THE EFFECTS OF E L E C T R O C O N V U L S I V E T H E R A P Y IN A N A N I M A L M O D E L OF PARKINSON'S DISEASE: MECHANISMS OF A POTENTIAL ADJUNCT T R E A T M E N T by ELISSA M A R I E STROME B.Sc , Trent University, 1999 M.Sc. The University of British Columbia, 2001 A THESIS SUBMITTED IN PARTIAL F U L F I L L M E N T OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPY i n THE F A C U L T Y OF G R A D U A T E STUDIES (Neuroscience) THE UNIVERSITY OF BRITISH C O L U M B I A June 2006 © Elissa Marie Strome, 2006 Abstract: Electroconvulsive therapy (ECT) is a widely used and effective treatment for mood disorders. ECT also appears to have positive effects in Parkinson's disease (PD), improving motor symptoms for several weeks. Some of the most consistent effects of electroconvulsive shock (ECS) in animals are enhancement of both monoamine neurotransmission and neurotrophic factor concentrations in limbic brain regions. We hypothesized that the mechanism of action of ECT in PD is similar to its proposed mechanism of action in depression, more specifically, that ECS in 6-hydroxydopamine (6-OHDA)-lesioned rats will : 1) improve motor behaviour; 2) enhance Dj and D3 (without changing D2) receptor binding; and 3) enhance striatal neurotrophic factor concentrations. We performed several pilot and validation studies to determine the appropriate animal models (non-human primate vs. rat model of PD) and to develop tools to investigate our hypotheses with (autoradiography with positron-emitting tracers and non-pharmacological motor evaluation). To examine our three specific hypotheses, we treated 6-OHDA-lesioned rats with ECS or sham treatment and examined their motor behaviour using two non-pharmacological behavioural tests; the Cylinder Test, which evaluates forelimb function, and the Tapered/Ledged Beam-Walking Test, which examines hindlimb function. After the course of ECS or sham treatment, the animals were sacrificed, and their brains were removed and processed for either dopamine receptor binding or neurotrophic factor concentration. ECS treatment significantly improved hindlimb function, but had no apparent impact on forelimb function. ECS treatment also enhanced striatal Di and D3 receptor binding, without affecting D2 binding. Finally, repeated ECS treatment decreased brain-derived neurotrophic factor (BDNF) concentrations in the prefrontal cortex (PFC), but increased BDNF in the hippocampus and striatum. Basic fibroblast growth factor (FGF-2) concentrations were increased in the striatum, whereas glial cell line-derived neurotrophic factor protein was significantly decreased in the PFC. This body of work provides the first thorough investigation of the effects of repeated ECS treatment in the 6-OHDA-lesioned rat, showing that it improves motor function, enhances D A neurotransmission via upregulation of the Di and D3 receptors, and increases striatal BDNF and FGF-2, and the results support the continued use and study of ECT as an adjunctive treatment for PD. i i Table of Contents: Abstract i i Table of Contents i i i List of Tables vii List of Figures viii List of Symbols and Abbreviations x Acknowledgments xi Dedication.. : xi i Co-Authorship Statement xiii Chapter 1: Introduction Literature Review 1 Parkinson's Disease 1 Electroconvulsive Therapy 1 Electroconvulsive Shock as an Animal Model of ECT 2 Overview and Objectives 7 Model and Tool Development 8 Hypotheses 9 Working Hypothesis 9 Hypothesis 1 12 Hypothesis 2 12 Hypothesis 3 13 References .. 15 Chapter 2: Electroconvulsive shock decreases binding to 5-HT2 receptors in nonhuman 18 primates: an in vivo positron emission tomography study with [ F]setoperone Preamble 25 Introduction 25 Materials and Methods 27 Subjects 27 ECS Procedure 27 PET Procedure 27 Data Analysis 28 Statistical Analysis 28 Results 29 Discussion 34 References 38 Chapter 3 : Quantitative in vitro phosphor imaging using [3H] and [18F] radioligands: the effects of chronic desipramine treatment on serotonin 5-HT2 receptors Preamble 43 Introduction 43 Materials and Methods 44 Subjects 44 5 - H T 2 Binding: [18F]Setoperone.. 44 5- H T 2 Binding: [3H]Ketanserin 45 [18F]Setoperone Standard Curves 45 [1 8F] Setoperone Binding Analysis 46 Resolution of [ 1 8F] Phosphor Imaging 46 [3H]Ketanserin Binding Analysis 48 Statistical Analysis 48 Results 50 Technical Issues of [1 8F] Phosphor Imaging 50 Fit and Reproducibility of [1 8F] Standard Curves 50 Resolution of [ l 8F] Phosphor Imaging 50 Technical Issues of [3H] Phosphor Imaging 53 Reusability of Tritium-Sensitive Phosphor Screens 53 Effects of DMI Treatment on 5-HT 2 Binding 54 Treatment and Regional Effects 54 Comparison of PET and Tritiated Ligands 56 Discussion • 56 Phosphor Imaging with PET Ligands 60 References 62 Chapter 4: Evaluation of the integrity of the dopamine system in a rodent model of Parkinson's disease: small animal PET compared to behavioral assessment and autoradiography Preamble 65 Introduction 65 Materials and Methods 67 Subjects 67 6- OHDA Lesioning 67 Tapered/Ledged Beam-Walking Test 67 MicroPET Imaging 68 Autoradiography 69 Statistical Analysis 69 Results 70 Discussion 75 References 78 Chapter 5: Electroconvulsive shock enhances Dj and D 3 receptor binding and improves motor behaviour in 6-OHDA-Iesioned rats Preamble 82 Introduction 83 Materials and Methods 84 Subjects 84 6-OHDA Lesioning 84 iv Cylinder Use Asymmetry Test 84 Tapered/Ledged Beam-Walking Test 85 Electroconvulsive Shock Treatment 85 Vesicular Monoamine Transporter-2 Binding 86 Di Receptor Binding ; 86 D2 Receptor Binding 86 D 3 Receptor Binding 86 Data Analysis... 