry 30(6): 1146-1149.

Johns, T. G. and C. C. Bernard (1999). "The structure and function of myelin oligodendrocyte glycoprotein." J Neurochem 72(1): 1-9.

Jones, K. A., D. P. Srivastava, J. A. Allen, R. T. Strachan, B. L. Roth and P. Penzes (2009). "Rapid modulation of spine morphology by the 5-HT2A serotonin receptor through kalirin-7 signaling." Proc Natl Acad Sci U S A 106(46): 19575-19580.

Jones, K. R. and L. F. Reichardt (1990). "Molecular cloning of a human gene that is a member of the nerve growth factor family." Proc Natl Acad Sci U S A 87(20): 8060-8064.

285

286

Kanai, Y. and M. A. Hediger (2004). "The glutamate/neutral amino acid transporter family SLC1: molecular, physiological and pharmacological aspects." Pflugers Arch 447(5): 469-479.

Kantrowitz, J. and D. C. Javitt (2012). "Glutamatergic transmission in schizophrenia: from basic research to clinical practice." Curr Opin Psychiatry 25(2): 96-102.

Karayiorgou, M., M. Altemus, B. L. Galke, D. Goldman, D. L. Murphy, J. Ott and J. A. Gogos (1997). "Genotype determining low catechol-O-methyltransferase activity as a risk factor for obsessive-compulsive disorder." Proc Natl Acad Sci U S A 94(9): 4572-4575.

Karayiorgou, M., C. Sobin, M. L. Blundell, B. L. Galke, L. Malinova, P. Goldberg, J. Ott and J. A. Gogos (1999). "Family-based association studies support a sexually dimorphic effect of COMT and MAOA on genetic susceptibility to obsessive-compulsive disorder." Biol Psychiatry 45(9): 1178-1189.

Kariuki-Nyuthe, C., B. Gomez-Mancilla and D. J. Stein (2014). "Obsessive compulsive disorder and the glutamatergic system." Curr Opin Psychiatry 27(1): 32-37.

Katerberg, H., D. C. Cath, D. A. Denys, P. Heutink, A. Polman, F. C. van Nieuwerburgh, D. L. Deforce, Z. Bochdanovits, A. J. van Balkom and J. A. den Boer (2010). "The role of the COMT Val(158)Met polymorphism in the phenotypic expression of obsessive-compulsive disorder." Am J Med Genet B Neuropsychiatr Genet 153B(1): 167-176.

Katerberg, H., K. L. Delucchi, S. E. Stewart, C. Lochner, D. A. Denys, D. E. Stack, J. M. Andresen, J. E. Grant, S. W. Kim, K. A. Williams, J. A. den Boer, A. J. van Balkom, J. H. Smit, P. van Oppen, A. Polman, M. A. Jenike, D. J. Stein, C. A. Mathews and D. C. Cath (2010). "Symptom dimensions in OCD: item-level factor analysis and heritability estimates." Behav Genet 40(4): 505-517.

Katerberg, H., C. Lochner, D. C. Cath, P. de Jonge, Z. Bochdanovits, J. C. Moolman-Smook, S. M. Hemmings, P. D. Carey, D. J. Stein, D. Sondervan, J. A. Boer, A. J. van Balkom, A. Polman and P. Heutink (2009). "The role of the brain-derived neurotrophic factor (BDNF) val66met variant in the phenotypic expression of obsessive-compulsive disorder (OCD)." Am J Med Genet B Neuropsychiatr Genet 150B(8): 1050-1062.

286

287

Kato, M. and A. Serretti (2010). "Review and meta-analysis of antidepressant pharmacogenetic findings in major depressive disorder." Mol Psychiatry 15(5): 473-500.

Katzman, M. A., P. Bleau, P. Blier, P. Chokka, K. Kjernisted, M. Van Ameringen, a. Canadian Anxiety Guidelines Initiative Group on behalf of the Anxiety Disorders Association of Canada/Association Canadienne des troubles, U. McGill, M. M. Antony, S. Bouchard, A. Brunet, M. Flament, S. Grigoriadis, S. Mendlowitz, K. O'Connor, K. Rabheru, P. M. Richter, M. Robichaud and J. R. Walker (2014). "Canadian clinical practice guidelines for the management of anxiety, posttraumatic stress and obsessive-compulsive disorders." BMC Psychiatry 14 Suppl 1: S1.

Kavanagh, D. H., S. Dwyer, M. C. O'Donovan and M. J. Owen (2013). "The ENCODE project: implications for psychiatric genetics." Mol Psychiatry 18(5): 540-542.

Kendler, K. S., M. Gatz, C. O. Gardner and N. L. Pedersen (2007). "Clinical indices of familial depression in the Swedish Twin Registry." Acta Psychiatr Scand 115(3): 214-220.

Kessler, R. C., W. T. Chiu, O. Demler, K. R. Merikangas and E. E. Walters (2005). "Prevalence, severity, and comorbidity of 12-month DSM-IV disorders in the National Comorbidity Survey Replication." Arch Gen Psychiatry 62(6): 617-627.

Kessler, R. C., K. A. McGonagle, S. Zhao, C. B. Nelson, M. Hughes, S. Eshleman, H. U. Wittchen and K. S. Kendler (1994). "Lifetime and 12-month prevalence of DSM-III-R psychiatric disorders in the United States. Results from the National Comorbidity Survey." Arch Gen Psychiatry 51(1): 8-19.

Keuthen, N. J., C. Fraim, T. Deckersbach, D. D. Dougherty, L. Baer and M. A. Jenike (2001). "Longitudinal follow-up of naturalistic treatment outcome in patients with trichotillomania." J Clin Psychiatry 62(2): 101-107.

Kieling, C., H. Baker-Henningham, M. Belfer, G. Conti, I. Ertem, O. Omigbodun, L. A. Rohde, S. Srinath, N. Ulkuer and A. Rahman (2011). "Child and adolescent mental health worldwide: evidence for action." Lancet 378(9801): 1515-1525.

287

288

Kim, E., S. Naisbitt, Y. P. Hsueh, A. Rao, A. Rothschild, A. M. Craig and M. Sheng (1997). "GKAP, a novel synaptic protein that interacts with the guanylate kinase-like domain of the PSD-95/SAP90 family of channel clustering molecules." J Cell Biol 136(3): 669-678.

Kim, S. A., J. H. Kim, M. Park, I. H. Cho and H. J. Yoo (2007). "Family-based association study between GRIK2 polymorphisms and autism spectrum disorders in the Korean trios." Neurosci Res 58(3): 332-335.

Kim, S. J., H. S. Lee and C. H. Kim (2005). "Obsessive-compulsive disorder, factor-analyzed symptom dimensions and serotonin transporter polymorphism." Neuropsychobiology 52(4): 176- 182.

Kim, S. J., K. Namkoong, J. I. Kang and C. H. Kim (2009). "Association of a 5-HT1Dbeta receptor gene polymorphism with obsessive-compulsive disorder in Korean male subjects." Neuropsychobiology 59(2): 96-99.

Kinnear, C., D. J. Niehaus, S. Seedat, J. C. Moolman-Smook, V. A. Corfield, G. Malherbe, A. Potgieter, C. Lombard and D. J. Stein (2001). "Obsessive-compulsive disorder and a novel polymorphism adjacent to the oestrogen response element (ERE 6) upstream from the COMT gene." Psychiatr Genet 11(2): 85-87.

Klaffke, S., I. R. Konig, F. Poustka, A. Ziegler, J. Hebebrand and O. Bandmann (2006). "Brain- derived neurotrophic factor: a genetic risk factor for obsessive-compulsive disorder and Tourette syndrome?" Mov Disord 21(6): 881-883.

Koch, K., T. J. Reess, O. G. Rus, C. Zimmer and M. Zaudig (2014). "Diffusion tensor imaging (DTI) studies in patients with obsessive-compulsive disorder (OCD): a review." J Psychiatr Res 54: 26-35.

Kolenikov, S. (2001). "DENORMIX: Stata module to perform decomposition of normal mixture." from http://ideas.repec.org/c/boc/bocode/s416605.html.

Koo, M. S., E. J. Kim, D. Roh and C. H. Kim (2010). "Role of dopamine in the pathophysiology and treatment of obsessive-compulsive disorder." Expert Rev Neurother 10(2): 275-290.

288

289

Koran, L. M., G. L. Hanna, E. Hollander, G. Nestadt, H. B. Simpson and A. American Psychiatric (2007). "Practice guideline for the treatment of patients with obsessive-compulsive disorder." Am J Psychiatry 164(7 Suppl): 5-53.

Kornstein, S. G. (1997). "Gender differences in depression: implications for treatment." J Clin Psychiatry 58 Suppl 15: 12-18.

Krapivinsky, G., I. Medina, L. Krapivinsky, S. Gapon and D. E. Clapham (2004). "SynGAP- MUPP1-CaMKII synaptic complexes regulate p38 MAP kinase activity and NMDA receptor- dependent synaptic AMPA receptor potentiation." Neuron 43(4): 563-574.

Krebs, M. O., O. Guillin, M. C. Bourdell, J. C. Schwartz, J. P. Olie, M. F. Poirier and P. Sokoloff (2000). "Brain derived neurotrophic factor (BDNF) gene variants association with age at onset and therapeutic response in schizophrenia." Mol Psychiatry 5(5): 558-562.

Krupp, M., A. Weinmann, P. R. Galle and A. Teufel (2006). "Actin binding LIM protein 3 (abLIM3)." Int J Mol Med 17(1): 129-133.

Krystal, J. H., G. Sanacora, H. Blumberg, A. Anand, D. S. Charney, G. Marek, C. N. Epperson, A. Goddard and G. F. Mason (2002). "Glutamate and GABA systems as targets for novel antidepressant and mood-stabilizing treatments." Mol Psychiatry 7 Suppl 1: S71-80.

Kuelz, A. K., F. Hohagen and U. Voderholzer (2004). "Neuropsychological performance in obsessive-compulsive disorder: a critical review." Biol Psychol 65(3): 185-236.

Kwon, J. S., J. H. Jang, J. S. Choi and D. H. Kang (2009). "Neuroimaging in obsessive- compulsive disorder." Expert Rev Neurother 9(2): 255-269.

Kwon, J. S., Y. H. Joo, H. J. Nam, M. Lim, E. Y. Cho, M. H. Jung, J. S. Choi, B. Kim, D. H. Kang, S. Oh, T. Park and K. S. Hong (2009). "Association of the glutamate transporter gene SLC1A1 with atypical antipsychotics-induced obsessive-compulsive symptoms." Arch Gen Psychiatry 66(11): 1233-1241.

Lahiri, D. K. and J. I. Nurnberger, Jr. (1991). "A rapid non-enzymatic method for the preparation of HMW DNA from blood for RFLP studies." Nucleic Acids Res 19(19): 5444.

289

290

Laje, G., D. M. Cannon, A. S. Allen, J. M. Klaver, S. A. Peck, X. Liu, H. K. Manji, W. C. Drevets and F. J. McMahon (2010). "Genetic variation in HTR2A influences serotonin transporter binding potential as measured using PET and [11C]DASB." Int J Neuropsychopharmacol 13(6): 715-724.

Lander, E. S., L. M. Linton, B. Birren, C. Nusbaum, M. C. Zody, J. Baldwin, K. Devon, K. Dewar, M. Doyle, W. FitzHugh, R. Funke, D. Gage, K. Harris, A. Heaford, J. Howland, L. Kann, J. Lehoczky, R. LeVine, P. McEwan, K. McKernan, J. Meldrim, J. P. Mesirov, C. Miranda, W. Morris, J. Naylor, C. Raymond, M. Rosetti, R. Santos, A. Sheridan, C. Sougnez, N. Stange-Thomann, N. Stojanovic, A. Subramanian, D. Wyman, J. Rogers, J. Sulston, R. Ainscough, S. Beck, D. Bentley, J. Burton, C. Clee, N. Carter, A. Coulson, R. Deadman, P. Deloukas, A. Dunham, I. Dunham, R. Durbin, L. French, D. Grafham, S. Gregory, T. Hubbard, S. Humphray, A. Hunt, M. Jones, C. Lloyd, A. McMurray, L. Matthews, S. Mercer, S. Milne, J. C. Mullikin, A. Mungall, R. Plumb, M. Ross, R. Shownkeen, S. Sims, R. H. Waterston, R. K. Wilson, L. W. Hillier, J. D. McPherson, M. A. Marra, E. R. Mardis, L. A. Fulton, A. T. Chinwalla, K. H. Pepin, W. R. Gish, S. L. Chissoe, M. C. Wendl, K. D. Delehaunty, T. L. Miner, A. Delehaunty, J. B. Kramer, L. L. Cook, R. S. Fulton, D. L. Johnson, P. J. Minx, S. W. Clifton, T. Hawkins, E. Branscomb, P. Predki, P. Richardson, S. Wenning, T. Slezak, N. Doggett, J. F. Cheng, A. Olsen, S. Lucas, C. Elkin, E. Uberbacher, M. Frazier, R. A. Gibbs, D. M. Muzny, S. E. Scherer, J. B. Bouck, E. J. Sodergren, K. C. Worley, C. M. Rives, J. H. Gorrell, M. L. Metzker, S. L. Naylor, R. S. Kucherlapati, D. L. Nelson, G. M. Weinstock, Y. Sakaki, A. Fujiyama, M. Hattori, T. Yada, A. Toyoda, T. Itoh, C. Kawagoe, H. Watanabe, Y. Totoki, T. Taylor, J. Weissenbach, R. Heilig, W. Saurin, F. Artiguenave, P. Brottier, T. Bruls, E. Pelletier, C. Robert, P. Wincker, D. R. Smith, L. Doucette-Stamm, M. Rubenfield, K. Weinstock, H. M. Lee, J. Dubois, A. Rosenthal, M. Platzer, G. Nyakatura, S. Taudien, A. Rump, H. Yang, J. Yu, J. Wang, G. Huang, J. Gu, L. Hood, L. Rowen, A. Madan, S. Qin, R. W. Davis, N. A. Federspiel, A. P. Abola, M. J. Proctor, R. M. Myers, J. Schmutz, M. Dickson, J. Grimwood, D. R. Cox, M. V. Olson, R. Kaul, C. Raymond, N. Shimizu, K. Kawasaki, S. Minoshima, G. A. Evans, M. Athanasiou, R. Schultz, B. A. Roe, F. Chen, H. Pan, J. Ramser, H. Lehrach, R. Reinhardt, W. R. McCombie, M. de la Bastide, N. Dedhia, H. Blocker, K. Hornischer, G. Nordsiek, R. Agarwala, L. Aravind, J. A. Bailey, A. Bateman, S. Batzoglou, E. Birney, P. Bork, D. G. Brown, C. B. Burge, L. Cerutti, H. C. Chen, D. Church, M. Clamp, R. R. Copley, T. Doerks, S. R. Eddy, E. E. 290

291

Eichler, T. S. Furey, J. Galagan, J. G. Gilbert, C. Harmon, Y. Hayashizaki, D. Haussler, H. Hermjakob, K. Hokamp, W. Jang, L. S. Johnson, T. A. Jones, S. Kasif, A. Kaspryzk, S. Kennedy, W. J. Kent, P. Kitts, E. V. Koonin, I. Korf, D. Kulp, D. Lancet, T. M. Lowe, A. McLysaght, T. Mikkelsen, J. V. Moran, N. Mulder, V. J. Pollara, C. P. Ponting, G. Schuler, J. Schultz, G. Slater, A. F. Smit, E. Stupka, J. Szustakowski, D. Thierry-Mieg, J. Thierry-Mieg, L. Wagner, J. Wallis, R. Wheeler, A. Williams, Y. I. Wolf, K. H. Wolfe, S. P. Yang, R. F. Yeh, F. Collins, M. S. Guyer, J. Peterson, A. Felsenfeld, K. A. Wetterstrand, A. Patrinos, M. J. Morgan, P. de Jong, J. J. Catanese, K. Osoegawa, H. Shizuya, S. Choi, Y. J. Chen and C. International Human Genome Sequencing (2001). "Initial sequencing and analysis of the human genome." Nature 409(6822): 860-921.

Landeros-Weisenberger, A., M. H. Bloch, B. Kelmendi, R. Wegner, J. Nudel, P. Dombrowski, C. Pittenger, J. H. Krystal, W. K. Goodman, J. F. Leckman and V. Coric (2010). "Dimensional predictors of response to SRI pharmacotherapy in obsessive-compulsive disorder." J Affect Disord 121(1-2): 175-179.

LaSalle, V. H., K. R. Cromer, K. N. Nelson, D. Kazuba, L. Justement and D. L. Murphy (2004). "Diagnostic interview assessed neuropsychiatric disorder comorbidity in 334 individuals with obsessive-compulsive disorder." Depress Anxiety 19(3): 163-173.

Le Hellard, S. and V. M. Steen (2014). "Genetic architecture of cognitive traits." Scand J Psychol.

Leckman, J. F., D. Denys, H. B. Simpson, D. Mataix-Cols, E. Hollander, S. Saxena, E. C. Miguel, S. L. Rauch, W. K. Goodman, K. A. Phillips and D. J. Stein (2010). "Obsessive- compulsive disorder: a review of the diagnostic criteria and possible subtypes and dimensional specifiers for DSM-V." Depress Anxiety 27(6): 507-527.

Leckman, J. F., D. E. Grice, L. C. Barr, A. L. de Vries, C. Martin, D. J. Cohen, C. J. McDougle, W. K. Goodman and S. A. Rasmussen (1994). "Tic-related vs. non-tic-related obsessive compulsive disorder." Anxiety 1(5): 208-215.

291

292

Leckman, J. F., D. E. Grice, J. Boardman, H. Zhang, A. Vitale, C. Bondi, J. Alsobrook, B. S. Peterson, D. J. Cohen, S. A. Rasmussen, W. K. Goodman, C. J. McDougle and D. L. Pauls (1997). "Symptoms of obsessive-compulsive disorder." Am J Psychiatry 154(7): 911-917.

Leckman, J. F., S. L. Rauch and D. Mataix-Cols (2007). "Symptom dimensions in obsessive- compulsive disorder: implications for the DSM-V." CNS Spectr 12(5): 376-387, 400.

Lee, D. H. and R. A. Linker (2012). "The role of myelin oligodendrocyte glycoprotein in autoimmune demyelination: a target for multiple sclerosis therapy?" Expert Opin Ther Targets 16(5): 451-462.

Lee, J. L., B. J. Everitt and K. L. Thomas (2004). "Independent cellular processes for hippocampal memory consolidation and reconsolidation." Science 304(5672): 839-843.

Lennertz, L., P. E. Franke, H. J. Grabe, F. Rampacher, S. Schulze-Rauschenbach, V. Guttenthaler, S. Ruhrmann, R. Pukrop, J. Klosterkotter, P. Falkai, W. Maier, M. Wagner and R. Mossner (2013). "The functional coding variant Asn107Ile of the neuropeptide S receptor gene (NPSR1) influences age at onset of obsessive-compulsive disorder." Int J Neuropsychopharmacol 16(9): 1951-1958.

Lennertz, L., M. Wagner, H. J. Grabe, P. E. Franke, V. Guttenthaler, F. Rampacher, S. Schulze- Rauschenbach, A. Vogeley, J. Benninghoff, S. Ruhrmann, R. Pukrop, J. Klosterkotter, P. Falkai, W. Maier and R. Mossner (2014). "5-HT3 receptor influences the washing phenotype and visual organization in obsessive-compulsive disorder supporting 5-HT3 receptor antagonists as novel treatment option." Eur Neuropsychopharmacol 24(1): 86-94.

Lesch, K. P., D. Bengel, A. Heils, S. Z. Sabol, B. D. Greenberg, S. Petri, J. Benjamin, C. R. Muller, D. H. Hamer and D. L. Murphy (1996). "Association of anxiety-related traits with a polymorphism in the serotonin transporter gene regulatory region." Science 274(5292): 1527- 1531.

Lett, T. A., T. J. Wallace, N. I. Chowdhury, A. K. Tiwari, J. L. Kennedy and D. J. Muller (2012). "Pharmacogenetics of antipsychotic-induced weight gain: review and clinical implications." Mol Psychiatry 17(3): 242-266.

292

293

Levinson, D. F., G. S. Zubenko, R. R. Crowe, R. J. DePaulo, W. S. Scheftner, M. M. Weissman, P. Holmans, W. N. Zubenko, S. Boutelle, K. Murphy-Eberenz, D. MacKinnon, M. G. McInnis, D. H. Marta, P. Adams, S. Sassoon, J. A. Knowles, J. Thomas and J. Chellis (2003). "Genetics of recurrent early-onset depression (GenRED): design and preliminary clinical characteristics of a repository sample for genetic linkage studies." Am J Med Genet B Neuropsychiatr Genet 119B(1): 118-130.

Levy, E. R., J. F. Powell, V. J. Buckle, Y. P. Hsu, X. O. Breakefield and I. W. Craig (1989). "Localization of human monoamine oxidase-A gene to Xp11.23-11.4 by in situ hybridization: implications for Norrie disease." Genomics 5(2): 368-370.

Lewis, A. (1936). "Problems of Obsessional Illness: (Section of Psychiatry)." Proc R Soc Med 29(4): 325-336.

Li, B., J. H. Sun, T. Li and Y. C. Yang (2012). "Neuropsychological study of patients with obsessive-compulsive disorder and their parents in China: searching for potential endophenotypes." Neurosci Bull 28(5): 475-482.

Li, J., J. Cui, X. Wang, J. Ma, H. Niu, X. Ma, X. Zhang and S. Liu (2015). "An association study between DLGAP1 rs11081062 and EFNA5 rs26728 polymorphisms with obsessive-compulsive disorder in a Chinese Han population." Neuropsychiatr Dis Treat 11: 897-905.

Li, J. and L. Ji (2005). "Adjusting multiple testing in multilocus analyses using the eigenvalues of a correlation matrix." Heredity (Edinb) 95(3): 221-227.

Li, J. M., C. L. Lu, M. C. Cheng, S. U. Luu, S. H. Hsu and C. H. Chen (2013). "Genetic analysis of the DLGAP1 gene as a candidate gene for schizophrenia." Psychiatry Res 205(1-2): 13-17.

Li, S., J. H. Zhao, J. Luan, R. N. Luben, S. A. Rodwell, K. T. Khaw, K. K. Ong, N. J. Wareham and R. J. Loos (2010). "Cumulative effects and predictive value of common obesity- susceptibility variants identified by genome-wide association studies." Am J Clin Nutr 91(1): 184-190.

Ligon, K. L., S. P. Fancy, R. J. Franklin and D. H. Rowitch (2006). "Olig gene function in CNS development and disease." Glia 54(1): 1-10. 293

294

Lin, K. M. and F. Cheung (1999). "Mental health issues for Asian Americans." Psychiatr Serv 50(6): 774-780.

Liu, J., G. Li, X. Peng, B. Liu, B. Yin, X. Tan, M. Fan, W. Fan, B. Qiang and J. Yuan (2004). "The cloning and preliminarily functional analysis of the human neurotrimin gene." Sci China C Life Sci 47(2): 158-164.

Liu, S., Y. Liu, H. Wang, R. Zhou, J. Zong, C. Li, X. Zhang and X. Ma (2011). "Association of catechol-O-methyl transferase (COMT) gene -287A/G polymorphism with susceptibility to obsessive-compulsive disorder in Chinese Han population." Am J Med Genet B Neuropsychiatr Genet 156B(4): 393-400.

Liu, S. G., X. H. Zhang, Y. Y. Yin, M. J. Wang, F. Y. Che and X. Ma (2011). "An association analysis between 5-HTTLPR polymorphism and obsessive-compulsive disorder, Tourette syndrome in a Chinese Han population." CNS Neurosci Ther 17(6): 793-795.

Lochner, C., S. M. Hemmings, C. J. Kinnear, J. C. Moolman-Smook, V. A. Corfield, J. A. Knowles, D. J. Niehaus and D. J. Stein (2004). "Corrigendum to "gender in obsessive- compulsive disorder: clinical and genetic findings" [Eur. Neuropsychopharmacol. 14 (2004) 105- 113]." Eur Neuropsychopharmacol 14(5): 437-445.

Lochner, C., S. M. Hemmings, C. J. Kinnear, J. C. Moolman-Smook, V. A. Corfield, J. A. Knowles, D. J. Niehaus and D. J. Stein (2004). "Gender in obsessive-compulsive disorder: clinical and genetic findings." Eur Neuropsychopharmacol 14(2): 105-113.

