<<

Characterization of Mice with Altered Transporter and Vesicular 2 Levels

by

Shababa Tanzeel Masoud

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Pharmacology and Toxicology University of Toronto

© Copyright by Shababa Tanzeel Masoud 2017

Characterization of Mice with Altered and Vesicular Monoamine Transporter 2 Levels

Shababa Tanzeel Masoud

Doctor of Philosophy

Department of Pharmacology and Toxicology

University of Toronto

2017 Abstract

Dopamine is a key neurotransmitter that regulates motor coordination and dysfunction of the dopamine system gives rise to Parkinson’s disease. Nigrostriatal dopamine neurons are vulnerable to various genetic and environmental insults, suggesting that these cells are inherently at-risk. A cell-specific risk factor for these neurons is the neurotransmitter, dopamine itself. If intracellular dopamine is not appropriately sequestered into vesicles, it can accumulate in the cytosol. Cytosolic dopamine is highly reactive and can trigger oxidative stress, leading to cellular toxicity. Cytosolic dopamine levels are modulated by the plasma membrane dopamine transporter (DAT) that takes up dopamine from the extracellular space, and the vesicular monoamine transporter 2 (VMAT2) that stores dopamine into vesicles. In this thesis, we altered

DAT and VMAT2 levels to investigate the detrimental consequences of potentially amplifying cytosolic dopamine in transgenic mice. Project 1 focused on selective over-expression of DAT in dopaminergic cells of transgenic mice (DAT-tg). DAT-tg mice displayed phenotypes of dopaminergic damage: increased dopamine-specific oxidative stress, L-DOPA-reversible fine motor deficits and enhanced sensitivity to toxicant insult, suggesting that increasing DAT- mediated dopamine uptake is detrimental for dopamine cells. As an extension of Project 1,

ii

Project 2 focused on mice that simultaneously over-express DAT and under-express VMAT2

(DAT-tg/VMAT2-kd mice). These animals were hypothesized to demonstrate exacerbated dopaminergic toxicity due to buildup of cytosolic dopamine caused by increased uptake and decreased packaging. While DAT-tg/VMAT2-kd mice displayed detrimental phenotypes (poor survival, decreased body weight, reduced dopamine tissue content and release) and compensatory changes (increased dopamine receptors and metabolism), they did not show dopamine cell loss. This is due to unexpected loss of phenotypes in DAT-tg mice from a new colony that no longer displayed dopaminergic neurodegeneration. Thus, instead of Parkinsonian behavior, DAT-tg/VMAT2-kd mice showed novel phenotypes such as hyperactivity and improved fine-motor and cognitive skills compared to other genotypes. DAT-tg/VMAT2-kd mice were also highly sensitive to -induced locomotion. Hence, in the absence of neurodegeneration, altered DAT and VMAT2 levels produced unique behavioral changes in

DAT-tg/VMAT2-kd mice, shedding light on the complex function of the dopamine system.

Collectively, these studies demonstrate how perturbations in dopamine compartmentalization can impact dopamine homeostasis and behavior.

iii

Acknowledgments

First, I would like to thank my supervisor, Dr. Ali Salahpour, for his immense guidance and mentorship during my Ph.D. As one of his first students, I have had the privilege of learning from him directly and seeing the lab grow over the years. His enthusiasm for science, positive outlook, understanding nature and approachability make him a truly unique supervisor.

To my committee members, Drs. Peter G. Wells, José Nobrega, W. M. Burnham and David S. Riddick: you have been my guiding light throughout this Ph.D. You have challenged, supported and encouraged me. I am eternally grateful for having the best Ph.D. supervisory committee I could ever hope for. A special thank you to Dr. David S. Riddick for playing the dual role of my co-supervisor and thesis reader. You have always had the time to check up on me, provide constructive criticism and guide me in the right direction. Also, a special thank you to Dr. W. M Burnham – I started my scientific journey in your lab as a 4th year project student and since then, I have shared a great working relationship with you as the TA for PCL475. Thank you for your kindness and for always having my best interest in mind.

To Dr. Amy Ramsey, thank you for offering your expertise and advice throughout my Ph.D. To our collaborators: Drs. Gary W. Miller, Jason Richardson, Jonathan Brotchie and Andrei Starostin – I truly appreciate your invaluable technical help with my projects. I would like to gratefully acknowledge Dr. Salah El Mestikawy for being my external examiner. A special mention for Lien Nguyen, my undergraduate project student, for her useful contribution to these experiments. I am also grateful for my sources of funding from Parkinson Society of Canada, Canadian Institutes of Health Research and the University of Toronto.

Wendy Horsfall, you are the backbone of our lab – I cannot thank you enough for sharing your knowledge and being so patient with us. Marija Milenkovic and Dr. Laura Vecchio, thank you for helping me every day and being my voice of reason. To all members of the Salahpour and Ramsey labs, I am grateful to have shared this journey with you.

Finally, I would like to extend my deepest gratitude to my family. To my parents, Chowdhury A. Masud and Shabina M. Masud - you never doubted me even for a moment. You stood by me as pillars of strength throughout all my struggles and I will forever remain grateful. To Nafees, you supported me in every way imaginable. Thank you for being my teammate.

iv

Table of Contents

Acknowledgments...... iv

Table of Contents ...... v

List of Figures ...... x

List of Tables ...... xiii

List of Appendices ...... xiv

List of Publications ...... xv

List of Abbreviations ...... xvi

Chapter 1 Introduction ...... 1

Introduction ...... 1

1.1 Statement of Research Problem ...... 1

1.2 Literature Review...... 2

1.2.1 Dopamine function in the brain ...... 2

1.2.1.1 Nigrostriatal pathway and movement ...... 3

1.2.1.2 Other dopaminergic pathways ...... 8

1.2.2 Dopamine homeostasis ...... 9

1.2.2.1 Synthesis ...... 9

1.2.2.2 Release ...... 11

1.2.2.3 Degradation ...... 12

1.2.3 Dopamine transport ...... 14

1.2.3.1 Plasma membrane transport ...... 14

1.2.3.2 Vesicular membrane transport ...... 15

1.2.4 Dopamine compartmentalization and its effects ...... 17

1.2.4.1 Extracellular dopamine ...... 17

1.2.4.1.1 Dopamine Receptors...... 17

v

1.2.4.2 Intracellular dopamine ...... 20

1.2.4.2.1 Cytosolic dopamine ...... 20

Reactivity ...... 21

Toxicity ...... 24

1.2.5 Classical drugs that interact with the dopamine system ...... 28

1.2.5.1 Enzyme ligands ...... 29

1.2.5.2 DAT ligands ...... 30

1.2.5.3 VMAT2 ligands ...... 33

1.2.5.4 Dopamine receptor ligands ...... 35

1.2.6 Parkinson’s disease ...... 37

1.2.6.1 Symptoms ...... 38

1.2.6.2 Pathology ...... 39

1.2.6.3 Therapy ...... 41

1.2.6.4 Etiology ...... 42

1.2.6.5 Vulnerability of nigrostriatal dopaminergic cells ...... 44

1.2.6.5.1 Role of cytosolic dopamine in Parkinson’s disease...... 47

1.2.6.5.2 Role of dopamine transporters in Parkinson’s disease ...... 49

1.2.7 Animal models with altered transporter levels ...... 52

1.2.7.1 DAT-knockout mice ...... 54

1.2.7.2 DAT-overexpressing transgenic mice ...... 55

1.2.7.3 VMAT2-knockout homozygote mice ...... 56

1.2.7.4 VMAT2-knockout heterozygote mice ...... 57

1.2.7.5 VMAT2-knockdown mice ...... 58

1.2.7.6 VMAT2-overexpressing mice ...... 61

vi

1.3 Rationale, Hypothesis and Aims ...... 62

Chapter 2 Materials and Methods ...... 64

Materials and Methods ...... 64

2.1 Mice ...... 64

2.1.1 Generation of DAT-tg mice (Project 1) ...... 64

2.1.2 Generation of DAT-tg/VMAT2-kd mice (Project 2) ...... 64

2.1.3 Body weight ...... 65

2.1.4 Survival ...... ……………………………………………………………………....66

2.2 Biochemistry ...... 66

2.2.1 Western blots ...... 66

2.2.2 Quantitative reverse transcriptase PCR ...... 67

2.2.3 Immunohistochemistry ...... 68

2.3 Neurochemistry ...... 68

2.3.1 High performance liquid chromatography (HPLC) ...... 68

2.3.2 Fast-scan cyclic voltammetry (FSCV) ...... 69

2.4 Stereology ...... 70

2.5 Radioligand binding ...... 72

2.6 Behavioral Assessments...... 73

2.6.1 Open field locomotor activity ...... 73

2.6.2 Wire-hang test ...... 74

2.6.3 Challenging beam traversal task ...... 74

2.6.4 Puzzle box ...... 76

2.6.5 Elevated plus maze ...... 77

2.6.6 Abnormal Involuntary Movements Scale ...... 78

2.7 Drug treatment ...... 79

2.7.1 MPTP ...... 79 vii

2.7.2 Dopaminergic drugs ...... 79

2.8 Statistics ...... 80

Chapter 3 Results ...... 81

Results ...... 81

3.1 Characterization of DAT over-expressing transgenic mice ...... 81

3.1.1 Presynaptic dopamine homeostasis ...... 81

3.1.2 Markers of oxidative stress ...... 84

3.1.3 Motor behavior ...... 90

3.1.4 Response to MPTP-induced dopaminergic damage ...... 95

3.2 Characterization of mice that over-express DAT and under-express VMAT2 ...... 98

3.2.1 Confirmation of transporter levels ...... 98

3.2.2 Fitness ...... 102

3.2.3 Presynaptic dopamine homeostasis ...... 106

3.2.4 Integrity of dopamine neurons ...... 116

3.2.5 Dopamine receptor levels ...... 120

3.2.6 Baseline behavior ...... 123

3.2.7 Response to dopaminergic drugs...... 135

Chapter 4 Discussion ...... 148

Discussion ...... 148

4.1 Project 1: Characterization of DAT-tg mice ...... 148

4.2 Project 2: Characterization of DAT-tg/VMAT2-kd mice ...... 153

4.2.1 Discrepancy between original DAT-tg mice and DAT-tg mice from the DAT- tg/VMAT2-kd colony ...... 159

4.2.2 Hypothesis revisited ...... 161

4.3 Conclusion ...... 162

4.4 Technical Challenges ...... 164

viii

4.5 Future Directions ...... 165

References ...... 168

Appendix 1 ...... 195

Appendix 2 ...... 201

Copyright Acknowledgements...... 208

ix

List of Figures

CHAPTER 1

Figure 1-1. Dopaminergic pathways of the brain...... 3

Figure 1-2. Direct and indirect pathways of the basal ganglia...... 5

Figure 1-3. Synthetic pathway of dopamine ...... 11

Figure 1-4. Degradation pathways for dopamine...... 13

Figure 1-5. Dopamine transport in the presynaptic neuron ...... 14

Figure 1-6. Generation of reactive oxygen species in dopamine cells ...... 23

Figure 1-7. Substrates for DAT cause selective damage to dopamine neurons...... 32

CHAPTER 2

Figure 2-1. Wire-hang test apparatus...... 74

Figure 2-2. Challenging beam traversal task...... 75

Figure 2-3. Puzzle box apparatus...... 76

Figure 2-4. Schematic image of elevated plus maze...... 78

CHAPTER 3: Project 1 - DAT-tg mice

Figure 3-1. DAT protein expression in the striatum of DAT-tg mice...... 82

Figure 3-2. Metabolite to dopamine ratios in the striatum of DAT-tg mice...... 83

Figure 3-3. VMAT2 protein expression in the striatum of DAT-tg mice...... 84

x

Figure 3-4. Protein carbonylation in the striatum of DAT-tg mice...... 86

Figure 3-5. Protein nitrosylation and MnSOD levels in DAT-tg mice...... 87

Figure 3-6. Cysteinyl adducts of dopamine and its metabolites in DAT-tg mice...... 89

Figure 3-7. Motor behavior of DAT-tg mice...... 91

Figure 3-8. Challenging beam traversal task in DAT-tg mice with L-DOPA treatment...... 93

Figure 3-9. Baseline behaviors of DAT-tg mice stratified by sex...... 94

Figure 3-10. Effect of MPTP treatment on TH protein levels in DAT-tg mice...... 96

Figure 3-11. Effect of MPTP on striatal dopamine tissue content of DAT-tg mice...... 97

CHAPTER 3: Project 2 - DAT-tg/VMAT2-kd mice

Figure 3-12. DAT protein expression in the striatum...... 99

Figure 3-13. VMAT2 protein levels in the striatum...... 100

Figure 3-14. DAT and VMAT2 mRNA expression in the midbrain...... 101

Figure 3-15. Survival curve from birth to 12 weeks of age...... 104

Figure 3-16. Body weight of adult mice...... 105

Figure 3-17. Striatal tissue content of dopamine and its metabolites...... 107

Figure 3-18. . Metabolite-to-dopamine ratios in the striatum...... 108

Figure 3-19. Electrically evoked dopamine release and uptake in the dorsal striatum...... 112

Figure 3-20. TH protein expression in the striatum...... 113

Figure 3-21. MAO-B protein expression in the striatum...... 114

xi

Figure 3-22. Stereological counts of TH+ cells in the SNc...... 117

Figure 3-23. Stereological counts of TH+ and NeuN+ cells in SNpc...... 118

Figure 3-24. Stereological counts of TH+ and Nissl+ cells in SNpc...... 119

Figure 3-25. Dopamine receptor levels in the striatum...... 122

Figure 3-26. Open field locomotion and stereotypy...... 126

Figure 3-27. Locomotor activity of 12-month old mice...... 127

Figure 3-28. Locomotor activity of DAT-tg/VMAT2-het mice...... 128

Figure 3-29. Fine motor skill evaluated using the challenging beam traversal task...... 129

Figure 3-30. Executive function evaluated using the puzzle box...... 131

Figure 3-31. Anxiety-like behavior assessed using elevated plus maze...... 133

Figure 3-32. Amphetamine-induced locomotion...... 137

Figure 3-33. Amphetamine-induced stereotypy...... 138

Figure 3-34. Abnormal involuntary movements (AIM) induced by 2 mg/kg of amphetamine. 139

Figure 3-35. Locomotor effect of 5 mg/kg amphetamine on WT and DAT-tg mice...... 140

Figure 3-36. Locomotion induced by DAT inhibitors, and ...... 142

Figure 3-37. Apomorphine-induced stereotypy...... 143

Figure 3-38. Effect of SKF 81297, L-DOPA and saline on locomotor activity of DAT VMAT2 mice...... 146

xii

List of Tables

Table 1-1. Summary of mouse models with genetically altered DAT or VMAT2 levels...... 53

Table 2-1. Description of tasks on the puzzle box test...... 77

Table 2-2. List of dopaminergic drugs administered ...... 80

Table 3-1. Summary of DAT and VMAT2 expression in DAT VMAT2 mice...... 102

Table 3-2. Summary of overall fitness of DAT VMAT2 mice...... 106

Table 3-3. Summary of presynaptic dopamine homeostasis in DAT VMAT2 mice...... 115

Table 3-4. Summary of dopamine cell counts in SNpc of DAT VMAT2 mice...... 119

Table 3-5. Summary of dopamine receptor levels in the striatum of DAT VMAT2 mice...... 122

Table 3-6. Summary of baseline motor and non-motor behaviors in DAT VMAT2 mice...... 134

xiii

List of Appendices

Appendix 1: Project 2 - Additional experiments……………………………………………195

Appendix 2: Low copy DAT-tg mice……………………………………………………… 201

xiv

List of Publications

Masoud ST, Vecchio LM, Bergeron Y, Hossain MM, Nguyen LT, Bermejo MK, Kile B, Sotnikova TD, Siesser WB, Gainetdinov RR, Wightman RM, Caron MG, Richardson JR, Miller GW, Ramsey AJ, Cyr M, Salahpour A. Increased expression of the dopamine transporter leads to loss of dopamine neurons, oxidative stress and l-DOPA reversible motor deficits. Neurobiol Dis. 2015; 74: 66-75.

Lohr KM*, Masoud ST*, Salahpour A, Miller GW. Membrane transporters as mediators of synaptic dopamine dynamics: implications for disease. Eur J Neurosci. 2017; 45 (1): 20-33.

Trossbach SV*, Bader V*, Hecher L, Pum ME, Masoud ST, Prikulis I, Schäble S, de Souza Silva MA, Su P, Boulat B, Chwiesko C, Poschmann G, Stühler K, Lohr KM, Stout KA, Oskamp A, Godsave SF, Müller-Schiffmann A, Bilzer T, Steiner H, Peters PJ, Bauer A, Sauvage M, Ramsey AJ, Miller GW, Liu F, Seeman P, Brandon NJ, Huston JP, Korth C. Misassembly of full-length Disrupted-in-Schizophrenia 1 protein is linked to altered dopamine homeostasis and behavioral deficits. Mol Psychiatry. 2016; 21 (11): 1561-1572.

Medvedev IO, Ramsey AJ, Masoud ST, Bermejo MK, Urs N, Sotnikova TD, Beaulieu JM, Gainetdinov RR, Salahpour A. D1 dopamine receptor coupling to PLCβ regulates forward locomotion in mice. J Neurosci. 2013; 33 (46): 18125-18133.

*co-first author

xv

List of Abbreviations

AADC aromatic L-amino acid decarboxylase

ADHD attention deficit hyperactivity disorder

AMPT α-methyl-para-

ATP adenosine triphosphate

BAC bacterial artificial chromosome cAMP cyclic adenosine monophosphate

CNS central nervous system

COMT catechol-O-methyltransferase

CREB cAMP response element-binding protein

DAG diacylglycerol

DARPP-32 dopamine- and cAMP-regulated neuronal phosphoprotein

DAT dopamine transporter

DAT-KO dopamine transporter knock-out

DAT-tg dopamine transporter over-expressing transgenic

DAT-tg/VMAT2-kd dopamine transporter overexpressing and vesicular monoamine transporter 2 knockdown

DAT-tg/VMAT2-het dopamine transporter overexpressing and vesicular monoamine transporter 2 heterozygote

DOPAC 3,4-dihydroxyphenylacetic acid

xvi

DOPAL 3,4-dihydroxyphenylacetaldehyde

DOPET 3,4-dihydroxyphenylethanol

FSCV fast scan cyclic voltammetry

GPe globus pallidus external

GPi globus pallidus internal

GPCR G protein coupled receptor

HPLC-EC High performance liquid chromatography with electrochemical detection

5-HT

5-HT 2A Serotonin 2A receptor

HVA homovanillic acid

IP3 inositol trisphosphate

LC locus coeruleus

L-DOPA L-3,4-dihydroxyphenylalanine

LRRK2 leucine-rich repeat kinase 2

MAO monoamine oxidase

MDMA 3,4- methylenedioxymethamphetamine

MnSOD Manganese superoxide dismutase

MPTP 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine

3MT 3-methoxytyramine

NET transporter

xvii

PBS phosphate-buffered saline

PINK1 PTEN-induced kinase 1

PKC protein kinase C

PVDF polyvinylidene difluoride

RGS Regulators of G protein signaling

ROS reactive oxygen species

SN substantia nigra

SNpc substantia nigra pars compacta

SNpr substantia nigra pars reticulata

STN subthalamic nucleus

TH tyrosine hydroxylase

VMAT2 vesicular monoamine transporter 2

VMAT2-het vesicular monoamine transporter 2 knock-out heterozygote

VMAT2-kd vesicular monoamine transporter 2 knock-down

VMAT2-OE vesicular monoamine transporter 2 over-expressor

VTA ventral tegmental area

WT wild-type

xviii

1

Chapter 1 Introduction

Introduction 1.1 Statement of Research Problem

Dopamine neurotransmission is important for a variety of physiological functions including motor coordination and reward-based learning. On the other hand, malfunction of the dopamine system gives rise to disorders such as Parkinson’s disease. While the pathological loss of nigrostriatal dopaminergic neurons in Parkinson’s disease is well established, the etiology of this neurodegeneration typically remains unknown. Both genetic and environmental insults have been implicated in causing selective damage to dopaminergic neurons even though in most instances, their mechanism of toxicity could theoretically have more widespread effects. This suggests that nigrostriatal dopamine neurons possess a unique phenotype with inherent characteristics that render them susceptible to challenges.

In fact, the endogenous neurotransmitter dopamine itself can act as a cell-specific risk factor for dopaminergic cells. When intracellular dopamine accumulates in the cytosolic space, it is highly prone to reactions that give rise to oxidative stress. In particular, cytosolic dopamine has been shown to produce reactive oxygen species and unstable quinones via metabolic, enzyme- dependent and autoxidation reactions (Graham, 1978; Graham and Gutknecht, 1978; Stokes et al., 1999; Ramkissoon and Wells, 2011). Using mostly in vitro and a few in vivo systems, previous studies have documented the potentially toxic effects of cytosolic dopamine accumulation (Filloux and Townsend, 1993; Hastings et al., 1996; Chen et al., 2008; Mosharov et al., 2009). However, typically, these studies injected exogenous dopamine into brain regions or engineered non-dopaminergic cells to take up the neurotransmitter. Thus, it is unclear whether dopaminergic cells that routinely handle this neurotransmitter and are capable of degrading it, may also succumb to cytosolic dopamine-induced toxicity.

Cytosolic dopamine levels are modulated by two key proteins: the dopamine transporter (DAT) and the vesicular monoamine transporter 2 (VMAT2). DAT is located on the presynaptic membrane of dopaminergic neurons and functions in the rapid uptake of dopamine from the

2 extracellular space into the nerve terminal. VMAT2 is located on the vesicular membrane of monoaminergic cells and functions to sequester intracellular neurotransmitters into vesicles for release. In a simplistic sense, DAT acts to increase cytosolic dopamine levels whereas VMAT2 acts to decrease it. In this work, we propose to use these two transporters as tools to manipulate the cytosolic pool of dopamine in vivo. Previously in our laboratory, we generated transgenic mice (DAT-tg) that over-express DAT specifically in dopaminergic neurons (Salahpour et al., 2008). DAT over-expression led to greater dopamine uptake and loss of midbrain dopamine neurons, presumably due to the detrimental effects of cytosolic dopamine accumulation (Masoud et al., 2015). In this work, we investigated phenotypes of DAT-tg mice and assessed markers of oxidative stress since cytosolic dopamine reactivity typically causes oxidative damage. Moreover, in a second project, we generated animals with simultaneously increased DAT and decreased VMAT2 levels to further enhance cytosolic dopamine accumulation in vivo. Instead of applying external, non-physiological concentrations of dopamine like previous studies, we modify endogenous dopamine compartmentalization and assess its impact on the function of the dopamine system. Results from this work will shed light on the role of DAT, VMAT2 and cytosolic dopamine in the inherent vulnerability of nigrostriatal dopamine neurons to insult.

1.2 Literature Review

1.2.1 Dopamine function in the brain

Dopamine was first discovered in the brain almost 60 years ago (Montagu, 1957; Carlsson et al., 1958) and was thought to merely act as an intermediate in norepinephrine and epinephrine synthesis. After years of research challenging this notion, the field finally began to recognize dopamine itself as a key neurotransmitter (Carlsson et al., 1957). Currently, dopamine is one of the most studied neurotransmitters due to its role in diverse physiological functions as well as its contribution to disease states such as Parkinson’s disease, schizophrenia and attention deficit hyperactivity disorder (ADHD). In the brain, dopamine is involved in a variety of functions including locomotion, cognition, motivation, neuroendocrine regulation and response to reward. Outside the central nervous system (CNS), dopamine plays important roles in regulating vascular, adrenal, renal, cardiac and immune function (Missale et al., 1998; Chang et al., 2006; Buttarelli et al., 2011).

3

Although dopamine performs several functions, dopamine neurons account for only a minute percentage (less than 0.001%) of all neurons and are confined to a few discrete regions of the brain (Surmeier et al., 2010). In the human brain, dopaminergic cell bodies have been detected in the substantia nigra (SN), ventral tegmental area (VTA), hypothalamus (posterior, arcuate nucleus, mammillothalamic tract), zona incerta, periventricular nucleus and olfactory bulb (Fuxe, 1965; Björklund and Dunnett, 2007). From these small regions, dopamine neurons project to various structures in the brain to exert their effects. Classically, dopaminergic projections in the brain are divided into four major pathways: 1) nigrostriatal, 2) mesolimbic, 3) mesocortical and 4) tuberoinfundibular. The majority of these pathways (nigrostriatal, mesolimbic and mesocortical) originate in the midbrain which includes the SN and VTA. The midbrain contains the majority (75%) of dopamine neurons which corresponds to 400,000 to 600,000 cells in adult humans (German et al., 1983; Pakkenberg et al., 1991).

Figure 1-1. Dopaminergic pathways of the brain. A schematic of the 4 major dopamine pathways in the brain showing where they originate and the structures they project to. Image adapted from Genetic Science Learning Center, 2013.

1.2.1.1 Nigrostriatal pathway and movement

The SN is divided into two parts: pars compacta (pc) and pars reticulata (pr). The SNpc consists of densely packed, neuromelanin-containing dopaminergic cell bodies that appear darker than

4 surrounding tissue, hence justifying the name substantia nigra, which is Latin for “black substance”. On the other hand, the SNpr primarily contains diffuse GABAergic neurons. The nigrostriatal pathway refers to dopamine cells that originate in the SNpc and project to the dorsal striatum, alternatively known as the caudate nucleus and putamen. These dopamine cells are also classified as A9 neurons according to the nomenclature proposed in 1964 that initially identified discrete dopamine-containing cell groups in the brain using immunofluorescence (Dahlstroem and Fuxe, 1964; Fuxe, 1965). Functionally, the nigrostriatal pathway is primarily responsible for controlling voluntary movement. Indeed, degeneration of these neurons leads to the motor symptoms that are characteristic of Parkinson’s disease, highlighting the essential role of the nigrostriatal pathway in motor function.

In order to describe how the nigrostriatal pathway regulates motor activity, it is important to understand the role of this pathway in the basal ganglia motor loop. The basal ganglia are a collection of distinct yet interconnected nuclei within the brain that act together to perform multiple functions, the most notable of which is movement control (Obeso et al., 1997). The basal ganglia include subcortical structures such as the striatum (caudate/putamen), globus pallidus internal and external (GPi, GPe), SN and subthalamic nucleus (STN). The basal ganglia also have strong connections with the thalamus and cortex. There are two central basal ganglia pathways that modulate movement: the direct and indirect pathways (Calabresi et al., 2014). In general, the direct pathway facilitates motor activity by removing inhibition on the thalamus and allowing it to excite the cortex and initiate movement. Conversely, the indirect pathway reduces unwanted motor activity by enhancing inhibition of the thalamus which prevents subsequent activation of motor cortices. The balance of these pathways allow for the selection of appropriate voluntary movements. Nigrostriatal dopamine neurons can influence both the direct and indirect pathways of movement. In particular, dopaminergic neurons from the SNpc synapse on to medium spiny GABAergic neurons in the striatum. When dopamine is released, it can activate D1 dopamine receptors on inhibitory GABAergic neurons. These neurons project to the GPi and inhibit its activity. Normally, the GPi provides tonic inhibition of the thalamocortical circuit. However, when dopamine activates the direct pathway, the GPi is strongly inhibited by the striatum, leading to disinhibition of the thalamus, thus allowing it to excite the motor cortex and initiate movement. Dopamine can also modulate the indirect pathway of movement by acting on D2-expressing GABAergic neurons in the striatum. In the indirect pathway, striatal neurons

5 project to the GPe and inhibit its activity. Typically, the GPe is responsible for tonic inhibition of the STN. However, when striatal neurons transiently inhibit the GPe, this releases the STN and allows it to excite the GPi. Activating the GPi leads to greater inhibition of thalamocortical circuits, preventing movement. When dopamine is released, it can inhibit D2-expressing striatal neurons, weakening the downstream effects of the indirect pathway. Hence, nigrostriatal dopamine neurons encourage the direct pathway via D1 receptors and suppress the indirect pathway via D2 receptors. The net effect of these actions by dopamine is to facilitate movement. In general, nigrostriatal dopamine acts as a crucial modulator of the basal ganglia motor loop.

Figure 1-2. Direct and indirect pathways of the basal ganglia. Excitatory input is shown as (+) and inhibitory input is shown as (-). Adapted from Neuroscience, 4th edition (Figure 18.8, Part 2). (Purves et al., 2008).

Besides involvement of the nigrostriatal dopaminergic pathway, other structures also influence the overall motor loop. For instance, the SN receives input from, and also projects to, the STN allowing for negative feedback mechanisms that can regulate the amount of dopamine released.

6

The SNpr also participates in basal ganglia connections as one of the major output structures of the striatum. In addition to the globus pallidus, the striatum also sends GABAergic projections to the SNpr, forming the striatonigral pathway. The SNpr also receives input from radiating dendrites of the SNpc (dopaminergic) as well as the GPe (GABAergic) (Deniau et al., 2007; Beaulieu and Gainetdinov, 2011). Hence, the SNpr is in a unique position to integrate various basal ganglia signals and send efferent projections to the thalamus, brain stem and superior colliculus via predominantly, GABAergic output neurons (Deniau et al., 2007). Finally, both the direct and indirect pathways of movement are under cortical control since the striatum receives input from the cortex.

Nigrostriatal dopamine neurons possess several unique characteristics that distinguish them from other types of neurons. Structurally, these neurons are highly branched and support enormous unmyelinated axonal fields (Matsuda et al., 2009). In humans, it has been estimated that each SNpc dopamine cell gives rise to approximately 370,000 synapses in the striatum (Arbuthnott and Wickens, 2007). In rats, each nigrostriatal axon forms 100,000 to 245,000 synapses, which is orders of magnitude higher than other basal ganglia cells: medium spiny neurons produce 300- 500 synapses and striatal GABAergic interneurons form around 5,000 synapses (Bolam and Pissadaki, 2012). In fact, even dopamine cells of the VTA produce far fewer synapses (12,000 to 30,000) than their nigral neighbors, highlighting the exceptional morphological phenotype of SNpc dopamine cells (Moss and Bolam, 2009; Bolam and Pissadaki, 2012). As a result of this extensive axonal arborization, relatively few nigral dopamine neurons can provide dense innervation of a large target area, the striatum. In order to maintain this axonal complexity, the energetic demands of nigrostriatal neurons are exceptionally high. Energy is required for cytoskeleton maintenance, axonal transport, action potential propagation and synaptic transmission. Indeed, in comparison to VTA dopamine cells, SNpc dopamine neurons have higher density of axonal mitochondria, greater rate of mitochondrial oxidative phosphorylation and elevated ATP production (Pacelli et al., 2015).

Aside from structural complexity, nigrostriatal dopamine neurons also display a distinctive physiological phenotype. Unlike most neurons, these cells are spontaneously active. Even in the absence of synaptic input, SNpc dopamine neurons generate regular action potentials at a slow frequency of 2-4Hz (Guzman et al., 2009). This self-generated pacemaking activity is thought to be responsible for maintaining baseline dopamine levels in the striatum. Most pacemakers,

7 including VTA dopamine neurons, rely on monovalent cations such as sodium for their pacemaking activity (Khaliq and Bean, 2010). However, adult SNpc dopamine neurons also engage voltage dependent L-type Ca channels containing the rare Cav1.3 subunit. This allows the channel to open at relatively hyperpolarized membrane potentials. Hence, calcium enters these cells at subthreshold membrane potentials, allowing for rhythmic oscillations to drive pacemaking in between spikes. Typically, intracellular calcium concentration is under tight homeostatic control due to its involvement in a variety of cellular processes. In most cells, calcium levels are manageable because the ion enters the cell only during evoked action potentials. However, nigrostriatal dopamine neurons experience a constant influx of calcium due to autonomous pacemaking. Therefore, these cells have increased pressure to regulate calcium levels and are more likely to accumulate intraneuronal calcium that can have detrimental effects (Surmeier et al., 2010; Bolam and Pissadaki, 2012). Inside the cell, calcium is buffered by membrane, mitochondrial and endoplasmic reticulum pumps that are metabolically expensive, making nigrostriatal neurons particularly reliant on ATP generation. Thus, similar to the extensive axonal arborization, this unusual calcium-dependent, tonic firing also imposes high energetic demands on nigrostriatal neurons.

Lastly, the most obvious factor that differentiates dopamine neurons from other cells is the neurotransmitter dopamine itself. While extracellular dopamine serves important functions in signaling and neurotransmission, intracellular dopamine also has significant consequences. As a highly reactive molecule, cytosolic dopamine can be auto-oxidized or undergo enzymatic reactions to produce volatile intermediates such as dopamine-quinones and 3,4- dihydroxyphenylacetaldehyde (DOPAL). These reactive derivatives of dopamine can modify cellular proteins, lipids and nucleic acids producing oxidative stress and damage (Graham, 1978; Burke et al., 2003). While all dopaminergic neurons contain dopamine, nigrostriatal neurons are suggested to intrinsically handle higher amounts of the neurotransmitter. In fact, when treated with the precursor of dopamine, L-3,4-dihydroxyphenylalanine (L-DOPA), studies reveal that SN neurons display 2 to 3 times higher accumulation of cytosolic dopamine in comparison to their counterparts in the VTA (Mosharov et al., 2009). Increased content of cytosolic dopamine in nigrostriatal neurons can affect cellular health and lead to toxicity, as discussed in detail in subsequent chapters.

8

In summary, nigrostriatal neurons play an essential role in voluntary movement by participating in the basal ganglia motor loop. These cells also have a distinctive phenotype that sets them apart from other cells as well as other dopaminergic pathways. While complex axonal branching, calcium-dependent pacemaking and high cytosolic dopamine content are unique and necessary features of nigrostriatal dopamine neurons, they can also act as risk factors for these cells (Mosharov et al., 2009; Surmeier et al., 2010; Bolam and Pissadaki, 2012). In fact, healthy humans demonstrate approximately a 40% loss of midbrain dopamine neurons between 40 and 60 years of age, suggesting that these cells are inherently vulnerable (Bogerts et al., 1983; Chinta and Andersen, 2005). Nigrostriatal dopamine neurons are prone to oxidative stress due to the handling of a reactive neurotransmitter and heavy dependence on mitochondrial oxidative phosphorylation to meet their energetic demands. These cell-specific factors may contribute to the vulnerability of nigrostriatal neurons not only in normal aging, but also in disorders such as Parkinson’s disease.

1.2.1.2 Other dopaminergic pathways

In addition to the nigrostriatal pathway, the mesolimbic and mesocortical pathways also originate in the midbrain (Björklund and Dunnett, 2007). However, instead of the SN, the cell bodies of these dopaminergic projections are contained in the VTA. In particular, the mesolimbic tract mainly sends projections to the nucleus accumbens, as well as the amygdala and hippocampus. The nucleus accumbens is a major component of the ventral striatum and plays an important role in reward and motivation. Thus, mesolimbic dopamine is involved in modulating response to rewarding stimuli and is strongly implicated in the behavioral effects of reinforcing drugs. It has been shown that psychostimulants such as cocaine and amphetamine stimulate release of mesolimbic dopamine, whereas withdrawal of these drugs dampens dopamine transmission (Adinoff, 2004; Sulzer, 2011). Mesocortical dopamine neurons primarily innervate the prefrontal cortex in addition to the cingulate and perirhinal cortices. This pathway is involved in cognitive processes such as attention, executive function, learning and memory. Since mesolimbic and mesocortical pathways are closely related and can share overlapping functions, they are collectively referred to as the mesocorticolimbic system. Finally, the fourth classical dopaminergic pathway in the brain is the tuberoinfundibular pathway. These dopamine cells arise from the arcuate and periventricular nuclei of the hypothalamus and send axons to the infundibular region, also known as the median eminence of the hypothalamus. Dopamine is

9 released in the capillary circulation that connects the hypothalamus to the pituitary gland, where it influences hormonal release. In particular, dopamine negatively regulates the release of prolactin, a hormone involved in lactation and reproductive functions. In summary, dopamine is a vital neurotransmitter that performs a variety of functions through different neuronal pathways.

1.2.2 Dopamine homeostasis

Dopamine belongs to a family of catecholamines which is part of a larger class of neurotransmitters known as monoamines. Monoamines are synthesized from particular amino acids and structurally contain an amino group that is connected to an aromatic ring through an ethyl chain. Monoamines include histamine, serotonin, dopamine, epinephrine and norepinephrine. The latter 3 compounds are further classified as catecholamines because they possess a catechol group (which is a benzene ring with 2 hydroxyl groups), that is conjugated with the side chain amine. Catecholamines are derived from the aromatic amino acid, l-tyrosine. Tyrosine can directly be obtained from protein-rich dietary sources or synthesized from the essential amino acid, . For dopaminergic cells, production of dopamine is the ultimate objective, however for other catecholaminergic systems, it serves as an intermediate step. Indeed, dopamine is a precursor in the sequential synthesis of norepinephrine and epinephrine. Specifically, dopamine is converted to norepinephrine by dopamine β hydroxylase while norepinephrine is transformed to epinephrine by the enzyme, N- methyltransferase.

1.2.2.1 Synthesis

The life-cycle of dopamine spans multiple stages including synthesis, vesicular storage, release, uptake and degradation. Synthesis of dopamine is a two-step process that occurs in the cytosol of catecholaminergic cells. The first step involves addition of a hydroxyl group on the phenol ring of the amino acid, L-tyrosine, to convert it to L-DOPA. This reaction is catalyzed by the rate- limiting enzyme, tyrosine hydroxylase (TH) and uses molecular oxygen (O2), iron (Fe) and tetrahydrobiopterin (BH4) as cofactors (Daubner et al., 2011). L-DOPA is then rapidly converted to dopamine by DOPA decarboxylase, generally known as aromatic L-amino acid decarboxylase. This reaction requires pyridoxal phosphate, the active form of vitamin B6, as a cofactor and generates CO2 as a by-product of decarboxylation. Synthesis of dopamine is tightly

10 regulated because it is a major contributor to overall dopamine homeostasis within a cell. As the rate-limiting enzyme in dopamine production, TH expression and activity are under complex regulatory control. For instance, levels of fully synthesized neurotransmitter can influence TH activity, allowing for feedback mechanisms to control intracellular dopamine accumulation

(Daubner et al., 2011). Specifically, dopamine competes with the TH cofactor BH4, to bind iron at the catalytic site of TH. Thus, in the presence of dopamine, the essential cofactor BH4 cannot associate with TH, leading to reversible inhibition of the synthetic enzyme. This provides negative feedback and inhibits further production of dopamine. Activity of TH can also be regulated by the protein’s state of phosphorylation. Phosphorylation of TH at particular sites (Ser19, 31, 40) can enhance its activity and lead to greater production of dopamine, whereas dephosphorylation of TH is correlated with reduced dopamine synthesis. Various signals can influence the phosphorylation of TH. For example, increased extracellular dopamine levels activate the D2 autoreceptor which inhibits phosphorylation of TH at Ser40 and thereby, dampens dopamine production (Lindgren et al., 2001). Conversely, membrane depolarization leads to influx of calcium that activates calcium-dependent kinases which phosphorylate TH and increase dopamine synthesis (Salvatore et al., 2016). Hence, the process of dopamine production is highly responsive to diverse stimuli. Once dopamine is synthesized within the cytosolic space, it is readily sequestered into dense core vesicles by the vesicular monoamine transporter 2 (VMAT2). Vesicular storage of dopamine serves the dual function of protecting the neurotransmitter from degradation and maintaining a high concentration of dopamine for eventual release. The process of vesicular storage is discussed in detail in a subsequent section.

11

Figure 1-3. Synthetic pathway of dopamine Adapted from Carlson, Physiology of Behavior 11th ed. (Carlson, 2012)

1.2.2.2 Release

Vesicular dopamine is released from the presynaptic neuron into the synaptic cleft through the process of exocytosis. This release is triggered by the arrival of an action potential that stimulates the dopaminergic nerve terminal. Depolarization causes voltage-gated calcium channels to open, increasing the presynaptic calcium concentration. Influx of calcium produces a cascade of intracellular events, including the mobilization of dopamine-containing vesicles. These vesicles migrate towards the presynaptic membrane where they are docked and primed. Subsequently, the vesicular membrane fuses with the plasma membrane, releasing dopamine contents into the synaptic cleft. This exocytotic dopamine release is dependent on generation of action potentials and calcium influx as demonstrated by in vitro experiments. In particular, studies show that evoked dopamine release can be blocked by: 1) , an inhibitor of voltage dependent sodium channels and 2) removal of extracellular calcium (Chen and Rice, 2001). Exocytosis is the predominant mechanism of dopamine release and is common to other neurotransmitters as

12 well. While typical neurotransmitter release occurs at the axon terminal (i.e. striatum in nigrostriatal pathway), dopamine can also be released from soma and dendrites in the SN and VTA. Similar to striatal dopamine release, midbrain somatodendritic dopamine release is also reported to regulate voluntary movement through basal ganglia circuits.

Presynaptic dopamine release plays an instrumental role in overall dopamine neurotransmission. Intensity of the dopamine signal relies on multiple factors including the amount of dopamine released, the time course of release events and the neuronal firing rate. Effects of extracellular dopamine on pre- and post-synaptic dopamine receptors are discussed in a subsequent section.

1.2.2.3 Degradation

To terminate the actions of released extracellular dopamine, the neurotransmitter must be removed from the synaptic cleft. This is achieved through 2 processes: 1) recycling dopamine back into the presynaptic neuron through the dopamine transporter (DAT), after which it can be re-packaged into vesicles or degraded and 2) metabolism of dopamine by glial cells. In either case, degradation serves as the final step in the life-cycle of dopamine. Degradation not only concludes the effects of dopamine but also limits buildup of the neurotransmitter to maintain homeostasis.

Within the presynaptic neuron, if dopamine is not sequestered into vesicles, it is available for degradation by metabolic enzymes in the cytosolic space (Eisenhofer et al., 2004a). Dopamine accumulates in the cytosol during synthesis, following extracellular or as a result of vesicular leakage. One of the key enzymes involved in monoamine catabolism is monoamine oxidase (MAO). MAO exists in two forms: MAO-A and MAO-B. In humans, dopamine is mostly metabolized by MAO-B, which is located on the outer mitochondrial membrane (Glover et al., 1977). MAO-B catalyzes the oxidative deamination of dopamine to produce the aldehyde, DOPAL as well as hydrogen peroxide. Both products are highly reactive and can contribute to oxidative stress in the cell (Goldstein et al., 2013). DOPAL can be deactivated to its corresponding , DOPET by aldehyde reductase. However, the more prevalent reaction is the rapid oxidation of DOPAL to a carboxylic acid, DOPAC by aldehyde dehydrogenase. DOPAC is one of the major intracellular metabolites of dopamine. Another important enzyme involved in dopamine degradation is catechol-O-methyltransferase (COMT) which is mainly expressed in glial cells. COMT transfers a methyl group donated from S-adenosylmethionine to a

13 hydroxyl group on DOPAC, generating another major metabolite, homovanillic acid (HVA). In the striatum, metabolism of dopamine primarily begins in the presynaptic neuron. However, a small proportion of circulating extracellular dopamine can also be taken up by glial cells. Since glia express both MAO and COMT, dopamine can be sequentially degraded to DOPAC and HVA as discussed. In an alternative, less significant metabolic pathway, COMT acts on dopamine before MAO. In this case, dopamine is methylated to 3-methoxytyramine (3MT) and then deaminated and oxidized to HVA. Additionally, some reports suggest that COMT is also expressed on post-synaptic neurons where it could participate in metabolism of released dopamine (Elsworth and Roth, 1997). Hence, prevalence of specific dopamine metabolites and preference of particular catabolic pathways depend on the abundance, activity and localization of key metabolic enzymes in different brain regions.

Figure 1-4. Degradation pathways for dopamine. Image adapted from Pérez-Mañá et al., 2015. COMT, catechol-O-methyltransferase; MAO, monoamine oxidase; ALDH, aldehyde dehydrogenase; ADH, alcohol dehydrogenase.

14

1.2.3 Dopamine transport

Transport of dopamine across different cellular compartments is an integral process that contributes to dopamine homeostasis, compartmentalization and neurotransmission. Cellular transport of dopamine occurs at 2 levels: 1) the plasma membrane and 2) the vesicular membrane.

Figure 1-5. Dopamine transport in the presynaptic neuron Dopamine levels are modulated by 2 transporters: the dopamine transporter (DAT, shown in blue) on the plasma membrane and the vesicular monoamine transporter 2 (VMAT2, shown in green) on the vesicular membrane. Adapted from Rilstone et al., 2013, NEJM (Rilstone et al., 2013).

1.2.3.1 Plasma membrane transport

Transport of dopamine across the plasma membrane is mediated by the dopamine transporter (DAT, SLC6A3), a membrane protein located on dopaminergic cells. Similar to other monoamine transporters, DAT has 12 transmembrane domains with intracellular amino- and carboxyl- termini and belongs to the SLC6A family of Na+/Cl--dependent symporters (Gainetdinov and Caron, 2003). In particular, DAT couples the active transport of dopamine with the movement of one Cl- and two Na+ ions along the concentration gradient. This concentration gradient is created by the plasma membrane Na+/K+ ATPase and serves as the driving force for

15

DAT-mediated dopamine uptake (Kanner and Schuldiner, 1987; Gether et al., 2006). Dopamine translocation across the plasma membrane occurs as a result of conformational changes in DAT. The uptake cycle begins when DAT is open to the extracellular space in an outward facing state (Reith et al., 2015). In this conformation, Na+ and Cl- ions bind to DAT and prepare the transporter for dopamine binding. Upon binding of dopamine, the extracellular gate closes, generating an occluded DAT state. Importantly, dopamine binding induces a conformational change allowing the transporter to open on the cytosolic side. In this inward facing state, dopamine and the ions dissociate from DAT. Finally, the cycle is reset once DAT returns to the outward facing conformation (Reith et al., 2015).

The function of DAT is to rapidly transport dopamine from the extracellular space into the cytosol of the presynaptic neuron. At the plasma membrane, DAT is located peri-synaptically, where it removes extracellular dopamine and provides spatial and temporal control of the dopamine signal (Hersch et al., 1997; Jones et al., 1998a; Cragg and Rice, 2004). In dopaminergic brain regions such as the striatum, DAT provides the principal mechanism of clearing extracellular dopamine and terminating neurotransmission (Giros et al., 1996). Aside from modulating the dynamics of released dopamine, DAT is also responsible for recycling the neurotransmitter back into the dopaminergic cell, allowing it to be reused (Sotnikova et al., 2006). By loading the presynaptic neuron with dopamine, DAT directly contributes to the buildup of cytosolic dopamine and indirectly influences vesicular dopamine as well. Accumulation of cytosolic dopamine can produce neurotoxicity as discussed in the next section. Hence, DAT is a key player in dopamine compartmentalization that can have significant consequences for the presynaptic neuron. Collectively, DAT regulates the concentrations of both 1) extracellular dopamine at the synapse and 2) intracellular dopamine within the presynaptic neuron.

1.2.3.2 Vesicular membrane transport

Another transporter that plays an essential role in maintaining dopamine homeostasis and neurotransmission is the vesicular monoamine transporter 2 (VMAT2, SLC18A2) (Erickson and Eiden, 1993; Wimalasena, 2011). VMAT2 is a membrane protein that is expressed on synaptic vesicles of monoaminergic neurons. Structurally, it contains 12 transmembrane helices with cytosolic amino- and carboxyl- termini. VMAT2 is responsible for transporting intracellular

16 monoamines such as dopamine, norepinephrine, epinephrine, serotonin and histamine from the cytosolic space into synaptic vesicles. Synaptic vesicles are small spherical lipid bilayers that are approximately 40nm in diameter. These vesicles are filled with neurotransmitters at the nerve terminal where they are released through exocytosis upon stimulation of the cell. VMAT2 belongs to the SLC18 family of transporter proteins that also include VMAT1 and the vesicular transporter. VMAT1 (SLC18A1) is predominantly located in neuroendocrine cells of the peripheral nervous system including chromaffin cells of the adrenal gland and melatonin- synthesizing cells of the pineal gland (Lawal and Krantz, 2013). Conversely, VMAT2 is primarily expressed in monoaminergic neurons of the CNS as well as platelets, β pancreatic cells and histaminergic cells of the gastric mucosa (Peter et al., 1995).

The process of vesicular filling serves dual functions as it accumulates dopamine for eventual release and also controls buildup of cytosolic levels. It is estimated that vesicular concentrations of monoamines are 10,000 fold higher than cytoplasmic levels due to VMAT2 loading (Parsons, 2000). VMAT2 packages high concentrations of dopamine within small vesicles through active transport which relies heavily on the electrochemical gradient generated by the vesicular H+- ATPase. Using the energy from ATP hydrolysis, the vesicular H+-ATPase preferentially moves H+ ions into vesicles. This establishes an acidic environment (pH 5.5) within the vesicular lumen and creates proton and electrochemical gradients across the vesicular membrane that serve as an energy source for VMAT2 activity. Specifically, as a H+-antiporter, VMAT2 couples the uptake of each dopamine molecule with the expulsion of 2 protons from the vesicular lumen. Transport is initiated by the efflux of the first H+ ion from the vesicular lumen which alters the conformation of the transporter and enables binding of dopamine on the cytosolic side (Wimalasena, 2011). Following translocation of the second proton, the transporter undergoes a conformational switch to move dopamine from the cytosolic side to the luminal side (Wimalasena, 2011). This type of VMAT2 uptake cycle applies to other monoamines as well.

Several factors can influence vesicular uptake including: 1) magnitude of the transmembrane proton and electrochemical gradients, 2) cytoplasmic concentrations of neurotransmitter and 3) VMAT2 expression and activity (Wimalasena, 2011). Vesicular storage is a dynamic process because although VMAT2 actively loads dopamine into vesicles, the neurotransmitter also passively leaks through vesicular membrane back into the cytoplasm at a constant rate. It is estimated that 90% of leaked molecules are re-captured into vesicles by VMAT2 and the

17 remaining 10% persist in the cytosol where they can be degraded (Eisenhofer et al., 2004b). Hence, overall vesicular content of dopamine is determined by a balance of VMAT2-uptake and passive leakage. Appropriate vesicular storage is fundamental to extracellular as well as cytosolic dopamine dynamics. Quantal release of dopamine has been shown to be closely tied to the expression of VMAT2, where increased VMAT2 levels lead to larger vesicular stores and greater dopamine release and knock-down of VMAT2 translates to lower dopamine release (Caudle et al., 2007; Lohr et al., 2014). Aside from influencing extracellular dopamine levels, vesicular storage is also a crucial mechanism of maintaining low levels of cytosolic dopamine and protecting cells from dopamine-induced toxicity as discussed in subsequent sections.

1.2.4 Dopamine compartmentalization and its effects

Appropriate compartmentalization of dopamine is essential to neuronal homeostasis. At a cellular level, dopamine can exist in distinct compartments: 1) extracellular and 2) intracellular that is further divided into vesicular and cytosolic fractions. Movement of dopamine between these compartments is a dynamic process that is mediated by DAT and VMAT2. Notably, based on the compartment, dopamine produces different effects that have important consequences for the cell. Generally, extracellular dopamine is given most importance as it plays a pivotal role in dopamine signaling. However, intracellular dopamine, specifically the cytosolic portion, has been shown to influence neuronal health and potentially contribute to the vulnerability of dopaminergic cells.

1.2.4.1 Extracellular dopamine

Once dopamine is released into the extracellular space, it participates in neurotransmission by acting on specific receptors.

1.2.4.1.1 Dopamine Receptors

Dopamine receptors belong to the superfamily of G protein coupled receptors (GPCRs). GPCRs are membrane proteins containing 7 transmembrane domains, an extracellular amino terminal and intracellular carboxyl tail (Kobilka, 2007). GPCRs can exist and function as monomers or oligomeric complexes (Angers et al., 2002). These metabotropic receptors receive signals from

18 the extracellular environment and respond by activating intracellular signal transduction pathways. Notably, GPCRs are associated with a heterotrimeric G protein complex consisting of α, β and γ subunits. Whether this coupling occurs before or after ligand binding to the GPCR is a matter of controversy (Kobilka, 2007; Qin et al., 2011). Nonetheless, once the GPCR is activated by an , the receptor undergoes a conformational change that catalyzes the exchange of GDP for GTP on the Gα subunit. This activates the G protein, and according to the classical theory, causes Gα to dissociate from the receptor and the Gβγ dimer (Gilman, 1987; Digby et al., 2006). However, some studies also indicate that physical dissociation of subunits may not be necessary for signaling (Levitzki and Klein, 2002; Bunemann et al., 2003). Importantly, activated Gα and Gβγ subunits then bind different intracellular proteins and propagate the signal via second messengers. Specific signaling pathways are dependent on the type of G protein that the receptor is coupled to. Signal transmission can be terminated by the GTPase activity of Gα that hydrolyzes GTP to GDP and converts the receptor to an inactive conformation. In this state, G protein subunits may re-associate and bind to the GPCR once again. Regulators of G protein signaling (RGS) are proteins that can accelerate the GTPase activity of Gα, thus, encouraging G- protein inactivation and termination of downstream signaling pathways (Beaulieu and Gainetdinov, 2011).

There are at least 5 different types of dopamine receptors including D1, D2 (which exists in 2 isoforms; long and short), D3, D4 and D5. Classically, dopamine receptors are divided into two families, D1 and D2, based on their structure, sequence homology, pharmacology and most importantly, signaling properties (Kebabian and Calne, 1979). Typically, dopamine receptors signal through G-proteins that are associated with adenylyl cyclase, an enzyme that converts ATP to cyclic adenosine monophosphate (cAMP). cAMP is a second messenger that regulates proteins such as protein kinase A (PKA). When activated, PKA phosphorylates downstream targets including ion channels, CREB and DARPP-32 that can amplify the signal. The D1-like family, consisting of D1 and D5, signal through Gαs/olf to stimulate adenylyl cyclase and PKA activity. Conversely, the D2-like family, including D2 long, D2 short, D3 and D4, are coupled to

Gαi/o which inhibits adenylyl cyclase and reduces PKA activity. Generally, dopamine activation of D1 receptors produces a stimulatory effect whereas D2 receptors produce an inhibitory effect.

Aside from cAMP-mediated signaling, D1-like receptors can also engage Gαq which regulates phospholipase C (PLC) (Sahu et al., 2009; Medvedev et al., 2013). Upon activation, PLC leads

19 to synthesis of inositol trisphosphate (IP3) and diacylglycerol (DAG). These second messengers activate protein kinase C (PKC) and mobilize intracellular calcium stores, triggering a cascade of downstream effects. While traditionally, dopamine receptors function as GPCRs, accumulating evidence suggests that they also engage G protein-independent pathways, such as β-arrestin signaling. Studies demonstrate that β-arrestin 2 contributes to D2 receptor signaling by regulating the Akt/glycogen synthase kinase 3 (GSK-3) pathway (Beaulieu et al., 2005).