87 Statistical Analysis 87 Results 87 Cylinder Test 87 Tapered/Ledged Beam-Walking Test 87 Vesicular Monoamine Transporter-2 Binding 87 Di Receptor Binding 89 D2 Receptor Binding 89 D 3 Receptor Binding 89 Discussion • 92 References 95 Chapter 6: The effects of electroconvulsive shock on GDNF, BDNF, and FGF-2 concentrations in the 6-OHDA-lesioned rat brain Preamble 101 Introduction 101 Materials and Methods 102 Subjects 102 6-OHDA Lesioning 102 Cylinder Use Asymmetry Test 103 Electroconvulsive Shock Treatment 103 Tissue Processing 103 GDNF and BDNF ELISA 104 FGF-2 ELISA 104 Data Analysis 104 Results •. 105 Cylinder Test 105 GDNF ELISA 105 BDNF ELISA 105 FGF-2 ELISA 105 Discussion 110 References 114 Chapter 7: General Discussion Preamble :120 Model and Tool Development 120 Model Development 120 Tool Development 122 The Effects of Repeated ECS on the D A System in Parkinsonian Rats 123 Status of Working Hypothesis 123 Hypothesis 1 123 Hypothesis 2 124 v Hypothesis 3 125 Synthesis of the Findings 126 Limitations of the Studies 126 Directions for Future Study 130 Significance of the Findings to PD 132 Conclusions 133 References 134 Appendix A: The use of anaesthesia for repeated ECS treatment in the rat Introduction 142 Materials and Methods 142 Results 143 Discussion 147 References 148 Appendix B: The effects of repeated ECS under ketamine anesthesia on BDNF mRNA expression Introduction .' 150 Materials and Methods 150 Results '. : 150 Discussion 150 References 152 List of Tables: Table 1.1: Effects of ECT in Parkinson's disease or parkinsonism 3 Table 1.2: Changes in D A receptors following repeated ECS 4 Table 1.3: Brief review of the literature on the effects of repeated ECS on the neurotrophic factors BDNF, FGF-2, and GDNF 6 Table 2.1: Polynomial contrasts to determine the pattern of the changes in 5 - H T 2 binding after a course of ECS 32 Table 3.1: Characteristics of the point source series used to create Figure 3.2 48 Table 3.2: Determination of the spatial resolution of the Multisensitive phosphor screens 52 Table 3.3: Decreased 5-HT 2 binding after chronic treatment with DMI detected using [18F]setoperone 55 Table 3.4: Correlation between [18F]setoperone and [3H]ketanserin binding 55 Table 4.1: Comparison of percent lesion after unilateral infusion of 20 ug (High Dose) or 2 ug (Low Dose) 6-OHDA, as measured by striatal [ U C]DTBZ binding in vivo by microPET and in vitro with autoradiography 71 vii List of Figures: Figure 1.1: Working hypothesis to explain the mechanism of action of ECT in PD 10 Figure 2.1: Relationship between seizure threshold and seizure length across the course of ECS 29 Figure 2.2: [18F]Setoperone binding to 5 - H T 2 receptors throughout the cortex in seven adult male rhesus monkey at baseline (pre-ECS), 24 hours, 1 week and 4-6 weeks after the last ECS session 31 Figure 2.3: Percent change in [18F]setoperone binding to 5 - H T 2 receptors 33 Figure 3.1: The same [18F]standard displayed three different ways, illustrating the spillover effect 47 Figure 3.2: Point sources printed on paper with [1 8F] ink 49 Figure 3.3: Fit of the curves derived from the [1 8F] standards 51 Figure 3.4: Spillover correction decreases the variability of the slopes of standard curves from different days 52 Figure 3.5: Background subtraction corrects for repeated use of tritium sensitive phosphor screens 53 Figure 3.6: Background subtraction compensates for accumulation of moisture by tritium-sensitive phosphor screens for up to three months after first opening 54 Figure 3.7: Decreased 5-HT 2 binding after chronic DMI treatment 57 Figure 3.8: Autoradiographic detection of decreased 5-HT2 binding using phosphor imaging and two different radioligands 58 Figure 4.1: The tapered/ledged beam-walking test apparatus 68 Figure 4.2: Relationships between the three measures of striatal D A integrity 72 Figure 4.3: TB test scores for each hindlimb in individual animals after severe (A-G) or mild (H-M) unilateral 6-OHDA lesioning 73 Figure 4.4: Coronal images of striatal [ n C ] D T B Z binding through the head and brain with microPET (left) and in brain only with autoradiography (right) 74 Figure 5.1: Forelimb use asymmetry scores before lesioning, and before and 48 h after repeated ECS or sham treatment 88 Figure 5.2: TB test data for the hindlimb contralateral to the lesion on the narrow section of the beam in ECS- and sham-treated rats before and 48 h after treatment 88 viii Figure 5.3: D A receptor binding values in the dorsal and ventral striatum in ECS- and sham-treated rats : 90-91 Figure 6.1: Forelimb use asymmetry scores before lesioning, and before and 48 h after repeated ECS or sham treatment 106 Figure 6.2: The effects of repeated ECS treatment on GDNF concentrations in the brain 107 Figure 6.3: The effects of repeated ECS treatment on BDNF concentrations in the brain 108 Figure 6.4: The effects of repeated ECS treatment on FGF-2 concentrations in the brain 109 Figure 7.1: Summary of the results of this body of work in the context of the working hypothesis 127 Figure 7.2: D A and BDNF synergism within the striatum after ECS treatment 129 Figure A . l : Mean charge dose (± SEM) required to elicit a seizure during the course of ECS under one of four conditions 144 Figure A.2: Mean seizure length (± SEM) during the course of ECS under one of four conditions 144 Figure A.3: Probability of eliciting a seizure with tonic hindlimb extension (THLE) under four different conditions as the course of ECS progressed 145 Figure A.4: .Repeated ECS treatment under ketamine increases frontal cortex [18F]setoperone binding 145 Figure A.5: Frontal cortex 5-HT2 receptor binding in rats treated with ECS or sham under ketamine anaesthesia as measured by [18F]setoperone 146 Figure B . l : The effects of repeated ECS treatment under ketamine anaesthesia on BDNF mRNA expression in the piriform cortex 151 ix List of Symbols and Abbreviations: 3-MHPG = 3-mefhoxy-4-hydroxyphenyl-glycol 5- HT = 5-hydroxytryptamine (serotonin) 6- OHDA = 