Lochner, C., S. M. Hemmings, C. J. Kinnear, D. Nel, S. M. Hemmings, S. Seedat, J. C. Moolman-Smook and D. J. Stein (2008). "Cluster analysis of obsessive-compulsive symptomatology: identifying obsessive-compulsive disorder subtypes." Isr J Psychiatry Relat Sci 45(3): 164-176.

Luleyap, H., D. Onatoglu, A. Tahiroglu, D. Alptekin, M. Yilmaz, S. Cetiner, A. Pazarbasi, I. Unal and A. Avci (2012). "Association between Obsessive Compulsive Disorder and Tumor Necrosis Factor-alpha Gene -308 (G>A) and -850 (C>T) Polymorphisms in Turkish Children." Balkan J Med Genet 15(2): 61-66.

294

295

Mac Master, F. P., M. S. Keshavan, E. L. Dick and D. R. Rosenberg (1999). "Corpus callosal signal intensity in treatment-naive pediatric obsessive compulsive disorders." Prog Neuropsychopharmacol Biol Psychiatry 23(4): 601-612.

MacDonald, M. (1981). Mystical Bedlam: Madness, Anxiety, and Healing in Seventeenth- Century England, Cambridge University Press.

MacGillivray, S., B. Arroll, S. Hatcher, S. Ogston, I. Reid, F. Sullivan, B. Williams and I. Crombie (2003). "Efficacy and tolerability of selective serotonin reuptake inhibitors compared with tricyclic antidepressants in depression treated in primary care: systematic review and meta- analysis." BMJ 326(7397): 1014.

Mack, N. A., A. P. Porter, H. J. Whalley, J. P. Schwarz, R. C. Jones, A. S. Khaja, A. Bjartell, K. I. Anderson and A. Malliri (2012). "beta2-syntrophin and Par-3 promote an apicobasal Rac activity gradient at cell-cell junctions by differentially regulating Tiam1 activity." Nat Cell Biol 14(11): 1169-1180.

Mackay, T. F. (2014). "Epistasis and quantitative traits: using model organisms to study gene- gene interactions." Nat Rev Genet 15(1): 22-33.

MacMaster, F. P. (2010). "Translational neuroimaging research in pediatric obsessive- compulsive disorder." Dialogues Clin Neurosci 12(2): 165-174.

Maina, G., G. Rosso, R. Zanardini, F. Bogetto, M. Gennarelli and L. Bocchio-Chiavetto (2010). "Serum levels of brain-derived neurotrophic factor in drug-naive obsessive-compulsive patients: a case-control study." J Affect Disord 122(1-2): 174-178.

Malhotra, A. K., C. U. Correll, N. I. Chowdhury, D. J. Muller, P. K. Gregersen, A. T. Lee, A. K. Tiwari, J. M. Kane, W. W. Fleischhacker, R. S. Kahn, R. A. Ophoff, H. Y. Meltzer, T. Lencz and J. L. Kennedy (2012). "Association between common variants near the melanocortin 4 receptor gene and severe antipsychotic drug-induced weight gain." Arch Gen Psychiatry 69(9): 904-912.

Marazziti, D. and G. Consoli (2010). "Treatment strategies for obsessive-compulsive disorder." Expert Opin Pharmacother 11(3): 331-343.

295

296

Marquez, L., B. Camarena, S. Hernandez, C. Loyzaga, L. Vargas and H. Nicolini (2013). "Association study between BDNF gene variants and Mexican patients with obsessive- compulsive disorder." Eur Neuropsychopharmacol 23(11): 1600-1605.

Martinowich, K. and B. Lu (2008). "Interaction between BDNF and serotonin: role in mood disorders." Neuropsychopharmacology 33(1): 73-83.

Martoni, R. M., C. Brombin, A. Nonis, G. C. Salgari, A. Buongiorno, M. C. Cavallini, E. Galimberti and L. Bellodi (2014). "Evaluating effect of symptoms heterogeneity on decision- making ability in obsessive-compulsive disorder." Psychiatry Clin Neurosci.

Mas, S., M. Pagerols, P. Gasso, A. Ortiz, N. Rodriguez, A. Morer, M. T. Plana, A. Lafuente and L. Lazaro (2014). "Role of GAD2 and HTR1B genes in early-onset obsessive-compulsive disorder: results from transmission disequilibrium study." Genes Brain Behav 13(4): 409-417.

Masys, D., D. Baker, A. Butros and K. E. Cowles (2002). "Giving patients access to their medical records via the internet: the PCASSO experience." J Am Med Inform Assoc 9(2): 181- 191.

Mataix-Cols, D. (2006). "Deconstructing obsessive-compulsive disorder: a multidimensional perspective." Curr Opin Psychiatry 19(1): 84-89.

Mataix-Cols, D., M. Boman, B. Monzani, C. Ruck, E. Serlachius, N. Langstrom and P. Lichtenstein (2013). "Population-based, multigenerational family clustering study of obsessive- compulsive disorder." JAMA Psychiatry 70(7): 709-717.

Mataix-Cols, D., I. M. Marks, J. H. Greist, K. A. Kobak and L. Baer (2002). "Obsessive- compulsive symptom dimensions as predictors of compliance with and response to behaviour therapy: results from a controlled trial." Psychother Psychosom 71(5): 255-262.

Mataix-Cols, D., E. Nakatani, N. Micali and I. Heyman (2008). "Structure of obsessive- compulsive symptoms in pediatric OCD." J Am Acad Child Adolesc Psychiatry 47(7): 773-778.

296

297

Mataix-Cols, D., S. L. Rauch, L. Baer, J. L. Eisen, D. M. Shera, W. K. Goodman, S. A. Rasmussen and M. A. Jenike (2002). "Symptom stability in adult obsessive-compulsive disorder: data from a naturalistic two-year follow-up study." Am J Psychiatry 159(2): 263-268.

Mataix-Cols, D., S. L. Rauch, P. A. Manzo, M. A. Jenike and L. Baer (1999). "Use of factor- analyzed symptom dimensions to predict outcome with serotonin reuptake inhibitors and placebo in the treatment of obsessive-compulsive disorder." Am J Psychiatry 156(9): 1409-1416.

Mataix-Cols, D., M. C. Rosario-Campos and J. F. Leckman (2005). "A multidimensional model of obsessive-compulsive disorder." Am J Psychiatry 162(2): 228-238.

Mathews, C. A., J. A. Badner, J. M. Andresen, B. Sheppard, J. A. Himle, J. E. Grant, K. A. Williams, D. A. Chavira, A. Azzam, M. Schwartz, V. I. Reus, S. W. Kim, E. H. Cook and G. L. Hanna (2012). "Genome-wide linkage analysis of obsessive-compulsive disorder implicates chromosome 1p36." Biol Psychiatry 72(8): 629-636.

Mathews, C. A., T. Greenwood, J. Wessel, A. Azzam, H. Garrido, D. A. Chavira, U. Chandavarkar, M. Bagnarello, M. Stein and N. J. Schork (2008). "Evidence for a heritable unidimensional symptom factor underlying obsessionality." Am J Med Genet B Neuropsychiatr Genet 147B(6): 676-685.

Mathews, C. A., C. M. Nievergelt, A. Azzam, H. Garrido, D. A. Chavira, J. Wessel, M. Bagnarello, V. I. Reus and N. J. Schork (2007). "Heritability and clinical features of multigenerational families with obsessive-compulsive disorder and hoarding." Am J Med Genet B Neuropsychiatr Genet 144B(2): 174-182.

Mathis, M. A., P. Alvarenga, G. Funaro, R. C. Torresan, I. Moraes, A. R. Torres, M. L. Zilberman and A. G. Hounie (2011). "Gender differences in obsessive-compulsive disorder: a literature review." Rev Bras Psiquiatr 33(4): 390-399.

Matsui, A., M. Tran, A. C. Yoshida, S. S. Kikuchi, M. U, M. Ogawa and T. Shimogori (2013). "BTBD3 controls dendrite orientation toward active axons in mammalian neocortex." Science 342(6162): 1114-1118.

297

298

Matsunaga, H., K. Hayashida, N. Kiriike, K. Maebayashi and D. J. Stein (2010). "The clinical utility of symptom dimensions in obsessive-compulsive disorder." Psychiatry Res 180(1): 25-29.

Matsunaga, H., N. Kiriike, T. Matsui, A. Miyata, Y. Iwasaki, K. Fujimoto, S. Kasai and M. Kojima (2000). "Gender differences in social and interpersonal features and personality disorders among Japanese patients with obsessive-compulsive disorder." Compr Psychiatry 41(4): 266- 272.

Matsunaga, H., K. Maebayashi, K. Hayashida, K. Okino, T. Matsui, T. Iketani, N. Kiriike and D. J. Stein (2008). "Symptom structure in Japanese patients with obsessive-compulsive disorder." Am J Psychiatry 165(2): 251-253.

Mattheisen, M., J. F. Samuels, Y. Wang, B. D. Greenberg, A. J. Fyer, J. T. McCracken, D. A. Geller, D. L. Murphy, J. A. Knowles, M. A. Grados, M. A. Riddle, S. A. Rasmussen, N. C. McLaughlin, E. L. Nurmi, K. D. Askland, H. D. Qin, B. A. Cullen, J. Piacentini, D. L. Pauls, O. J. Bienvenu, S. E. Stewart, K. Y. Liang, F. S. Goes, B. Maher, A. E. Pulver, Y. Y. Shugart, D. Valle, C. Lange and G. Nestadt (2014). "Genome-wide association study in obsessive- compulsive disorder: results from the OCGAS." Mol Psychiatry.

May-Tolzmann, U. (1998). "Obsessional neurosis: a nosographic innovation by Freud." Hist Psychiatry 9(35): 335-353.

McDougle, C. J. (1997). "Update on pharmacologic management of OCD: agents and augmentation." J Clin Psychiatry 58 Suppl 12: 11-17.

McDougle, C. J., W. K. Goodman, J. F. Leckman, N. C. Lee, G. R. Heninger and L. H. Price (1994). "Haloperidol addition in fluvoxamine-refractory obsessive-compulsive disorder. A double-blind, placebo-controlled study in patients with and without tics." Arch Gen Psychiatry 51(4): 302-308.

McDougle, C. J., W. K. Goodman, J. F. Leckman and L. H. Price (1993). "The psychopharmacology of obsessive compulsive disorder. Implications for treatment and pathogenesis." Psychiatr Clin North Am 16(4): 749-766.

298

299

McGovern, D. P., M. R. Jones, K. D. Taylor, K. Marciante, X. Yan, M. Dubinsky, A. Ippoliti, E. Vasiliauskas, D. Berel, C. Derkowski, D. Dutridge, P. Fleshner, D. Q. Shih, G. Melmed, E. Mengesha, L. King, S. Pressman, T. Haritunians, X. Guo, S. R. Targan, J. I. Rotter and I. B. D. G. C. International (2010). "Fucosyltransferase 2 (FUT2) non-secretor status is associated with Crohn's disease." Hum Mol Genet 19(17): 3468-3476.

McGrath, L. M., D. Yu, C. Marshall, L. K. Davis, B. Thiruvahindrapuram, B. Li, C. Cappi, G. Gerber, A. Wolf, F. A. Schroeder, L. Osiecki, C. O'Dushlaine, A. Kirby, C. Illmann, S. Haddad, P. Gallagher, J. A. Fagerness, C. L. Barr, L. Bellodi, F. Benarroch, O. J. Bienvenu, D. W. Black, M. H. Bloch, R. D. Bruun, C. L. Budman, B. Camarena, D. C. Cath, M. C. Cavallini, S. Chouinard, V. Coric, B. Cullen, R. Delorme, D. Denys, E. M. Derks, Y. Dion, M. C. Rosario, V. Eapen, P. Evans, P. Falkai, T. V. Fernandez, H. Garrido, D. Geller, H. J. Grabe, M. A. Grados, B. D. Greenberg, V. Gross-Tsur, E. Grunblatt, G. A. Heiman, S. M. Hemmings, L. D. Herrera, A. G. Hounie, J. Jankovic, J. L. Kennedy, R. A. King, R. Kurlan, N. Lanzagorta, M. Leboyer, J. F. Leckman, L. Lennertz, C. Lochner, T. L. Lowe, G. J. Lyon, F. Macciardi, W. Maier, J. T. McCracken, W. McMahon, D. L. Murphy, A. L. Naarden, B. M. Neale, E. Nurmi, A. J. Pakstis, M. T. Pato, C. N. Pato, J. Piacentini, C. Pittenger, Y. Pollak, V. I. Reus, M. A. Richter, M. Riddle, M. M. Robertson, D. Rosenberg, G. A. Rouleau, S. Ruhrmann, A. S. Sampaio, J. Samuels, P. Sandor, B. Sheppard, H. S. Singer, J. H. Smit, D. J. Stein, J. A. Tischfield, H. Vallada, J. Veenstra-VanderWeele, S. Walitza, Y. Wang, J. R. Wendland, Y. Y. Shugart, E. C. Miguel, H. Nicolini, B. A. Oostra, R. Moessner, M. Wagner, A. Ruiz-Linares, P. Heutink, G. Nestadt, N. Freimer, T. Petryshen, D. Posthuma, M. A. Jenike, N. J. Cox, G. L. Hanna, H. Brentani, S. W. Scherer, P. D. Arnold, S. E. Stewart, C. A. Mathews, J. A. Knowles, E. H. Cook, D. L. Pauls, K. Wang and J. M. Scharf (2014). "Copy number variation in obsessive-compulsive disorder and tourette syndrome: a cross-disorder study." J Am Acad Child Adolesc Psychiatry 53(8): 910-919.

McKay, D., J. Piacentini, S. Greisberg, F. Graae, M. Jaffer and J. Miller (2006). "The structure of childhood obsessions and compulsions: dimensions in an outpatient sample." Behav Res Ther 44(1): 137-146.

299

300

McMahon, F. J., S. Buervenich, D. Charney, R. Lipsky, A. J. Rush, A. F. Wilson, A. J. Sorant, G. J. Papanicolaou, G. Laje, M. Fava, M. H. Trivedi, S. R. Wisniewski and H. Manji (2006). "Variation in the gene encoding the serotonin 2A receptor is associated with outcome of antidepressant treatment." Am J Hum Genet 78(5): 804-814.

Menzies, L., S. R. Chamberlain, A. R. Laird, S. M. Thelen, B. J. Sahakian and E. T. Bullmore (2008). "Integrating evidence from neuroimaging and neuropsychological studies of obsessive- compulsive disorder: the orbitofronto-striatal model revisited." Neurosci Biobehav Rev 32(3): 525-549.

Menzies, L., G. B. Williams, S. R. Chamberlain, C. Ooi, N. Fineberg, J. Suckling, B. J. Sahakian, T. W. Robbins and E. T. Bullmore (2008). "White matter abnormalities in patients with obsessive-compulsive disorder and their first-degree relatives." Am J Psychiatry 165(10): 1308-1315.

Mercader, J. M., M. Ribases, M. Gratacos, J. R. Gonzalez, M. Bayes, R. de Cid, A. Badia, F. Fernandez-Aranda and X. Estivill (2007). "Altered brain-derived neurotrophic factor blood levels and gene variability are associated with anorexia and bulimia." Genes Brain Behav 6(8): 706-716.

Miguel, E. C., L. Baer, B. J. Coffey, S. L. Rauch, C. R. Savage, R. L. O'Sullivan, K. Phillips, C. Moretti, J. F. Leckman and M. A. Jenike (1997). "Phenomenological differences appearing with repetitive behaviours in obsessive-compulsive disorder and Gilles de la Tourette's syndrome." Br J Psychiatry 170: 140-145.

Miguel, E. C., Y. A. Ferrao, M. C. Rosario, M. A. Mathis, A. R. Torres, L. F. Fontenelle, A. G. Hounie, R. G. Shavitt, A. V. Cordioli, C. H. Gonzalez, K. Petribu, J. B. Diniz, D. M. Malavazzi, R. C. Torresan, A. L. Raffin, E. Meyer, D. T. Braga, S. Borcato, C. Valerio, L. N. Gropo, S. Prado Hda, E. A. Perin, S. I. Santos, H. Copque, M. C. Borges, A. P. Lopes, E. D. Silva and D. Brazilian Research Consortium on Obsessive-Compulsive Spectrum (2008). "The Brazilian Research Consortium on Obsessive-Compulsive Spectrum Disorders: recruitment, assessment instruments, methods for the development of multicenter collaborative studies and preliminary results." Rev Bras Psiquiatr 30(3): 185-196.

300

301

Miguita, K., Q. Cordeiro, J. Siqueira-Roberto, R. G. Shavitt, J. C. Castillo, A. R. Castillo, E. C. Miguel and H. Vallada (2007). "Association analysis between a VNTR intron 8 polymorphism of the dopamine transporter gene (SLC6A3) and obsessive- compulsive disorder in a Brazilian sample." Arq Neuropsiquiatr 65(4A): 936-941.

Milad, M. R. and S. L. Rauch (2012). "Obsessive-compulsive disorder: beyond segregated cortico-striatal pathways." Trends Cogn Sci 16(1): 43-51.

Molteni, R., A. Cattaneo, F. Calabrese, F. Macchi, J. D. Olivier, G. Racagni, B. A. Ellenbroek, M. Gennarelli and M. A. Riva (2010). "Reduced function of the serotonin transporter is associated with decreased expression of BDNF in rodents as well as in humans." Neurobiol Dis 37(3): 747-755.

Monteiro, P. and G. Feng (2015). "Learning from Animal Models of Obsessive-Compulsive Disorder." Biol Psychiatry.

Monzani, B., F. Rijsdijk, M. Anson, A. C. Iervolino, L. Cherkas, T. Spector and D. Mataix-Cols (2012). "A twin study of body dysmorphic concerns." Psychol Med 42(9): 1949-1955.

Monzani, B., F. Rijsdijk, L. Cherkas, J. Harris, N. Keuthen and D. Mataix-Cols (2012). "Prevalence and heritability of skin picking in an adult community sample: a twin study." Am J Med Genet B Neuropsychiatr Genet 159B(5): 605-610.

Monzani, B., F. Rijsdijk, J. Harris and D. Mataix-Cols (2014). "The structure of genetic and environmental risk factors for dimensional representations of DSM-5 obsessive-compulsive spectrum disorders." JAMA Psychiatry 71(2): 182-189.

Morley, K. I. and W. D. Hall (2004). "Using pharmacogenetics and pharmacogenomics in the treatment of psychiatric disorders: some ethical and economic considerations." J Mol Med (Berl) 82(1): 21-30.

Mossner, R., S. Daniel, D. Albert, A. Heils, O. Okladnova, A. Schmitt and K. P. Lesch (2000). "Serotonin transporter function is modulated by brain-derived neurotrophic factor (BDNF) but not nerve growth factor (NGF)." Neurochem Int 36(3): 197-202.

301

302

Mossner, R., S. Walitza, F. Geller, A. Scherag, L. Gutknecht, C. Jacob, L. Bogusch, H. Remschmidt, M. Simons, B. Herpertz-Dahlmann, C. Fleischhaker, E. Schulz, A. Warnke, A. Hinney, C. Wewetzer and K. P. Lesch (2006). "Transmission disequilibrium of polymorphic variants in the tryptophan hydroxylase-2 gene in children and adolescents with obsessive- compulsive disorder." Int J Neuropsychopharmacol 9(4): 437-442.

Mossner, R., S. Walitza, K. P. Lesch, F. Geller, N. Barth, H. Remschmidt, F. Hahn, B. Herpertz- Dahlmann, C. Fleischhaker, E. Schulz, A. Warnke, A. Hinney and C. Wewetzer (2005). "Brain- derived neurotrophic factor V66M polymorphism in childhood-onset obsessive-compulsive disorder." Int J Neuropsychopharmacol 8(1): 133-136.

Motazacker, M. M., B. R. Rost, T. Hucho, M. Garshasbi, K. Kahrizi, R. Ullmann, S. S. Abedini, S. E. Nieh, S. H. Amini, C. Goswami, A. Tzschach, L. R. Jensen, D. Schmitz, H. H. Ropers, H. Najmabadi and A. W. Kuss (2007). "A defect in the ionotropic glutamate receptor 6 gene (GRIK2) is associated with autosomal recessive mental retardation." Am J Hum Genet 81(4): 792-798.

Muglia, P., A. M. Vicente, M. Verga, N. King, F. Macciardi and J. L. Kennedy (2003). "Association between the BDNF gene and schizophrenia." Mol Psychiatry 8(2): 146-147.

Muller, D. J., E. J. Brandl, R. Hwang, A. K. Tiwari, J. E. Sturgess, C. C. Zai, J. A. Lieberman, J. L. Kennedy and M. A. Richter (2012). "The AmpliChip(R) CYP450 test and response to treatment in schizophrenia and obsessive compulsive disorder: a pilot study and focus on cases with abnormal CYP2D6 drug metabolism." Genet Test Mol Biomarkers 16(8): 897-903.

Mumford, D. B. (1992). "Detection of psychiatric disorders among Asian patients presenting with somatic symptoms." Br J Hosp Med 47(3): 202-204.

Mundo, E., M. A. Richter, F. Sam, F. Macciardi and J. L. Kennedy (2000). "Is the 5-HT(1Dbeta) receptor gene implicated in the pathogenesis of obsessive-compulsive disorder?" Am J Psychiatry 157(7): 1160-1161.

302

303

Mundo, E., M. A. Richter, G. Zai, F. Sam, J. McBride, F. Macciardi and J. L. Kennedy (2002). "5HT1Dbeta Receptor gene implicated in the pathogenesis of Obsessive-Compulsive Disorder: further evidence from a family-based association study." Mol Psychiatry 7(7): 805-809.

Murphy, D. L., P. R. Moya, M. A. Fox, L. M. Rubenstein, J. R. Wendland and K. R. Timpano (2013). "Anxiety and affective disorder comorbidity related to serotonin and other neurotransmitter systems: obsessive-compulsive disorder as an example of overlapping clinical and genetic heterogeneity." Philos Trans R Soc Lond B Biol Sci 368(1615): 20120435.

Murphy, D. L., K. R. Timpano, M. G. Wheaton, B. D. Greenberg and E. C. Miguel (2010). "Obsessive-compulsive disorder and its related disorders: a reappraisal of obsessive-compulsive spectrum concepts." Dialogues Clin Neurosci 12(2): 131-148.

Nagase, T., K. Ishikawa, M. Suyama, R. Kikuno, M. Hirosawa, N. Miyajima, A. Tanaka, H. Kotani, N. Nomura and O. Ohara (1998). "Prediction of the coding sequences of unidentified human genes. XII. The complete sequences of 100 new cDNA clones from brain which code for large proteins in vitro." DNA Res 5(6): 355-364.

Nair, A. and R. Howard (2013). "ENCODE and a new landscape for psychiatric genetics." Br J Psychiatry 203(2): 84-85.

Nakamae, T. (2011). "[Diversity of obsessive-compulsive disorder and pharmacotherapy associated with obsessive-compulsive spectrum disorders]." Seishin Shinkeigaku Zasshi 113(10): 1016-1025.

Narayanaswamy, J. C., B. Viswanath, A. Veshnal Cherian, S. Bada Math, T. Kandavel and Y. C. Janardhan Reddy (2012). "Impact of age of onset of illness on clinical phenotype in OCD." Psychiatry Res 200(2-3): 554-559.

Neafsey, P., G. Ginsberg, D. Hattis and B. Sonawane (2009). "Genetic polymorphism in cytochrome P450 2D6 (CYP2D6): Population distribution of CYP2D6 activity." J Toxicol Environ Health B Crit Rev 12(5-6): 334-361.

Nestadt, G., O. J. Bienvenu, G. Cai, J. Samuels and W. W. Eaton (1998). "Incidence of obsessive-compulsive disorder in adults." J Nerv Ment Dis 186(7): 401-406. 303

304

Nestadt, G., C. Z. Di, M. A. Riddle, M. A. Grados, B. D. Greenberg, A. J. Fyer, J. T. McCracken, S. L. Rauch, D. L. Murphy, S. A. Rasmussen, B. Cullen, A. Pinto, J. A. Knowles, J. Piacentini, D. L. Pauls, O. J. Bienvenu, Y. Wang, K. Y. Liang, J. F. Samuels and K. B. Roche (2009). "Obsessive-compulsive disorder: subclassification based on co-morbidity." Psychol Med 39(9): 1491-1501.

Nestadt, G., M. Grados and J. F. Samuels (2010). "Genetics of obsessive-compulsive disorder." Psychiatr Clin North Am 33(1): 141-158.