Dopamine receptors are expressed on both pre- and post-synaptic neurons. Pre-synaptic receptors on dopaminergic neurons allow these cells to regulate their own function through negative feedback mechanisms. In response to changes in extracellular dopamine, autoreceptors can adjust neuronal firing rate, dopamine synthesis and release accordingly (Missale et al., 1998). Autoreceptors are present along the dopaminergic neuron and therefore can respond to both terminal and somatodendritic dopamine release. Presynaptic dopamine receptors belong to the class of D2 receptors while D1 receptors are exclusively post-synaptic. Generally, D2 autoreceptors are activated by a lower concentration of dopamine than post-synaptic receptors, allowing for high sensitivity to extracellular dopamine levels (Elsworth and Roth, 1997). Stimulation of autoreceptors leads to reduction of neuronal firing, inhibition of dopamine synthesis and diminished release of dopamine. Taken together, these actions dampen extracellular dopamine signaling. With regards to the nigrostriatal pathway, D1 receptors are expressed on medium spiny neurons that project to the GPi and constitute the direct pathway of movement, while D2 receptors are expressed on striatal projections to the GPe which is the indirect pathway. Hence, extracellular dopamine promotes movement by stimulating the direct pathway via D1 and suppressing the indirect pathway via D2 receptors. Generally, activation of post-synaptic D1 receptors has a stimulatory effect on locomotion. However, effects of D2 receptors are more complex since they are expressed both pre-and post-synaptically. While activation of post-synaptic D2 receptors promotes locomotor activity, stimulation of D2 autoreceptors produces the opposite effect.

Although D1 and D2 receptors are typically divided into two distinct families, recent evidence suggests that their actions may be interconnected. When D1 and D2 receptors were co-expressed in the same cell, dual stimulation elevated intracellular calcium via a pathway that could not be activated by either receptor individually (Lee et al., 2004). These findings led to the discovery of D1-D2 heteromeric receptor complexes in the brain, specifically the striatum, that were found to

20

be coupled to Gαq/11 (So et al., 2005; Rashid et al., 2007). Through this signaling pathway, concurrent agonist binding to both receptors activates PLC and causes release of intracellular calcium, which then stimulates Ca2+/calmodulin-dependent protein kinase II, an important mediator of synaptic plasticity and learning. Blockade of D1 or D2 receptors with antagonists prevented this cascade, illustrating the necessity of both receptor types for rapid activation of the

Gαq/11 pathway (Rashid et al., 2007). Despite ongoing controversy regarding dopamine receptor heterodimerization, some studies suggests that these heteromers may play important roles in pathological conditions such as schizophrenia, and drug addiction (Grymek et al., 2009; Pei et al., 2010; Perreault et al., 2010; Hasbi et al., 2011).

1.2.4.2 Intracellular dopamine

Although extracellular dopamine serves important functions in dopamine signaling, the majority of synaptic dopamine is stored intracellularly within dopamine neurons. Intracellular dopamine is divided into two compartments: vesicular and cytosolic. Vesicular dopamine is a reflection of overall dopamine tissue content because at any given moment, most neurotransmitters are stored within vesicles. Dopamine is accumulated in vesicles for eventual release. In fact, vesicular dopamine has been shown to directly determine the amount of neurotransmitter released from a cell (Caudle et al., 2007; Lohr et al., 2014). Hence, vesicular dopamine not only represents the largest cellular repository of dopamine, it also impacts neurotransmitter signaling. The process of vesicular storage is dynamic and involves active uptake as well as passive leakage as discussed in the next section. Importantly, when dopamine is sequestered into vesicles, it is protected from metabolic reactions that can occur in the cytosol.

1.2.4.2.1 Cytosolic dopamine

Cytosolic dopamine represents a small fraction of presynaptic dopamine since the neurotransmitter is usually readily packaged into vesicles. However, there are multiple circumstances when dopamine can accumulate in the cytosolic space: 1) during synthesis, 2) following reuptake from the extracellular space, and 3) after vesicular leakage. In the cytoplasm, dopamine is exposed to various reactions that can propagate oxidative stress and potentially have damaging consequences for the dopaminergic cell.

21

Reactivity

Dopamine is a highly reactive molecule that can undergo enzymatic reactions or direct auto- oxidation. The predominant metabolic pathway of cytosolic dopamine involves deamination by the enzyme, MAO. This reaction gives rise to 2 products: 1) DOPAL, a volatile aldehyde and 2) hydrogen peroxide, a reactive oxygen species (ROS) (Stokes et al., 1999). If hydrogen peroxide is not rapidly eliminated by anti-oxidant pathways such as glutathione peroxidase, it can react with transition metals, such as iron to generate more reactive oxidants (Halliwell, 1992). In addition, the other product of dopamine metabolism, DOPAL has been shown to produce quinones and radical species, cause protein cross-linking and damage mitochondria (Kristal et al., 2001; Rees et al., 2009; Anderson et al., 2011). Moreover, in the presence of hydrogen peroxide, DOPAL generates highly reactive hydroxyl radicals that can cause further macromolecular damage (Burke et al., 2004). This demonstrates the synergistic and potentially harmful effects of these dopamine metabolites. In addition, emerging in vitro and in vivo evidence suggest that DOPAL is toxic to cells (Dauer et al., 2002). Injection of low doses of DOPAL in rat SN resulted in loss of TH staining, which is indicative of dopaminergic toxicity (Burke et al., 2003). One hypothesis also postulates that DOPAL contributes to the loss of dopamine-containing terminals in Parkinson’s disease (Goldstein et al., 2011, 2013). Indeed, post-mortem analysis of patients reveal increased DOPAL:DOPAC ratios, suggesting impaired detoxification of this reactive metabolite (Goldstein et al., 2011). Taken together, the normal degradation of cytosolic dopamine directly produces chemicals that can propagate oxidative stress in dopaminergic cells.

Aside from metabolic deamination, cytosolic dopamine is also prone to oxidation reactions. These reactions yield ROS (such as superoxide, hydroxyl radicals and hydrogen peroxide) and quinone products (Tse et al., 1976; Barzilai et al., 2001). In particular, the catechol ring of dopamine can undergo oxidation to produce electron deficient dopamine-quinones that are highly unstable. These chemical species readily react with nucleophilic sulfhydryl groups on free cysteine, glutathione or cysteinyl residues of proteins (Graham, 1978; Hastings and Zigmond, 1994; Stokes et al., 1999). Addition reactions between dopamine quinones and cysteine predominantly occur at the number 5 position of the catechol ring, leading to the formation of 5- S-cysteinyl-dopamine, an indicator of dopaminergic oxidative stress (Fornstedt et al., 1986). Conjugation between dopamine-derived quinones and cysteine residues on glutathione can

22 reduce the levels and effectiveness of this important antioxidant. Furthermore, given that cysteinyl residues often reside at the active site of proteins, covalent modification of these residues by dopamine-quinones can alter protein structure and inhibit normal function. Aside from dopamine, its catechol-containing precursor, L-DOPA, and metabolite, DOPAC, are also capable of forming quinones that bind cysteinyl residues (Fornstedt et al., 1986).

Oxidation of cytosolic dopamine can occur spontaneously or via enzymatic activation. Enzymes that catalyze the conversion of dopamine to dopamine-quinones include: 1) tyrosinase, which is involved in melanin formation, 2) prostaglandin H synthase, also known as cyclooxygenase, which possesses peroxidase activity and catalyzes prostaglandin production, 3) oxidase, which is involved in purine catabolism and generates ROS and 4) lipoxygenase, which mediates metabolism of eicosanoids (Korytowski et al., 1987; Rosei et al., 1994; Hastings, 1995; Foppoli et al., 1997; Gonçalves et al., 2009; Ramkissoon and Wells, 2011). Thus, diverse enzymatic pathways are capable of dopamine oxidation, highlighting the reactivity of this neurotransmitter. In fact, even in the absence of catalysts, dopamine can be directly autoxidized to produce superoxide anions and reactive quinones (Graham and Gutknecht, 1978; Barzilai et al., 2001).

Superoxide can be converted to H2O2 by superoxide dismutase or it can react with nitric oxide, to generate peroxynitrite, a highly reactive nitrogen species. In comparison to other catecholamines, the dopamine molecule displays the highest rate of autoxidation, suggesting that it is most likely to spontaneously form reactive quinones (Graham and Gutknecht, 1978). Conversely, dopamine exhibits the slowest rate of internal cyclization, a process by which the quinone reacts with its own side chain amine group producing leukochromes or aminochromes. This slow detoxification pathway prolongs the longevity of the dopamine quinone, allowing it access to react with sulfhydryl groups on cellular macromolecules and generate cysteinyl adducts (Graham, 1978). Indeed, rates of addition reactions between dopamine-quinones and external sulfhydryl groups on glutathione or cysteine residues is at least 3 fold higher than internal cyclization reactions (Tse et al., 1976). Moreover, cellular conditions such as the presence of transition metals (e.g. iron, copper or manganese) and a pro-oxidant background (e.g. hydroxyl radical) can accelerate dopamine autoxidation. In particular, dopaminergic neurons routinely produce hydrogen peroxide as a result of MAO-mediated metabolism. Through the Fenton reaction, H2O2 reacts with transition metals such as iron to produce reactive hydroxyl radicals that can greatly enhance dopamine oxidation rates (Nappi et al., 1995). Given that dopaminergic neurons are also an

23 abundant source of iron, the contribution of this transition metal in promoting dopamine oxidation can be significant (Halliwell and Gutteridget, 1984; Velez-Pardo et al., 1997). Furthermore, increased iron content was found in the substantia nigra of Parkinson’s disease patients versus control subjects, suggesting that iron-mediated production of reactive dopamine intermediates may impact disease pathogenesis (Sofic et al., 1988; Dexter et al., 1989b).

Figure 1-6. Generation of reactive oxygen species in dopamine cells Dopamine neurons are inherently prone to oxidative stress due to a variety of reactions that convert dopamine to ROS. (Top) Enzyme dependent and independent reactions of cytosolic dopamine (DA) lead to the production of cytotoxic molecules including superoxide anions (O2•), dopamine–quinone species (SQ•), and hydroxyl radicals (OH•). Hydrogen peroxide (H2O2), a by-product of dopamine metabolism by monoamine oxidase (MAO), can lead to formation of hydroxyl radicals via the Fenton reaction. Antioxidant systems include glutathione (GSH) peroxidase, glutathione reductase, superoxide dismutase and catalase. (Bottom) An imbalance between the production and elimination of ROS may propagate oxidative stress and render dopamine cells vulnerable to cell death. These pathways may play a role in the pathogenesis of neurodegenerative disorders such as Parkinson’s disease. Image adapted from Lotharius and Brundin, 2002.

24

In summary, when dopamine accumulates in the cytosolic space, it is available for a variety of reactions including metabolic deamination, enzymatic oxidation and autoxidation. Due to an unstable catechol ring, the structure of dopamine easily lends itself to the formation of radical species, quinones and reactive intermediates. These products can propagate oxidative stress and compromise normal cell function by binding and inactivating important macromolecules. As such, cytosolic dopamine inherently possesses the potential to cause toxicity in cells. In fact, one of the most toxic compounds known to damage dopaminergic neurons is 6-hydroxydopamine, a structural analog of dopamine that differs from the neurotransmitter only by a single hydroxyl group.

Toxicity

The theory that accumulation of cytosolic dopamine can be deleterious to neuronal function and survival is one that has been discussed for some time. In 1978, a study by Graham et al showed that dopamine was much more cytotoxic to neuroblastoma cells than the other catecholamines, epinephrine and norepinephrine (Graham and Gutknecht, 1978). This toxicity correlates with dopamine’s higher rate of autoxidation and lower rate of internal cyclization, strongly implicating oxidative damage and quinone formation as a mechanism for dopamine-induced cell death (Graham, 1978). Indeed, in a catecholaminergic cell line, application of dopamine produced signs of oxidative stress such as lipid peroxidation, DNA base damage and increased intracellular peroxides (Masserano et al., 1996, 2000). These changes were accompanied by dopamine-induced apoptosis that could be inhibited by: 1) catalase, an anti-oxidant enzyme that catalyzes the decomposition of H2O2, and 2) N-acetylcysteine, a precursor in the formation of the anti-oxidant, glutathione. Furthermore, a study by Lai et al confirmed that dopamine-induced cytotoxicity in catecholaminergic neuroblastoma cells could be reversed by application of various anti-oxidants such as glutathione, L-ascorbic acid and N-acetylcysteine or anti-oxidant enzymes such as catalase and superoxide dismutase (Lai and Yu, 1997). The connection between dopamine-induced toxicity and oxidative stress is further supported by a study showing that pretreatment with a glutathione-depleting compound, L-buthionine sulfoximine, exacerbates the detrimental effects of dopamine in a neuroblastoma cell line (Stokes et al., 2000). This indicates a circular relationship where dopamine reactivity can lead to oxidative stress, but a pro-oxidant background can also further promote dopaminergic toxicity.

25

A number of other studies lend additional support to the toxic potential of dopamine in vitro. Ziv et al proposed that physiologically-relevant concentrations of dopamine were capable of causing apoptosis-like cell death in cultured embryonic sympathetic neurons as shown by axonal degeneration, nuclear fragmentation and severe shrinkage of cell bodies (Ziv et al., 1994; Barzilai et al., 2001). In this system, dopamine was the most toxic in comparison to norepinephrine, epinephrine and serotonin, highlighting the inherent ability of this neurotransmitter to cause cellular damage (Zilkha-Falb et al., 1997). Similar results of dopamine-induced cell death were also observed in mesencephalic, dorsal root ganglion, cortical and striatal primary neuronal cultures (Michel and Hefti, 1990; Tanaka et al., 1991; Alagarsamy et al., 1997; McLaughlin et al., 2002). In a study on human neuroblastoma cells that selectively take up dopamine, treatment with the neurotransmitter produced changes in cell morphology, shrinkage, atrophy, accumulation of apoptotic particles and cell death (Simantov et al., 1996). However, when dopamine uptake was inhibited by the application of antisense DAT-specific oligonucleotides, it dose-dependently decreased the toxic effects of dopamine. This clearly suggests that dopamine uptake into the cytosol is necessary for dopaminergic toxicity (Simantov et al., 1996; Porat et al., 2001).

The most convincing evidence linking cellular damage and cytosolic dopamine was provided by Mosharov and colleagues in 2009. This research had some unique advantages: 1) it is the only study that directly measured cytoplasmic dopamine using a novel technique, intracellular patch electrochemistry and 2) it used cells that naturally handle dopamine: cultured cells from the ventral midbrain which mostly contain dopaminergic neurons (Mosharov et al., 2009). Initially, they treated these cells with increasing concentrations of L-DOPA, the precursor of dopamine, and reported corresponding increases in cytosolic dopamine levels, as expected. Importantly, they also showed that L-DOPA-induced surges in cytosolic dopamine were correlated with progressive loss of TH-positive dopaminergic neurons. Interestingly, dopaminergic neurons from the SN were more susceptible to toxicity than VTA neurons, as observed in Parkinson’s disease. This vulnerability was related to greater accumulation of cytosolic dopamine in SN versus VTA cells when treated with L-DOPA. Pharmacologic and genetic manipulations revealed a close correlation between cytoplasmic dopamine content and cell death. For instance, treatment with pargyline, an inhibitor of MAO, effectively terminated the metabolism of dopamine to DOPAC, thus increasing cytosolic dopamine levels and enhancing neuronal loss when treated with L-

26

DOPA. Conversely, blockade of dopamine synthesis using NSD-1015 or benserazide, inhibitors of DOPA decarboxylase, reduced cytosolic dopamine levels and prevented L-DOPA induced cell death. Genetic over-expression of VMAT2, enhanced vesicular uptake of dopamine and significantly decreased cytosolic dopamine levels. Consequently, this protected cells from L- DOPA mediated neurotoxicity. Taken together, these results convincingly demonstrate that cytosolic levels of dopamine can directly determine toxic outcomes in dopaminergic cells. Despite possessing the unique technical advantage of being able to measure cytosolic dopamine, it should be noted that in this study, cells needed to be preloaded with L-DOPA for cytosolic dopamine levels to reach the threshold for detection. This highlights the difficulty of measuring such a small and transient pool of dopamine even in neuronal cultures.

Expanding on the findings of in vitro studies, in vivo research has also validated the ability of dopamine to cause detrimental effects in intact biological systems. The simplest of these studies involve direct application of dopamine to the brain. In an extreme example, intracerebroventricular injection of dopamine in rats pretreated with pargyline, an MAO inhibitor, caused dose-dependent mortality of animals (Ben-Shachar et al., 1995). Conversely, in the absence of pargyline, dopamine-induced mortality was reduced, suggesting that blocking dopamine degradation allows the neurotransmitter to accumulate and propagate toxicity. In these rats, dopamine also inhibited activity of NADH dehydrogenase, a crucial enzyme involved in mitochondrial respiration. This reveals the ability of dopamine to modulate the function of the electron transport chain, which may play a role in its toxicity. Several studies have assessed the effects of dopamine injections in the striatum as it is rich in dopaminergic nerve terminals. In rats, intrastriatal application of dopamine produces apoptotic cell death and DNA damage while activating transcription factors that are responsive to oxidative stress (Hattori et al., 1998; Luo et al., 1999). These changes are proportional to the concentration of dopamine applied and the length of exposure, indicating dose-dependent effects of dopamine in triggering cellular toxicity (Hattori et al., 1998). In other studies, following intrastriatal injection of dopamine in rats, both free and protein-bound cysteinyl-dopamine and cysteinyl-DOPAC were markedly increased (22 to 37-fold), suggesting enhanced oxidation of the neurotransmitter and subsequent conjugation with cellular macromolecules (Hastings et al., 1996; Rabinovic et al., 2000). This was correlated with dose-dependent loss of TH-immunoreactivity and gliosis surrounding the injection site (Filloux and Townsend, 1993; Hastings et al., 1996). Furthermore, these authors also

27 demonstrated selective degeneration of dopaminergic terminals as evidenced by 1) reduction in the dopamine-specific marker, DAT which correlates with loss of TH and 2) amino-cupric silver staining showing neuronal degeneration specifically in the area of reduced TH signal (Rabinovic et al., 2000). This illustrates that even in dopaminergic neurons that are equipped with protective mechanisms to handle dopamine such as vesicular storage and degradation, an overload of the molecule can instigate cellular damage. Interestingly, extracellular dopamine reached peak levels within 30 minutes post-injection after which it rapidly declined, pointing to DAT-mediated uptake of dopamine as a mechanism of clearing extracellular levels and accumulating the neurotransmitter exclusively in dopaminergic cells (Rabinovic et al., 2000). Also, dopamine- induced oxidative modifications seemed to precede the earliest signs of degeneration suggesting that oxidative stress plays a causal role in dopaminergic toxicity (Rabinovic et al., 2000). This idea is further cemented by findings showing that the extent of dopamine-induced protein modification and neurodegeneration can be rescued by co-injection of the antioxidants glutathione and ascorbate, strongly implicating oxidative stress as the mechanism of cell death (Hastings et al., 1996).

Direct measurement of cytosolic dopamine levels has not yet been possible in vivo due to technical constraints. Despite this limitation, other indicators can be used to indirectly gauge cytosolic dopamine levels in vivo, such a metabolite to dopamine ratios, presence of cysteinyl- dopamine adducts and activity of dopamine transporters. To study the effects of cytosolic dopamine in vivo, Chen et al generated transgenic mice with ectopic and inducible DAT expression in the forebrain (Chen et al., 2008). In these mice, GABAergic striatal neurons were engineered to take up extracellular dopamine released from nigrostriatal dopaminergic terminals. Since striatal neurons lack the regulatory mechanisms to effectively sequester and metabolize dopamine, once taken up, the neurotransmitter accumulates in the cytoplasm. These mice showed significant signs of oxidative stress in the striatum as demonstrated by dramatic increases in cysteinyl dopamine and cysteinyl DOPAC and reduction of glutathione content. Oxidative modifications were accompanied by progressive neurodegeneration in the striatum and impaired locomotor activity. In order to assess whether the amount of cytosolic dopamine could impact cellular and behavioral outcomes in these mice, dopamine supply was enhanced through L-DOPA treatment. L-DOPA accelerated neurodegeneration, exacerbated loss of body weight and further deteriorated motor function. Conversely, when unilateral 6-hydroxydopamine lesions

28 to the medial forebrain bundle were used to reduce dopaminergic input to the striatum, motor dysfunction was attenuated. These results show that exposure to dopamine is the single determining factor in producing neurodegeneration and motor disability in these animals. The damaging effects of cytosolic dopamine are evident in these engineered striatal cells since they lack the protective mechanisms to appropriately sequester and manage the neurotransmitter. Various other animal models have been developed with altered levels of dopamine transporters such as VMAT2-knockdown and DAT-overexpressing transgenic mice (Caudle et al., 2007; Salahpour et al., 2008). By modifying the transport and compartmentalization of dopamine, the neurotransmitter can be forced to accumulate within the cytosolic compartment. While published results from these studies are discussed in subsequent sections, novel findings are also characterized in this thesis.

In summary, a substantial body of evidence suggests that cytosolic dopamine is highly reactive and can cause toxicity in vitro and in vivo. However, most of these studies possess important caveats: 1) they exogenously apply non-physiological concentrations of dopamine or its precursor, and 2) they use systems that are not intrinsically equipped to handle dopamine. Thus, it is unclear whether the findings from these studies are relevant to physiological conditions of dopaminergic neurons. The primary aim of this thesis is to address these limitations and investigate the effects of altered dopamine transport and potential cytosolic dopamine accumulation in vivo.

1.2.5 Classical drugs that interact with the dopamine system

Since dopamine neurotransmission plays important roles in brain function and is also implicated in various disease conditions, the dopamine system serves as an important pharmacological target. Various drugs can modify dopamine synthesis, degradation, transport or receptor function, to ultimately impact dopaminergic signaling and related behaviors. Dopaminergic drugs are used therapeutically, recreationally or as research tools to manipulate and investigate dopaminergic function in vitro and in vivo. The following section describes selected classical drugs that affect the dopamine system by interacting with enzymes, transporters or receptors.

29

1.2.5.1 Enzyme ligands

Dopamine synthesis and degradation are fully dependent upon the activity of key enzymes. Different drugs can bind to these enzymes and alter their function, thus directly affecting dopamine homeostasis. The initial step in dopamine production in the conversion of tyrosine to L-DOPA by the rate-limiting enzyme, TH. A well-known inhibitor of TH activity is the compound, α-methyl-para-tyrosine (AMPT) (Brogden et al., 1981). As a structural relative of tyrosine, AMPT competes for the substrate binding site and blocks enzymatic activity, thus preventing dopamine synthesis (Bloemen et al., 2008). Since TH is also involved in norepinephrine production, application of AMPT results in depletion of dopamine, norepinephrine and their metabolites. Therapeutically, AMPT is approved for clinical use in pheochromocytoma, a rare catecholamine-secreting tumor in the adrenal gland that leads to hypertensive crisis (Brogden et al., 1981). In healthy humans, an acute dose of AMPT has been reported to produce slightly negative mood, anxiety, sleepiness and decreased attention and alertness (Bloemen et al., 2008). For research purposes, AMPT is often used as a pharmacological challenge to evaluate the consequence of catecholamine depletion on outcomes of interest.

The second enzyme involved in dopamine anabolism is DOPA decarboxylase. Two commonly used inhibitors of DOPA decarboxylase are benserazide and carbidopa. Both these drugs act peripherally as they are unable to cross the blood brain barrier. Hence, these drugs block the conversion of L-DOPA to dopamine in the body without affecting the brain. Clinically, they are used as adjunctive therapy in combination with L-DOPA for the management of Parkinson’s disease (Birkmayer and Hornykiewicz, 1961). Parkinson’s disease is characterized by reduced dopaminergic transmission. Therefore, to replenish dopamine levels, its precursor, L-DOPA, is applied since dopamine itself cannot cross the blood brain barrier and produces peripheral side effects. Normally, exogenously administered L-DOPA would be converted to dopamine in the periphery before it reaches the brain. However, by co-administering benserazide or carbidopa, peripheral decarboxylation of L-DOPA can be blocked, allowing the dopamine precursor to enter the CNS where it can be converted to dopamine and exert its therapeutic effect. Further details on Parkinson’s disease and its therapy are included in a later section. It should be noted that DOPA decarboxylase is generally known as aromatic L-amino acid decarboxylase (AADC) because it catalyzes several decarboxylation reactions and is involved in serotonin synthesis as

30 well. Thus, in addition to inhibiting dopamine production, benserazide and carbidopa can also prevent the decarboxylation of 5-hydroxytryptophan to serotonin in the periphery.

The major pathways responsible for dopamine degradation involve the enzyme, MAO. MAO not only catalyzes the inactivation of monoamines, but is also involved in metabolizing exogenous compounds, such as MPTP. In humans, MAO is found within the brain and in the periphery and exists in two forms: MAO-A and MAO-B. MAO-B mainly degrades dopamine and phenethylamine, a trace amine, while MAO-A also metabolizes serotonin and norepinephrine (Glover et al., 1977; Di Monte et al., 1996). Drugs that target MAO can be non-specific or selective for a particular isoform. Non-selective and irreversible MAO inhibitors such as isocarboxacid, phenelzine and are clinically used as antidepressants and anxiolytics. Since dopamine, norepinephrine and serotonin signaling contribute to normal mood, by blocking their metabolism, MAO inhibitors elevate levels of these neurotransmitters. Selective MAO-B inhibitors include selegiline (also known as deprenyl), and pargyline. Selegiline and rasagiline are used in the treatment of Parkinson’s disease because they specifically increase dopamine levels by blocking MAO-B-mediated metabolism. At higher doses, some of these drugs lose their selectivity and also bind to MAO-A. Selegiline can also be used to combat acute MPTP exposure (Przedborski et al., 2001). MPTP is converted to its toxic metabolite, MPP+ by MAO-B. Therefore, administration of selegiline prevents formation of the active metabolite and reduces toxicity. A possible side effect of MAO inhibitor use is hypertension due to increased catecholamine levels.

1.2.5.2 DAT ligands

DAT is the primary target for many compounds including psychostimulants, medications and neurotoxicants (Miller et al., 1999b; Torres et al., 2003). Since uptake of dopamine is dependent on DAT, pharmacological manipulation of DAT can produce profound effects on dopamine neurotransmission. Two classical psychostimulants that operate by altering DAT function are cocaine and amphetamine. Cocaine binds to DAT and blocks the transport of dopamine from the extracellular space to the presynaptic neuron (Ritz et al., 1987). Cocaine is a competitive inhibitor of dopamine transport because its binding site overlaps with dopamine’s site of action, precluding the endogenous substrate from binding (Beuming et al., 2008). As a result, dopamine accumulates in the extracellular space where it can reinforce downstream signaling. Conversely,

31 (amphetamine, , MDMA) compete with dopamine to enter dopaminergic cells, acting as a substrate for DAT (Sulzer et al., 2005). Once inside the cell, amphetamine disrupts the proton gradient required for vesicular storage thus causing dopamine to leak from the vesicles into the cytoplasm (Sulzer et al., 1995). Ultimately, accumulation of cytosolic dopamine in combination with the actions of amphetamine on DAT cause a reversal of the transporter, resulting in efflux of intracellular dopamine into the extracellular space. This DAT-mediated release of dopamine produces a surge in dopamine signaling in response to amphetamine. While cocaine and amphetamine can also produce other effects in the CNS, it is the manipulation of DAT function that directly enhances dopamine neurotransmission and is thought to underlie the reinforcing properties of these psychostimulants (Donovan et al., 1999; Howell and Kimmel, 2008). In addition to enhancing extracellular dopamine levels, these psychostimulants have also been shown to activate phasic dopamine signaling events causing release of dopamine that contributes to drug reinforcement (Aragona et al., 2008; Wanat et al., 2009; Daberkow et al., 2013).

Along with these examples, several compounds can inhibit DAT function with varying levels of selectivity and potency. Initially, it was postulated that all DAT inhibitors would have cocaine- like and reinforcing properties (Ritz et al., 1987). However, over the past 10-15 years, accumulating evidence has challenged this notion, showing heterogeneity among DAT inhibitors (Schmitt et al., 2013). In fact, different compounds preferentially bind and stabilize distinct structural states of DAT. Typical DAT inhibitors such as cocaine and methylphenidate have been shown to stabilize the outward facing conformation and produce locomotor stimulation and behavioral reinforcement (Loland et al., 2007). However, atypical DAT inhibitors such as , and (GBR12909) tend to promote occluded/inward facing conformations (Schmitt et al., 2013). Interestingly, these compounds also lack cocaine-like behavioral effects and possess limited rewarding properties (Schmitt and Reith, 2011). Recently, ligands that bind to allosteric sites on DAT have been identified and shown to block dopamine uptake as well (Janowsky et al., 2016). Hence, it seems that the specific pharmacological profile of a drug and its behavioral effects are heavily dependent on how the drug interacts with DAT and which structural conformation is favored. In general, DAT antagonists lock the transporter in a particular structural state, preventing the conformational transitions that are required to shuttle dopamine across the plasma membrane (Reith et al., 2015). Taken together, these data exemplify

32

1) the diversity of DAT ligands and 2) the responsiveness of DAT to different types of pharmacological manipulation.

As a plasma membrane transporter, DAT also provides a gate of entry into dopaminergic cells. The most potent dopaminergic toxicants, 6-hydroxydopamine and MPP+, are substrates of DAT (Gainetdinov et al., 1997; Miller et al., 1999b; Schober, 2004). These compounds are used to mimic symptoms of Parkinson’s disease in animal models because they cause robust degeneration of dopaminergic cells. Since 6-hydroxydopamine is a structural analog of dopamine, it can hijack the DAT-mediated uptake mechanism to access dopamine cells. It should be noted that 6-hydroxydopamine is also a substrate for the norepinephrine transporter (NET) and thus, must be administered specifically to dopaminergic regions to exert selective toxicity. With regards to MPTP, after crossing the blood brain barrier, this compound is converted to its toxic metabolite MPP+ by glial MAO-B. MPP+ is specifically translocated into dopaminergic neurons by DAT (Javitch et al., 1985). Once these toxins accumulate in dopaminergic cells, they cause oxidative stress and mitochondrial dysfunction, culminating in neurotoxicity (Miller et al., 1999a; Simola et al., 2007; Abdulwahid Arif and Ahmad Khan, 2010). Hence, DAT provides a molecular gateway for toxicants to selectively access and damage dopaminergic cells.

Figure 1-7. Substrates for DAT cause selective damage to dopamine neurons. Mechanisms of MPTP and 6OHDA induced toxicity in dopamine cells. Image adapted from Rangel-Barajas et al., 2015.

33

In summary, various compounds produce significant effects in the brain as a result of their actions on DAT. Given this rich pharmacology, manipulation of DAT function occurs in different ways: 1) by inhibiting DAT and causing buildup of extracellular dopamine, 2) by reversing DAT and causing release of dopamine and 3) by acting as substrate of DAT and using the transporter to specifically access dopamine cells. Interestingly, compounds that increase DAT activity are currently lacking.

1.2.5.3 VMAT2 ligands

Vesicular storage of dopamine can be modified by drugs that act on VMAT2. Similar to DAT pharmacology, VMAT ligands include inhibitors, psychostimulants and toxins. However, since VMAT transports serotonin and norepinephrine in addition to dopamine, drugs that manipulate this transporter can produce more widespread effects.

Reserpine and are two well-established inhibitors of VMAT function. is a potent VMAT inhibitor that binds at, or very close to, the cytoplasmic monoamine binding site of VMAT. Reserpine acts on both VMAT1 and VMAT2 although it has a higher affinity for the latter. Reserpine-induced inhibition of VMAT is long-lasting as it is thought to bind irreversibly. By disrupting monoamine uptake and storage in vesicles, reserpine and other VMAT inhibitors can substantially diminish monoamine signaling. Previously, reserpine was used to treat hypertension since it reduces catecholamine signaling in the peripheral sympathetic nervous system (Freis, 1954). However, a reported side effect of reserpine treatment was depression due to depletion of monoamines in the CNS. Indeed, the monoamine hypothesis of depression was derived, at least in part, from the negative effect of reserpine on mood. In rats, reserpine administration is used to model Parkinson’s disease since it produces profound hypokinesia and rigidity (Colpaert, 1987). This highlights the crucial role of VMAT2 in maintaining appropriate dopaminergic tone for locomotion. In contrast to reserpine, tetrabenazine is relatively selective for VMAT2, has a shorter half-life and reversibly binds to a site that is distinct from the substrate binding site. Clinically, tetrabenazine is used for symptomatic control of hyperkinetic disorders such as Huntington’s disease (Paleacu, 2007). Its therapeutic effect in controlling involuntary movements is at least partially mediated by VMAT2 inhibition and the consequent dampening of monoaminergic, and particularly dopaminergic, transmission. Other drugs such as and structurally related compounds, have also been shown to inhibit VMAT2 by binding the same

34 site as tetrabenazine. Lobeline derivatives are suggested to reduce the addictive effects of methamphetamine by decreasing drug-induced dopamine release (Wilhelm et al., 2008; Nickell et al., 2010).

Aside from acting on plasma membrane transporters, psychostimulants like amphetamine can also influence the movement of substrates across the vesicular membrane. Although the precise mechanisms of vesicular involvement are controversial, evidence suggests multiple ways that amphetamine and its derivatives, methamphetamine and MDMA, can interact with VMAT2: 1) by acting as a substrate to gain access to the vesicular lumen, 2) by inhibiting dopamine uptake from the cytosol, 3) by dissipating the proton gradient that drives vesicular monoamine uptake and 4) by promoting efflux of transmitters from vesicles (Sulzer et al., 2005; Lawal and Krantz, 2013; Nickell et al., 2014). In addition, amphetamines have been shown to displace reserpine or tetrabenazine binding to VMAT2, indicating that they interact at overlapping sites on the transporter. However, since amphetamines are highly lipophilic, they can also potentially permeate membranes without the engagement of carriers. According to one hypothesis, since amphetamine is a weak base, once it enters the vesicle, it becomes protonated in the acidic environment of the vesicular lumen. Thus, amphetamine binds free protons and alkanizes the vesicular interior, disrupting the activity of the proton pump that is necessary for monoamine uptake. Another hypothesis suggests that since amphetamines are themselves transported by VMAT2, they cause dopamine release via a carrier-mediated exchange mechanism (Partilla et al., 2006). Regardless of the particular mechanism, administration of amphetamine increases cytosolic dopamine levels by 5-fold as measured by intracellular patch electrochemistry (Mosharov et al., 2003). These results demonstrate the ability of amphetamine to displace dopamine from vesicles into the cytosol. Following this increase in cytoplasmic dopamine, DAT activity is reversed leading to amphetamine-evoked dopamine release. Hence, amphetamine acts by not only engaging DAT, but also manipulating vesicular transport to deplete dopamine stores and redistribute the neurotransmitter to the cytosol. It should be noted that amphetamines also produce other pharmacological effects such as NET reversal and MAO inhibition, which can contribute to their mechanism of action.

Toxins can also act as ligands for VMAT2. Interestingly, the VMAT sequence shares close homology with bacterial toxin-extruding antiporters, suggesting that VMAT has evolved from proteins that function to protect the cell from exogenous compounds. Hence, it is not surprising

35 that VMAT2 also possesses a neuroprotective role in monoaminergic cells. Indeed, VMAT was identified on the basis of conferring resistance to MPP+ toxicity (Liu et al., 1992; Stern-Bach et al., 1992). MPP+, the active metabolite of MPTP, causes cell death by inhibiting complex I of the mitochondrial electron transport chain, disrupting energy production and generating ROS. MPP+ is not only a substrate of DAT, but it is also sequestered into vesicles by VMAT2. This prevents MPP+ from interacting with mitochondria and causing cellular damage. VMAT2 expression is closely correlated with the extent of MPTP-induced toxicity: mice with low VMAT2 levels are particularly vulnerable while those with high VMAT2 levels are protected (Gainetdinov et al., 1998; Mooslehner et al., 2001; Lohr et al., 2014, 2016). Besides MPP+, VMAT2 is also a target for environmental toxins such as organochlorine pesticides (e.g. heptachlor), structurally-related polychlorinated biphenyls and brominated flame retardants. These compounds have been reported to bind and inhibit VMAT2 function thus, reducing vesicular dopamine uptake and storage.

In summary, while the physiological role of VMAT2 is to package monoamines, this transporter also responds to drugs that have important consequences for dopamine compartmentalization and transmission. There exists a bidirectional relationship between VMAT2 and its ligands: certain drugs manipulate VMAT2 function and vesicular dopamine while other compounds are sequestered by VMAT2 for cellular protection.

1.2.5.4 Dopamine receptor ligands

The drugs described so far can impact synaptic dopamine levels by altering dopamine uptake, storage, synthesis or degradation. However, another large class of drugs acts on specialized dopamine receptors without directly modifying neurotransmitter levels. These ligands are classified according to: 1) the type of dopamine receptors they target and 2) their actions at the receptor. Dopamine receptors are typically divided into 2 families: D1 (includes D1 and D5) and D2 (includes D2, D3 and D4). While ligands generally show preference for one of the two families of dopamine receptors, they are rarely selective for a single receptor type within the same family.

Apomorphine is a non-selective dopamine receptor agonist that binds both receptor types but has higher affinity for D2. In humans, apomorphine has been used in the treatment of Parkinson’s disease, as it stimulates dopamine receptor signaling despite low extracellular dopamine levels in

36 patients (Deleu et al., 2004). However, apomorphine also has powerful emetic effects probably due to its actions on dopamine receptors in the chemoreceptor trigger zone. In rodents, administration of apomorphine produces stereotypy and climbing behavior, which is regarded as a readout of striatal dopamine receptor activation (Protais et al., 1976). D2 receptor include quinpirole, bromocriptine, carbergoline, pramipexole, lisuride, and others. Although these agents may also act at other receptors at higher doses, they show strong affinity for D2-like receptors. Clinically, many of these drugs are used as adjunctive therapy or even monotherapy in Parkinson’s disease since they can directly activate D2 receptors in the striatum to restore dopaminergic signaling even though presynaptic dopamine-releasing cells have degenerated (Hisahara and Shimohama, 2011). Specifically, D2 agonists suppress the activity of the indirect pathway, thus promoting locomotion. In comparison to L-DOPA, the traditional treatment for Parkinson’s disease, certain D2 agonists have much longer half-lives making them an attractive alternative. However, D2 receptors function as both: 1) presynaptic autoreceptors on dopamine cells and 2) post-synaptic receptors on striatal GABAergic neurons (De Mei et al., 2009). Depending on the dose, some D2 agonists such as apomorphine, have been reported to preferentially bind autoreceptors versus post-synaptic receptors (Skirboll et al., 1979). Autoreceptor activation dampens dopamine synthesis, release and neuronal firing, ultimately diminishing the dopamine signal, whereas post-synaptic receptor stimulation promotes dopamine signaling pathways and encourages motor behavior. Hence, some D2 agonists can produce paradoxical behavioral effects due to engagement of both types of D2 receptors (Skirboll et al., 1979).

Another important class of clinically relevant drugs that interact with the dopamine system are D2 antagonists used in the treatment of schizophrenia. According to the dopamine hypothesis of schizophrenia, positive symptoms such as hallucinations and delusions, are caused by over- activity of the mesolimbic dopaminergic pathway (Howes and Kapur, 2009). In fact, this hypothesis emerged from the discovery that 1) major drugs are D2 blockers and 2) clinical effectiveness of these drugs was directly correlated with their affinity for D2 receptors (Creese et al., 1976; Seeman et al., 1976; Howes and Kapur, 2009). The first generation of typical antipsychotic drugs such as , chlorpromazine and fluphenazine, were effective against psychosis by blocking post-synaptic D2-mediated transmission in the nucleus accumbens. However, these drugs also antagonize D2 signaling in other pathways that terminate

37 in the striatum and hypothalamus, producing adverse effects such as extrapyramidal symptoms and hyperprolactinemia, respectively. As such, second generation atypical antipsychotic drugs (clozapine, risperidone, olanzapine, ) were generated with a presumably ameliorated side effect profile. These drugs are also primarily D2 antagonists although they act at other receptors such as 5HT2A (serotonin 2A) as well. For research purposes, raclopride and spiperone are commonly used to bind and inhibit D2 receptors for radioligand binding and other experiments.

There are also selective ligands for D1-like receptors. D1 agonists include dihydrexidines and such as SKF81297. shows anti-parkinsonian effects in MPTP- treated monkeys, suggesting that stimulation of D1 signaling in the direct pathway can compensate for presynaptic dopaminergic damage (Taylor et al., 1991). SKF 81297 treatment in WT animals has been reported to produce stimulant-like effects such as hyperactivity and self- administration, showcasing the role of dopamine neurotransmission in motor behavior and reward (Weed and Woolverton, 1995). The synthetic compound, SCH23390 was the first selective D1 antagonist and has been a useful research tool (Bourne, 2006). Several D1 ligands also show affinity for D2-like receptors. For example, various typical and atypical antipsychotic drugs block D1 receptors in addition to D2 receptors.

1.2.6 Parkinson’s disease

Parkinson’s disease is the most common neurodegenerative movement disorder in humans that affects approximately 1% of the population over the age of 60 (de Lau and Breteler, 2006). It is a progressive disease and its incidence significantly increases with age (Dauer and Przedborski, 2003). The clinical features of Parkinson’s disease were described by James Parkinson in his “Essay on the Shaking Palsy” in 1817 (Parkinson, 1817). However, it took over a century to link this disease to the neurotransmitter, dopamine. Indeed, dopamine was first synthesized in 1910 and was identified in the mammalian brain in the late 1950s (Carlsson et al., 1958; Hornykiewicz, 1986). Even after its discovery, dopamine remained in the shadows of the other two popular catecholamines, norepinephrine and epinephrine, as an intermediary. A major breakthrough occurred in the 1960s when Ehringer and Hornykiewicz demonstrated that post- mortem brains of Parkinson’s disease patients showed almost a complete loss of dopamine in the caudate putamen (Ehringer and Hornykievicz, 1960). This link between Parkinson’s disease and

38 striatal dopamine depletion established the essential physiological role of this neurotransmitter in controlling motor function.

1.2.6.1 Symptoms

Overall, Parkinson’s disease is a progressive neurological disorder that affects multiple neurotransmitter systems and is often coupled with psychiatric, cognitive, sensory and autonomic symptoms. However, the cardinal symptom of the disease is the impairment of voluntary movement that is attributed to reduced dopaminergic tone in the basal ganglia (Dauer and Przedborski, 2003). This arises from the specific loss of dopaminergic neurons projecting from the SNpc to the striatum. Clinically, Parkinson’s disease is characterized by motor deficits including muscle rigidity, postural instability, impaired gait, resting tremor, bradykinesia (slowness of movement) and ultimately, akinesia (loss of movement) (Jankovic, 2008). Deficits in initiating and executing voluntary movements significantly impact the patient’s quality of life as everyday tasks become difficult to perform. Motor symptoms only become apparent when dopaminergic tone in the striatum is depleted by ~80% and ~60% of nigrostriatal dopamine neurons have degenerated. Since motor disability serves as the most robust clinical feature of the disease, typically, when patients are diagnosed, the majority of SNpc dopamine cells have already been lost. This reduces the therapeutic window for intervention and emphasizes the need to recognize other symptoms of Parkinson’s disease that may arise earlier.

Although clinical diagnosis of Parkinson’s disease relies on the presence of motor deficits, many patients also experience non-motor symptoms that substantially contribute to their disability. Even James Parkinson’s early description of the disease identifies non-motor symptoms (Parkinson, 1817). These include sleep abnormalities such as disrupted nocturnal sleep, excessive daytime somnolence and REM sleep behavioral disorder, which occurs in a third of patients (Schenck et al., 1996; Olson et al., 2000; Chaudhuri et al., 2006). Patients also report neuropsychiatric problems such as anhedonia, apathy, depression, anxiety and cognitive impairment. While depressive symptoms can partially be attributed to reaction upon diagnosis of the disease, it is believed that monoaminergic deficiency plays a prominent role (Poewe, 2008). The rate of dementia increases with older patients and is 6-times higher in those with Parkinson’s disease compared to healthy individuals (Emre, 2003; Chaudhuri et al., 2006). Also, a variety of symptoms are related to autonomic dysfunction such as orthostatic hypotension, sexual

39 impairment and constipation, one of the most common non-motor symptoms (Poewe, 2008). A prospective study on 7000 men showed that those with initial constipation were 3-times more likely to develop Parkinson’s disease after 10 years (Abbott et al., 2001). A particularly notable olfactory symptom is hyposmia, the reduced ability to detect and discriminate odors. Hyposmia affects up to 90% of patients and may be used as a preclinical marker for Parkinson’s disease (Chaudhuri et al., 2006). Although various non-motor symptoms are evident prior to motor impairment and diagnosis, others reveal themselves with disease progression. While dopamine depletion contributes to the development of motor disability, widespread pathology in other neurotransmitter systems are likely responsible for non-motor symptoms of Parkinson’s disease.

1.2.6.2 Pathology

The pathological hallmark of Parkinson’s disease is a loss of nigrostriatal dopaminergic neurons. Interestingly, dopamine neurons in the adjacent VTA are relatively spared, suggesting a specific vulnerability of SNpc dopamine cells. Degeneration of SNpc dopamine neurons leads to diminished dopaminergic innervation of the striatum, which disrupts the basal ganglia motor loop. Specifically, D1 receptors on GABAergic medium spiny neurons are not adequately stimulated while D2-expressing striatal neurons are not sufficiently inhibited by dopamine. The net effect of this nigrostriatal imbalance is to enhance the inhibitory output of the basal ganglia to the thalamus, thus subsequently reducing cortical activity which impedes voluntary movement. These changes eventually produce the motor deficits that characterize Parkinson’s disease. Studies report that the reduction of dopamine terminals in the striatum is greater than the loss of dopamine cell bodies in the SNpc (Bernheimer et al., 1973). This suggests that terminals are more sensitive to damage in Parkinson’s disease and cell body degeneration may occur in a “dying back” process that is initiated in the axon terminals (Dauer and Przedborski, 2003; Cheng et al., 2010). It is interesting to note that DAT is mostly present in dopaminergic terminals and serves as a gateway for dopamine and toxicants to enter the cytosol.

Another prominent pathological feature and diagnostic marker of Parkinson’s disease is the presence of neuronal cytoplasmic protein inclusions known as Lewy bodies. These inclusions were first described by Frederic Lewy in the early 1900s and are also found in other diseases such as Lewy body dementia (Holdorff, 2002). Histologically, typical Lewy bodies appear as eosinophilic spherical masses surrounded by a halo of radiating fibrils. They are composed of

40 proteins such as ubiquitin and most notably, α-synuclein. α-Synuclein is abundant in the brain and is predominantly found in neuron terminals where it acts as a synaptic modulator with chaperoning abilities (Souza et al., 2000). α-Synuclein knockout mice show deficits in vesicle mobilization and synaptic transmission, highlighting the physiological role of this protein in synaptic function (Cabin et al., 2002). Also, α-synuclein interacts with various cellular components such as SNARE complexes, chaperone proteins, tubulin and DAT (Norris et al., 2004; Burré et al., 2010). Natively, α-synuclein is proposed to exist in a soluble form as an unstructured monomer or tetramer (Bartels et al., 2011; Fauvet et al., 2012). However, in pathological conditions, the protein undergoes conformational changes and misfolding to aggregate into high molecular weight oligomers and insoluble fibrils that give rise to Lewy bodies (Spillantini et al., 1997; Norris et al., 2004). Lewy bodies have been proposed to contribute to disease pathogenesis in multiple ways: 1) abnormal aggregates of α-synuclein reduce availability of the normal protein and thus, disrupt its physiological effects and 2) intracellular proteinaceous inclusions directly obstruct cellular functioning leading to degeneration (Norris et al., 2004; Luk et al., 2012). In Parkinson’s disease, Lewy bodies are not only found in the remaining dopaminergic cells of the SNpc, but have also been detected in other neurotransmitter systems (noradrenergic, serotonergic and ) and other brain regions including the olfactory bulb, locus coeruleus (LC), raphe nucleus, dorsal nucleus of the vagus, pedunculopontine nucleus, hypothalamus, nucleus basalis, cerebral cortex and autonomic ganglia (Forno, 1996; Norris et al., 2004). Interestingly, a recent study has shown that α-synuclein fibrils travel from cell-to-cell, providing a mechanism for the spread of Lewy bodies throughout interconnected brain structures (Luk et al., 2012).

Similar to Lewy body pathology, neurodegeneration in Parkinson’s disease is also not limited to the SN. While the extent of nigrostriatal degeneration is dramatic and undeniably responsible for motor deficiency in Parkinson’s disease, neuronal loss also extends to other brain regions. Degeneration of noradrenergic cells of the LC is comparable to the SN and tends to precede dopaminergic degeneration (Ehringer and Hornykievicz, 1960; Zarow et al., 2003). Noradrenergic loss is postulated to contribute to motor deficits as well since nigral dopamine neurons can receive noradrenergic input through α2-adrenergic receptors (Delaville et al., 2011). Furthermore, activation of the LC has been shown to alter firing of SN neurons (Grenhoff et al., 1993). In mice, loss of norepinephrine produced greater motor deficits than MPTP treatment

41 which causes specific damage to dopaminergic cells. This highlights a possible contribution of norepinephrine in controlling motor activity (Rommelfanger et al., 2007). The LC also plays important roles in cognition, circadian rhythm and mood. Thus, loss of these cells in Parkinson’s disease could give rise to non-motor symptoms such as cognitive impairment, sleep disorders, anxiety and depression (Gesi et al., 2000; Delaville et al., 2011; Del Tredici and Braak, 2013). Aside from catecholamine systems, degeneration has also been reported in other types of cells such as serotonergic cells of the raphe nucleus and cholinergic cells of the nucleus basalis and dorsal nucleus of the vagus (Jellinger, 1991; Bohnen and Albin, 2011). It is interesting to note, that in addition to cell loss, Lewy body pathology is evident in most, if not all, of these pathways as well. Importantly, these pathways connect to diverse brain regions and can account for the variety of functional deficits seen in Parkinson’s disease. For instance, loss of hippocampal structures and cholinergic cortical inputs have been reported to contribute to increased rates of dementia in older patients with Parkinson’s disease (Braak et al., 1996; Dauer and Przedborski, 2003). Also, serotonergic lesions are speculated to contribute to depressive symptoms in Parkinson’s disease.

In summary, Parkinson’s disease is a multi-system disorder with widespread pathology and a spectrum of symptoms. However, the cardinal diagnostic symptom of this disease is impairment of voluntary movement that is caused by degeneration of nigrostriatal dopamine neurons. Loss of SNpc dopamine cells is often coupled with degeneration and Lewy body pathology in extranigral regions as well.

1.2.6.3 Therapy

The symptoms of Parkinson’s disease are predominantly treated using dopamine replacement therapy. L-DOPA, the precursor of dopamine, is the gold standard pharmacological treatment. It is typically administered with AADC inhibitors (e.g. carbidopa or benserazide) to block peripheral metabolism and allow maximal amounts of the precursor to reach the brain, where it is converted to dopamine. The therapeutic effects of L-DOPA were revealed almost concurrently with the discovery that striatal dopamine levels were severely depleted in Parkinson’s disease patients (Hornykiewicz, 1986). In fact, the of a dopamine precursor in rescuing the motor deficits of Parkinson’s disease provided the final evidence for the involvement of striatal dopaminergic transmission in motor control (Hornykiewicz, 1986). Despite its popularity as the

42 treatment of choice since the 1960s when it was first introduced in Parkinson’s disease, L-DOPA administration has noteworthy limitations: 1) it only provides symptomatic control without addressing underlying pathology or disease progression 2) with chronic use, efficacy is often diminished, producing motor fluctuations and 3) long-term use can also lead to the development of adverse effects such as dyskinesias (Hornykiewicz, 1986; Marsden, 1994). Dopamine receptor agonists are also used in Parkinson’s disease predominantly as adjunctive therapy. When patients no longer respond adequately to pharmacotherapy, surgical approaches, namely deep brain stimulation, is used to manage late stages of the disease. In particular, high frequency deep brain simulation of target regions such as the subthalamic nucleus or globus pallidus internal, ameliorates the function of the basal ganglia motor loop in Parkinson’s disease (The Deep-Brain Stimulation for Parkinson’s Disease Study Group, 2001). However, none of these treatment approaches tackle the neurodegeneration that gives rise to motor symptoms in Parkinson’s disease. Furthermore, since it is a progressive disorder, the pathology is ongoing even when patients are on medication. Thus, it is important to attack the root cause of the disease to prevent or at least, decelerate the loss of nigrostriatal dopamine neurons. In order to accomplish this task, the cause of dopaminergic neurodegeneration in Parkinson’s disease needs to be better understood.

1.2.6.4 Etiology

While the pathological loss of nigrostriatal dopaminergic neurons is well-established in Parkinson’s disease, the etiology of this degeneration remains elusive in the majority of cases. (Surmeier et al., 2010). Approximately 90% of Parkinson’s disease is termed sporadic or idiopathic, without a known cause. Over the last decade, the role of genetics in Parkinson’s disease pathogenesis has been progressively explored. Using linkage and genome wide association analyses, different genetic variants have been identified that either cause familial forms of the disease or are associated with increased risk of developing “sporadic” Parkinson’s disease (Hardy et al., 2006; Shulman et al., 2011; Klein and Westenberger, 2012). The first locus shown to cause Parkinson’s disease was SNCA, the gene responsible for generating α-synuclein. Given the significant role of α-synuclein in Lewy bodies, a pathological marker of Parkinson’s disease, it is not surprising to find that some familial forms of the disorder are caused by dominantly-inherited mutations or multiple copies of the SNCA gene. Missense mutations in leucine-rich repeat kinase 2 (LRRK2) are also commonly associated with autosomal dominant

43 parkinsonism, whereas mutations in genes such as parkin, PTEN-induced kinase 1 (PINK1) and DJ-1 give rise to early-onset autosomal recessive forms of the disease (Nuytemans et al., 2010). Since exclusively monogenic cases of Parkinson’s disease are rare, the contribution of genetic polymorphisms as risk factors has also been considered using genome wide association analyses. Variants in SNCA, LRRK2 and β-glucocerebrosidase (GBA) genes have been associated with an increased susceptibility of developing Parkinson’s disease, although these polymorphisms have also been detected in asymptomatic individuals (Shulman et al., 2011). Given the presence of genetic heterogeneity and variants with incomplete penetrance, genetic mutations only account for 5-10% of Parkinson’s disease cases (Dauer and Przedborski, 2003; Hardy et al., 2006; Shulman et al., 2011).

In addition to genetic influence, there is also substantial evidence for an environmental component in Parkinson’s disease. In particular, the landmark discovery of MPTP (1-methyl-4- phenyl-1,2,3,6-tetrahydropyridine) in the 1980s brought to the forefront a role of exogenous toxicants in Parkinson’s disease pathogenesis (Langston et al., 1983; Schober, 2004). MPTP was accidentally generated as a by-product in an illegal attempt to synthesize MPPP, an drug. After using MPTP-contaminated drugs intravenously, young adults developed rapid-onset, irreversible and chronic parkinsonism. Due to their dramatic symptoms and complete inability to move, they were referred to as the “frozen addicts”. Post-mortem analyses revealed specific damage to dopaminergic neurons of the substantia nigra, identical to advanced stage Parkinson’s disease (Langston et al., 1983). It was later discovered that the metabolite of MPTP, MPP+, is a substrate for DAT that selectively gains access to dopamine cells through the plasma membrane transporter. Once inside the cell, MPP+ disrupts complex I of the electron transport chain, inhibits energy production and exacerbates oxidative stress eventually leading to cell death. Since its discovery, MPTP has been frequently used to model Parkinson’s disease in animal research due to its ability to cause potent and selective toxicity in nigrostriatal dopamine neurons. Aside from MPTP, epidemiological and case-control studies have suggested that environmental conditions such as exposure to pesticides, residing in rural areas and drinking well water could be significant risk factors for developing Parkinson’s disease (Rajput et al., 1987; Semchuk et al., 1992; Priyadarshi et al., 2000; Tanner et al., 2011). Indeed, a meta-analysis of 19 distinct studies found that exposure to pesticides approximately doubled the risk of disease (Priyadarshi et al., 2000). Furthermore, a dose dependent relationship has been reported between

44 lifetime cumulative exposure to paraquat, a widely used herbicide, and susceptibility to Parkinson’s disease (Liou et al., 1997). Residential exposure to maneb (a fungicide) and paraquat, was reported to increase the risk of Parkinson’s disease by 75% in a Californian study (Costello et al., 2009). Chronic treatment of animal models with pesticides such as rotenone was shown to recapitulate many of the fundamental features of Parkinson’s disease, including selective nigrostriatal neuronal degeneration, hypokinesia, and cytoplasmic Lewy body-like inclusions (Greenamyre et al., 2000; Alam and Schmidt, 2002; Cicchetti et al., 2009). Together, these studies highlight the significance of environmental risk factors in the etiology of Parkinson’s disease.