6-hydroxydopamine A N O V A = analysis of variance A N C O V A = analysis of covariance BDI = Beck Depression Inventory BDNF = brain derived neurotrophic factor BP = binding potential cAMP = cyclic adenosine monophosphate CREB = cAMP response element-binding protein CSF = cerebrospinal fluid D A = dopamine DAT = dopamine transporter DBS = deep brain stimulation D L U = digital light units DMI = desipramine DTBZ = dihydrotetrabenazine ECS = electroconvulsive shock ECT = electroconvulsive therapy ELISA = enzyme-linked immunosorbent assay FGF-2 = basic fibroblast growth factor F W H M = full-width at half-maximum G A B A = y amino butyric acid GDNF = glial cell line-derived neurotrophic factor H A M - D = Hamilton Depression Rating Scale H V A = homovanillic acid ICV = intracerebroventricular K d = receptor affinity LID = L-DOPA-induced dyskinesia M A P = maximum a posteriori M A P kinase = mitogen-activated protein kinase MPTP = l-mefhyl-4-phenyl-l,2,3,6-tetrahydropyridine NAcc = nucleus accumbens N P Y = neuropeptide Y PD = Parkinson's disease PET = positron emission tomography . PFC = prefrontal cortex P K A = protein kinase A PSF = point-spread function ROD = relative optical density ROI = region of interest SN = substantia nigra SNpc = substantia nigra pars compacta SPECT = single photon emission computed tomography SSRI = selective serotonin reuptake inhibitor TB test = tapered/ledged beam-walking test T C A = tricyclic antidepressant T H = tyrosine hydroxylase T H L E = tonic hindlimb extension UPDRS = Unified Parkinson Disability Rating Scale V M A T 2 = vesicular monoamine transporter 2 Acknowledgements: I'd like to thank my supervisor, Dr. Doris Doudet, for her support during the course of my graduate studies, as a mentor, friend, and finally colleague. I am grateful for all of the many opportunities that Doris gave me to do good research, and know that she was the key to my success. Thanks to my mom and dad, for encouraging my independence from a very young age, for teaching me about the importance of family, and for supporting me in everything that I have ever done. I owe a great debt of gratitude to all of my colleagues at the Pacific Parkinson's Research Centre and the UBC/TRIUMF PET program. Over the years, I have benefited from the expertise and hard work of so many people; the principal investigators, Tom Ruth, Jon Stoessl, and Vesna Sossi; technicians and technologists, Jessica Grant, Rick Kornelsen, Salma Jivan, Carolyn English, Caroline Williams, Ken Buckley, Siobhan McCormick, and all the staff at the Animal Resource Unit; and my fellow graduate students, Joe Flores and Ivan Cepeda. Thanks also to the members of my supervisory committee, Drs. Tony Phillips, Cathy Rankin, and Wolf Tetzlaff, and ex officio member Dr. Athanasios Zis, for sharing their expertise, and for their guidance, support, and constructive critique of my research project. During the course of this body of research, I personally received funding from the Natural Sciences and Engineering Research Council of Canada (NSERC), and from the Michael Smith Foundation for Health Research (MSFHR), and we received funding for the project from the Canadian Institutes of Health Research (CIHR). Those scholarship and research dollars were greatly appreciated, and, I think, well spent. I'd also like to recognize my research subjects, without whom none of the questions within these pages could have been addressed. Thanks to Erin, for many years of friendship and stimulating scientific discussions. Thanks to Carey, for being my kindred spirit. And of course, I have to thank Dave, for being there. xi Dedication: To my grandfather, whose behaviour inspired my childhood curiosity, and motivated my career choice as an adult. xii Co-Authorship Statement: Much of the work included in Chapters 2-6 of this thesis was performed in collaboration with my research supervisor, mentors, fellow graduate students, and colleagues at the University of British Columbia. Every attempt has been made to give credit where it is due, either by sharing co-authorship in the published manuscripts, or by acknowledging assistance at the end of the manuscripts or in the Acknowledgments section of this thesis. In general, I performed 70-90% of the work required to produce a publishable manuscript independently, including most of the data collection, analysis, and interpretation, and all of the writing, save for comments or suggestions from my collaborators on the draft manuscripts. xiii Chapter 1 Introduction Literature Review: Parkinson's Disease Parkinson's disease (PD) is a progressive and debilitating neurodegenerative disorder affecting over 100,000 Canadians. The hallmark of the disease is the degeneration of the dopamine (DA)-containing neurons of the substantia nigra pars compacta (SNpc), but other neuropathological symptoms can also be detected at post-mortem, including the formation of Lewy bodies, and the depletion of the other monoamine neurotransmitters serotonin (5-HT) and norepinephrine (reviewed in Lang and Lozano, 1998a). The primary symptoms of PD include tremor, rigidity, bradykinesia (slow movement), hypokinesia (paucity of movement) and postural instability with gait disturbance. The early motor symptoms of PD are treated effectively with the D A precursor L-DOPA in combination with a peripheral metabolism inhibitor (carbidopa), and/or other adjunctive medications such as D A agonists. Within 5-7 years, however, the response to L-DOPA is diminished, and patients experience "wearing-off or "on-off" phenomena, and/or dyskinesia (abnormal involuntary movements; see Lang and Lozano, 1998b for review). Surgical interventions, either lesioning (e.g. in pallidotomy) or implanting a high frequency stimulator (known as deep brain stimulation, or DBS) in the major basal ganglia