Nestadt, G., J. Samuels, M. Riddle, O. J. Bienvenu, 3rd, K. Y. Liang, M. LaBuda, J. Walkup, M. Grados and R. Hoehn-Saric (2000). "A family study of obsessive-compulsive disorder." Arch Gen Psychiatry 57(4): 358-363.

Nestadt, G., J. Samuels, M. A. Riddle, K. Y. Liang, O. J. Bienvenu, R. Hoehn-Saric, M. Grados and B. Cullen (2001). "The relationship between obsessive-compulsive disorder and anxiety and affective disorders: results from the Johns Hopkins OCD Family Study." Psychol Med 31(3): 481-487.

Neves-Pereira, M., E. Mundo, P. Muglia, N. King, F. Macciardi and J. L. Kennedy (2002). "The brain-derived neurotrophic factor gene confers susceptibility to bipolar disorder: evidence from a family-based association study." Am J Hum Genet 71(3): 651-655.

Nibuya, M., S. Morinobu and R. S. Duman (1995). "Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments." J Neurosci 15(11): 7539-7547.

Nicolini, H., P. Arnold, G. Nestadt, N. Lanzagorta and J. L. Kennedy (2009). "Overview of genetics and obsessive-compulsive disorder." Psychiatry Res 170(1): 7-14.

Niitsu, T., C. Fabbri, F. Bentini and A. Serretti (2013). "Pharmacogenetics in major depression: a comprehensive meta-analysis." Prog Neuropsychopharmacol Biol Psychiatry 45: 183-194.

Novak, C. E., N. J. Keuthen, S. E. Stewart and D. L. Pauls (2009). "A twin concordance study of trichotillomania." Am J Med Genet B Neuropsychiatr Genet 150B(7): 944-949.

304

305

Numakawa, T., S. Suzuki, E. Kumamaru, N. Adachi, M. Richards and H. Kunugi (2010). "BDNF function and intracellular signaling in neurons." Histol Histopathol 25(2): 237-258.

Nyholt, D. R. (2004). "A simple correction for multiple testing for single-nucleotide polymorphisms in linkage disequilibrium with each other." Am J Hum Genet 74(4): 765-769.

Obsessive Compulsive Cognitions Working, G. (2001). "Development and initial validation of the obsessive beliefs questionnaire and the interpretation of intrusions inventory." Behav Res Ther 39(8): 987-1006.

Obsessive Compulsive Cognitions Working, G. (2003). "Psychometric validation of the Obsessive Beliefs Questionnaire and the Interpretation of Intrusions Inventory: Part I." Behav Res Ther 41(8): 863-878.

Obsessive Compulsive Cognitions Working, G. (2005). "Psychometric validation of the obsessive belief questionnaire and interpretation of intrusions inventory--Part 2: Factor analyses and testing of a brief version." Behav Res Ther 43(11): 1527-1542.

Orvaschel, H. (1990). "Early onset psychiatric disorder in high risk children and increased familial morbidity." J Am Acad Child Adolesc Psychiatry 29(2): 184-188.

O'Sullivan, R. L., N. J. Keuthen, G. A. Christenson, C. S. Mansueto, D. J. Stein and S. E. Swedo (1997). "Trichotillomania: behavioral symptom or clinical syndrome?" Am J Psychiatry 154(10): 1442-1449.

Overbeek, T., K. Schruers, E. Vermetten and E. Griez (2002). "Comorbidity of obsessive- compulsive disorder and depression: prevalence, symptom severity, and treatment effect." J Clin Psychiatry 63(12): 1106-1112.

Pallanti, S. and G. Grassi (2014). "Pharmacologic treatment of obsessive-compulsive disorder comorbidity." Expert Opin Pharmacother 15(17): 2543-2552.

Pallanti, S., G. Grassi, E. D. Sarrecchia, A. Cantisani and M. Pellegrini (2011). "Obsessive- compulsive disorder comorbidity: clinical assessment and therapeutic implications." Front Psychiatry 2: 70.

305

306

Pallanti, S. and E. Hollander (2014). "Pharmacological, experimental therapeutic, and transcranial magnetic stimulation treatments for compulsivity and impulsivity." CNS Spectr 19(1): 50-61.

Pallanti, S., E. Hollander, C. Bienstock, L. Koran, J. Leckman, D. Marazziti, M. Pato, D. Stein, J. Zohar and O. C. D. C. International Treatment Refractory (2002). "Treatment non-response in OCD: methodological issues and operational definitions." Int J Neuropsychopharmacol 5(2): 181-191.

Paschen, W., C. D. Blackstone, R. L. Huganir and C. A. Ross (1994). "Human GluR6 kainate receptor (GRIK2): molecular cloning, expression, polymorphism, and chromosomal assignment." Genomics 20(3): 435-440.

Paschen, W., J. C. Hedreen and C. A. Ross (1994). "RNA editing of the glutamate receptor subunits GluR2 and GluR6 in human brain tissue." J Neurochem 63(5): 1596-1602.

Patton, J. H., M. S. Stanford and E. S. Barratt (1995). "Factor structure of the Barratt impulsiveness scale." J Clin Psychol 51(6): 768-774.

Pauls, D. L. (2010). "The genetics of obsessive-compulsive disorder: a review." Dialogues Clin Neurosci 12(2): 149-163.

Pauls, D. L., A. Abramovitch, S. L. Rauch and D. A. Geller (2014). "Obsessive-compulsive disorder: an integrative genetic and neurobiological perspective." Nat Rev Neurosci 15(6): 410- 424.

Pe'er, I., R. Yelensky, D. Altshuler and M. J. Daly (2008). "Estimation of the multiple testing burden for genomewide association studies of nearly all common variants." Genet Epidemiol 32(4): 381-385.

Perez-Torrado, R., D. Yamada and P. A. Defossez (2006). "Born to bind: the BTB protein- protein interaction domain." Bioessays 28(12): 1194-1202.

Perneger, T. V. (1998). "What's wrong with Bonferroni adjustments." BMJ 316(7139): 1236- 1238.

306

307

Perroud, N., M. Guipponi, A. Pertusa, M. A. Fullana, A. C. Iervolino, L. Cherkas, T. Spector, D. Collier and D. Mataix-Cols (2011). "Genome-wide association study of hoarding traits." Am J Med Genet B Neuropsychiatr Genet 156(2): 240-242.

Pham-Dinh, D., M. G. Mattei, J. L. Nussbaum, G. Roussel, P. Pontarotti, N. Roeckel, I. H. Mather, K. Artzt, K. F. Lindahl and A. Dautigny (1993). "Myelin/oligodendrocyte glycoprotein is a member of a subset of the immunoglobulin superfamily encoded within the major histocompatibility complex." Proc Natl Acad Sci U S A 90(17): 7990-7994.

Pinto, A., J. L. Eisen, M. C. Mancebo, B. D. Greenberg, R. L. Stout and S. A. Rasmussen (2007). "Taboo thoughts and doubt/checking: a refinement of the factor structure for obsessive- compulsive disorder symptoms." Psychiatry Res 151(3): 255-258.

Pinto, A., B. D. Greenberg, M. A. Grados, O. J. Bienvenu, 3rd, J. F. Samuels, D. L. Murphy, G. Hasler, R. L. Stout, S. L. Rauch, Y. Y. Shugart, D. L. Pauls, J. A. Knowles, A. J. Fyer, J. T. McCracken, J. Piacentini, Y. Wang, V. L. Willour, B. Cullen, K. Y. Liang, R. Hoehn-Saric, M. A. Riddle, S. A. Rasmussen and G. Nestadt (2008). "Further development of YBOCS dimensions in the OCD Collaborative Genetics study: symptoms vs. categories." Psychiatry Res 160(1): 83-93.

Pinto, D., A. T. Pagnamenta, L. Klei, R. Anney, D. Merico, R. Regan, J. Conroy, T. R. Magalhaes, C. Correia, B. S. Abrahams, J. Almeida, E. Bacchelli, G. D. Bader, A. J. Bailey, G. Baird, A. Battaglia, T. Berney, N. Bolshakova, S. Bolte, P. F. Bolton, T. Bourgeron, S. Brennan, J. Brian, S. E. Bryson, A. R. Carson, G. Casallo, J. Casey, B. H. Chung, L. Cochrane, C. Corsello, E. L. Crawford, A. Crossett, C. Cytrynbaum, G. Dawson, M. de Jonge, R. Delorme, I. Drmic, E. Duketis, F. Duque, A. Estes, P. Farrar, B. A. Fernandez, S. E. Folstein, E. Fombonne, C. M. Freitag, J. Gilbert, C. Gillberg, J. T. Glessner, J. Goldberg, A. Green, J. Green, S. J. Guter, H. Hakonarson, E. A. Heron, M. Hill, R. Holt, J. L. Howe, G. Hughes, V. Hus, R. Igliozzi, C. Kim, S. M. Klauck, A. Kolevzon, O. Korvatska, V. Kustanovich, C. M. Lajonchere, J. A. Lamb, M. Laskawiec, M. Leboyer, A. Le Couteur, B. L. Leventhal, A. C. Lionel, X. Q. Liu, C. Lord, L. Lotspeich, S. C. Lund, E. Maestrini, W. Mahoney, C. Mantoulan, C. R. Marshall, H. McConachie, C. J. McDougle, J. McGrath, W. M. McMahon, A. Merikangas, O. Migita, N. J. Minshew, G. K. Mirza, J. Munson, S. F. Nelson, C. Noakes, A. Noor, G. Nygren, G. Oliveira, K. 307

308

Papanikolaou, J. R. Parr, B. Parrini, T. Paton, A. Pickles, M. Pilorge, J. Piven, C. P. Ponting, D. J. Posey, A. Poustka, F. Poustka, A. Prasad, J. Ragoussis, K. Renshaw, J. Rickaby, W. Roberts, K. Roeder, B. Roge, M. L. Rutter, L. J. Bierut, J. P. Rice, J. Salt, K. Sansom, D. Sato, R. Segurado, A. F. Sequeira, L. Senman, N. Shah, V. C. Sheffield, L. Soorya, I. Sousa, O. Stein, N. Sykes, V. Stoppioni, C. Strawbridge, R. Tancredi, K. Tansey, B. Thiruvahindrapduram, A. P. Thompson, S. Thomson, A. Tryfon, J. Tsiantis, H. Van Engeland, J. B. Vincent, F. Volkmar, S. Wallace, K. Wang, Z. Wang, T. H. Wassink, C. Webber, R. Weksberg, K. Wing, K. Wittemeyer, S. Wood, J. Wu, B. L. Yaspan, D. Zurawiecki, L. Zwaigenbaum, J. D. Buxbaum, R. M. Cantor, E. H. Cook, H. Coon, M. L. Cuccaro, B. Devlin, S. Ennis, L. Gallagher, D. H. Geschwind, M. Gill, J. L. Haines, J. Hallmayer, J. Miller, A. P. Monaco, J. I. Nurnberger, Jr., A. D. Paterson, M. A. Pericak-Vance, G. D. Schellenberg, P. Szatmari, A. M. Vicente, V. J. Vieland, E. M. Wijsman, S. W. Scherer, J. S. Sutcliffe and C. Betancur (2010). "Functional impact of global rare copy number variation in autism spectrum disorders." Nature 466(7304): 368-372.

Pittenger, C. (2013). "Disorders of memory and plasticity in psychiatric disease." Dialogues Clin Neurosci 15(4): 455-463.

Plani-Lam, J. H., T. C. Chow, K. L. Siu, W. H. Chau, M. H. Ng, S. Bao, C. T. Ng, P. Sham, D. K. Shum, E. Ingley, D. Y. Jin and Y. Q. Song (2015). "PTPN21 exerts pro-neuronal survival and neuritic elongation via ErbB4/NRG3 signaling." Int J Biochem Cell Biol 61: 53-62.

Pooley, E. C., N. Fineberg and P. J. Harrison (2007). "The met(158) allele of catechol-O- methyltransferase (COMT) is associated with obsessive-compulsive disorder in men: case- control study and meta-analysis." Mol Psychiatry 12(6): 556-561.

Poyurovsky, M., E. Michaelovsky, A. Frisch, G. Knoll, I. Amir, B. Finkel, F. Buniak, H. Hermesh and R. Weizman (2005). "COMT Val158Met polymorphism in schizophrenia with obsessive-compulsive disorder: a case-control study." Neurosci Lett 389(1): 21-24.

Pruim, R. J., R. P. Welch, S. Sanna, T. M. Teslovich, P. S. Chines, T. P. Gliedt, M. Boehnke, G. R. Abecasis and C. J. Willer (2010). "LocusZoom: regional visualization of genome-wide association scan results." Bioinformatics 26(18): 2336-2337.

308

309

Purcell, S., S. S. Cherny and P. C. Sham (2003). "Genetic Power Calculator: design of linkage and association genetic mapping studies of complex traits." Bioinformatics 19(1): 149-150.

Purcell, S., B. Neale, K. Todd-Brown, L. Thomas, M. A. Ferreira, D. Bender, J. Maller, P. Sklar, P. I. de Bakker, M. J. Daly and P. C. Sham (2007). "PLINK: a tool set for whole-genome association and population-based linkage analyses." Am J Hum Genet 81(3): 559-575.

Qin, H., J. F. Samuels, Y. Wang, Y. Zhu, M. A. Grados, M. A. Riddle, B. D. Greenberg, J. A. Knowles, A. J. Fyer, J. T. McCracken, D. L. Murphy, S. A. Rasmussen, B. A. Cullen, J. Piacentini, D. Geller, S. E. Stewart, D. Pauls, O. J. Bienvenu, F. S. Goes, B. Maher, A. E. Pulver, D. Valle, C. Lange, M. Mattheisen, N. C. McLaughlin, K. Y. Liang, E. L. Nurmi, K. D. Askland, G. Nestadt and Y. Y. Shugart (2015). "Whole-genome association analysis of treatment response in obsessive-compulsive disorder." Mol Psychiatry.

Qin, H.-D., Y. Wang, M. A. Grados, M. A. Riddle, B. D. Greenberg, J. A. Knowles, A. J. Fyer, J. T. McCracken, D. L. Murphy, S. A. Rasmussen, B. Cullen, J. Placentini, D. Geller, D. Pauls, E. Stewart, O. J. Bienvenu, Y. Chen, F. S. Goes, B. Maher, J. F. Samuels, G. Nestadt and Y. Y. Shugart (2013). Whole genome association study identified novel antidepressant response loci for the treatment of obsessive-compulsive disorder with selective sertonin re-uptake inhibitors. 52nd American College of Neuropsychopharmacology, Hollywood, Florida, USA.

Rachman, S. (2009). "Psychological treatment of anxiety: the evolution of behavior therapy and cognitive behavior therapy." Annu Rev Clin Psychol 5: 97-119.

Radua, J., M. Grau, O. A. van den Heuvel, M. Thiebaut de Schotten, D. J. Stein, E. J. Canales- Rodriguez, M. Catani and D. Mataix-Cols (2014). "Multimodal voxel-based meta-analysis of white matter abnormalities in obsessive-compulsive disorder." Neuropsychopharmacology 39(7): 1547-1557.

Rajender, G., M. S. Bhatia, K. Kanwal, S. Malhotra, T. B. Singh and D. Chaudhary (2011). "Study of neurocognitive endophenotypes in drug-naive obsessive-compulsive disorder patients, their first-degree relatives and healthy controls." Acta Psychiatr Scand 124(2): 152-161.

309

310

Raju, T. N. (1999). "The Nobel chronicles. 1958: George Wells Beadle (1903-89), Edward Lawrie Tatum (1909-75) and Joshua Lederberg (b 1925)." Lancet 353(9169): 2082.

Ramamoorthy, S., A. L. Bauman, K. R. Moore, H. Han, T. Yang-Feng, A. S. Chang, V. Ganapathy and R. D. Blakely (1993). "Antidepressant- and cocaine-sensitive human serotonin transporter: molecular cloning, expression, and chromosomal localization." Proc Natl Acad Sci U S A 90(6): 2542-2546.

Ranta, S., Y. Zhang, B. Ross, E. Takkunen, A. Hirvasniemi, A. de la Chapelle, T. C. Gilliam and A. E. Lehesjoki (2000). "Positional cloning and characterisation of the human DLGAP2 gene and its exclusion in progressive epilepsy with mental retardation." Eur J Hum Genet 8(5): 381- 384.

Rasmussen, S. A. and J. L. Eisen (1992). "The epidemiology and clinical features of obsessive compulsive disorder." Psychiatr Clin North Am 15(4): 743-758.

Rauch, S. L. and C. R. Savage (1997). "Neuroimaging and neuropsychology of the striatum. Bridging basic science and clinical practice." Psychiatr Clin North Am 20(4): 741-768.

Real, E., Gratacos, M., Alonso, P., et al. (2010). Pharmacological resistance level in OCD patients without a stressful life event at onset of the disorder is associated with a polymorphism of the glutamate transporter gene (SLC1A1). 18th World Congress of Psychiatric Genetics, Athens, Greece.

Real, E., M. Gratacos, V. Soria, G. Escaramis, P. Alonso, C. Segalas, M. Bayes, R. de Cid, J. M. Menchon and X. Estivill (2009). "A brain-derived neurotrophic factor haplotype is associated with therapeutic response in obsessive-compulsive disorder." Biol Psychiatry 66(7): 674-680.

Redies, C., N. Hertel and C. A. Hubner (2012). "Cadherins and neuropsychiatric disorders." Brain Res 1470: 130-144.

Redrobe, J. P., C. P. MacSweeney and M. Bourin (1996). "The role of 5-HT1A and 5-HT1B receptors in antidepressant drug actions in the mouse forced swimming test." Eur J Pharmacol 318(2-3): 213-220.

310

311

Reich, A., C. Spering, K. Gertz, C. Harms, E. Gerhardt, G. Kronenberg, K. A. Nave, M. Schwab, S. C. Tauber, A. Drinkut, K. Harms, C. P. Beier, A. Voigt, S. Gobbels, M. Endres and J. B. Schulz (2011). "Fas/CD95 regulatory protein Faim2 is neuroprotective after transient brain ischemia." J Neurosci 31(1): 225-233.

Reindl, M., F. Di Pauli, K. Rostasy and T. Berger (2013). "The spectrum of MOG - associated demyelinating diseases." Nat Rev Neurol 9(8): 455-461.

Rende, R., B. Birmaher, D. Axelson, M. Strober, M. K. Gill, S. Valeri, L. Chiappetta, N. Ryan, H. Leonard, J. Hunt, S. Iyengar and M. Keller (2007). "Childhood-onset bipolar disorder: Evidence for increased familial loading of psychiatric illness." J Am Acad Child Adolesc Psychiatry 46(2): 197-204.

Reynolds, G. P. (2012). "The pharmacogenetics of symptom response to antipsychotic drugs." Psychiatry Investig 9(1): 1-7.

Ripke, S., C. O'Dushlaine, K. Chambert, J. L. Moran, A. K. Kahler, S. Akterin, S. E. Bergen, A. L. Collins, J. J. Crowley, M. Fromer, Y. Kim, S. H. Lee, P. K. Magnusson, N. Sanchez, E. A. Stahl, S. Williams, N. R. Wray, K. Xia, F. Bettella, A. D. Borglum, B. K. Bulik-Sullivan, P. Cormican, N. Craddock, C. de Leeuw, N. Durmishi, M. Gill, V. Golimbet, M. L. Hamshere, P. Holmans, D. M. Hougaard, K. S. Kendler, K. Lin, D. W. Morris, O. Mors, P. B. Mortensen, B. M. Neale, F. A. O'Neill, M. J. Owen, M. P. Milovancevic, D. Posthuma, J. Powell, A. L. Richards, B. P. Riley, D. Ruderfer, D. Rujescu, E. Sigurdsson, T. Silagadze, A. B. Smit, H. Stefansson, S. Steinberg, J. Suvisaari, S. Tosato, M. Verhage, J. T. Walters, C. Multicenter Genetic Studies of Schizophrenia, D. F. Levinson, P. V. Gejman, K. S. Kendler, C. Laurent, B. J. Mowry, M. C. O'Donovan, M. J. Owen, A. E. Pulver, B. P. Riley, S. G. Schwab, D. B. Wildenauer, F. Dudbridge, P. Holmans, J. Shi, M. Albus, M. Alexander, D. Campion, D. Cohen, D. Dikeos, J. Duan, P. Eichhammer, S. Godard, M. Hansen, F. B. Lerer, K. Y. Liang, W. Maier, J. Mallet, D. A. Nertney, G. Nestadt, N. Norton, F. A. O'Neill, G. N. Papadimitriou, R. Ribble, A. R. Sanders, J. M. Silverman, D. Walsh, N. M. Williams, B. Wormley, C. Psychosis Endophenotypes International, M. J. Arranz, S. Bakker, S. Bender, E. Bramon, D. Collier, B. Crespo-Facorro, J. Hall, C. Iyegbe, A. Jablensky, R. S. Kahn, L. Kalaydjieva, S. Lawrie, C. M. Lewis, K. Lin, D. H. Linszen, I. Mata, A. McIntosh, R. M. Murray, R. A. Ophoff, J. Powell, D. 311

312

Rujescu, J. Van Os, M. Walshe, M. Weisbrod, D. Wiersma, C. Wellcome Trust Case Control, P. Donnelly, I. Barroso, J. M. Blackwell, E. Bramon, M. A. Brown, J. P. Casas, A. P. Corvin, P. Deloukas, A. Duncanson, J. Jankowski, H. S. Markus, C. G. Mathew, C. N. Palmer, R. Plomin, A. Rautanen, S. J. Sawcer, R. C. Trembath, A. C. Viswanathan, N. W. Wood, C. C. Spencer, G. Band, C. Bellenguez, C. Freeman, G. Hellenthal, E. Giannoulatou, M. Pirinen, R. D. Pearson, A. Strange, Z. Su, D. Vukcevic, P. Donnelly, C. Langford, S. E. Hunt, S. Edkins, R. Gwilliam, H. Blackburn, S. J. Bumpstead, S. Dronov, M. Gillman, E. Gray, N. Hammond, A. Jayakumar, O. T. McCann, J. Liddle, S. C. Potter, R. Ravindrarajah, M. Ricketts, A. Tashakkori-Ghanbaria, M. J. Waller, P. Weston, S. Widaa, P. Whittaker, I. Barroso, P. Deloukas, C. G. Mathew, J. M. Blackwell, M. A. Brown, A. P. Corvin, M. I. McCarthy, C. C. Spencer, E. Bramon, A. P. Corvin, M. C. O'Donovan, K. Stefansson, E. Scolnick, S. Purcell, S. A. McCarroll, P. Sklar, C. M. Hultman and P. F. Sullivan (2013). "Genome-wide association analysis identifies 13 new risk loci for schizophrenia." Nat Genet 45(10): 1150-1159.

Riviere, J. B., L. Xiong, A. Levchenko, J. St-Onge, C. Gaspar, Y. Dion, P. Lesperance, G. Tellier, F. Richer, S. Chouinard, G. A. Rouleau and G. Montreal Tourette Study (2009). "Association of intronic variants of the BTBD9 gene with Tourette syndrome." Arch Neurol 66(10): 1267-1272.

Robins, E. and S. B. Guze (1970). "Establishment of diagnostic validity in psychiatric illness: its application to schizophrenia." Am J Psychiatry 126(7): 983-987.

Rocha, F. F., L. Malloy-Diniz, N. V. Lage and H. Correa (2010). "Positive association between MET allele (BDNF Val66Met polymorphism) and obsessive-compulsive disorder." Rev Bras Psiquiatr 32(3): 323-324.

Roessler, E., Y. Ma, M. V. Ouspenskaia, F. Lacbawan, C. Bendavid, C. Dubourg, P. A. Beachy and M. Muenke (2009). "Truncating loss-of-function mutations of DISP1 contribute to holoprosencephaly-like microform features in humans." Hum Genet 125(4): 393-400.

Roig, B., C. Virgos, N. Franco, L. Martorell, J. Valero, J. Costas, A. Carracedo, A. Labad and E. Vilella (2007). "The discoidin domain receptor 1 as a novel susceptibility gene for schizophrenia." Mol Psychiatry 12(9): 833-841.

312

313

Rosario-Campos, M. C., E. C. Miguel, S. Quatrano, P. Chacon, Y. Ferrao, D. Findley, L. Katsovich, L. Scahill, R. A. King, S. R. Woody, D. Tolin, E. Hollander, Y. Kano and J. F. Leckman (2006). "The Dimensional Yale-Brown Obsessive-Compulsive Scale (DY-BOCS): an instrument for assessing obsessive-compulsive symptom dimensions." Mol Psychiatry 11(5): 495-504.