In summary, although the symptomatology and pathology of Parkinson’s disease are well- elucidated, the precise cause of this disorder remains unknown in the majority of cases. It is a complex and multifactorial disease that is influenced by age, genetics and the environment.

1.2.6.5 Vulnerability of nigrostriatal dopaminergic cells

Despite recent advances in identifying factors that increase the risk of developing Parkinson’s disease, the question remains as to why the specific population of nigrostriatal dopaminergic neurons are the most susceptible to insult. Both genetic and environmental factors could theoretically have widespread implications on different neuronal populations of the CNS, however they tend to produce particularly selective effects. For instance, parkin encodes a protein involved in the ubiquitin-proteasome system that mediates protein degradation, however how mutations in this gene accounts for specific dopaminergic cell loss is unknown (Paris et al., 2009). Genetic studies also reveal that mutations in PINK1 are associated with early onset Parkinson’s disease suggesting that the loss of this mitochondrial protein kinase has very particular implications on the nigrostriatal tract of neurons in comparison to other cell types (Nuytemans et al., 2010). In the case of environmental toxicants, apart from MPTP, which is metabolized to form a substrate for DAT, other compounds associated with increasing Parkinson’s disease susceptibility do not display any precise characteristics that would target them to dopaminergic neurons only. For instance, rotenone is a highly lipophilic compound that can cross the blood brain barrier and inhibit complex I of the electron transport chain (Alam and Schmidt, 2002; McCormack et al., 2002; Richardson et al., 2005; Ramachandiran et al., 2007). Although rotenone can theoretically accumulate in the entire brain, it causes selective damage to

45 dopaminergic neurons of the SNpc. Indeed, it is interesting to note that a variety of genetic and environmental factors with diverse mechanisms of action, all seem to converge in damaging a small group of discrete neurons in the SNpc which gives rise to the symptoms of Parkinson’s disease. This indicates that apart from general risk factors, there must exist “cell-specific” factors that render nigrostriatal dopamine neurons highly sensitive to toxicity (Surmeier et al., 2010). In order to better understand the etiology of nigrostriatal neurodegeneration in Parkinson’s disease, it is important to uncover the inherent characteristics of these dopamine neurons that permit them to be easily targeted by genetic and exogenous insults.

As previously discussed, dopamine neurons of the SNpc display several unique features such as: 1) L-type calcium channel dependent pacemaking activity, 2) morphological complexity of axon terminals, 3) high bioenergetic demands, 4) increased basal oxidative stress and 5) an extremely reactive cytosolic substrate, dopamine. These characteristics may shape the intrinsic vulnerability of SNpc dopamine neurons to insult. Indeed, even healthy humans experience around a 40% reduction in nigrostriatal dopamine neurons between the ages of 40 and 60, illustrating the susceptibility of these cells to age-dependent degeneration (Bogerts et al., 1983; Stark and Pakkenberg, 2004; Chinta and Andersen, 2005). In Parkinson’s disease, the differential vulnerability of SNpc dopamine neurons could be mediated by the distinctive physiological nature of these cells. Unlike most neurons of the brain, adult SNpc dopamine cells rely on L-type calcium channels with a Cav1.3 subunit, to generate rhythmic action potentials (Guzman et al., 2009). A case-control study demonstrates that subjects prescribed L-type calcium channel blockers for the treatment of hypertension, were 27% less likely to develop Parkinson’s disease, indicating a neuroprotective effect (Ritz et al., 2010). Furthermore, in vitro and in vivo inhibition of L-type calcium channels protects SNpc dopamine cells from damage induced by rotenone and MPTP, two toxicants used to model Parkinson’s disease (Chan et al., 2007). These data suggest that the unique dependence of SNpc neurons on L-type calcium channels may contribute to their vulnerability.

In addition, nigrostriatal dopamine neurons display an extensive, unmyelinated axonal arbor that is orders of magnitude larger than other neurons (Bolam and Pissadaki, 2012). Together, autonomous pacemaking and massive axonal complexity impose high metabolic costs on these cells, rendering them particularly sensitive to any perturbation in mitochondrial energy production (Pissadaki and Bolam, 2013). Indeed, mitochondrial dysfunction is suggested to

46 participate in the mechanisms of toxicity underlying Parkinson’s disease (Keeney, 2006; Winklhofer and Haass, 2010). Post mortem analyses of Parkinson’s disease patients indicate a significant reduction of mitochondrial complex I in the substantia nigra (Mizuno et al., 1989; Schapira et al., 1990; Janetzky et al., 1994). Moreover, mitochondria derived from patients show increased oxidative damage of complex I catalytic subunits that correlate with reduced functionality (Keeney, 2006). Interestingly, several genetic and environmental risk factors associated with Parkinson’s disease, have also been shown to directly or indirectly affect mitochondrial integrity. The most obvious examples are rotenone and MPP+, two toxicants that have been implicated to cause Parkinson’s disease in humans and are commonly used to model the disorder in animals. Both compounds exert their toxicity by inhibiting complex I of the electron transport chain, reducing ATP production and enhancing generation of reactive oxygen and nitrogen species (Greenamyre et al., 2003; Winklhofer and Haass, 2010). The genes, PINK1 and parkin, that give rise to familial forms of Parkinson’s disease, have also been shown to regulate mitochondrial function (Clark et al., 2006; Deng et al., 2008; Poole et al., 2008). In summary, nigrostriatal dopamine neurons are heavily dependent on mitochondrial ATP production to meet their high energy demands and therefore, any interruption of mitochondrial activity seems to differentially affect these cells. Furthermore, accumulating evidence of impaired mitochondrial function in Parkinson’s disease suggests that it may participate in disease pathogenesis.

Although the precise mechanisms underlying nigrostriatal degeneration in Parkinson’s disease remain to be fully elucidated, oxidative stress has emerged as a crucial player (Dias et al., 2013). Oxidative stress is defined as an imbalance between the production of ROS and the ability of anti-oxidant mechanisms to detoxify these volatile chemicals. The resulting disequilibrium produces detrimental consequences as reactive oxygen and nitrogen species modify key cellular macromolecules, disrupt their function and eventually lead to cell death. Postmortem tissue from Parkinson’s disease patients demonstrate extensive oxidative and nitrosative injury to dopaminergic regions like the SN as indicated by: 1) increased protein carbonyls, 2) reduced levels of antioxidants: glutathione, glutathione peroxidase and catalase 3) elevated protein adducts, 4) enhanced lipid peroxidation and 5) increased oxidative modifications of DNA and RNA molecules (Dexter et al., 1989a; Sian et al., 1994; Yoritaka et al., 1996; Alam et al., 1997a, 1997b; Zhang et al., 1999; Asanuma et al., 2003). Moreover, almost all toxicants that are

47 associated with nigrostriatal damage and are used to model Parkinson’s disease – such as MPTP, 6-hydroxydopamine, rotenone and paraquat – induce oxidative stress as their mechanism of toxicity. These findings strongly suggest that nigrostriatal dopamine neurons are particularly sensitive to oxidative injury and may be exposed to increased basal levels of oxidative stress. As the primary cellular consumer of oxygen and armed with several redox enzymes, mitochondria represent a major source of ROS. While the electron transport chain transfers electrons onto molecular oxygen to generate superoxide anions, mitochondria also contain antioxidant defense systems to detoxify the ROS generated. However, in cases of mitochondrial dysfunction, as observed in Parkinson’s disease, antioxidant defense mechanisms become compromised while ROS production is exacerbated to give rise to oxidative stress (Lin and Beal, 2006; Yan et al., 2013). Perhaps the most notable source of ROS in nigrostriatal dopamine neurons is the endogenous neurotransmitter, dopamine itself. As previously mentioned, MAO-B-mediated degradation of dopamine routinely gives rise to H2O2. Also, TH-dependent synthesis of dopamine has been shown to catalyze production of ROS through hydroxylation reactions in vitro (Haavik et al., 1997). Furthermore, as a highly unstable molecule, dopamine is exposed to oxidation reactions in the cytosolic space, which contribute to generation of reactive quinones and radical species, as previously discussed. Thus, it is possible that the constant handling of cytosolic dopamine may render SNpc neurons particularly susceptible to oxidative stress and predispose them to degeneration in Parkinson’s disease.

1.2.6.5.1 Role of cytosolic dopamine in Parkinson’s disease

Accumulation of cytosolic dopamine has been shown to produce cellular toxicity in vitro and in vivo, as described in previous sections. Given the toxic potential of cytosolic dopamine and the differential loss of nigrostriatal dopaminergic neurons in Parkinson’s disease, it is possible that cytosolic dopamine may play a role in disease pathogenesis. Several findings indicate a link between cytosolic dopamine reactivity and Parkinson’s disease pathology. For instance cysteinyl adducts of dopamine, L-DOPA and DOPAC are significantly increased in the SN of Parkinson’s disease patients, demonstrating increased oxidation of cytosolic dopamine, its precursor and metabolite (Spencer et al., 1998). These cysteinyl conjugates have been shown to elevate ROS generation and enhance DNA base modification, causing neuronal damage (Spencer et al., 2002). Hence, in Parkinson’s disease, nigrostriatal neurons are exposed to dopamine-induced oxidative modifications that can potentially contribute to their degeneration. Dopamine turnover is also

48 increased in the brains of Parkinson’s disease patients (Goldstein et al., 2011, 2013). Since intracellular metabolism specifically occurs on the cytosolic fraction of dopamine, enhanced turnover may represent efforts to detoxify higher basal levels of cytosolic dopamine in Parkinson’s disease patients. While dopamine neurons of the SN undergo substantial degradation in Parkinson’s disease, those in the VTA are relatively spared. Although both subsets of neurons manage and transmit dopamine, nigral cells are more susceptible suggesting a unique physiological nature. Indeed, when treated with L-DOPA, studies reveal that SN neurons accumulate 2 to 3 times higher levels of cytosolic dopamine in comparison to their counterparts in the VTA (Mosharov et al., 2009). This clearly illustrates that not all dopamine neurons are created equal. Furthermore, it sheds light on the possibility that since SNpc neurons intrinsically handle higher quantities of cytosolic dopamine, they probably also experience greater oxidative stress and therefore, become more sensitive to insult in Parkinson’s disease versus VTA neurons. The continuous oxidative trauma present in SNpc dopamine cells may account for the exquisite vulnerability of these neurons to complex I inhibitors (e.g. rotenone, MPTP) or general inducers of oxidative stress (e.g. paraquat) in Parkinson’s disease (Tanner et al., 2011).

Although nigrostriatal damage is the most striking feature of Parkinson’s disease, other notable signs include the presence of α-synuclein filled Lewy bodies and considerable degeneration of noradrenergic neurons in the LC (Dauer and Przedborski, 2003). Interestingly, studies suggest that cytosolic dopamine levels can potentially impact these factors as well (Sulzer, 2001). Xu et al report that α-synuclein exhibits neurotoxic effects in dopaminergic neurons while providing neuroprotection in non-dopaminergic cells (Xu et al., 2002). This consolidates the role of cell- specific risk factors that are unique to dopaminergic neurons. Moreover, α-synuclein-induced apoptosis of dopaminergic neurons was shown to be dependent on the presence of dopamine, since blocking dopamine synthesis protected these cells from toxicity (Xu et al., 2002). In addition, α-synuclein-transfected dopamine neurons showed marked increases in ROS and application of antioxidants inhibited α-synuclein-induced cell death (Xu et al., 2002). This suggests that the toxic actions of dopamine and α-synuclein both converge at a common mechanism of generating oxidative stress which can lead to cell death. Another study demonstrates that dopamine can form oxidative adducts with α-synuclein (Conway et al., 2001). These reactive adducts stabilize the toxic protofibril form of α-synuclein, while inhibiting the formation of benign fibrils. As a result, in the presence of cytosolic dopamine and an oxidative

49 environment, α-synuclein assumes a pathogenic role (Conway et al., 2001). α-Synuclein is not only the main component of Lewy bodies, but has also been shown to cause familial forms of Parkinson’s disease. These results indicate that cytoplasmic dopamine concentration plays an important role in triggering the pathological accumulation of α-synuclein in Parkinson’s disease.

A significant loss of LC noradrenergic cells has also been detected in Parkinson’s disease. This deterioration is held responsible for some non-motor phenotypes such as REM sleep disturbances, a common early symptom of Parkinson’s disease. Interestingly, a shared attribute between noradrenergic and dopaminergic neurons is the molecule dopamine, which is the direct precursor to norepinephrine. In noradrenergic neurons, dopamine is converted to norepinephrine by dopamine β hydroxylase in the vesicular lumen. Hence, noradrenergic neurons also possess an intracellular pool of dopamine, like dopaminergic neurons. Thus, theoretically, the loss of noradrenergic neurons in Parkinson’s disease could be related to the fact that these cells also handle the highly unstable and reactive molecule, dopamine, that can instigate oxidative stress. Overall, the damaging effects of cytosolic dopamine can potentially contribute to various aspects of Parkinson’s disease pathology. Indeed, cytosolic dopamine-induced toxicity not only offers an explanation for the differential susceptibility of nigrostriatal dopamine neurons, but may also mediate α-synuclein pathology and noradrenergic cell loss in Parkinson’s disease.

1.2.6.5.2 Role of dopamine transporters in Parkinson’s disease

Accumulation of dopamine within the cytoplasm is controlled by two transporters: DAT and VMAT2. DAT increases the cytosolic pool of dopamine by taking it up from the extracellular space while VMAT2 reduces cytosolic accumulation by sequestering intracellular dopamine into vesicles. Since cytosolic dopamine is postulated to contribute to the vulnerability of nigrostriatal dopamine neurons in Parkinson’s disease, the balance of DAT and VMAT2 activity may also impact disease pathogenesis. Furthermore, these transporters are also targeted by various drugs including toxicants that have been implicated in Parkinson’s disease.

In general, the DAT protein sequence appears to be highly conserved, possibly as an evolutionary mechanism to preserve appropriate function of the dopamine system (Vandenbergh et al., 2000). The first genetic condition directly caused by loss-of-function mutations in the DAT gene is DAT deficiency syndrome, a complex motor disorder of progressive parkinsonism- dystonia that typically manifests in infancy and severely reduces life expectancy (Kurian et al.,

50

2009, 2011; Ng et al., 2014). The striking phenotypes in these patients convincingly demonstrate the significance of DAT genetics in controlling motor behavior. Unlike DAT deficiency syndrome, concrete evidence of a causal link between genetic DAT mutations and Parkinson’s disease is lacking. This is probably because the etiology of Parkinson’s disease is multifactorial and genetic mutations account for only a small proportion of cases (5-10%) (Dauer and Przedborski, 2003; Sulzer, 2007). However, several lines of evidence suggest that DAT may act as a risk factor in Parkinson’s disease. For instance, neuroanatomical analyses indicate that regions of the human brain containing the highest levels of DAT protein – the caudate and putamen – are most sensitive to damage in Parkinson’s disease (Bernheimer et al., 1973; Miller et al., 1997). The pattern of dopaminergic cell loss in the midbrain also appears to parallel the expression of DAT; nigral neurons display higher DAT mRNA than VTA neurons, which are relatively spared in Parkinson’s disease (Uhl et al., 1994). These findings indicate a correlation between DAT expression and vulnerability to insult in Parkinson’s disease. The potential role of DAT in enhancing susceptibility of dopamine neurons is two-fold: first, it functions to increase the pool of cytosolic dopamine, which is highly reactive and second, it allows toxicants such as MPTP selective access to dopaminergic cells. Hence, DAT activity could sensitize dopamine neurons to both intrinsic oxidative stress as well as extrinsic environmental insult. In fact, a study by Ritz et al demonstrates that DAT genetic variants in combination with pesticide exposure can increase risk of Parkinson’s disease by several fold (Ritz et al., 2009). These DAT variants include single nucleotide polymorphisms in the 5’ region as well as variable number tandem repeats at the 3’ region of the gene. Although the functional consequences of these DAT variants are unclear, these results highlight the synergistic influence of DAT in Parkinson’s disease especially in conjunction with environmental insults (Kelada et al., 2006; Sulzer, 2007). In a rare example, DAT coding variants were also identified in an individual with comorbid early-onset parkinsonism and ADHD (Hansen et al., 2014). In vitro, these variants resulted in reduced dopamine uptake capacity, indicating a role of DAT function in disease pathogenesis (Hansen et al., 2014). In summary, while DAT mutations typically produce drastic childhood-onset motor syndromes like DAT deficiency syndrome, in a progressive age-related disorder like Parkinson’s disease, DAT is more likely to play a modulatory role in combination with other risk factors.

Similar to DAT, mutations within the VMAT2 coding region are rare, consolidating the fundamental role of these transporters in neurotransmission. In a unique case, members of a

51 consanguineous family were discovered to possess a particular VMAT2 mutation that compromised vesicular transport of monoamines (Rilstone et al., 2013). These members suffered from infantile-onset parkinsonism with severe cognitive, autonomic and psychiatric disturbances, reflecting defects in monoamine transmission (Rilstone et al., 2013). Once again, similar to DAT, these rare VMAT2 mutations give rise to dramatic pediatric-onset movement disorders that confirm the necessity of VMAT2 function for appropriate motor behavior. Various findings suggest that vesicular function is involved in mechanisms leading to nigrostriatal degeneration in Parkinson’s disease. Vesicular uptake of dopamine and tetrabenazine (VMAT2 ligand) binding were both severely reduced in isolated synaptic vesicles from Parkinson’s disease patients, even after correcting for dopamine terminal loss (Pifl et al., 2014). This defect in VMAT2 function can impair vesicular dopamine storage, causing it to accumulate in the cytosolic space where it is exposed to oxidative reactions. Hence, reduced VMAT2 activity in Parkinsonian patients can potentially influence disease progression. Conversely, gain-of-function haplotypes in the VMAT2 promotor region were found to decrease the risk of Parkinson’s disease in females (Glatt et al., 2006). In another study, two specific polymorphisms in the VMAT2 promotor sequence were also found to confer a reduced risk of developing sporadic Parkinson’s disease (Brighina et al., 2013). These studies indicate that increased VMAT2 function is protective for Parkinson’s disease probably by 1) sequestering intracellular dopamine into vesicles and thereby, reducing cytosolic dopamine content and 2) isolating toxicants such as MPP+ away from cellular machinery. In addition, toxicants such as organochlorine pesticides and polybrominated biphenyl compounds that have been associated with Parkinson’s disease and detected in post-mortem brains of patients, have also been shown to inhibit VMAT2 activity and produce nigrostriatal damage (Bemis and Seegal, 2004; Richardson and Miller, 2004; Hatcher et al., 2008; Guillot and Miller, 2009; Cannon and Greenamyre, 2011; Hatcher-Martin et al., 2012; Bradner et al., 2013; Wilson et al., 2014). Hence, environmental toxicants may exert part of their damaging effects in dopaminergic cells by altering VMAT2 activity. In general, VMAT2 protects dopaminergic neurons from endogenous and exogenous insults and dysregulation of its function may contribute to the vulnerability of dopaminergic neurons in Parkinson’s disease.

52

1.2.7 Animal models with altered transporter levels

The function of DAT and VMAT2 in regulating dopamine dynamics is best elucidated by animal models with varying levels of these transporters. Over the past few years, several such mouse models have been generated allowing for controlled titration of transporter expression and elucidation of its effects on the dopaminergic system. Findings from these animal models and from this thesis, have been summarized in Table 1-1 (Lohr et al., 2017) .

53

Table 1-1. Summary of mouse models with genetically altered DAT or VMAT2 levels.

54

1.2.7.1 DAT-knockout mice

The critical role of DAT in maintaining appropriate dopaminergic function is clearly demonstrated by DAT knockout mice (DAT-KO). Genetic ablation of this plasma membrane transporter produces dramatic changes in extracellular and intracellular dopamine dynamics (Giros et al., 1996; Jaber et al., 1997; Jones et al., 1998a). DAT-KO mice display 5-fold elevated extracellular dopamine levels due to lack of uptake. Additionally, dopamine remains in the extracellular space 300 times longer since diffusion is the only mechanism to clear the neurotransmitter in DAT-KO mice. Conversely, intracellular dopamine content is reduced by 95% demonstrating that DAT-mediated recycling of dopamine is chiefly responsible for maintaining presynaptic dopamine levels (Sotnikova et al., 2005). Due to depleted intracellular stores, evoked dopamine release is also diminished by 75% in DAT-KO mice (Jones et al., 1998a). These neurochemical changes illustrate the vital role of DAT in balancing dopamine levels across different cellular compartments.

Furthermore, lack of DAT activity also triggers compensatory alterations in other pre- and post- synaptic markers of the dopamine system. Striatal post-synaptic D1 and D2 receptors are downregulated by 60% and 40% respectively, to adapt to high extracellular dopamine (Ghisi et al., 2009). Presynaptic D2 autoreceptors are also desensitized, disrupting regulatory negative feedback mechanisms (Giros et al., 1996; Jones et al., 1999). Levels of dopamine metabolites, HVA and 3-MT, are increased suggesting that dopamine degradation may be altered in these mice (Jones et al., 1998a). Without DAT-mediated dopamine recycling, presynaptic dopamine levels in DAT-KO mice are solely dependent on synthesis by TH. Paradoxically, while TH expression is reduced, dopamine synthesis rates are doubled, highlighting major adaptive changes in attempts to stabilize dopamine levels in DAT-KO mice (Jones et al., 1998a; Jaber et al., 1999). Behaviorally, these animals show spontaneous hyperlocomotion and impaired habituation as a result of increased extracellular dopamine (Giros et al., 1996). DAT-KO mice also display disturbances in cognition and sensorimotor gating (Ralph et al., 2001; Barr et al., 2003; Yamashita et al., 2006; Weiss et al., 2007). Pharmacologically, DAT-KO mice are insensitive to the classical stimulant actions of cocaine and amphetamine but show paradoxical calming effects instead (Giros et al., 1996; Jones et al., 1998b; Gainetdinov et al., 1999). In particular, when treated with cocaine or amphetamine, dopamine release and locomotor activity

55 are not enhanced in DAT-KO mice, validating that transporter function is compulsory for psychostimulant effects. Also, DAT-KO mice are completely resistant to nigrostriatal damage induced by MPTP demonstrating that DAT-mediated uptake of MPP+ is required for neurotoxic effects (Gainetdinov et al., 1997; Bezard et al., 1999). Collectively, mice lacking DAT show dramatic neurochemical, adaptive and behavioral changes in the dopamine system. Investigation of DAT-KO mice has contributed essential knowledge on the physiological role of DAT as well as the importance of this transporter as a pharmacological target.

1.2.7.2 DAT-overexpressing transgenic mice

On the other end of the spectrum, our laboratory has created transgenic mice that over-express DAT allowing for in vivo analysis of increased dopamine uptake. Specifically, DAT over- expressing transgenic (DAT-tg) mice were generated by pronuclear injection of a bacterial artificial chromosome (BAC) containing the 40-kb mouse DAT locus along with 80kb of flanking DNA sequences (Salahpour et al., 2008). Since the promotor region of DAT is not well- characterized, this approach allows for DAT transgene expression to be driven by the endogenous promotor. Hence, DAT is selectively over-expressed in dopaminergic neurons as confirmed by immunohistochemical data showing similar tissue localization of DAT between wild type (WT) and DAT-tg mice, although the extent of expression is higher in the latter as expected (Salahpour et al., 2008). Southern blots estimate that DAT-tg mice display 3-fold higher genomic DAT expression in comparison to WT mice. Since WT animals contain 2 endogenous copies of DAT, this suggests that DAT-tg mice possess a total of 6 DAT copies consisting of 2 endogenous and 4 transgenic copies. Similar to genomic levels, DAT-tg mice also display approximately 3-fold greater total DAT protein in the striatum. However, when DAT levels were assessed specifically in the synaptic plasma membrane fraction, the increase was much more modest (30%) suggesting that not all transgenic DAT is expressed at the plasma membrane (Salahpour et al., 2008). Consistent with this, the amount of functional DAT in DAT- tg mice was increased by 38% as measured by radioligand binding. This translates to approximately a 50% increase in the rate of dopamine uptake attesting to enhanced DAT activity in DAT-tg mice. Due to greater dopamine clearance, these mice display about a 40% reduction in extracellular dopamine levels (Salahpour et al., 2008). To compensate for this reduction in dopaminergic tone, DAT-tg mice also demonstrate a 30% increase in D1 and 60% increase in D2 receptors in the striatum (Ghisi et al., 2009). In response to dopamine receptor agonists such as

56 apomorphine (non-selective) or a combination of SKF 81297 (D1 agonist) and quinpirole (D2 agonist), DAT-tg mice demonstrate increased climbing behavior further supporting enhanced dopamine receptor function in these animals. Furthermore, when treated with amphetamine, DAT-tg mice show marked enhancement of dopamine release and concomitantly increased locomotor activity. These results indicate that DAT over-expression enhances the sensitivity of these animals to the psychostimulant effects of amphetamine (Salahpour et al., 2008).

While the effects of DAT over-expression on extracellular dopamine, post-synaptic receptors and psychostimulant response have been summarized in two manuscripts (Salahpour et al., 2008; Ghisi et al., 2009), its consequences on presynaptic dopamine dynamics were unclear when I began my doctoral thesis. Unpublished findings suggested that although increased DAT- mediated uptake is expected to enhance dopamine accumulation in the presynaptic neuron, both striatal dopamine tissue content and evoked dopamine release were reduced in DAT-tg mice. Furthermore, stereological counts of dopaminergic neurons in the SN and VTA revealed a 30- 40% loss in DAT-tg animals compared to WT mice. In light of a large body of literature suggesting that accumulation of cytosolic dopamine can produce neurotoxicity, we investigated whether increased DAT expression may lead to greater intracellular loading of dopamine and provide a mechanism for the neuronal loss in DAT-tg mice. The fine motor behavior of these mice was also assessed as a readout of nigrostriatal dopaminergic function. In addition, the sensitivity of DAT-tg mice to exogenous toxicant insult was evaluated by treating them with MPTP, a compound known to cause Parkinson’s disease. Further details on these experiments and their results are outlined in subsequent chapters and summarized in a manuscript (Masoud et al., 2015). In summary, DAT-tg mice represent a useful in vivo model to understand the consequences of increased dopamine uptake and probably, cytosolic dopamine accumulation, in neurons that routinely handle this neurotransmitter.

1.2.7.3 VMAT2-knockout homozygote mice

In addition to DAT, a series of studies have investigated the role of VMAT2 in vivo by varying levels of the transporter in genetically modified mice. The first and most extreme example of this is genetic ablation of VMAT2 expression in VMAT2-knockout (VMAT2-KO) mice (Fon et al., 1997; Takahashi et al., 1997; Wang et al., 1997). Strikingly, lack of VMAT2 results in postnatal death with most mice dying within 1-3 days after birth. VMAT2-KO mice also appear small,

57 feed poorly, are hypoactive and show severely stunted growth. These results emphasize the physiological necessity of normal VMAT2 function for survival and development. While immunohistochemical measures of dopaminergic cell bodies and projections appear normal, whole brain tissue content of dopamine, norepinephrine and serotonin are drastically reduced by over 95% in VMAT2-KO mice (Fon et al., 1997; Wang et al., 1997). Although rates of synthesis are almost doubled in these animals, it cannot compensate for the loss of vesicular storage, which is the main determinant of monoamine content in the brain. Despite the severe decline in monoamine levels, metabolite concentrations are unchanged or increased, suggesting enhanced monoamine degradation in VMAT2-KO mice. Since neurotransmitters are no longer protected in vesicles, they are extremely vulnerable to metabolic reactions in the cytosol. Furthermore, electrically-evoked dopamine efflux was completely abolished in striatal slices from VMAT2- KO animals, highlighting the critical role of vesicular loading in exocytotic neurotransmitter release. Hence, as a regulator of vesicular uptake, storage and release, VMAT2 activity directly impacts both intracellular and extracellular dopamine levels. Interestingly, treatment with amphetamine, enhances locomotion, feeding and survival of VMAT2-KO mice (Fon et al., 1997). Since amphetamine produces non-vesicular dopamine release, treatment with this drug circumvents the lack of VMAT2-mediated exocytotic neurotransmitter release, alleviating the severe symptoms in VMAT2-KO mice (Fon et al., 1997). Given that VMAT2 is involved in regulating all monoamines, effects on serotonergic and noradrenergic systems may also influence the phenotypes in these mice.

1.2.7.4 VMAT2-knockout heterozygote mice

Since homozygote VMAT2-KO mice survive for only a few days, detailed experimental analyses could not be conducted on these animals. Instead, heterozygote VMAT2-knockout (VMAT2-het) mice containing one functional VMAT2 allele were studied as they survive to adulthood. VMAT2-het mice show normal development and are indistinguishable from WT mice in appearance and locomotion (Takahashi et al., 1997). Western blots confirm that in comparison to WT animals, VMAT2-het mice show 50% VMAT2 protein expression, as expected. Although the reduction in VMAT2 produces alterations in dopamine homeostasis, the changes are less severe than homozygote VMAT2-KO mice. For instance, VMAT2-het mice display 50% lower dopamine uptake in striatal vesicular preparations, indicating a reduction in VMAT2-uptake activity that parallels the decrease in expression. Extracellular dopamine in the striatum is also

58 reduced by 40% in heterozygotes suggesting impaired VMAT2-mediated dopamine release. While there are conflicting reports regarding tissue levels, in general, striatal dopamine content seems to be reduced by approximately 25%, demonstrating diminished VMAT2 storage capacity in these mice (Wang et al., 1997). However, DOPAC levels were increased by 36%, suggesting an enhancement of dopamine turnover similar to VMAT2-KO mice. In addition to moderate changes in dopamine homeostasis, these mice display striking phenotypes in response to drugs that target the dopamine system. For example, VMAT2-het mice show pronounced hyperactivity when treated with psychostimulants such as cocaine or amphetamine in comparison to WT mice (Takahashi et al., 1997; Wang et al., 1997). However, despite the behavioral effect, amphetamine-induced dopamine release is diminished in VMAT2-het mice (Wang et al., 1997). A possible explanation of the locomotor sensitivity in these mice is post-synaptic receptor up- regulation, which may have developed to compensate for chronically lower extracellular dopamine levels. Indeed, pretreatment with either SCH23390 (D1 antagonist) or raclopride (D2 antagonist), prevented cocaine-induced hyperlocomotion in both WT and VMAT2-het mice, indicating that this response is driven by dopamine receptor function (Wang et al., 1997). In addition, administration of MPTP, a Parkinson’s disease-inducing neurotoxin, produced greater dopaminergic damage in VMAT2-het mice than WT animals as indicated by: 1) loss of nigral dopaminergic neurons, 2) reduction of striatal dopamine tissue content and 3) decreased DAT protein expression, a marker of dopaminergic nerve terminals (Takahashi et al., 1997; Gainetdinov et al., 1998). Enhanced susceptibility to toxicant insult demonstrates the protective role of vesicular transport in packaging exogenous compounds like MPP+ into vesicles to prevent their interaction with cellular machinery.

1.2.7.5 VMAT2-knockdown mice

Interestingly, in an attempt to knockout the VMAT2 gene, serendipitous recombination events gave rise to transgenic mice that express only 5% of normal VMAT2 levels (Mooslehner et al., 2001; Caudle et al., 2007). These VMAT2-knockdown (VMAT2-kd) mice possess a hypomorphic VMAT2 allele with insertion of the neomycin cassette in the third intron of the VMAT2 gene. Unlike VMAT2-KO mice, these animals are viable into adulthood, allowing for long-term assessment of vesicular deficiency. When these mice were first created by Mooslehner and colleagues, they unintentionally used an inbred strain of C57BL/6 mice that was later found to be lacking the α-synuclein gene locus (Mooslehner et al., 2001; Specht and Schoepfer, 2001).

59

Since α-synuclein is a ubiquitous and important protein that can interact with cytosolic dopamine and contribute to Parkinson’s disease, deletion of this gene limited the utility of these mice. Subsequently, the Miller laboratory at Emory University strategically bred mice that were heterozygous for the α-synuclein and VMAT2 genes to eliminate all traces of the α-synuclein mutation and generate VMAT2-kd mice on a normal α-synuclein background (Caudle et al., 2007). While most of these studies were performed on VMAT2-kd mice maintained on a C57BL/6 and 129SV mixed genetic background (Caudle et al., 2007; Guillot et al., 2008; Taylor et al., 2009, 2014), one recent manuscript and the work outlined in this thesis pertain to VMAT2- kd mice that were back-crossed to C57BL/6 for several generations (Lohr et al., 2016).

Even though 5% VMAT2 protein expression allows VMAT2-kd mice to survive into adulthood, they display prominent age-related neurochemical, compensatory and behavioral changes as well as altered response to toxicants. Concurrent with reduced VMAT2 expression, functional vesicular uptake of dopamine is also decreased by 80% in VMAT2-kd mice (Caudle et al., 2007). Tissue levels of dopamine, serotonin and norepinephrine are dramatically diminished (over 80%) throughout the brain as a result of depleted vesicular stores (Mooslehner et al., 2001; Caudle et al., 2007; Taylor et al., 2014). Furthermore, the reduction in striatal dopamine content is age-dependent with 6 and 12 month old animals showing progressively lower dopamine levels than 2 month old VMAT2-kd mice (Caudle et al., 2007). Reduced monoamine tissue content was accompanied by increased monoamine turnover as indicated by higher DOPAC/dopamine, HVA/dopamine, 5HIAA/5HT (serotonin) and DHPG/norepinephrine ratios in various brain regions (striatum, cortex, hippocampus), similar to VMAT2-KO and VMAT2-het mice (Mooslehner et al., 2001; Caudle et al., 2007; Taylor et al., 2009, 2014). In comparison to WT animals, stimulated dopamine release was also significantly lower in VMAT2-kd mice implying that deficient vesicular filling results in smaller quantal release of neurotransmitters that can dampen extracellular levels (Lohr et al., 2016). As a mechanism to compensate for reduced dopaminergic tone, VMAT2-kd mice show enhanced TH activity to upregulate dopamine synthesis (Caudle et al., 2007).

Presynaptically, a severe reduction in vesicular storage is expected to result in accumulation of dopamine in the cytoplasm, which can produce negative consequences for the cell. When the integrity of dopaminergic neurons was assessed in VMAT2-kd mice lacking the α-synuclein gene, no evidence of dopamine cell loss was found (Mooslehner et al., 2001). However,

60

VMAT2-deficient mice with normal α-synuclein expression showed progressive nigrostriatal degeneration (Caudle et al., 2007). Specifically, in comparison to WT mice, TH-positive cells of the SN were reduced by 12% and 26% in 18 and 24-month old VMAT2-kd mice, respectively (Caudle et al., 2007). However, neighboring dopamine neurons of the VTA were spared, as seen in Parkinson’s disease (Taylor et al., 2014). This suggests that cytosolic dopamine in combination with the presence of α-synuclein can exacerbate dopaminergic toxicity in the SN. Previous studies have shown that cytosolic dopamine can undergo oxidative modifications to interact with α-synuclein and stabilize the protofibril form of the protein which is neurotoxic (Conway et al., 2001). Indeed, aged VMAT2-kd mice show pathological accumulation of α- synuclein in the SN, which is a hallmark of Lewy body pathology in Parkinson’s disease (Caudle et al., 2007). Notably, markers of dopaminergic oxidative stress, cysteinyl-DOPA and cysteinyl- DOPAC are also increased in the striatum of VMAT2-kd mice prior to the onset of dopaminergic neurodegeneration. Taken together, these results demonstrate that vesicular deficiency can cause dopamine to buildup in the cytosolic space where it produces oxidative damage and eventually leads to loss of nigrostriatal dopamine cells. Behaviorally, VMAT2-kd mice show deficits in novelty-induced locomotion that are reversed by L-DOPA, the precursor of dopamine and principal treatment for Parkinson’s disease. Hence, this establishes that the motor deficiency in VMAT2-kd mice is due to reduced dopaminergic tone. Aside from basal dopaminergic toxicity, these mice are also particularly sensitive to the effects of the neurotoxin, MPTP. In particular, MPTP treatment produces 1) decreases in DAT levels, a marker of dopaminergic nerve terminals in the striatum and 2) loss of TH-immunopositive cells of the SNpc (Mooslehner et al., 2001; Lohr et al., 2016). MPTP-induced damage is exacerbated in VMAT2-kd mice compared to their WT littermates, showing that reduced VMAT2 expression enhances the vulnerability of dopaminergic cells to toxicant insult, as demonstrated by both VMAT2-kd and VMAT2-het mice. Thus, VMAT2 confers cellular protection by sequestering reactive cytosolic dopamine and exogenous toxins into vesicular compartments.

Since dopamine is not the only monoamine transported by VMAT2, VMAT2-kd mice also display phenotypes that relate to noradrenergic and serotonergic transmission. Preceding nigrostriatal cell loss, these mice show progressive noradrenergic neurodegeneration in the LC (Taylor et al., 2014). Interestingly, dopamine is generated in the cytosol of these neurons as well since it is the direct precursor of norepinephrine. In Parkinson’s disease, motor symptoms are

61 attributed to dopaminergic cell loss, while non-motor symptoms are often due to noradrenergic cell loss. Similarly, VMAT2-kd mice also show non-motor deficits such as progressively diminished olfactory discrimination, altered sleep latency and delayed gastrointestinal emptying (Taylor et al., 2009). In addition, these mice also display anxiety-like and depressive behavior as assessed on the elevated plus maze and forced swim test, respectively (Taylor et al., 2009). These phenotypes are a reflection of disrupted serotonergic, dopaminergic and noradrenergic transmission. In summary, given the crucial role of VMAT2 in regulating intracellular and extracellular monoamine concentrations, reduced vesicular storage leads to a variety of detrimental consequences in VMAT2-kd mice.

1.2.7.6 VMAT2-overexpressing mice

Taken together, the evidence from VMAT2-KO, VMAT2-het and VMAT2-kd mice highlight the adverse effects of decreased VMAT2 levels, suggesting that enhancing VMAT2 function can potentially be beneficial for dopaminergic cells. A BAC transgenic approach was used to generate mice that over-express VMAT2 in monoaminergic cells (VMAT2-OE) (Lohr et al., 2014). Physiologically, VMAT2-OE mice seem healthy and have normal body weight. As expected, these animals display increased VMAT2 expression in the nigrostriatal dopaminergic pathway, as well as serotonergic and noradrenergic cell bodies. Genomic quantitative PCR results suggest that VMAT2-OE mice have incorporated 3 copies of the BAC to possess a total of 5 copies of the VMAT2 gene. This translates to increases in VMAT2 mRNA (3.5 fold) and VMAT2 protein in striatal homogenates (3-fold) and vesicular fractions (3-fold). Functionally, VMAT2-OE mice display 2-fold higher vesicular dopamine uptake, 56% larger maximal vesicular capacity for dopamine and 33% greater dopamine vesicle volume in comparison to WT mice. These results demonstrate that increased VMAT2 expression is capable of inducing functional changes in dopamine vesicular transport and storage. Since the majority of dopamine in the brain is stored within vesicles, greater vesicular volume in VMAT2-OE mice is reflected in a 21% higher dopamine tissue content in the striatum, which is an indication of presynaptic dopamine stores. Moreover, enhanced vesicular efficiency also promotes dopamine neurotransmission in the striatum as evidenced by increased stimulated dopamine release (84%) in slices and greater extracellular dopamine levels (44%). Interestingly, other dopaminergic markers such as striatal DAT and TH protein expression, number of TH-positive cells in the midbrain and dopamine metabolite levels, remain unchanged in these mice, suggesting absence

62 of major compensatory modifications to the dopamine system. Behaviorally, these mice show 41% increased locomotor activity selectively in the active, dark cycle, indicating that elevated vesicular filling of dopamine can enhance motor behavior. Regarding tests of anxiety-like behaviors, VMAT2-OE mice portray no changes on the elevated plus maze, however they depict reduced basal anxiety on the marble burying task. In addition, these animals show reduced depressive-like behavior on the forced swim task in comparison to WT mice. In combination, VMAT2-OE mice portray improved outcomes on motor, anxiety and depression measures, unlike VMAT2-kd mice. Lastly, VMAT2-OE mice are also relatively resistant to MPTP-induced neurotoxicity since they showed smaller decreases in 1) TH and DAT protein expression in the striatum and 2) TH+ positive cell bodies in the SNpc, when compared to WT mice (Lohr et al., 2014, 2016). Thus, upregulation of VMAT2 protects dopaminergic cells from toxicant insult while reduced transporter expression is damaging as shown by VMAT2-het and VMAT2-kd mice. Together, the spectrum of mouse models with altered VMAT2 expression summarize the significance of vesicular storage in preserving the health of vulnerable nigrostriatal dopamine neurons and maintaining appropriate neurotransmission for monoaminergic behaviors.

1.3 Rationale, Hypothesis and Aims

As discussed, nigrostriatal dopamine neurons are inherently susceptible to a wide variety of insults. A cell-specific risk factor that may contribute to their intrinsic vulnerability is the highly reactive cytosolic neurotransmitter, dopamine. Studies show that buildup of dopamine in the cytosol can produce oxidative stress and deleterious consequences for the cell. However, these studies typically apply exogenous concentrations of dopamine or use non-dopaminergic systems that are not equipped to handle the neurotransmitter. To address these limitations, we propose to alter endogenous dopamine compartmentalization in genetically modified mice and investigate its effects on dopaminergic function. In particular, by enhancing dopamine uptake through DAT over-expression (Project 1) and reducing dopamine storage through VMAT2 knockdown (Project 2), we aim to increase cytosolic dopamine levels and investigate its outcomes on dopamine homeostasis, dopamine-related behaviors and response to dopaminergic drugs and toxicants. Based on previous findings indicating the toxic potential of cytosolic dopamine reactivity, our hypothesis is as follows:

63

Hypothesis: Genetic manipulations of transporter expression that potentially increase cytosolic dopamine levels, will lead to dopaminergic toxicity (e.g. loss of dopamine cells, altered dopamine homeostasis, oxidative stress, poor motor behavior).

This thesis has been separated into two projects:

Project 1. Aim: To evaluate the consequences of dopamine transporter (DAT) over- expression on the dopamine system of transgenic mice (DAT-tg) and assess their response to MPTP

Previous characterization of DAT-tg mice reveals enhanced uptake of dopamine and 36% reduction in the number of midbrain dopamine neurons. In this Project, we extended previous findings by investigating: markers of oxidative stress (as a potential mechanism of dopaminergic neurodegeneration), level of dopamine metabolism and motor behavior of DAT-tg mice. In addition, their sensitivity to exogenous insult was also evaluated by administering a Parkinson’s disease-inducing toxicant, MPTP.

Project 2. Aim: To investigate the dual effect of DAT over-expression and VMAT2 knock- down on the dopamine system of genetically modified mice

Adult DAT-tg mice display a 36% loss of dopamine neurons while aged (24 month old) VMAT2-kd mice also show comparable (26%) loss of nigrostriatal dopamine neurons. In this project, DAT-tg and VMAT2-kd mice were intercrossed to generate animals that would potentially accumulate more cytosolic dopamine and display greater toxicity than either genotype alone. Characterization of these mice included assessments of overall fitness, presynaptic and postsynaptic markers of the dopamine system, baseline behaviors and response to dopaminergic drugs.

64

Chapter 2 Materials and Methods

Materials and Methods

The following sections include all methods used for Projects 1 and 2. Some sections may apply exclusively to one project.

2.1 Mice

2.1.1 Generation of DAT-tg mice (Project 1)

Generation of DAT-tg mice using BAC transgenesis has been described in Salahpour et al., 2008. Briefly, transgenic animals were created by pronuclear injection of a BAC containing the DAT locus and 80kb of upstream and downstream genomic sequences. This approach was used since the promotor region of the DAT gene was not well characterized. The DNA was isolated from the BAC (obtained from Genome Sciences) and injected in pronuclei of C57BL/6J embryos at the Duke Transgenic Mouse Facility. Once a positive transgene founder was identified using PCR-based genotyping, it was further bred to generate the mouse colony. For most experiments, adult (3-5 months old) DAT-tg mice and their wild-type (WT) littermates were used, unless otherwise specified. Animals were age and sex-matched across groups. Animals were provided food and water ad libitum and maintained on a 12-hour light-dark cycle. Experiments were conducted in accordance with the Canadian Council on Animal Care and approved by the Faculty of Medicine Animal Care Committee at the University of Toronto.

2.1.2 Generation of DAT-tg/VMAT2-kd mice (Project 2)

DAT-tg/VMAT2-kd double transgenic mice were generated by interbreeding DAT-tg and VMAT2-kd mice. Generation of DAT-tg mice has been discussed above. VMAT2-kd mice were generously donated from our collaborator, Dr. Miller at Emory University. VMAT2-kd animals were generated using gene targeting as previously described (Caudle et al., 2007). Briefly, insertion of the neomycin cassette in the third intron of the VMAT2 gene results in hypomorphic mice that only show 5% of normal VMAT2 protein expression (Caudle et al., 2007). Initial generation of these mice occurred on an inbred strain of C57BL/6 mice that was later found to be

65 lacking the α-synuclein gene locus (Mooslehner et al., 2001; Specht and Schoepfer, 2001). Subsequently, the Miller laboratory at Emory University strategically bred mice that were heterozygous for the α-synuclein and VMAT2 genes to eliminate all traces of the α-synuclein mutation and generate VMAT2-kd mice on a normal α-synuclein background (Caudle et al., 2007). These VMAT2-kd mice were maintained on a C57BL/6 and 129SV mixed genetic background. In particular, DAT tg/VMAT2-kd mice were produced from two rounds of cross breeding. First DAT-tg mice (normal VMAT2) were crossed with VMAT2-kd (VMAT2 -/-) animals to produce DAT-tg/VMAT2-heterozygous mice (DAT-tg/VMAT2 +/-). These DAT- tg/VMAT2 +/- mice were then crossed with VMAT2 +/- (normal DAT, DAT-ntg) animals producing 6 possible genotypes:

DAT-ntg/ VMAT2 +/+ (WT, 12.5%), DAT-ntg/ VMAT2 -/- (VMAT2-kd, 12.5%), DAT-tg/ VMAT+/+ (DAT-tg, 12.5%), DAT-tg/VMAT2 -/- (DAT-tg/VMAT2-kd, 12.5%), DAT- tg/VMAT2 +/- (25%) and DAT-ntg/VMAT2+/- (25%) mice.

This breeding strategy yields the 4 necessary genotypes that are used for experiments: WT, DAT-tg, VMAT2-kd and DAT-tg/VMAT2-kd, indicating that littermates can serve as experimental controls. However, the probability of obtaining an animal of each genotype is only 1/8 or 12.5%. Collectively these mice are referred to as the DAT VMAT2 colony. Since the original VMAT2-kd mice were on a mixed background, mice were back-crossed to C57BL/6J for several generations to produce a mouse colony exclusively on the C57BL/6 background.

For most experiments, adult (2-4 month old) DAT VMAT2 mice were used, unless otherwise specified. Animals were age and sex-matched across groups. Animals were provided food and water ad libitum and maintained on a 12-hour light-dark cycle. Experiments were conducted in accordance with the Canadian Council on Animal Care and approved by the Faculty of Medicine Animal Care Committee at the University of Toronto.

2.1.3 Body weight

Adult DAT VMAT2 mice (2 to 4 months old) were weighed from all 4 genotypes: WT, DAT-tg, VMAT2-kd and DAT-tg/VMAT2-kd. The average age of animals from each genotype was calculated and matched across groups. Results were stratified according to sex since male mice tend to be larger than female mice.

66

2.1.4 Survival

Kaplan-Meier survival curves were generated using retrospective data from DAT VMAT2 animals born between 2012 and 2015. Survival data were assimilated from mice that were naturally found dead and mice that were intentionally sacrificed to conduct experiments. The age at which the animal was found dead or was sacrificed was noted. The analysis focused on the time frame between birth and 12 weeks of age, after which adult mice are usually sacrificed for experiments. Results were shown for both sexes combined and for each sex separately. Results were tabulated and analysed using GraphPad Prism 6.

2.2 Biochemistry

2.2.1 Western blots

Western blots were used to quantify expression of various proteins in striatal tissue. Western blots were performed as previously described (Masoud et al., 2015). The striatum was dissected and tissue was mechanically homogenized in RIPA buffer with protease inhibitors. For most proteins, samples were centrifuged at 15,000 rpm for 15 minutes and the supernatant was used to analyze protein concentration (BCA protein assay, Pierce). However, for VMAT2 and its corresponding loading control, GAPDH, striatal tissue was mechanically homogenized in 320mM sucrose, 5mM HEPES buffer with protease inhibitors. Homogenized samples were centrifuged at 3500rpm for 5 minutes and the supernatant was again centrifuged at 14,000rpm for 1 hour. The pellet was resuspended in homogenization buffer and used to analyze protein concentration (BCA protein assay, Pierce).

Protein extracts (20-30ug) were separated by 8.5 - 10% SDS/PAGE and transferred onto polyvinylidene difluoride (PVDF) membranes. Nonspecific binding was blocked using either 5- 7.5% milk, 5% BSA (specifically Na/K ATPase) or Rockland blocking buffer (specifically for DAT). Immunoblots were incubated overnight at 4°C with the following primary antibodies: rat anti-DAT (1:750, Millipore), rabbit anti-TH (1:3000, Millipore), rabbit anti-VMAT2 (1:20,000, obtained from Miller lab, Lohr et al., 2014), mouse anti-manganese superoxide dismutase (MnSOD, 1:1000, BD Transduction), goat anti-MAOb (1:1000, Santa Cruz), mouse anti- GAPDH (1:4000, Sigma), mouse anti-α-tubulin (1:2000, Hybridoma Bank) and rabbit anti- sodium/potassium ATPase (Na/K ATPase, 1:2000, Cell Signaling). Species appropriate

67 secondary antibodies (1:5000, Alexa Fluor 680 or IRDye 800CW, Rockland) were used and blots were developed using the LI-COR Odyssey Imaging System (LI-COR). Densitometric analysis of protein bands were performed using Image-J software (National Institutes of Health). Immunoblots of loading controls (GAPDH, α-tubulin, Na/K ATPase) were used to normalize protein loading across samples.

Protein carbonyl and 3-nitrotyrosine levels were also evaluated in DAT-tg mice (Project 1) using western blots. The striatum was dissected and synaptic plasma membrane (SPM) fractions were prepared at 4oC using protease inhibitors according to Salahpour et al., 2008. Briefly, striata from 3-4 mice were combined, homogenized in 4mM HEPES /0.32M sucrose buffer (pH7.4) and centrifuged at 900 x g. The resulting supernatant was centrifuged at 10,000 × g. The pellet was resuspended in 0.32 M sucrose/HEPES and lysed with water. Membranes were layered on a discontinuous sucrose gradient, ultracentrifuged at 200,000 x g (2hrs), and the 1.2M sucrose interphase was collected. The SPM fraction was added to 0.32M sucrose, centrifuged at 200,000 x g (30min), and the pellet was resuspended in 50mM HEPES/2mM EDTA solution. Protein concentration was determined using the BCA Protein Assay (Thermo Scientific). For protein carbonyl detection, the SPM samples (20ug) were further derivatized to 2,4- dinitrophenylhydrazone (DNP) by reaction with 2,4-dinitrophenylhydrazine (DNPH) according to the Oxyblot Protein Oxidation Detection Kit (Millipore). Western blots were used to quantify both DNP levels (from SPM derivatized samples) and 3-nitrotyrosine levels (from SPM samples). Proteins were separated by 10% SDS/PAGE and transferred onto PVDF membranes. Membranes were incubated with primary antibodies (rabbit anti-DNP, 1:300, Millipore or mouse anti-3-nitrotyrosine, 1:350, Abcam) and corresponding secondary antibodies (1:5000, Rockland). Immunostaining was developed using the LI-COR and quantified using Image-J. DNP levels were used as a measure of protein carbonylation.

2.2.2 Quantitative reverse transcriptase PCR

Quantitative reverse transcriptase PCR was used to determine mRNA expression of DAT and VMAT2 in DAT VMAT2 mice. Since these transporters are dopaminergic markers, mRNA was isolated from the midbrain which contains dopamine cell bodies. Brain regions were microdissected and homogenized in Tri-Reagent (BioShop) to isolate RNA. RNA isolation steps were performed as previously described (Rio et al., 2010). Briefly, homogenates are centrifuged and

68 chloroform is added to isolate RNA in the aqueous phase. The collected RNA phase is concentrated in a pellet and dissolved in DEPC water. RNA concentration was measured using optical density readings (260/280 nm). cDNA was constructed from RNA samples using SuperScript III Reverse Transcriptase according to manufacturer’s protocol (Invitrogen). Primer sets were generated for each gene of interest (including a housekeeping gene, phosphoglycerate kinase 1, PGK1) and verified for the presence of target transcripts using PCR. Finally, quantitative PCR was performed using sample cDNA, primers, SYBR Green Dye (Invitrogen) and the Applied Biosystems 7500 Real-Time PCR System. Relative expression of target genes was quantified using the ΔΔCt method (as described in Livak & Schmittgen, 2001) and normalized to PGK1 levels. Final results were reported as a ratio of WT expression.

2.2.3 Immunohistochemistry

Mice were anesthetized and intra-cardially perfused with 4% paraformaldehyde. Brains were removed, stored in 30% sucrose for at least 24 hours (cryoprotection) and sectioned to 50µm coronal sections using a Leica cryostat. Striatal sections were 1) quenched using 0.5% sodium borohydride, 2) rinsed, 3) blocked using 10% normal goat serum, 3% fish gelatin and 0.1% Triton X-100 and 4) incubated with primary rabbit anti-TH antibody (1:500, Millipore) overnight. Sections were then rinsed and incubated with the appropriate anti-rabbit secondary antibody for 1 hour (IRdye 800 or AF680 1:5000, Rockland Inc.). Sections were mounted on to slides and cover-slipped. Immunofluorescence was visualized using the LI-COR Odyssey Imaging System (LI-COR).

2.3 Neurochemistry

2.3.1 High performance liquid chromatography (HPLC)

HPLC with electrochemical detection (HPLC-EC) was used to measure dopamine, DOPAC and HVA levels in striatal tissue. Dissected striata were homogenized in 0.1M perchloric acid and centrifuged (9,400 x g for 10 min at 4oC). The supernatant was filtered through a 0.22µm membrane (Millipore). Samples were analyzed using a Hypersil Gold C18 column (150 x 3mm; 5µm; Thermo Scientific) and a LC-4C Amperometric Detector (BASi) set at an oxidizing potential of +0.75V. The mobile phase contained 24mM Na2HPO4, 3.6mM 1-octanesulfonic acid, 30mM citric acid, 0.14mM EDTA in 19% methanol, adjusted to pH 4.7 using concentrated NaOH. After the column was equilibrated with the mobile phase, appropriate electrochemical

69 separation of the following compounds was confirmed using standard solutions: dopamine, DOPAC, HVA, serotonin, 5-hydroxyindoleacetic acid (5-HIAA), a metabolite of serotonin, and 2,3-dihydroxybenzoic acid (DHBA), an internal standard that was added to every sample. Calibration curves were generated using increasing concentrations of dopamine, DOPAC, HVA and DHBA for quantification of these chemicals in brain tissue. Area under the curve was used to estimate concentration. Dopamine, DOPAC and HVA tissue content was normalized to DHBA levels. Metabolite-to-dopamine ratios were calculated by dividing metabolite tissue concentration by the tissue levels of dopamine for each animal.