output structures are effective as subsequent steps in the treatment course (Walter and Vitek, 2004), but these treatments are not without risk, do not provide indefinite relief, and may introduce further side effects. With increasing disease severity also come symptoms outside the realm of movement including sleep and cognitive disturbances, as well as affective disorders, predominantly in the form of major depression. Prevalence studies suggest that 30-40% of PD patients suffer from comorbid major depression (Tandberg et al., 1996; Slaughter et al., 2001). Although traditional antidepressants such as tricyclic antidepressants or selective serotonin reuptake inhibitors (SSRIs) are often prescribed to treat depressive symptoms in PD, their use, safety, and efficacy have not been thoroughly investigated. In fact, a recent meta-analysis examining eleven studies concludes that the placebo effect is as strong as the true drug effect after antidepressant treatment for depression in PD (Weintraub et al., 2005). In addition, as in the normal depressed population, some patients may not respond well to antidepressant drugs, and there are concerns about the development of side effects after the use of SSRIs in PD because they can cause movement disorders in and of themselves (Caley, 1997; Gerber and Lynd, 1998). Electroconvulsive therapy (ECT) is a safe, effective, and widely used alternative for all patients who find no relief from their depression after trying several antidepressant drugs, or for those who develop negative side effects. Electroconvulsive Therapy ECT has a long and controversial history in psychiatry, and as such, it is often used only as a last resort to treat many psychiatric disorders. Several recent meta-analyses have shown, however, that ECT is more effective at treating depression than antidepressant drugs (UK ECT Review Group, 2003; Kho et al., 2003; Pagnin-et al., 2004). It is particularly effective in depression, certain types of schizophrenia, mania, and catatonia (Fink, 2001). ECT is the treatment of choice for depression in pregnant women, and in psychiatric emergencies. There are few contraindications (intracerebral mass or recent myocardial infarction), and it is associated with very low morbidity and mortality. Although the risks associated with ECT are 1 minimal, the treatment is associated with one consistent side effect, in the form of memory disturbance. While patients do typically experience both retrograde and/or anterograde memory disturbance, the loss is transient and involves only the weeks surrounding the treatment (reviewed in Rami-Gonzalez et al., 2001). Indeed, the controversy surrounding the use of ECT, especially in the United States, stems primarily from the memory disturbance that is associated with^he treatment. Modern ECT procedures, including the use of anaesthesia, oxygenation, and lower currents, however, have greatly reduced the severity of the cognitive side-effects. While there have been a few reports of profound and long-lasting memory disturbance in isolated individuals, the evidence for brain damage after ECT is weak. ECT is a controversial treatment, and some authors have suggested that along with the risks and benefits of treatment, informed consent protocols should include some discussion about this controversy (Reisner, 2003). Serendipitously, ECT treatment in depressed patients with PD has been found to not only alleviate the depressive symptoms, but to also have beneficial effects on the motor symptoms of PD. Many case studies, open trials, and a few sham-controlled double blind investigations, totalling over 200 patients studied to date, suggest that ECT significantly improves the motor symptoms of PD in over half of these patients, independent of the antidepressant effect (Kennedy et al., 2003; see Table 1.1 for a summary of the literature). The effects are most pronounced in patients with more severe PD or debilitating side effects (e.g. "on-off phenomenon; Balldin et al., 1981; Andersen et al., 1987). Interestingly, the antiparkinsonian effects are often observed within 3-5 treatments (see Table 1.1), whereas the antidepressant effects take 6-12 treatments to emerge. The amelioration of the motor symptoms of PD by ECT has been reported to last from weeks to months, and can be extended with maintenance treatments (Aarsland et al., 1997; Fall and Granerus, 1999). Animal studies have been extensively employed to try to understand the mechanism of action of ECT. In animal models, electroconvulsive shock (ECS) has a wide range of effects, from general changes in cerebral blood flow and metabolism to specific changes in neurotransmitter systems, neuroendocrine function, and gene expression (reviewed in Fochtmann, 1994). Most studies to date have been devoted to the study of the mechanism of action of ECT in depression, with intense focus on monoamine neurotransmitter systems in limbic brain regions. Although almost all aspects of neurotransmission seem to be affected by acute ECS treatment, more important are the cumulative effects of repeated treatments, since it takes several treatments to see either an antidepressant or an antiparkinsonian effect in patients. Electroconvulsive Shock as an Animal Model of ECT One of the earliest observations after repeated ECS is that it enhances monoamine-mediated behaviours, particularly, those induced by 5-HT- or DA-stimulating drugs (Green et a l , 1983a; Green et al., 1983b; Goodwin et al., 1984; Metz and Heal, 1986; and see Table 1.2). Further to those behavioural observations, using in vivo microdialysis, enhancements in the levels of 5-HT and D A in their target brain regions were observed after repeated ECS (Nomikos et al., 1991; Zis et al., 1991; Yoshida et al., 1998a; Yoshida et al., 1998b), as well as increases in the post-synaptic receptors for these receptors, notably the 5-HT 2 receptor in the cortex and hippocampus (Kellar et al., 1981; Green et al., 1983a; Stockmeier and Kellar, 1986; Biegon and Israeli, 1987; Pandey et al., 1992; Butler et al., 1993; Burnet et al., 1999) and D A receptors in the dorsal striatum and nucleus accumbens (NAcc; see Table 1.2 for review). For a long time, the changes in 5-HT and D A neurotransmission were used to explain the effectiveness of ECT as an antidepressant treatment, and only recently have the downstream effects of this monoamine activation been investigated. 