Rosenberg, D. R., M. S. Keshavan, E. L. Dick, W. W. Bagwell, F. P. MacMaster and B. Birmaher (1997). "Corpus callosal morphology in treatment-naive pediatric obsessive compulsive disorder." Prog Neuropsychopharmacol Biol Psychiatry 21(8): 1269-1283.

Rosenbloom, K. R., C. A. Sloan, V. S. Malladi, T. R. Dreszer, K. Learned, V. M. Kirkup, M. C. Wong, M. Maddren, R. Fang, S. G. Heitner, B. T. Lee, G. P. Barber, R. A. Harte, M. Diekhans, J. C. Long, S. P. Wilder, A. S. Zweig, D. Karolchik, R. M. Kuhn, D. Haussler and W. J. Kent (2013). "ENCODE data in the UCSC Genome Browser: year 5 update." Nucleic Acids Res 41(Database issue): D56-63.

Roses, A. D. (2000). "Pharmacogenetics and the practice of medicine." Nature 405(6788): 857- 865.

Ross, J., J. Badner, H. Garrido, B. Sheppard, D. A. Chavira, M. Grados, J. M. Woo, P. Doo, P. Umana, E. Fournier, S. S. Murray and C. A. Mathews (2011). "Genomewide linkage analysis in Costa Rican families implicates chromosome 15q14 as a candidate region for OCD." Hum Genet 130(6): 795-805.

Rosso, G., U. Albert, G. F. Asinari, F. Bogetto and G. Maina (2012). "Stressful life events and obsessive-compulsive disorder: clinical features and symptom dimensions." Psychiatry Res 197(3): 259-264.

Rotge, J. Y., B. Aouizerate, J. Tignol, B. Bioulac, P. Burbaud and D. Guehl (2010). "The glutamate-based genetic immune hypothesis in obsessive-compulsive disorder. An integrative approach from genes to symptoms." Neuroscience 165(2): 408-417.

313

314

Rothstein, J. D., L. Martin, A. I. Levey, M. Dykes-Hoberg, L. Jin, D. Wu, N. Nash and R. W. Kuncl (1994). "Localization of neuronal and glial glutamate transporters." Neuron 13(3): 713- 725.

Rouquier, S., J. B. Lowe, R. J. Kelly, A. L. Fertitta, G. G. Lennon and D. Giorgi (1995). "Molecular cloning of a human genomic region containing the H blood group alpha(1,2)fucosyltransferase gene and two H locus-related DNA restriction fragments. Isolation of a candidate for the human Secretor blood group locus." J Biol Chem 270(9): 4632-4639.

Rubinsztein, D. C., J. Leggo, M. Chiano, A. Dodge, G. Norbury, E. Rosser and D. Craufurd (1997). "Genotypes at the GluR6 kainate receptor locus are associated with variation in the age of onset of Huntington disease." Proc Natl Acad Sci U S A 94(8): 3872-3876.

Rucklidge, J. J. (2008). "Gender differences in ADHD: implications for psychosocial treatments." Expert Rev Neurother 8(4): 643-655.

Ruscio, A. M., D. J. Stein, W. T. Chiu and R. C. Kessler (2010). "The epidemiology of obsessive-compulsive disorder in the National Comorbidity Survey Replication." Mol Psychiatry 15(1): 53-63.

Sabb, F. W., C. E. Bearden, D. C. Glahn, D. S. Parker, N. Freimer and R. M. Bilder (2008). "A collaborative knowledge base for cognitive phenomics." Mol Psychiatry 13(4): 350-360.

Sadock, B. J., H. I. Kaplan and V. A. Sadock (2007). Kaplan and Sadock's Synopsis of Psychiatry, Lippincott Williams & Wilkins.

Saito, Y., K. Nobuhara, G. Okugawa, K. Takase, T. Sugimoto, M. Horiuchi, C. Ueno, M. Maehara, N. Omura, H. Kurokawa, K. Ikeda, N. Tanigawa, S. Sawada and T. Kinoshita (2008). "Corpus callosum in patients with obsessive-compulsive disorder: diffusion-tensor imaging study." Radiology 246(2): 536-542.

Sampaio, A. S., J. Fagerness, J. Crane, M. Leboyer, R. Delorme, D. L. Pauls and S. E. Stewart (2011). "Association between polymorphisms in GRIK2 gene and obsessive-compulsive disorder: a family-based study." CNS Neurosci Ther 17(3): 141-147.

314

315

Sampaio, A. S., A. G. Hounie, K. Petribu, C. Cappi, I. Morais, H. Vallada, M. C. do Rosario, S. E. Stewart, J. Fargeness, C. Mathews, P. Arnold, G. L. Hanna, M. Richter, J. Kennedy, L. Fontenelle, C. A. de Braganca Pereira, D. L. Pauls and E. C. Miguel (2015). "COMT and MAO- A polymorphisms and obsessive-compulsive disorder: a family-based association study." PLoS One 10(3): e0119592.

Samuels, J., Y. Y. Shugart, M. A. Grados, V. L. Willour, O. J. Bienvenu, B. D. Greenberg, J. A. Knowles, J. T. McCracken, S. L. Rauch, D. L. Murphy, Y. Wang, A. Pinto, A. J. Fyer, J. Piacentini, D. L. Pauls, B. Cullen, S. A. Rasmussen, R. Hoehn-Saric, D. Valle, K. Y. Liang, M. A. Riddle and G. Nestadt (2007). "Significant linkage to on chromosome 14 in families with obsessive-compulsive disorder: results from the OCD Collaborative Genetics Study." Am J Psychiatry 164(3): 493-499.

Samuels, J. F., O. J. Bienvenu, 3rd, A. Pinto, A. J. Fyer, J. T. McCracken, S. L. Rauch, D. L. Murphy, M. A. Grados, B. D. Greenberg, J. A. Knowles, J. Piacentini, P. A. Cannistraro, B. Cullen, M. A. Riddle, S. A. Rasmussen, D. L. Pauls, V. L. Willour, Y. Y. Shugart, K. Y. Liang, R. Hoehn-Saric and G. Nestadt (2007). "Hoarding in obsessive-compulsive disorder: results from the OCD Collaborative Genetics Study." Behav Res Ther 45(4): 673-686.

Sanavio, E. (1988). "Obsessions and compulsions: the Padua Inventory." Behav Res Ther 26(2): 169-177.

Satoh, K., H. Yanai, T. Senda, K. Kohu, T. Nakamura, N. Okumura, A. Matsumine, S. Kobayashi, K. Toyoshima and T. Akiyama (1997). "DAP-1, a novel protein that interacts with the guanylate kinase-like domains of hDLG and PSD-95." Genes Cells 2(6): 415-424.

Satzinger, H. (2008). "Theodor and Marcella Boveri: chromosomes and cytoplasm in heredity and development." Nat Rev Genet 9(3): 231-238.

Saxena, S. (2011). "Pharmacotherapy of compulsive hoarding." J Clin Psychol 67(5): 477-484.

Saxena, S., A. L. Brody, J. M. Schwartz and L. R. Baxter (1998). "Neuroimaging and frontal- subcortical circuitry in obsessive-compulsive disorder." Br J Psychiatry Suppl(35): 26-37.

315

316

Saxena, S. and S. L. Rauch (2000). "Functional neuroimaging and the neuroanatomy of obsessive-compulsive disorder." Psychiatr Clin North Am 23(3): 563-586.

Scheggia, D., S. Sannino, M. L. Scattoni and F. Papaleo (2012). "COMT as a drug target for cognitive functions and dysfunctions." CNS Neurol Disord Drug Targets 11(3): 209-221.

Schindler, K. M., M. A. Richter, J. L. Kennedy, M. T. Pato and C. N. Pato (2000). "Association between homozygosity at the COMT gene locus and obsessive compulsive disorder." Am J Med Genet 96(6): 721-724.

Schizophrenia Working Group of the Psychiatric Genomics, C. (2014). "Biological insights from 108 schizophrenia-associated genetic loci." Nature 511(7510): 421-427.

Schooler, C., A. J. Revell, K. R. Timpano, M. Wheaton and D. L. Murphy (2008). "Predicting genetic loading from symptom patterns in obsessive- compulsive disorder: a latent variable analysis." Depress Anxiety 25(8): 680-688.

Schulze, T. G., H. Fangerau and P. Propping (2004). "From degeneration to genetic susceptibility, from eugenics to genethics, from Bezugsziffer to LOD score: the history of psychiatric genetics." Int Rev Psychiatry 16(4): 246-259.

Scimemi, A., H. Tian and J. S. Diamond (2009). "Neuronal transporters regulate glutamate clearance, NMDA receptor activation, and synaptic plasticity in the hippocampus." J Neurosci 29(46): 14581-14595.

Sen, S., R. M. Nesse, S. F. Stoltenberg, S. Li, L. Gleiberman, A. Chakravarti, A. B. Weder and M. Burmeister (2003). "A BDNF coding variant is associated with the NEO personality inventory domain neuroticism, a risk factor for depression." Neuropsychopharmacology 28(2): 397-401.

Serretti, A., M. Kato and J. L. Kennedy (2008). "Pharmacogenetic studies in depression: a proposal for methodologic guidelines." Pharmacogenomics J 8(2): 90-100.

316

317

Setou, M., T. Nakagawa, D. H. Seog and N. Hirokawa (2000). "Kinesin superfamily motor protein KIF17 and mLin-10 in NMDA receptor-containing vesicle transport." Science 288(5472): 1796-1802.

Shaffer, L. G., A. Theisen, B. A. Bejjani, B. C. Ballif, A. S. Aylsworth, C. Lim, M. McDonald, J. W. Ellison, D. Kostiner, S. Saitta and T. Shaikh (2007). "The discovery of microdeletion syndromes in the post-genomic era: review of the methodology and characterization of a new 1q41q42 microdeletion syndrome." Genet Med 9(9): 607-616.

Sham, P. C. and S. M. Purcell (2014). "Statistical power and significance testing in large-scale genetic studies." Nat Rev Genet 15(5): 335-346.

Sharma, V. and D. Biswas (2012). "Cobalamin deficiency presenting as obsessive compulsive disorder: case report." Gen Hosp Psychiatry 34(5): 578 e577-578.

Sheehan, D. V., Y. Lecrubier, K. H. Sheehan, P. Amorim, J. Janavs, E. Weiller, T. Hergueta, R. Baker and G. C. Dunbar (1998). "The Mini-International Neuropsychiatric Interview (M.I.N.I.): the development and validation of a structured diagnostic psychiatric interview for DSM-IV and ICD-10." J Clin Psychiatry 59 Suppl 20: 22-33;quiz 34-57.

Shibata, H., A. Shibata, H. Ninomiya, N. Tashiro and Y. Fukumaki (2002). "Association study of polymorphisms in the GluR6 kainate receptor gene (GRIK2) with schizophrenia." Psychiatry Res 113(1-2): 59-67.

Shmelkov, S. V., A. Hormigo, D. Jing, C. C. Proenca, K. G. Bath, T. Milde, E. Shmelkov, J. S. Kushner, M. Baljevic, I. Dincheva, A. J. Murphy, D. M. Valenzuela, N. W. Gale, G. D. Yancopoulos, I. Ninan, F. S. Lee and S. Rafii (2010). "Slitrk5 deficiency impairs corticostriatal circuitry and leads to obsessive-compulsive-like behaviors in mice." Nat Med 16(5): 598-602, 591p following 602.

Shuang, M., J. Liu, M. X. Jia, J. Z. Yang, S. P. Wu, X. H. Gong, Y. S. Ling, Y. Ruan, X. L. Yang and D. Zhang (2004). "Family-based association study between autism and glutamate receptor 6 gene in Chinese Han trios." Am J Med Genet B Neuropsychiatr Genet 131B(1): 48- 50.

317

318

Shuchman, M., P. C. Hebert, R. Kale, B. Sibbald, K. Flegel and N. MacDonald (2008). "Bringing a research base to psychiatry." CMAJ 178(10): 1257-1260.

Shugart, Y. Y., Y. Wang, J. F. Samuels, M. A. Grados, B. D. Greenberg, J. A. Knowles, J. T. McCracken, S. L. Rauch, D. L. Murphy, S. A. Rasmussen, B. Cullen, R. Hoehn-Saric, A. Pinto, A. J. Fyer, J. Piacentini, D. L. Pauls, O. J. Bienvenu, M. A. Riddle, K. Y. Liang and G. Nestadt (2009). "A family-based association study of the glutamate transporter gene SLC1A1 in obsessive-compulsive disorder in 378 families." Am J Med Genet B Neuropsychiatr Genet 150B(6): 886-892.

Singh, A. B., C. A. Bousman, C. Ng and M. Berk (2014). "Antidepressant pharmacogenetics." Curr Opin Psychiatry 27(1): 43-51.

Skarphedinsson, G., S. Compton, P. H. Thomsen, B. Weidle, K. Dahl, J. B. Nissen, N. C. Torp, K. Hybel, K. H. Melin, R. Valderhaug, T. Wentzel-Larsen and T. Ivarsson (2015). "Tics Moderate Sertraline, but Not Cognitive-Behavior Therapy Response in Pediatric Obsessive- Compulsive Disorder Patients Who Do Not Respond to Cognitive-Behavior Therapy." J Child Adolesc Psychopharmacol 25(5): 432-439.

Sklar, P., S. B. Gabriel, M. G. McInnis, P. Bennett, Y. Lim, G. Tsan, S. Schaffner, G. Kirov, I. Jones, M. Owen, N. Craddock, J. R. DePaulo and E. S. Lander (2002). "Family-based association study of 76 candidate genes in bipolar disorder: BDNF is a potential risk locus. Brain-derived neutrophic factor." Mol Psychiatry 7(6): 579-593.

Slassi, A. (2002). "Recent advances in 5-HT1B/1D receptor antagonists and agonists and their potential therapeutic applications." Curr Top Med Chem 2(6): 559-574.

Smith, C. P., S. Weremowicz, Y. Kanai, M. Stelzner, C. C. Morton and M. A. Hediger (1994). "Assignment of the gene coding for the human high-affinity glutamate transporter EAAC1 to 9p24: potential role in dicarboxylic aminoaciduria and neurodegenerative disorders." Genomics 20(2): 335-336.

Snider, L. A. and S. E. Swedo (2003). "Post-streptococcal autoimmune disorders of the central nervous system." Curr Opin Neurol 16(3): 359-365.

318

319

Somia, N. V., M. J. Schmitt, D. E. Vetter, D. Van Antwerp, S. F. Heinemann and I. M. Verma (1999). "LFG: an anti-apoptotic gene that provides protection from Fas-mediated cell death." Proc Natl Acad Sci U S A 96(22): 12667-12672.

Soomro, G. M., D. Altman, S. Rajagopal and M. Oakley-Browne (2008). "Selective serotonin re- uptake inhibitors (SSRIs) versus placebo for obsessive compulsive disorder (OCD)." Cochrane Database Syst Rev(1): CD001765.

Sparkes, R. S., N. Lan, I. Klisak, T. Mohandas, A. Diep, T. Kojis, C. Heinzmann and J. C. Shih (1991). "Assignment of a serotonin 5HT-2 receptor gene (HTR2) to human chromosome 13q14- q21 and mouse chromosome 14." Genomics 9(3): 461-465.

Speliotes, E. K., C. J. Willer, S. I. Berndt, K. L. Monda, G. Thorleifsson, A. U. Jackson, H. Lango Allen, C. M. Lindgren, J. Luan, R. Magi, J. C. Randall, S. Vedantam, T. W. Winkler, L. Qi, T. Workalemahu, I. M. Heid, V. Steinthorsdottir, H. M. Stringham, M. N. Weedon, E. Wheeler, A. R. Wood, T. Ferreira, R. J. Weyant, A. V. Segre, K. Estrada, L. Liang, J. Nemesh, J. H. Park, S. Gustafsson, T. O. Kilpelainen, J. Yang, N. Bouatia-Naji, T. Esko, M. F. Feitosa, Z. Kutalik, M. Mangino, S. Raychaudhuri, A. Scherag, A. V. Smith, R. Welch, J. H. Zhao, K. K. Aben, D. M. Absher, N. Amin, A. L. Dixon, E. Fisher, N. L. Glazer, M. E. Goddard, N. L. Heard-Costa, V. Hoesel, J. J. Hottenga, A. Johansson, T. Johnson, S. Ketkar, C. Lamina, S. Li, M. F. Moffatt, R. H. Myers, N. Narisu, J. R. Perry, M. J. Peters, M. Preuss, S. Ripatti, F. Rivadeneira, C. Sandholt, L. J. Scott, N. J. Timpson, J. P. Tyrer, S. van Wingerden, R. M. Watanabe, C. C. White, F. Wiklund, C. Barlassina, D. I. Chasman, M. N. Cooper, J. O. Jansson, R. W. Lawrence, N. Pellikka, I. Prokopenko, J. Shi, E. Thiering, H. Alavere, M. T. Alibrandi, P. Almgren, A. M. Arnold, T. Aspelund, L. D. Atwood, B. Balkau, A. J. Balmforth, A. J. Bennett, Y. Ben-Shlomo, R. N. Bergman, S. Bergmann, H. Biebermann, A. I. Blakemore, T. Boes, L. L. Bonnycastle, S. R. Bornstein, M. J. Brown, T. A. Buchanan, F. Busonero, H. Campbell, F. P. Cappuccio, C. Cavalcanti-Proenca, Y. D. Chen, C. M. Chen, P. S. Chines, R. Clarke, L. Coin, J. Connell, I. N. Day, M. den Heijer, J. Duan, S. Ebrahim, P. Elliott, R. Elosua, G. Eiriksdottir, M. R. Erdos, J. G. Eriksson, M. F. Facheris, S. B. Felix, P. Fischer-Posovszky, A. R. Folsom, N. Friedrich, N. B. Freimer, M. Fu, S. Gaget, P. V. Gejman, E. J. Geus, C. Gieger, A. P. Gjesing, A. Goel, P. Goyette, H. Grallert, J. Grassler, D. M. Greenawalt, C. J. Groves, V. Gudnason, C.

319

320

Guiducci, A. L. Hartikainen, N. Hassanali, A. S. Hall, A. S. Havulinna, C. Hayward, A. C. Heath, C. Hengstenberg, A. A. Hicks, A. Hinney, A. Hofman, G. Homuth, J. Hui, W. Igl, C. Iribarren, B. Isomaa, K. B. Jacobs, I. Jarick, E. Jewell, U. John, T. Jorgensen, P. Jousilahti, A. Jula, M. Kaakinen, E. Kajantie, L. M. Kaplan, S. Kathiresan, J. Kettunen, L. Kinnunen, J. W. Knowles, I. Kolcic, I. R. Konig, S. Koskinen, P. Kovacs, J. Kuusisto, P. Kraft, K. Kvaloy, J. Laitinen, O. Lantieri, C. Lanzani, L. J. Launer, C. Lecoeur, T. Lehtimaki, G. Lettre, J. Liu, M. L. Lokki, M. Lorentzon, R. N. Luben, B. Ludwig, Magic, P. Manunta, D. Marek, M. Marre, N. G. Martin, W. L. McArdle, A. McCarthy, B. McKnight, T. Meitinger, O. Melander, D. Meyre, K. Midthjell, G. W. Montgomery, M. A. Morken, A. P. Morris, R. Mulic, J. S. Ngwa, M. Nelis, M. J. Neville, D. R. Nyholt, C. J. O'Donnell, S. O'Rahilly, K. K. Ong, B. Oostra, G. Pare, A. N. Parker, M. Perola, I. Pichler, K. H. Pietilainen, C. G. Platou, O. Polasek, A. Pouta, S. Rafelt, O. Raitakari, N. W. Rayner, M. Ridderstrale, W. Rief, A. Ruokonen, N. R. Robertson, P. Rzehak, V. Salomaa, A. R. Sanders, M. S. Sandhu, S. Sanna, J. Saramies, M. J. Savolainen, S. Scherag, S. Schipf, S. Schreiber, H. Schunkert, K. Silander, J. Sinisalo, D. S. Siscovick, J. H. Smit, N. Soranzo, U. Sovio, J. Stephens, I. Surakka, A. J. Swift, M. L. Tammesoo, J. C. Tardif, M. Teder- Laving, T. M. Teslovich, J. R. Thompson, B. Thomson, A. Tonjes, T. Tuomi, J. B. van Meurs, G. J. van Ommen, V. Vatin, J. Viikari, S. Visvikis-Siest, V. Vitart, C. I. Vogel, B. F. Voight, L. L. Waite, H. Wallaschofski, G. B. Walters, E. Widen, S. Wiegand, S. H. Wild, G. Willemsen, D. R. Witte, J. C. Witteman, J. Xu, Q. Zhang, L. Zgaga, A. Ziegler, P. Zitting, J. P. Beilby, I. S. Farooqi, J. Hebebrand, H. V. Huikuri, A. L. James, M. Kahonen, D. F. Levinson, F. Macciardi, M. S. Nieminen, C. Ohlsson, L. J. Palmer, P. M. Ridker, M. Stumvoll, J. S. Beckmann, H. Boeing, E. Boerwinkle, D. I. Boomsma, M. J. Caulfield, S. J. Chanock, F. S. Collins, L. A. Cupples, G. D. Smith, J. Erdmann, P. Froguel, H. Gronberg, U. Gyllensten, P. Hall, T. Hansen, T. B. Harris, A. T. Hattersley, R. B. Hayes, J. Heinrich, F. B. Hu, K. Hveem, T. Illig, M. R. Jarvelin, J. Kaprio, F. Karpe, K. T. Khaw, L. A. Kiemeney, H. Krude, M. Laakso, D. A. Lawlor, A. Metspalu, P. B. Munroe, W. H. Ouwehand, O. Pedersen, B. W. Penninx, A. Peters, P. P. Pramstaller, T. Quertermous, T. Reinehr, A. Rissanen, I. Rudan, N. J. Samani, P. E. Schwarz, A. R. Shuldiner, T. D. Spector, J. Tuomilehto, M. Uda, A. Uitterlinden, T. T. Valle, M. Wabitsch, G. Waeber, N. J. Wareham, H. Watkins, C. Procardis, J. F. Wilson, A. F. Wright, M. C. Zillikens, N. Chatterjee, S. A. McCarroll, S. Purcell, E. E. Schadt, P. M. Visscher, T. L. Assimes, I. B. Borecki, P. Deloukas, C. S. Fox, L. C. Groop, T. Haritunians, D. J. Hunter, R. C. Kaplan, K.

320

321

L. Mohlke, J. R. O'Connell, L. Peltonen, D. Schlessinger, D. P. Strachan, C. M. van Duijn, H. E. Wichmann, T. M. Frayling, U. Thorsteinsdottir, G. R. Abecasis, I. Barroso, M. Boehnke, K. Stefansson, K. E. North, M. I. McCarthy, J. N. Hirschhorn, E. Ingelsson and R. J. Loos (2010). "Association analyses of 249,796 individuals reveal 18 new loci associated with body mass index." Nat Genet 42(11): 937-948.

Stein, D. J. (2013). "What is a mental disorder? A perspective from cognitive-affective science." Can J Psychiatry 58(12): 656-662.

Stein, D. J., E. W. Andersen and K. F. Overo (2007). "Response of symptom dimensions in obsessive-compulsive disorder to treatment with citalopram or placebo." Rev Bras Psiquiatr 29(4): 303-307.

Stein, D. J., P. D. Carey, C. Lochner, S. Seedat, N. Fineberg and E. W. Andersen (2008). "Escitalopram in obsessive-compulsive disorder: response of symptom dimensions to pharmacotherapy." CNS Spectr 13(6): 492-498.

Stein, D. J., J. E. Grant, M. E. Franklin, N. Keuthen, C. Lochner, H. S. Singer and D. W. Woods (2010). "Trichotillomania (hair pulling disorder), skin picking disorder, and stereotypic movement disorder: toward DSM-V." Depress Anxiety 27(6): 611-626.

Steketee, G., R. Frost and K. Bogart (1996). "The Yale-Brown Obsessive Compulsive Scale: interview versus self-report." Behav Res Ther 34(8): 675-684.

Stergiakouli, E., J. Martin, M. L. Hamshere, K. Langley, D. M. Evans, B. St Pourcain, N. J. Timpson, M. J. Owen, M. O'Donovan, A. Thapar and G. Davey Smith (2015). "Shared genetic influences between attention-deficit/hyperactivity disorder (ADHD) traits in children and clinical ADHD." J Am Acad Child Adolesc Psychiatry 54(4): 322-327.