5-S-Cysteinyl-dopamine and 5-S-cysteinyl-DOPAC were measured in collaboration with the Richardson lab. Since sensitivity has been an issue for HPLC analysis of cysteinyl adducts, we designed two positive controls that were expected to demonstrate enhancement of cysteinyl modified products. First, WT mice were treated with 5mg/kg of reserpine (i.p.), a VMAT2 inhibitor, and sacrificed 16 hours later when brain tissues were harvested. Other groups have shown that this reserpine regimen increases formation of cysteinyl DA by 135% (Fornstedt and Carlsson, 1989; Hatcher et al., 2007). Second, Caudle et al. report that VMAT2-kd mice display increased levels of cysteinyl L-DOPA and cysteinyl DOPAC, therefore, they were used as a second positive control (Caudle et al., 2007). Striatal tissue was dissected from positive controls, WT and DAT-tg mice and rapidly frozen in liquid nitrogen. Frozen striatal samples were shipped to the Richardson lab for HPLC-EC analysis that has been previously described elsewhere (Caudle et al., 2007; Hatcher et al., 2007). Briefly, samples were sonicated in 0.1M perchloric acid containing 347µM sodium bisulfite and 134µM EDTA. Homogenates were centrifuged, filtered and separated on a C18 column. The electrochemical detector was set at an oxidizing potential of +0.65V. The mobile phase was MD-TM (ESA) containing 2mM NaCl and adjusted to pH 2.1 using concentrated HCl. Quantification of all neurochemicals was conducted by referring to calibration curves constructed from pure standards (purity >98%; dopamine, DOPAC, HVA and DHBA from Sigma Aldrich; 5-S-cysteinyl-dopamine and 5-S-cysteinyl- DOPAC from NIMH Chemical Repository).

2.3.2 Fast-scan cyclic voltammetry (FSCV)

FSCV was performed in slice preparations of DAT VMAT2 mice to determine electrically- evoked dopamine release and uptake in the dorsal striatum. These studies were conducted in

70 collaboration with the Miller lab at Emory University as they possess the necessary equipment and expertise for FSCV. We sent live animals to the Miller lab and they performed FSCV according to previously described methods (Kile et al., 2012; Lohr et al., 2014). Briefly, mice were anesthetized, decapitated, and coronal slices (300µm) from the striatum were cut and maintained in cold artificial cerebral spinal fluid (pH 7.4, 95% O2 / 5% CO2). Recordings were performed in a slice perfusion chamber at 37oC (Warner Instruments). Dopamine release was electrically stimulated by biphasic (2 ms per phase) constant-current (350µA) pulses generated from a tungsten-bipolar electrode on the surface of the slice. The carbon-fiber detection microelectrode was placed 75-100µm into the slice and 100-200µm away from the stimulating electrode. Carbon-fiber microelectrodes were calibrated with dopamine standards. For each animal, four different sites were sampled in the dorsal striatum with 5-min intervals between stimulations. The waveform for dopamine detection consisted of a −0.4 V holding potential versus an Ag/AgCl reference electrode. The applied voltage ramp ranged from −0.4 V to 1.0 V. Dopamine release and uptake measures were extracted using nonlinear regression analysis. Data were analyzed using two redundant, yet different methods. The first simplistic method approximates “dopamine release” using the peak amplitude and “dopamine clearance” using the decay-time constant, tau (where lower tau, measured in seconds, implies faster clearance) (Yorgason et al., 2011). The second method, recently proposed by Hoffman and colleagues, uses curve modeling to determine release and uptake parameters (Hoffman et al., 2016).

2.4 Stereology

Stereological counts of dopaminergic neurons in the SNpc of DAT VMAT2 mice were performed in collaboration with 2 different laboratories (Miller and Brotchie) in 3 independent experiments to ascertain the results. Dopamine neurons were identified with a dopaminergic marker, TH and a neuronal marker, NeuN or Nissl. The stereological techniques used have been described in detail previously and are briefly summarized for each collaboration (Lohr et al., 2014, 2015; Taylor et al., 2014).

With the Miller lab at Emory University, we sent them whole brains that were transcardially perfused with 4% paraformaldehyde and cryoprotected in sucrose. Tissue was serially sectioned at a thickness of 40 μm (24 μm after staining/dehydrating) using a freezing sliding microtome. Coronal sections were stained according to Lohr et al, 2014 using rabbit-anti TH (AB152,

71

Millipore) and counterstained with 0.1% cresyl violet. Of all sections that contained the SNpc region, every 6th section was counted using the optical fractionator method (frames were 50 x 50 μm, counting grid was 120 x120 μm) in the Stereo Investigator software (MicroBrightField, Colchester, VT). This method is not affected by changes in the volume of the structure sampled. Boundaries of the SNpc were outlined under magnification of the 4× objective with reference to a mouse brain atlas (Franklin and Paxinos, 2012). Stereological counts were performed under a 40× objective using guard zones of 2 μm. Immunoreactive neurons were only counted if the recognizable profile came into focus within the counting frame. This method certifies a uniform, random and systematic cell count (Gundersen coefficients of error were less than 0.1).

With the Brotchie lab at Toronto Western Research Institute, we sent them perfused and sectioned tissue for staining and analysis. The mouse brain was perfused with 0.9% saline and heparin to clear out any blood in the tissue. The brain was then fixed with 4% paraformaldehyde and cryoprotected using serial sucrose concentration solutions (15% and 30%). Relevant tissue from the midbrain was sectioned using the cryostat (40 μm thickness) and collected in antifreeze solution. To perform immunohistochemical staining, endogenous peroxidase activity of the tissue was quenched and non-specific binding sites were blocked. Primary antibodies used: mouse anti-NeuN (Chemicon International, MAB377, 1:1000) and rabbit anti-TH (Chemicon International, AB152, 1:2000). Appropriate secondary antibodies were used. Vector DAB was used to stain NeuN positive neurons and Vector Blue alkaline phosphatase was used to immunolabel TH positive cells. Sections were mounted, clarified and cover-slipped for stereological counting using the StereoInvestigator software on equipment configured by MicroBrightField. For each brain, every 4th section from the SNpc was counted, producing a total of 6 sections/brain. TH and NeuN cells were counted simultaneously in brightfield. For each mouse, SNpc boundaries were delineated by closely tracing around the region of TH+ cells and excluding the SNpr and VTA, according to a mouse brain atlas (Franklin and Paxinos, 2012). Counting parameters were as follows: guard zone 2 μm, dissector height (Z) 20 μm, counting frame 175 x 175 μm, sampling grid 275 x 275 μm, section evaluation interval 4. The right SNpc was counted for every animal.

72

2.5 Radioligand binding

Radioligand binding was used to assess D1 and D2 receptor levels in the striatum as described previously (Ghisi et al., 2009). Striatal tissues were rapidly dissected and homogenized in lysis buffer (50 mM Tris–HCl (pH 7.4), 120mMNaCl, 1mMEDTA) containing protease inhibitors. The homogenate was centrifuged (1000 rpm for 10 min at 4 °C) to remove nuclei. The resulting supernatant was centrifuged (40,000 g for 20 min at 4 °C), the pellet was resuspended in lysis buffer and centrifuged again under the same conditions. The final pellet was resuspended in assay buffer (50 mM Tris–HCl (pH 7.4), 120 mM NaCl, 5 mM KCl, 2 mM CaCl2, 1 mM MgCl2). Protein concentration of membranes was determined using BCA protein assay (Pierce).

For D1 receptor saturation experiments, prepared striatal membranes (1.2 μg/μl, 50 μl) were incubated with [3H]-SCH23390, a D1 (50 μl, 16 nM) and (100 nM, 50 μl), a serotonin receptor antagonist, to prevent radioligand binding to these receptors. This reaction was performed in assay buffer (200 μl) at room temperature for 1 hour. In parallel reactions, nonspecific binding was measured using non-radiolabeled flupenthixol (10 μM), a dopamine receptor antagonist. For D2 receptor saturation experiments, prepared striatal membranes (0.5 μg/μl, 150 μl) were incubated with [3H]-spiperone, a D2 receptor antagonist (50 μl, 3 nM). This reaction was performed in assay buffer (250 μl) at room temperature for 2 hours. In parallel reactions, nonspecific binding was measured using non-radiolabeled haloperidol (6 μM), a D2 antagonist.

All reactions were terminated by filtration over Brandel GF/C glass fiber filters and washing with cold assay buffer. Filters were incubated overnight in high flash point scintillation cocktail (5 ml, Lefko-Fluor). Radioactivity was counted using a liquid scintillation counter. Counts of non-specific binding were subtracted from total binding to obtain specific [3H]-SCH23390 or [3H]- Spiperone binding, which corresponds to D1 or D2 binding, respectively. Radioactivity counts were converted to fmol/mg tissue for final results.

73

2.6 Behavioral Assessments

2.6.1 Open field locomotor activity

Baseline motor behavior of untreated animals was assessed using open field activity chambers. Open-field locomotor activity and stereotypy were measured using the VersaMax Animal Activity Monitoring System (Omnitech Electronics). Mice were placed in acrylic chambers (20cm x 20cm x 45cm) and infrared light sensors were used to track movement. Dim light is maintained throughout testing and external noise is minimized. Locomotor activity was measured as: distance traveled (measured in cm), number of horizontal movements (discrete movements must be separated by at least 1 second), horizontal activity (number of beam interruptions that occur in the vertical sensor) and vertical activity (number of beam interruptions that occur in the vertical sensor). Stereotypic behavior is defined as repetitive movements such as grooming, head bobbing etc. The software identifies stereotypic behavior when an animal breaks the same beam or set of beams repeatedly. Stereotypy was measured as stereotypy count (number of repetitive beam breaks) and stereotypy number (number of times the monitor observes stereotypic behavior in an animal, a break of one second or more is required to separate stereotypic episodes). These parameters are recorded in five minute intervals over a two-hour period to assess baseline behavior. Data are presented in two ways: 1) in 5 minute increments to display behavior over time or 2) as a sum of the 2-hour period to represent total activity.

Aside from baseline measures, open field activity is also assessed in response to drug treatment. For these experiments, first the animal is allowed to habituate to the chamber while activity is monitored typically for 60 minutes, unless otherwise denoted. Then, the animal is removed from the activity chamber and injected with the drug of choice. The animal is returned to the activity chamber immediately following injection and activity is monitored for another 90 minutes. When data parameters are analyzed over time, the entire 150 minutes of monitoring is shown (60-minute habituation plus 90 minutes post-injection), to gauge the animal’s behavior before and after drug administration. However, to measure total activity (e.g. total distance traveled), only the 90-minute period following drug injection is summed to demonstrate drug-induced effects on behavior.

74

2.6.2 Wire-hang test

The wire-hang test, an assessment of motor strength, was conducted by placing a mouse on a wire cage lid and shaking the lid slightly to make the animal grip the wires. Then the lid was inverted and suspended above a clean cage containing bedding. The latency of the mouse to fall off the grid was measured. Trials were stopped if the mouse remained on the lid for over 10 minutes. Average values were calculated from two trials (at least 15 minutes apart).

Figure 2-1. Wire-hang test apparatus. Mice are suspended on a wire lid above an open clean cage containing bedding. Latency of mice to let go of the wire lid and fall into the cage below is recorded. Image adapted from Stanford Medicine, 2016.

2.6.3 Challenging beam traversal task

The challenging beam traversal is a test of fine motor skill. It was conducted in bright light according to the method described by Fleming et al., 2004. Animals were trained to traverse the length of a Plexiglas beam consisting of four sections (25 cm each, 1 m total length) decreasing in width from 3.5 cm to 0.5 cm by 1 cm increments. Mice were trained for two days (3 trials each) to traverse the beam that led to the animal's home cage. On the third day (test day), a mesh grid (1 cm squares) of corresponding width was placed over the beam surface. A space of approximately 1 cm separated the grid from the surface of the beam. Animals were videotaped

75 while traversing the grid-surfaced beam over three trials, separated by at least 5 minutes. For each trial, animals were allowed a maximum of 5 minutes to traverse the beam. Video files were recorded using a camcorder for later manual scoring. Scoring was performed blind. The video was viewed frame-by-frame to detect errors defined as paw slips through the mesh-grid and paws placed on the side rather than the top of the grid during forward motion. Other parameters recorded include number of steps to cross the beam and latency to traverse the beam. Number of errors, number of steps, errors per step and time to traverse the beam were quantified per trial and averaged over three trials for each animal.

Figure 2-2. Challenging beam traversal task. (A) Side view of apparatus on test day (Day 3) showing beam and mesh grid on top. Representative frames from video recordings demonstrating errors (denoted by arrows) such as (B) front paw slip, (C) rear paw placed on the side of the grid rather than the surface and (D) both paws misplaced. Images taken by Laura Vecchio at Salahpour lab, University of Toronto.

For Project 1, the effect of L-DOPA treatment on fine motor ability was also evaluated using the challenging beam traversal task. In this case, animals were trained for 2 days without any drug treatment. On the test day (Day 3), animals were treated with 12.5 mg/kg of benserazide (i.p.),

76 followed 20 minutes later by 25 mg/kg of L-DOPA (i.p.). In the control group, animals received two 0.9% saline injections separated by 20 minutes. Testing on the challenging beam began 10 minutes after the second injection. This treatment regimen has previously been used to assess L- DOPA effects on challenging beam motor behavior (Hwang et al., 2005). Data were analyzed as mentioned before.

2.6.4 Puzzle box

The puzzle box strives to assess executive functions in rodents (Ben Abdallah et al., 2011). Using different tasks, the puzzle box evaluates multiple aspects of cognition including problem- solving, short- and long-term memory. The puzzle box is divided into 2 compartments: a large, brightly-lit start box (58 x 28 x 27.5 cm) and a smaller, dark, enclosed goal box (14 x 28 x 27.5 cm), that are connected by a door and underpass.

Figure 2-3. Puzzle box apparatus. Images showing top view of the puzzle box which is divided into 2 compartments (start box and goal box) that are connected by an underpass. Image adapted from Ben Abdallah et al., 2011.

77

Since mice are comfortable in dark, enclosed spaces, when they are first introduced to the bright start box, they develop a preference to “escape” to the dark goal box. Over 9 trials, mice are required to complete escape tasks of increasing difficulty to reach the goal box within a limited amount of time. The paradigm consists of 3 tasks per day for 3 consecutive days (total of 9 tasks, T1-9), during which mice are trained to solve a puzzle by removing obstructions at the underpass to reach the goal box. Between each trial (on the same day), the mouse is returned to its home cage for two minutes. A brief description of the tasks is shown in the table below.

Table 2-1. Description of tasks on the puzzle box test.

For each task, latency to escape to the goal box is recorded. The timer is stopped when the two hind legs of the mouse have entered the goal box. Each task lasts for a maximum of 300 seconds (5 minutes) and mice that do not complete the task within that time frame are assigned the maximum score (300 seconds). For all tasks, shorter latency to escape indicates better performance. Mice that cannot complete the first task on Day 1 (T1) are eliminated from the study.

2.6.5 Elevated plus maze

The elevated plus maze is used to measure anxiety-like behavior in rodents. The maze consists of 4 arms (each 30.5 cm long, 5.0 cm wide) connected by a central zone. Two opposite arms are

78 open (no walls) while the other two opposite arms are enclosed with walls (15.25 cm high). Typically, animals prefer to remain in the dim enclosed arms, hence, exploration of open arms is thought to represent reduced anxiety. This test was performed under 2 light conditions: 1) dim light (15-16 lux) and 2) ambient light (210–240 lux), using different groups of mice. At the beginning of the test, mice are placed in the center zone, facing an open arm. The test lasts for 8 minutes, during which mice explore the maze. Biobserve Viewer3 software was used to track animal movement and the amount of time spent in each arm. Data are reported as percent of total time spent in 1) open arms and 2) closed arms. Total time is defined as time spent in all arms as well as the center zone. The software also estimates distance traveled during the test using video tracking.

Figure 2-4. Schematic image of elevated plus maze. Image shows side view of elevated plus maze (2 open and 2 closed arms). Reproduced from Stoelting Co., 2017.

2.6.6 Abnormal Involuntary Movements Scale

The Abnormal Involuntary Movement Scale (AIMS) is used to assess drug-induced dyskinesias. We used this scale to evaluate the effects of 2mg/kg of amphetamine in DAT VMAT2 mice. Animals were placed in an acrylic box and allowed to habituate to the environment for 60 mins.

79

Then they were removed from the box, injected with amphetamine (2 mg/kg) and returned to the box. Their activity was monitored for 75 minutes following injection. In particular, the behavior of each animal was video-recorded for 1 minute intervals every 15 minutes. The first recording was taken right before drug injection to obtain a baseline measure. A total of 6 recordings (15 minutes apart) were taken of each animal for manual scoring. Scoring identified abnormal movements that lacked purpose (rather than akinesia). A score was assigned for 4 individual categories of abnormal movements: 1) locomotor (abnormal locomotion such as crouching low while walking, full body tremors, backwards movement), axial (abnormal postures of the head and trunk; head bobbing), limb (abnormal movements of paw; resting paw tremor, kicking out back legs while walking), and orolingual (vacuous chewing, biting, tongue thrusting). A score of 1 to 4 was assigned for each category based on the duration of an abnormal movement in a 60 second recording (0-15 seconds = 1; 16-30 seconds = 2; 31-45 seconds = 3; and 46-60 seconds = 4). For each time point, the minimum aggregated score from all categories is 4 while the maximum is 16. For an overall effect of the drug, the scores for each time point are summed (minimum 24, maximum 96).

2.7 Drug treatment

2.7.1 MPTP

MPTP hydrochloride (Sigma Aldrich) was dissolved in phosphate buffered saline (PBS) and administered (i.p. 0.1 ml/ 10g body weight) twice, 10 hours apart, at a dose of 15 or 30 mg/kg of body weight. Animals were sacrificed after seven days and brains were harvested for biochemical and neurochemical analyses.

2.7.2 Dopaminergic drugs

Several drugs that interact with the dopamine system (e.g. psychostimulants, dopamine receptor agonists) were acutely injected into drug-naïve animals prior to conducting behavioral analyses. The table below provides a summary of the dose, route of administration, dissolving vehicle and source of each drug that was used.

80

Table 2-2. List of dopaminergic drugs administered All injections were administered at a concentration of 0.1ml per 10g of body weight. i.p., intraperitoneal; s.c., subcutaneous.

Drug Dose (mg/kg) Injection Vehicle Solution Source Amphetamine 0.5, 1, 2 or 5 i.p. PBS R&D Systems Europe Cocaine 20 i.p. 0.9% saline (NaCl) Medisca Methylphenidate 5 i.p. PBS Tocris Bioscience Apomorphine 2 s.c. 0.1% ascorbic acid Sigma Aldrich in distilled H2O SKF-81297 2 i.p. 0.9% saline (NaCl) Sigma Aldrich L-DOPA (methyl ester) 25 i.p. 0.9% saline (NaCl) Sigma Aldrich

Benserazide 12.5 i.p. 0.9% saline (NaCl) Sigma Aldrich (co-administered with L-DOPA)

2.8 Statistics

All data are shown as mean ± SEM. Data were statistically analyzed by two tailed t-tests, one- way ANOVA with Bonferroni post hoc tests, or two-way ANOVA with Bonferroni post hoc tests, as appropriate. GraphPad Prism and SPSS software were used for graphs and statistical analyses. Significance is reported at p<0.05.

81

Chapter 3 Results Results

Thesis results have been divided into two main sections depending on the mouse model that is characterized: 1) DAT over-expressing transgenic mice and 2) mice that simultaneously over- express DAT and under-express VMAT2.

3.1 Characterization of DAT over-expressing transgenic mice

The majority of the results shown in this section have been published (Masoud et al., 2015). Since DAT-tg mice were first generated by Dr. Salahpour at Duke University before I began my doctoral thesis, the following characteristics of these mice have already been summarized in two previous publications: 1) expression and function of DAT, 2) response to psychostimulants and 3) post-synaptic dopamine receptor function (Salahpour et al., 2008; Ghisi et al., 2009). Furthermore, prior to this thesis work, there were three interesting unpublished findings from DAT-tg mice; these animals showed: 1) a 33% reduction in striatal dopamine tissue content, 2) a 72% decrease in electrically-evoked dopamine release from striatal slices and 3) 32-36% and 28- 30% loss of TH positive neurons in the SN and VTA, respectively. These results suggested that DAT over-expression produces detrimental effects in midbrain dopamine neurons of DAT-tg mice. Outlined below are the experiments conducted to expand on previous findings and provide new results regarding underlying mechanisms, behavioral outcomes and response to toxicants.

3.1.1 Presynaptic dopamine homeostasis

The first set of experiments were designed to evaluate how dopamine is handled within the presynaptic neuron of DAT-tg mice. Initially, the expression of DAT in the striatum of DAT-tg mice was confirmed using western blots (Fig. 3.1). As expected, DAT protein levels were significantly higher in DAT-tg mice in comparison to WT animals.

82

Figure 3-1. DAT protein expression in the striatum of DAT-tg mice. DAT western blot and densitometry analysis (N=3). DAT levels were corrected for loading using Na/K ATPase and normalized to WT expression. Data shown are means ± SEM. ** p<0.01.

After confirming that transgenic animals over-express DAT, we investigated whether greater DAT-mediated dopamine uptake may result in accumulation of dopamine within the presynaptic neuron. Functional characterization of DAT had previously indicated faster dopamine uptake in striatal slices from DAT-tg mice (Salahpour et al., 2008). This suggests that the neurotransmitter could potentially buildup in the cytosolic space of the presynaptic neuron, which can produce deleterious consequences for the cell. Since direct in vivo measurement of cytosolic dopamine is not technically feasible, indirect parameters are used to gauge cytosolic dopamine levels. In particular, the cytosolic fraction of dopamine is exposed to various metabolic reactions. While dopamine tissue content in the striatum was previously measured in these animals, levels of DOPAC and HVA, the major metabolites of dopamine, were unknown. Therefore, using HPLC- EC (high performance liquid chromatography with electrochemical detection), we assessed metabolite levels in relation to dopamine content, as an indicator of dopamine degradation. DAT-tg animals showed a 60% increase in the DOPAC/dopamine ratio (Fig. 3.2A) and a 38% increase in the HVA/dopamine ratio (Fig. 3.2B), suggesting a higher turnover of dopamine in these animals. Furthermore, in collaboration with the Goldstein lab at the National Institutes of Health, we assessed levels of DOPAL, a volatile and potentially toxic metabolite of dopamine, in the striatum of DAT-tg mice (Goldstein et al., 2013). As illustrated in Figure 3.2C, DAT-tg mice show a trend (p=0.05) towards a higher DOPAL-to-dopamine ratio in comparison to WT mice.

83

Given the propensity of DOPAL to cause oxidative damage, a possible enhancement of DOPAL content in these mice may contribute to dopaminergic toxicity. In summary, since intraneuronal metabolism of dopamine occurs specifically in the cytosolic space, enhanced dopamine turnover in DAT-tg mice implies buildup of cytosolic dopamine.

Figure 3-2. Metabolite to dopamine ratios in the striatum of DAT-tg mice. Ratio of (A) DOPAC-to-dopamine, (B) HVA-to-dopamine and (C) DOPAL-to-dopamine tissue content (N=10-11). DOPAL-to-dopamine ratios were measured in collaboration with Dr. Goldstein at NIH. Data shown are means ± SEM. **p<0.01.

Cytosolic dopamine levels are regulated by various mechanisms including plasma membrane uptake via DAT, metabolic reactions and VMAT2-mediated vesicular storage. Since the majority of intracellular dopamine is sequestered within vesicles and VMAT2 plays a crucial role in maintaining low levels of cytosolic dopamine, VMAT2 protein levels were evaluated in the striatum of DAT-tg mice. As shown in Figure 3.3, transgenic animals displayed 30% lower VMAT2 protein levels than WT mice. While this decrease may reflect the concurrent loss of dopaminergic neurons in DAT-tg mice, it also implies that reduced vesicular storage could contribute to buildup of cytosolic dopamine in these animals.

84

Figure 3-3. VMAT2 protein expression in the striatum of DAT-tg mice. VMAT2 western blot and densitometry analysis of striatal tissue from WT and DAT-tg mice (N=4). VMAT2-knockdown (VMAT2-kd) samples were used as a negative control to identify the specific VMAT2 band. VMAT2 levels were corrected for loading using GAPDH and normalized to WT expression. Data shown are means ± SEM. **p<0.01.

Taken together, these data indicate possible accumulation of cytosolic dopamine in presynaptic dopaminergic neurons of DAT-tg mice due to 1) increased expression of DAT, the protein responsible for transporting extracellular dopamine into the cytosolic space, 2) increased metabolite-to-dopamine ratios, suggesting presence of dopamine in the cytoplasm where it can be degraded and 3) decreased expression of VMAT2, the protein responsible for sequestering cytosolic dopamine into vesicles.

3.1.2 Markers of oxidative stress

According to previous findings, DAT-tg mice display reductions in dopamine tissue content and electrically-evoked dopamine release in the striatum, despite greater uptake of dopamine (Salahpour et al., 2008). Furthermore, stereological counts of midbrain dopamine neurons corroborated these findings by revealing a concurrent loss of dopamine cells in DAT-tg mice. Although these data demonstrated compromised integrity of dopamine neurons in transgenic animals, potential mechanisms underlying this damage were unexplored. A large body of literature suggests that cytosolic dopamine is highly reactive and can induce oxidative stress.

85

Also, enhanced dopamine metabolism in DAT-tg mice can generate ROS as by-products. Therefore, we investigated whether the spontaneous loss of dopaminergic neurons in DAT-tg mice may be associated with oxidative damage (Graham, 1978; Hastings et al., 1996; Stokes et al., 1999). Several markers of oxidative stress such as protein carbonylation, nitrosylation and cysteinyl adducts were explored in addition to anti-oxidant mechanisms.

ROS can react with biological molecules such as proteins and modify their structure. Carbonyl groups are often formed on protein side chains as a result of direct or indirect oxidative reactions (Dalle-Donne et al., 2003). Since these carbonyl groups are relatively stable and easily detectable, carbonylation is commonly used as a general marker of protein oxidation. The level of protein carbonylation in the striatum of WT and DAT-tg mice was assessed using synaptic plasma membrane preparations as they are enriched with mitochondrial membranes which contain proteins that are particularly sensitive to oxidative damage. First, 3-5 month old (13-20 weeks) mice were used since dopaminergic neurodegeneration is evident at that age, however no differences were detected between genotypes (Figure 3.4A). Then, we postulated that if oxidative stress was causing dopaminergic cell loss, evidence of oxidative damage would be expected to precede the onset of neurodegeneration. Therefore, protein carbonyl levels were evaluated in younger mice (6-8 weeks), however, no changes were observed (Figure 3.4B).

86

Figure 3-4. Protein carbonylation in the striatum of DAT-tg mice. Western blots and quantification of protein carbonyls in synaptic plasma membrane fractions from the striatum of WT and DAT-tg mice. Striata from 3-4 mice were pooled per sample. Animals were (A) 13-20 weeks (3-5 months) old or (B) 6-8 weeks old. Data presented as mean ± SEM.

87

Aside from ROS, reactive nitrogen species such as peroxynitrite can also interact with proteins and alter their structure. These radical species mediate the nitration of susceptible tyrosine residues on proteins to produce 3-nitrotyrosine, another general biomarker of nitrosative and oxidative damage in cells. Assessment of 3-nitrotyrosine levels in striatal synaptic plasma membrane fractions showed no differences between WT and DAT-tg mice using western blots (Figure 3.5A). In general, oxidative stress arises as a result of anti-oxidant mechanisms being overwhelmed by the production of reactive oxygen and nitrogen species. Therefore, in addition to evaluating markers of oxidative damage, anti-oxidant enzyme levels were also investigated to gauge the overall redox environment. Specifically, we measured protein expression of manganese superoxide dismutase (MnSOD), a key mitochondrial enzyme that detoxifies superoxide radicals generated through respiration. In total striatal homogenates, no differences were detected in MnSOD protein levels between WT and DAT-tg mice (Figure 3.5B).

Figure 3-5. Protein nitrosylation and MnSOD levels in DAT-tg mice. (A) Western blot and quantification of 3-nitrotyrosine in striatal synaptic plasma membrane fractions from WT and DAT-tg mice. Striata from 3-4 mice were pooled per sample. Adult mice

88

(3-5 months old) were used. (B) Western blot and quantification of manganese superoxide (MnSOD) protein levels in total striatal extracts from WT and DAT-tg mice (N=6-7). Data presented as mean ± SEM.

While protein carbonylation, nitrosylation and MnSOD levels are used to judge the general level of oxidative stress, these measures are not specific for dopaminergic cells. Since DAT over- expression selectively affects dopaminergic neurons in DAT-tg mice, we needed a sensitive method to specifically measure dopaminergic oxidative modifications. The formation of cysteinyl adducts on dopamine and its metabolites result from reactions between volatile dopamine-quinones and cysteine residues. Importantly, cysteinyl-modified dopamine, L-DOPA and DOPAC are indicative of oxidative stress occurring particularly within dopaminergic neurons, where these substrates are located (Graham, 1978; Fornstedt and Carlsson, 1989; Hastings and Zigmond, 1994). Hence, we designed an HPLC protocol to electrochemically separate cysteinyl adducts of dopamine, L-DOPA and DOPAC from other commonly found neurochemicals in the brain. After optimizing this technique, we were able to achieve isolation of 9 neurochemicals including the afore-mentioned cysteinyl adducts using HPLC-EC as shown in Figure 3.6A. Although this protocol was effective in measuring these chemicals in standard solutions of known concentrations, due to the lack of adequate sensitivity, we were unable to detect trace levels of cysteinyl adducts from brain tissue. Therefore, we collaborated with Dr. Jason Richardson from Rutgers University as his laboratory is equipped with sensitive electrochemical detectors that are optimized for HPLC analysis of cysteinyl adducts. We sent him striatal tissue from 3-5 month old WT and DAT-tg mice, as well as reserpine-treated WT mice as positive controls since reserpine administration has been shown to enhance formation of cysteinyl adducts of dopamine (Fornstedt and Carlsson, 1989). DAT-tg mice exhibit a 35% increase in 5-S-cysteinyl-dopamine (p< 0.05, Fig. 3.6B), in addition to a 62% increase in 5-S- cysteinyl-DOPAC levels (p< 0.01, Fig. 3.6C). Elevated tissue content of cysteinyl-dopamine and cysteinyl-DOPAC suggests that oxidative stress may underlie the dopaminergic cell loss observed in these mice.

89

Figure 3-6. Cysteinyl adducts of dopamine and its metabolites in DAT-tg mice. (A) Representative HPLC traces showing 9 peaks corresponding to 9 individual chemicals separated using HPLC-EC and eluted at distinct time points. The blue trace corresponds to a solution containing all 9 chemicals while the black trace corresponds to a solution containing cysteinyl L-DOPA, cysteinyl-DOPAC, cysteinyl-dopamine and the internal control, DHBA. Although this method was appropriate for separation, it could not be used for detection in brain tissue. HClO4, perchloric acid; DHBA, 2,3-dihydroxybenzoic acid (internal control); DA,

90 dopamine; 5HIAA, 5-hydroxyindoleacetic acid; 5HT, serotonin. Quantification of (B) 5-S- cysteinyl-dopamine and (C) 5-S-cysteinyl-DOPAC tissue content in the striatum of WT and DAT-tg mice (N=9-10) was performed in collaboration with Dr. Richardson at Rutgers University. Data shown are means ± SEM. *p<0.05, **p<0.01.

In summary, general markers of oxidative stress such as protein carbonylation, nitrosylation and levels of the ubiquitous anti-oxidant enzyme, MnSOD, are unchanged in DAT-tg mice. However, specific markers of dopaminergic oxidative stress (cysteinyl dopamine and cysteinyl DOPAC) are significantly elevated in DAT over-expressing animals. These data suggest that the oxidative damage in these mice specifically arises from dopaminergic cells instead of non- dopaminergic sources. Previous experiments on DAT-tg mice demonstrate signs of dopaminergic toxicity including reduced dopamine tissue content, reduced dopamine release and decreased number of midbrain dopamine neurons. The concurrent presence of dopamine-specific oxidative stress in these animals suggests that it may play a role in propagating dopaminergic damage.

3.1.3 Motor behavior

Since the nigrostriatal dopamine pathway is heavily involved in controlling motor activity, we assessed whether dopaminergic cell loss in DAT-tg mice had any influence on their baseline motor behavior. First, open-field locomotion was measured for two hours and no changes were detected in total distance traveled (Fig. 3.7A) or stereotypy (Fig. 3.7B) in DAT-tg mice. Second, animals were assessed using the wire-hang test, a measure of muscle strength where rodents are inverted on a wire grid and suspended above a cage until they fall off (Luk et al., 2012; Oaks et al., 2013). A previous paper has reported that mice with dopaminergic degeneration display deficits on the wire hang test (Luk et al., 2012) even though there were no differences in gross locomotion. Since DAT-tg mice also do not display deficits in general locomotor activity, this test may represent a more sensitive measure of motor coordination. As shown in Figure 3.7C, DAT-tg mice showed 36% shorter latency to fall off the wire in comparison to their WT counterparts (p< 0.05), demonstrating compromised motor strength.

91

Figure 3-7. Motor behavior of DAT-tg mice. (A) Total distance traveled and (B) stereotypy counts from WT and DAT-tg mice tested in open field activity monitors for two hours (N=25-28). Stereotypy counts are defined as the number of beam breaks detected on the infrared monitor during stereotypic behavior. (C) Average latency of mice to fall off the wire in the wire-hang test (N=37-40). Data shown are means ± SEM. *p<0.05.

92

Third, the challenging beam traversal task was used to test fine motor skills of DAT-tg mice. These experiments were performed by Laura Vecchio, Lien Nguyen and myself. In this task, animals traverse a progressively narrow beam in order to reach their home cage. Any slips or misplaced paws during traversal are scored as errors. Results are summarized as number of errors, steps, errors per step and time to traverse the beam. This task is particularly responsive to motor deficits that arise from nigrostriatal dopamine dysfunction (Drucker-Colín and García- Hernández, 1991; Fleming et al., 2004). In addition to testing DAT-tg mice at baseline (saline treatment), we also administered L-DOPA and benserazide prior to behavioral assessment. Since L-DOPA is the precursor to dopamine, we evaluated whether replenishing dopaminergic tone in DAT-tg mice can alter their outcomes on the challenging beam traversal task. Saline-treated DAT-tg mice showed a 50% increase in number of errors (slips and misplaced paws) and a 47% increase in errors per step while traversing the beam (p< 0.01, Fig. 3.8A and p< 0.01, Fig. 3.8C, respectively). However, when treated with L-DOPA, DAT-tg animals performed significantly better as demonstrated by decreased errors, fewer steps taken and lower errors per step in comparison to saline-treated transgenic mice (p< 0.01, Fig. 3.8A; p< 0.05, Fig. 3.8B and p< 0.05, Fig. 3.8C, respectively). Across all groups, there were no differences in time to traverse the beam (Fig. 3.8D). Collectively, results from these behavioral tests indicate that although DAT-tg mice do not show any changes in gross locomotion, they display significant deficits in fine motor coordination. Moreover, L-DOPA-treatment can reverse these deficits, suggesting that the motor deficiency in DAT-tg mice is due to loss of dopaminergic cells and reduced dopamine tone.

.

93

Figure 3-8. Challenging beam traversal task in DAT-tg mice with L-DOPA treatment. Animals were injected with benserazide (12.5 mg/kg), followed 20 minutes later by L-DOPA (25 mg/kg). Control animals were injected with 0.9% saline separated by 20 minutes. Mice (N=8-13) were tested on the challenging beam traversal task (3 trials) 10 minutes after the second injection. (A) Number of errors (including slips and misplaced paws) made while traversing the beam. (B) Number of steps taken to traverse beam. (C) Number of errors per step taken. (D) Time to traverse the beam. Data shown are means ± SEM. *p<0.05, **p<0.01.

While conducting behavioral analyses on DAT-tg mice, some sex differences were noted especially in the wire hang task. Qualitatively, female mice seemed capable of hanging on the wire for much longer periods than male mice. To highlight this difference, the initial wire hang data were stratified by sex as shown in Figure 3.9C. Additionally, baseline data for locomotor, stereotypy and challenging beam traversal were also sex-stratified (see Figure 3.9A, B and D,

94 respectively) to provide a comprehensive overview of sex differences in these behaviors. Female mice seemed to show greater locomotor activity than male mice regardless of genotype. In the behavioral tests that showed significant differences between WT and DAT-tg mice (wire hang and challenging beam), similar trends were also recapitulated in the sex-stratified data.

Figure 3-9. Baseline behaviors of DAT-tg mice stratified by sex. (A) Total distance traveled and (B) stereotypy counts from WT and DAT-tg mice tested in open field activity monitors for two hours (N=10-15 per sex per genotype). Stereotypy counts are defined as the number of beam breaks detected on the infrared monitor during stereotypic behavior. (C) Average latency of mice to fall off the wire in the wire-hang test (n= 18-21 per sex per genotype). (D) Number of errors (including slips and misplaced paws) made while traversing the challenging beam (N=11-12 per sex per genotype). Differences are denoted by lines comparing two groups. Data shown are means ± SEM. * p<0.05, **p<0.01, ***p<0.001.

95

3.1.4 Response to MPTP-induced dopaminergic damage

Sensitivity of DAT-tg mice to exogenous toxicant insult was investigated using MPTP, a compound shown to cause selective damage to dopaminergic neurons in humans as well as animal models. MPTP is converted to its toxic metabolite MPP+, which enters dopamine cells through DAT and inhibits mitochondrial complex I, eventually causing cell death. Since DAT-tg mice basally show evidence of dopaminergic damage, we investigated whether they would also be more vulnerable to toxicant insult. Specifically, using two doses of MPTP, 15 and 30 mg/kg of body weight, we assessed expression of TH, the synthetic enzyme for dopamine and marker of dopaminergic cells in the striatum. TH protein expression was evaluated qualitatively by immunohistochemistry (Fig. 3.10 A) and quantitatively using western blots (Fig. 3.10 B, C). At 15 mg/kg of MPTP, DAT-tg mice displayed lower TH immunofluorescence (Fig. 3.10 A) and protein levels (p< 0.05, Fig. 3.10 C) than WT animals (Fig. 3.10 A, B). Indeed, in WT mice, this dose of MPTP did not elicit any significant change in TH immunoreactivity (Fig. 3.10 A) or protein levels (Fig. 3.10 B) when compared to saline treatment. At 30mg/kg of MPTP, TH immunofluorescence was decreased in both WT and DAT-tg mice (Fig. 3.10 A) however, the extent of reduction was greater in DAT-tg mice as quantified by western blot analysis (Fig. 3.10 B, C). In particular, TH levels were reduced by 65% in transgenic animals (p< 0.001, Fig. 3.10 C) in contrast to only 28% in WT animals (p< 0.01, Fig. 3.10 B), when compared to saline treatment. These results demonstrate that DAT-tg mice are more vulnerable to MPTP treatment and exhibit sensitivity at doses that do not significantly affect WT animals.

96

Figure 3-10. Effect of MPTP treatment on TH protein levels in DAT-tg mice. Adult mice were treated with saline, 15 or 30 mg/kg of MPTP. (A) Immunohistochemical analysis of tyrosine hydroxylase (TH) in the striatum of WT and DAT-tg mice treated with saline, 15 or 30 mg/kg of MPTP. Representative TH-labeled (black) coronal sections are shown. Western blot analysis of TH protein expression in the striatum of (B) WT and (C) DAT-tg mice treated with saline, 15 or 30 mg/kg of MPTP (N=3-4). TH levels were corrected for loading

97

using α-tubulin and normalized to WT expression. Data shown are means ± SEM. Differences are in comparison to saline-treated animals. * p<0.05, **p<0.01, ***p<0.001.

Next, striatal dopamine tissue content was measured to assess the integrity of dopaminergic nerve terminals in MPTP-treated mice. At both 15 and 30 mg/kg of MPTP, the respective reductions in dopamine tissue content were greater in DAT-tg mice compared to WT controls, indicating that increased DAT levels exacerbate MPTP-induced neurotoxicity (15 mg/kg MPTP, p< 0.05; 30 mg/kg MPTP, p< 0.01; Fig. 3.11). A difference in striatal dopamine content was also detected between saline-treated WT and DAT-tg mice (p< 0.01, Fig. 3.11), corroborating the basal reduction in dopamine tissue levels previously observed in untreated transgenic animals.

Figure 3-11. Effect of MPTP on striatal dopamine tissue content of DAT-tg mice. Relative striatal dopamine tissue content is shown for mice treated with saline, 15 or 30 mg/kg of MPTP (N=7-9). Levels are represented as percent of WT saline-treated mice. Significant differences are in comparison to WT mice at each dose. Data shown are means ± SEM. *p<0.05; **p<0.01.

98

In summary, characterization of mice with increased DAT levels has revealed important changes in presynaptic dopamine dynamics such as enhanced dopamine metabolism and reduced VMAT2 levels, which point to the possible accumulation of cytosolic dopamine. Dopaminergic markers of oxidative stress are also elevated in DAT-tg mice, suggesting that oxidative damage may play a role in the loss of midbrain dopamine neurons observed in these mice. Behaviorally, although transgenic animals display normal gross locomotion, their fine motor skills and motor strength are compromised. Interestingly, deficits in motor coordination can be reversed with L- DOPA treatment, implicating reduced dopaminergic tone as the culprit underlying motor deficiencies in DAT-tg mice. Finally, transgenic animals are highly vulnerable to dopaminergic damage induced by MPTP, indicating an important role of DAT in mediating toxicant injury.

3.2 Characterization of mice that over-express DAT and under- express VMAT2

This section encapsulates results from the second project of my doctoral thesis focusing on transgenic mice that simultaneously over-express DAT and under-express VMAT2. These animals were generated by crossbreeding DAT-tg and VMAT2-kd mice in a separate colony (Caudle et al., 2007; Salahpour et al., 2008). This breeding scheme gave rise to all genotypes so littermates can serve as controls. In this project, mice are segregated into 4 genotypes of interest: wild-type (WT), DAT over-expression (DAT-tg), VMAT2-knockdown (VMAT2-kd) and DAT over-expression combined with VMAT2-knockdown (DAT-tg/VMAT2-kd). Collectively, these animals will be referred to as “DAT VMAT2” mice in this thesis. Since results are reported for multiple genotypes and various comparisons are possible, at the end of every subsection, a summary table is provided that outlines the specific experimental results for each genotype.

3.2.1 Confirmation of transporter levels

It should be noted that previous publications demonstrate that in DAT-tg mice, DAT over- expression is restricted to dopaminergic neurons because expression of the transgene is guided by the DAT promotor (Salahpour et al., 2008). On the other hand, in VMAT2-kd mice, VMAT2 expression is reduced in all monoaminergic cells (Caudle et al., 2007).

99

In the first set of experiments, the intended expression of DAT and VMAT2 proteins were assessed across the 4 genotypes of mice. DAT and VMAT2 protein levels in the striatum were analyzed using western blots. RNA expression of the 2 transporters was determined using quantitative reverse transcriptase PCR in the midbrain.

Striatal DAT protein levels were increased in DAT-tg and DAT-tg/VMAT2-kd mice, as expected (Figure 3.12). The degree of DAT over-expression varies between 150-175% of WT levels (Figure 3.12). However, in the original report on DAT-tg mice, DAT levels were increased to 300% of WT levels (Salahpour et al., 2008). Hence, the level of DAT over-expression in DAT VMAT2 mice is not as high as the original DAT-tg mice.

Figure 3-12. DAT protein expression in the striatum. (A) Representative DAT western blot. (B) Quantification of DAT protein using densitometry. (N=8). DAT levels were corrected for loading using Na/K ATPase and normalized to WT expression. Data presented as mean ± SEM. Statistical comparisons are against WT mice. *p<0.05; ***p<0.001.

100

Striatal VMAT2 protein levels were drastically reduced in VMAT2-kd and DAT-tg/VMAT2-kd mice, as expected (Figure 3.13). VMAT2-kd animals showed a 90% reduction in VMAT2 levels when compared to WT, which is similar to earlier reports (Caudle et al., 2007). In the previous characterization of DAT-tg mice, these animals showed a 25% reduction in VMAT2 protein in comparison to WT animals (Figure 3.3), as a likely reflection of dopaminergic neurodegeneration and/or reduced vesicular storage capacity (Masoud et al., 2015). However, this effect was not observed in Figure 3.13 using DAT-tg mice derived from the DAT VMAT2 colony. The western blot was repeated to include a larger sample size (N=4-5, not shown) and still no difference in VMAT2 levels was detected between WT and DAT-tg mice, indicating disparity between original DAT-tg mice and DAT-tg animals from the DAT VMAT2 colony.

Figure 3-13. VMAT2 protein levels in the striatum. (A) VMAT2 Western blot. (B) VMAT2 protein quantification using densitometry (N=2-4). VMAT2 levels were corrected for loading using GAPDH and normalized to WT expression. Data presented as mean ± SEM. Statistical comparisons are against WT mice. ***p<0.001.

Next, RNA levels of these transporters were assessed in the midbrain which contains nigral dopaminergic cell bodies. DAT mRNA was significantly reduced in the midbrain of DAT-tg and DAT-tg/VMAT2-kd mice (Figure 3.14). Since DAT protein is over-expressed in these animals (Figure 3.12), a reduction in mRNA levels is unexpected. Similarly, VMAT2 mRNA expression

101 is increased in VMAT2-kd and DAT-tg/VMAT2-kd mice (Figure 3.14) although these mice show drastic reductions in VMAT2 protein levels (Figure 3.13). mRNA expression is consistently in the opposite direction of protein levels suggesting that 1) mRNA and protein levels do not necessarily correlate and 2) the observed changes in mRNA levels may serve as a compensatory mechanism to normalize the genetic manipulations in these mice.

Figure 3-14. DAT and VMAT2 mRNA expression in the midbrain. (A) DAT mRNA and (B) VMAT2 mRNA relative to wild type (N=6). qPCR results were normalized to the housekeeping gene phosphoglycerate kinase 1 (PGK1). DAT mRNA quantification was corroborated using 2 separate primer sets targeting different parts of the DAT gene. Data presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05; **p<0.01; ***p<0.001.

102

Table 3-1. Summary of DAT and VMAT2 expression in DAT VMAT2 mice. Comparisons are against WT mice.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Transporter protein levels DAT protein Increased 150% (not No change Increased 300% like Salahpour et al, 2008) VMAT2 protein No change (25% Decreased Decreased decrease Masoud et al, 2015) Transporter mRNA levels DAT mRNA Decreased No change Decreased VMAT2 mRNA No change Increased Increased

In summary, DAT-tg/VMAT2-kd mice show higher DAT and lower VMAT2 protein levels, as expected. While protein and mRNA changes are inconsistent, proteins are the functional unit for these transporters and they are appropriately changed in DAT-tg/VMAT2-kd mice. Of note, DAT-tg mice do not replicate the level of DAT over-expression that was previously published, suggesting that the phenotype of these mice may be less severe than the original DAT-tg mice. DAT-tg mice also did not replicate decreased VMAT2 protein levels, which was a reflection of dopamine cell loss. More results presented below support the notion of attenuated dopaminergic toxicity in these DAT-tg mice in comparison to the original DAT-tg mice. Possible reasons for this discrepancy include genetic background, nutrition and breeding differences which are outlined in the Discussion.

3.2.2 Fitness

After confirming that DAT and VMAT2 protein expression was altered in DAT VMAT2 mice as expected, we evaluated whether these animals showed any gross phenotypic changes. Qualitatively, we had observed that DAT-tg/VMAT2-kd mice appear smaller than their littermates and often die prematurely. Since these animals seem more fragile, they are usually weaned at a later age and their rodent chow is routinely supplemented with peanut butter and safflower seeds. Due to these qualitative observations, the survival and body weight of these mice were investigated as indicators of their overall fitness.

103

A survival curve was generated using retrospective data from 2012 to 2015. Data were assimilated from 1) animals that were naturally found dead within the colony, and 2) animals that were intentionally sacrificed to conduct experiments. The survival analysis focused on the time frame between birth and 12 weeks of age, after which animals are usually sacrificed for experiments. DAT-tg/VMAT2-kd mice demonstrate significantly poorer survival than the other genotypes (Figure 3.15 A). In particular, by the end of 12 weeks, survival of DAT-tg/VMAT2-kd mice is reduced by 46%. Furthermore, we had previously observed that more male DAT- tg/VMAT2-kd mice were found dead within the colony, compared to their female counterparts. Therefore, the survival data were separated by sex. Survival of male DAT-tg/VMAT2-kd mice was significantly reduced by 67% at the end of 12 weeks (Figure 3.16 B). Female DAT- tg/VMAT2-kd also showed a trend towards reduced survival, however, the effect did not appear as striking as the males (Figure 3.16 C).

104

Figure 3-15. Survival curve from birth to 12 weeks of age. Kaplan-Meier curves were generated using retrospective data from animals born between 2012 and 2015. Survival curve for (A) All mice, (B) Male mice, (C) Female mice. The number of

105 mice that died naturally (natural) or were sacrificed for experiments after 12 weeks (sac’d) are shown for each genotype. In a few cases, the sex of the animal was not noted at the time of death and therefore those cases were excluded from the sex-stratified analysis. Statistical differences are in comparison to WT, DAT-tg and VMAT2-kd mice. ***p<0.001.

Next, body weight of 2-4 month old animals was compared across genotypes. Since male and female mice differ in weight – these data were also stratified by sex. DAT-tg/VMAT2-kd mice were significantly lighter than WT mice of the same sex (Figure 3.16). Male VMAT2-kd mice were also lighter than their WT counterparts, which has been previously reported (Mooslehner et al., 2001).

Figure 3-16. Body weight of adult mice. (A) Male mice and (B) female mice (N=21-27 per genotype per sex). Data presented as mean ± SEM. Statistical differences are in comparison to wild type mice unless otherwise denoted. *p<0.05; ***p<0.001.

106

Table 3-2. Summary of overall fitness of DAT VMAT2 mice. Comparisons are against WT mice.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Fitness Survival No change No change 50% reduction (67% in males) Body weight No change Reduced in males Reduced in males & (similar to females Mooslehner et al, 2001)

In summary, DAT over-expression combined with VMAT2 knockdown leads to impaired fitness as evidenced by reduced survival and body weight of DAT-tg/VMAT2-kd mice. Given that survival and body weight are complex traits and dopamine plays a variety of roles in the CNS (e.g. locomotion, reward, motivation, lactation) as well as peripheral areas, the precise cause of these changes is unclear. In addition, male DAT-tg/VMAT2-kd mice seem to be more susceptible since as they depict greater decline in survival and body weight in comparison to their female counterparts. While the fitness data were stratified by sex, biochemical, neurochemical and behavioral results (shown below) are with both sexes combined because: 1) typically, the sample size for these experiments are relatively low, therefore, dividing the data by sex would diminish the power to detect differences, 2) the chance of obtaining an animal with a particular genotype and a particular sex is 1-in-12 according to the breeding scheme for DAT VMAT2 mice. Due to these low odds, it is not always feasible to collect enough animals and conduct experiments in a timely manner and 3) as shown in Figure 3.15 B, male DAT- tg/VMAT2-kd are highly susceptible to premature death, therefore, it is difficult to assess parameters exclusively in male mice after 12 weeks of age, when most experiments are conducted.

3.2.3 Presynaptic dopamine homeostasis

Presynaptic neurotransmitter homeostasis is maintained by various processes including release, uptake, synthesis and metabolism. Aspects of these processes were evaluated to determine how dopamine is handled in the presynaptic neurons of DAT VMAT2 mice. First, given that DAT and VMAT2 are involved in packaging and transporting dopamine, levels of dopamine and its

107 major metabolites, DOPAC and HVA, were assessed in the striatum of DAT VMAT2 mice using HPLC with electrochemical detection. As shown in Figure 3.17, dopamine levels are significantly reduced in all 3 genotypes that have altered DAT and/or VMAT2 levels in comparison to WT mice. In particular, VMAT2-kd mice show a 95% reduction in dopamine tissue content that is corroborated by previous characterization of these mice (Caudle et al., 2007). Both VMAT2-kd and DAT-tg/VMAT2-kd mice show similar drastic reductions in dopamine levels, indicating that VMAT2 is crucial in maintaining intracellular dopamine content. In addition, DAT-tg mice show 21% reduction in dopamine levels compared to WT mice. Previously, Dr. Salahpour had found 33% lower dopamine tissue content in DAT-tg mice which was explained by a similar loss of dopamine neurons (Masoud et al., 2015). However, in this case, there is no evidence for dopaminergic cell loss (see stereology results Figures 3.22- 3.24). Hence, reduced dopamine tissue content in these DAT-tg may be a reflection of terminal changes in the striatum while cell bodies in the substantia nigra remain intact. Unlike dopamine tissue content, metabolite levels are not as prominently altered in DAT VMAT2 mice. DOPAC levels are increased in DAT-tg mice and slightly reduced in DAT-tg/VMAT2-kd mice while HVA tissue content is unchanged across all 4 genotypes (Figure 3.17).

Figure 3-17. Striatal tissue content of dopamine and its metabolites. (A) Dopamine, (B) DOPAC and (C) HVA levels assessed in striatal tissue from DAT VMAT2 mice (N=6-13) using HPLC-EC. Data presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05; **p<0.01; ***p<0.001.

108

Second, metabolite-to-dopamine ratios were calculated and used as indicators of dopamine turnover in the striatum (Figure 3.18). VMAT2-kd and DAT-tg/VMAT2-kd mice display greatly increased DOPAC/dopamine and HVA/dopamine ratios in comparison to WT mice. DAT-tg mice also show trends towards increased metabolite-to-dopamine ratios, as previously observed (Masoud et al., 2015). While the increased ratios are clearly a function of reduced dopamine tissue levels in these mice (Figure 3.17 A), they also indicate enhanced dopamine metabolism which could serve as a mechanism to control cytosolic buildup of dopamine.

Figure 3-18. . Metabolite-to-dopamine ratios in the striatum. (A) DOPAC-to-dopamine (DOPAC/DA) and (B) HVA-to-dopamine (HVA/DA) ratios calculated from striatal tissue content levels (N=6-13). Data presented as mean ± SEM. Statistical differences are in comparison to WT mice. ***p<0.001.