2 Table 1.1: Effects of ECT in Parkinson's disease or parkinsonism; Reference (in chronological order) Type of study Results Lebensohn and Jenkins, 1975 Case study Amelioration of depression and parkinsonism in two patients with severe PD after 4 ECT Dysken et al., 1976 Case study Alleviation of bradykinesia and rigidity in one patient after 12 ECT (noticeable improvement by 6th ECT) Ananth et al., 1979 Case study Amelioration of drug-induced parkinsonism in one patient after 6 ECT (immediate effects after 1st ECT) Balldin etal., 1981 Open trial Marked improvement of "on-off' symptoms in 5/9 patients, lasting 2-41 weeks Andersen et al., 1987 Placebo-controlled, double-blind study Patients treated with ECT show prolonged "on" phases, decreased motor severity (short-lasting effects); no changes CSF metabolites Atre-Vaidya and Jampala, 1988 Case study Improvement of mania and parkinsonism in one patient after 12 ECT Roth et al, 1988 Case study Improvement of parkinsonism in one patient after 10 ECT (mania and tardive dyskinesia require an extra 7 ECT); anti-depressant and -PD effects last at least 4 months Goswami et al., 1989 Longitudinal triphasic Amelioration of neuroleptic-induced parkinsonism in 9 schizophrenics after 9 ECT Douyon et al., 1989 Open trial Significant improvement in motor function after mean 7 ECT in 7 patients with PD (no depression); significant improvement in 5 patients after only 2 ECT Lauterbach and Moore, 1990 Case study Significant improvement of both parkinsonism and dystonia in one patient after 9 ECT; UPDRS, dystonia, BDI and Ham-D scores fall Stern, 1991 Case study Improvement in PD symptoms after 2nd of 8 ECT; lasts 8 wks Zervas and Fink, 1992 Open trial Improvement in rigidity, tremor, bradykinesia, "on-off' in 4 patients with severe PD after 3-6 ECT; 1 UPDRS score 20-40%; lasts 4-6 wks; L-DOPA doses reduced Friedman and Gordon, 1992 Open trial 3 out of 4 PD patients show markedly improved motor function lasting as long as 8 wks Fall et al., 1995 Open trial 16 non-depressed PD patients - all show anti-PD effect lasting days-weeks (50% of patients) or 3-18 months (50% of patients); sig. T CSF HVA and NPY; patients with longest lasting effect show lower CSF 3-MHPG than patients with shorter lasting effect Pridmore and Pollard, 1996 Open trial 12 patients with PD - 3 no effect, 5 mild effect (2 wks -30 months), 4 marked effect (10 wks - 35 months) Nymeyer and Grossberg, 2002 Case study Marked improvement after 2nd ECT; L-DOPA must be reduced to eliminate delirium Aarsland etal., 1997 Case study Maintenance ECT prevents relapse in 2 patients with severe PD Moellentine et al., 1997 Retrospective study 14/25 PD patients show at least transient improvement in motor function Fall and Granerus, 1999 Case study 2 PD patients show even further improvement with maintenance ECT after a course Fall et al., 2000 SPECT imaging of DAT Motor symptoms improve; No change in [123I]-P-CIT uptake after a course of ECT in 6 patients; those with the best uptake improve the most Shulman, 2004 Case study Maintenance ECT for 4 years in late stage PD improves symptoms and mobility Legend: 3-MHPG = 3-methoxy-4-hydroxyphenyl-glycol; BDI = Beck Depression Inventory; CSF = cerebrospinal fluid; HAM-D = Hamilton Depression Rating Scale; HVA = homovanillic acid; NPY = neuropeptide Y; SPECT = single photon emission computed tomography; UPDRS = Unified Parkinson Disability Rating Scale. Table 1.2: Changes in DA receptors followine repeated ECS: Author ECS parameters Technique Effects of ECS BEHAVIOUR: Green etal., 1977 6-OHDA 150 V, Is 10/10 d halothane Amphetamine-induced locomotion/ circling T locomotion normal animals, T circling 6-OHDA-treated Wielosz, 1981 150 mA, 0.3 s 7/7 d Amphetamine-, apomorphine-induced behaviour T spontaneous and drug-induced activity (lasts 5-10 d); no change stereotypical behaviour Green et al., 1983a mice 90 V, 1 s 5/10 d halothane Apomorphine-induced activity Sig. T apomorphine-induced activity Smith and Sharp, 1997 150 V, 1s 5/10 d halothane Agonist-induced behaviour (activity, locomotion, grooming, sniffing, rearing) No change D, or D 2 agonists alone; sig. T activity, locomotion, sniffing w. concomitant admin. D i and D 2 agonist or apomorphine (lasts 3 wks) Andrade et al., 2002 30 or 120 mC 5/5 d Apomorphine-induced activity Activity increases only in rats treated with high dose ECS Zarrindast et al., 2004 150 mA 8/16 d Apomorphine-, amphetamine-, SKF 38393- and quinpirole-induced activity 7-10 d post-ECS T apomorphine locomotion @ 7 d, T SKF grooming @ 9 d; no change quinpirole yawning MICRODIALYSIS: Glue etal., 1990 100 mA, 10 s 8/8 d chloral hydrate (ECS 1 and 8) Striatal interstitial cone, of DA (during anaesthesia) Sig. T basal DA Zisetal., 1991 150 V, 0.75 s 8/15 d Striatal interstitial cone, of DA and metabolites (freely moving) Sig. T basal DOPAC and HVA; trend to T DA Brannan et al., 1993 6-OHDA in SN 100 V, 2 s chloral hydrate 8/16 d Striatal interstitial cone, of DA and metabolites (during anaesthesia) t basal DOPAC, i basal DA in lesioned striatum Yoshida et al., 1998a 125 V, 1.2 s 8/8 d Striatal interstitial cone, of DA and metabolites (freely moving) Sig. t basal DOPAC and HVA; trend to t DA -1^ Author ECS parameters Technique Effects of ECS RECEPTOR BINDING: Bergstrom and Kellar, 1979a 150 mA, 0.2 s corneal 7/7 d Binding to striatal homogenates ; [3HJ spiroperidol No change D 2 B ^ or Ka Reches et al., 1984) 150 V, 1.5 s 12/28 d Binding to striatal homogenates; [JH] spiperone No effect D 2 Klimek and Nielsen, 1987 100 mA, 1 s 6/11 d Binding to mesolimbic homogenates; [JH]SCH 23390 i D i B ^ (but n= 3 and no proper control group) Fochtmann et al., 1989 80 mA, 0.5 s 8/16 d Binding to sections; [ l2 iI]SCH 23390 29% t D, binding SN NowakandZak, 1989 150 mA, 0.5 s 10/10 d Binding to striatal homogenates; [JH]SCH 23390 T D[ binding BarkaietaL, 1990 150 V, 1s 8/8 d Binding to sections; [JH]SCH 23390 and spiroperidol T Bmax D 2 accumbens, amygdala (also T K Relative distance Figure 3.2: Point sources printed on paper with [18F] ink. A) Actual phosphor images of the point sources used to examine the spatial resolution of the Multisensitive phosphor screens. The point sources are 3.30, 2.41, 1.65, 0.95 and 0.64 mm in diameter, and the distance between each spot is four times the diameter center-to-center. B) A representative point-spread function, determined from the smallest diameter point sources in A). The X-axis has been normalized to mean ± SD = 0 ± 1. After fitting to a Gaussian distribution, and knowing the pixel size of the system, we estimate the spatial resolution of the system to be about 470 um FWHM. Results: I. Technical Issues of [18F1 Phosphor Imaging Fit and Reproducibility of [ F] Standard Curves: The Spillover Effect Displaying the image relative to the spot in question ("spillover correction") when determining the optical density of a particular point on the standard curve is critical to the production of accurate standard curves with excellent goodness of fit and high reproducibility. Because the standards are created on a two-dimensional surface, where the known value is the amount of activity in each spot, and because the shapes of the standard spots and distribution of activity within the spot can vary, the entire area must be taken into account to create the standard curve. This means that inaccurate delineation of the standard spot can greatly affect the fit of the standard curve, and consequently the quantification of the data. Figure 3.3A shows an example of what the curve looks like after spillover correction (squares) compared to i f the optical density is measured when the image is displayed in the default mode (triangles). The fit of the line is better after spillover correction (Linear regression: squares r 2 = 0.99, S y x = 0.99, triangles r 2 = 0.99, S y x = 1.24). By displaying the image relative to the spot in question, we remove spillover in the hottest spots, and are better able to delineate the coldest spots, thereby defining the standard spots more accurately. Equally importantly, Figs. 3.3B and 3.3C illustrate how spillover correction reduces the variability across standards made on a single day (Fig. 3.3B mean r 2, S y. x = 0.99, 5.62, Fig. 3.3C mean r 2, S y. x = 0.99, 2.78). Not only does spillover correction of the standard curves decrease the variability across standards made on a single day, but it also decreases variability across days. 3.4A shows the mean values of the standard curves produced on two different days without spillover correction, and Fig. 3.4B shows the same curves after spillover correction. Because the software default displays the image with pixel values at the extremes of the distribution removed, and because pixels of high activity "spillover" onto neighbouring pixels, the ROD in an ROI must also be determined when the image is displayed relative to the region in question (so that the minimum and maximum pixel values are also those of the region in question). Correction for spillover also helps in determining the appropriate size of the ROIs. 18 Resolution of [ F] Phosphor Imaging Figure 3.2A shows a representative image from the point source experiments. The resolution of the software and printer were the limiting factors in this experiment, and we could not accurately produce smaller diameter spots. Points in every diameter series were easily visually resolvable. Each point source produced a Gaussian activity distribution (Fig. 3.2B). Based on the mean distribution of the 15 point sources from the smallest series of points (diameter = 0.64 mm), we estimate the resolution to be about 470 um full-width at half-maximum (FWHM; table 3.2). 50 A) 15.0 Figure 3.3: Fit of the curves derived from the [ F] standards. A) After spillover correction of the data in Fig. 3.1, by displaying the image relative to each point in the curve before taking optical density readings, the data are a better fit to the standard curve (squares) than if optical density is measured without partial volume correction (triangles). B) Standard curves created on the same day, but quantified without spillover correction. C) The same set of curves as in B), but with spillover correction. By compensating for spillover, a set of standards created on the same day, but apposed to different screens shows both a better fit to the points in the curves, and less variability in slope. Figure 3.4: Spillover correction decreases the variability of the slopes of standard curves from different days. A) Mean standard curves from two different days measured without spillover correction. B) Mean standard curves from two different days measured with spillover correction. Table 3.2: Determination of the spatial resolution of the Multisensitive phosphor screens. Point-spread functions were determined for each of the spots in the smallest diameter series, and fit to a Gaussian distribution. From the F W H M of the PSF, the known diameter of the point source, and the known pixel size of 42.3 um, we estimate