Stewart, S. E., J. A. Fagerness, J. Platko, J. W. Smoller, J. M. Scharf, C. Illmann, E. Jenike, N. Chabane, M. Leboyer, R. Delorme, M. A. Jenike and D. L. Pauls (2007). "Association of the SLC1A1 glutamate transporter gene and obsessive-compulsive disorder." Am J Med Genet B Neuropsychiatr Genet 144B(8): 1027-1033.

321

322

Stewart, S. E., C. Mayerfeld, P. D. Arnold, J. R. Crane, C. O'Dushlaine, J. A. Fagerness, D. Yu, J. M. Scharf, E. Chan, F. Kassam, P. R. Moya, J. R. Wendland, R. Delorme, M. A. Richter, J. L. Kennedy, J. Veenstra-VanderWeele, J. Samuels, B. D. Greenberg, J. T. McCracken, J. A. Knowles, A. J. Fyer, S. L. Rauch, M. A. Riddle, M. A. Grados, O. J. Bienvenu, B. Cullen, Y. Wang, Y. Y. Shugart, J. Piacentini, S. Rasmussen, G. Nestadt, D. L. Murphy, M. A. Jenike, E. H. Cook, D. L. Pauls, G. L. Hanna and C. A. Mathews (2013). "Meta-analysis of association between obsessive-compulsive disorder and the 3' region of neuronal glutamate transporter gene SLC1A1." Am J Med Genet B Neuropsychiatr Genet 162B(4): 367-379.

Stewart, S. E., J. Platko, J. Fagerness, J. Birns, E. Jenike, J. W. Smoller, R. Perlis, M. Leboyer, R. Delorme, N. Chabane, S. L. Rauch, M. A. Jenike and D. L. Pauls (2007). "A genetic family- based association study of OLIG2 in obsessive-compulsive disorder." Arch Gen Psychiatry 64(2): 209-214.

Stewart, S. E., M. C. Rosario, L. Baer, A. S. Carter, T. A. Brown, J. M. Scharf, C. Illmann, J. F. Leckman, D. Sukhodolsky, L. Katsovich, S. Rasmussen, W. Goodman, R. Delorme, M. Leboyer, N. Chabane, M. A. Jenike, D. A. Geller and D. L. Pauls (2008). "Four-factor structure of obsessive-compulsive disorder symptoms in children, adolescents, and adults." J Am Acad Child Adolesc Psychiatry 47(7): 763-772.

Stewart, S. E., D. Yu, J. M. Scharf, B. M. Neale, J. A. Fagerness, C. A. Mathews, P. D. Arnold, P. D. Evans, E. R. Gamazon, L. K. Davis, L. Osiecki, L. McGrath, S. Haddad, J. Crane, D. Hezel, C. Illman, C. Mayerfeld, A. Konkashbaev, C. Liu, A. Pluzhnikov, A. Tikhomirov, C. K. Edlund, S. L. Rauch, R. Moessner, P. Falkai, W. Maier, S. Ruhrmann, H. J. Grabe, L. Lennertz, M. Wagner, L. Bellodi, M. C. Cavallini, M. A. Richter, E. H. Cook, Jr., J. L. Kennedy, D. Rosenberg, D. J. Stein, S. M. Hemmings, C. Lochner, A. Azzam, D. A. Chavira, E. Fournier, H. Garrido, B. Sheppard, P. Umana, D. L. Murphy, J. R. Wendland, J. Veenstra-VanderWeele, D. Denys, R. Blom, D. Deforce, F. Van Nieuwerburgh, H. G. Westenberg, S. Walitza, K. Egberts, T. Renner, E. C. Miguel, C. Cappi, A. G. Hounie, M. Conceicao do Rosario, A. S. Sampaio, H. Vallada, H. Nicolini, N. Lanzagorta, B. Camarena, R. Delorme, M. Leboyer, C. N. Pato, M. T. Pato, E. Voyiaziakis, P. Heutink, D. C. Cath, D. Posthuma, J. H. Smit, J. Samuels, O. J. Bienvenu, B. Cullen, A. J. Fyer, M. A. Grados, B. D. Greenberg, J. T. McCracken, M. A. Riddle,

322

323

Y. Wang, V. Coric, J. F. Leckman, M. Bloch, C. Pittenger, V. Eapen, D. W. Black, R. A. Ophoff, E. Strengman, D. Cusi, M. Turiel, F. Frau, F. Macciardi, J. R. Gibbs, M. R. Cookson, A. Singleton, C. North American Brain Expression, J. Hardy, U. K. B. E. Database, A. T. Crenshaw, M. A. Parkin, D. B. Mirel, D. V. Conti, S. Purcell, G. Nestadt, G. L. Hanna, M. A. Jenike, J. A. Knowles, N. Cox and D. L. Pauls (2013). "Genome-wide association study of obsessive-compulsive disorder." Mol Psychiatry 18(7): 788-798.

Strober, M., W. Morrell, J. Burroughs, C. Lampert, H. Danforth and R. Freeman (1988). "A family study of bipolar I disorder in adolescence. Early onset of symptoms linked to increased familial loading and lithium resistance." J Affect Disord 15(3): 255-268.

Studer, L., C. Spenger, R. W. Seiler, C. A. Altar, R. M. Lindsay and C. Hyman (1995). "Comparison of the effects of the neurotrophins on the morphological structure of dopaminergic neurons in cultures of rat substantia nigra." Eur J Neurosci 7(2): 223-233.

Suen, P. C., K. Wu, E. S. Levine, H. T. Mount, J. L. Xu, S. Y. Lin and I. B. Black (1997). "Brain-derived neurotrophic factor rapidly enhances phosphorylation of the postsynaptic N- methyl-D-aspartate receptor subunit 1." Proc Natl Acad Sci U S A 94(15): 8191-8195.

Suliman, S., S. M. Hemmings and S. Seedat (2013). "Brain-Derived Neurotrophic Factor (BDNF) protein levels in anxiety disorders: systematic review and meta-regression analysis." Front Integr Neurosci 7: 55.

Sullivan, P. F., M. C. Neale and K. S. Kendler (2000). "Genetic epidemiology of major depression: review and meta-analysis." Am J Psychiatry 157(10): 1552-1562.

Sumitani, S., S. Ueno, Y. Ishimoto, T. Taniguchi, M. Tomotake, I. Motoki, K. Yamauchi and T. Ohmori (2006). "[Clinical features of patients with obsessive-compulsive disorder showing different pharmacological responses]." Seishin Shinkeigaku Zasshi 108(12): 1282-1292.

Summerfeldt, L. J., M. A. Richter, M. M. Anthony, V. M. Huta and R. P. Swinson (1997). Symptom structure in OCD: factor analytic evidence for subgroups. 1997 Annual Meeting New Research Program and Abstracts, Washington, DC, American Psychiatric Association.

323

324

Summerfeldt, L. J., M. A. Richter, M. M. Antony and R. P. Swinson (1999). "Symptom structure in obsessive-compulsive disorder: a confirmatory factor-analytic study." Behav Res Ther 37(4): 297-311.

Swedo, S. E., J. F. Leckman and N. R. Rose (2012). "From Research Subgroup to Clinical Syndrome: Modifying the PANDAS

Criteria to Describe PANS (Pediatric Acute-onset Neuropsychiatric

Syndrome)." Pediatrics & Therapeutics 2(2): 113.

Szeszko, P. R., B. A. Ardekani, M. Ashtari, A. K. Malhotra, D. G. Robinson, R. M. Bilder and K. O. Lim (2005). "White matter abnormalities in obsessive-compulsive disorder: a diffusion tensor imaging study." Arch Gen Psychiatry 62(7): 782-790.

Taj, M. J. R., B. Viswanath, M. Purushottam, T. Kandavel, Y. C. Janardhan Reddy and S. Jain (2013). "DRD4 gene and obsessive compulsive disorder: do symptom dimensions have specific genetic correlates?" Prog Neuropsychopharmacol Biol Psychiatry 41: 18-23.

Takebayashi, H., Y. Nabeshima, S. Yoshida, O. Chisaka, K. Ikenaka and Y. Nabeshima (2002). "The basic helix-loop-helix factor olig2 is essential for the development of motoneuron and oligodendrocyte lineages." Curr Biol 12(13): 1157-1163.

Tanaka, T., P. Scheet, B. Giusti, S. Bandinelli, M. G. Piras, G. Usala, S. Lai, A. Mulas, A. M. Corsi, A. Vestrini, F. Sofi, A. M. Gori, R. Abbate, J. Guralnik, A. Singleton, G. R. Abecasis, D. Schlessinger, M. Uda and L. Ferrucci (2009). "Genome-wide association study of vitamin B6, vitamin B12, folate, and homocysteine blood concentrations." Am J Hum Genet 84(4): 477-482.

Taylor, S. (2011). "Early versus late onset obsessive-compulsive disorder: evidence for distinct subtypes." Clin Psychol Rev 31(7): 1083-1100.

Taylor, S. (2013). "Molecular genetics of obsessive-compulsive disorder: a comprehensive meta- analysis of genetic association studies." Mol Psychiatry 18(7): 799-805.

324

325

Team, R. C. (2013). R: A language and environment for statistical computing. Vienna, Austria, R Foundation for Statistical Computing.

Tek, C. and B. Ulug (2001). "Religiosity and religious obsessions in obsessive-compulsive disorder." Psychiatry Res 104(2): 99-108.

Thomas, K. L., S. Davis, S. P. Hunt and S. Laroche (1996). "Alterations in the expression of specific glutamate receptor subunits following hippocampal LTP in vivo." Learn Mem 3(2-3): 197-208.

Thordarson, D. S., A. S. Radomsky, S. Rachman, R. Shafran, C. N. Sawchuk and A. Ralph Hakstian (2004). "The Vancouver Obsessional Compulsive Inventory (VOCI)." Behav Res Ther 42(11): 1289-1314.

Timpano, K. R., N. B. Schmidt, M. G. Wheaton, J. R. Wendland and D. L. Murphy (2011). "Consideration of the BDNF gene in relation to two phenotypes: hoarding and obesity." J Abnorm Psychol 120(3): 700-707.

Ting, J. T. and G. Feng (2011). "Neurobiology of obsessive-compulsive disorder: insights into neural circuitry dysfunction through mouse genetics." Curr Opin Neurobiol 21(6): 842-848.

Tiwari, A. K., R. P. Souza and D. J. Muller (2009). "Pharmacogenetics of anxiolytic drugs." J Neural Transm 116(6): 667-677.

Tolin, D. F., J. S. Abramowitz and G. J. Diefenbach (2005). "Defining response in clinical trials for obsessive-compulsive disorder: a signal detection analysis of the Yale-Brown obsessive compulsive scale." J Clin Psychiatry 66(12): 1549-1557.

Torres, A. R., L. F. Fontenelle, Y. A. Ferrao, M. C. do Rosario, R. C. Torresan, E. C. Miguel and R. G. Shavitt (2012). "Clinical features of obsessive-compulsive disorder with hoarding symptoms: a multicenter study." J Psychiatr Res 46(6): 724-732.

Torresan, R. C., A. T. Ramos-Cerqueira, R. G. Shavitt, M. C. do Rosario, M. A. de Mathis, E. C. Miguel and A. R. Torres (2013). "Symptom dimensions, clinical course and comorbidity in men and women with obsessive-compulsive disorder." Psychiatry Res 209(2): 186-195.

325

326

Tu, H. P., A. M. Ko, S. J. Wang, C. H. Lee, R. A. Lea, S. L. Chiang, H. C. Chiang, T. N. Wang, M. C. Huang, T. T. Ou, G. T. Lin and Y. C. Ko (2010). "Monoamine oxidase A gene polymorphisms and enzyme activity associated with risk of gout in Taiwan aborigines." Hum Genet 127(2): 223-229.

Tukel, R., O. Bozkurt, A. Polat, A. Genc and H. Atli (2006). "Clinical predictors of response to pharmacotherapy with selective serotonin reuptake inhibitors in obsessive-compulsive disorder." Psychiatry Clin Neurosci 60(4): 404-409.

Tukel, R., E. Ertekin, S. Batmaz, F. Alyanak, A. Sozen, B. Aslantas, H. Atli and I. Ozyildirim (2005). "Influence of age of onset on clinical features in obsessive-compulsive disorder." Depress Anxiety 21(3): 112-117.

Tukel, R., H. Gurvit, B. A. Ertekin, S. Oflaz, E. Ertekin, B. Baran, S. A. Kalem, P. E. Kandemir, F. A. Ozdemiroglu and F. Atalay (2012). "Neuropsychological function in obsessive-compulsive disorder." Compr Psychiatry 53(2): 167-175.

Tukel, R., H. Gurvit, B. Ozata, N. Ozturk, B. A. Ertekin, E. Ertekin, B. Baran, S. A. Kalem, D. Buyukgok and G. S. Direskeneli (2012). "Brain-derived neurotrophic factor gene Val66Met polymorphism and cognitive function in obsessive-compulsive disorder." Am J Med Genet B Neuropsychiatr Genet 159B(7): 850-858.

Tukel, R., H. Gurvit, N. Ozturk, B. Ozata, B. A. Ertekin, E. Ertekin, B. Baran, S. A. Kalem, D. Buyukgok and G. S. Direskeneli (2013). "COMT Val158Met polymorphism and executive functions in obsessive-compulsive disorder." J Neuropsychiatry Clin Neurosci 25(3): 214-221.

Tukel, R., B. Ozata, N. Ozturk, B. A. Ertekin, E. Ertekin and G. S. Direskeneli (2014). "The role of the brain-derived neurotrophic factor SNP rs2883187 in the phenotypic expression of obsessive-compulsive disorder." J Clin Neurosci 21(5): 790-793.

Tukel, R., A. Polat, A. Genc, O. Bozkurt and H. Atli (2004). "Gender-related differences among Turkish patients with obsessive-compulsive disorder." Compr Psychiatry 45(5): 362-366.

326

327

Turksoy, N., R. Bilici, A. Yalciner, Y. O. Ozdemir, I. Ornek, A. E. Tufan and A. Kara (2014). "Vitamin B12, folate, and homocysteine levels in patients with obsessive-compulsive disorder." Neuropsychiatr Dis Treat 10: 1671-1675.

Tyler, W. J., M. Alonso, C. R. Bramham and L. D. Pozzo-Miller (2002). "From acquisition to consolidation: on the role of brain-derived neurotrophic factor signaling in hippocampal- dependent learning." Learn Mem 9(5): 224-237.

Ullmer, C., K. Schmuck, A. Figge and H. Lubbert (1998). "Cloning and characterization of MUPP1, a novel PDZ domain protein." FEBS Lett 424(1-2): 63-68.

Umehara, H., S. Numata, A. Tajima, M. Kinoshita, S. Nakaaki, I. Imoto, S. Sumitani and T. Ohmori (2015). "No association between the COMT Val158Met polymorphism and the long- term clinical response in obsessive-compulsive disorder in the Japanese population." Hum Psychopharmacol. van Grootheest, D. S., D. I. Boomsma, J. M. Hettema and K. S. Kendler (2008). "Heritability of obsessive-compulsive symptom dimensions." Am J Med Genet B Neuropsychiatr Genet 147B(4): 473-478. van Grootheest, D. S., D. C. Cath, A. T. Beekman and D. I. Boomsma (2005). "Twin studies on obsessive-compulsive disorder: a review." Twin Res Hum Genet 8(5): 450-458.

Van Oppen, P., R. J. Hoekstra and P. M. Emmelkamp (1995). "The structure of obsessive- compulsive symptoms." Behav Res Ther 33(1): 15-23.

Veenstra-VanderWeele, J., S. J. Kim, D. Gonen, G. L. Hanna, B. L. Leventhal and E. H. Cook, Jr. (2001). "Genomic organization of the SLC1A1/EAAC1 gene and mutation screening in early- onset obsessive-compulsive disorder." Mol Psychiatry 6(2): 160-167.

Vetti, H. H., A. Molven, A. K. Eliassen and V. M. Steen (2010). "Is pharmacogenetic CYP2D6 testing useful?" Tidsskr Nor Laegeforen 130(22): 2224-2228.

327

328

Viswanath, B., Y. C. Janardhan Reddy, K. J. Kumar, T. Kandavel and C. R. Chandrashekar (2009). "Cognitive endophenotypes in OCD: a study of unaffected siblings of probands with familial OCD." Prog Neuropsychopharmacol Biol Psychiatry 33(4): 610-615.

Viswanath, B., J. C. Narayanaswamy, A. V. Cherian, Y. C. Reddy and S. B. Math (2011). "Is familial obsessive-compulsive disorder different from sporadic obsessive-compulsive disorder? A comparison of clinical characteristics, comorbidity and treatment response." Psychopathology 44(2): 83-89.

Viswanath, B., M. J. R. Taj, M. Purushottam, T. Kandavel, P. H. Shetty, Y. C. Reddy and S. Jain (2013). "No association between DRD4 gene and SRI treatment response in obsessive compulsive disorder: need for a novel approach." Asian J Psychiatr 6(4): 347-348.

Voyiaziakis, E., O. Evgrafov, D. Li, H. J. Yoon, P. Tabares, J. Samuels, Y. Wang, M. A. Riddle, M. A. Grados, O. J. Bienvenu, Y. Y. Shugart, K. Y. Liang, B. D. Greenberg, S. A. Rasmussen, D. L. Murphy, J. R. Wendland, J. T. McCracken, J. Piacentini, S. L. Rauch, D. L. Pauls, G. Nestadt, A. J. Fyer and J. A. Knowles (2011). "Association of SLC6A4 variants with obsessive- compulsive disorder in a large multicenter US family study." Mol Psychiatry 16(1): 108-120.

Vulink, N. C., H. G. Westenberg, F. van Nieuwerburgh, D. Deforce, S. B. Fluitman, J. S. Meinardi and D. Denys (2012). "Catechol-O-methyltranferase gene expression is associated with response to citalopram in obsessive-compulsive disorder." Int J Psychiatry Clin Pract 16(4): 277- 283.

Walitza, S., D. S. Bove, M. Romanos, T. Renner, L. Held, M. Simons, C. Wewetzer, C. Fleischhaker, H. Remschmidt, A. Warnke and E. Grunblatt (2012). "Pilot study on HTR2A promoter polymorphism, -1438G/A (rs6311) and a nearby copy number variation showed association with onset and severity in early onset obsessive-compulsive disorder." J Neural Transm 119(4): 507-515.

Walitza, S., Z. Marinova, E. Grunblatt, S. E. Lazic, H. Remschmidt, T. D. Vloet and J. R. Wendland (2014). "Trio study and meta-analysis support the association of genetic variation at the serotonin transporter with early-onset obsessive-compulsive disorder." Neurosci Lett 580: 100-103. 328

329

Walitza, S., A. Scherag, T. J. Renner, A. Hinney, H. Remschmidt, B. Herpertz-Dahlmann, E. Schulz, H. Schafer, K. W. Lange, C. Wewetzer and M. Gerlach (2008). "Transmission disequilibrium studies in early onset of obsessive-compulsive disorder for polymorphisms in genes of the dopaminergic system." J Neural Transm 115(7): 1071-1078.

Walitza, S., J. R. Wendland, E. Gruenblatt, A. Warnke, T. A. Sontag, O. Tucha and K. W. Lange (2010). "Genetics of early-onset obsessive-compulsive disorder." Eur Child Adolesc Psychiatry 19(3): 227-235.

Walitza, S., C. Wewetzer, A. Warnke, M. Gerlach, F. Geller, G. Gerber, T. Gorg, B. Herpertz- Dahlmann, E. Schulz, H. Remschmidt, J. Hebebrand and A. Hinney (2002). "5-HT2A promoter polymorphism -1438G/A in children and adolescents with obsessive-compulsive disorders." Mol Psychiatry 7(10): 1054-1057.

Wang, J., S. N. Jani-Sait, E. A. Escalon, A. J. Carroll, P. J. de Jong, I. R. Kirsch and P. D. Aplan (2000). "The t(14;21)(q11.2;q22) chromosomal translocation associated with T-cell acute lymphoblastic leukemia activates the BHLHB1 gene." Proc Natl Acad Sci U S A 97(7): 3497- 3502.

Wang, Y., C. A. Mathews, Y. Li, Z. Lin and Z. Xiao (2011). "Brain-derived neurotrophic factor (BDNF) plasma levels in drug-naive OCD patients are lower than those in healthy people, but are not lower than those in drug-treated OCD patients." J Affect Disord 133(1-2): 305-310.

Ward, L. D. and M. Kellis (2012). "HaploReg: a resource for exploring chromatin states, conservation, and regulatory motif alterations within sets of genetically linked variants." Nucleic Acids Res 40(Database issue): D930-934.

Watson, J. D. and F. H. Crick (1953). "The structure of DNA." Cold Spring Harb Symp Quant Biol 18: 123-131.

Weissman, M. M., R. C. Bland, G. J. Canino, S. Greenwald, H. G. Hwu, C. K. Lee, S. C. Newman, M. A. Oakley-Browne, M. Rubio-Stipec, P. J. Wickramaratne and et al. (1994). "The cross national epidemiology of obsessive compulsive disorder. The Cross National Collaborative Group." J Clin Psychiatry 55 Suppl: 5-10.

329

330

Weissman, M. M., P. Wickramaratne, K. R. Merikangas, J. F. Leckman, B. A. Prusoff, K. A. Caruso, K. K. Kidd and G. D. Gammon (1984). "Onset of major depression in early adulthood. Increased familial loading and specificity." Arch Gen Psychiatry 41(12): 1136-1143.

Weisstaub, N. V., M. Zhou, A. Lira, E. Lambe, J. Gonzalez-Maeso, J. P. Hornung, E. Sibille, M. Underwood, S. Itohara, W. T. Dauer, M. S. Ansorge, E. Morelli, J. J. Mann, M. Toth, G. Aghajanian, S. C. Sealfon, R. Hen and J. A. Gingrich (2006). "Cortical 5-HT2A receptor signaling modulates anxiety-like behaviors in mice." Science 313(5786): 536-540.

Weizman, S., X. Gonda, P. Dome and G. Faludi (2012). "Pharmacogenetics of antidepressive drugs: a way towards personalized treatment of major depressive disorder." Neuropsychopharmacol Hung 14(2): 87-101.

Welch, J. M., J. Lu, R. M. Rodriguiz, N. C. Trotta, J. Peca, J. D. Ding, C. Feliciano, M. Chen, J. P. Adams, J. Luo, S. M. Dudek, R. J. Weinberg, N. Calakos, W. C. Wetsel and G. Feng (2007). "Cortico-striatal synaptic defects and OCD-like behaviours in Sapap3-mutant mice." Nature 448(7156): 894-900.

Wendland, J. R., M. R. Kruse, K. R. Cromer and D. L. Murphy (2007). "A large case-control study of common functional SLC6A4 and BDNF variants in obsessive-compulsive disorder." Neuropsychopharmacology 32(12): 2543-2551.

Wendland, J. R., P. R. Moya, K. R. Timpano, A. P. Anavitarte, M. R. Kruse, M. G. Wheaton, R. F. Ren-Patterson and D. L. Murphy (2009). "A haplotype containing quantitative trait loci for SLC1A1 gene expression and its association with obsessive-compulsive disorder." Arch Gen Psychiatry 66(4): 408-416.

Westfall, P. H. and S. S. Young (1993). Resampling based multiple testing. New York, Wiley.

Wetterneck, C. T., T. E. Little, G. S. Chasson, A. H. Smith, J. M. Hart, M. A. Stanley and T. Bjorgvinsson (2011). "Obsessive-compulsive personality traits: how are they related to OCD severity?" J Anxiety Disord 25(8): 1024-1031.

330

331

Wilde, A., B. Meiser, P. B. Mitchell and P. R. Schofield (2010). "Public interest in predictive genetic testing, including direct-to-consumer testing, for susceptibility to major depression: preliminary findings." Eur J Hum Genet 18(1): 47-51.

Williams, M. J., M. S. Almen, R. Fredriksson and H. B. Schioth (2012). "What model organisms and interactomics can reveal about the genetics of human obesity." Cell Mol Life Sci 69(22): 3819-3834.

Williams, M. T., J. Elstein, E. Buckner, J. Abelson and J. Himle (2012). "Symptom Dimensions in Two Samples of Africans Americans with Obsessive-Compulsive Disorder." J Obsessive Compuls Relat Disord 1(3): 145-152.

Williams, M. T., S. G. Farris, E. N. Turkheimer, M. E. Franklin, H. B. Simpson, M. Liebowitz and E. B. Foa (2014). "The impact of symptom dimensions on outcome for exposure and ritual prevention therapy in obsessive-compulsive disorder." J Anxiety Disord 28(6): 553-558.