109

Since dopamine content in the presynaptic neuron is influenced by neurotransmitter release and uptake, these parameters were evaluated in collaboration with Dr. Miller at Emory University. We sent them mice from the DAT VMAT2 colony and they performed fast-scan cyclic voltammetry (FSCV) on brain slices to determine dopamine release and clearance. Dopamine release was electrically-evoked by a stimulating electrode and the resulting current was measured by a recording electrode. We focused on the dorsal striatum which is the major projection area for nigral dopamine neurons. Data were analyzed using two redundant, yet different methods. The first simplistic method approximates “dopamine release” using the peak amplitude and “dopamine clearance” using the decay-time constant, tau (where lower tau, measured in seconds, implies faster clearance) (Yorgason et al., 2011). The second method, recently proposed by Hoffman and colleagues, uses curve modeling to control for the interdependence between release and uptake and provide accurate representation of these parameters (Hoffman et al., 2016). In this analysis, higher rate of uptake implies faster clearance. Results are shown using both these methods in Figure 3.19. As indicated by both analyses, dopamine release in the dorsal striatum is significantly lower in DAT-tg, VMAT2-kd and DAT-tg/VMAT2-kd mice in comparison to WT animals (Fig 3.19 B, C). This corroborates previous data also showing decreased striatal dopamine tissue content in the 3 genotypes mentioned (Fig. 3.17 A). In VMAT2-kd and DAT- tg/VMAT2-kd mice, reduced vesicular storage of dopamine could lead to the observed reduction in dopamine release. For DAT-tg mice, previous characterization also reported decreased dopamine release due to the loss of dopamine cells (Masoud et al., 2015). However, these DAT- tg mice do not demonstrate neurodegeneration (see Figures 3.22-3.24 in next section), suggesting that reduced dopamine release in the striatum of these animals may indicate terminal damage in the absence of cell loss.

With regards to dopamine clearance, DAT-tg and DAT-tg/VMAT2-kd mice were expected to display higher uptake as a result of increased DAT expression. However, using tau measurements (Fig 3.19 D), VMAT2-kd and DAT-tg/VMAT2-kd mice display faster dopamine clearance compared to WT mice, while DAT-tg mice also show a non-significant trend. Previously, using Michaelis-Menten kinetics, the clearance of evoked dopamine (Vmax) was reported to be substantially faster in DAT-tg mice, attesting to a functional increase in DAT (Salahpour et al., 2008). However, in this experiment, even VMAT2-kd mice, which have normal levels of DAT, demonstrate faster dopamine clearance. Another technical consideration

110 that should be noted is that tau is a simplistic measure of dopamine uptake that could have been influenced by low dopamine release in these genotypes. If dopamine release is reduced, then conceivably, overall neurotransmitter clearance may also require a shorter period of time. Using the Hoffman analysis, no significant changes in dopamine uptake were observed across the genotypes, although VMAT2-kd and DAT-tg/VMAT2-kd mice showed trends towards higher dopamine uptake. In summary, DAT-tg mice demonstrate a trend towards increased dopamine uptake (using tau) while DAT-tg/VMAT2-kd mice display significantly enhanced dopamine clearance, attesting to functional over-expression of DAT in these mice.

111

112

Figure 3-19. Electrically evoked dopamine release and uptake in the dorsal striatum. Determined by FSCV in slice preparations. (A) Traces of dopamine currents recorded over time following a single-pulse stimulation. The ascending curve represents dopamine release while the descending curve represents dopamine clearance. Dopamine release is estimated by (B) peak amplitude (N=4-5) and (C) the Hoffman parameter, r/ke (N=2-4). Dopamine uptake/clearance is estimated by (D) the decay time constant, tau (N=3-5) and (E) the Hoffman parameter, ku (N=3- 4). For Hoffman modeling, individual data points must meet certain criteria in order to be included in the analysis and as a result, some data were excluded. Data presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05; **p<0.01; ***p<0.001.

Another key process that regulates presynaptic dopamine levels is synthesis. Since TH is the rate-limiting enzyme involved in dopamine production, TH protein expression and levels were measured in the striatum of DAT VMAT2 mice. As shown in Figure 3.20, there are no differences across the genotypes in 1) TH expression, as assessed by immunohistochemistry (A) or 2) TH protein levels as assessed by western blots (B, C). These results suggest that dopamine production is unlikely to be altered in mice with varying levels of DAT and VMAT2, although other factors such as TH activity would need to be evaluated to obtain conclusive knowledge of the synthetic pathway.

113

Figure 3-20. TH protein expression in the striatum. (A) TH immunohistochemistry. Representative TH-labeled (black) coronal sections are shown for each genotype. (B) TH western blot using GAPDH as a loading control. (B) Quantification of TH protein using densitometry. TH levels are normalized to WT expression. (N=3-4). Data presented as mean ± SEM

Lastly, a pilot experiment was conducted to further probe dopamine metabolism in the presynaptic neuron. Since metabolite-to-dopamine ratios were elevated in VMAT2-kd and DAT- tg/VMAT2-kd mice, we investigated protein expression of MAO-B, the enzyme responsible for degrading dopamine to DOPAC, as another indicator of dopamine metabolism. Increased DOPAC-to-dopamine ratios may suggest an up-regulation of MAO-B in these mice. In a trial experiment, a western blot of MAO-B protein was conducted using a new antibody in total striatal homogenates (Figure 3.21 A). Unfortunately, this antibody produced a variable signal with high background staining. Although statistical differences between the 4 genotypes were

114 lacking, there does seem to be a trend towards higher MAO-B protein levels in DAT-tg and DAT-tg/VMAT2-kd mice in comparison to WT animals (Figure 3.21 B). Despite multiple attempts with various antibodies, the MAO-b western blot could not be reliably reproduced.

Figure 3-21. MAO-B protein expression in the striatum. (A) MAO-B western blot. (B) Quantification of MAO-B protein using densitometry. MAO-B levels were corrected for loading using GAPDH and normalized to WT expression (N=2-4). Data presented as mean ± SEM.

115

Table 3-3. Summary of presynaptic dopamine homeostasis in DAT VMAT2 mice. Comparisons are against WT mice.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Striatal Neurochemistry Dopamine 21% decrease (33% 96% decrease 95% decrease decrease Masoud et al, 2015) DOPAC Increased No change decrease HVA No change No change No change Evoked dopamine release and uptake Striatal release Decreased Decreased Decreased Striatal uptake Trend increase (50% Increased (tau) Increased (tau) increase in Salahpour et al, 2008) Metabolism DOPAC/dopamine Trend increase. Increased (replicate Large increase Similar to Masoud et Taylor et al 2009, al 2015 2013) HVA/dopamine Trend increase. Large increase Large increase Similar to Masoud et al 2015 MAO-B protein levels Unchanged (Trend Unchanged Unchanged (Trend increase?) increase?) Synthesis TH protein expression Unchanged Unchanged Unchanged TH protein levels Unchanged Unchanged Unchanged

In summary, DAT-tg/VMAT2-kd mice show prominently reduced dopamine tissue content in the striatum as well as decreased dopamine release in both the striatum and nucleus accumbens, demonstrating mishandling of dopamine in the presynaptic neuron. Since both HPLC tissue content and electrically-evoked FSCV parameters predominantly reflect vesicular monoamine stores, knockdown of VMAT2 has a robust and similar effect in both VMAT2-kd and DAT- tg/VMAT2-kd mice. Therefore, any additional contribution of DAT over-expression in DAT- tg/VMAT2-kd mice is difficult to gauge using these techniques. Furthermore, although DAT-tg mice display a trend towards higher dopamine uptake, it does not seem as significant as

116 previously reported (Masoud et al., 2015). Similarly, while DAT-tg mice show reduced dopamine tissue content, it is not as pronounced as previous reports (Masoud et al., 2015). This is consistent with a lower degree of DAT over-expression in these mice as well (Figure 3.12). Once again, collectively, these results suggest that the DAT-tg mice from the DAT VMAT2 colony do not display as severe phenotypes as the original DAT-tg line. This is corroborated by stereology results where DAT-tg mice no longer display dopamine cell loss (see Figures 3.22 – 3.24). Regarding dopamine metabolism, DAT-tg/VMAT2-kd mice demonstrate increased metabolite-to-dopamine ratios, suggesting that enhanced dopamine turnover may represent a compensatory mechanism to counteract the buildup of cytosolic dopamine. Lastly, while presynaptic dopamine dynamics are altered in DAT VMAT2 mice, TH levels appear to be unchanged.

3.2.4 Integrity of dopamine neurons

Previously, Dr. Salahpour demonstrated a 36% reduction of nigral dopamine neurons in 3-5 month old DAT-tg mice (Masoud et al., 2015), while Caudle et al. reported a 26% loss of nigral dopamine neurons in 22 month old VMAT2-kd mice (Caudle et al., 2007). Thus, it was hypothesized that DAT-tg/VMAT2-kd mice will display greater cell loss than either genotype alone, since the dual effect of DAT over-expression and VMAT2-knockdown could lead to more accumulation of cytosolic dopamine and cause toxicity.

To test this hypothesis, we conducted stereological counts of dopaminergic neurons in the SNpc using TH as a dopaminergic marker and NeuN or Nissl as a neuronal marker. For these experiments, we collaborated with two research groups that possess the necessary equipment, analytical software and expertise to routinely perform stereology on dopaminergic neurons. Counters were blind to genotypes and 3-5 month old animals were used for stereology. At this age, we expect DAT-tg mice to show cell loss (Masoud et al., 2015) and VMAT2-kd mice to show no change. We repeated stereology in 3 independent experiments as outlined below.

The first attempt was a collaboration with Dr. Miller’s lab at Emory University in 2011. We sent them perfused and fixed tissue for stereology and TH counts were performed by Carlos Lazo, a new post-doc at the time. No differences were found between the 4 genotypes in the SNpc

117

(Figure 3.22). DAT-tg mice were meant to serve as a control since we previously reported 36% neuronal loss in these mice, however this effect was not replicated.

Figure 3-22. Stereological counts of TH+ cells in the SNpc. Counts performed using 3-5 month old DAT VMAT2 mice (N=7-8) in collaboration with Carlos Lazo from Miller lab, Emory University.

Second, we collaborated with Dr. Brotchie’s lab at Toronto Western Krembil Research Institute in 2014. Initially, I started to count TH and NeuN positive cells in the SNpc using their software, however due to high background staining, positive cells were difficult to distinguish and results were variable. Therefore, to enhance experimental efficiency, we sent perfused, fixed and sectioned tissue to Gabriela Reyes at the Brotchie lab for staining and stereological counting since she regularly performs such experiments. The results for TH and NeuN are shown in Figure 3.23. No differences were found between genotypes and the previously published cell loss in DAT-tg mice (Masoud et al., 2015) was not replicated. For NeuN, DAT-tg mice showed a trend towards lower counts in comparison to WT mice although this was not reflected in TH counts (Figure 3.23 A, B).

118

Figure 3-23. Stereological counts of TH+ and NeuN+ cells in SNpc. Counts of (A) TH and (B) NeuN positive cells in the SNpc of 3-5 month old DAT VMAT2 mice. (N=5-10). Data presented as mean ± SEM. Collaboration with Gabriela Reyes from Brotchie lab, Krembil Institute, Toronto Western Hospital.

Third, we collaborated with Dr. Miller’s lab at Emory University again in 2015. This time, stereological counting was performed by Amy Dunn, who is experienced at counting dopamine neurons in genetic models. For this experiment, we also included additional controls: midbrain sections from 18-24-month old WT and VMAT2-kd mice. At this age, VMAT2-kd mice are expected to show approximately 20% reduction in TH+ and Nissl+ cells in comparison to WT mice (Caudle et al., 2007). As shown in Figure 3.24, the cell loss in aged VMAT2-kd is replicated. However, dopaminergic cell loss in DAT-tg mice is still not replicated. Similar to the 2 previous attempts, no differences were detected between any of the DAT VMAT2 genotypes.

119

Figure 3-24. Stereological counts of TH+ and Nissl+ cells in SNpc. Counts of (A) TH and (B) Nissl positive cells in the SNpc of 3-5 month old DAT VMAT2 mice (N=6-8). Collaboration with Amy Dunn at Miller lab, Emory University. Aged (18-24 month old) WT and VMAT2-kd mice (N=3-4) were included as controls (shown in blue). Statistical difference between aged VMAT2-kd mice in comparison to aged wild type control animals. Data presented as mean ± SEM. *p<0.05; **p<0.01.

Table 3-4. Summary of dopamine cell counts in SNpc of DAT VMAT2 mice. Comparisons are against WT mice.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Dopamine cell counts

SNpc TH No change (does No change as No change 1) Carlos (Miller lab) not replicate 36% expected (Caudle et 2) Gaby (Brotchie lab) loss in Masoud et al, 2007) 3) Amy (Miller lab) al, 2015)

SNpc Nissl/NeuN No change (does No change as No change 1) Carlos (Miller lab) not replicate 32% expected (Caudle et 2) Gaby (Brotchie lab) loss in Masoud et al, 2007) 3) Amy (Miller lab) al, 2015)

120

In summary, DAT-tg/VMAT2-kd did not show any change in dopamine cell number, which was unexpected. These experiments were repeated 3 times not only to ascertain the result, but also because of the following challenges: 1) Stereology proved to be an unreliable technique. In fact, we sent the same slides to be counted by the same researcher (Brotchie lab) twice and obtained opposing results, highlighting the variability of this technique. Due to these technical issues, our confidence in the results was diminished. 2) DAT-tg mice no longer showed dopamine cell loss as previously published (Masoud et al., 2015). The premise of the DAT VMAT2 project was based on the finding that DAT- tg mice show dopaminergic degeneration at 3-5 months of age and we hypothesized that by concurrently decreasing VMAT2 in these mice, the degeneration would be exacerbated. In 3 independent experiments, DAT-tg mice failed to show dopamine cell loss or even a trend towards reduced cell number. This result consolidated our suspicion that DAT-tg mice from the DAT VMAT2 colony lacked the degree of dopaminergic damage that was previously noted. Possible reasons for these differences is that original DAT-tg mice and our current DAT-tg mice belong to separate colonies that may have differences in genetic background/modifiers, nutrition and breeding. These issues are further examined in the discussion. However, keeping in mind that DAT-tg mice no longer show dopamine cell loss and VMAT2-kd mice only demonstrate cell loss at 18 months or older, it is not surprising to find that 3-5 month old DAT-tg/VMAT2-kd mice do not show signs of neurodegeneration.

3.2.5 Dopamine receptor levels

After assessing presynaptic dopamine homeostasis and neuron integrity, post-synaptic dopamine receptor levels were investigated in DAT VMAT2 mice. Since DAT-tg, VMAT2-kd and DAT- tg/VMAT2-kd show reduced dopaminergic tissue content (Figure 3.17 A) and release (Figure 3.19 B, C), it was predicted that post-synaptic dopamine receptors would be up-regulated in these mice to compensate. It has been previously shown that DAT-tg mice have a 30% increase in D1 receptors and a 62% increase in D2 receptors (Ghisi et al., 2009). As an additional control, a few DAT-KO samples were also included. DAT-KO mice are hyperdopaminergic due to the

121 lack of dopamine uptake and therefore, D1 and D2 receptors are downregulated by 60% and 40%, respectively (Ghisi et al., 2009).

D1 and D2 receptor binding assays were performed on striatal membranes from DAT VMAT2 mice. For D1 binding, there was noticeable variability within genotypes perhaps because frozen striatal samples were used for these experiments. No significant differences were detected between the 4 genotypes (Figure 3.25 A). In particular, the previously-reported increase in D1 levels of DAT-tg mice was no longer observed, consistent with other results (e.g. stereology) that also failed to recapitulate earlier findings. However, a 67% reduction was replicated in DAT-KO samples suggesting that the experimental conditions were suitable for detecting changes in D1 binding. Statistical comparisons between WT control and DAT-KO samples are not shown due to small sample size (N=1-2). Regarding D2 binding, VMAT2-kd and DAT-tg/VMAT2-kd samples show increased D2 levels in comparison to WT mice (Figure 3.25 B). However, once again, the previously reported 62% increase in D2 levels was not replicated in DAT-tg mice (Ghisi et al., 2009). DAT-KO mice showed 26% lower D2 binding in comparison to WT mice. This decrease is not as large (40%) as previously reported (Ghisi et al., 2009), however with 1-2 samples, at least a similar trend was observed.

122

Figure 3-25. Dopamine receptor levels in the striatum. Receptor levels determined by radioligand binding. (A) D1 and (B) D2 receptor levels (N=4-6 samples, each sample consists of striatal tissue from 3-4 mice). DAT knockout (DATKO) samples were included as controls (N=2 samples, each sample contains striatal tissue from 4 mice, shown in blue). Statistical comparisons between WT control and DAT-KO mice were not performed due to low sample size (N=1 for WT control sample). Differences are in comparison to respective wild type samples. Data presented as mean ± SEM. *p<0.05.

Table 3-5. Summary of dopamine receptor levels in the striatum of DAT VMAT2 mice. Comparisons are against WT mice.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Receptor levels D1 No change – did not Trend increase Trend increase replicate 30% increase (Ghisi et al., 2009) D2 No change – did not Increased Increased replicate 62% increase (Ghisi et al., 2009)

123

In summary, DAT-tg/VMAT2-kd mice show increased D2 levels and a possible trend towards increased D1 binding. DAT-tg/VMAT2-kd mice show 1) low dopamine tissue content and 2) decreased dopamine release, therefore, as a compensatory mechanism, upregulation of dopamine receptors was anticipated in these animals. Since similar upregulation of dopamine receptors was also observed in VMAT2-kd mice, enhanced receptor levels may not be sufficient to explain some phenotypes that are unique to DAT-tg/VMAT2-kd mice such as basal hyperactivity (see Figure 3.26 below). Once again, DAT-tg mice did not show expected increases in receptor levels. However, since presynaptic dopaminergic cell loss was not replicated in the DAT-tg mice from the DAT VMAT2 colony (Figures 3.22-3.24), compensatory changes in postsynaptic dopamine receptors are also less likely to occur. Taken together, these DAT-tg mice consistently display less severe phenotypes than the original DAT-tg mice.

3.2.6 Baseline behavior

Dopamine signaling mediates a variety of behaviors such as locomotion, cognition and reward. Since DAT-tg/VMAT2-kd mice show significant changes in dopamine handling, we investigated whether these changes had any effect on their baseline behaviors. In particular, the following behaviors were examined: 1) open field locomotion, 2) fine motor skills using challenging beam traversal, 3) executive function using puzzle box and 4) anxiety-like behavior using elevated plus maze.

Initially, basal locomotion was assessed in open-field activity monitors and the following parameters were measured for two hours: 1) total distance traveled, 2) number of horizontal movements, 3) horizontal activity, defined as the number of beam interruptions that occur in the horizontal sensor, 4) vertical activity, defined as the number of beam interruptions that occur in the vertical sensor, 5) stereotypy count, the number of times the animal breaks the same beam(s) repeatedly during stereotypic behavior and 6) stereotypy number, the number of times the animal is monitored to engage in stereotypic behavior. Collectively, these parameters provide a holistic view of an animal’s locomotor behavior and are summarized in Figure 3.26. DAT-tg/VMAT2-kd mice are hyperactive as evidenced by a 5-fold increase in total distance traveled, higher number of movements and greater horizontal activity in comparison to WT, DAT-tg and VMAT2-kd

124 mice (Figure 3.26 B, C, D). Furthermore, when distance traveled is measured in 5 minute intervals, DAT-tg/VMAT2-kd mice are more active than other genotypes for the first hour, however during the second hour, their movement begins to decline and they apparently habituate to the environment, similar to other genotypes (Figure 3.26 A). These data suggest that the hyperlocomotion of DAT-tg/VMAT2-kd mice is a reflection of increased exploration in a novel environment. Conversely, DAT-tg/VMAT2-kd mice show no significant changes in vertical activity, whereas VMAT2-kd display lower vertical activity that WT mice (Figure 3.26 E). Stereotypy refers to repetitive movements such as grooming that have been shown to involve stimulation of striatal dopamine receptors (Delfs and Kelley, 1990). While baseline stereotypy counts were unchanged between the 4 genotypes, DAT-tg/VMAT2-kd mice show a trend towards increased stereotypic behavior. Also, in comparison to VMAT2-kd mice, DAT- tg/VMAT2-kd mice display a greater number of stereotypic movements. In conclusion, the basal locomotion of DAT-tg/VMAT2-kd mice is significantly different from other genotypes demonstrating that the dual effects of DAT over-expression and VMAT2 knockdown produces unique motor phenotypes in these mice.

125

126

Figure 3-26. Open field locomotion and stereotypy. Adult mice were assessed using automated VersaMax software (N=13-18). (A) Distance traveled over time. Differences are in comparison to WT mice. Sum of (B) total distance traveled, (C) number of horizontal movements, (D) horizontal activity (the number of beam interruptions that occur in the horizontal sensor), (E) vertical activity (the number of beam interruptions that occur in the vertical sensor), (F) total stereotypy count (the number of times the animal breaks the same beam(s) repeatedly) and (G) stereotypy number (the number of times an animal is monitored to engage in stereotypic behavior with a minimal interval of 1 second). Measures were obtained over a two-hour period using VersaMax software. Data presented as mean ± SEM. Statistical differences are as denoted. In the case of multiple significant differences, the top asterisk denotes comparison to WT mice, the middle asterisk refers to comparison against DAT- tg mice and the bottom asterisk is versus VMAT2-kd animals. *p<0.05; **p<0.01; ***p<0.001.

Given the striking and unexpected phenotype of hyperactivity in DAT-tg/VMAT2-kd mice, two additional pilot experiments were conducted to further explore the robustness of this result. First, the locomotor behavior of aged (12-month old) DAT VMAT2 mice was assessed. This is the only experiment where aged mice were used. As shown in Figure 3.27, the hyperactivity of DAT-tg/VMAT2-kd mice persists even at 12 months of age since they travel twice as much as age-matched WT mice. However, in comparison to adult DAT-tg/VMAT2-kd animals (3-5 months old, Fig 3.26 B), the degree of hyperactivity is diminished in aged DAT-tg/VMAT2-kd mice.

127

Figure 3-27. Locomotor activity of 12-month old mice.

Total distance traveled over 2 hours in open-field activity monitors (N=5-8). Data were assessed using VersaMax software and are represented as mean ± SEM. Statistical differences are in comparison to WT (top) and DAT-tg (bottom) mice. *p<0.05; **p<0.01.

Second, we evaluated whether the degree of VMAT2-knockdown can affect the baseline locomotor phenotype of DAT-tg/VMAT2-kd mice. Clearly, manipulating the expression of a single transporter, DAT or VMAT2, is not sufficient to induce hyperlocomotion since both DAT-tg and VMAT2-kd mice show similar locomotor activity as WT mice. In DAT-tg/VMAT2- kd mice, DAT is over-expressed by 50% while VMAT2 is simultaneously under-expressed by 95%. In this experiment, we posed the question whether a less dramatic reduction in VMAT2 levels could be coupled with DAT over-expression to recapitulate the hyperactivity of DAT- tg/VMAT2-kd mice. Fortunately, due to the breeding scheme used in the DAT VMAT2 colony, half of the animals produced are heterozygous for the VMAT2 locus. Hence, we had access to VMAT2-het mice that display 50% of VMAT2 levels and have been extensively characterized by other groups. In this pilot experiment, the locomotor activity of adult (3-5 month old) WT

128

mice was compared with DAT-tg/VMAT2-het animals, that should show 50% higher DAT and 50% lower VMAT2 expression. These mice were used as a tool to further explore the robustness of the locomotor findings from DAT-tg/VMAT2-kd mice. Interestingly, unlike DAT-tg/ VMAT2-kd animals (95% lower VMAT2), DAT-tg/VMAT2-het mice (50% lower VMAT2) are not hyperactive (Fig 3.28). These results suggest that on a background of DAT over-expression, a 50% reduction in VMAT2 levels is not sufficient in producing hyperlocomotion and instead, a 95% reduction is likely necessary. Hence, a threshold effect can be observed, where a high degree of VMAT2-knockdown is required for the locomotor phenotype to manifest.

Figure 3-28. Locomotor activity of DAT-tg/VMAT2-het mice.

Total distance traveled over 2 hours in open-field activity monitors (N=4-8). Data were assessed using VersaMax software and are represented as mean ± SEM.

After evaluating gross locomotion, fine motor skill of DAT VMAT2 mice was assessed using the challenging beam traversal task. Briefly, animals were trained to traverse a beam of narrowing width leading to the home cage. On the test day, a mesh grid was placed on top of the beam and any slips or misplaced paws were counted as errors. As shown in Figure 3.29A, DAT-tg mice display significantly higher errors compared to WT, VMAT2-kd and DAT-tg/VMAT2-kd mice. These findings corroborate previous results from Figure 3.8 where DAT-tg mice showed a 50%

129 increase in the number of errors compared to WT mice (Masoud et al., 2015). Findings from the challenging beam traversal task suggest that DAT over-expression negatively impacts fine motor skill whereas concurrent VMAT2-knockdown seems to rescue this deficit in DAT-tg/VMAT2-kd mice. No differences were detected in the latency to traverse the beam across the 4 genotypes due to high variability (Figure 3.29 B). Although DAT-tg/VMAT2-kd mice showed increased locomotion in open-field analyses (Fig 3.26B), they did not appear to be hyperactive during this task and instead showed a trend towards longer latency to cross the beam compared to WT mice (Fig 3.29 B). Perhaps the short duration of the challenging beam task (5 minutes maximum) combined with 2 previous days of training, allows animals to become habituated to the environment. In conclusion, despite their basal open-field hyperlocomotion, the fine motor skills of DAT-tg/VMAT2-kd mice remain intact in the challenging beam traversal task. In fact, the behavioral deficits in DAT-tg mice are ameliorated in DAT-tg/VMAT2-kd mice, alluding to a beneficial effect.

Figure 3-29. Fine motor skill evaluated using the challenging beam traversal task. DAT VMAT2 mice were trained on this task for 2 days and results were collected on the third test day as an average of 3 trials (N=13-15). (A) Number of errors while traversing the beam.

130

Errors represent slips or misplaced paws while traversing the beam. (B) Latency to cross the beam. Data presented as mean ± SEM. For statistical differences, the top asterisk denotes a comparison to WT mice, the middle asterisk refers to comparison against VMAT2-kd mice and the bottom asterisk is versus DAT-tg/VMAT2-kd animals. *p<0.05; **p<0.01.

Next, executive function of these mice was examined using the puzzle box, a problem-solving test in which mice are required to complete escape tasks of increasing difficulty within a limited amount of time (Ben Abdallah et al., 2011). The box is divided into 2 compartments: a brightly lit start box and a dark goal box that are connected by a door and underpass. The test consists of 3 tasks per day for 3 days producing a total of 9 tasks (T1-9). Time to solve the puzzle and escape to the goal box is recorded for each trial, where shorter latency corresponds with better performance. A brief description of each task is shown in Table 2.1. Collectively, the puzzle box provides information on problem-solving, short-term and long-term memory.

The results for each task of the puzzle box test are shown in Figure 3.30. DAT-tg/VMAT2-kd mice do not perform differently from WT mice on any of the tasks. However, interestingly, VMAT2-kd mice are significantly worse than the other 3 genotypes on T5 and T8, when a novel puzzle is introduced. VMAT2-kd mice also display poorer outcomes on T9 in comparison to DAT-tg and DAT-tg/VMAT2-kd mice. Typically, T9 is a test of short term memory since the mice are required to solve the same puzzle as T8. However, since several of the VMAT2-kd mice failed to solve the new task in T8, the higher latency in T9 may represent their ongoing struggle with the difficult puzzle. Taken together, problem-solving ability is diminished in VMAT2-kd mice however, DAT-overexpression seems to reverse this deficit since DAT- tg/VMAT2-kd mice show normal problem-solving abilities. In summary, executive function of DAT-tg/VMAT2-kd mice is intact as assessed by the puzzle box test.

131

Figure 3-30. Executive function evaluated using the puzzle box. Latency of mice to enter the goal box is shown (N=6-13). Data presented as mean ± SEM. Statistical differences are denoted for each task as follows: T5: ** vs WT, ** vs DAT-tg, ** vs DAT-tg/VMAT2-kd; T8: * vs WT, *** vs DAT-tg, *** vs DAT-tg/VMAT2-kd; T9: ** vs DAT- tg, ** vs DAT-tg/VMAT2-kd. *p<0.05; **p<0.01; ***p<0.001.

Lastly, baseline anxiety-like behavior of DAT VMAT2 mice was measured using the elevated plus maze. The elevated plus maze consists of 2 enclosed arms, 2 open arms and a center zone. Typically, animals prefer to remain in the closed arms and exploration of open arms is thought to represent reduced anxiety. This test was conducted under 2 light conditions: 1) dim (15-16 lx) and 2) bright (210-240 lx) light. Separate groups of animals were used for each condition. Notably, depending on the amount of ambient light, the results of the elevated plus maze were different. In dim light, there were no changes in the percent time spent in closed arms (Figure 3.31 B), however both VMAT2-kd and DAT-tg/VMAT2-kd mice spent significantly more time exploring open arms than WT mice (Fig 3.31 A). This suggests that VMAT2-kd and DAT- tg/VMAT2-kd mice are less anxious than their WT counterparts. During this task, the locomotor activity of the 4 genotypes were similar, discounting any possible effect of DAT-tg/VMAT2-kd hyperactivity in influencing these results (Fig 3.31 C). On the other hand, in bright light conditions, there were no differences between the genotypes in time spent in the open or closed arms (Fig 3.31 D, E). Contrary to the previous result, VMAT2-kd mice showed a trend towards

132 less time spent in open arms (Fig 3.31 D), suggesting that the level of anxiety may be increased in these mice. Interestingly, while the percent of time spent in open arms by DAT-tg/VMAT2-kd mice was similar in dim (Fig 3.31 A) and bright conditions (Fig 3.31 D) (roughly 20%), the other 3 genotypes displayed variable behavior depending on the level of light. WT and DAT-tg mice seemed to spend a larger percent of time in the open arms under bright light (15-18%) versus dim light (10%) while VMAT2-kd mice displayed the opposite trend. Hence, DAT-tg/VMAT2- kd mice appear to be relatively insensitive to ambient light conditions during the elevated plus maze, whereas the behavior of other genotypes is altered. However, locomotor activity of VMAT2-kd and DAT-tg/VMAT2-kd mice was reduced compared to WT mice in bright light (Fig 3.31 F). In general, results from the elevated plus maze were dependent on the light conditions and while reduced anxiety-like behavior was observed in VMAT2-kd and DAT- tg/VMAT2-kd mice under dim light, these differences were no longer present under bright light. Previously, VMAT2-kd mice were shown to display higher anxiety-like behavior on the elevated plus maze, however there are important methodological differences between those experiments and this study (Taylor et al., 2009). In the previous case, the light-dark cycle of the mice was reversed for weeks prior to testing, such that they were kept in darkness during the day. Then on the test day, when they would normally be accustomed to darkness, they were exposed to bright light during the elevated plus maze task. Given the potent effect of light conditions on this task as demonstrated by my experiments, it is highly likely that the difference in results obtained is directly attributable to the cycle reversal and light conditions chosen for each particular experiment.

133

Figure 3-31. Anxiety-like behavior assessed using elevated plus maze. This test was conducted using different groups of animals in dim (15-16 lx, N=10-14) and bright (210-240 lx, N=5-10) light. Percent of total time spent in (A, D) open arms or (B, E) closed arms and (C, F) distance traveled during the task are shown for each light condition (Dim: A-C, Bright: D-F). Total time is a sum of the time spent in open arms, closed arms and the center zone. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. **p<0.01.

134

Table 3-6. Summary of baseline motor and non-motor behaviors in DAT VMAT2 mice. Comparisons are against WT mice unless otherwise denoted.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Gross locomotion Distance traveled Unchanged Unchanged Increased (5X) Number of movements Unchanged Unchanged Increased Horizontal activity Unchanged Unchanged Increased Vertical activity Unchanged Reduced Unchanged Aged locomotion Unchanged Unchanged Increased (2X) Stereotypy Stereotypy count Unchanged Unchanged Trend towards increase Stereotypy number Unchanged Unchanged Higher than VMAT2- kd mice Fine motor skill Challenging beam Increased errors Unchanged Unchanged in traversal similar to Masoud et comparison to WT. al, 2015. Lower than DAT-tg. Latency to cross beam Unchanged Unchanged Trend increase Executive function Puzzle box Unchanged Deficits in problem Unchanged from WT, solving tasks (T5, Enhanced versus T8) VMAT2-kd mice Anxiety - Elevated plus maze Dim light Unchanged Less anxious Less anxious Bright light Unchanged Trend towards Unchanged increased anxiety Locomotion Unchanged Unchanged in dim Unchanged in dim light, reduced in light, reduced in bright light bright light

With regards to motor behavior, DAT-tg/VMAT2-kd are hyperactive and show normal fine motor skills, which were not expected. However, in light of the fact that nigral dopamine cells are intact in these mice, it is unlikely that they would suffer from poor motor performance. Hyperactivity is a unique and robust phenotype of DAT-tg/VMAT2-kd mice that is observed in locomotor activity chambers, but not necessarily in the home cage (qualitative observation).

135

Hence, novelty of the activity chamber may act as a stimulus to induce hyperlocomotion in these animals. In addition, DAT-tg/VMAT2-kd mice show better executive function than VMAT2-kd mice and improved fine motor skills compared to DAT-tg mice, suggesting that altering both DAT and VMAT2 may be beneficial for these behaviors. Aside from voluntary movement, dopamine plays crucial roles in attention, executive function, motivation and reward which may be affecting these behaviors. Interestingly, DAT-tg/VMAT2-kd mice are similar to VMAT2-kd mice in neurochemistry (Figure 3.17, 3.19) and receptor levels (Figure 3.25), however behaviorally, these mice are unique and distinct from VMAT2-kd mice. This suggests that the balance of DAT and VMAT2 is crucial for behavioral outputs. Lastly, DAT-tg mice show higher errors on the challenging beam traversal task as previously reported (Masoud et al., 2015). This suggests that fine motor skill is still affected in these DAT-tg mice, despite lack of dopamine cell loss (Figures 3.22-3.24).

3.2.7 Response to dopaminergic drugs

In the last set of experiments, we investigated the response of DAT VMAT2 mice to different drugs that interact with the dopamine system including 1) psychostimulants, 2) dopamine receptor agonists and 3) the dopamine precursor L-DOPA. These drugs were used as tools to investigate dopaminergic function of DAT VMAT2 mice.

Psychostimulants such as amphetamine, cocaine and methylphenidate have been shown to exert their behavioral effects predominantly by altering dopamine transport mechanisms. In particular, amphetamine increases extracellular levels of dopamine by acting on both DAT and VMAT2, two transporters that have been altered in DAT-tg/VMAT2-kd mice. It has previously been reported that DAT-tg mice show increased amphetamine-induced hyperactivity (Salahpour et al., 2008) while Mooslehner and colleagues demonstrate increased stereotypic behavior in VMAT2- kd mice treated with amphetamine (Mooslehner et al., 2001). Therefore, it was hypothesized that DAT-tg/VMAT2-kd mice will show an even more exaggerated response to amphetamine.

Mice were habituated in the locomotor activity chamber for 60 minutes and then injected with 0.5, 1, or 2 mg/kg of amphetamine. Locomotor activity and stereotypy counts were recorded for 90 minutes post injection. At the lowest dose of 0.5 mg/kg, there was no effect of amphetamine

136 on total distance traveled (Fig 3.32 A) or stereotypy counts (Fig 3.33 A) in WT, DAT-tg and VMAT2-kd mice. However, DAT-tg/VMAT2-kd mice displayed higher locomotor (Fig 3.32 A, D) and stereotypic (Fig 3.33 A, D) activity after drug administration. In response to 1mg/kg of amphetamine, DAT-tg and VMAT2-Kd mice show non-significant enhancements in locomotor activity (Fig 3.32 B) and stereotypy (Fig 3.33 B) while DAT-tg/VMAT2-kd mice display substantial increases in comparison to WT mice (Fig 3.32 D, 3.33 D). Finally, at 2mg/kg of amphetamine, DAT-tg and VMAT2-Kd mice show significantly more locomotor activity (Fig 3.32 C) and stereotypy (Fig 3.33 C) than WT animals, as expected. However, there is no difference in total distance traveled or stereotypy between WT and DAT-tg/VMAT2-kd mice (Figure 3.32 D, 3.33 D). At this dose, DAT-tg/VMAT2-kd mice displayed peculiar movements such as tremors, jerking, backward locomotion and tongue protrusions which were measured using the Abnormal Involuntary Movements (AIMs) scale. An AIMs score is assigned based on the duration of abnormal movements that are divided into 4 categories: locomotor, axial, limb or orolingual. The minimum score for each time point is 4 and the maximum score is 16 (according to the 4 categories). As shown in Fig 3.34 A and B, only DAT-tg/VMAT2-kd mice display severe AIMs in response to 2 mg/kg of amphetamine. The lack of locomotor activity in these mice at 2mg/kg of amphetamine is due to their abnormal movements which do not permit normal locomotion. Therefore, DAT-tg/VMAT2-kd mice display an inverse relationship between total distance traveled and AIMs score, where at peak AIMs scores, locomotion is negligible and when AIMs begin to subside about an hour post-injection, locomotor activity also begins to resume (Fig 3.34 C).

137

Figure 3-32. Amphetamine-induced locomotion. Mice were habituated to the chamber for 60 minutes, injected with amphetamine and monitored for an additional 90 minutes (N=8-12). Arrow denotes time of injection. Distance traveled over time in response to (A) 0.5, (B) 1 and (C) 2 mg/kg of amphetamine. (D) Sum of distance traveled post injection for all doses. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05, **p<0.01, p<0.001.

138

Figure 3-33. Amphetamine-induced stereotypy. Mice were habituated to the chamber for 60 minutes, injected with amphetamine and monitored for an additional 90 minutes (N=8-12). Stereotypy counts over time in response to (A) 0.5, (B) 1 and (C) 2 mg/kg of amphetamine. Arrow denotes time of injection. (D) Sum of stereotypy counts post injection for all doses. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05, **p<0.01, p<0.001.

139

Figure 3-34. Abnormal involuntary movements (AIM) induced by 2 mg/kg of amphetamine. Animals were habituated to the chamber for 60 minutes and then injected with drug (N=5-7). Arrow denotes time of injection. Behavior of animals was recorded for 1 minute immediately before injection and in 15 minute intervals up to 75 minutes post-injection. Hence, a total of 6 recordings were scored manually on the AIMs scale on 4 categories: locomotor, orolingual, axial and limb peculiarities. For each time point, the minimum score is 4 and the maximum is 16. A score is assigned based on duration of abnormal movements during the test period. (A) AIMs score over time. (B) Total AIMs score (sum of all time points). The dashed line represents the minimum AIMs score. (C) Relationship between AIMs score and distance traveled by DAT- tg/VMAT2-kd mice over time. Locomotion and AIMs were assessed in different cohorts of mice due to technical limitations of videotaping animals while in the locomotor activity chamber. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05, **p<0.01, p<0.001.

140

Collectively, these results demonstrate that DAT-tg/VMAT2-kd mice are exquisitely sensitive to the effects of amphetamine, such that they display AIMs at doses that normally evoke hyperactivity in other genotypes. To further explore this behavior, we tested whether higher doses of amphetamine would elicit similar abnormal movements in other genotypes. In a pilot experiment, after injection of 5mg/kg amphetamine, WT mice were hyperactive whereas DAT-tg mice displayed a short burst of activity followed by decreased locomotion (Fig 3.35). Qualitatively, during the periods of reduced locomotor activity, these DAT-tg animals displayed abnormal behavior reminiscent of DAT-tg/VMAT2-kd mice treated with 2mg/kg of amphetamine (data not quantified). For comparison purposes, a single DAT-tg/VMAT2-kd mouse was also treated with 5mg/kg of amphetamine and it displayed lack of locomotor activity (Fig 3.35) and several abnormal movements, as expected. Although the abnormal movements appeared to be more extreme in DAT-tg/VMAT2-kd mice, this experiment demonstrates that at high enough doses, other genotypes can also develop this behavior in response to amphetamine.

Figure 3-35. Locomotor effect of 5 mg/kg amphetamine on WT and DAT-tg mice. WT and DAT-tg mice were habituated to the chamber for 60 minutes, injected with 5mg/kg amphetamine and monitored for an additional 90 minutes (N=4-5). A single DAT-tg/VMAT2-kd mouse was also included for comparison (dashed red line). (A) Distance traveled over time. Arrow denotes time of injection. (B) Sum of distance traveled post-injection. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05, **p<0.01, p<0.001.

141

After assessing the effects of amphetamine, which causes reversal of DAT, the response of DAT VMAT2 mice to DAT inhibitors, cocaine and methylphenidate, was investigated. Both these compounds are psychostimulants that cause behavioral hyperactivity in rodents. It was previously shown that DAT-tg mice are similar to WT animals in their behavioral response to cocaine, while they are hypersensitive to the effects of amphetamine (Salahpour et al., 2008). Conversely, VMAT2-het mice displayed increased horizontal activity versus WT animals when treated with 20 mg/kg of cocaine (Wang et al., 1997). Given these previously published findings, it was predicted that DAT-tg mice will not display increased sensitivity to cocaine while VMAT2-kd mice will. As shown in Figure 3.36, when treated with 20 mg/kg of cocaine, 1) DAT-tg animals displayed a moderate increase in activity that was similar to WT mice and 2) VMAT2-kd mice demonstrated significant hyperactivity, as expected (Fig 3.36). Furthermore, DAT-tg/VMAT2-kd mice were also hyperactive in comparison to WT mice, however, they behaved similarly to VMAT2-kd mice as demonstrated in Figure 3.36 B. Analogous behavioral results were observed in response to another DAT inhibitor, methylphenidate (5 mg/kg). Similar to cocaine, both VMAT2-kd and DAT-tg/VMAT2-kd mice were hyperactive when treated with methylphenidate (Fig 3.36 D). DAT-tg/VMAT2-kd mice appeared to demonstrate a slightly greater response than VMAT2-kd mice, however, the difference between the two genotypes was not significant (Fig 3.36 D). These results suggest that unlike a DAT-reversing agent like amphetamine which produces uniquely enhanced sensitivity in DAT-tg/VMAT2-kd mice (Fig 3.32 D, 3.33 D), DAT inhibitors induce hyperactivity to the same extent in both VMAT2-kd and DAT-tg/VMAT2-kd mice. This suggests that knockdown of VMAT2 is particularly relevant for locomotor response to DAT inhibition, whereas over-expression of DAT does not particularly affect this phenotype. It is important to note that amphetamine causes release of dopamine via DAT while cocaine and methylphenidate prevent the uptake of dopamine following activity- dependent vesicular release. This mechanistic difference may underlie the behavioral differences described above.

142

Figure 3-36. Locomotion induced by DAT inhibitors, cocaine and methylphenidate. Mice were habituated to the chamber for 60 minutes, injected with cocaine (N=6) or methylphenidate (N=4-8) and monitored for an additional 90 minutes. Distance traveled over time in response to (A) cocaine or (C) methylphenidate. Sum of total distance traveled after administration of (B) cocaine or (D) methylphenidate. Arrow denotes time of injection. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05, **p<0.01, p<0.001.

Next, the effect of dopamine receptor agonists on DAT VMAT2 mice was assessed as an indirect measure of receptor function. Similar to previous studies, animals were habituated to the activity chamber for 60 minutes, drug was injected and locomotor activity was monitored for 90 minutes. First, the non-selective D1/D2 receptor agonist, apomorphine was used. It has previously been

143 shown that in rodents, administration of apomorphine produces stereotypy, which is regarded as a readout of striatal dopamine receptor activation (Protais et al., 1976). Hence, we assessed stereotypic behavior (e.g. repetitive chewing/grooming) of DAT VMAT2 mice in response to 2 mg/kg of apomorphine (Fig 3.37). Qualitatively, following drug administration, some DAT- tg/VMAT2-kd mice engaged in extreme biting behavior that results in self-injury and bleeding which is uncommon at this dose. Quantitatively, this behavior was partially represented as increased stereotypy counts in DAT-tg/VMAT2-kd mice compared to WT animals after administration of apomorphine (Fig 3.37). However, other quantitative scales such as the AIM scale may have been better suited to fully capture this behavior. VMAT2-kd mice also demonstrate a non-significant trend towards higher stereotypy when treated with apomorphine. In summary, increased apomorphine-induced stereotypy in DAT-tg/VMAT2-kd mice suggests that receptor function is up-regulated, corroborating previous radioligand binding assessments of D2 receptor levels (Fig 3.25) in these mice.

Figure 3-37. Apomorphine-induced stereotypy. Mice were habituated to the chamber for 60 minutes, injected with 2mg/kg of apomorphine s.c. and monitored for an additional 90 minutes (N=8-9). Sum of stereotypy counts over 90 minutes following injection of apomorphine. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05.

144

Second, we evaluated the effect of a selective D1 agonist, SKF81297, on the locomotor activity of DAT VMAT2 mice compared to saline treatment. As shown in Figure 3.38 B, VMAT2-kd and DAT-tg/VMAT2-kd mice seem to show higher locomotor activity than WT mice at certain time points after SKF-81297 administration (see time course 105-135 minutes). However, when total activity of saline-treated DAT-tg/VMAT2-kd mice is compared to SKF81297-treated DAT- tg/VMAT2-kd mice, there are no statistical differences (Fig 3.38 D). Similarly, when the dopamine precursor and anti-Parkinsonian drug, L-DOPA (25 mg/kg) was tested in DAT- tg/VMAT2-kd mice, no differences were noted in total distance traveled compared to saline treatment (Figure 3.38 D). Since DAT-tg/VMAT2-kd mice are basally hyperactive, it translates to greater distance traveled even after saline injections (Fig 3.38 A). Due to increased locomotion of saline-treated DAT-tg/VMAT2-kd mice, the effects of SKF-81297 and L-DOPA are masked in these animals at the doses tested. Unlike amphetamine (Fig 3.32, 3.33), cocaine and methylphenidate (Fig 3.36), which produce substantial increases in the locomotor activity of DAT-tg/VMAT2-kd mice that surpasses their basal hyperactivity, the effects of SKF-81297 and L-DOPA are more modest and therefore, difficult to distinguish from the spontaneous activity of DAT-tg/VMAT2-kd mice in novel environments (Fig 3.38 B). It is interesting to note that in DAT-tg/VMAT2-kd mice, the time course of activity following saline treatment (Fig 3.38 A, downward sloping) appears different from SKF-81297 or L-DOPA treatment (Fig 3.38 B, C; peak after injection), suggesting some behavioral effects of these drugs (versus saline) that are not fully captured in the quantitative analysis (Fig 3.38 D).

145

146

Figure 3-38. Effect of SKF 81297, L-DOPA and saline on locomotor activity of DAT VMAT2 mice. Mice were habituated to the activity chamber for 60 minutes and injected with either (A) saline (0.9% NaCl), (B) 2 mg/kg of SKF-81297 or (C) 25 mg/kg L-DOPA with 12.5 mg/kg benserazide (peripheral DOPA decarboxylase blocker). Distance traveled post-injection was monitored for an additional 90 minutes. Different cohorts of mice were used for each drug (N=7-11). Figures A-C show distance traveled over time where the arrow denotes time of injection. Data are presented as mean ± SEM. Statistical differences are in comparison to WT mice. (D) Sum of total distance traveled after drug administration. For each genotype, the effect of a drug is compared to saline administration. Statistical differences are reported against saline treatment. Data are presented as mean ± SEM. *p<0.05, **p<0.01, ***p<0.001.

Table 3-7. Summarized effects of dopaminergic drugs on behavior of DAT VMAT2 mice. Comparisons are against WT mice, unless otherwise denoted.

Experiment DAT-tg VMAT2-kd DAT-tg/VMAT2-kd Saline No change No change Increased (basal Locomotor hyperactivity) Psychostimulants Amphetamine Increased Increased Increased locomotion Locomotor locomotion and locomotion and and stereotypy at 0.5 Stereotypy stereotypy at stereotypy at and 1 mg/kg. AIMs at AIMs 2mg/kg (similar to 2mg/kg 2mg/kg Salahpour et al, 2008) Cocaine No change (similar Increased Increased (similar to Locomotor to (Salahpour et al., VMAT2-kd mice) 2008)

Methylphenidate No change (similar Increased Increased (similar to to (Salahpour et al., VMAT2-kd mice) 2008) Dopamine receptor agonists Apomorphine No change No change Increased Stereotypy SKF-81297 No change Increased Similar increase as Locomotor saline Dopamine precursor LDOPA No change No change Similar increase as Locomotor saline

147

In summary, VMAT2/DAT-tg mice are exquisitely sensitive to the effects of amphetamine at very low doses which do not affect other genotypes (e.g. 0.5mg/kg). While we expected these animals to show a left-shifted dose response curve, the development of abnormal involuntary movements (jerking, tremor, orolingual) at a dose of 2mg/kg of amphetamine was not anticipated. DAT-tg and VMAT2-kd mice also showed enhanced locomotor and stereotypic response to amphetamine in comparison to WT mice, as expected, however the effect in DAT- tg/VMAT2-kd mice was synergistic. Unlike amphetamine, DAT inhibitors like cocaine and methylphenidate, produced significant hyperactivity in both VMAT2-kd and DAT-tg/VMAT2- kd mice. Both of these genotypes demonstrate higher dopamine receptor levels which may explain their locomotor response to DAT inhibitors. The non-selective dopamine receptor agonist, apomorphine, produced increased stereotypy in DAT-tg/VMAT2-kd mice, also indicating enhanced receptor function in these mice. However, locomotor effects of the D1 agonist, SKF-81297 and the dopamine precursor, L-DOPA, could not be distinguished from saline treatment in DAT-tg/VMAT2-kd mice due to their basal hyperactivity. It should be noted that qualitatively, the time courses for SKF-81297 and L-DOPA induced locomotion in DAT- tg/VMAT2-kd mice show distinct patterns (Fig 3.38 B -C, peak after injection) in comparison to saline injection (Fig. 3.38 A, continuous downward trend). Collectively, DAT-tg/VMAT2-kd mice display a selective and robust response to amphetamine, indicating intense behavioral sensitivity to DAT reversal.

148

Chapter 4 Discussion Discussion

4.1 Project 1: Characterization of DAT-tg mice

In this study, we report that over-expression of DAT is capable of triggering oxidative stress, dopamine neuron loss and L-DOPA reversible motor deficits in DAT-tg mice. Previously, ectopic expression of DAT was shown to cause death of non-dopaminergic cells, presumably due to their inability to properly handle cytotoxic dopamine (Chen et al., 2008). However, we demonstrate that even in dopamine cells that are inherently equipped with the molecular machinery to properly store, metabolize and release dopamine, an increase in DAT expression can lead to higher dopamine uptake and damaging consequences. Aside from our work, previous studies using plasmid and lentiviral techniques have also reported that DAT over-expression can increase dopamine uptake and alter downstream behaviors (Martres et al., 1998; Adriani et al., 2009). Transgenic mice expressing DAT under the TH promoter showed higher DAT levels, greater dopamine uptake and modest, but significant reductions in striatal dopamine tissue content (Donovan et al., 1999), similar to DAT-tg mice. In comparison to these studies, our BAC transgenic approach to over-express DAT has several important advantages including: 1) robust, long-term DAT expression, 2) selectivity for dopaminergic neurons using the DAT promoter and 3) lack of injection and transfection-related complications. Collectively, this body of work shows that increased DAT activity can significantly impact and change dopamine homeostasis.

The dopamine system is notoriously sensitive to endogenous and exogenous challenges (Langston et al., 1983; Hastings et al., 1996; Mosharov et al., 2009). Therefore, 46% higher dopamine uptake in DAT-tg mice (Salahpour et al., 2008) produces dramatic effects on dopamine homeostasis, cell survival, oxidative stress and motor behaviors, as noted in our study (Masoud et al., 2015). These results highlight the physiological importance of tightly regulating cytosolic dopamine levels since moderate deviations in dopamine compartmentalization can directly impact neuronal survivability. Another example of this is the VMAT2-kd mice. These animals express only 5% of normal VMAT2 protein and display decreased dopamine tissue content, nigrostriatal neurodegeneration and increased levels of cysteinyl-catechols (Caudle et al., 2007), similar to DAT-tg mice. Physiologically, VMAT2-kd mice are deficient in

149 sequestering intracellular dopamine into vesicles while DAT-tg mice have excess dopamine uptake (Caudle et al., 2007; Salahpour et al., 2008). In addition to higher uptake, DAT-tg mice also display reduced VMAT2 expression, suggesting that vesicular storage of dopamine could also be compromised. Taken together, the genetic manipulations in VMAT2-kd and DAT-tg mice effectively act to increase the cytosolic pool of dopamine. This buildup of cytosolic dopamine could be a common pathway that is responsible for the basal loss of dopamine neurons and oxidative stress evident in both VMAT2-kd and DAT-tg mice.

There are several observations supporting the hypothesis that accumulation of cytosolic dopamine results in loss of dopaminergic neurons in DAT-tg mice. First, results from DAT-KO animals highlight the critical role of DAT in loading the presynaptic neuron with dopamine (Giros et al., 1996; Jones et al., 1998a) In DAT-KO mice, lack of uptake leads to 5-times higher extracellular dopamine levels and extremely low dopamine tissue content (5%), indicating depleted intracellular stores. Conversely, in DAT-tg mice, higher levels of functional DAT lead to a 46% increase in dopamine uptake and a 40% decrease in extracellular dopamine, suggesting that the neurotransmitter is accumulating in the presynaptic neuron (Salahpour et al., 2008). However, despite the likely buildup of dopamine within each dopaminergic cell, DAT-tg mice display a 33% reduction in overall dopamine tissue content as a direct consequence of 30-36% loss of dopamine neurons. Secondly, we report higher metabolite-to-dopamine ratios in DAT-tg mice. Since DOPAC is a direct product of cytosolic dopamine metabolism, a 60% increase in the DOPAC/dopamine ratio could indicate that a greater proportion of dopamine is present in the cytosol and not sequestered into vesicles (Di Monte et al., 1996). Elevated metabolite-to- dopamine ratios also imply enhanced dopamine turnover that could be a compensatory mechanism to tackle the buildup of intracellular dopamine (Zigmond et al., 2002). Thirdly, increased levels of 5-S-cysteinyl-dopamine and 5-S-cysteinyl-DOPAC were detected in the striatum of DAT-tg mice. These cysteinyl-modified adducts have been suggested to arise from the oxidation of cytosolic dopamine and its metabolites (Graham, 1978; Fornstedt and Carlsson, 1989; Hastings and Zigmond, 1994). Not only are cysteinyl adducts a direct consequence of cytosolic dopamine reactivity, they are also capable of independently inducing further neuronal damage (Spencer et al., 2002). Next, lower VMAT2 protein expression in DAT-tg mice also suggests potential buildup of cytosolic dopamine. Although this decrease may be a reflection of dopaminergic cell loss per se, nonetheless, reduced VMAT2 levels can negatively impact

150 vesicular storage and disable these mice from handling increased dopamine uptake caused by DAT over-expression. Lastly, accumulation of cytosolic dopamine has been suggested to have deleterious effects on cell survival (Caudle et al., 2007; Chen et al., 2008; Mosharov et al., 2009) that is clearly reflected in the loss of dopamine neurons in DAT-tg mice (Masoud et al., 2015). Collectively, these observations suggest that DAT over-expression most likely leads to high cytosolic levels of dopamine, thereby producing the downstream detrimental effects observed in DAT-tg mice.