the resolution of the system to be about 470 um. Direction of Mean ± SD Diameter of the Resolution (um) measurement PSF FWHM (nm) point source (urn) Horizontal 789.14 ±23.89 640 461.67 Vertical 798.59 ±25.49 640 477.65 Both 793.86 ±24.33 640 469.70 52 II. Technical issues of [3Hlphosphor imaging Reusability of Tritium-Sensitive Phosphor Screens Although tritium-sensitive phosphor screens lose their sensitivity over time (due to moisture accumulation), the loss of sensitivity can be corrected for to a certain extent. The accumulated moisture affects the screen homogeneously, in the form of increased background noise. We examined the loss of sensitivity by exposing one tritium-sensitive phosphor screen to the same set of slides in two consecutive exposures. Figure 3.5A shows the data from several background regions after a 3 d exposure upon first removing the screen from the shrink-wrap packaging that it is shipped in, and after a subsequent 3 d exposure (scanned 6 d after opening). There is a significant increase in the optical density (DLU/mm 2) of the background regions (to.05(2),i2 = 4.75, p < 0.0005). However, after correcting the data from the ROIs placed on brain tissue slices collected on those two consecutive exposures by subtracting the appropriate background regions, there are no significant differences between the data collected on the two days (Fig. 3.5B; to.os(2),i2= 0.09, p > 0.93). To analyze the reusability of the tritium-sensitive phosphor screens over even longer time periods, we apposed the plastic [3H]microscales to the screens at several time-points after the screens were first opened. After decay correction, linear regression analysis of the mean DLU/mm 2 values from eight microscales, indicates that the response of the screens remains linear for at least three months after the first use, but. that the slopes of the lines are significantly different (Fig. 3.6A; linear regression r 2 ,= 1.0 for all; analysis of covariance (ANCOVA) F0.o5(2), 3,24 = 5.47, p < 0.005). However, when background is subtracted from the data, there is no significant difference between the slopes of the standard curves (Fig. 3.6B; A N C O V A F 0.o5(2), 3 ,24= 2.42, p> 0.09). A) B) 6500' 5500' 5000-1 4500-150000-1 3 a 50000-D A Y 3 D A Y 3 • • • • • IK."- 1 D A Y 6 Figure 3.5: Background subtraction corrects for repeated use of tritium sensitive phosphor screens. A) Data from several background regions obtained from the same tritium-sensitive phosphor screen 3 and 6 days after opening. There is a significant increase in background noise with subsequent exposures (p < 0.0005). D L U = digital light units. B) After subtracting the appropriate background regions from the ROIs placed on the tissue slices, there is no significant difference between the two exposures. 53 A) B) F i g u r e 3 . 6 : B a c k g r o u n d s u b t r a c t i o n compensates f o r a c c u m u l a t i o n o f m o i s t u r e b y t r i t i u m -sensi t ive p h o s p h o r screens f o r u p to t h r e e m o n t h s af ter f i r s t o p e n i n g . A) Data from a set of eight [3H]microscales apposed to the phosphor screens at several time points without background subtraction. Even after three months since the first exposure, the response remains linear. However, the slopes of the lines are significantly different. B ) The same data with background subtracted from each point on the standard curve. There is now no significant difference between the slopes of any of the lines. I I I . Ef fects o f D M I T r e a t m e n t o n 5-HT? B i n d i n g Treatment and Regional Effects Chronic treatment with DMI decreased 5-HT 2 binding in all cortical regions examined, as measured by both [3H]ketanserin and [18F]setoperone. The tissue for all three binding studies ([3H]ketanserin and [18F]setoperone low and high specific activity) came from the same ten animals (n = 5 per group), using consecutive slides for the examination of regional effects. The placement of ROIs was standardized for a particular area of cortex, and was consistent across all three binding studies. Table 3.3 shows the corrected binding data for several cortical regions collected on two different days, using low and high specific activity [18F]setoperone. The data for each region were compared with a paired t-test, and no significant differences were found. Therefore, in regions where binding was performed on both days, the mean binding value for each animal was used for further analyses (frontal and anterior and posterior parietal cortices) (table 3.4). Two-way analysis of variance (ANOVA) with repeated measures indicates [18F]setoperone binding was significantly decreased by chronic D M I treatment in all cortical regions examined (Fo.osp), \,i = 42.16, p < 0.0003), and that there were significant regional differences (F0.o5(2), 5,35 - 56.95, p < 0.0001). Post-hoc testing (Scheffe test) indicates that binding in the prefrontal and occipital cortices was different than all other regions. There was also a significant treatment/region interaction (Fn.osp), 5,35 = 3.85, p < 0.007). The right panel of table 3.4 shows the [3H]ketanserin data. Again, chronic DMI treatment significantly decreased 5 - H T 2 binding in all cortical regions examined (Fo.osp), 1,7 = 13.65, p < 0.008), and again, there was a significant regional effect (Fo.osp), 5,35 = 21.41, p < 0.0001), with post-hoc testing indicating significantly lower binding in the occipital cortex than all other regions, and higher binding in the prefrontal cortex than the posterior parietal and occipital cortices. There was also a significant treatment/region interaction (F0.05(2), 5,35= 3.04, p < 0.02). 54 Table 3.3: Decreased 5-HT2 binding after chronic treatment with DMI detected using [18F]setoperone. 