Willour, V. L., Y. Yao Shugart, J. Samuels, M. Grados, B. Cullen, O. J. Bienvenu, 3rd, Y. Wang, K. Y. Liang, D. Valle, R. Hoehn-Saric, M. Riddle and G. Nestadt (2004). "Replication study supports evidence for linkage to 9p24 in obsessive-compulsive disorder." Am J Hum Genet 75(3): 508-513.

Wu, H., X. Wang, Z. Xiao, S. Yu, L. Zhu, D. Wang, K. Jiang, Z. Wang, T. Zhang and D. Fralick (2013). "Association between SLC1A1 gene and early-onset OCD in the Han Chinese population: a case-control study." J Mol Neurosci 50(2): 353-359.

Wu, K., G. L. Hanna, P. Easter, J. L. Kennedy, D. R. Rosenberg and P. D. Arnold (2013). "Glutamate system genes and brain volume alterations in pediatric obsessive-compulsive disorder: a preliminary study." Psychiatry Res 211(3): 214-220.

Wu, K., G. L. Hanna, D. R. Rosenberg and P. D. Arnold (2012). "The role of glutamate signaling in the pathogenesis and treatment of obsessive-compulsive disorder." Pharmacol Biochem Behav 100(4): 726-735.

331

332

Wu, K. D., D. Watson and L. A. Clark (2007). "A self-report version of the Yale-Brown Obsessive-Compulsive Scale Symptom Checklist: psychometric properties of factor-based scales in three samples." J Anxiety Disord 21(5): 644-661.

Wu, L., X. Zhao, Y. Shen, M. X. Zhang, Y. Yan, D. Hou, L. Meng, J. Liu, H. Cheng and J. Mi (2015). "Promoter methylation of fas apoptotic inhibitory molecule 2 gene is associated with obesity and dyslipidaemia in Chinese children." Diab Vasc Dis Res 12(3): 217-220.

Wu, M. X., Z. Ao, K. V. Prasad, R. Wu and S. F. Schlossman (1998). "IEX-1L, an apoptosis inhibitor involved in NF-kappaB-mediated cell survival." Science 281(5379): 998-1001.

Wu, P. L., H. S. Tang, H. Y. Lane, C. A. Tsai and G. E. Tsai (2011). "Sarcosine therapy for obsessive compulsive disorder: a prospective, open-label study." J Clin Psychopharmacol 31(3): 369-374.

Xu, Z. and J. A. Taylor (2009). "SNPinfo: integrating GWAS and candidate gene information into functional SNP selection for genetic association studies." Nucleic Acids Res 37(Web Server issue): W600-605.

Zai, G., P. Arnold, J. Strauss, N. King, E. Burroughs, M. A. Richter and J. L. Kennedy (2005). "No association between brain-derived neurotrophic factor gene and obsessive-compulsive disorder." Psychiatr Genet 15(4): 235.

Zai, G., Y. B. Bezchlibnyk, M. A. Richter, P. Arnold, E. Burroughs, C. L. Barr and J. L. Kennedy (2004). "Myelin oligodendrocyte glycoprotein (MOG) gene is associated with obsessive-compulsive disorder." Am J Med Genet B Neuropsychiatr Genet 129B(1): 64-68.

Zai, G., E. J. Brandl, D. J. Muller, M. A. Richter and J. L. Kennedy (2014). "Pharmacogenetics of antidepressant treatment in obsessive-compulsive disorder: an update and implications for clinicians." Pharmacogenomics 15(8): 1147-1157.

Zai, G., T. Sicard, E. Burroughs, J. L. Kennedy and M. A. Richter (2012). "An exploration of the oligodendrocyte lineage transcription factor 2 gene and obsessive-compulsive disorder." Psychiatr Genet 22(3): 149.

332

333

Zai, G., C. C. Zai, P. D. Arnold, N. Freeman, E. Burroughs, J. L. Kennedy and M. A. Richter (2015). "Meta-analysis and association of brain-derived neurotrophic factor (BDNF) gene with obsessive-compulsive disorder." Psychiatr Genet 25(2): 95-96.

Zhang, J., Y. Chen, K. Zhang, H. Yang, Y. Sun, Y. Fang, Y. Shen and Q. Xu (2010). "A cis- phase interaction study of genetic variants within the MAOA gene in major depressive disorder." Biol Psychiatry 68(9): 795-800.

Zhang, X., J. Liu, J. Cui and C. Liu (2013). "Study of symptom dimensions and clinical characteristics in Chinese patients with OCD." J Affect Disord 151(3): 868-874.

Zhang, X., J. Liu, Y. Guo, W. Jiang and J. Yu (2015). "Association Study between Oligodendrocyte Transcription Factor 2 Gene and Obsessive-Compulsive Disorder in a Chinese Han Population." Depress Anxiety.

Zhou, J., H. F. Bradford and G. M. Stern (1994). "The stimulatory effect of brain-derived neurotrophic factor on dopaminergic phenotype expression of embryonic rat cortical neurons in vitro." Brain Res Dev Brain Res 81(2): 318-324.

Zohar, J., M. Chopra, Y. Sasson, R. Amiaz and D. Amital (2000). "Obsessive compulsive disorder: serotonin and beyond." World J Biol Psychiatry 1(2): 92-100.

Zohar, J., J. L. Kennedy, E. Hollander and L. M. Koran (2004). "Serotonin-1D hypothesis of obsessive-compulsive disorder: an update." J Clin Psychiatry 65 Suppl 14: 18-21.

Zuchner, S., J. R. Wendland, A. E. Ashley-Koch, A. L. Collins, K. N. Tran-Viet, K. Quinn, K. C. Timpano, M. L. Cuccaro, M. A. Pericak-Vance, D. C. Steffens, K. R. Krishnan, G. Feng and D. L. Murphy (2009). "Multiple rare SAPAP3 missense variants in trichotillomania and OCD." Mol Psychiatry 14(1): 6-9.

333

334

Appendices

Appendix I: Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score and Symptom Checklist.

Appendix II: Family History Inventory for OCD.

FHI-OCD.

Table 1. Variable Description of Family History Inventory for OCD.

FHI Algorithm. Algorithm to Determine Diagnosis for Family History.

Appendix III: Pharmacogenetics Questionnaire.

Appendix IV: Supplementary Tables.

Table 1. Extensive Summary Table of Genetic Studies of OCD Phenotypes.

Table 2. Extensive Summary Table of Studies of OCD Symptom Dimensions.

Table 3. Extensive Summary Table of Genetic Studies of OCD Symptom Dimensions.

334

335

Appendix I

Yale-Brown Obsessive-Compulsive Scale (Y-BOCS) Severity Score and Symptom Checklist

335

336

Appendix II

Family History Inventory for OCD

336

337

Table 1. Variable Description of the Family History Inventory in OCD. Variable Description prob_id proband ID number targ_id target_id targ_rel target's relation to proband targ_r target's relation to proband targ_ini target initials targ_yb target year of birth targ_gen target gender inf_a_id informant "a" ID relative to the proband inf_a_in informant a initials inf_a_g informant a gender inf_a_yb informant "a" year of birth inf_a_yi informant "a" year of interview inf_a_st_rc informant "a" status relative to target (recoded) inf_a_re reliability of information from informant a inf_b_id informant "b" ID relative to the proband inf_b_in informant "b" initials inf_b_g informant "b" gender inf_b_yb informant "b" year of birth inf_b_yi informant "b" year of interview inf_b_st_rc informant "b" status relative to target inf_b_re reliability of information from informant b inf_c_id informant "c" ID relative to the proband inf_c_in informant "c" initials inf_c_g informant "c" gender inf_c_yb informant "c" year of birth inf_c_yi informant "c" year of interview inf_c_st_rc informant "c" status relative to target (recoded) inf_c_re reliability of information from informant c inf_d_id informant "d" ID relative to the proband inf_d_in informant "d" initials inf_d_g informant "d" gender inf_d_yb informant "d" year of birth inf_d_yi informant "d" year of interview inf_d_st_rc informant "d" status relative to target (recoded) inf_d_re reliability of information from informant d ocd1_a OCD question 1 (a - first informant) ocd2_a OCD question 2 (a - first informant) ocd3_a OCD question 3 (a - first informant) ocd4_a OCD question 4 (a - first informant) 337

338

tic4_a TIC question 4 (a - first informant) tic5_a TIC question 5 (a - first informant) tic6_a TIC question 6 (a - first informant) tic7_a TIC question 7 (a - first informant) trich8_a TRICH question 8 (a - first informant) trich9_a TRICH question 9 (a - first informant) skin10_a SKIN PICK question 10 (a - first informant) nail11_a NAIL BITE question 11 (a - first informant) bdd12_a BDD question 12 (a - first informant) bdd13_a BDD question 13 (a - first informant) eat14_a Eating disorder question 14 (a - first informant) eat15_a Eating disorder15_a(a_first informant) eat16_a Eating disorder16_a(a_first informant) eat17_a Eating disorder17_a (a_first informant) eat18_a Eating disorder18_a (a_first informant) eat19_a Eating disorder19_a (a_first informant) ocd1_b OCD question 1 (b - second informant) ocd2_b OCD question 2 (b - second informant) ocd3_b OCD question 3 (b - second informant) ocd4_b OCD question 4 (b- second informant) tic4_b TIC question 4 (b - second informant) tic5_b TIC question 5 (b - second informant) tic6_b TIC question 6 (b - second informant) tic7_b TIC question 7 (b - second informant) trich8_b TRICH question 8 (b -second informant) trich9_b TRICH question 9 (b -second informant) skin10_b SKIN PICK question 10 (b - second informant) nail11_b NAIL BITE question 11(b - second informant) bdd12_b bdd12_b(b-second informant) bdd13_b bdd13_b(b_second informant) eat14_b Eating disorder14_b (b_second informant) eat15_b Eating disorder15_b(b_second informant) eat16_b Eating disorder16_b(b-second informant) eat17_b Eating disorder17_b(b_second informant) eat18_b Eating disorder18_b(b_second informant) eat19_b Eating disorder19_b(b_second informant) ocd1_c OCD question 1 (c - third informant) ocd2_c OCD question 2 (c - third informant) ocd3_c OCD question 3 (c - third informant) ocd4_c OCD question 4 (c- third informant) tic4_c TIC question 4 (c - third informant)

338

339

tic5_c TIC question 5 (c - third informant) tic6_c TIC question 6 (c - third informant) tic7_c TIC question 7 (c - third informant) trich8_c TRICH question 8 (c - third informant) trich9_c TRICH question 9 (c - third informant) skin10_c SKIN PICK question 10 (c - third informant) nail11_c NAIL BITE question 11 (c - third informant) bdd12_c bdd12_c(c-third informant) bdd_13c bdd_13c (c-third informant) eat14_c Eating disorder14_c(c_third informant) eat15_c Eating disorder15_c(c_third informant) eat16_c Eating disorder16_c(c_third informant) eat17_c Eating disorder17_c(c_third informant) eat18_c Eating disorder18_c(c_third informant) eat19_c Eating disorder19_c(c_third informant) ocd1_d OCD question 1 (d - fourth informant) ocd2_d OCD question 2 (d - fourth informant) ocd3_d OCD question 3 (d - fourth informant) ocd4_d OCD question 4 (d - fourth informant) tic4_d TIC question 4 (d - fourth informant) tic5_d TIC question 5 (d - fourth informant) tic6_d TIC question 6 (d - fourth informant) tic7_d TIC question 7 (d - fourth informant) trich8_d TRICH question 8 (d - fourth informant) trich9_d TRICH question 9 (d - fourth informant) skin10_d SKIN PICK question 10 (d - fourth informant) nail11_d NAIL BITE question 11 (d - fourth informant) bdd12_d bdd12_d(d-fourth informant) bdd_13d bdd_13d (d_fourth informant) eat14_d Eating disorder14_d (d_fourth informant) eat15_d Eating disorder15_d (d_fourth informant) eat16_d Eating disorder16_d (d_fourth informant) eat17_d Eating disorder17_d (d_fourth informant) eat18_d Eating disorder18_d (d_fourth informant) eat19_d Eating disorder19_d (d_fourth informant)

339

340

Family History Inventory Algorithm

OCD Diagnosis in First-Degree Family Members:

0=unaffected 1=possible OCD 2=probable OCD 3=OCD only 4=hoarding only 5=OCD+hoarding

Reliability measure to see how reliable the informants were for these below (if reliability of informant = 2, 3, or 5): If q1 & q2 & q4 = yes, then definite OCD If (q1 or q2) & q4 = yes, then definite OCD If q1 = yes, then possible OCD If q2 = yes, then probable OCD If q4 = yes, then definite OCD If q3 = yes, then hoarding+probable OCD If q3 & q4 = yes, then hoarding+probable OCD If q1 & q2 & q3 & q4 = yes, then OCD+hoarding

No need reliability measure for these below: - Above algorithm can all be used here if >=2 informants said the same thing - If 2 informants gave q2 = yes, then definite OCD - If 2 informants gave q1 = yes, then probable OCD - If 2 informants gave q1 & q2 = yes, then definite OCD - If 2 informants gave q3 = yes and others = no, then hoarding without OCD - If 2 informants gave (q1 & q3 = yes) or (q2 & q3 = yes) or (q3 & q4 = yes), then OCD+hoarding

Father: If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and ((father_ocd1_a=0 and father_ocd2_a=0 and father_ocd3_a=0) or father_ocd4_a=0)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and ((father_ocd1_b=0 and father_ocd2_b=0 and father_ocd3_b=0) or father_ocd4_b=0)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and ((father_ocd1_c=0 and father_ocd2_c=0 and father_ocd3_c=0) or father_ocd4_c=0)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and ((father_ocd1_d=0 and father_ocd2_d=0 and father_ocd3_d=0) or father_ocd4_d=0)]], then Father_OCD=0

If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_ocd1_a=1 and father_ocd2_a=0 and father_ocd3_a=0)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_ocd1_b=1 and father_ocd2_b=0 and father_ocd3_b=0)] or

340

341

[(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_ocd1_c=1 and father_ocd2_c=0 and father_ocd3_c=0)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_ocd1_d=1 and father_ocd2_d=0 and father_ocd3_d=0)]], then Father_OCD=1

If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_ocd1_a=0 and father_ocd2_a=1 and father_ocd3_a=0)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_ocd1_b=0 and father_ocd2_b=1 and father_ocd3_b=0)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_ocd1_c=0 and father_ocd2_c=1 and father_ocd3_c=0)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_ocd1_d=0 and father_ocd2_d=1 and father_ocd3_d=0)]], then Father_OCD=2

If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and ((father_ocd1_a=1 and father_ocd2_a=1 and father_ocd3_a=0) or father_ocd4_a=1)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and ((father_ocd1_b=1 and father_ocd2_b=1 and father_ocd3_b=0) or father_ocd4_b=1)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and ((father_ocd1_c=1 and father_ocd2_c=1 and father_ocd3_c=0) or father_ocd4_c=1)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and ((father_ocd1_d=1 and father_ocd2_d=1 and father_ocd3_d=0) or father_ocd4_d=1)]], then Father_OCD=3

If [(father_ocd2_a=1 and father_ocd2_b=1 and father_ocd3_a=0 and father_ocd3_b=0) or (father_ocd2_a=1 and father_ocd2_c=1 and father_ocd3_a=0 and father_ocd3_c=0) or (father_ocd2_a=1 and father_ocd2_d=1 and father_ocd3_a=0 and father_ocd3_d=0) or (father_ocd2_b=1 and father_ocd2_c=1 and father_ocd3_b=0 and father_ocd3_c=0) or (father_ocd2_b=1 and father_ocd2_d=1 and father_ocd3_b=0 and father_ocd3_d=0) or (father_ocd2_c=1 and father_ocd2_d=1 and father_ocd3_c=0 and father_ocd3_d=0)], then Father_OCD=3

If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_ocd1_a=0 and father_ocd2_a=0 and father_ocd3_a=1)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_ocd1_b=0 and father_ocd2_b=0 and father_ocd3_b=1)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_ocd1_c=0 and father_ocd2_c=0 and father_ocd3_c=1)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_ocd1_d=0 and father_ocd2_d=0 and father_ocd3_d=1)]], then Father_OCD=4

If [(father_ocd3_a=1 and father_ocd3_b=1 and father_ocd1_a=0 and father_ocd2_a=0 and father_ocd1_b=0 and father_ocd2_b=0) or (father_ocd3_a=1 and father_ocd3_c=1 and father_ocd1_a=0 and father_ocd2_a=0 and father_ocd1_c=0 and father_ocd2_c=0) or (father_ocd3_a=1 and father_ocd3_d=1 and father_ocd1_a=0 and father_ocd2_a=0 and father_ocd1_d=0 and father_ocd2_d=0) or (father_ocd3_b=1 and father_ocd3_c=1 and father_ocd1_b=0 and father_ocd2_b=0 and father_ocd1_c=0 and father_ocd2_c=0) or

341

342

(father_ocd3_b=1 and father_ocd3_d=1 and father_ocd1_b=0 and father_ocd2_b=0 and father_ocd1_d=0 and father_ocd2_d=0) or (father_ocd3_c=1 and father_ocd3_d=1 and father_ocd1_c=0 and father_ocd2_c=0 and father_ocd1_d=0 and father_ocd2_d=0)], then Father_OCD=4

If [[(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and ((father_ocd1_a=1 and father_ocd2_a=1 and father_ocd3_a=1) or father_ocd4_a=1)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and ((father_ocd1_b=1 and father_ocd2_b=1 and father_ocd3_b=1) or father_ocd4_b=1)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and ((father_ocd1_c=1 and father_ocd2_c=1 and father_ocd3_c=1) or father_ocd4_c=1)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and ((father_ocd1_d=1 and father_ocd2_d=1 and father_ocd3_d=1) or father_ocd4_d=1)]], then Father_OCD=5

If [(father_ocd2_a=1 and father_ocd2_b=1 and father_ocd3_a=1and father_ocd3_b=1) or (father_ocd2_a=1 and father_ocd2_c=1 and father_ocd3_a=1 and father_ocd3_c=1) or (father_ocd2_a=1 and father_ocd2_d=1 and father_ocd3_a=1 and father_ocd3_d=1) or (father_ocd2_b=1 and father_ocd2_c=1 and father_ocd3_b=1 and father_ocd3_c=1) or (father_ocd2_b=1 and father_ocd2_d=1 and father_ocd3_b=1 and father_ocd3_d=1) or (father_ocd2_c=1 and father_ocd2_d=1 and father_ocd3_c=1 and father_ocd3_d=1)], then Father_OCD=5

If [(father_ocd4_a=1 and father_ocd4_b=1 and father_ocd3_a=1and father_ocd3_b=1) or (father_ocd4_a=1 and father_ocd4_c=1 and father_ocd3_a=1 and father_ocd3_c=1) or (father_ocd4_a=1 and father_ocd4_d=1 and father_ocd3_a=1 and father_ocd3_d=1) or (father_ocd4_b=1 and father_ocd4_c=1 and father_ocd3_b=1 and father_ocd3_c=1) or (father_ocd4_b=1 and father_ocd4_d=1 and father_ocd3_b=1 and father_ocd3_d=1) or (father_ocd4_c=1 and father_ocd4_d=1 and father_ocd3_c=1 and father_ocd3_d=1)], then Father_OCD=5

Replace father for mother, relative 1 (rel1), relative 2 (rel2), and up to relative 8 (rel8) - mother and relative 6 have informant a, b, and c only - relative 7 and 8 have informant a only - please adjust algorithm accordingly and eliminate informant d for mother and relative 6, and eliminate informant b, c, d for relative 7 and 8

Mother: If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and ((mother_ocd1_a=0 and mother_ocd2_a=0 and mother_ocd3_a=0) or mother_ocd4_a=0)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and ((mother_ocd1_b=0 and mother_ocd2_b=0 and mother_ocd3_b=0) or mother_ocd4_b=0)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and ((mother_ocd1_c=0 and mother_ocd2_c=0 and mother_ocd3_c=0) or mother_ocd4_c=0)], then Mother_OCD=0

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_ocd1_a=1 and mother_ocd2_a=0 and mother_ocd3_a=0)] or

342

343

[(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_ocd1_b=1 and mother_ocd2_b=0 and mother_ocd3_b=0)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_ocd1_c=1 and mother_ocd2_c=0 and mother_ocd3_c=0)], then Mother_OCD=1

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_ocd1_a=0 and mother_ocd2_a=1 and mother_ocd3_a=0)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_ocd1_b=0 and mother_ocd2_b=1 and mother_ocd3_b=0)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_ocd1_c=0 and mother_ocd2_c=1 and mother_ocd3_c=0)], then Mother_OCD=2

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and ((mother_ocd1_a=1 and mother_ocd2_a=1 and mother_ocd3_a=0) or mother_ocd4_a=1))] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and ((mother_ocd1_b=1 and mother_ocd2_b=1 and mother_ocd3_b=0) or mother_ocd4_b=1))] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and ((mother_ocd1_c=1 and mother_ocd2_c=1 and mother_ocd3_c=0) or mother_ocd4_c=1))], then Mother_OCD=3

If [(mother_ocd2_a=1 and mother_ocd2_b=1 and mother_ocd3_a=0 and mother_ocd3_b=0) or (mother_ocd2_a=1 and mother_ocd2_c=1 and mother_ocd3_a=0 and mother_ocd3_c=0) or (mother_ocd2_b=1 and mother_ocd2_c=1 and mother_ocd3_b=0 and mother_ocd3_c=0), then Mother_OCD=3

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_ocd1_a=0 and mother_ocd2_a=0 and mother_ocd3_a=1)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_ocd1_b=0 and mother_ocd2_b=0 and mother_ocd3_b=1)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_ocd1_c=0 and mother_ocd2_c=0 and mother_ocd3_c=1)], then Mother_OCD=4

If [(mother_ocd3_a=1 and mother_ocd3_b=1 and mother_ocd1_a=0 and mother_ocd2_a=0 and mother_ocd1_b=0 and mother_ocd2_b=0) or (mother_ocd3_a=1 and mother_ocd3_c=1 and mother_ocd1_a=0 and mother_ocd2_a=0 and mother_ocd1_c=0 and mother_ocd2_c=0) or (mother_ocd3_b=1 and mother_ocd3_c=1 and mother_ocd1_b=0 and mother_ocd2_b=0 and mother_ocd1_c=0 and mother_ocd2_c=0)], then Mother_OCD=4

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and ((mother_ocd1_a=1 and mother_ocd2_a=1 and mother_ocd3_a=1) or mother_ocd4_a=1)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and ((mother_ocd1_b=1 and mother_ocd2_b=1 and mother_ocd3_b=1) or mother_ocd4_b=1)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and ((mother_ocd1_c=1 and mother_ocd2_c=1 and mother_ocd3_c=1) or mother_ocd4_c=1)], then Mother_OCD=5

If [(mother_ocd2_a=1 and mother_ocd2_b=1 and mother_ocd3_a=1and mother_ocd3_b=1) or (mother_ocd2_a=1 and mother_ocd2_c=1 and mother_ocd3_a=1 and mother_ocd3_c=1) or

343

344

(mother_ocd2_b=1 and mother_ocd2_c=1 and mother_ocd3_b=1 and mother_ocd3_c=1), then Mother_OCD=5

If [(mother_ocd4_a=1 and mother_ocd4_b=1 and mother_ocd3_a=1and mother_ocd3_b=1) or (mother_ocd4_a=1 and mother_ocd4_c=1 and mother_ocd3_a=1 and mother_ocd3_c=1) or (mother_ocd4_b=1 and mother_ocd4_c=1 and mother_ocd3_b=1 and mother_ocd3_c=1), then Mother_OCD=5

Replace mother for relative 6 (rel6) - mother and relative 6 have informant a, b, and c only - relative 7 and 8 have informant a only - please adjust algorithm accordingly and eliminate informant d for mother and relative 6, and eliminate informant b, c, d for relative 7 and 8

Relative 7 (rel7): If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and ((rel7_ocd1_a=0 and rel7_ocd2_a=0 and rel7_ocd3_a=0) or rel7_ocd4_a=0)], then Relative7_OCD=0

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_ocd1_a=1 and rel7_ocd2_a=0 and rel7_ocd3_a=0)], then Relative7_OCD=1

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_ocd1_a=0 and rel7_ocd2_a=1 and rel7_ocd3_a=0)], then Relative7_OCD=2

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and ((rel7_ocd1_a=1 and rel7_ocd2_a=1 and rel7_ocd3_a=0) or rel7_ocd4_a=1))], then Relative7_OCD=3