We also demonstrated that DAT-tg mice are highly sensitive to MPTP-induced neurotoxicity. Indeed, when treated with MPTP, DAT-tg mice showed greater reductions in striatal TH levels and dopamine tissue content compared to WT animals. MPP+, the toxic metabolite of MPTP, is a substrate for DAT and therefore, causes selective damage to dopaminergic cells (Langston et al., 1984; Chiba et al., 1985; Ramsay et al., 1986; Gainetdinov et al., 1997; Schober, 2004). While the dependence of MPTP neurotoxicity on DAT function has previously been demonstrated (Gainetdinov et al., 1997; Bezard et al., 1999; Miller et al., 1999b; Schober, 2004), our results indicate a synergistic interaction between environmental and genetic risk factors that could have broader implications for complex pathological conditions such as Parkinson’s disease (Cannon and Greenamyre, 2013). In Parkinson’s disease, both genetic mutations and environmental conditions have been documented to increase disease risk (Priyadarshi et al., 2000; Hardy et al., 2006; Martin et al., 2011; Cannon and Greenamyre, 2013). Moreover, animal models that depend on a single type of insult seldom recapitulate the full spectrum of the disorder (Beal, 2010). Although genes such as PINK1, DJ1 and PARK2 (parkin) have been implicated in familial forms of Parkinson’s disease, mutating or knocking-out these genes in most animal models does not reproduce dopaminergic cell loss (Goldberg et al., 2003; Yamaguchi and Shen, 2007; Gispert et al., 2009). Conversely, while acute toxicant treatment (e.g. MPTP or 6-hydroxydopamine) can produce abrupt neurodegeneration, it does not address the underlying disease mechanism of a chronic and progressive disorder like Parkinson’s disease (Schober, 2004). Given the shortcomings of these individual approaches, the convergence of genetic as well as environmental insults may be more representative of idiopathic Parkinson’s disease that is hypothesized to arise from multiple hits (Sulzer, 2007; Cannon and Greenamyre, 2013). Our results lend support to this idea by showing that genetic over-expression of DAT combined with exogenous exposure to MPTP, aggravates toxicity to dopamine neurons.

151

Although the effect of genetic mutations on DAT expression is unclear in humans, a correlation study reports that DAT genetic variants in combination with exposure to exogenous compounds (e.g. pesticides) can potentiate the risk of developing Parkinson’s disease by 3- or 4-fold (Ritz et al., 2009). This highlights the significance of genetic and environmental interactions in the pathology of Parkinson’s disease.

The cellular, neurochemical and behavioral changes observed in DAT-tg mice recapitulate important features of Parkinson’s disease. Firstly, loss of midbrain dopamine neurons and reduced dopamine tissue content in the striatum of DAT-tg mice capture the major pathological characteristics of Parkinson’s disease (Dauer and Przedborski, 2003). However, it should be noted that Parkinson’s disease is characterized by selective nigrostriatal degeneration, whereas DAT-tg mice also demonstrate loss of VTA dopamine neurons. This is probably due to transgenic over-expression of DAT in the VTA, which enhances the vulnerability of this region in DAT-tg mice. Physiologically, VTA neurons do not express as much DAT as SNpc neurons and therefore, the VTA is relatively spared from damage in Parkinson’s disease (Blanchard et al., 1994). The relationship between DAT expression and neurodegeneration is supported by a study in Parkinson’s disease patients showing that brain regions containing the highest levels of DAT protein – the caudate and putamen – are also the most sensitive to damage (Miller et al., 1997). In addition, a recent meta-analysis has identified the DAT gene as a risk factor for Parkinson’s disease in certain populations (Zhai et al., 2014). Secondly, oxidative stress has long been postulated to be involved in the development of Parkinson’s disease (Fahn and Cohen, 1992) and we report that DAT-tg mice display increased levels of cysteinyl-dopamine and cysteinyl- DOPAC, two markers that are also elevated in the SN of Parkinson’s disease patients (Spencer et al., 1998). Thirdly, increased dopamine turnover in the transgenic mice mirrors elevated metabolite-to-dopamine ratios that have been reported in Parkinson’s disease patients (Birkmayer and Hornykiewicz, 1962; Zigmond et al., 2002). In addition, both DAT-tg mice and Parkinson’s disease patients show reductions in VMAT2 protein expression in comparison to control samples (Miller et al., 1999a). Behaviorally, DAT-tg mice do not exhibit any deficits in gross locomotion, probably because the level of cell loss in these animals is not sufficient to cause major motor disturbances. In Parkinson’s disease patients, motor deficits are only evident when greater than 70% of dopaminergic tone is lost in the striatum (Bernheimer et al., 1973). However, results from the wire-hang test and challenging beam traversal task clearly

152 demonstrate that fine motor coordination, balance and strength are compromised in DAT-tg mice similar to Parkinson’s disease patients. Other studies on dopaminergic dysfunction have shown that these two tests are sensitive to motor impairment even in the absence of gross locomotor changes (Hwang et al., 2005; Luk et al., 2012). Furthermore, not only do DAT-tg mice display motor disturbances on the challenging beam traversal; these deficits are also reversed by L- DOPA, the principal treatment for motor symptoms of Parkinson’s disease. This suggests that dopamine neuronal loss in DAT-tg mice leads to motor deficits that can be reversed by restoring dopaminergic tone. Hence, parallel to Parkinson’s disease patients, DAT-tg mice also demonstrate motor behaviors that are responsive to L-DOPA treatment. Given these overlapping results, we postulate that the mishandling of cytosolic dopamine exhibited by DAT-tg mice could provide important insights on the unique vulnerability of dopamine cells in Parkinson’s disease.

In conclusion, in Project 1, we used transgenic mice that selectively over-express DAT in dopaminergic neurons to investigate the effects of cytosolic dopamine accumulation in vivo. As shown by our results, moderate increases in DAT function cause spontaneous dopaminergic cell loss, oxidative stress and fine motor impairment that is reversed by L-DOPA treatment. These results suggest that the integrity of dopamine neurons depends heavily on the ability of DAT to maintain proper homeostatic control of presynaptic dopamine. Since dopaminergic cells are selectively damaged by a broad variety of genetic and environmental insults, it demonstrates that these cells are inherently at risk. Our results imply that buildup of cytosolic dopamine, a highly reactive and potentially toxic molecule, may underlie the cell-specific vulnerability of dopaminergic neurons to damage. We propose that dopamine uptake through DAT, maintains a constant cytosolic pool of this neurotransmitter that can propagate oxidative stress in dopamine cells. This type of chronic damage may render these neurons vulnerable to degeneration, especially if coupled with other genetic or environmental insults that are linked with the pathogenesis of Parkinson’s disease. Since DAT-tg mice display spontaneous neuronal loss and heightened toxicity in response to MPTP, these mice provide a useful tool to study the effects of endogenous and exogenous challenges on dopamine cells.

153

4.2 Project 2: Characterization of DAT-tg/VMAT2-kd mice

The goal of this project was to investigate the dual effect of DAT over-expression and VMAT2 knockdown in genetically modified mice. As an extension of Project 1, the rationale for this study also arose from a body of literature showing that cytosolic dopamine is highly reactive and can produce neurotoxicity by triggering oxidative stress (Graham, 1978; Chen et al., 2008; Mosharov et al., 2009). In particular, as demonstrated in Project 1, over-expression of DAT and greater dopamine uptake produces damaging outcomes in DAT-tg mice including oxidative stress, spontaneous loss of midbrain dopamine neurons and fine motor deficits (Masoud et al., 2015). In addition, VMAT2-kd mice from the Miller lab at Emory University, exhibit reduced vesicular storage of dopamine which also translates to dopaminergic damage as assessed by diminished striatal dopamine, evidence of oxidative stress and loss of nigrostriatal dopamine neurons in aged mice (Caudle et al., 2007). In both these mouse models, potential accumulation of cytosolic dopamine due to increased uptake or decreased vesicular packaging, leads to deleterious consequences. However, the level of dopaminergic cell loss in these mouse models is moderate (around 30% for each genotype) and does not reach the extent of damage that is typically observed in Parkinson’s disease patients (around 70%) (Caudle et al., 2007; Sulzer, 2007; Kordower et al., 2013; Masoud et al., 2015). Therefore, stemming from the results of Project 1 (DAT-tg mice) and in collaboration with the Miller lab (VMAT2-kd mice), we interbred DAT-tg and VMAT2-kd mice to generate double transgenic animals (DAT- tg/VMAT2-kd mice) that were hypothesized to have greater accumulation of cytosolic dopamine and consequently, demonstrate exacerbated symptoms of dopaminergic damage that may better resemble the pathophysiology of Parkinson’s disease.

We systematically characterized the dopamine system of DAT-tg/VMAT2-kd mice by evaluating presynaptic dopamine homeostasis, survival of dopamine neurons, post-synaptic dopamine receptors, basal dopamine-mediated behaviors and behavioral response to dopaminergic drugs. We hypothesized that DAT-tg/VMAT2-kd mice will show phenotypes associated with dopamine toxicity such as loss of dopamine cells, reduced dopamine tissue content, upregulation of dopamine receptors and poor motor behavior. While dopaminergic cell loss was not observed, some phenotypes of dopamine dysregulation were recapitulated in DAT- tg/VMAT2-kd mice. For instance, in comparison to WT animals, DAT-tg/VMAT2-kd mice display: 1) 95% reduction in striatal dopamine tissue content, 2) 85% reduction in evoked-

154 dopamine release, 3) 15- to 20-fold increase in DOPAC/dopamine and HVA/dopamine ratios, suggesting higher dopamine turnover, 4) 33% increase in striatal D2 receptor levels, indicating compensatory up-regulation of this receptor to tackle reduced dopaminergic tone, 5) 46% reduction in survival of 12-week old mice and 6) decreased adult body weight. Taken together, these findings illustrate that DAT-tg/VMAT2-kd mice have considerable impairments as a result of mishandling dopamine. However, some of these changes were not as pronounced as hypothesized, and more importantly, DAT-tg/VMAT2-kd mice did not show any signs of dopaminergic cell loss or compromised motor ability, two hallmarks of Parkinson’s disease. The lack of Parkinsonian phenotypes in DAT-tg/VMAT2-kd mice is related to a complication that arose during this study where DAT-tg mice (that were used to breed DAT-tg/VMAT2-kd mice), no longer displayed some previously-observed phenotypes of dopaminergic toxicity. This issue is discussed in detail in the next section. The following discussion focuses on current findings from DAT-tg/VMAT2-kd mice.

It is interesting to note that for most biochemical and neurochemical assessments, DAT- tg/VMAT2-kd mice displayed similar results as VMAT2-kd mice. Indeed, on pre- and post- synaptic measures of the dopamine system (such as dopamine tissue content, release, uptake, metabolite-to-dopamine ratios and receptor binding), VMAT2-kd and DAT-tg/VMAT2-kd mice were generally indistinguishable. In animals with VMAT2 knockdown (regardless of DAT expression), the 95% reduction in VMAT2 levels severely compromised their vesicular storage capacity as illustrated by drastic reductions in dopamine tissue content and evoked-dopamine release. Due to the diminished dopaminergic tone, dopamine receptors were also up-regulated in VMAT2-kd and DAT-tg/VMAT2-kd mice to compensate for reduced dopamine release. In addition, VMAT2-kd and DAT-tg/VMAT2-kd mice demonstrated increased metabolite-to- dopamine ratios, suggesting that dopamine was being metabolized in the cytosolic space as a consequence of reduced vesicular storage. The impact of VMAT2 knockdown on dopamine tissue content, evoked dopamine release and metabolite-to-dopamine ratios were so robust, that they masked any additional contribution of DAT over-expression in DAT-tg/VMAT2-kd mice, probably due to ceiling or basement effects. For instance, DOPAC/dopamine and HVA/dopamine ratios are enhanced by 18 to 23-fold in VMAT2-kd mice, potentially reaching a ceiling effect beyond which further increase in dopamine metabolism may not be physiologically feasible or technically detectable in DAT-tg/VMAT2-kd mice. In comparison to WT animals,

155

DAT-tg mice also showed moderate reductions in dopamine tissue content and release, and trends towards faster dopamine uptake, suggesting that over-expression of DAT translates to functional changes in the dopamine system of DAT-tg mice. However, the consequences of DAT over-expression in animals with reduced VMAT2 levels were difficult to perceive perhaps due to more modest effects. Between the two genetic manipulations, DAT over-expression and VMAT2 knockdown, the latter produced a stronger impact due to multiple reasons: 1) VMAT2 is essential for survival since ablation of this gene results in post-natal death whereas DAT-KO mice survive into adulthood, 2) VMAT2 levels were drastically reduced to 5% of normal WT levels, whereas DAT expression was enhanced by 50-75% and 3) VMAT2 is expressed in all monoaminergic cells and therefore, VMAT2-knockdown affected multiple neurotransmitter systems whereas DAT over-expression was confined to the dopaminergic system. Hence, on measures that are heavily influenced by vesicular storage capacity (such as striatal dopamine tissue content and electrically-evoked dopamine release), knockdown of VMAT2 produced similar effects in both VMAT2-kd and DAT-tg/VMAT2-kd mice, regardless of their DAT expression.

However, DAT-tg/VMAT2-kd mice also display unique behavioral phenotypes that distinguish them from other mice. First, they are hyperactive in open-field locomotion as evidenced by ~5- fold greater distance traveled, higher horizontal activity and number of horizontal movements than WT, DAT-tg and VMAT2-kd mice. Interestingly, genetic ablation of DAT also produces hyperactivity in mice (Giros et al., 1996). However, DAT-KO mice show impaired habituation which is not evident in DAT-tg/VMAT2-kd mice. Furthermore, during other short tests such as the elevated plus maze, puzzle box or challenging beam traversal, DAT-tg/VMAT2-kd mice do not appear hyperactive and can solve the task at hand, suggesting that their hyper- exploratory behavior is context-dependent. Second, although DAT-tg mice display fine motor deficits on the challenging beam traversal task, these impairments are reversed in DAT-tg/VMAT2-kd mice. This suggests that reducing VMAT2 levels in DAT-tg mice improves their fine motor skills. Third, in the puzzle box, while VMAT2-kd mice perform worse than all other genotypes during problem-solving tasks (T5, T8), these deficits are rescued in DAT-tg/VMAT2-kd mice. This suggests that increasing DAT expression in VMAT2-kd mice ameliorates their cognitive deficits. Collectively, judging from these behavioral results, it seems that altering the balance of DAT and VMAT2 produces hyperactivity in novel environments and improves fine-motor skill and

156 executive function of DAT-tg/VMAT2-kd mice compared to their counterparts. While these striking results were not expected, they highlight the complexity of the dopamine system in determining behavioral outputs. Previous assessments of dopamine tissue content, release, uptake and receptor levels suggested that DAT-tg/VMAT2-kd mice closely resemble VMAT2- kd animals, however, their behavioral outcomes are distinctly unique. While up-regulated striatal dopamine receptors in DAT-tg/VMAT2-kd mice could partially explain their hyperactivity, VMAT2-kd mice also express similar neurochemical and receptor changes but do not exhibit increased locomotion, indicating a paradox. This disconnect between neurochemical and behavioral analyses raises an important issue in the study of mouse models. Although neurochemical analyses offer significant value in understanding aspects of the underlying system, they are limited by several factors: 1) they are conducted in specific tissues, brain regions or slices, 2) they are post-mortem examinations, and 3) typically, they capture a snapshot of the system at a particular time point. Conversely, behavioral assessments occur in intact, living organisms and measure outcomes that are integrated from multiple neural systems. Hence, isolated biochemical/neurochemical assessments may not adequately explain behavioral changes that arise from complex interactions between several pathways. Indeed, dopamine plays important roles in a variety of functions including motor control, attention, cognition, motivation and reward; all of which can contribute to the behaviors tested.

In addition to changes in baseline behaviors, DAT-tg/VMAT2-kd mice also display differential locomotor responses to drugs that interact with the dopamine system. Most notably, DAT- tg/VMAT2-kd mice are exquisitely sensitive to the effects of amphetamine, a psychostimulant that reverses the activity of DAT, ultimately causing release of dopamine. At very low doses of amphetamine (0.5 mg/kg), DAT-tg/VMAT2-kd mice display significantly increased locomotion and stereotypy while at higher doses (2 mg/kg), these animals demonstrate abnormal involuntary movements – giving rise to a left-shifted dose-response curve. Both DAT-tg and VMAT2-kd mice also depict heightened sensitivity to the stimulant effects of amphetamine, albeit to a lesser extent than DAT-tg/VMAT2-kd mice. Hence, animals with concurrent DAT over-expression and VMAT2 knockdown exhibit a truly synergistic and robust response to amphetamine. Conversely, DAT inhibitors, such as cocaine or methylphenidate, produce increased locomotor responses in both DAT-tg/VMAT2-kd and VMAT2-kd mice, indicating a lack of selectivity for the double transgenic animals. Cocaine-induced behavioral effects cannot be explained by dopamine

157 release, as indicated by preliminary FSCV experiments conducted in the presence of cocaine (Appendix 1). However, the differential response of DAT VMAT2 mice to amphetamine versus cocaine/methylphenidate may be explained by the drugs’ mechanisms of action. In the case of DAT inhibitors, dopamine is released through normal activity-dependent vesicular mechanisms but its uptake is blocked, allowing the extracellular neurotransmitter to interact with post- synaptic receptors. Since basal evoked-dopamine release and dopamine receptor levels are similarly altered in VMAT2-kd and DAT-tg/VMAT2-kd mice, this could explain their similar behavioral responses to cocaine and methylphenidate. In addition, previous characterization of DAT-tg mice revealed that they show no differences in comparison to WT mice when treated with DAT blockers (Salahpour et al., 2008). This suggests that drug-induced blockade of dopamine uptake is not affected by DAT over-expression, therefore, DAT-tg/VMAT2-kd mice also effectively behave as VMAT2-kd animals when treated with DAT blockers such as cocaine or methylphenidate. Amphetamine, on the other hand, acts by dissipating the vesicular proton gradient which forces dopamine to accumulate in the cytosolic space. This buildup of cytosolic dopamine along with the actions of amphetamine, reverse the activity of DAT, causing non- vesicular dopamine efflux (Sulzer et al., 1995, 2005). Physiologically, DAT-tg/VMAT2-kd mice are anticipated to accumulate more dopamine in the cytosolic space than DAT-tg or VMAT2-kd mice alone, due to the dual effect of greater dopamine uptake combined with reduced vesicular storage. If indeed cytosolic dopamine levels are higher in DAT-tg/VMAT2-kd mice, then amphetamine-induced reversal of DAT would also release larger amounts of the neurotransmitter in the extracellular space in comparison to other genotypes. This enhanced DAT-mediated dopamine release in DAT-tg/VMAT2-kd mice could explain their robust locomotor response to amphetamine.

In Project 2, by simultaneously altering DAT and VMAT2 levels in DAT-tg/VMAT2-kd mice, we created an imbalance in dopamine compartmentalization that produced unique and unexpected phenotypes. As mentioned, DAT-tg/VMAT2-kd mice are hyperactive, perform better than other genotypes on fine motor skill and executive function tasks, and are highly responsive to amphetamine treatment. Although we measured several aspects of pre and post-synaptic dopamine dynamics including dopamine tissue content, dopamine release and uptake, dopamine receptor levels, number of midbrain dopaminergic neurons and metabolite-to-dopamine ratios, these parameters cannot fully explain the behavioral results obtained. One possible explanation

158 for these results is that simultaneously increasing DAT and reducing VMAT2 levels produces a buildup of cytosolic dopamine that eventually reverses DAT activity, causing dopamine to leak out of the cell. This idea is supported by previous studies that provide evidence for DAT- mediated reverse transport of dopamine (Leviel, 2001). First, as discussed, amphetamine’s mechanism of action relies on its ability to promote non-exocytotic release of dopamine by reversing the activity of DAT (Sulzer et al., 1995; Leviel, 2001). This reveals that as a transporter, DAT is capable of moving dopamine in the opposite direction, at least in the presence of amphetamine. Second, when a human DAT coding variant, Ala559Val, was introduced in cells, it was shown to exhibit spontaneous DAT-mediated outward efflux of dopamine (Mazei-Robison et al., 2008). Furthermore, knock-in mice generated from this DAT variant displayed higher extracellular dopamine levels, consistent with DAT-mediated leakage of dopamine (Mergy et al., 2014). These results suggest that structural modification of DAT can give rise to a transporter that constitutively releases dopamine into the extracellular space. Third, using the giant dopamine neuron of the pond snail Planorbis corneus, it was illustrated that injection of dopamine within the cytosol leads to neurotransmitter efflux (Sulzer et al., 1995). Specifically, this efflux was blocked by the DAT inhibitor, , indicating that the release of dopamine was DAT-dependent. Hence, in this system, increased cytosolic dopamine concentrations were sufficient to induce DAT-mediated dopamine release. Collectively, these findings lend support to the hypothesis that DAT-tg/VMAT2-kd mice may endogenously leak dopamine via DAT due to buildup of cytosolic levels of the neurotransmitter. Reversal of DAT would lead to increased extracellular dopamine levels that can explain novelty-induced hyperlocomotion of DAT-tg/VMAT2-kd mice as well as their enhanced performance on fine motor and cognitive tests compared to other genotypes. Furthermore, enhanced response of DAT-tg/VMAT2-kd mice to amphetamine may reflect their basal sensitivity towards DAT reversal. From our current FSCV results, dopamine release is diminished in DAT-tg/VMAT2-kd mice, rather than increased, which would be expected with DAT reversal. However, it is important to note that electrical stimulation produces vesicular, exocytotic dopamine release which was assessed by FSCV, whereas DAT reversal would lead to non-vesicular, non- exocytotic, transporter-mediated dopamine leakage, which could not be measured with this method. While DAT reversal is an intriguing hypothesis to explain the unique behaviors of DAT-tg/VMAT2-kd mice, there are also some caveats to this theory that should be considered. For instance, unlike DAT-tg/VMAT2-kd mice, knock-in DAT Val559 mice, which display

159

DAT-mediated dopamine efflux, do not exhibit overt hyperactivity. Instead, they display a context-dependent darting phenotype (Mergy et al., 2014). Also, DAT Val 559 mice are less sensitive to the effects of amphetamine whereas DAT-tg/VMAT2-kd mice show robust behavioral responses to this drug. Differences between DAT Val 559 and DAT-tg/VMAT2-kd mice are possible because in the former model, dopamine efflux is caused by a rare DAT variant, whereas in the latter, the transporter is functioning normally however an imbalance in dopamine compartmentalization may lead to efflux (Mergy et al., 2014). In summary, other parameters such as extracellular dopamine levels need to be evaluated in DAT-tg/VMAT2-kd mice in order to test the hypothesis of DAT reversal.

Another possible hypothesis to explain the unexpected phenotypes of DAT-tg/VMAT2-kd mice relates to the modulation of dopamine’s signal-to-noise ratio. Signal is defined as the action of dopamine on its intended synaptic receptors while noise is defined as extra-synaptic effects of the neurotransmitter once released. In DAT-tg/VMAT2-kd mice, due to reduced vesicular storage, the quantal release of dopamine is decreased. Conversely, due to enhanced uptake, dopamine is quickly removed from the peri-synaptic space. We initially hypothesized that these genetic manipulations will reduce the absolute quantity of dopamine signaling and cause deleterious downstream effects. However, it is also possible that aside from absolute differences, the relative imbalance of DAT and VMAT2 acts to focus the dopamine signal by releasing a finite amount of neurotransmitter that acts preferentially on synaptic receptors and is efficiently taken up before it diffuses and acts off-target. Hence, DAT-tg/VMAT2-kd mice show improvements in fine motor skill, problem solving and exploration that are not observed in other genotypes. This is another possible explanation for the results obtained however, several other parameters need to be tested to confirm this hypothesis.

4.2.1 Discrepancy between original DAT-tg mice and DAT-tg mice from the DAT-tg/VMAT2-kd colony

The most important challenge encountered during Project 2 was that DAT-tg mice from the DAT VMAT2 breeding colony did not fully replicate some findings obtained from the original DAT- tg mice (Salahpour et al., 2008; Ghisi et al., 2009; Masoud et al., 2015). Some phenotypes were present but attenuated: such as DAT over-expression (1.75-fold vs 3-fold), reduced dopamine

160 tissue content (21% lower vs 33%), enhanced dopamine uptake (39% higher trend vs 60%), reduced electrically-evoked dopamine release (62% lower vs 72%), and increased errors on the challenging beam traversal task (45% increase vs 50%). Other phenotypes were completely absent: such as loss of nigrostriatal dopamine neurons, decreased striatal VMAT2 protein levels and upregulation of striatal D1 and D2 receptors. Indeed, stereological counts of midbrain dopamine neurons were repeated in three independent experiments to ascertain the result that DAT-tg mice no longer demonstrated dopaminergic cell loss. Taken together, these data suggest that DAT-tg mice from the DAT VMAT2 colony consistently demonstrate less dopaminergic toxicity than the original DAT-tg mice. Since these animals no longer show dopaminergic cell loss, many of its downstream and compensatory changes are also not replicated, such as decreased VMAT2 protein and upregulated dopamine receptors. However, even in the absence of dopaminergic neurodegeneration, these DAT-tg mice still display reduced striatal dopamine tissue content, reduced striatal dopamine release and increased fine motor errors, providing evidence of modest dopaminergic damage at the terminals that may not have affected the cell bodies. It is possible that the level of DAT over-expression and functional dopamine uptake in these mice is not high enough to reach the threshold for cellular toxicity (as indicated by 1.75- fold more DAT protein in these mice instead of 3-fold more DAT protein in the original DAT-tg mice). Given that these DAT-tg mice are from the DAT VMAT2 line and previous DAT-tg mice belonged to a separate colony, there can be many possible reasons behind this discrepancy:

1. Genetic background: Original DAT-tg mice were purely on a C57BL/6 background. DAT-tg mice were then crossed with VMAT2-kd mice that were on a mixed C57BL/6 and 129SV background. Through successive generations of breeding, DAT-tg/VMAT2- kd mice were backcrossed to C57BL/6. However, it is possible that by interbreeding from different genetic backgrounds, some genetic modifiers may have been introduced that dampened the phenotypes of these mice. An example of this is illustrated by results from our own laboratory, where we found differences in cocaine response depending on whether C57BL/6 mice were obtained from Jackson or Charles River. This highlights the potent effect of specific genetic backgrounds on dopaminergic phenotypes. 2. Nutrition: Since DAT-tg/VMAT2-kd mice are fragile and die prematurely, we supplement these mice (and their cage littermates) with peanut butter and safflower seeds from birth. The original DAT-tg mice did not receive any dietary supplementation. Given

161

that there is a bidirectional relationship between dopamine signaling and feeding behavior/reward, it is possible that nutritional differences may contribute towards phenotypic outcomes. 3. Breeding: Original DAT-tg mice were generated by breeding DAT-nTg (WT) mice with DAT-Tg mice which produced 50% DAT-nTg and 50% DAT-Tg mice. In the DAT VMAT2 colony, DAT-tg mice are generated by breeding +/- VMAT2-kd/ DAT-tg mice with +/- VMAT2-kd/DAT-nTg mice which produces 6 possible genotypes. Hence DAT- tg mice from the DAT VMAT2 colony are conceived and reared by parents that are heterozygotes for VMAT2, a gene essential for monoamine storage. In comparison, parents of the original DAT-tg mice were normal for the VMAT2 locus. Difference in the breeding pairs may impact prenatal and postnatal development of the progeny, accounting for differences in their phenotypes.

4.2.2 Hypothesis revisited

Keeping in mind that DAT-tg mice have lost their neurodegenerative phenotype, it is not surprising that DAT-tg/VMAT2-kd mice also did not show evidence of prominent dopaminergic toxicity such as loss of dopamine cells or poor motor ability. Our initial hypothesis hinged on the finding that DAT-tg mice demonstrate dopaminergic cell loss (Masoud et al., 2015) and since that effect was lost, our hypothesis regarding Parkinsonian effects in DAT-tg/VMAT2-kd mice could no longer be supported. Due to dampened effects of DAT over-expression in these DAT-tg mice, many features of DAT-tg/VMAT2-kd mice closely mimicked VMAT2-kd animals, instead. Nonetheless, we discovered several unique and robust phenotypes in DAT-tg/VMAT2- kd mice (such as reduced survival, basal hyperactivity, supersensitivity to amphetamine) that provide novel information regarding dopamine function. In fact, in models that show neurodegeneration, most of the changes observed are symptoms or compensations of the cell loss. By studying animals on a non-neurodegenerative background, we can objectively evaluate the changes in dopamine function as a result of DAT and VMAT2 expression, without the complication of altered dopamine neuron numbers across genotypes.

In summary, the goal for Project 2 was to generate mice that simultaneously over-express DAT and under-express VMAT2. Numerous experiments were conducted to characterize the

162 dopamine system and its behavioral output in these mice. Our initial hypothesis regarding enhanced dopaminergic toxicity in DAT-tg/VMAT2-kd mice was not supported by the results mainly because DAT-tg mice no longer demonstrated dopaminergic cell loss. Loss of previously-observed phenotypes in DAT-tg mice posed a major obstacle in interpreting the results from this project, however, we executed our plan of study and uncovered interesting phenotypes in DAT-tg/VMAT2-kd mice that shed light on the complex functioning of the dopamine system.

4.3 Conclusion

The overall aim of this thesis was to investigate the effects of altered dopamine compartmentalization on the function of the dopamine system and related behaviors in genetically modified mice. In particular, we aspired to amplify cytosolic compartmentalization of dopamine by enhancing dopamine uptake through DAT over-expression and reducing dopamine vesicular storage through VMAT2 knockdown.

Project 1 focused on DAT over-expressing transgenic mice that displayed detrimental outcomes including loss of midbrain dopamine neurons, oxidative stress, L-DOPA reversible motor deficits and enhanced vulnerability to MPTP-induced toxicity. These results clearly demonstrate that increasing dopamine uptake and probable accumulation of cytosolic dopamine is harmful for dopaminergic cells. Although these neurons are equipped to store, metabolize and release dopamine, a modest modification in dopamine transport produces damaging consequences. Hence, Project 1 provided evidence that enhancing cytosolic dopamine is sufficient to cause Parkinson’s disease-like damage in mice. This implies that constantly handling a reactive neurotransmitter like dopamine can render dopaminergic cells inherently vulnerable, which may contribute to their heightened susceptibility to insult in Parkinson’s disease.

Project 2 focused on mice that simultaneously over-express DAT and under-express VMAT2. This project was meant to carry forward the findings of Project 1 by further amplifying cytosolic dopamine levels, however, due to reasons discussed above, DAT-tg/VMAT2-kd mice did not show Parkinsonian features like dopamine cell loss or motor deficits. First, this suggests that there exists a threshold of toxicity that must be breached in order to achieve dopamine cell loss. Original DAT-tg mice showed higher protein expression of DAT and greater dopamine uptake, which translated to loss of dopamine cells (Masoud et al., 2015), whereas DAT-tg mice from the

163

DAT VMAT2 colony displayed comparatively lower DAT over-expression and less pronounced increase in dopamine uptake, which translated to dampened phenotypes and intact nigrostriatal dopamine neurons. This phenomenon regarding the extent of DAT over-expression has previously been observed as well. Original DAT-tg mice were estimated to contain 6 copies of the DAT gene (2 endogenous and 4 from the BAC) based on DAT Southern blot analysis and they demonstrated dopamine cell loss (Salahpour et al., 2008; Masoud et al., 2015). Interestingly, when these DAT-tg animals (6 copies of DAT) were crossed with DAT-KO mice (0 copies of DAT), their progeny lost 1 functional copy of the DAT gene due to incorporation of the DAT- KO allele. Hence, these mice were expected to display 5 copies of the DAT gene (1 endogenous and 4 from the BAC) and remarkably, they no longer displayed dopamine cell loss (Salahpour unpublished). Once again, these results demonstrate a threshold effect where a certain level of DAT over-expression is required to manifest changes in dopamine cell survival. Comparison of Project 1 and 2 also provides similar information, suggesting that loss of dopamine cells only occurs once the degree of cytosolic dopamine accumulation exceeds a threshold beyond which the cell can no longer combat dopamine-induced toxicity. Second, gathering parallel data from DAT-tg mice with or without dopaminergic neurodegeneration, allows us to directly differentiate between phenotypes that are dependent on dopamine cell loss (such as upregulation of dopamine receptors, only observed in original DAT-tg mice) and those that are not (such as presence of fine motor deficits, observed in both original DAT-tg mice and those from the DAT VMAT2 colony).

Moreover, results from Project 2 also offered novel and unexpected insight regarding the role of dopamine compartmentalization in behavioral outcomes. Despite a 95% reduction in striatal dopamine tissue content and 62% decrease in evoked dopamine release, DAT-tg/VMAT2-kd mice were hyperactive and performed better than their counterparts in fine-motor and problem- solving tasks. The mice were also extremely sensitive to amphetamine-induced locomotion, showing abnormal involuntary movements at a dose of 2mg/kg of amphetamine. These findings highlight the immense reserve capacity of the dopamine system in sustaining behavior despite drastic alterations in dopamine homeostasis. In one sense, these findings also offer hope, of being able to modulate and enhance motor activity in a system that was designed to be hypodopaminergic. Evidently, the relationship between dopamine neurotransmission and behavior is far more complex than we had previously anticipated, giving rise to new hypotheses

164 that could potentially explain these results (e.g. DAT reversal). Thus, characteristics of DAT- tg/VMAT2-kd mice reveal untapped potential of the dopamine system to withstand and adapt to modifications in dopamine compartmentalization, while “improving” behavioral outcomes. This not only broadens our understanding of the dopamine system, but also provokes new thoughts about ways to target this system therapeutically.

Collectively, the results of this thesis have confirmed our predictions regarding cytosolic dopamine-induced toxicity (Project 1) and also enriched our existing knowledge of the dopamine system by revealing novel behavioral findings in mice with altered dopamine compartmentalization (Project 2).

4.4 Technical Challenges

During the course of Projects 1 and 2, a few technical challenges were faced that complicated our interpretation of the results obtained. The first obstacle we encountered was regarding BAC transgenic over-expression of the DAT gene in mice. We discovered that there was a spontaneous loss of genomic DAT copy numbers in transgenic mice (see Appendix 2). This problem was first found in the DAT-tg colony (Project 1) and then independently discovered in the DAT-tg/VMAT2-kd colony (Project 2). During successive rounds of breeding, extra copies of the DAT transgene can be lost producing “low” copy DAT-tg mice. Typically, DAT-tg mice possess 6-8 copies of the DAT gene as determined by genomic quantitative PCR of the DAT locus. However, “low” copy DAT-tg and DAT-tg/VMAT2-kd mice only possess one extra copy of the DAT gene (3 copies total). This dampens the effects of DAT over-expression and produces phenotypic variability. To control this, we routinely performed genomic qPCR to check for DAT copy number. All results shown in this thesis are from “high” copy DAT-tg mice (except in Appendix 2 which shows data from “low” copy DAT-tg mice).

Secondly, in Project 2, the technique of stereology was found to be unreliable. We conducted stereological counts of midbrain dopamine neurons in collaboration with other laboratories that routinely perform these experiments. We repeated stereology in 3 independent experiments to ascertain the results. In one case, we sent the same slides to be counted by the same researcher (blind to the genotypes) twice - and opposing results were obtained. Thus, even when controlling for the user and tissue staining, this technique showed high variability. Since stereology was a

165 central part of this project, challenges with this technique, compounded with altered phenotypes in DAT-tg mice, created a major hurdle in Project 2.

4.5 Future Directions

The results from this thesis, in particular Project 2, have raised several interesting questions that warrant further exploration. With regards to Project 1, DAT-tg mice displayed dopaminergic damage that is postulated to be caused by accumulation of cytosolic dopamine. However, since cytosolic dopamine cannot be measured in vivo, it would be interesting to generate primary neuronal cultures of dopaminergic neurons from DAT-tg mice and use intracellular patch electrochemistry to measure dopamine levels in the cytosolic compartment (Mosharov et al., 2003, 2009). This experiment would answer the question of how much cytosolic dopamine accumulation occurs in DAT-tg mice. In addition, DAT-tg mice display enhanced sensitivity to MPTP-induced toxicity. MPP+ is a substrate for DAT, therefore heightened toxicity in DAT-tg mice may simply reflect increased access of the toxicant into dopaminergic neurons. To further explore the interaction between genetic and environmental insults in disease pathogenesis, it is important to test the effect of a toxicant in DAT-tg mice that is independent of DAT in its activity. We attempted these experiments with the pesticide, rotenone, however these studies were not successful due to 1) technical complications in administering a highly lipophilic drug and 2) relative insensitivity of mice to rotenone-induced damage. Nonetheless, it would be useful to investigate the response of DAT-tg mice to other environmental toxicants that have been implicated in Parkinson’s disease, such as maneb. This would answer the question whether mishandling dopamine can predispose DAT-tg mice to environmental insults and shed light on the multiple hit hypothesis of Parkinson’s disease (Sulzer, 2007).

With regards to Project 2, there are several avenues that can be pursued to further elaborate on the current results. First, function of the dopamine system can be further explored in DAT- tg/VMAT2-kd mice. Although dopamine release, uptake, tissue content, cell number and receptor levels have been evaluated in DAT-tg/VMAT2-kd mice, extracellular dopamine levels and firing pattern of dopamine neurons should also be evaluated. These parameters are instrumental in explaining the unique behavioral phenotypes of DAT-tg/VMAT2-kd mice, such as motor hyperactivity, improved fine motor skill and problem solving ability compared to other genotypes. Furthermore, if the theory of DAT reversal in DAT-tg/VMAT2-kd mice is true, then

166 we would expect extracellular dopamine levels to be increased as a result of DAT-mediated dopamine efflux. Midbrain dopamine neurons fire spontaneously in pacemaking and/or burst modes (Grace and Bunney, 1984). The endogenous firing pattern is important as it leads to dopamine release and contributes to dopamine-related behaviors, such as reward-based learning (Schultz et al., 1997; Reynolds et al., 2001). In fact, dopamine itself can regulate the firing activity of dopamine neurons via feedback mechanisms (Paladini et al., 2003). Since the dynamics of evoked dopamine release and uptake are drastically altered in DAT-tg/VMAT2-kd mice, this could affect the firing pattern of dopamine cells. Hence, it would be informative to investigate dopamine neuronal firing activity and extracellular dopamine levels in DAT- tg/VMAT2-kd mice to explain some of their behavioral phenotypes.

Second, a robust finding of DAT-tg/VMAT2-kd that warrants examination is their poor survival. Adult (12-week old) DAT-tg/VMAT2-kd mice demonstrate a 40-50% reduction in survival when compared to age-matched WT, DAT-tg and VMAT2-kd mice. In addition, DAT- tg/VMAT2-kd mice are also significantly smaller in size than WT animals. The dramatic impairment in survival selectively affects animals that concurrently over-express DAT and under-express VMAT2, indicating a synergistic negative effect of these two manipulations on fitness. Although survival and body weight are complex traits that are influenced by several factors including feeding, nutrition and development, it is evident that altering the balance of dopamine compartmentalization produces detrimental outcomes in DAT-tg/VMAT2-kd mice. Thus, further work unraveling the impact of both neuronal and peripheral dopamine signaling on overall fitness is warranted.

Third, in comparison to female mice, survival and body weight measures seem to be more drastically diminished in male DAT-tg/VMAT2-kd mice, indicating a significant role of sex in determining response. In this thesis, neurochemical and behavioral data could not be stratified by sex due to relatively smaller sample sizes and technical constraints including: 1) reduced survival of male mice, 2) low probability (1/16) of obtaining an animal of a particular genotype and sex and 3) small litter sizes (presumably because of dopamine and prolactin dysregulation). However, it would be useful to investigate if other dopaminergic measures also show a sex bias using larger sample sizes that are powered to detect those differences. Moreover, uncovering the mechanisms that underlie the vulnerability of male DAT-tg/VMAT2-kd mice (and protection of females), is particularly relevant because in humans, the likelihood of developing Parkinson’s

167 disease is also higher in men, suggesting a specific vulnerability of males to dopaminergic malfunction (Van Den Eeden et al., 2003; Wooten et al., 2004).

Fourth, premature death of DAT-tg/VMAT2-kd mice adds a layer of complication in studying these animals, since most experiments are conducted at or after 12 weeks of age when 46% of DAT-tg/VMAT2-kd mice have already died. As a result, characterization of these mice inadvertently occurs on the remaining animals that have survived to adulthood. It is conceivable that those animals that naturally died beforehand may have expressed more extreme changes that led to their demise in comparison to the mice that survived. This could also explain why the phenotypes observed in DAT-tg/VMAT2-kd mice were less severe than hypothesized. We anticipated drastic modifications in dopamine cell number, tissue content and receptor binding beyond the levels observed in DAT-tg and VMAT2-kd mice alone. Perhaps these synergistic changes were not observed because our studies were focused on DAT-tg/VMAT2-kd mice that survived into adulthood and may have successfully compensated for their genetic shortcomings. Thus, it would be useful to examine animals at earlier time points using non-invasive techniques (e.g. behavior, imaging) and track their survival to determine whether those animals that die by 12 weeks of age, indicate signs of toxicity during early development. Also, closely tracking mouse survival may allow for collection of viable tissue as soon as an animal is found dead. This way, dopaminergic markers can be assessed in specific brain regions and compared to age- matched mice.

Lastly, our experiments in Project 2 have focused primarily on the dopamine system. However, reduction of VMAT2 levels in VMAT2-kd and DAT-tg/VMAT2-kd mice should affect all monoaminergic systems where the protein is normally expressed. Furthermore, in addition to dopaminergic neurodegeneration, VMAT2-kd mice are also reported to display 1) loss of noradrenergic cells in the LC and 2) disruption of serotonin signaling (Taylor et al., 2014; Alter et al., 2016). Some of the behavioral tests conducted on DAT VMAT2 mice focused on attention, anxiety and cognitive phenotypes that are likely to be influenced by changes in noradrenergic and serotonergic signaling. Therefore, it would be interesting to gauge the function of these other monoamine systems in DAT-tg/VMAT2-kd mice. Although DAT over-expression is confined to dopaminergic neurons, one can imagine that crippling the dopamine system of DAT-tg/VMAT2-kd mice may unveil compensatory changes in noradrenergic or serotonergic systems of these mice.

168

References

Abbott, R.D., Petrovitch, H., White, L.R., Masaki, K.H., Tanner, C.M., Curb, J.D., Grandinetti, A., Blanchette, P.L., Popper, J.S., and Ross, G.W. (2001). Frequency of bowel movements and the future risk of Parkinson’s disease. Neurology 57, 456–462.

Ben Abdallah, N.M.B., Fuss, J., Trusel, M., Galsworthy, M.J., Bobsin, K., Colacicco, G., Deacon, R.M.J., Riva, M.A., Kellendonk, C., Sprengel, R., Lipp, H.P., and Gass, P. (2011). The puzzle box as a simple and efficient behavioral test for exploring impairments of general cognition and executive functions in mouse models of schizophrenia. Exp. Neurol. 227, 42– 52.

Abdulwahid Arif, I., and Ahmad Khan, H. (2010). Environmental toxins and Parkinson’s disease: putative roles of impaired electron transport chain and oxidative stress. Toxicol. Ind. Health 26, 121–128.

Adinoff, B. (2004). Neurobiologic processes in drug reward and addiction. Harv. Rev. Psychiatry 12, 305–320.

Adriani, W., Boyer, F., Gioiosa, L., Macrì, S., Dreyer, J.-L., and Laviola, G. (2009). Increased impulsive behavior and risk proneness following lentivirus-mediated dopamine transporter over-expression in rats’ nucleus accumbens. Neuroscience 159, 47–58.

Alagarsamy, S., Phillips, M., Pappas, T., and Johnson, K.M. (1997). Dopamine neurotoxicity in cortical neurons. Drug Alcohol Depend. 48, 105–111.

Alam, M., and Schmidt, W.J. (2002). Rotenone destroys dopaminergic neurons and induces parkinsonian symptoms in rats. Behav. Brain Res. 136, 317–324.

Alam, Z.I., Daniel, S.E., Lees, A.J., Marsden, D.C., Jenner, P., and Halliwell, B. (1997a). A generalised increase in protein carbonyls in the brain in Parkinson’s but not incidental Lewy body disease. J. Neurochem. 69, 1326–1329.

Alam, Z.I., Jenner, A., Daniel, S.E., Lees, A.J., Cairns, N., Marsden, C.D., Jenner, P., and Halliwell, B. (1997b). Oxidative DNA damage in the parkinsonian brain: an apparent selective increase in 8-hydroxyguanine levels in substantia nigra. J. Neurochem. 69, 1196– 1203.

Alter, S.P., Stout, K.A., Lohr, K.M., Taylor, T.N., Shepherd, K.R., Wang, M., Guillot, T.S., and Miller, G.W. (2016). Reduced vesicular monoamine transport disrupts serotonin signaling but does not cause serotonergic degeneration. Exp. Neurol. 275, 17–24.

Anderson, D.G., Mariappan, S.V.S., Buettner, G.R., and Doorn, J.A. (2011). Oxidation of 3,4- dihydroxyphenylacetaldehyde, a toxic dopaminergic metabolite, to a semiquinone radical and an ortho-quinone. J. Biol. Chem. 286, 26978–26986.

Angers, S., Salahpour, A., and Bouvier, M. (2002). Dimerization: an emerging concept for G

169

protein-coupled receptor ontogeny and function. Annu. Rev. Pharmacol. Toxicol. 42, 409– 435.

Aragona, B.J., Cleaveland, N.A., Stuber, G.D., Day, J.J., Carelli, R.M., and Wightman, R.M. (2008). Preferential enhancement of dopamine transmission within the nucleus accumbens shell by cocaine is attributable to a direct increase in phasic dopamine release events. J. Neurosci. 28, 8821–8831.

Arbuthnott, G.W., and Wickens, J. (2007). Space, time and dopamine. Trends Neurosci. 30, 62– 69.

Asanuma, M., Miyazaki, I., and Ogawa, N. (2003). Dopamine- or L-DOPA-induced neurotoxicity: The role of dopamine quinone formation and tyrosinase in a model of Parkinson’s disease. Neurotox. Res. 5, 165–176.

Barr, A.M., Lehmann-Masten, V., Paulus, M., Gainetdinov, R.R., Caron, M.G., and Geyer, M.A. (2003). The Selective Serotonin-2A Receptor Antagonist M100907 Reverses Behavioral Deficits in Dopamine Transporter Knockout Mice. Neuropsychopharmacology 29, 221–228.

Bartels, T., Choi, J.G., and Selkoe, D.J. (2011). α-Synuclein occurs physiologically as a helically folded tetramer that resists aggregation. Nature 477, 107–110.

Barzilai, A., Melamed, E., and Shirvan, A. (2001). Is there a rationale for neuroprotection against dopamine toxicity in Parkinson’s disease? Cell. Mol. Neurobiol. 21, 215–235.

Beal, M.F. (2010). Parkinson’s disease: a model dilemma. Nature 466, S8–S10.

Beaulieu, J.-M., and Gainetdinov, R.R. (2011). The Physiology, Signaling, and Pharmacology of Dopamine Receptors. Pharmacol. Rev. 63, 182–217.

Beaulieu, J.-M., Sotnikova, T.D., Marion, S., Lefkowitz, R.J., Gainetdinov, R.R., and Caron, M.G. (2005). An Akt/β-Arrestin 2/PP2A Signaling Complex Mediates Dopaminergic Neurotransmission and Behavior. Cell 122, 261–273.

Bemis, J.C., and Seegal, R.F. (2004). PCB-induced inhibition of the vesicular monoamine transporter predicts reductions in synaptosomal dopamine content. Toxicol. Sci. 80, 288–295.

Ben-Shachar, D., Zuk, R., and Glinka, Y. (1995). Dopamine Neurotoxicity: Inhibition of Mitochondrial Respiration. J. Neurochem. 64, 718–723.

Bernheimer, H., Birkmayer, W., Hornykiewicz, O., Jellinger, K., and Seitelberger, F. (1973). Brain dopamine and the syndromes of Parkinson and Huntington Clinical, morphological and neurochemical correlations. J. Neurol. Sci. 20, 415–455.

Beuming, T., Kniazeff, J., Bergmann, M.L., Shi, L., Gracia, L., Raniszewska, K., Newman, A.H., Javitch, J.A., Weinstein, H., Gether, U., and Loland, C.J. (2008). The binding sites for cocaine and dopamine in the dopamine transporter overlap. Nat. Neurosci. 11, 780–789.

Bezard, E., Gross, C.E., Fournier, M.C., Dovero, S., Bloch, B., and Jaber, M. (1999). Absence of

170

MPTP-induced neuronal death in mice lacking the dopamine transporter. Exp Neurol 155, 268–273.

Birkmayer, W., and Hornykiewicz, O. (1961). [The L-3,4-dioxyphenylalanine (DOPA)-effect in Parkinson-akinesia]. Wien. Klin. Wochenschr. 73, 787–788.

Birkmayer, W., and Hornykiewicz, O. (1962). [The L-dihydroxyphenylalanine (L-DOPA) effect in Parkinson’s syndrome in man: On the pathogenesis and treatment of Parkinson akinesis]. Arch Psychiatr Nervenkr Z Gesamte Neurol Psychiatr 203, 560–574.

Björklund, A., and Dunnett, S.B. (2007). Dopamine neuron systems in the brain: an update. Trends Neurosci. 30, 194–202.

Blanchard, V., Raisman-Vozari, R., Vyas, S., Michel, P.P., Javoy-Agid, F., Uhl, G., and Agid, Y. (1994). Differential expression of tyrosine hydroxylase and membrane dopamine transporter genes in subpopulations of dopaminergic neurons of the rat mesencephalon. Mol. Brain Res. 22, 29–38.

Bloemen, O., de Koning, M., Boot, E., Booij, J., and van Amelsvoort, T.A. (2008). Challenge and Therapeutic Studies Using Alpha-Methyl-para-Tyrosine (AMPT) in Neuropsychiatric Disorders: A Review. Cent. Nerv. Syst. Agents Med. Chem. 8, 249–256.

Bogerts, B., Häntsch, J., and Herzer, M. (1983). A morphometric study of the dopamine- containing cell groups in the mesencephalon of normals, Parkinson patients, and schizophrenics. Biol. Psychiatry 18, 951–969.

Bohnen, N.I., and Albin, R.L. (2011). The cholinergic system and Parkinson disease. Behav. Brain Res. 221, 564–573.

Bolam, J.P., and Pissadaki, E.K. (2012). Living on the edge with too many mouths to feed: why dopamine neurons die. Mov. Disord. 27, 1478–1483.

Bourne, J.A. (2006). SCH 23390: The First Selective Dopamine D1-Like Receptor Antagonist. CNS Drug Rev. 7, 399–414.

Braak, H., Braak, E., Yilmazer, D., de Vos, R.A.I., Jansen, E.N.H., and Bohl, J. (1996). Pattern of brain destruction in Parkinson’s and Alzheimer’s diseases. J. Neural Transm. 103, 455– 490.

Bradner, J.M., Suragh, T.A., Wilson, W.W., Lazo, C.R., Stout, K.A., Kim, H.M., Wang, M.Z., Walker, D.I., Pennell, K.D., Richardson, J.R., Miller, G.W., and Caudle, W.M. (2013). Exposure to the polybrominated diphenyl ether mixture DE-71 damages the nigrostriatal dopamine system: role of dopamine handling in neurotoxicity. Exp. Neurol. 241, 138–147.

Brighina, L., Riva, C., Bertola, F., Saracchi, E., Fermi, S., Goldwurm, S., and Ferrarese, C. (2013). Analysis of vesicular monoamine transporter 2 polymorphisms in Parkinson’s disease. Neurobiol Aging 34, 1712.e9-13.

Brogden, R.N., Heel, R.C., Speight, T.M., and Avery, G.S. (1981). alpha-Methyl-p-tyrosine: a

171

review of its pharmacology and clinical use. Drugs 21, 81–89.

Bunemann, M., Frank, M., and Lohse, M.J. (2003). Gi protein activation in intact cells involves subunit rearrangement rather than dissociation. Proc. Natl. Acad. Sci. 100, 16077–16082.

Burke, W.J., Li, S.W., Chung, H.D., Ruggiero, D.A., Kristal, B.S., Johnson, E.M., Lampe, P., Kumar, V.B., Franko, M., Williams, E.A., and Zahm, D.S. (2004). Neurotoxicity of MAO Metabolites of Catecholamine Neurotransmitters: Role in Neurodegenerative Diseases. Neurotoxicology 25, 101–115.

Burke, W.J., Li, S.W., Williams, E.A., Nonneman, R., and Zahm, D.S. (2003). 3,4- Dihydroxyphenylacetaldehyde is the toxic dopamine metabolite in vivo: implications for Parkinson’s disease pathogenesis. Brain Res. 989, 205–213.

Burré, J., Sharma, M., Tsetsenis, T., Buchman, V., Etherton, M.R., and Südhof, T.C. (2010). Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science 329, 1663–1667.

Buttarelli, F.R., Fanciulli, A., Pellicano, C., and Pontieri, F.E. (2011). The dopaminergic system in peripheral blood lymphocytes: from physiology to pharmacology and potential applications to neuropsychiatric disorders. Curr. Neuropharmacol. 9, 278–288.

Cabin, D.E., Shimazu, K., Murphy, D., Cole, N.B., Gottschalk, W., McIlwain, K.L., Orrison, B., Chen, A., Ellis, C.E., Paylor, R., Lu, B., and Nussbaum, R.L. (2002). Synaptic vesicle depletion correlates with attenuated synaptic responses to prolonged repetitive stimulation in mice lacking alpha-synuclein. J. Neurosci. 22, 8797–8807.

Calabresi, P., Picconi, B., Tozzi, A., Ghiglieri, V., and Di Filippo, M. (2014). Direct and indirect pathways of basal ganglia: a critical reappraisal. Nat. Neurosci. 17, 1022–1030.

Cannon, J.R., and Greenamyre, J.T. (2011). The role of environmental exposures in neurodegeneration and neurodegenerative diseases. Toxicol. Sci. 124, 225–250.

Cannon, J.R., and Greenamyre, J.T. (2013). Gene–environment interactions in Parkinson’s disease: Specific evidence in humans and mammalian models. Neurobiol. Dis. 57, 38–46.

Carlson, N.R. (2012). Physiology of behavior 11th edition (Pearson).

Carlsson, A., Lindqvist, M., and Magnusson, T. (1957). 3,4-Dihydroxyphenylalanine and 5- hydroxytryptophan as reserpine antagonists. Nature 180, 1200.

Carlsson, A., Lindqvist, M., Magnusson, T., and Waldeck, B. (1958). On the presence of 3- hydroxytyramine in brain. Science 127, 471.

Caudle, W.M., Richardson, J.R., Wang, M.Z., Taylor, T.N., Guillot, T.S., McCormack, A.L., Colebrooke, R.E., Di Monte, D.A., Emson, P.C., and Miller, G.W. (2007). Reduced Vesicular Storage of Dopamine Causes Progressive Nigrostriatal Neurodegeneration. J. Neurosci. 27, 8138–8148.

172

Chan, C.S., Guzman, J.N., Ilijic, E., Mercer, J.N., Rick, C., Tkatch, T., Meredith, G.E., and Surmeier, D.J. (2007). “Rejuvenation” protects neurons in mouse models of Parkinson’s disease. Nature 447, 1081–1086.

Chandler, K.J., Chandler, R.L., Broeckelmann, E.M., Hou, Y., Southard-Smith, E.M., and Mortlock, D.P. (2007). Relevance of BAC transgene copy number in mice: transgene copy number variation across multiple transgenic lines and correlations with transgene integrity and expression. Mamm. Genome 18, 693–708.

Chang, H.-Y., Grygoruk, A., Brooks, E.S., Ackerson, L.C., Maidment, N.T., Bainton, R.J., and Krantz, D.E. (2006). Overexpression of the Drosophila vesicular monoamine transporter increases motor activity and courtship but decreases the behavioral response to cocaine. Mol. Psychiatry 11, 99–113.