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_ocd1_a=0 and rel7_ocd2_a=0 and rel7_ocd3_a=1)], then Relative7_OCD=4

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and ((rel7_ocd1_a=1 and rel7_ocd2_a=1 and rel7_ocd3_a=1) or rel7_ocd4_a=1)], then Relative7_OCD=5

Replace rel7 with rel8

Trichotillomania (0=unaffected and 1=trichotillomania):

Father: If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_trich8_a=0 and father_trich9_a=0)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_trich8_b=0 and father_trich9_b=0)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_trich8_c=0 and father_trich9_c=0)] or

344

345

[(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_trich8_d=0 and father_trich9_d=0)], then Father_Trichotillomania=0

If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_trich8_a=1 or father_trich9_a=1)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_trich8_b=1 or father_trich9_b=1)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_trich8_c=1 or father_trich9_c=1)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_trich8_d=1 or father_trich9_d=1)], then Father_Trichotillomania=1

If [(father_trich8_a=1 and father_trich8_b=1) or (father_trich9_a=1 and father_trich9_b=1)] or [(father_trich8_a=1 and father_trich8_c=1) or (father_trich9_a=1 and father_trich9_c=1)] or [(father_trich8_a=1 and father_trich8_d=1) or (father_trich9_a=1 and father_trich9_d=1)] or [(father_trich8_b=1 and father_trich8_c=1) or (father_trich9_b=1 and father_trich9_c=1)] or [(father_trich8_b=1 and father_trich8_d=1) or (father_trich9_b=1 and father_trich9_d=1)] or [(father_trich8_c=1 and father_trich8_d=1) or (father_trich9_c=1 and father_trich9_d=1)], then Father_Trichotillomania=1

Mother: If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_trich8_a=0 and mother_trich9_a=0)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_trich8_b=0 and mother_trich9_b=0)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_trich8_c=0 and mother_trich9_c=0)], then Mother_Trichotillomania=0

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_trich8_a=1 or mother_trich9_a=1)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_trich8_b=1 or mother_trich9_b=1)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_trich8_c=1 or mother_trich9_c=1)], then Mother_Trichotillomania=1

If [(mother_trich8_a=1 and mother_trich8_b=1) or (mother_trich9_a=1 and mother_trich9_b=1)] or [(mother_trich8_a=1 and mother_trich8_c=1) or (mother_trich9_a=1 and mother_trich9_c=1)] or [(mother_trich8_b=1 and mother_trich8_c=1) or (mother_trich9_b=1 and mother_trich9_c=1)], then Mother_Trichotillomania=1

Rel7: If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_trich8_a=0 and rel7_trich9_a=0)], then Relative7_Trichotillomania=0

345

346

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_trich8_a=1 or rel7_trich9_a=1)], then Relative7_Trichotillomania=1

Skin Picking (0=unaffected and 1=skin picking):

Father: If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and father_skin10_a=0] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and father_skin10_b=0] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and father_skin10_c=0] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and father_skin10_d=0], then Father_Skin=0

If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and father_skin10_a=1] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and father_skin10_b=1] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and father_skin10_c=1] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and father_skin10_d=1], then Father_Skin=1

If (father_skin10_a=1 and father_skin10_b=1) or (father_skin10_a=1 and father_skin10_c=1) or (father_skin10_a=1 and father_skin10_d=1) or (father_skin10_b=1 and father_skin10_c=1) or (father_skin10_b=1 and father_skin10_d=1) or (father_skin10_c=1 and father_skin10_d=1), then Father_Skin=1

Mother: If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and mother_skin10_a=0] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and mother_skin10_b=0] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and mother_skin10_c=0], then Mother_Skin=0

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and mother_skin10_a=1] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and mother_skin10_b=1] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and mother_skin10_c=1], then Mother_Skin=1

If (mother_skin10_a=1 and mother_skin10_b=1) or (mother_skin10_a=1 and mother_skin10_c=1) or (mother_skin10_b=1 and mother_skin10_c=1), then Mother_Skin=1

Rel7: If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and rel7_skin10_a=0], then Relative7_Skin=0

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and rel7_skin10_a=1], then Relative7_Skin=1

346

347

Body Dysmorphic Disorder (0=unaffected and 1=BDD):

Father: If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_bdd10_a=0 and father_bdd13_a=0)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_bdd10_b=0 and father_bdd13_b=0)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_bdd10_c=0 and father_bdd13_c=0)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_bdd10_d=0 and father_bdd13_d=0)], then Father_BDD=0

If [(father_inf_a_re=2 or father_inf_a_re=3 or father_inf_a_re=5) and (father_bdd10_a=1 or father_bdd13_a=1)] or [(father_inf_b_re=2 or father_inf_b_re=3 or father_inf_b_re=5) and (father_bdd10_b=1 or father_bdd13_b=1)] or [(father_inf_c_re=2 or father_inf_c_re=3 or father_inf_c_re=5) and (father_bdd10_c=1 or father_bdd13_c=1)] or [(father_inf_d_re=2 or father_inf_d_re=3 or father_inf_d_re=5) and (father_bdd10_d=1 or father_bdd13_d=1)], then Father_BDD=1

If [(father_bdd10_a=1 and father_bdd10_b=1) or (father_bdd13_a=1 and father_bdd13_b=1)] or [(father_bdd10_a=1 and father_bdd10_c=1) or (father_bdd13_a=1 and father_bdd13_c=1)] or [(father_bdd10_a=1 and father_bdd10_d=1) or (father_bdd13_a=1 and father_bdd13_d=1)] or [(father_bdd10_b=1 and father_bdd10_c=1) or (father_bdd13_b=1 and father_bdd13_c=1)] or [(father_bdd10_b=1 and father_bdd10_d=1) or (father_bdd13_b=1 and father_bdd13_d=1)] or [(father_bdd10_c=1 and father_bdd10_d=1) or (father_bdd13_c=1 and father_bdd13_d=1)], then Father_BDD=1

Mother: If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_bdd10_a=0 and mother_bdd13_a=0)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_bdd10_b=0 and mother_bdd13_b=0)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_bdd10_c=0 and mother_bdd13_c=0)], then Mother_BDD=0

If [(mother_inf_a_re=2 or mother_inf_a_re=3 or mother_inf_a_re=5) and (mother_bdd10_a=1 or mother_bdd13_a=1)] or [(mother_inf_b_re=2 or mother_inf_b_re=3 or mother_inf_b_re=5) and (mother_bdd10_b=1 or mother_bdd13_b=1)] or [(mother_inf_c_re=2 or mother_inf_c_re=3 or mother_inf_c_re=5) and (mother_bdd10_c=1 or mother_bdd13_c=1)], then Mother_BDD=1

347

348

If [(mother_bdd10_a=1 and mother_bdd10_b=1) or (mother_bdd13_a=1 and mother_bdd13_b=1)] or [(mother_bdd10_a=1 and mother_bdd10_c=1) or (mother_bdd13_a=1 and mother_bdd13_c=1)] or [(mother_bdd10_b=1 and mother_bdd10_c=1) or (mother_bdd13_b=1 and mother_bdd13_c=1)], then Mother_BDD=1

Rel7: If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_bdd10_a=0 and rel7_bdd13_a=0)], then Relative7_BDD=0

If [(rel7_inf_a_re=2 or rel7_inf_a_re=3 or rel7_inf_a_re=5) and (rel7_bdd10_a=1 or rel7_bdd13_a=1)], then Relative7_BDD=1

348

349

Appendix III

Pharmacogenetics Questionnaire

349

350

Appendix IV

Table 1. Extensive Summary Table of Genetic Studies of OCD Phenotypes.

Study Year OCD Phenotype Methodology Sample Size Gene Polymorphism Results Met allele and Met/Met genotype Karayiorgou et 73 OCD and 148 associated with risk for OCD in 1997 Gender Case-control study COMT rs4680 (Val158Met) al. controls males (Met/Met: OR=5.18, P=0.0002) DAT1 40-bp VNTR, DRD2 DRD4 2/4 as protective genotype >100 OCD matched TaqI (rs1800497), DRD3 Billett et al. 1998 Gender Case-control study to OCD (P=0.021) and not to controls MscI (Ser9Gly; rs6280), significant with gender and DRD4 48-bp VNTR Gender, with/without 109 OCD and 107 Cavallini et al. 1998 Case-control study HTR2C Cys23Ser (rs6318) Not significant tics controls Homozygotes were preferentially Schindler et al. 2000 Gender TDT 72 OCD trios COMT rs4680 (Val158Met) transmitted to OCD (P=0.017) Gender, age of onset Higher frequency of allele 2 (C; (≤16 years as high activity) in OCD males Camarena et 122 OCD and 124 MAOA EcoRV (C1460T; 2001 childhood onset), Case-control study (P=0.002-0.0053) and allele 1 (T; al. controls rs1137070) with/without comorbid low activity) associated with OCD MDD females (P=0.0007) 101 Caucasian OCD A allele more frequent in OCD Enoch et al. 2001 Gender Case-control study HTR2A G-1438A (rs6311) and 138 controls females than controls (P=0.015) 56 OCD individuals Veenstra- Sequence the LOD (27 with lifetime Early-onset (ranging SLC1A1 (no evidence for VanderWeele 2001 score peak region definite OCD) from 7 Not significant from 3-14 years) functional mutation) et al. of chromosome 9q families ascertained through OCD children 48 OCD (Afrikaner) COMT rs4680 and ERE6 C- COMT rs4680 associated with Kinnear et al. 2001 Gender Case-control study and 48 controls >T total OCD sample (P=0.03) Alsobrook et Met allele associated with OCD 2002 Gender TDT 56 OCD trios COMT rs4680 al. females (P=0.048)

350

351

55 OCD children and A allele associated with early- HTR2A G-1438G/A Walitza et al. 2002 Early-onset, gender Association study adolescents, and 223 onset total OCD sample (P=0.048) (rs6311) controls and OCD females (P=0.03) 56 OCD individuals (27 with lifetime Early-onset (ranging Genome-wide Chromosome 9q where Evidence for linkage on Hanna et al. 2002 definite OCD) from 7 from 3-14 years) linkage analysis SLC1A1 is located chromosome 9q (LOD score=2) families ascertained through OCD children Two-step analysis Long/long genotype associated OC symptom of the 5HTT 180 OCD and 112 Cavallini et al. 2002 SLC6A4 5HTTLPR with repeating/counting dimensions insertion/deletion controls (P=0.0013) polymorphism rs988748 C allele (P=0.0005- BDNF rs988748, 0.003), rs2049046 T allele rs2049046, rs6265 (P=0.012-0.02), rs6265 Val allele Early- (<18 years) vs. (Val66Met), (CA)n, Hall et al. 2003 TDT 164 OCD triads (P=0.0005-0.001), and (CA)n C late-onset rs2763965, rs2352802, allele (P= 0.0005-0.0025) rs972096, rs1387145, overtransmitted in early-onset rs1464896, and rs2140887 subset and total OCD samples 252 OCD total DRD4 VNTR, DAT VNTR, sample (180 Association study COMT NlaIII (rs4680), DRD4 VNTR A7 allele associated Hemmings et Early- (≤15 years) vs. Caucasian OCD 2004 comparing early- HTR2A MspI (T102C; with late-onset Caucasian OCD al. late-onset subset and 80 vs. late-onset OCD rs6313), and HTR1B HincII (P=0.0218) Afrikaner OCD subset (G861C; rs6296) for genetic analysis)

351

352

OCD males associated with earlier age of onset (P=0.035) and more aggressive OCD symptoms DRD4 VNTR, DAT VNTR, (P=0.011); HTR1B rs6296 C/C 220 OCD total SLC6A4 5HTTLPR, HTR1B genotype associated with Gender, early- vs. sample (178 G861C (rs6296), HTR2A Afrikaner OCD males than late-onset, OCD Genetic association Caucasian OCD MspI (T102C; rs6313), TH Lochner et al. 2004 controls (P0.026); MAOA symptoms, and Axis I study subset and 81 Val81Met (rs6356), COMT rs1137070 C/C genotype comorbidity Afrikaner OCD subset Val158Met (rs4680), and associated with OCD than controls for genetic analysis) MAOA C1460T (EcoRV; (P=0.047) and more frequent in rs1137070) OCD females (P=0.009) but T allele more frequent in OCD males (P=0.008) Early-onset (mean 64 OCD trios TPH1 rs1800532, SLC6A4 age at Walitza et al. 2004 TDT (children and 5HTTLPR, and HTR1B Not significant onset=11.57±3.0 adolescents) rs6296 years) Childhood-onset (mean age at 67 OCD children and Mossner et al. 2005 TDT BDNF Val66Met (rs6265) Not significant onset=11.52±3.0 adolescents years) 24 OCD females, 112 S allele (s/l or s/s) frequency was non-impulsive Baca-Garćia et Genetic association lowest in OCD and highest in 2005 Female controls, 118 SLC6A4 5HTTLPR al. study impulsive suicide attempters impulsive suicidal (P=0.003) patients Met allele (P=0.013) and Met/Met 113 genotype (P=0.0098) were more Poyurovsky et schizophrenia+OCD, frequent in OCD males than 2005 Gender Case-control study COMT rs4680 Val158Met) al. 79 OCD, and 171 schizophrenia+OCD and controls controls but did not survive correction for multiple comparison Early-onset (mean rs4570625-rs4565946 G-C 71 OCD trios age at TPH2 rs4570625 and haplotype was overtransmitted Mossner et al. 2006 TDT (children and onset=11.73±2.9 rs4565946 (P=0.035) and G-T haplotype was adolescents) years) undertransmitted (P=0.022) in

352

353

early-onset OCD

5HTTLPR short allele associated Age of onset, positive with female OCD patients SLC6A4 5HTTLPR, HTR1B family history of OCD, (P=0.014) and HTR2A G allele Candidate gene 156 OCD and 134 G861C (rs6296), and Denys et al. 2006a clinical subtypes, and G/G genotype associated with study controls HTR2A G1438A (MspI; comorbidity, and positive family history of OCD rs6311) symptom severity (P=0.039-0.043) and earlier onset of OCD (P=0.038) DRD2 rs1800497 A2 allele COMT rs4680 (Val158Met) 159 OCD and 159 (P=0.020) and COMT rs4680 Met Denys et al. 2006b Gender Case-control study and DRD2 TaqIA controls allele (P=0.035) associated with (rs1800497) OCD males rs301434 (P=0.0007-0.002), SLC1A1 rs1980943, rs301435 (P=0.001-0.009), rs3780415, rs7856209, rs3087879 (P=0.002-0.006) and Candidate gene 157 OCD and Arnold et al. 2006 Gender rs3780412, rs301430, the rs301434-rs301435 haplotype study relatives (total 476) rs301979, rs301434, (P=0.0006-0.001) associated with rs301435, and rs3087879 transmission of male and total OCD sample SLC1A1 rs10814987, rs10739062, rs1980943, rs3780412 (P=0.04) and rs301430 Early-onset (≤18 rs10115600, rs10758629, (P=0.03) associated with early- Dickel et al. 2006 TDT, FBAT 71 OCD trios years), gender rs3780412, rs301979, onset; rs3780412 associated with rs301430, rs301434, and early-onset males (P=0.002) rs301443

353

354

ACE Alu ins/del associated with total OCD sample (P=0.037); HTR2C rs6318 (P<0.001), COMT rs362204 (P=0.040), BDNF rs6265 (P=0.036), and ACE Alu ins/del (P=0.049) associated with OCD males; GRIN2B rs890 (P=0.013) and BDNF rs6265 (P=0.036) associated with symptom severity in total OCD HTR2A rs6311 and 6313, sample; DRD1 rs4532 associated HTR1B rs6296, HTR6 with symptom severity in OCD rs1805054, HTR2C rs6318, males (P=0.045); BDNF rs6265 DAT VNTR, DRD1 A-48G associated with symptom severity (rs4532), DRD2 rs1800497, in OCD females (P=0.013-0.045); DRD3 rs6280, DRD4 HTR2A rs6313 (P=0.015), COMT Gender, symptom rs1800955 and 48-bp rs362204 (P=0.040), DAT VNTR severity, age at VNTR, COMT rs2097603, (P<0.001), DRD3 rs6280 Hemmings et onset, co-morbidity Case-control study 132 OCD (Afrikaner) 2006 rs4680, and rs362204, (P=0.022), GRIN2B rs890 al. (i.e., mental illness, and meta-analysis and 218 controls GRIN2B rs1806191 and (P=0.012), BDNF rs6265 tics), primary rs890, BDNF rs6265, (P=0.028), and PLC-γ1 rs8192797 symptom dimension rs2049046, and rs988748, (P=0.030) associated with age at HOXB8 rs2303486, ESRα onset in OCD males and GRIN2B rs9340799 and rs2234693, rs1806191 (P=0.049), HOXB8 INPP-1 rs1882891, PLC-γ1 rs2303486 (P=0.024), and PLC-γ1 rs8192797, and ACE Alu rs8192797 (P=0.039) in OCD ins/del females; DAT VNTR (P<0.001) and PLC-γ1 rs8192797 (P=0.005) associated with age at onset in total OCD sample; COMT rs2097603 (P=0.012-0.018) and HTR6 rs1805054 (P=0.008) associated with OCD+MDD; COMT rs4680 (P=0.004-0.026) and BDNF rs6265 (P=0.035- 0.059) associated with OCD+/- tics; DRD4 rs1800955 (P=0.043-

354

355

0.124), COMT rs4680 (P=0.006- 0.008), and ESRα rs9340799 (P=0.002-0.004) and rs2234693 (P=0.041-0.054) associated with OCD+/-hoarding

355

356

rs2228622 A allele (permuted P=0.045) and rs3780412 C allele (permuted P=0.045) were SLC1A1 rs10974625, overtransmitted to OCD males but rs2228622, rs3780413, Family-based not females; overtransmission of rs3780412, rs12682807, association 66 OCD families with rs12682897-rs2072657-rs301430 Stewart et al. 2007 Gender rs2073657, rs301430, candidate gene 57 trios A-T-T haplotype (permuted rs301979, rs301434, study P=0.0043) and undertransmission rs301435, rs3087879, of A-T-C haplotype (permuted rs301437, and rs301440 P=0.0015) to total OCD sample but main effect to OCD males (global permuted P=0.0031)

Symptom severity, childhood (<18 years) vs adult onset, positive familial history, symptom BDNF Val66Met (rs6265), BDNF Val66Met (rs6265), dimension factors, SLC6A4 5HTTLPR, SLC6A4 5HTTLPR, rs25531, Wendland et Candidate gene 347 OCD and 749 2007 total number of rs25531, Intron 2 VNTR STin2, and I425V/I425L = not al. study controls comorbid disorders, Stin2, and I425V/I425L significant with factors, early/late total number of (rs28914832) AAO, or familiality comorbid anxiety disorders, suicidality, and experience of trauma HTR2A rs6311 associated with OCD and tic disorder (P=0.05) at HTR2A rs6311, HTR1B a trend level; SLC6A4 5HTTLPR rs6296, SLC6A4 Ile425Val Gender, early-onset Candidate gene 54 parent-child OCD associated with females (P=0.03); Dickel et al. 2007 (rs28914832), SLC6A4 (≤18 years), tics study trios pooled analysis of 5 replication 5HTTLPR, and BDNF studies of SLC6A4 5HTTLPR rs6265 supports association with OCD (P=0.02)

356

357

Significant difference between OCD and controls (P=0.0003), OCD males and male controls Camarena et 210 OCD and 202 2007 Gender, tics Case-control study DRD4 48-bp VNTR (P=0.0051), OCD females and al. controls (Mexican) female controls (P=0.0275), and OCD with tics and controls (P=0.003) 2-repeat (P<0.001) and 3-repeat 208 OCD and 865 (P=0.002) associated with OCD Miguita et al. 2007 Gender, age at onset Case-control study DAT/SLC6A3 VNTR controls males but no overall significant results Independent sample: Met allele associated with OCD 87 OCD and 327 (Independent sample: P=0.026; Case-control study controls; meta- Pooley et al. 2007 Gender COMT rs4680 meta-analysis: P=0.005) and and meta-analysis analysis sample: 606 associated with OCD males OCD and 1302 (P<0.0001) controls COMT Val158Met (rs4680) Met/Met genotype in Caucasian DRD4 VNTR, DAT VNTR, subset and earlier onset SLC6A4 5HTTLPR, HTR1B associated with higher G861C (rs6269), HTR2A obsessional/checking cluster Age of onset, OCD Genetic association T102C (rs6313), TH Lochner et al. 2008 261 OCD scores (P<0.001); no significant symptoms study Val81Met (rs6356), COMT results for DRD4 VNTR, DAT Val158Met (rs4680), and VNTR, 5HTTLPR, HTR1B G861C MAOA C1460T (ECORV; (rs6269), HTR2A T102C (rs6313), rs1137070) TH Val81Met (rs6356), and MAOA C1460T (ECORV) Early-onset (mean DRD4 VNTR, COMT DRD4 VNTR 4-repeat alleles was age of 69 trios (chlidren and Walitza et al. 2008 TDT Val158Met (rs4680), and under-transmitted to early-onset onset=11.7±2.9 adolescents) DAT1 VNTR OCD (P=0.03; global P=0.047) years)

357

358

64 (Walitza et al., Significant association between 2004) and 54 (Dickel the 1-allele and OCD in family- et al., 2007) OCD Bloch et al. 2008 Early-onset, ethnicity Meta-analysis SLC6A4 5HTTLPR based association sutdies and in trios and 15 studies studies involving children (P<0.05) involving Caucasian and Caucasians (P=0.03) OCD

BDNF rs11030096, rs925946, rs10501087, rs6265, rs12273363, rs908867, rs1491850, and rs1491851; NTRK2 rs1201364, rs1201363, rs1187332, rs999244, rs1147193, rs1439050, rs1187352, rs1187356, BDNF rs6265-rs12273363- rs1187362, rs1209068, rs908867-rs1491850-rs1491851 rs3780632, rs4877880, G-C-G-C-T as protective rs1662695, rs1187286, haplotype to OCD (permutated rs716893, rs1187274, P=0.006), NTRK2 rs10512159- Candidate gene 215 OCD and 342 rs7816, rs1867283, rs1948308-rs2378672-rs1387926- Alonso et al. 2008 Gender study controls rs7855888, rs11140783, rs1387924-rs3739570-rs1490403 rs2120266, rs10116287, G-C-G-G-A-C-A as protective rs10868232, rs4877289, haplotype to OCD (permutated rs1443445, rs920776, P=0.0006), and NTRK2 rs995861, rs1545286, rs2378672 associated with OCD rs7045900, rs2277192, females (P<0.0001) rs2277193, rs4412435, rs10868241, rs4361832, rs11795386, rs12000011, rs10512159, rs1948308, rs2378672, rs1387926, rs1387924, rs3739570, rs1490403, rs10780695, rs1073049, and rs1586681

358

359

Met allele associated with OCD 112 OCD (South risk to males (P=0.047) and Hemmings et 2008 Gender, age at onset Case-control study African Afrikaner) and BDNF Val66Met (rs6265) Val/Val genotype associated with al. 140 controls more severe OCD symptoms in females (P=0.031) 419 subclinical OCD Met/Met genotype associated with Katerberg et al. 2009 Age of onset, gender Case-control study BDNF Val66Met (rs6265) and 650 controls later onset in women (P=0.002) rs301443 associated with total SLC1A1 rs10814987, OCD sample (P=0.000067) and rs10739062, rs1980943, OCD males (P=0.00027); rs10115600, rs3780415, Genetic association 378 OCD families rs10739062-rs1980943 G-G Shugart et al. 2009 Gender rs2228622, rs3780412, study with 1950 subjects haplotype (P=0.046) and rs301430, rs301979, rs3087879-rs3011443 G-C rs301434, rs301435, haplotype (P=0.024) associated rs3087879, and rs301443 with total OCD sample 167 OCD males G allele (P=0.004) and G/G Gender - male, Kim et al. 2009 Case-control study (Korean) and 107 HTR1B rs6296 (G861C) genotype (P=0.007) associated symptom dimension controls with OCD males Val allele associated with OCD Gender, age of onset, males (P=0.039) and interaction of symptom severity, 373 OCD and 462 genotype and female with lower Katerberg et al. 2010 Case-control study COMT rs4680 (Val158Met) symptom dimension, controls scores on somatic and sensory family history phenomena symptom dimension (P=0.018) SLC6A4 rs1487971, rs2020930, rs4392119, 5HTTLPR, 5HTTLPR- rs25531, rs25533, Family-based rs2020933, rs2066713, Intron 2 VNTR 9/10 alleles Voyiaziakis et association 278 OCD families 2011 Gender rs6355 (G56A), intron 2 associated with OCD females al. candidate gene with 1241 individuals VNTR STin2, rs140699, (global corrected P=0.0069) study (FBAT) rs140700, intron 7 (GAAA)n, rs4583306, I425V (rs28914832), rs6353 (T439T), and rs7224199