Chaudhuri, K.R., Healy, D.G., and Schapira, A.H. (2006). Non-motor symptoms of Parkinson’s disease: diagnosis and management. Lancet Neurol. 5, 235–245.

Chen, B.T., and Rice, M.E. (2001). Novel Ca2+ dependence and time course of somatodendritic dopamine release: substantia nigra versus striatum. J. Neurosci. 21, 7841–7847.

Chen, L., Ding, Y., Cagniard, B., Van Laar, A.D., Mortimer, A., Chi, W., Hastings, T.G., Kang, U.J., and Zhuang, X. (2008). Unregulated Cytosolic Dopamine Causes Neurodegeneration Associated with Oxidative Stress in Mice. J. Neurosci. 28, 425–433.

Cheng, H.-C., Ulane, C.M., and Burke, R.E. (2010). Clinical progression in Parkinson’s disease and the neurobiology of axons. Ann. Neurol. 67, 715–725.

Chiba, K., Peterson, L.A., Castagnoli, K.P., Trevor, A.J., and Castagnoli, N. (1985). Studies on the molecular mechanism of bioactivation of the selective nigrostriatal toxin 1-methyl-4- phenyl-1,2,3,6-tetrahydropyridine. Drug Metab. Dispos. 13, 342–347.

Chinta, S.J., and Andersen, J.K. (2005). Dopaminergic neurons. Int. J. Biochem. Cell Biol. 37, 942–946.

Cicchetti, F., Drouin-Ouellet, J., and Gross, R.E. (2009). Environmental toxins and Parkinson’s disease: what have we learned from pesticide-induced animal models? Trends Pharmacol. Sci. 30, 475–483.

Clark, I.E., Dodson, M.W., Jiang, C., Cao, J.H., Huh, J.R., Seol, J.H., Yoo, S.J., Hay, B.A., and Guo, M. (2006). Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin. Nature 441, 1162–1166.

Colpaert, F. (1987). Pharmacological characteristics of tremor, rigidity and hypokinesia induced by reserpine in rat. Neuropharmacology 26, 1431–1440.

Conway, K.A., Rochet, J.C., Bieganski, R.M., and Lansbury, P.T. (2001). Kinetic stabilization of the alpha-synuclein protofibril by a dopamine-alpha-synuclein adduct. Science 294, 1346– 1349.

173

Costello, S., Cockburn, M., Bronstein, J., Zhang, X., and Ritz, B. (2009). Parkinson’s Disease and Residential Exposure to Maneb and Paraquat From Agricultural Applications in the Central Valley of California. Am. J. Epidemiol. 169, 919–926.

Cragg, S.J., and Rice, M.E. (2004). DAncing past the DAT at a DA synapse. Trends Neurosci. 27, 270–277.

Creese, I., Burt, D., and Snyder, S. (1976). Dopamine receptor binding predicts clinical and pharmacological potencies of antischizophrenic drugs. Science 192, 481–483.

Daberkow, D.P., Brown, H.D., Bunner, K.D., Kraniotis, S.A., Doellman, M.A., Ragozzino, M.E., Garris, P.A., and Roitman, M.F. (2013). Amphetamine paradoxically augments exocytotic dopamine release and phasic dopamine signals. J. Neurosci. 33, 452–463.

Dahlstroem, A., and Fuxe, K. (1964). Evidence for the existence of monoamine-containing neurons in the central nervous system. I. Demonstration of monoamine in the cell bodies of brain stem neurons. Acta Physiol. Scand. Suppl. 232, 1–55.

Dalle-Donne, I., Rossi, R., Giustarini, D., Milzani, A., and Colombo, R. (2003). Protein carbonyl groups as biomarkers of oxidative stress. Clin. Chim. Acta 329, 23–38.

Daubner, S.C., Le, T., and Wang, S. (2011). Tyrosine hydroxylase and regulation of dopamine synthesis. Arch. Biochem. Biophys. 508, 1–12.

Dauer, W., Kholodilov, N., Vila, M., Trillat, A.-C., Goodchild, R., Larsen, K.E., Staal, R., Tieu, K., Schmitz, Y., Yuan, C.A., Rocha, M., Jackson-Lewis, V., Hersch, S., Sulzer, D., Przedborski, S., Burke, R., and Hen, R. (2002). Resistance of alpha -synuclein null mice to the parkinsonian neurotoxin MPTP. Proc. Natl. Acad. Sci. U. S. A. 99, 14524–14529.

Dauer, W., and Przedborski, S. (2003). Parkinson’s disease: mechanisms and models. Neuron 39, 889–909.

Delaville, C., Deurwaerdère, P. De, and Benazzouz, A. (2011). Noradrenaline and Parkinson’s disease. Front. Syst. Neurosci. 5, 31.

Deleu, D., Hanssens, Y., and Northway, M.G. (2004). Subcutaneous apomorphine : an evidence- based review of its use in Parkinson’s disease. Drugs Aging 21, 687–709.

Delfs, J.M., and Kelley, A.E. (1990). The role of D1 and D2 dopamine receptors in oral stereotypy induced by dopaminergic stimulation of the ventrolateral striatum. Neuroscience 39, 59–67.

Deng, H., Dodson, M.W., Huang, H., and Guo, M. (2008). The Parkinson’s disease genes pink1 and parkin promote mitochondrial fission and/or inhibit fusion in Drosophila. Proc. Natl. Acad. Sci. 105, 14503–14508.

Deniau, J.M., Mailly, P., Maurice, N., and Charpier, S. (2007). The pars reticulata of the substantia nigra: a window to basal ganglia output. Prog. Brain Res. 160, 151–172.

174

Dexter, D.T., Carter, C.J., Wells, F.R., Javoy-Agid, F., Agid, Y., Lees, A., Jenner, P., and Marsden, C.D. (1989a). Basal lipid peroxidation in substantia nigra is increased in Parkinson’s disease. J. Neurochem. 52, 381–389.

Dexter, D.T., Wells, F.R., Lees, A.J., Agid, F., Agid, Y., Jenner, P., and Marsden, C.D. (1989b). Increased nigral iron content and alterations in other metal ions occurring in brain in Parkinson’s disease. J. Neurochem. 52, 1830–1836.

Dias, V., Junn, E., and Mouradian, M.M. (2013). The role of oxidative stress in Parkinson’s disease. J. Parkinsons. Dis. 3, 461–491.

Digby, G.J., Lober, R.M., Sethi, P.R., and Lambert, N.A. (2006). Some G protein heterotrimers physically dissociate in living cells. Proc. Natl. Acad. Sci. U. S. A. 103, 17789–17794.

Donovan, D.M., Miner, L.L., Perry, M.P., Revay, R.S., Sharpe, L.G., Przedborski, S., Kostic, V., Philpot, R.M., Kirstein, C.L., Rothman, R.B., Schindler, C.W., and Uhl, G.R. (1999). Cocaine reward and MPTP toxicity: Alteration by regional variant dopamine transporter overexpression. Mol. Brain Res. 73, 37–49.

Drucker-Colín, R., and García-Hernández, F. (1991). A new motor test sensitive to aging and dopaminergic function. J. Neurosci. Methods 39, 153–161.

Van Den Eeden, S.K., Tanner, C.M., Bernstein, A.L., Fross, R.D., Leimpeter, A., Bloch, D.A., and Nelson, L.M. (2003). Incidence of Parkinson’s Disease: Variation by Age, Gender, and Race/Ethnicity. Am. J. Epidemiol. 157, 1015–1022.

Ehringer, H., and Hornykievicz, O. (1960). Distribution of noradrenaline and dopamine (3- hydroxytyramine) in the human brain and their behavior in diseases of the extrapyramidal system. Klin. Wochenschr. 38, 1236–1239.

Eisenhofer, G., Kopin, I.J., and Goldstein, D.S. (2004a). Catecholamine Metabolism: A Contemporary View with Implications for Physiology and Medicine. Pharmacol. Rev. 56, 331–349.

Eisenhofer, G., Kopin, I.J., and Goldstein, D.S. (2004b). Leaky catecholamine stores: undue waste or a stress response coping mechanism? Ann NY Acad Sci 1018, 224–230.

Elsworth, J.D., and Roth, R.H. (1997). Dopamine Synthesis, Uptake, Metabolism, and Receptors: Relevance to Gene Therapy of Parkinson’s Disease. Exp. Neurol. 144, 4–9.

Emre, M. (2003). Dementia associated with Parkinson’s disease. Lancet. Neurol. 2, 229–237.

Erickson, J.D., and Eiden, L.E. (1993). Functional identification and molecular cloning of a human brain vesicle monoamine transporter. J Neurochem 61, 2314–2317.

Fahn, S., and Cohen, G. (1992). The oxidant stress hypothesis in Parkinson’s disease: evidence supporting it. Ann. Neurol. 32, 804–812.

Fauvet, B., Mbefo, M.K., Fares, M.-B., Desobry, C., Michael, S., Ardah, M.T., Tsika, E., Coune,

175

P., Prudent, M., Lion, N., Eliezer, D., Moore, D.J., Schneider, B., Aebischer, P., El-Agnaf, O.M., Masliah, E., and Lashuel, H.A. (2012). α-Synuclein in central nervous system and from erythrocytes, mammalian cells, and Escherichia coli exists predominantly as disordered monomer. J. Biol. Chem. 287, 15345–15364.

Filloux, F., and Townsend, J.J. (1993). Pre- and postsynaptic neurotoxic effects of dopamine demonstrated by intrastriatal injection. Exp. Neurol. 119, 79–88.

Fleming, S.M., Salcedo, J., Fernagut, P.-O., Rockenstein, E., Masliah, E., Levine, M.S., and Chesselet, M.-F. (2004). Early and Progressive Sensorimotor Anomalies in Mice Overexpressing Wild-Type Human -Synuclein. J. Neurosci. 24, 9434–9440.

Fon, E.A., Pothos, E.N., Sun, B.C., Killeen, N., Sulzer, D., and Edwards, R.H. (1997). Vesicular transport regulates monoamine storage and release but is not essential for amphetamine action. Neuron 19, 1271–1283.

Foppoli, C., Coccia, R., Cini, C., and Rosei, M.A. (1997). Catecholamines oxidation by xanthine oxidase. Biochim. Biophys. Acta 1334, 200–206.

Forno, L.S. (1996). Neuropathology of Parkinson’s disease. J. Neuropathol. Exp. Neurol. 55, 259–272.

Fornstedt, B., and Carlsson, A. (1989). A marked rise in 5-S-cysteinyl-dopamine levels in guinea-pig striatum following reserpine treatment. J. Neural Transm. 76, 155–161.

Fornstedt, B., Rosengren, E., and Carlsson, A. (1986). Occurrence and distribution of 5-S- cysteinyl derivatives of dopamine, dopa and dopac in the brains of eight mammalian species. Neuropharmacology 25, 451–454.

Franklin, K.B.J., and Paxinos, G. (2012). Paxinos and Franklin’s The mouse brain in stereotaxic coordinates 4th edition (Academic Press).

Freis, E.D. (1954). Mental depression in hypertensive patients treated for long periods with large doses of reserpine. N Engl J Med 251, 1006–1008.

Fuxe, K. (1965). Evidence for the existence of monoamine neurons in the central nervous system. IV. Distribution of monoamine nerve terminals in the central nervous system. Acta Physiol. Scand. Suppl. 247, 37+.

Gainetdinov, R.R., and Caron, M.G. (2003). Monoamine transporters: from genes to behavior. Annu. Rev. Pharmacol. Toxicol. 43, 261–284.

Gainetdinov, R.R., Fumagalli, F., Jones, S.R., and Caron, M.G. (1997). Dopamine transporter is required for in vivo MPTP neurotoxicity: evidence from mice lacking the transporter. J. Neurochem. 69, 1322–1325.

Gainetdinov, R.R., Fumagalli, F., Wang, Y.M., Jones, S.R., Levey, a I., Miller, G.W., and Caron, M.G. (1998). Increased MPTP neurotoxicity in vesicular monoamine transporter 2 heterozygote knockout mice. J. Neurochem. 70, 1973–1978.

176

Gainetdinov, R.R., Wetsel, W.C., Jones, S.R., Levin, E.D., Jaber, M., and Caron, M.G. (1999). Role of serotonin in the paradoxical calming effect of psychostimulants on hyperactivity. Science 283, 397–401.

Genetic Science Learning Center (2013). The Reward Pathway Reinforces Behavior. http://learn.genetics.utah.edu/content/addiction/rewardbehavior/

German, D.C., Schlusselberg, D.S., and Woodward, D.J. (1983). Three-dimensional computer reconstruction of midbrain dopaminergic neuronal populations: from mouse to man. J. Neural Transm. 57, 243–254.

Gesi, M., Soldani, P., Giorgi, F.S., Santinami, A., Bonaccorsi, I., and Fornai, F. (2000). The role of the locus coeruleus in the development of Parkinson’s disease. Neurosci. Biobehav. Rev. 24, 655–668.

Gether, U., Andersen, P.H., Larsson, O.M., and Schousboe, A. (2006). Neurotransmitter transporters: molecular function of important drug targets. Trends Pharmacol. Sci. 27, 375– 383.

Ghisi, V., Ramsey, A.J., Masri, B., Gainetdinov, R.R., Caron, M.G., and Salahpour, A. (2009). Reduced D2-mediated signaling activity and trans-synaptic upregulation of D1 and D2 dopamine receptors in mice overexpressing the dopamine transporter. Cell. Signal. 21, 87–94.

Gilman, A.G. (1987). G Proteins: Transducers of Receptor-Generated Signals. Annu. Rev. Biochem. 56, 615–649.

Giros, B., Jaber, M., Jones, S.R., Wightman, R.M., and Caron, M.G. (1996). Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking the dopamine transporter. Nature 379, 606–612.

Gispert, S., Ricciardi, F., Kurz, A., Azizov, M., Hoepken, H.-H., Becker, D., Voos, W., Leuner, K., Müller, W.E., Kudin, A.P., Kunz, W.S., Zimmermann, A., Roeper, J., Wenzel, D., Jendrach, M., García-Arencíbia, M., Fernández-Ruiz, J., Huber, L., Rohrer, H., Barrera, M., Reichert, A.S., Rüb, U., Chen, A., Nussbaum, R.L., and Auburger, G. (2009). Parkinson Phenotype in Aged PINK1-Deficient Mice Is Accompanied by Progressive Mitochondrial Dysfunction in Absence of Neurodegeneration. PLoS One 4, e5777.

Glatt, C.E., Wahner, A.D., White, D.J., Ruiz-Linares, A., and Ritz, B. (2006). Gain-of-function haplotypes in the vesicular monoamine transporter promoter are protective for Parkinson disease in women. Hum. Mol. Genet. 15, 299–305.

Glover, V., Sandler, M., Owen, F., and Riley, G.J. (1977). Dopamine is a monoamine oxidase B substrate in man. Nature 265, 80–81.

Goldberg, M.S., Fleming, S.M., Palacino, J.J., Cepeda, C., Lam, H.A., Bhatnagar, A., Meloni, E.G., Wu, N., Ackerson, L.C., Klapstein, G.J., Gajendiran, M., Roth, B.L., Chesselet, M.-F., Maidment, N.T., Levine, M.S., and Shen, J. (2003). Parkin-deficient Mice Exhibit Nigrostriatal Deficits but Not Loss of Dopaminergic Neurons. J. Biol. Chem. 278, 43628– 43635.

177

Goldstein, D.S., Sullivan, P., Holmes, C., Kopin, I.J., Basile, M.J., and Mash, D.C. (2011). Catechols in post-mortem brain of patients with Parkinson disease. Eur. J. Neurol. 18, 703– 710.

Goldstein, D.S., Sullivan, P., Holmes, C., Miller, G.W., Alter, S., Strong, R., Mash, D.C., Kopin, I.J., and Sharabi, Y. (2013). Determinants of buildup of the toxic dopamine metabolite DOPAL in Parkinson’s disease. J. Neurochem. 126, 591–603.

Gonçalves, L.L., Ramkissoon, A., and Wells, P.G. (2009). Prostaglandin H synthase-1-catalyzed bioactivation of neurotransmitters, their precursors, and metabolites: Oxidative DNA damage and electron spin resonance spectroscopy studies. Chem. Res. Toxicol. 22, 842–852.

Grace, A.A., and Bunney, B.S. (1984). The control of firing pattern in nigral dopamine neurons: burst firing. J. Neurosci. 4, 2877–2890.

Graham, D.G. (1978). Oxidative pathways for catecholamines in the genesis of neuromelanin and cytotoxic quinones. Molec Pharm 14, 633–643.

Graham, D.G., and Gutknecht, F. (1978). Autoxidation Toxicity versus Covalent Binding of Quinones as the Mechanism of Toxicity of Dopamine, 6-Hydroxydopamine, and Related Compounds toward C1300 Neuroblastoma Cells in Vitro. Molec Pharm 14, 644–653.

Greenamyre, J.T., Betarbet, R., and Sherer, T.B. (2003). The rotenone model of Parkinson’s disease: genes, environment and mitochondria. Parkinsonism Relat. Disord. 9, 59–64.

Greenamyre, J.T., Betarbet, R., Sherer, T.B., MacKenzie, G., Garcia-Osuna, M., and Panov, A. V. (2000). Chronic systemic pesticide exposure reproduces features of Parkinson’s disease. Nat. Neurosci. 3, 1301–1306.

Grenhoff, J., Nisell, M., Ferré, S., Aston-Jones, G., and Svensson, T.H. (1993). Noradrenergic modulation of midbrain dopamine cell firing elicited by stimulation of the locus coeruleus in the rat. J. Neural Transm. 93, 11–25.

Grymek, K., Łukasiewicz, S., Faron-Góreckaa, A., Tworzydlo, M., Polit, A., and Dziedzicka- Wasylewska, M. (2009). Role of silent polymorphisms within the dopamine D1 receptor associated with schizophrenia on D1-D2 receptor hetero-dimerization. Pharmacol. Rep. 61, 1024–1033.

Guillot, T.S., and Miller, G.W. (2009). Protective actions of the vesicular monoamine transporter 2 (VMAT2) in monoaminergic neurons. Mol. Neurobiol. 39, 149–170.

Guillot, T.S., Shepherd, K.R., Richardson, J.R., Wang, M.Z., Li, Y., Emson, P.C., and Miller, G.W. (2008). Reduced vesicular storage of dopamine exacerbates methamphetamine-induced neurodegeneration and astrogliosis. J Neurochem 106, 2205–2217.

Guzman, J.N., Sánchez-Padilla, J., Chan, C.S., and Surmeier, D.J. (2009). Robust Pacemaking in Substantia Nigra Dopaminergic Neurons. J. Neurosci. 29, 11011–11019.

Haavik, J., Almás, B., and Flatmark, T. (1997). Generation of Reactive Oxygen Species by

178

Tyrosine Hydroxylase: A Possible Contribution to the Degeneration of Dopaminergic Neurons? J. Neurochem. 68, 328–332.

Halliwell, B. (1992). Reactive oxygen species and the central nervous system. J. Neurochem. 59, 1609–1623.

Halliwell, B., and Gutteridget, J.M.C. (1984). Oxygen toxicity, oxygen radicals, transition metals and disease. Biochem. J 219, 1–14.

Hansen, F.H., Skjørringe, T., Yasmeen, S., Arends, N. V., Sahai, M.A., Erreger, K., Andreassen, T.F., Holy, M., Hamilton, P.J., Neergheen, V., Karlsborg, M., Newman, A.H., … Jentsch, T. (2014). Missense dopamine transporter mutations associate with adult parkinsonism and ADHD. J. Clin. Invest. 124, 3107–3120.

Hardy, J., Cai, H., Cookson, M.R., Gwinn-Hardy, K., and Singleton, A. (2006). Genetics of Parkinson’s disease and parkinsonism. Ann. Neurol. 60, 389–398.

Hasbi, A., O’Dowd, B.F., and George, S.R. (2011). Dopamine D1-D2 receptor heteromer signaling pathway in the brain: emerging physiological relevance. Mol. Brain 4, 26.

Hastings, T.G. (1995). Enzymatic oxidation of dopamine: the role of prostaglandin H synthase. J. Neurochem. 64, 919–924.

Hastings, T.G., Lewis, D.A., and Zigmond, M.J. (1996). Role of oxidation in the neurotoxic effects of intrastriatal dopamine injections. Proc. Natl. Acad. Sci. U. S. A. 93, 1956–1961.

Hastings, T.G., and Zigmond, M.J. (1994). Identification of catechol-protein conjugates in neostriatal slices incubated with [3H]dopamine: impact of ascorbic acid and glutathione. J. Neurochem. 63, 1126–1132.

Hatcher-Martin, J.M., Gearing, M., Steenland, K., Levey, A.I., Miller, G.W., and Pennell, K.D. (2012). Association between polychlorinated biphenyls and Parkinson’s disease neuropathology. Neurotoxicology 33, 1298–1304.

Hatcher, J.M., Pennell, K.D., and Miller, G.W. (2008). Parkinson’s disease and pesticides: a toxicological perspective. Trends Pharmacol. Sci. 29, 322–329.

Hatcher, J.M., Richardson, J.R., Guillot, T.S., McCormack, A.L., Di Monte, D.A., Jones, D.P., Pennell, K.D., and Miller, G.W. (2007). Dieldrin exposure induces oxidative damage in the mouse nigrostriatal dopamine system. Exp. Neurol. 204, 619–630.

Hattori, A., Luo, Y., Umegaki, H., Munoz, J., and Roth, G.S. (1998). Intrastriatal injection of dopamine results in DNA damage and apoptosis in rats. Neuroreport 9, 2569–2572.

Hersch, S.M., Yi, H., Heilman, C.J., Edwards, R.H., and Levey, A.I. (1997). Subcellular localization and molecular topology of the dopamine transporter in the striatum and substantia nigra. J. Comp. Neurol. 388, 211–227.

179

Hisahara, S., and Shimohama, S. (2011). Dopamine Receptors and Parkinson’s Disease. Int. J. Med. Chem. 2011, Article ID 403039. 16 pgs.

Hoffman, A.F., Spivak, C.E., and Lupica, C.R. (2016). Enhanced Dopamine Release by Dopamine Transport Inhibitors Described by a Restricted Diffusion Model and Fast-Scan Cyclic Voltammetry. ACS Chem. Neurosci. 7, 700–709.

Holdorff, B. (2002). Friedrich Heinrich Lewy (1885-1950) and his work. J. Hist. Neurosci. 11, 19–28.

Hornykiewicz, O. (1986). A Quarter Century of Brain Dopamine Research. In Dopaminergic Systems and Their Regulation, (London: Palgrave Macmillan UK), pp. 3–18.

Howell, L.L., and Kimmel, H.L. (2008). Monoamine transporters and psychostimulant addiction. Biochem Pharmacol 75, 196–217.

Howes, O.D., and Kapur, S. (2009). The dopamine hypothesis of schizophrenia: version III--the final common pathway. Schizophr. Bull. 35, 549–562.

Hwang, D.-Y., Fleming, S.M., Ardayfio, P., Moran-Gates, T., Kim, H., Tarazi, F.I., Chesselet, M.-F., and Kim, K.-S. (2005). 3,4-Dihydroxyphenylalanine Reverses the Motor Deficits in Pitx3-Deficient Aphakia Mice: Behavioral Characterization of a Novel Genetic Model of Parkinson’s Disease. J. Neurosci. 25, 2132–2137.

Jaber, M., Dumartin, B., Sagné, C., Haycock, J.W., Roubert, C., Giros, B., Bloch, B., and Caron, M.G. (1999). Differential regulation of tyrosine hydroxylase in the basal ganglia of mice lacking the dopamine transporter. Eur. J. Neurosci. 11, 3499–3511.

Jaber, M., Jones, S., Giros, B., and Caron, M.G. (1997). The dopamine transporter: a crucial component regulating dopamine transmission. Mov. Disord. 12, 629–633.

Janetzky, B., Hauck, S., Youdim, M.B.H., Riederer, P., Jellinger, K., Pantucek, F., Zo¨chling, R., Boissl, K.W., and Reichmann, H. (1994). Unaltered aconitase activity, but decreased complex I activity in substantia nigra pars compacta of patients with Parkinson’s disease. Neurosci. Lett. 169, 126–128.

Jankovic, J. (2008). Parkinson’s disease: clinical features and diagnosis. J. Neurol. Neurosurg. Psychiatry 79, 368–376.

Janowsky, A., Tosh, D.K., Eshleman, A.J., and Jacobson, K.A. (2016). Rigid Adenine Nucleoside Derivatives as Novel Modulators of the Human Sodium Symporters for Dopamine and Norepinephrine. J. Pharmacol. Exp. Ther. 357, 24–35.

Javitch, J.A., D’Amato, R.J., Strittmatter, S.M., and Snyder, S.H. (1985). Parkinsonism-inducing neurotoxin, N-methyl-4-phenyl-1,2,3,6 -tetrahydropyridine: uptake of the metabolite N- methyl-4-phenylpyridine by dopamine neurons explains selective toxicity. Proc. Natl. Acad. Sci. U. S. A. 82, 2173–2177.

Jellinger, K.A. (1991). Pathology of Parkinson’s disease. Changes other than the nigrostriatal

180

pathway. Mol. Chem. Neuropathol. 14, 153–197.

John, C.E., and Jones, S.R. (2007). Voltammetric characterization of the effect of monoamine uptake inhibitors and releasers on dopamine and serotonin uptake in mouse caudate-putamen and substantia nigra slices. Neuropharmacology 52, 1596–1605.

Johnson, M.A., Rajan, V., Miller, C.E., and Wightman, R.M. (2006). Dopamine release is severely compromised in the R6/2 mouse model of Huntington’s disease. J. Neurochem. 97, 737–746.

Jones, S.R., Gainetdinov, R.R., Hu, X.T., Cooper, D.C., Wightman, R.M., White, F.J., and Caron, M.G. (1999). Loss of autoreceptor functions in mice lacking the dopamine transporter. Nat. Neurosci. 2, 649–655.

Jones, S.R., Gainetdinov, R.R., Jaber, M., Giros, B., Wightman, R.M., and Caron, M.G. (1998a). Profound neuronal plasticity in response to inactivation of the dopamine transporter. Proc. Natl. Acad. Sci. U. S. A. 95, 4029–4034.

Jones, S.R., Gainetdinov, R.R., Wightman, R.M., and Caron, M.G. (1998b). Mechanisms of amphetamine action revealed in mice lacking the dopamine transporter. J. Neurosci. 18, 1979–1986.

Kanner, B.I., and Schuldiner, S. (1987). Mechanism of transport and storage of neurotransmitters. CRC Crit. Rev. Biochem. 22, 1–38.

Kebabian, J.W., and Calne, D.B. (1979). Multiple receptors for dopamine. Nature 277, 93–96.

Keeney, P.M. (2006). Parkinson’s Disease Brain Mitochondrial Complex I Has Oxidatively Damaged Subunits and Is Functionally Impaired and Misassembled. J. Neurosci. 26, 5256– 5264.

Kelada, S.N.P., Checkoway, H., Kardia, S.L.R., Carlson, C.S., Costa-Mallen, P., Eaton, D.L., Firestone, J., Powers, K.M., Swanson, P.D., Franklin, G.M., Longstreth, W.T., Weller, T.S., Afsharinejad, Z., and Costa, L.G. (2006). 5’ and 3’ region variability in the dopamine transporter gene (SLC6A3), pesticide exposure and Parkinson’s disease risk: A hypothesis- generating study. Hum. Mol. Genet. 15, 3055–3062.

Khaliq, Z.M., and Bean, B.P. (2010). Pacemaking in dopaminergic ventral tegmental area neurons: depolarizing drive from background and voltage-dependent sodium conductances. J. Neurosci. 30, 7401–7413.

Kile, B.M., Walsh, P.L., McElligott, Z.A., Bucher, E.S., Guillot, T.S., Salahpour, A., Caron, M.G., and Wightman, R.M. (2012). Optimizing the temporal resolution of fast-scan cyclic voltammetry. ACS Chem. Neurosci. 3, 285–292.

Klein, C., and Westenberger, A. (2012). Genetics of Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2, a008888.

Kobilka, B.K. (2007). G protein coupled receptor structure and activation. Biochim. Biophys.

181

Acta 1768, 794–807.

Kordower, J.H., Olanow, C.W., Dodiya, H.B., Chu, Y., Beach, T.G., Adler, C.H., Halliday, G.M., and Bartus, R.T. (2013). Disease duration and the integrity of the nigrostriatal system in Parkinson’s disease. Brain 136, 2419–2431.

Korytowski, W., Sarna, T., Kalyanaraman, B., and Sealy, R.C. (1987). Tyrosinase-catalyzed oxidation of dopa and related catechol(amine)s: a kinetic electron spin resonance investigation using spin-stabilization and spin label oximetry. Biochim. Biophys. Acta 924, 383–392.

Kristal, B.S., Conway, A.D., Brown, A.M., Jain, J.C., Ulluci, P.A., Li, S.W., and Burke, W.J. (2001). Selective dopaminergic vulnerability: 3,4-dihydroxyphenylacetaldehyde targets mitochondria. Free Radic. Biol. Med. 30, 924–931.

Kurian, M.A., Li, Y., Zhen, J., Meyer, E., Hai, N., Christen, H.-J., Hoffmann, G.F., Jardine, P., von Moers, A., Mordekar, S.R., O’Callaghan, F., Wassmer, E., Wraige, E., Dietrich, C., Lewis, T., Hyland, K., Heales, S., Sanger, T., Gissen, P., Assmann, B.E., Reith, M.E.A., and Maher, E.R. (2011). Clinical and molecular characterisation of hereditary dopamine transporter deficiency syndrome: an observational cohort and experimental study. Lancet. Neurol. 10, 54–62.

Kurian, M.A., Zhen, J., Cheng, S.-Y., Li, Y., Mordekar, S.R., Jardine, P., Morgan, N. V, Meyer, E., Tee, L., Pasha, S., Wassmer, E., Heales, S.J.R., Gissen, P., Reith, M.E.A., and Maher, E.R. (2009). Homozygous loss-of-function mutations in the gene encoding the dopamine transporter are associated with infantile parkinsonism-dystonia. J. Clin. Invest. 119, 1595– 1603.

Lai, C.-T., and Yu, P.H. (1997). Dopamine- and l-β-3,4-dihydroxyphenylalanine hydrochloriDe (l-Dopa)-induced cytotoxicity towards catecholaminergic neuroblastoma SH-SY5Y Cells. Biochem. Pharmacol. 53, 363–372.

Langston, J.W., Ballard, P., Tetrud, J.W., and Irwin, I. (1983). Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science 219, 979–980.

Langston, J.W., Irwin, I., Langston, E.B., and Forno, L.S. (1984). Pargyline prevents MPTP- induced parkinsonism in primates. Science 225, 1480–1482. de Lau, L.M., and Breteler, M.M. (2006). Epidemiology of Parkinson’s disease. Lancet Neurol. 5, 525–535.

Lawal, H.O., and Krantz, D.E. (2013). SLC18: Vesicular neurotransmitter transporters for monoamines and acetylcholine. Mol. Aspects Med. 34, 360–372.

Lee, S.P., So, C.H., Rashid, A.J., Varghese, G., Cheng, R., Lança, A.J., O’Dowd, B.F., and George, S.R. (2004). Dopamine D1 and D2 receptor Co-activation generates a novel phospholipase C-mediated calcium signal. J. Biol. Chem. 279, 35671–35678.

Leviel, V. (2001). The reverse transport of DA, what physiological significance? Neurochem. Int. 38, 83–106.

182

Levitzki, A., and Klein, S. (2002). G-Protein Subunit Dissociation Is not an Integral Part of G- Protein Action. ChemBioChem 3, 815–818.

Lin, M.T., and Beal, M.F. (2006). Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 443, 787–795.

Lindgren, N., Xu, Z.-Q.D., Herrera-Marschitz, M., Haycock, J., Hökfelt, T., and Fisone, G. (2001). Dopamine D 2 receptors regulate tyrosine hydroxylase activity and phosphorylation at Ser40 in rat striatum. Eur. J. Neurosci. 13, 773–780.

Liou, H.H., Tsai, M.C., Chen, C.J., Jeng, J.S., Chang, Y.C., Chen, S.Y., and Chen, R.C. (1997). Environmental risk factors and Parkinson’s disease: a case-control study in Taiwan. Neurology 48, 1583–1588.

Liu, Y., Peter, D., Roghani, A., Schuldiner, S., Privé, G.G., Eisenberg, D., Brecha, N., and Edwards, R.H. (1992). A cDNA that suppresses MPP+ toxicity encodes a vesicular amine transporter. Cell 70, 539–551.

Livak, K.J., and Schmittgen, T.D. (2001). Analysis of relative gene expression data using real- time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25, 402–408.

Lohr, K.M., Bernstein, A.I., Stout, K.A., Dunn, A.R., Lazo, C.R., Alter, S.P., Wang, M., Li, Y., Fan, X., Hess, E.J., Yi, H., Vecchio, L.M., Goldstein, D.S., Guillot, T.S., Salahpour, A., and Miller, G.W. (2014). Increased vesicular monoamine transporter enhances dopamine release and opposes Parkinson disease-related neurodegeneration in vivo. Proc. Natl. Acad. Sci. 111, 9977–9982.

Lohr, K.M., Chen, M., Hoffman, C.A., McDaniel, M.J., Stout, K.A., Dunn, A.R., Wang, M., Bernstein, A., and Miller, G.W. (2016). Vesicular monoamine transporter 2 (VMAT2) level regulates MPTP vulnerability and clearance of excess dopamine in mouse striatal terminals. Toxicol. Sci. 153, 79–88.

Lohr, K.M., Masoud, S.T., Salahpour, A., and Miller, G.W. (2017). Membrane transporters as mediators of synaptic dopamine dynamics: implications for disease. Eur. J. Neurosci. 45, 20– 33.

Lohr, K.M., Stout, K.A., Dunn, A.R., Wang, M., Salahpour, A., Guillot, T.S., and Miller, G.W. (2015). Increased Vesicular Monoamine Transporter 2 (VMAT2; Slc18a2) Protects against Methamphetamine Toxicity. ACS Chem. Neurosci. 6, 790–799.

Loland, C.J., Desai, R.I., Zou, M.-F., Cao, J., Grundt, P., Gerstbrein, K., Sitte, H.H., Newman, A.H., Katz, J.L., and Gether, U. (2007). Relationship between Conformational Changes in the Dopamine Transporter and Cocaine-Like Subjective Effects of Uptake Inhibitors. Mol. Pharmacol. 73, 813–823.

Lotharius, J., and Brundin, P. (2002). Pathogenesis of parkinson’s disease: dopamine, vesicles and α-synuclein. Nat. Rev. Neurosci. 3, 932–942.

Luk, K.C., Kehm, V., Carroll, J., Zhang, B., Brien, P.O., Trojanowski, J.Q., and Lee, V.M.

183

(2012). Pathological a-Synuclein Transmission in Nontransgenic Mice. Science 338, 949– 953.

Luo, Y., Hattori, A., Munoz, J., Qin, Z.H., and Roth, G.S. (1999). Intrastriatal dopamine injection induces apoptosis through oxidation-involved activation of transcription factors AP- 1 and NF-kappaB in rats. Mol. Pharmacol. 56, 254–264.

Marsden, C.D. (1994). Problems with long-term levodopa therapy for Parkinson’s disease. Clin. Neuropharmacol. 17, S32-44.

Martin, I., Dawson, V.L., and Dawson, T.M. (2011). Recent Advances in the Genetics of Parkinson’s Disease. Annu. Rev. Genomics Hum. Genet. 12, 301–325.

Martres, M.P., Demeneix, B., Hanoun, N., Hamon, M., and Giros, B. (1998). Up- and down- expression of the dopamine transporter by plasmid DNA transfer in the rat brain. Eur. J. Neurosci. 10, 3607–3616.

Masoud, S.T., Vecchio, L.M., Bergeron, Y., Hossain, M.M., Nguyen, L.T., Bermejo, M.K., Kile, B., Sotnikova, T.D., Siesser, W.B., Gainetdinov, R.R., Wightman, R.M., Caron, M.G., Richardson, J.R., Miller, G.W., Ramsey, A.J., Cyr, M., and Salahpour, A. (2015). Increased expression of the dopamine transporter leads to loss of dopamine neurons, oxidative stress and l-DOPA reversible motor deficits. Neurobiol. Dis. 74, 66–75.

Masserano, J.M., Baker, I., Venable, D., Gong, L., Zullo, S.J., Merril, C.R., and Wyatt, R.J. (2000). Dopamine induces cell death, lipid peroxidation and DNA base damage in a catecholaminergic cell line derived from the central nervous system. Neurotox. Res. 1, 171– 179.

Masserano, J.M., Gong, L., Kulaga, H., Baker, I., and Wyatt, R.J. (1996). Dopamine induces apoptotic cell death of a catecholaminergic cell line derived from the central nervous system. Mol. Pharmacol. 50, 1309–1315.

Matsuda, W., Furuta, T., Nakamura, K.C., Hioki, H., Fujiyama, F., Arai, R., and Kaneko, T. (2009). Single nigrostriatal dopaminergic neurons form widely spread and highly dense axonal arborizations in the neostriatum. J. Neurosci. 29, 444–453.

Mazei-Robison, M.S., Bowton, E., Holy, M., Schmudermaier, M., Freissmuth, M., Sitte, H.H., Galli, A., and Blakely, R.D. (2008). Anomalous dopamine release associated with a human dopamine transporter coding variant. J. Neurosci. 28, 7040–7046.

McCormack, A.L., Thiruchelvam, M., Manning-Bog, A.B., Thiffault, C., Langston, J.W., Cory- Slechta, D.A., and Di Monte, D.A. (2002). Environmental risk factors and Parkinson’s disease: selective degeneration of nigral dopaminergic neurons caused by the herbicide paraquat. Neurobiol. Dis. 10, 119–127.

McLaughlin, B.A., Nelson, D., Erecińska, M., and Chesselet, M.-F. (2002). Toxicity of Dopamine to Striatal Neurons In Vitro and Potentiation of Cell Death by a Mitochondrial Inhibitor. J. Neurochem. 70, 2406–2415.

184

Medvedev, I.O., Ramsey, A.J., Masoud, S.T., Bermejo, M.K., Urs, N., Sotnikova, T.D., Beaulieu, J.-M., Gainetdinov, R.R., and Salahpour, A. (2013). D1 dopamine receptor coupling to PLCβ regulates forward locomotion in mice. J. Neurosci. 33, 18125–18133.

De Mei, C., Ramos, M., Iitaka, C., and Borrelli, E. (2009). Getting specialized: presynaptic and postsynaptic dopamine D2 receptors. Curr. Opin. Pharmacol. 9, 53–58.

Mergy, M.A., Gowrishankar, R., Gresch, P.J., Gantz, S.C., Williams, J., Davis, G.L., Wheeler, C.A., Stanwood, G.D., Hahn, M.K., and Blakely, R.D. (2014). The rare DAT coding variant Val559 perturbs DA neuron function, changes behavior, and alters in vivo responses to psychostimulants. Proc. Natl. Acad. Sci. U. S. A. 111, E4779-88.

Michel, P.P., and Hefti, F. (1990). Toxicity of 6-hydroxydopamine and dopamine for dopaminergic neurons in culture. J. Neurosci. Res. 26, 428–435.

Miller, G.W., Erickson, J.D., Perez, J.T., Penland, S.N., Mash, D.C., Rye, D.B., and Levey, A.I. (1999a). Immunochemical analysis of vesicular monoamine transporter (VMAT2) protein in Parkinson’s disease. Exp. Neurol. 156, 138–148.

Miller, G.W., Gainetdinov, R.R., Levey, a I., and Caron, M.G. (1999b). Dopamine transporters and neuronal injury. Trends Pharmacol. Sci. 20, 424–429.

Miller, G.W., Staley, J.K., Heilman, C.J., Perez, J.T., Mash, D.C., Rye, D.B., and Levey, A.I. (1997). Immunochemical analysis of dopamine transporter protein in Parkinson’s disease. Ann. Neurol. 41, 530–539.

Missale, C., Nash, S.R., Robinson, S.W., Jaber, M., and Caron, M.G. (1998). Dopamine Receptors: From Structure to Function. Physiol. Rev. 78, 189–225.

Mizuno, Y., Ohta, S., Tanaka, M., Takamiya, S., Suzuki, K., Sato, T., Oya, H., Ozawa, T., and Kagawa, Y. (1989). Deficiencies in Complex I subunits of the respiratory chain in Parkinson’s disease. Biochem. Biophys. Res. Commun. 163, 1450–1455.

Montagu, K.A. (1957). Catechol compounds in rat tissues and in brains of different animals. Nature 180, 244–245.

Di Monte, D.A., DeLanney, L.E., Irwin, I., Royland, J.E., Chan, P., Jakowec, M.W., and Langston, J.W. (1996). Monoamine oxidase-dependent metabolism of dopamine in the striatum and substantia nigra of L-DOPA-treated monkeys. Brain Res. 738, 53–59.

Mooslehner, K.A., Chan, P.M., Xu, W., Liu, L., Smadja, C., Humby, T., Allen, N.D., Wilkinson, L.S., and Emson, P.C. (2001). Mice with very low expression of the vesicular monoamine transporter 2 gene survive into adulthood: potential mouse model for parkinsonism. Mol Cell Biol 21, 5321–5331.

Mosharov, E. V., Larsen, K.E., Kanter, E., Phillips, K.A., Wilson, K., Schmitz, Y., Krantz, D.E., Kobayashi, K., Edwards, R.H., and Sulzer, D. (2009). Interplay between Cytosolic Dopamine, Calcium, and α-Synuclein Causes Selective Death of Substantia Nigra Neurons. Neuron 62, 218–229.

185

Mosharov, E. V, Gong, L.-W., Khanna, B., Sulzer, D., and Lindau, M. (2003). Intracellular patch electrochemistry: regulation of cytosolic catecholamines in chromaffin cells. J. Neurosci. 23, 5835–5845.

Moss, J., and Bolam, J.P. (2009). The Relationship between Dopaminergic Axons and Glutamatergic Synapses in the Striatum: Structural Considerations. In Dopamine Handbook, (Oxford University Press), pp. 49–60.

Nappi, A.J., Vass, E., Prota, G., and Memoli, S. (1995). The effects of hydroxyl radical attack on dopa, dopamine, 6-hydroxydopa, and 6-hydroxydopamine. Pigment Cell Res. 8, 283–293.

Ng, J., Zhen, J., Meyer, E., Erreger, K., Li, Y., Kakar, N., Ahmad, J., Thiele, H., Kubisch, C., Rider, N.L., Morton, D.H., Strauss, K.A., Puffenberger, E.G., D’Agnano, D., Anikster, Y., Carducci, C., Hyland, K., Rotstein, M., Leuzzi, V., Borck, G., Reith, M.E.A., and Kurian, M.A. (2014). Dopamine transporter deficiency syndrome: phenotypic spectrum from infancy to adulthood. Brain 137, 1107–1119.

Nickell, J.R., Krishnamurthy, S., Norrholm, S., Deaciuc, G., Siripurapu, K.B., Zheng, G., Crooks, P.A., and Dwoskin, L.P. (2010). Lobelane inhibits methamphetamine-evoked dopamine release via inhibition of the vesicular monoamine transporter-2. J. Pharmacol. Exp. Ther. 332, 612–621.

Nickell, J.R., Siripurapu, K.B., Vartak, A., Crooks, P.A., and Dwoskin, L.P. (2014). The vesicular monoamine transporter-2: an important pharmacological target for the discovery of novel therapeutics to treat methamphetamine abuse. Adv. Pharmacol. 69, 71–106.

Norris, E.H., Giasson, B.I., and Lee, V.M.-Y. (2004). α-Synuclein: Normal Function and Role in Neurodegenerative Diseases. Curr. Top. Dev. Biol. 60, 17–54.

Nuytemans, K., Theuns, J., Cruts, M., and Van Broeckhoven, C. (2010). Genetic etiology of Parkinson disease associated with mutations in the SNCA, PARK2, PINK1, PARK7, and LRRK2 genes: A mutation update. Hum. Mutat. 31, 763–780.

Oaks, A.W., Frankfurt, M., Finkelstein, D.I., and Sidhu, A. (2013). Age-Dependent Effects of A53T Alpha-Synuclein on Behavior and Dopaminergic Function. PLoS One 8, e60378.

Obeso, J.A., Rodriguez, M.C., and DeLong, M.R. (1997). Basal ganglia pathophysiology. A critical review. Adv. Neurol. 74, 3–18.

Olson, E.J., Boeve, B.F., and Silber, M.H. (2000). Rapid eye movement sleep behaviour disorder: demographic, clinical and laboratory findings in 93 cases. Brain 123, 331–339.

Pacelli, C., Giguère, N., Bourque, M.-J., Lévesque, M., Slack, R.S., and Trudeau, L.-É. (2015). Elevated Mitochondrial Bioenergetics and Axonal Arborization Size Are Key Contributors to the Vulnerability of Dopamine Neurons. Curr. Biol. 25, 2349–2360.

Pakkenberg, B., Møller, A., Gundersen, H.J., Mouritzen Dam, A., and Pakkenberg, H. (1991). The absolute number of nerve cells in substantia nigra in normal subjects and in patients with Parkinson’s disease estimated with an unbiased stereological method. J. Neurol. Neurosurg.

186

Psychiatry 54, 30–33.

Paladini, C.A., Robinson, S., Morikawa, H., Williams, J.T., and Palmiter, R.D. (2003). Dopamine controls the firing pattern of dopamine neurons via a network feedback mechanism. Proc. Natl. Acad. Sci. 100, 2866–2871.

Paleacu, D. (2007). Tetrabenazine in the treatment of Huntington’s disease. Neuropsychiatr. Dis. Treat. 3, 545–551.

Paris, I., Lozano, J., Perez-Pastene, C., Muñoz, P., and Segura-Aguilar, J. (2009). Molecular and neurochemical mechanisms in PD pathogenesis. Neurotox. Res. 16, 271–279.

Parkinson, J. (1817). An Essay on the Shaking Palsy (London: Whittingham and Rowland).

Parsons, S.M. (2000). Transport mechanisms in acetylcholine and monoamine storage. FASEB J. 14, 2423–2434.

Partilla, J.S., Dempsey, A.G., Nagpal, A.S., Blough, B.E., Baumann, M.H., and Rothman, R.B. (2006). Interaction of amphetamines and related compounds at the vesicular monoamine transporter. J. Pharmacol. Exp. Ther. 319, 237–246.

Pei, L., Li, S., Wang, M., Diwan, M., Anisman, H., Fletcher, P.J., Nobrega, J.N., and Liu, F. (2010). Uncoupling the dopamine D1-D2 receptor complex exerts antidepressant-like effects. Nat. Med. 16, 1393–1395.

Pérez-Mañá, C., Farré, M., Pujadas, M., Mustata, C., Menoyo, E., Pastor, A., Langohr, K., and de la Torre, R. (2015). Ethanol induces hydroxytyrosol formation in humans. Pharmacol. Res. 95–96, 27–33.

Perreault, M.L., Hasbi, A., Alijaniaram, M., Fan, T., Varghese, G., Fletcher, P.J., Seeman, P., O’Dowd, B.F., and George, S.R. (2010). The dopamine D1-D2 receptor heteromer localizes in / neurons: increased high affinity state following amphetamine and in schizophrenia. J. Biol. Chem. 285, 36625–36634.

Peter, D., Liu, Y., Sternini, C., de Giorgio, R., Brecha, N., and Edwards, R.H. (1995). Differential expression of two vesicular monoamine transporters. J. Neurosci. 15, 6179–6188.

Pifl, C., Rajput, A., Reither, H., Blesa, J., Cavada, C., Obeso, J. a, Rajput, A.H., and Hornykiewicz, O. (2014). Is Parkinson’s disease a vesicular dopamine storage disorder? Evidence from a study in isolated synaptic vesicles of human and nonhuman primate striatum. J. Neurosci. 34, 8210–8218.

Pissadaki, E.K., and Bolam, J.P. (2013). The energy cost of action potential propagation in dopamine neurons: clues to susceptibility in Parkinson’s disease. Front. Comput. Neurosci. 7, 13.

Poewe, W. (2008). Non-motor symptoms in Parkinson’s disease. Eur. J. Neurol. 15 Suppl 1, 14– 20.

187

Poole, A.C., Thomas, R.E., Andrews, L.A., McBride, H.M., Whitworth, A.J., and Pallanck, L.J. (2008). The PINK1/Parkin pathway regulates mitochondrial morphology. Proc. Natl. Acad. Sci. 105, 1638–1643.

Porat, S., Premkumar, A., and Simantov, R. (2001). Dopamine induces phenotypic differentiation or apoptosis in a dose-dependent fashion: involvement of the dopamine transporter and p53. Dev. Neurosci. 23, 432–440.

Priyadarshi, A., Khuder, S.A., Schaub, E.A., and Shrivastava, S. (2000). A meta-analysis of Parkinson’s disease and exposure to pesticides. Neurotoxicology 21, 435–440.

Protais, P., Costentin, J., and Schwartz, J.C. (1976). Climbing behavior induced by apomorphine in mice: a simple test for the study of dopamine receptors in striatum. Psychopharmacology (Berl). 50, 1–6.

Przedborski, S., Przedborski, S., Naini, A.B., Naini, A.B., Akram, M., and Akram, M. (2001). The parkinson toxin 1-methyl-4-phenyl-1,2,3,6-tetrahydopyridine (MPTP): a technical review of its utility and safety. J. Neurochem. 1265–1274.

Purves, D., Augustine, G.J., Fitzpatrick, D., Hall, W.C., LaMantia, A.-S., McNamara, J.O., and White, L.E. (2008). Neuroscience 4th edition (Sinauer Associates Inc.).

Qin, K., Dong, C., Wu, G., and Lambert, N.A. (2011). Inactive-state preassembly of G(q)- coupled receptors and G(q) heterotrimers. Nat. Chem. Biol. 7, 740–747.

Rabinovic, A.D., Lewis, D.A., and Hastings, T.G. (2000). Role of oxidative changes in the degeneration of dopamine terminals after injection of neurotoxic levels of dopamine. Neuroscience 101, 67–76.

Rajput, A.H., Uitti, R.J., Stern, W., Laverty, W., O’Donnell, K., O’Donnell, D., Yuen, W.K., and Dua, A. (1987). Geography, drinking water chemistry, pesticides and herbicides and the etiology of Parkinson’s disease. Can. J. Neurol. Sci. 14, 414–418.

Ralph, R.J., Paulus, M.P., Fumagalli, F., Caron, M.G., and Geyer, M.A. (2001). Prepulse inhibition deficits and perseverative motor patterns in dopamine transporter knock-out mice: differential effects of D1 and D2 receptor antagonists. J. Neurosci. 21, 305–313.

Ramachandiran, S., Hansen, J.M., Jones, D.P., Richardson J.R., J.R., and Miller, G.W. (2007). Divergent mechanisms of paraquat, MPP+, and rotenone toxicity: Oxidation of thioredoxin and caspase-3 activation. Toxicol. Sci. 95, 163–171.

Ramkissoon, A., and Wells, P.G. (2011). Human prostaglandin H synthase (hPHS)-1- and hPHS- 2-dependent bioactivation, oxidative macromolecular damage, and cytotoxicity of dopamine, its precursor, and its metabolites. Free Radic. Biol. Med. 50, 295–304.

Ramsay, R.R., Dadgar, J., Trevor, A., and Singer, T.P. (1986). Energy-driven uptake of N- methyl-4-phenylpyridine by brain mitochondria mediates the neurotoxicity of MPTP. Life Sci. 39, 581–588.

188

Rangel-Barajas, C., Coronel, I., and Florán, B. (2015). Dopamine Receptors and Neurodegeneration. Aging Dis. 6, 349–368.

Rashid, A.J., So, C.H., Kong, M.M.C., Furtak, T., El-Ghundi, M., Cheng, R., O’Dowd, B.F., and George, S.R. (2007). D1-D2 dopamine receptor heterooligomers with unique pharmacology are coupled to rapid activation of Gq/11 in the striatum. Proc. Natl. Acad. Sci. 104, 654–659.

Rees, J.N., Florang, V.R., Eckert, L.L., and Doorn, J.A. (2009). Protein reactivity of 3,4- dihydroxyphenylacetaldehyde, a toxic dopamine metabolite, is dependent on both the aldehyde and the catechol. Chem. Res. Toxicol. 22, 1256–1263.

Reith, M.E.A., Blough, B.E., Hong, W.C., Jones, K.T., Schmitt, K.C., Baumann, M.H., Partilla, J.S., Rothman, R.B., and Katz, J.L. (2015). Behavioral, biological, and chemical perspectives on atypical agents targeting the dopamine transporter. Drug Alcohol Depend. 147, 1–19.

Reynolds, J.N.J., Hyland, B.I., and Wickens, J.R. (2001). A cellular mechanism of reward- related learning. Nature 413, 67–70.

Richardson, J.R., and Miller, G.W. (2004). Acute exposure to aroclor 1016 or 1260 differentially affects dopamine transporter and vesicular monoamine transporter 2 levels. Toxicol. Lett. 148, 29–40.

Richardson, J.R., Quan, Y., Sherer, T.B., Greenamyre, J.T., and Miller, G.W. (2005). Paraquat neurotoxicity is distinct from that of MPTP and rotenone. Toxicol. Sci. 88, 193–201.

Rilstone, J.J., Alkhater, R.A., and Minassian, B.A. (2013). Brain Dopamine–Serotonin Vesicular Transport Disease and Its Treatment. N. Engl. J. Med. 368, 543–550.

Rio, D.C., Ares, M., Hannon, G.J., and Nilsen, T.W. (2010). Purification of RNA Using TRIzol (TRI Reagent). Cold Spring Harb. Protoc. 6, pdb.prot5439.

Ritz, B., Rhodes, S.L., Qian, L., Schernhammer, E., Olsen, J.H., and Friis, S. (2010). L-type calcium channel blockers and parkinson disease in Denmark. Ann. Neurol. 67, 600–606.

Ritz, B.R., Manthripragada, A.D., Costello, S., Lincoln, S.J., Farrer, M.J., Cockburn, M., and Bronstein, J. (2009). Dopamine transporter genetic variants and pesticides in Parkinson’s disease. Environ. Health Perspect. 117, 964–969.

Ritz, M.C., Lamb, R.J., Goldberg, S.R., and Kuhar, M.J. (1987). Cocaine receptors on dopamine transporters are related to self-administration of cocaine. Science 237, 1219–1223.

Rommelfanger, K.S., Edwards, G.L., Freeman, K.G., Liles, L.C., Miller, G.W., and Weinshenker, D. (2007). Norepinephrine loss produces more profound motor deficits than MPTP treatment in mice. Proc. Natl. Acad. Sci. U. S. A. 104, 13804–13809.

Rosei, M.A., Blarzino, C., Foppoli, C., Mosca, L., and Coccia, R. (1994). Lipoxygenase- catalyzed oxidation of catecholamines. Biochem. Biophys. Res. Commun. 200, 344–350.

Sahu, A., Tyeryar, K.R., Vongtau, H.O., Sibley, D.R., and Undieh, A.S. (2009). D5 Dopamine

189

Receptors are Required for Dopaminergic Activation of Phospholipase C. Mol. Pharmacol. 75, 447–453.

Salahpour, A., Ramsey, A.J., Medvedev, I.O., Kile, B., Sotnikova, T.D., Holmstrand, E., Ghisi, V., Nicholls, P.J., Wong, L., Murphy, K., Sesack, S.R., Wightman, R.M., Gainetdinov, R.R., and Caron, M.G. (2008). Increased amphetamine-induced hyperactivity and reward in mice overexpressing the dopamine transporter. Proc. Natl. Acad. Sci. U. S. A. 105, 4405–4410.

Salvatore, M.F., Calipari, E.S., and Jones, S.R. (2016). Regulation of tyrosine hydroxylase expression and phosphorylation in dopamine transporter-deficient mice. ACS Chem. Neurosci. 7, 941–951.

Schapira, A.H., Cooper, J.M., Dexter, D., Clark, J.B., Jenner, P., and Marsden, C.D. (1990). Mitochondrial complex I deficiency in Parkinson’s disease. J. Neurochem. 54, 823–827.

Schenck, C.H., Bundlie, S.R., and Mahowald, M.W. (1996). Delayed emergence of a parkinsonian disorder in 38% of 29 older men initially diagnosed with idiopathic rapid eye movement sleep behaviour disorder. Neurology 46, 388–393.

Schmitt, K.C., and Reith, M.E.A. (2011). The atypical stimulant and nootropic modafinil interacts with the dopamine transporter in a different manner than classical cocaine-like inhibitors. PLoS One 6, e25790.

Schmitt, K.C., Rothman, R.B., and Reith, M.E.A. (2013). Nonclassical pharmacology of the dopamine transporter: atypical inhibitors, allosteric modulators, and partial substrates. J. Pharmacol. Exp. Ther. 346, 2–10.

Schober, A. (2004). Classic toxin-induced animal models of Parkinson’s disease: 6-OHDA and MPTP. Cell Tissue Res. 318, 215–224.

Schultz, W., Dayan, P., and Montague, P.R. (1997). A Neural Substrate of Prediction and Reward. Science (80-. ). 275.

Seeman, P., Chau-Wong, M., Tedesco, J., and Wong, K. (1976). Dopamine receptors in human and calf brains, using [3H]apomorphine and an antipsychotic drug. Proc. Natl. Acad. Sci. U. S. A. 73, 4354–4358.

Semchuk, K.M., Love, E.J., and Lee, R.G. (1992). Parkinson’s disease and exposure to agricultural work and pesticide chemicals. Neurology 42, 1328–1335.

Shulman, J.M., De Jager, P.L., and Feany, M.B. (2011). Parkinson’s Disease: Genetics and Pathogenesis. Annu. Rev. Pathol. Mech. Dis. 6, 193–222.

Sian, J., Dexter, D.T., Lees, A.J., Daniel, S., Agid, Y., Javoy-Agid, F., Jenner, P., and Marsden, C.D. (1994). Alterations in glutathione levels in Parkinson’s disease and other neurodegenerative disorders affecting basal ganglia. Ann. Neurol. 36, 348–355.

Simantov, R., Blinder, E., Ratovitski, T., Tauber, M., Gabbay, M., and Porat, S. (1996). Dopamine-induced apoptosis in human neuronal cells: inhibition by nucleic acids antisense to

190

the dopamine transporter. Neuroscience 74, 39–50.

Simola, N., Morelli, M., and Carta, A.R. (2007). The 6-hydroxydopamine model of Parkinson’s disease. Neurotox. Res. 11, 151–167.

Skirboll, L., Grace, A., and Bunney, B. (1979). Dopamine auto- and postsynaptic receptors: electrophysiological evidence for differential sensitivity to dopamine agonists. Science 206, 80–82.

So, C.H., Varghese, G., Curley, K.J., Kong, M.M.C., Alijaniaram, M., Ji, X., Nguyen, T., O’Dowd, B.F., and George, S.R. (2005). D1 and D2 Dopamine Receptors Form Heterooligomers and Cointernalize after Selective Activation of Either Receptor. Mol. Pharmacol. 68, 568–578.

Sofic, E., Riederer, P., Heinsen, H., Beckmann, H., Reynolds, G.P., Hebenstreit, G., and Youdim, M.B. (1988). Increased iron (III) and total iron content in post mortem substantia nigra of parkinsonian brain. J. Neural Transm. 74, 199–205.

Sotnikova, T.D., Beaulieu, J.-M., Gainetdinov, R.R., and Caron, M.G. (2006). Molecular biology, pharmacology and functional role of the plasma membrane dopamine transporter. CNS Neurol. Disord. Drug Targets 5, 45–56.

Sotnikova, T.D., Beaulieu, J.M., Barak, L.S., Wetsel, W.C., Caron, M.G., and Gainetdinov, R.R. (2005). Dopamine-independent locomotor actions of amphetamines in a novel acute mouse model of parkinson disease. PLoS Biol. 3, e271.

Souza, J.M., Giasson, B.I., Lee, V.M., and Ischiropoulos, H. (2000). Chaperone-like activity of synucleins. FEBS Lett. 474, 116–119.

Specht, C.G., and Schoepfer, R. (2001). Deletion of the alpha-synuclein locus in a subpopulation of C57BL/6J inbred mice. BMC Neurosci. 2, 11.

Spencer, J.P.E., Jenner, P., Daniel, S.E., Lees, A.J., Marsden, D.C., and Halliwell, B. (1998). Conjugates of Catecholamines with Cysteine and GSH in Parkinson’s Disease: Possible Mechanisms of Formation Involving Reactive Oxygen Species. J. Neurochem. 71, 2112– 2122.

Spencer, J.P.E., Whiteman, M., Jenner, P., and Halliwell, B. (2002). 5-s-Cysteinyl-conjugates of catecholamines induce cell damage, extensive DNA base modification and increases in caspase-3 activity in neurons. J. Neurochem. 81, 122–129.

Spillantini, M.G., Schmidt, M.L., Lee, V.M., Trojanowski, J.Q., Jakes, R., and Goedert, M. (1997). Alpha-synuclein in Lewy bodies. Nature 388, 839–840.

Stanford Medicine (2016). Sensorimotor Tests - Behavioral and Functional Neuroscience Laboratory. http://med.stanford.edu/sbfnl/services/bm/sm.html

Stark, A.K., and Pakkenberg, B. (2004). Histological changes of the dopaminergic nigrostriatal system in aging. Cell Tissue Res. 318, 81–92.

191

Stern-Bach, Y., Keen, J.N., Bejerano, M., Steiner-Mordoch, S., Wallach, M., Findlay, J.B., and Schuldiner, S. (1992). Homology of a vesicular amine transporter to a gene conferring resistance to 1-methyl-4-phenylpyridinium. Proc. Natl. Acad. Sci. U. S. A. 89, 9730–9733.

Stoelting Co. (2017). Neuroscience Physiology Research Equipment - Psychological Assessment. http://www.stoeltingco.com/elevated-plus-maze.html

Stokes, A.H., Hastings, T.G., and Vrana, K.E. (1999). Cytotoxic and genotoxic potential of dopamine. J. Neurosci. Res. 55, 659–665.

Stokes, A.H., Lewis, D.Y., Lash, L.H., Jerome, W.G., Grant, K.W., Aschner, M., and Vrana, K.E. (2000). Dopamine toxicity in neuroblastoma cells: role of glutathione depletion by l- BSO and apoptosis. Brain Res. 858, 1–8.

Sulzer, D. (2001). alpha-synuclein and cytosolic dopamine: stabilizing a bad situation. Nat. Med. 7, 1280–1282.

Sulzer, D. (2007). Multiple hit hypotheses for dopamine neuron loss in Parkinson’s disease. Trends Neurosci. 30, 244–250.

Sulzer, D. (2011). How Addictive Drugs Disrupt Presynaptic Dopamine Neurotransmission. Neuron 69, 628–649.

Sulzer, D., Chen, T.K., Lau, Y.Y., Kristensen, H., Rayport, S., and Ewing, A. (1995). Amphetamine redistributes dopamine from synaptic vesicles to the cytosol and promotes reverse transport. J. Neurosci. 15, 4102–4108.

Sulzer, D., Sonders, M.S., Poulsen, N.W., and Galli, A. (2005). Mechanisms of neurotransmitter release by amphetamines: A review. Prog. Neurobiol. 75, 406–433.

Surmeier, D.J., Guzman, J.N., Sanchez-Padilla, J., and Goldberg, J.A. (2010). Chapter 4 – What causes the death of dopaminergic neurons in Parkinson’s disease? Prog. Brain Res. 183, 59– 77.

Takahashi, N., Miner, L.L., Sora, I., Ujike, H., Revay, R.S., Kostic, V., Jackson-Lewis, V., Przedborski, S., and Uhl, G.R. (1997). VMAT2 knockout mice: heterozygotes display reduced amphetamine-conditioned reward, enhanced amphetamine locomotion, and enhanced MPTP toxicity. Proc. Natl. Acad. Sci. U. S. A. 94, 9938–9943.

Tanaka, M., Sotomatsu, A., Kanai, H., and Hirai, S. (1991). Dopa and dopamine cause cultured neuronal death in the presence of iron. J. Neurol. Sci. 101, 198–203.

Tanner, C.M., Kamel, F., Ross, G.W., Hoppin, J.A., Goldman, S.M., Korell, M., Marras, C., Bhudhikanok, G.S., Kasten, M., Chade, A.R., Comyns, K., Richards, M.B., Meng, C., Priestley, B., Fernandez, H.H., Cambi, F., Umbach, D.M., Blair, A., Sandler, D.P., and Langston, J.W. (2011). Rotenone, paraquat, and Parkinson’s disease. Environ. Health Perspect. 119, 866–872.

Taylor, J.R., Lawrence, M.S., Redmond, D.E., Elsworth, J.D., Roth, R.H., Nichols, D.E., and

192

Mailman, R.B. (1991). Dihydrexidine, a full dopamine D1 agonist, reduces MPTP-induced parkinsonism in monkeys. Eur. J. Pharmacol. 199, 389–391.

Taylor, T.N., Alter, S.P., Wang, M., Goldstein, D.S., and Miller, G.W. (2014). Reduced vesicular storage of catecholamines causes progressive degeneration in the locus ceruleus. Neuropharmacology 76, 97–105.

Taylor, T.N., Caudle, W.M., Shepherd, K.R., Norrian, A., Jackson, C.R., Iuvone, P.M., Weinshenker, D., Greene, J.G., Miller, G.W., Noorian, A., Jackson, C.R., Iuvone, P.M., Weinshenker, D., Greene, J.G., and Miller, G.W. (2009). Nonmotor symptoms of Parkinson’s disease revealed in an animal model with reduced monoamine storage capacity. J. Neurosci. 29, 8103–8113.

The Deep-Brain Stimulation for Parkinson’s Disease Study Group (2001). Deep-Brain Stimulation of the Subthalamic Nucleus or the Pars Interna of the Globus Pallidus in Parkinson’s Disease. N. Engl. J. Med. 345, 956–963.

Torres, G.E., Gainetdinov, R.R., and Caron, M.G. (2003). Plasma membrane monoamine transporters: structure, regulation and function. Nat. Rev. Neurosci. 4, 13–25.

Del Tredici, K., and Braak, H. (2013). Dysfunction of the locus coeruleus-norepinephrine system and related circuitry in Parkinson’s disease-related dementia. J. Neurol. Neurosurg. Psychiatry 84, 774–783.

Tse, D.C.S., McCreery, R.L., and Adams, R.N. (1976). Potential oxidative pathways of brain catecholamines. J. Med. Chem. 19, 37–40.

Uhl, G.R., Walther, D., Mash, D., Faucheux, B., and Javoy-Agid, F. (1994). Dopamine transporter messenger RNA in Parkinson’s disease and control substantia nigra neurons. Ann. Neurol. 35, 494–498.

Vandenbergh, D.J., Thompson, M.D., Cook, E.H., Bendahhou, E., Nguyen, T., Krasowski, M.D., Zarrabian, D., Comings, D., Sellers, E.M., Tyndale, R.F., George, S.R., O’Dowd, B.F., and Uhl, G.R. (2000). Human dopamine transporter gene: coding region conservation among normal, Tourette’s disorder, and attention-deficit hyperactivity disorder populations. Mol. Psychiatry 5, 283–292.

Velez-Pardo, C., Jimenez Del Rio, M., Verschueren, H., Ebinger, G., and Vauquelin, G. (1997). Dopamine and iron induce apoptosis in PC12 cells. Pharmacol. Toxicol. 80, 76–84.

Wanat, M.J., Willuhn, I., Clark, J.J., and Phillips, P.E.M. (2009). Phasic dopamine release in appetitive behaviors and drug addiction. Curr. Drug Abuse Rev. 2, 195–213.

Wang, Y.-M., Gainetdinov, R.R., Fumagalli, F., Xu, F., Jones, S.R., Bock, C.B., Miller, G.W., Wightman, R.M., and Caron, M.G. (1997). Knockout of the Vesicular Monoamine Transporter 2 Gene Results in Neonatal Death and Supersensitivity to Cocaine and Amphetamine. Neuron 19, 1285–1296.

Weed, M.R., and Woolverton, W.L. (1995). The reinforcing effects of dopamine D1 receptor

193

agonists in rhesus monkeys. J. Pharmacol. Exp. Ther. 275, 1367–1374.

Weiss, S., Nosten-Bertrand, M., McIntosh, J.M., Giros, B., and Martres, M.-P. (2007). improves cognitive deficits of dopamine transporter knockout mice without long-term tolerance. Neuropsychopharmacology 32, 2465–2478.

Wilhelm, C.J., Johnson, R.A., Eshleman, A.J., and Janowsky, A. (2008). Lobeline effects on tonic and methamphetamine-induced dopamine release. Biochem. Pharmacol. 75, 1411–1415.

Wilson, W.W., Shapiro, L.P., Bradner, J.M., and Caudle, W.M. (2014). Developmental exposure to the organochlorine insecticide endosulfan damages the nigrostriatal dopamine system in male offspring. Neurotoxicology 44, 279–287.

Wimalasena, K. (2011). Vesicular monoamine transporters: structure-function, pharmacology, and medicinal chemistry. Med. Res. Rev. 31, 483–519.

Winklhofer, K.F., and Haass, C. (2010). Mitochondrial dysfunction in Parkinson’s disease. Biochim. Biophys. Acta - Mol. Basis Dis. 1802, 29–44.

Wooten, G.F., Currie, L.J., Bovbjerg, V.E., Lee, J.K., and Patrie, J. (2004). Are men at greater risk for Parkinson’s disease than women? J. Neurol. Neurosurg. Psychiatry 75, 637–639.

Xu, J., Kao, S.-Y., Lee, F.J.S., Song, W., Jin, L.-W., and Yankner, B.A. (2002). Dopamine- dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease. Nat. Med. 8, 600–606.

Yamaguchi, H., and Shen, J. (2007). Absence of dopaminergic neuronal degeneration and oxidative damage in aged DJ-1-deficient mice. Mol. Neurodegener. 2, 10.

Yamashita, M., Fukushima, S., Shen, H., Hall, F.S., Uhl, G.R., Numachi, Y., Kobayashi, H., and Sora, I. (2006). Norepinephrine transporter blockade can normalize the prepulse inhibition deficits found in dopamine transporter knockout mice. Neuropsychopharmacology 31, 2132– 2139.

Yan, M.H., Wang, X., and Zhu, X. (2013). Mitochondrial defects and oxidative stress in Alzheimer disease and Parkinson disease. Free Radic. Biol. Med. 62, 90–101.

Yorgason, J.T., España, R.A., and Jones, S.R. (2011). Demon voltammetry and analysis software: analysis of cocaine-induced alterations in dopamine signaling using multiple kinetic measures. J. Neurosci. Methods 202, 158–164.

Yoritaka, A., Hattori, N., Uchida, K., Tanaka, M., Stadtman, E.R., and Mizuno, Y. (1996). Immunohistochemical detection of 4-hydroxynonenal protein adducts in Parkinson disease. Proc. Natl. Acad. Sci. U. S. A. 93, 2696–2701.

Zarow, C., Lyness, S.A., Mortimer, J.A., and Chui, H.C. (2003). Neuronal loss is greater in the locus coeruleus than nucleus basalis and substantia nigra in Alzheimer and Parkinson diseases. Arch. Neurol. 60, 337–341.

194

Zhai, D., Li, S., Zhao, Y., and Lin, Z. (2014). SLC6A3 is a risk factor for Parkinson’s disease: A meta-analysis of sixteen years’ studies. Neurosci. Lett. 564, 99–104.

Zhang, J., Perry, G., Smith, M.A., Robertson, D., Olson, S.J., Graham, D.G., and Montine, T.J. (1999). Parkinson’s Disease Is Associated with Oxidative Damage to Cytoplasmic DNA and RNA in Substantia Nigra Neurons. Am. J. Pathol. 154, 1423–1429.

Zigmond, M.J., Hastings, T.G., and Perez, R.G. (2002). Increased dopamine turnover after partial loss of dopaminergic neurons: Compensation or toxicity? Park. Relat. Disord. 8, 389– 393.

Zilkha-Falb, R., Ziv, I., Nardi, N., Offen, D., Melamed, E., and Barzilai, A. (1997). Monoamine- Induced Apoptotic Neuronal Cell Death. Cell. Mol. Neurobiol. 17, 101–118.

Ziv, I., Melamed, E., Nardi, N., Luria, D., Achiron, A., Offen, D., and Barzilai, A. (1994). Dopamine induces apoptosis-like cell death in cultured chick sympathetic neurons--a possible novel pathogenetic mechanism in Parkinson’s disease. Neurosci. Lett. 170, 136–140.

195

Appendix 1

Project 2: Additional Experiments

A) In addition to the dorsal striatum, evoked-dopamine release and uptake were also assessed in the nucleus accumbens of DAT VMAT2 mice using FSCV (slice preparations). Similar to the dorsal striatum, dopamine release in the nucleus accumbens was significantly lower in DAT-tg, VMAT2-kd and DAT-tg/VMAT2-kd mice in comparison to WT animals (Appendix Fig 1.1B). However, in the nucleus accumbens (Appendix Figure 1.1A), overall dopamine levels across all genotypes appeared lower than the dorsal striatum (Fig 3.19 A), which is expected given the size and dopaminergic innervation of these regions. No differences were observed in dopamine clearance using tau (Appendix Fig 1.1C). Hoffman analysis of release and uptake were not performed due to multiple reasons: 1) lower evoked- dopamine levels combined with quick uptake yields trace shapes that are not ideal for Hoffman modeling in the nucleus accumbens, and 2) a caveat of using Hoffman modeling is that individual data points must meet certain criteria in order to be included in the analysis and as a result, some data had to be excluded yielding a sample size that is too low for statistical analysis.

196

Appendix Figure 1.1. Electrically-evoked dopamine release and uptake in the nucleus accumbens determined by FSCV in slice preparations. (A) Traces of dopamine currents recorded over time following a single-pulse stimulation. The ascending curve represents dopamine release while the descending curve represents dopamine clearance. Dopamine release is estimated by (B) peak amplitude (N=4-5). Dopamine clearance is estimated by (C) the decay time constant, tau (N=4-5). Data presented as mean ± SEM. Statistical differences are in comparison to WT mice. *p<0.05; **p<0.01.

197

B) For most dopaminergic drugs, the response of DAT VMAT2 mice was measured behaviorally using motor outputs. However, in the case of cocaine, we also performed FSCV measures of dopamine release and uptake in the dorsal striatum from brain slices. This experiment was performed in collaboration with Dr. Miller at Emory University. As before, redundant measures are used to estimate dopamine release (peak amplitude and Hoffman analysis of release) and uptake (tau and Hoffman parameter for clearance). Appendix Figure 1.2 shows evoked-dopamine release and uptake in response to 3 ascending concentrations of cocaine. These concentrations in slice preparations roughly correlate with 5-20 mg/kg of cocaine i.p. in mice (Johnson et al., 2006; John and Jones, 2007; Yorgason et al., 2011). In general, higher concentrations of cocaine tend to increase dopamine release (Appendix Fig 1.2 E, F) and decrease uptake (Appendix Fig 1.2 G, H) as a result of DAT inhibition. Since baseline parameters are vastly different across the 4 genotypes (see 0 µM cocaine races in Appendix Fig 1.2 A-D), release and uptake data were normalized to 0 µM cocaine and expressed as fold change (Appendix Fig 1.2 E-H). In the normalized data, dopamine release in DAT-tg mice was significantly more responsive to cocaine (Appendix Fig 1.2 E) than other genotypes. This result may reflect the fact that cocaine acts on the transporter that is over-expressed in DAT-tg mice. However, interestingly, DAT-tg mice are not different from WT animals in their behavioral response to cocaine as previously shown (Fig 3.36 B), suggesting that other factors aside from evoked-dopamine release may contribute to cocaine- induced locomotor response (such as up-regulation of dopamine receptors).

198

199

Appendix Figure 1.2. Effects of cocaine on electrically-evoked dopamine release and uptake in the dorsal striatum. These experiments were performed on brain slices using FSCV. Traces of dopamine currents in the presence of ascending doses of cocaine (1, 3 and 10 µM) were recorded over time following electrical stimulation in (A) WT, (B) DAT-tg, (C) VMAT2-kd and (D) DAT-tg/VMAT2-kd mice. The ascending curve represents dopamine release while the descending curve represents dopamine clearance. Dopamine release is estimated by (E) peak amplitude (N=4-5) and (F) the Hoffman parameter, r/ke (N=2-4). Dopamine uptake/clearance is estimated by (G) the decay time constant, tau (N=3-5) and (H) the Hoffman parameter, ku (N=3- 4). Data were normalized to 0 µM cocaine for each genotype. Results are presented as mean ± SEM. Statistical differences are in comparison to WT mice. ***p<0.001.

C) In the last experiment of this section, we conducted a pilot study to evaluate whether the behavioral effects of genotypic DAT over-expression and VMAT2 under-expression could be mimicked pharmacologically. DAT-tg/VMAT2-kd mice display several unique behaviors, the most striking of which is their basal hyperactivity. Typically, VMAT2 inhibitors such as reserpine or tetrabenazine, suppress locomotor activity in WT mice due to diminished neurotransmitter packaging which depletes vesicular release (Colpaert, 1987). However, our results show that, on a background of DAT over-expression, genetically reducing VMAT2 function produces the opposite effect - increasing locomotor activity. Therefore, in this preliminary experiment, tetrabenazine was used to assess whether pharmacological reduction of VMAT2 activity on a background of DAT over-expression (DAT-tg mice) would replicate the genetic condition of DAT-tg/VMAT2-kd mice. In particular, only WT and DAT-tg mice were treated with tetrabenazine since VMAT2-kd mice genetically express low levels of VMAT2. As shown in Appendix Figure 1.3, WT and DAT-tg mice behave similarly in response to tetrabenazine or vehicle. The VMAT2 inhibitor certainly does not seem to produce the extent of hyperactivity witnessed in DAT- tg/VMAT2-kd mice (Fig. 3.26). Another dose or preparation of tetrabenazine is needed to differentiate between drug and vehicle treatment. Nonetheless, these preliminary data suggest that long-term, genetic VMAT2-knockdown is required to produce basal hyperactivity in DAT-tg/VMAT2-kd mice that cannot be recapitulated with a single, acute dose of a VMAT2 inhibitor in DAT-tg mice.

200

Appendix Figure 1.3. Effect of tetrabenazine on locomotor activity of WT and DAT-tg mice. WT and DAT-tg mice were habituated to the activity chamber for 30 minutes, injected with the VMAT2 inhibitor, tetrabenazine (2 mg/kg i.p.) or vehicle solution consisting of 20% DMSO in PBS, and monitored for an additional 90 minutes (N=6-8). (A) Distance traveled over time. Arrow denotes time of injection. (B) Sum of total distance traveled after drug administration. Data are presented as mean ± SEM.

201

Appendix 2

Project 1: DAT-tg mice with low number of DAT copies

Midway through Project 1, we discovered that our mouse colony was housing two strains of DAT-tg mice: one with an expected “high” number of DAT copies and one with an unexpected “low” number of transgenic DAT copies. DAT copy number was determined using genomic quantitative PCR (qPCR) as shown below.

Appendix Figure 2.1. DAT copy number in WT and DAT-tg mice as assessed by genomic qPCR. All results were normalized to WT mice (shown in blue) which contain 2 endogenous copies of DAT. Number of DAT copies in DAT-tg mice was highly variable - ranging from 3-10 total copies. Based on copy number, DAT-tg mice were divided into 2 groups: 1) high copy number (shown in green, 6 or more total copies) and 2) low copy number (shown in red, typically containing 3 total copies). Experiment performed by Wendy Horsfall.

Initially, differences in DAT copy number were not detected because our routine mouse genotyping was performed using PCR (instead of qPCR), which is unable to differentiate between copy number. However, once qPCR revealed the presence of low copy DAT-tg mice, we retroactively tested genomic samples from all the mice that were in our colony since it was started. We discovered that one of the four founders that was brought to the University of

202

Toronto from Duke University (where DAT-tg mice were first generated), was a low copy DAT- tg mouse. Breeding that low copy DAT-tg mouse gave rise to more such mice in the colony. Thus far, all the results presented in this thesis used “high” copy number DAT-tg mice (between 6-10 total DAT copies as assessed by qPCR). However, some experiments were also inadvertently conducted on “low” copy number DAT-tg mice (3 total copies: 2 endogenous and 1 BAC) before they were discovered in the colony. Results from these experiments are shown in Appendix 2. Importantly, the number of DAT copies had significant impact on the phenotypes observed.

First, the effect of MPTP administration on dopamine tissue content was assessed in the striatum of low copy DAT-tg mice (Appendix Figure 2.2). No differences were observed between WT and DAT-tg mice at any dose, unlike previous results from high copy DAT-tg mice (Figure 3.11). Notably, saline-treated high copy DAT-tg mice showed 25% reduction in dopamine tissue levels compared to saline-treated WT mice (Fig 3.11) – an effect that is no longer observed in low-copy DAT-tg mice (Appendix Figure 2.2). This confirms that low copy DAT-tg mice do not show baseline changes in dopamine tissue content or enhanced sensitivity to MPTP-induced toxicity (unlike high copy DAT-tg mice).

Appendix Figure 2.2. Striatal dopamine tissue content is shown for WT and low copy DAT-tg mice treated with saline, 15 or 30 mg/kg of MPTP (n = 6-8). Data shown are means ± SEM.

203

Second, low copy DAT-tg mice were treated with rotenone, a pesticide associated with increased risk of Parkinson’s disease. Osmotic minipumps containing 7mg/kg/day of rotenone were surgically implanted in the subcutaneous cavity of WT and low copy DAT-tg mice. Drug infusion was continued for 28 days after which animals were sacrificed and brains were harvested to examine markers of dopaminergic damage in the striatum. Similar to MPTP results, no differences were observed between WT and low copy DAT-tg mice in dopamine tissue content (Appendix Figure 2.3 A) or TH immunofluorescence (Appendix Figure 2.3 C), when treated with rotenone or vehicle. In general, this regiment of rotenone administration seemed to be ineffective in producing toxicity since rotenone-treated WT mice were indistinguishable from their vehicle-treated counterparts. However, once again, low-copy DAT-tg mice lacked the reduction in dopamine tissue content (Appendix Figure 2.3 A) that was previously observed in high copy DAT-tg animals (see saline treatment in Figure 3.11).

Appendix Figure 2.3. Effect of rotenone treatment (7 mg/kg, 28-day infusion) on low copy DAT-tg mice. Rotenone is highly lipophilic therefore is was dissolved in a vehicle solution consisting of polyethylene glycol and DMSO in a 1:1 ratio. (A) Striatal dopamine tissue content after drug treatment (N=10-18). (B) Representative TH-labeled coronal images of the striatum

204

(one half) and (C) quantification of TH immunofluorescence (N=5-7), following drug treatment. Data presented as mean ± SEM.

Third, protein carbonylation was evaluated in untreated low-copy DAT-tg mice as a general marker of oxidative stress. Unlike the previous experiments where low copy DAT-tg mice did not demonstrate evidence of toxicity (Appendix Figures 2.2-2.3), in this experiment, low copy DAT-tg mice showed elevated levels of protein carbonylation, suggesting presence of general oxidative stress (Appendix Figure 2.4). Interestingly, high copy DAT-tg mice did not display any changes in protein carbonylation (Figure 3.4) when previously tested.

Appendix Figure 2.4. Protein carbonyl levels assessed in the striatum of low copy DAT-tg mice. (A) Western blot and (B) quantification of protein carbonyls in synaptic plasma membrane fractions from the striatum of WT and DAT-tg mice. Striata from 3-4 mice were pooled per sample. Data presented as mean ± SEM. *p<0.05.

In summary, DAT-tg mice with low number of DAT copies are significantly different from the original (high copy number) DAT-tg mice. Although low copy DAT-tg mice do not demonstrate reduced dopamine tissue content or enhanced sensitivity to MPTP insult, they show evidence of general oxidative damage. This suggests that one extra copy of DAT may lead to moderate oxidative stress in the absence of dopaminergic damage.

205

Project 2: DAT-tg mice with low number of DAT copies

In the DAT VMAT2 colony, routine screening of genomic DAT copy number independently identified low copy DAT-tg mice once again. This was a separate occurrence to the low copy DAT-tg mice found in Project 1. In this case, when we retroactively traced the emergence of the first low copy DAT-tg mouse in the DAT VMAT2 colony, it was found to have a high copy DAT-tg parent and littermates, suggesting that the loss of copy number occurred spontaneously. The variability in DAT copy number is shown below.

Appendix Figure 2.5. DAT copy number in DAT-tg mice from the DAT VMAT2 colony as determined by genomic qPCR. All results were normalized to WT mice (shown in blue) which contain 2 endogenous copies of DAT. Low copy DAT-tg mice from Project 1 (shown in orange, 3 total copies) were used as controls. DAT-tg mice from the DAT VMAT2 colony (shown in green) demonstrate variability in the number of DAT copies. In particular, 2 mice were found to have only 3 DAT copies (low copy mice, circled in red). Data presented as mean ± SD.

When baseline dopamine tissue content was assessed in the striatum of DAT VMAT2 mice, a mixture of 8 high and 2 low copy DAT-tg mice were used unintentionally. In this initial

206 experiment, DAT-tg mice showed a 17% reduction in dopamine tissue content compared to WT animals (Appendix Figure 2.6 A). However, once the 2 low copy DAT-tg mice were removed from the analysis, the difference rose to 21% (Appendix Figure 2.6 B). This suggests that including low copy DAT-tg mice dampened the effects of DAT over-expression on dopamine tissue content.

Appendix Figure 2.6. Comparison of striatal dopamine tissue content in DAT-tg mice with (A) mixed high and low DAT copy numbers versus (B) only high DAT copy numbers. A few VMAT2-kd mice were included as experimental controls. Data presented as mean ± SD. Statistical difference assessed between WT and DAT-tg mice. ***p<0.001.

Noting the apparent effect of DAT copy number on dopamine tissue content, a linear correlation between these two parameters was evaluated using WT, high-copy DAT-tg and low copy DAT- tg mice from the DAT VMAT2 colony (Appendix Figure 2.7). A significant negative correlation was found: higher DAT copy number was associated with lower dopamine tissue content. This suggests that greater DAT over-expression (high copy number) leads to greater dopaminergic damage (loss of dopamine tissue content) in mice, possibly due to the deleterious effects of cytosolic dopamine accumulation.

207

Appendix Figure 2.7. Linear regression of DAT copy number and dopamine tissue content assessed in WT, high-copy DAT-tg and low copy DAT-tg mice from the DAT VMAT2 colony. The regression analysis yielded a significant, negative slope (***p<0.001), indicating that the two variables are inversely correlated. R2 = 0.55.

In summary, the independent emergence of low copy DAT-tg mice in both Projects 1 and 2 indicates that BAC transgenic mouse models that incorporate multiple copies of a gene are sensitive to spontaneous loss of copy number over successive generations of breeding (Chandler et al., 2007). Furthermore, alterations in DAT copy number produce differences in dopaminergic phenotypes. This suggests that the level of DAT over-expression (or the number of DAT gene copies) is integral in determining dopaminergic outcomes.

208

Copyright Acknowledgements

ELSEVIER LICENSE TERMS AND CONDITIONS Feb 10, 2017

This Agreement between Shababa T Masoud ("You") and Elsevier ("Elsevier") consists of your license details and the terms and conditions provided by Elsevier and Copyright Clearance Center.

License Number 3938970614566

License date Licensed Content Publisher Elsevier Licensed Content Publication Neurobiology of Disease Licensed Content Title Increased expression of the dopamine transporter leads to loss of dopamine neurons, oxidative stress and l-DOPA reversible motor deficits Licensed Content Author S.T. Masoud,L.M. Vecchio,Y. Bergeron,M.M. Hossain,L.T. Nguyen,M.K. Bermejo,B. Kile,T.D. Sotnikova,W.B. Siesser,R.R. Gainetdinov,R.M. Wightman,M.G. Caron,J.R. Richardson,G.W. Miller,A.J. Ramsey,M. Cyr,A. Salahpour Licensed Content Date February 2015 Licensed Content Volume 74 Licensed Content Issue n/a Licensed Content Pages 10 Start Page 66 End Page 75 Type of Use reuse in a thesis/dissertation Portion full article Format both print and electronic Are you the author of this Elsevier Yes article? Will you be translating? No

Order reference number

209

Title of your thesis/dissertation Characterization of Mice with Altered Dopamine Transporter and Vesicular Monoamine Transporter 2 Levels Expected completion date Dec 2016 Estimated size (number of pages) 200 Elsevier VAT number GB 494 6272 12 Requestor Location Shababa T Masoud 1 King's College Circle - Rm 4302

Toronto, ON M5S 1A8 Canada Attn: Shababa T Masoud Total 0.00 CAD Terms and Conditions INTRODUCTION 1. The publisher for this copyrighted material is Elsevier. By clicking "accept" in connection with completing this licensing transaction, you agree that the following terms and conditions apply to this transaction (along with the Billing and Payment terms and conditions established by Copyright Clearance Center, Inc. ("CCC"), at the time that you opened your Rightslink account and that are available at any time at http://myaccount.copyright.com). GENERAL TERMS 2. Elsevier hereby grants you permission to reproduce the aforementioned material subject to the terms and conditions indicated. 3. Acknowledgement: If any part of the material to be used (for example, figures) has appeared in our publication with credit or acknowledgement to another source, permission must also be sought from that source. If such permission is not obtained then that material may not be included in your publication/copies. Suitable acknowledgement to the source must be made, either as a footnote or in a reference list at the end of your publication, as follows: "Reprinted from Publication title, Vol /edition number, Author(s), Title of article / title of chapter, Pages No., Copyright (Year), with permission from Elsevier [OR APPLICABLE SOCIETY COPYRIGHT OWNER]." Also Lancet special credit - "Reprinted from The Lancet, Vol. number, Author(s), Title of article, Pages No., Copyright (Year), with permission from Elsevier."

210

4. Reproduction of this material is confined to the purpose and/or media for which permission is hereby given. 5. Altering/Modifying Material: Not Permitted. However figures and illustrations may be altered/adapted minimally to serve your work. Any other abbreviations, additions, deletions and/or any other alterations shall be made only with prior written authorization of Elsevier Ltd. (Please contact Elsevier at [email protected]) 6. If the permission fee for the requested use of our material is waived in this instance, please be advised that your future requests for Elsevier materials may attract a fee. 7. Reservation of Rights: Publisher reserves all rights not specifically granted in the combination of (i) the license details provided by you and accepted in the course of this licensing transaction, (ii) these terms and conditions and (iii) CCC's Billing and Payment terms and conditions. 8. License Contingent Upon Payment: While you may exercise the rights licensed immediately upon issuance of the license at the end of the licensing process for the transaction, provided that you have disclosed complete and accurate details of your proposed use, no license is finally effective unless and until full payment is received from you (either by publisher or by CCC) as provided in CCC's Billing and Payment terms and conditions. If full payment is not received on a timely basis, then any license preliminarily granted shall be deemed automatically revoked and shall be void as if never granted. Further, in the event that you breach any of these terms and conditions or any of CCC's Billing and Payment terms and conditions, the license is automatically revoked and shall be void as if never granted. Use of materials as described in a revoked license, as well as any use of the materials beyond the scope of an unrevoked license, may constitute copyright infringement and publisher reserves the right to take any and all action to protect its copyright in the materials. 9. Warranties: Publisher makes no representations or warranties with respect to the licensed material. 10. Indemnity: You hereby indemnify and agree to hold harmless publisher and CCC, and their respective officers, directors, employees and agents, from and against any and all claims arising out of your use of the licensed material other than as specifically authorized pursuant to this license.

211

11. No Transfer of License: This license is personal to you and may not be sublicensed, assigned, or transferred by you to any other person without publisher's written permission. 12. No Amendment Except in Writing: This license may not be amended except in a writing signed by both parties (or, in the case of publisher, by CCC on publisher's behalf). 13. Objection to Contrary Terms: Publisher hereby objects to any terms contained in any purchase order, acknowledgment, check endorsement or other writing prepared by you, which terms are inconsistent with these terms and conditions or CCC's Billing and Payment terms and conditions. These terms and conditions, together with CCC's Billing and Payment terms and conditions (which are incorporated herein), comprise the entire agreement between you and publisher (and CCC) concerning this licensing transaction. In the event of any conflict between your obligations established by these terms and conditions and those established by CCC's Billing and Payment terms and conditions, these terms and conditions shall control. 14. Revocation: Elsevier or Copyright Clearance Center may deny the permissions described in this License at their sole discretion, for any reason or no reason, with a full refund payable to you. Notice of such denial will be made using the contact information provided by you. Failure to receive such notice will not alter or invalidate the denial. In no event will Elsevier or Copyright Clearance Center be responsible or liable for any costs, expenses or damage incurred by you as a result of a denial of your permission request, other than a refund of the amount(s) paid by you to Elsevier and/or Copyright Clearance Center for denied permissions. LIMITED LICENSE The following terms and conditions apply only to specific license types: 15. Translation: This permission is granted for non-exclusive world English rights only unless your license was granted for translation rights. If you licensed translation rights you may only translate this content into the languages you requested. A professional translator must perform all translations and reproduce the content word for word preserving the integrity of the article. 16. Posting licensed content on any Website: The following terms and conditions apply as follows: Licensing material from an Elsevier journal: All content posted to the web site must maintain the copyright information line on the bottom of each image; A hyper-text

212 must be included to the Homepage of the journal from which you are licensing at http://www.sciencedirect.com/science/journal/xxxxx or the Elsevier homepage for books at http://www.elsevier.com; Central Storage: This license does not include permission for a scanned version of the material to be stored in a central repository such as that provided by Heron/XanEdu. Licensing material from an Elsevier book: A hyper-text link must be included to the Elsevier homepage at http://www.elsevier.com . All content posted to the web site must maintain the copyright information line on the bottom of each image.

Posting licensed content on Electronic reserve: In addition to the above the following clauses are applicable: The web site must be password-protected and made available only to bona fide students registered on a relevant course. This permission is granted for 1 year only. You may obtain a new license for future website posting. 17. For journal authors: the following clauses are applicable in addition to the above: Preprints: A preprint is an author's own write-up of research results and analysis, it has not been peer-reviewed, nor has it had any other value added to it by a publisher (such as formatting, copyright, technical enhancement etc.). Authors can share their preprints anywhere at any time. Preprints should not be added to or enhanced in any way in order to appear more like, or to substitute for, the final versions of articles however authors can update their preprints on arXiv or RePEc with their Accepted Author Manuscript (see below). If accepted for publication, we encourage authors to link from the preprint to their formal publication via its DOI. Millions of researchers have access to the formal publications on ScienceDirect, and so links will help users to find, access, cite and use the best available version. Please note that Cell Press, The Lancet and some society-owned have different preprint policies. Information on these policies is available on the journal homepage. Accepted Author Manuscripts: An accepted author manuscript is the manuscript of an article that has been accepted for publication and which typically includes author- incorporated changes suggested during submission, peer review and editor-author communications.

213

Authors can share their accepted author manuscript:

  immediately o via their non-commercial person homepage or blog o by updating a preprint in arXiv or RePEc with the accepted manuscript o via their research institute or institutional repository for internal institutional uses or as part of an invitation-only research collaboration work-group o directly by providing copies to their students or to research collaborators for their personal use o for private scholarly sharing as part of an invitation-only work group on commercial sites with which Elsevier has an agreement   after the embargo period o via non-commercial hosting platforms such as their institutional repository o via commercial sites with which Elsevier has an agreement In all cases accepted manuscripts should:

  link to the formal publication via its DOI   bear a CC-BY-NC-ND license - this is easy to do   if aggregated with other manuscripts, for example in a repository or other site, be shared in alignment with our hosting policy not be added to or enhanced in any way to appear more like, or to substitute for, the published journal article. Published journal article (JPA): A published journal article (PJA) is the definitive final record of published research that appears or will appear in the journal and embodies all value-adding publishing activities including peer review co-ordination, copy-editing, formatting, (if relevant) pagination and online enrichment. Policies for sharing publishing journal articles differ for subscription and gold open access articles: Subscription Articles: If you are an author, please share a link to your article rather than the full-text. Millions of researchers have access to the formal publications on ScienceDirect, and so links will help your users to find, access, cite, and use the best available version. Theses and dissertations which contain embedded PJAs as part of the formal submission can be posted publicly by the awarding institution with DOI links back to the formal publications on ScienceDirect. If you are affiliated with a library that subscribes to ScienceDirect you have additional private sharing rights for others' research accessed under that agreement. This includes use

214 for classroom teaching and internal training at the institution (including use in course packs and courseware programs), and inclusion of the article for grant funding purposes. Gold Open Access Articles: May be shared according to the author-selected end-user license and should contain a CrossMark logo, the end user license, and a DOI link to the formal publication on ScienceDirect. Please refer to Elsevier's posting policy for further information. 18. For book authors the following clauses are applicable in addition to the above: Authors are permitted to place a brief summary of their work online only. You are not allowed to download and post the published electronic version of your chapter, nor may you scan the printed edition to create an electronic version. Posting to a repository: Authors are permitted to post a summary of their chapter only in their institution's repository. 19. Thesis/Dissertation: If your license is for use in a thesis/dissertation your thesis may be submitted to your institution in either print or electronic form. Should your thesis be published commercially, please reapply for permission. These requirements include permission for the Library and Archives of Canada to supply single copies, on demand, of the complete thesis and include permission for Proquest/UMI to supply single copies, on demand, of the complete thesis. Should your thesis be published commercially, please reapply for permission. Theses and dissertations which contain embedded PJAs as part of the formal submission can be posted publicly by the awarding institution with DOI links back to the formal publications on ScienceDirect.

Elsevier Open Access Terms and Conditions You can publish open access with Elsevier in hundreds of open access journals or in nearly 2000 established subscription journals that support open access publishing. Permitted third party re-use of these open access articles is defined by the author's choice of Creative Commons user license. See our open access license policy for more information. Terms & Conditions applicable to all Open Access articles published with Elsevier:

215

Any reuse of the article must not represent the author as endorsing the adaptation of the article nor should the article be modified in such a way as to damage the author's honour or reputation. If any changes have been made, such changes must be clearly indicated. The author(s) must be appropriately credited and we ask that you include the end user license and a DOI link to the formal publication on ScienceDirect. If any part of the material to be used (for example, figures) has appeared in our publication with credit or acknowledgement to another source it is the responsibility of the user to ensure their reuse complies with the terms and conditions determined by the rights holder. Additional Terms & Conditions applicable to each Creative Commons user license: CC BY: The CC-BY license allows users to copy, to create extracts, abstracts and new works from the Article, to alter and revise the Article and to make commercial use of the Article (including reuse and/or resale of the Article by commercial entities), provided the user gives appropriate credit (with a link to the formal publication through the relevant DOI), provides a link to the license, indicates if changes were made and the licensor is not represented as endorsing the use made of the work. The full details of the license are available at http://creativecommons.org/licenses/by/4.0. CC BY NC SA: The CC BY-NC-SA license allows users to copy, to create extracts, abstracts and new works from the Article, to alter and revise the Article, provided this is not done for commercial purposes, and that the user gives appropriate credit (with a link to the formal publication through the relevant DOI), provides a link to the license, indicates if changes were made and the licensor is not represented as endorsing the use made of the work. Further, any new works must be made available on the same conditions. The full details of the license are available at http://creativecommons.org/licenses/by-nc-sa/4.0. CC BY NC ND: The CC BY-NC-ND license allows users to copy and distribute the Article, provided this is not done for commercial purposes and further does not permit distribution of the Article if it is changed or edited in any way, and provided the user gives appropriate credit (with a link to the formal publication through the relevant DOI), provides a link to the license, and that the licensor is not represented as endorsing the use made of the work. The full details of the license are available at http://creativecommons.org/licenses/by-nc-nd/4.0. Any commercial reuse of Open

216

Access articles published with a CC BY NC SA or CC BY NC ND license requires permission from Elsevier and will be subject to a fee. Commercial reuse includes:

  Associating advertising with the full text of the Article   Charging fees for document delivery or access   Article aggregation   Systematic distribution via e-mail lists or share buttons Posting or linking by commercial companies for use by customers of those companies.

20. Other Conditions: v1.8

Questions? [email protected] or +1-855-239-3415 (toll free in the US) or +1-978-646-2777.

217

JOHN WILEY AND SONS LICENSE TERMS AND CONDITIONS Feb 10, 2017

This Agreement between Shababa T Masoud ("You") and John Wiley and Sons ("John Wiley and Sons") consists of your license details and the terms and conditions provided by John Wiley and Sons and Copyright Clearance Center.

License Number 4045251325759

License date Licensed Content Publisher John Wiley and Sons Licensed Content Publication European Journal of Neuroscience Licensed Content Title Membrane transporters as mediators of synaptic dopamine dynamics: implications for disease Licensed Content Author Kelly M. Lohr,Shababa T. Masoud,Ali Salahpour,Gary W. Miller Licensed Content Date Sep 2, 2016 Licensed Content Pages 14 Type of use Dissertation/Thesis Requestor type Author of this Wiley article Format Print and electronic Portion Full article Will you be translating? No Title of your thesis / dissertation Characterization of Mice with Altered Dopamine Transporter and Vesicular Monoamine Transporter 2 Levels Expected completion date Dec 2016 Expected size (number of pages) 200 Requestor Location Shababa T Masoud 1 King's College Circle - Rm 4302

Toronto, ON M5S 1A8 Canada Attn: Shababa T Masoud Publisher Tax ID EU826007151 Billing Type Invoice

218

Billing Address Shababa T Masoud 1 King's College Circle - Rm 4302

Toronto, ON M5S 1A8 Canada Attn: Shababa T Masoud Total 0.00 CAD Terms and Conditions TERMS AND CONDITIONS This copyrighted material is owned by or exclusively licensed to John Wiley & Sons, Inc. or one of its group companies (each a"Wiley Company") or handled on behalf of a society with which a Wiley Company has exclusive publishing rights in relation to a particular work (collectively "WILEY"). By clicking "accept" in connection with completing this licensing transaction, you agree that the following terms and conditions apply to this transaction (along with the billing and payment terms and conditions established by the Copyright Clearance Center Inc., ("CCC's Billing and Payment terms and conditions"), at the time that you opened your RightsLink account (these are available at any time at http://myaccount.copyright.com).

Terms and Conditions

 The materials you have requested permission to reproduce or reuse (the "Wiley Materials") are protected by copyright.

 You are hereby granted a personal, non-exclusive, non-sub licensable (on a stand- alone basis), non-transferable, worldwide, limited license to reproduce the Wiley Materials for the purpose specified in the licensing process. This license, and any CONTENT (PDF or image file) purchased as part of your order, is for a one- time use only and limited to any maximum distribution number specified in the license. The first instance of republication or reuse granted by this license must be completed within two years of the date of the grant of this license (although copies prepared before the end date may be distributed thereafter). The Wiley Materials shall not be used in any other manner or for any other purpose, beyond what is granted in the license. Permission is granted subject to an appropriate acknowledgement given to the author, title of the material/book/journal and the publisher. You shall also duplicate the copyright notice that appears in the Wiley publication in your use of the Wiley Material. Permission is also granted on the understanding that nowhere in the text is a previously published source

219

acknowledged for all or part of this Wiley Material. Any third party content is expressly excluded from this permission.

 With respect to the Wiley Materials, all rights are reserved. Except as expressly granted by the terms of the license, no part of the Wiley Materials may be copied, modified, adapted (except for minor reformatting required by the new Publication), translated, reproduced, transferred or distributed, in any form or by any means, and no derivative works may be made based on the Wiley Materials without the prior permission of the respective copyright owner.For STM Signatory Publishers clearing permission under the terms of the STM Permissions Guidelines only, the terms of the license are extended to include subsequent editions and for editions in other languages, provided such editions are for the work as a whole in situ and does not involve the separate exploitation of the permitted figures or extracts, You may not alter, remove or suppress in any manner any copyright, trademark or other notices displayed by the Wiley Materials. You may not license, rent, sell, loan, lease, pledge, offer as security, transfer or assign the Wiley Materials on a stand-alone basis, or any of the rights granted to you hereunder to any other person.

 The Wiley Materials and all of the intellectual property rights therein shall at all times remain the exclusive property of John Wiley & Sons Inc, the Wiley Companies, or their respective licensors, and your interest therein is only that of having possession of and the right to reproduce the Wiley Materials pursuant to Section 2 herein during the continuance of this Agreement. You agree that you own no right, title or interest in or to the Wiley Materials or any of the intellectual property rights therein. You shall have no rights hereunder other than the license as provided for above in Section 2. No right, license or interest to any trademark, trade name, service mark or other branding ("Marks") of WILEY or its licensors is granted hereunder, and you agree that you shall not assert any such right, license or interest with respect thereto

 NEITHER WILEY NOR ITS LICENSORS MAKES ANY WARRANTY OR REPRESENTATION OF ANY KIND TO YOU OR ANY THIRD PARTY, EXPRESS, IMPLIED OR STATUTORY, WITH RESPECT TO THE MATERIALS OR THE ACCURACY OF ANY INFORMATION CONTAINED IN THE MATERIALS, INCLUDING, WITHOUT LIMITATION, ANY IMPLIED WARRANTY OF MERCHANTABILITY, ACCURACY, SATISFACTORY QUALITY, FITNESS FOR A PARTICULAR PURPOSE, USABILITY, INTEGRATION OR NON-INFRINGEMENT AND ALL SUCH WARRANTIES ARE HEREBY EXCLUDED BY WILEY AND ITS LICENSORS AND WAIVED BY YOU.

 WILEY shall have the right to terminate this Agreement immediately upon breach of this Agreement by you.

 You shall indemnify, defend and hold harmless WILEY, its Licensors and their respective directors, officers, agents and employees, from and against any actual or

220

threatened claims, demands, causes of action or proceedings arising from any breach of this Agreement by you.

 IN NO EVENT SHALL WILEY OR ITS LICENSORS BE LIABLE TO YOU OR ANY OTHER PARTY OR ANY OTHER PERSON OR ENTITY FOR ANY SPECIAL, CONSEQUENTIAL, INCIDENTAL, INDIRECT, EXEMPLARY OR PUNITIVE DAMAGES, HOWEVER CAUSED, ARISING OUT OF OR IN CONNECTION WITH THE DOWNLOADING, PROVISIONING, VIEWING OR USE OF THE MATERIALS REGARDLESS OF THE FORM OF ACTION, WHETHER FOR BREACH OF CONTRACT, BREACH OF WARRANTY, TORT, NEGLIGENCE, INFRINGEMENT OR OTHERWISE (INCLUDING, WITHOUT LIMITATION, DAMAGES BASED ON LOSS OF PROFITS, DATA, FILES, USE, BUSINESS OPPORTUNITY OR CLAIMS OF THIRD PARTIES), AND WHETHER OR NOT THE PARTY HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES. THIS LIMITATION SHALL APPLY NOTWITHSTANDING ANY FAILURE OF ESSENTIAL PURPOSE OF ANY LIMITED REMEDY PROVIDED HEREIN.

 Should any provision of this Agreement be held by a court of competent jurisdiction to be illegal, invalid, or unenforceable, that provision shall be deemed amended to achieve as nearly as possible the same economic effect as the original provision, and the legality, validity and enforceability of the remaining provisions of this Agreement shall not be affected or impaired thereby.

 The failure of either party to enforce any term or condition of this Agreement shall not constitute a waiver of either party's right to enforce each and every term and condition of this Agreement. No breach under this agreement shall be deemed waived or excused by either party unless such waiver or consent is in writing signed by the party granting such waiver or consent. The waiver by or consent of a party to a breach of any provision of this Agreement shall not operate or be construed as a waiver of or consent to any other or subsequent breach by such other party.

 This Agreement may not be assigned (including by operation of law or otherwise) by you without WILEY's prior written consent.

 Any fee required for this permission shall be non-refundable after thirty (30) days from receipt by the CCC.

 These terms and conditions together with CCC's Billing and Payment terms and conditions (which are incorporated herein) form the entire agreement between you and WILEY concerning this licensing transaction and (in the absence of fraud) supersedes all prior agreements and representations of the parties, oral or written. This Agreement may not be amended except in writing signed by both parties. This Agreement shall be binding upon and inure to the benefit of the parties' successors, legal representatives, and authorized assigns.

221

 In the event of any conflict between your obligations established by these terms and conditions and those established by CCC's Billing and Payment terms and conditions, these terms and conditions shall prevail.

 WILEY expressly reserves all rights not specifically granted in the combination of (i) the license details provided by you and accepted in the course of this licensing transaction, (ii) these terms and conditions and (iii) CCC's Billing and Payment terms and conditions.

 This Agreement will be void if the Type of Use, Format, Circulation, or Requestor Type was misrepresented during the licensing process.

 This Agreement shall be governed by and construed in accordance with the laws of the State of New York, USA, without regards to such state's conflict of law rules. Any legal action, suit or proceeding arising out of or relating to these Terms and Conditions or the breach thereof shall be instituted in a court of competent jurisdiction in New York County in the State of New York in the United States of America and each party hereby consents and submits to the personal jurisdiction of such court, waives any objection to venue in such court and consents to service of process by registered or certified mail, return receipt requested, at the last known address of such party.

WILEY OPEN ACCESS TERMS AND CONDITIONS Wiley Publishes Open Access Articles in fully Open Access Journals and in Subscription journals offering Online Open. Although most of the fully Open Access journals publish open access articles under the terms of the Creative Commons Attribution (CC BY) License only, the subscription journals and a few of the Open Access Journals offer a choice of Creative Commons Licenses. The license type is clearly identified on the article. The Creative Commons Attribution License The Creative Commons Attribution License (CC-BY) allows users to copy, distribute and transmit an article, adapt the article and make commercial use of the article. The CC-BY license permits commercial and non- Creative Commons Attribution Non-Commercial License The Creative Commons Attribution Non-Commercial (CC-BY-NC)License permits use, distribution and reproduction in any medium, provided the original work is properly cited and is not used for commercial purposes.(see below)

Creative Commons Attribution-Non-Commercial-NoDerivs License

222

The Creative Commons Attribution Non-Commercial-NoDerivs License (CC-BY-NC- ND) permits use, distribution and reproduction in any medium, provided the original work is properly cited, is not used for commercial purposes and no modifications or adaptations are made. (see below) Use by commercial "for-profit" organizations Use of Wiley Open Access articles for commercial, promotional, or marketing purposes requires further explicit permission from Wiley and will be subject to a fee. Further details can be found on Wiley Online Library http://olabout.wiley.com/WileyCDA/Section/id-410895.html

Other Terms and Conditions:

v1.10 Last updated September 2015

Questions? [email protected] or +1-855-239-3415 (toll free in the US) or +1-978-646-2777.