359

360

A-T-A-T (rs11583978-rs7541937- 172 OCD (South SAPAP3 rs11583978, rs6662980-rs4652867) haplotype Boardman et Age at onset (as African white), 45 rs7541937, rs6662980, 2011 Case-control study associated with earlier age at al. continuous variable) TTM, and 153 rs4652867, rs11264172, onset when compared with C-G- controls rs7555884, and rs6682829 G-G haplotype (P=0.036) G/G genotype more frequent in Gender, age of onset OCD (P=0.00091) than controls; (<18 as early onset), 200 OCD and 403 G/G genotype more frequent in Liu et al. 2011a with/without tic, Y- Case-control study controls (Chinese COMT A-287G (rs2075507) OCD males (P<0.001) and OCD BOCS symptom Han) females (P=0.026) than controls dimensions but no significant association for onset, tic, or symptom dimensions 207 OCD, 108 Gender, early- (<18 Tourette syndrome Liu et al. 2011b years) vs late-onset, Case-control study SLC6A4 5HTTLPR Not significant trios, and 275 with/without tics controls rs6311 A allele with early-onset (P=0.005) and CNV one copy Symptom severity Candidate gene 136 OCD children HTR2A G-1438A (rs6311) (deletion) with very early onset Walitza et al. 2012 and early-onset association study and adolescents and CNV (2.5 years earlier; P=0.031 and increased severity (CY-BOCS 8.7 higher; P=0.004) 33 Caucasian OCD Childhood-onset (<18 Genomewide Genomewide linkage Chromosome 1p36, 2p14, 5q13, Mathews et al. 2012 families with 245 years) linkage analysis analyses 6p25, and 10p13 (HLOD score≥2) available individuals 49 OCD children and TNFα rs1800629 (G-308A) rs1800629 associated with OCD Lüleyap et al. 2012 Children Case-control study 58 controls and rs1799124 (C-850T) children (P<0.001) 7R allele greater in OCD than controls (P=0.03) and more in 173 OCD and 201 Gender, symptom OCD females (P=0.02) and a Taj et al. 2013 Case-control study controls (South DRD4 VNTR dimensions trend with 2R allele in Indian) symmetry/ordering factor (P=0.01 uncorrected)

360

361

Age of onset (18 244 early-onset OCD, SLC1A1 rs10491734, rs10491734 A allele (P=0.023) Wu et al. 2013 years as cut off for Case-control study 244 late-onset OCD, rs2228622, rs301430, and and A/A genotype (P=0.009) early- vs late-onset) and 244 controls rs301443 associated with early-onset Asn107 carriers with 4 years Age of onset (as earlier onset than Ile107 continuous variable) 232 OCD and 308 NPSR1 Asn107Ile (P=0.032) in OCD and Asn107 Lennertz et al. 2013 Case-control study and early onset controls (rs324981) associated with early onset OCD defined as <12 years when compared to controls (P=0.0004) rs6265 Val associated with OCD when compared to controls (P=0.0001) and rs1519480 with significant association (P=0.0001); Gender, age of onset 232 OCD, 111 first- rs7124442 C allele associated (as continuous BDNF rs6265, rs1519480, Márquez et al. 2013 Case-control study degree relatives, and with OCD females when variable), symptom and rs7124442 283 controls compared to female controls severity (P=0.0183) and C allele associated with male controls when compared to OCD males (P=0.0018)

361

362

266 SNPs in 35 genes (Dopamine: DRD2, DRD3, DRD4, SLC6A3, COMT, MAOA; Serotonin: HTR1B, HTR2A, HTR2C, SLC6A4, HTR1B rs2000292, GAD2 SLC18A1, TPH2; rs8190748 and rs992990 were Glutamate: GRIK2, Genetic over-transmitted from parents to GRIN2B, GRIA1, GRIA3, Early-onset transmission early-onset OCD (P=0.01-0.006); Mas et al. 2014 75 OCD trios SLC1A1, DLGAP3; GABA: ± disequilibrium HTR1B rs2000292 associated (11.83 3.09 years) GAD1, GAD2; BDNF: study with males (P=0.0006) and GAD2 NTRK2, NTRK3, BDNF, rs8190748 with females AKT1, GSK3β; (P=0.0006) Neuroregulin: NGFR, ERBB4, NRG1, OLIG1, OLIG2; Others: LMX1A, BDKRB2, CDH9, KCNN3, and EFNA5 Dallaspezia et Age of onset (as Genetic association C/C genotype associated with 2014 225 OCD SLC1A1 rs301430 al. continuous variable) study later onset (P=0.017) La allele was over-transmitted to Trio study and 103 OCD children early-onset OCD (P=0.0054) and Walitza et al. 2014 Early-onset SLC6A4 5HTTLPR meta-analysis and adolescents trios significant meta-analysis (P=0.00021)

362

363

Table 2. Extensive Summary Table of Studies of OCD Symptom Dimensions.

Analysis Number of Variance Study Year Scoring Factors Technique Subjects Explained PCA, category- Range=0- 3 factors: symmetry and hoarding, contamination and cleaning, pure Baer et al. 1994 level, current 107 OCD 48% 2 obsessions symptoms

Hantouche PCA, category- 3 factors (category-level): predominantly compulsive, predominantly 32.5% and 1996 N/A and item-level, 615 OCD obsessive, mixed Lancrenon current symptoms 17 factors (item-level) N/A PCA, category- 2 independent Leckman et Number of 4 factors: obsessions and checking, symmetry and ordering, 1997 level, lifetime OCD samples of 63.5% al. symptoms cleanliness and washing, hoarding symptoms 208 and 98 Number of symptoms PCA as per 100 OCD and Alsobrook II as per Leckman et al. 4 factors: aggressive/checking/sexual, hoarding, 1999 466 first-degree ~70% et al. Leckman (1997), category- symmetry/ordering/counting/ritual, contamination relatives et al. level (1997) Confirmatory 4 factors: aggressive/sexual/religious/somatic/checking, Summerfeldt Range=0- factor analysis, 1999 203 OCD symmetry/repeating/ordering/counting, contamination and N/A et al. 1 category-level, cleaning/washing, hoarding current symptoms PCA, category- Mataix-Cols Range=0- 5 factors: symmetry/ordering, hoarding, contamination/cleaning, 1999 level, current 354 OCD 65.5% et al. 2 aggressive/checking, sexual/religious obsessions symptoms Girishchandra PCA, item-level, 5 factors: contamination/cleaning, hoarding/symmetry/arranging, 2001 N/A 202 OCD (Indian) 34.9% and Khanna lifetime symptoms aggressive/doubts/checking, superstitions, sexual/religious 5 factors: contamination/cleaning/repeating, Range=0- PCA, current Tek and Ulug 2001 45 OCD symmetry/ordering/somatic, aggressive/counting/hoarding, 65.5% 1 symptoms sexual/religious, checking/hoarding PCA, (category- 5 factors: contamination/cleaning, hoarding, Range=0- 180 OCD and Cavallini et al. 2002 level), lifetime sexual/religious/aggressive/checking, symmetry/ordering, 59.87% 1 112 controls symptoms repeating/counting

363

364

PCA as per Mataix-Cols Range=0- Mataix-Cols et al. 5 factors: symmetry/ordering, hoarding, contamination/cleaning, 2002a 117 OCD N/A et al. 2 (1999), category- aggressive/checking, sexual/religious level PCA, category- Mataix-Cols Range=0- 5 factors: aggressive/checking, contamination/washing, 2002b level, current 153 OCD 63.7% et al. 1 symmetry/ordering, hoarding, sexual/somatic symptoms PCA, item-level, 4 factors: Feinstein et Range=0- 2003 current baseline 160 OCD symmetry/exactness/ordering/arranging/repeating/counting/touching, 54.1% al. 1 symptoms contamination/cleaning/aggressive/checking, hoarding, sexual/religious 5 factors: aggressive/sexual/religious obsessions, Range=0- PCA, item-level, Denys et al. 2004a 150 OCD contamination/cleaning, somatic/checking, symmetry/ordering, high 42.5% 2 current symptoms risk aggressive/checking 5 factors: contamination/cleaning, aggressive/sexual/religious/somatic, Range=0- PCA, item-level, Denys et al. 2004b 335 OCD somatic obsessions and checking, symmetry and prefectionism, high 41.7% 2 current symptoms risk aggressive and checking PCA, category- 4 factors: aggressive/sexual/religious/checking, Range=0- Hasler et al. 2005 level, lifetime 169 OCD symmetry/repeating/counting/ordering/arranging, 64.7% 1 symptoms contamination/cleaning, hoarding Range=0- PCA, category- 124 OCD and 4 factors: hoarding/repeating, contamination/cleaning, Kim et al. 2005 62.8% 2 level 171 controls aggressive/sexual, religious/somatic PCA, category- 4 factors: aggressive/religious/checking, Range=0- Hasler et al. 2006 level, lifetime 153 OCD symmetry/repeating/counting/ordering, contamination/cleaning, 65.0% 1 symptoms hoarding Exploratory factor Range=0- 5 factors: symmetry/ordering, hoarding, doubt/somatic/checking, Pinto et al. 2007 analysis (PCA), 293 OCD 65.5% 1 and 0-2 contamination/cleaning, taboo thoughts (aggressive/sexual/religious) category-level 221 definite or Dichotomous probable OCD exploratory factor Range=0- (129 definite or Cullen et al. 2007 analysis, 4 factors: pure obsessions, contamination, symmetry/order, hoarding 76.7% 1 probable OCD category-level, and 92 multiplex lifetime symptoms OCD families) Wendland et Range=0- PCA, category- 347 OCD and 4 factors: obsessions/checking, symmetry/ordering, 2007 N/A al. 1 level 749 controls contamination/cleaning, hoarding

364

365

Confirmatory factor analysis, 53 OCD, 96 non- Range=0- category-level, OCD psychiatric 3 factors: contamination/cleaning/repeating/checking/somatic, Wu et al. 2007 N/A 1 using self-report patients, and 419 aggressive/sexual/religious, hoarding/ordering/counting/symmetry version of Y- students BOCS Exploratory factor 5 factors: contamination/cleaning, harm/checking, Range=0- analysis, item- Stein et al. 2007 434 OCD aggressive/sexual/religious, hoarding/symmetry, N/A 2 level, current somatic/hypochondriacal symptoms PCA as per Baer (1994), Leckman et al. (1997) and Mataix-Cols et al. 4 factors: aggressive/sexual/religious/somatic/checking, Range=0- (2005), and Hasler et al. 2007 418 OCD symmetry/repeating/counting/ordering/arranging, 64.0% 1 cluster analysis contamination/cleaning, hoarding using Ward's method, category- level, lifetime symptoms Cluster analysis 3 clusters (2.1 linkage distance level): contamination/washing, using Ward's hoarding/symmetry/ordering, obsessional/checking; [6 clusters (1.5 Range=0- Lochner et al. 2008 method, item- 261 OCD linkage distance level but "clumpsy"): contamination/washing, N/A 1 level, current hoarding, symmetry/ordering/arranging/repeating, sexual, symptoms somatic/religious, aggressive/harm-related] PCA, category- Matsunaga et Range=0- 343 OCD 4 factors: contamination/washing, hoarding, 2008 level, current 57.7% al. 2 (Japanese) symmetry/repeating/ordering, aggressive/checking symptoms

Dichotomous Factor Analysis DFA-5 factors: taboo thoughts, symmetry/ordering, hoarding, (DFA) for item- N/A contamination/cleaning, doubt/checking Range=0- level analysis, Pinto et al. 2008 485 OCD 1 lifetime symptoms

PCA for PCA-5 factors: symmetry/ordering, taboo thoughts, hoarding, categorical-level 64.1% doubt/checking, contamination/cleaning analysis, lifetime

365

366

symptoms

Exploratory factor Range=0- 5 factors: contamination/cleaning, harm/checking, hoarding/symmetry, Stein et al. 2008 analysis, item- 466 OCD N/A 2 religious/sexual, somatic/hypochondriacal level Y-BOCS and Thoughts and Behaviour Inventory (TBI; Schooler et Range=0- Slattery et al., 5 factors: fear of hurting self/others, repetition/symmetry, 2008 398 OCD N/A al. 1 2004), contamination, checking/counting, hoarding confirmatory factor analysis, item-level, lifetime symptoms Factor analysis at item-level using tetrachoric Katerberg et Range=0- 373 OCD and 6 factors: taboo, contamination/cleaning, rituals and superstition, fear 2010a correlation 66% al. 1 462 controls of harm, intolerance of uncertainty, somatic and sensory phenomena coefficient estimates, lifetime symptoms

PCA with 5 factors (item-level): taboo, contamination/cleaning, doubts, 1224 OCD and 59% confirmatory superstitions/rituals, symmetry/hoarding Katerberg et Range=0- 52 OCD-affected 2010b factor analysis, al. 1 multigenerational item-level, lifetime families 4 factors (categorical-level): symptoms symmetry/ordering/arranging/counting/repeating, 65.5% aggressive/sexual/religious/checking, contamination/cleaning, hoarding Range=0- PCA, current 5 factors: contamination/cleaning, hoarding, symmetry/ordering, Jang et al. 2010 144 OCD 65.41% 2 symptoms aggressive/sexual/religious/checking, repeating/counting Range=0- PCA, lifetime 4 factors: symmetry/repeating/ordering/counting/checking, Albert et al. 2010 329 OCD 58% 1 symptoms aggressive/religious/sexual/somatic, contamination/cleaning, hoarding Range=0- PCA as per 5 factors as per Mataix-Cols et al. (1999): symmetry/ordering, 65.5% as 229 OCD and Alonso et al. 2011 2 as per Mataix-Cols et al. hoarding, contamination/cleaning, aggressive/checking, per 279 controls Mataix- (1999) sexual/religious obsessions Mataix-

366

367

Cols et al. Cols et al. (1999) (1999) PCA as per Leckman et al. 200 OCD and Range=0- 5 factors: hoarding, contamination/cleaning, symmetry/ordering, Liu et al. 2011 (1997), category- 403 controls 71.8% 1 aggressive/checking, somatic/repeating level, current (Chinese Han) symptoms 5 factors: doubts/checking/repeating, contamination/washing/cleaning, PCA, category- Cherian et al. 2012 N/A 545 OCD hoarding/collecting, symmetry/ordering, 62% level forbidden/sexual/religious/aggressive

As per Mataix- Cols et al. (1999) As per using Y-BOCS Mataix- 225 OCD and 5 factors (previously identified): symmetry/ordering, hoarding, Alonso et al. 2012 and DY-BOCS for N/A Cols et al. 279 controls contamination/cleaning, aggression/checking, sexual/religious dimensions but (1999) did not perform factor analysis

PCA on 74 OCD 2 African Range=0- subjects, American OCD 6 factors: contamination/washing, hoarding, sexuality/reassurance, Wiliams et al. 2012 59.06% 2 category-level, samples (74 and aggression/mental, symmetry/perfectionism, doubt/checking current symptoms 54) Brakoulias et Range=0- PCA, category- 5 factors: hoarding, contamination/cleaning, doubt/checking, 2013 154 OCD 64.9% al. 2 level symmetry/ordering, unacceptable/taboo thoughts 173 OCD and PCA, category- 5 factors: hoarding, doubts/checking, symmetry/ordering, Taj et al. 2013 N/A 201 controls 65% level contamination, forbidden thoughts (South Indian) 67.08% exploratory factor 5 factors (category-level): symmetry/arranging/repeating/counting, (category- analysis (PCA), contamination/cleaning, hoarding, aggressive/checking, Range=0- 512 OCD level) and Zhang et al. 2013 category- and somatic/religious/sexual; 6 factors (item-level): contamination/cleaning, 1 (Chinese) 66.10% item-level, lifetime symmetry/arranging/repeating/counting/superstition, hoarding, (item- symptoms doubt/checking, somatic, religious/sexual/mental level) 58 parent-child Lennertz et Range=0- PCA, category- 4 factors: symmetry/ordering/repeating/counting, 2014 trios, 236 OCD 53.9% al. 1 level contamination/washing, hoarding, aggressive/sexual/religious and 310 controls

367

368

Brakoulias et PCA, category- 5 factors: hoarding, contamination/cleaning, symmetry/ordering, 2014 N/A 154 OCD 64.9% al. level unacceptable/taboo thoughts, doubt/checking

Exploratory factor Range=0- analysis, 5 factors: doubts about harm/checking, unacceptable/taboo thoughts, Williams et al. 2014 238 OCD 68.5% 2 category-level, contamination/cleaning, hoarding, symmetry/ordering current symptoms

Multidimensional item response Range=0- 269 OCD and Martoni et al. 2015 theory model, 5 factors: hoarding, washing, symmetry, rituals, forbidden thoughts N/A 1 120 controls category-level, current symptoms

368

369

Table 3. Extensive Summary Table of Genetic Studies of OCD Symptom Dimensions.

Analysis Number of Variance Study Year Scoring Factors Genetics Technique Subjects Explained Alsobroo 1999 Number PCA as per 100 OCD and 4 factors: ~70% Segregation analysis k II et al. of Leckman et al. 466 first- aggressive/checking/sexual, with a relative risk of symptom (1997), degree hoarding, 1.7 for OCD or s as per category-level relatives symmetry/ordering/counting/ritual, subclinical OCD Leckman contamination symmetry/ordering et al. factor symptoms in (1997) relatives Cavallini 2002 Range=0 PCA, (category- 180 OCD and 5 factors: contamination/cleaning, 59.87% 5HTTLPR long/long et al. -1 level), lifetime 112 controls hoarding, genotype associated symptoms sexual/religious/aggressive/checking with higher , symmetry/ordering, repeating/counting repeating/counting factor scores (P=0.0013) Kim et al. 2005 Range=0 PCA, category- 124 OCD and 4 factors: hoarding/repeating, 62.8% 5HTTLPR = not -2 level 171 controls contamination/cleaning, significant aggressive/sexual, religious/somatic Hasler et 2006 Range=0 PCA, category- 153 OCD 4 factors: 65.0% 5HTTLPR short/short al. -1 level, lifetime aggressive/religious/checking, genotype and short symptoms symmetry/repeating/counting/orderin allele associated with g, contamination/cleaning, hoarding symmetry/repeating/co unting/ordering factor Wendlan 2007 Range=0 PCA, category- 347 OCD and 4 factors: obsessions/checking, N/A BDNF Val66Met d et al. -1 level 749 controls symmetry/ordering, (rs6265), SLC6A4 contamination/cleaning, hoarding 5HTTLPR, rs25531, STin2, and I425V/I425L = not significant with factors, early/late AAO, or familiality

369

370

Lochner 2008 Range=0 Cluster analysis 261 OCD 3 clusters (2.1 linkage distance N/A COMT Val158Met et al. -1 using Ward's level): contamination/washing, (rs4680) Met/Met method, item- hoarding/symmetry/ordering, genotype in Caucasian level, current obsessional/checking; [6 clusters subset and earlier symptoms (1.5 linkage distance level but onset associated higher "clumpsy"): contamination/washing, obsessional/checking hoarding, cluster scores symmetry/ordering/arranging/repeati (P<0.001); no ng, sexual, somatic/religious, significant results for aggressive/harm-related] DRD4 VNTR, DAT VNTR, 5HTTLPR, HTR1B G861C (rs6269), HTR2A T102C (rs6313), TH Val81Met (rs6356), and MAOA C1460T (ECORV) Schooler 2008 Range=0 Y-BOCS and 398 OCD 5 factors: fear of hurting self/others, N/A Class II OCD (more et al. -1 Thoughts and repetition/symmetry, contamination, severe OCD symptoms Behaviour checking/counting, hoarding across 5 factors) Inventory (TBI; associated with Slattery et al., significant family 2004), history, earlier onset, confirmatory and greater psychiatric factor analysis, comorbidity and item-level, impairment lifetime symptoms Katerber 2010a Range=0 Factor analysis 373 OCD and 6 factors: taboo, 66% COMT Val158Met g et al. -1 at item-level 462 controls contamination/cleaning, rituals and (rs4680) associated using superstition, fear of harm, with somatic and tetrachoric intolerance of uncertainty, somatic sensory phenomena correlation and sensory phenomena factor in females coefficient (P=0.024, not survived estimates, correction for multiple lifetime comparisons) symptoms

370

371

Katerber 2010b Range=0 PCA with 1224 OCD and 5 factors (item-level): taboo, 59% Heritabilities: 0.26 g et al. -1 confirmatory 52 OCD- contamination/cleaning, doubts, taboo, 0.24 factor analysis, affected superstitions/rituals, contamination/cleaning, item-level, multigeneration symmetry/hoarding 0.44 doubts, 0.34 lifetime al families rituals/superstitions, symptoms 0.35 hoarding/symmetry 4 factors (categorical-level): 65.5% Heritabilities: 0.00 symmetry/ordering/arranging/countin symmetry/counting, g/repeating, 0.58 aggressive/sexual/religious/checking aggressive/sexual/religi , contamination/cleaning, hoarding ous, 0.32 contamination/cleaning, 0.23 hoarding Alonso et 2011 Range=0 PCA as per 229 OCD and 5 factors as per Mataix-Cols et al. 65.5% as per 29 SNPs spanning the al. -2 as per Mataix-Cols et 279 controls (1999): symmetry/ordering, Mataix-Cols et estrogen receptor alpha Mataix- al. (1999) hoarding, contamination/cleaning, al. (1999) (ESR1) and estrogen Cols et aggressive/checking, receptor beta (ESR2) al. sexual/religious obsessions genes with ESR1 (1999) rs34535804 A allele associated with lower risk of contamination/cleaning factor symptoms (P=0.0001, remained significant after multiple testing correction) and ESR1 rs34535804*A- rs488133*C- rs9478245*C- rs2234693*C- rs9340799*G haplotype associated with less contamination/cleaning factor symptoms when compared to those without these

371

372

symptoms (P=0.018)

Liu et al. 2011 Range=0 PCA as per 200 OCD and 5 factors: hoarding, 71.8% COMT A-287G G/G -1 Leckman et al. 403 controls contamination/cleaning, genotype more frequent (1997), (Chinese Han) symmetry/ordering, in OCD (P=0.00091) category-level, aggressive/checking, than controls with no current somatic/repeating association in symptom symptoms dimensions Alonso et 2012 As per As per Mataix- 225 OCD and 5 factors (previously identified): N/A GRIN2B rs1805476 al. Mataix- Cols et al. 279 controls symmetry/ordering, hoarding, associated with OCD Cols et (1999) using Y- contamination/cleaning, males (P=0.002) and a al. BOCS and DY- aggression/checking, 4-SNP haplotype (1999) BOCS for sexual/religious (rs1805476.rs1805501. dimensions but rs1805502.rs1805477) did not perform associated with factor analysis contamination/cleaning symptoms (P=0.023) but no significant association for rs4522263, rs1558766, rs3026174, rs1805501, rs1805502 (T588C), rs1805476 (A5806C), rs1805477, rs890 (T5072G), rs1805246, and rs1806191

372

373

Taj et al. 2013 N/A PCA, category- 173 OCD and 5 factors: hoarding, doubts/checking, 65% DRD4 VNTR 7R allele level 201 controls symmetry/ordering, contamination, greater in OCD than (South Indian) forbidden thoughts controls (P=0.03) and more in OCD females (P=0.02) and a trend with 2R allele in symmetry/ordering factor (P=0.01 uncorrected) Lennertz 2014 Range=0 PCA, category- 58 parent-child 4 factors: 53.9% HTR3E rs7627615 G et al. -1 level trios, 236 OCD symmetry/ordering/repeating/countin allele associated with and 310 g, contamination/washing, hoarding, contamination/washing controls aggressive/sexual/religious factor (P=0.0001) but not significant for HTR3A rs1062613, HTR3B rs1176744, HTR3C rs6766410 and rs6807362, and HTR3D rs1000952

Scoring: investigators assigned a score of 1 if a symptom category was present and 0 if it was absent. Score 2 was assigned in some studies for the most prominent symptom category.

PCA: principal component analysis.

N/A: not available

Table adapted from Mataix-Cols et al. (2005) Am J Psychiatry 162:228-238.

373

374

Copyright Acknowledgements

374