5-HT 2 binding was significantly decreased throughout the cerebral cortex. After correcting for differences in specific activity, the data from two different days are not [18F]setoperone binding day 1 Incubation activity = 0.48 mCi SA = 265 Ci/mmol (pmol/cc) T O [ F]setoperone binding day 2 Incubation activity = 2.19 mCi SA = 1214 Ci/mmol (pmol/cc) Paired t-test P value DMI V E H % change DMI V E H % change Prefrontal (+3.7)a - - - 71.70 ± 10.18 115.2 ± 15.22 -37.76 -Frontal (+1.7) 56.3 ± 6.71 82.69 ± 19.01 -31.91 50.94 ± 5.49 79.20 ±8.18 -35.68 0.27 Anterior parietal (+0.7) 59.43 ± 5.80 87.41 ± 13.02 -32.01 58.82 + 6.63 93.95 ± 8.64 -37.39 0.51 Parietal (-1.3) 55.41 ±7.98 86.05 ± 15.62 -35.61 - - -Posterior parietal (-2.3) 49.74 ± 12.83 69.42 ± 14.57 -28.35 48.81 ±6.01 72.08 ± 13.90 -32.28 0.86 Occipital (-5.8) - - - 25.75 ±4.78 34.05 ±7.81 -24.38 -Data are mean ± SD. anumbers in parentheses are the coordinates in mm from Bregma (Paxinos and Watson, 1997). DMI = desipramine hydrochloride 15 mg/kg i.p for 17d, VEH = vehicle (sterile water). Table 3.4: Correlation between [18F]setoperone and [3H]ketanserin binding. After pooling the data from the [18F]setoperone studies, the data obtained with the PET and tritiated tracers are equivalent. j-n [ F]setoperone binding pooled (pmol/cc)b — [ H]ketanserin binding (pmol/cc)c DMI V E H % change DMI V E H % change Prefrontal (+3.7)a 71.70 ± 10.18 115.2 ± 15.22 -37.76 59.55 ± 16.27 98.49 ± 17.19 -39.16 Frontal (+1.7) 53.66 ± 5.41 80.95 ±11.90 -33.71 46.20 ± 8.45 85.67 ± 30.53 -46.07 Anterior parietal (+0.7) 59.13 ±5.36 90.68 ± 5.84 -34.79 60.23 ± 15.62 84.92 ± 20.21 -29.07 Parietal (-1.3) 55.41 ±7.98 86.05 ± 15.62 -20.5 58.55 ± 17.98 82.80 ± 26.25 -29.29 Posterior parietal (-2.3) 49.28 ± 6.93 70.75 ± 11.05 -30.35 46.89 ± 6.37 65.38 ± 22.54 -28.28 Occipital (-5.8) 25.75 ±4.78 34.05 ±7.81 -24.38 23.00 ±8.61 32.37 ± 5.89 -28.95 Data are mean ± SD. a numbers in parentheses are the coordinates in mm from Bregma (Paxinos and Watson, 1997). DMI = desipramine hydrochloride 15 mg/kg i.p for 17d, VEH = vehicle (sterile water).b Significant effect of treatment (p < 0.0009), (p < 0.0001), and significant treatment/region interaction (p < 0.007);c significant effect of treatment (p < 0.007), region (p < 0.0001)., and significant treatment/region interaction (p < 0.02). Figure 3.7 shows graphical examination of the treatment/region interaction effect detected in the binding data from both radioligands. The frontal and parietal cortices had similar receptor binding in the vehicle treated animals, whereas prefrontal and occipital cortex binding were much higher and lower respectively. Comparison of PET and Tritiated Ligands Paired t-tests indicate no significant differences in 5 - H T 2 binding in the frontal and parietal cortices as measured by [3H]ketanserin and [18F]setoperone. However, binding in the prefrontal and occipital cortex was slightly, but significantly lower with [3H]ketanserin compared to [18F]setoperone (p < 0.0042 and p < 0.04 respectively). Figure 3.8 shows representative images obtained with [3H]ketanserin and [1 8F] setoperone at the level of the frontal cortex. Discussion: The main purpose of this study was to demonstrate and validate the use of phosphor imaging systems for quantitative receptor binding autoradiographic studies with tritiated and positron-emitting ligands. In particular, we provide three new pieces of information: 1) by correcting for differences in specific activity, and taking precautions in the analysis of standard curves to limit variability, autoradiographic phosphor imaging data collected with positron-emitting ligands made on different days is highly reproducible; 2) using tracers for the same receptor subtype, the data obtained from receptor autoradiographic phosphor imaging with positron-emitting ligands corresponds well to that obtained with the more traditional tritiated ligands; and 3) although tritium-sensitive phosphor screens lose their sensitivity over time, with post-fixation to limit contamination and background subtraction, they can be reused at least seven times, and for at least three months after the first exposure to obtain quantitative data. This study has also confirmed the distribution of cortical 5 - H T 2 receptors. Autoradiographic studies with [3H]ketanserin indicate high 5-HT 2 binding in layers III, IV and V of the cortex in a rostro-caudal gradient in humans, monkeys, and rats (Pazos et al., 1985; Hoyer et a l , 1986; Lidow et al., 1989; Appel et al., 1990). The distribution of 5-HT 2 binding in our study agrees closely with the literature. However, although the numerical binding values are comparable to some reports (Lidow et al., 1989; Appel et al., 1990), they are tenfold lower than the data from others (Pazos et al., 1985; Hoyer et a l , 1986). The cause of this discrepancy is difficult to determine, as the only difference in incubation conditions in these studies is the drug used for non-specific binding. However, the close concurrence of our data with that of other studies in rat (Appel et al., 19,90) validates the use of tritium-sensitive phosphor screens as an appropriate alternative to film. Although we found essentially the same results with [3H]ketanserin and [1 8F] setoperone, the binding values for [3H]ketanserin were lower in both the DMI and vehicle treated animals in the prefrontal cortex (table 3.4). This may be a result of the post-fixation of the [3H]ketanserin bound slides in paraformaldehyde vapour, as slightly decreased binding and significantly decreased affinity (Kd) have been reported in the literature after this post-fixation protocol (Liberatore et al., 1999; Pavey et a l , 2002). Alternatively, the discrepancy may result from differences in binding kinetics of the two ligands. In any case, the relative changes in 5-HT 2 binding measured with [3H]ketanserin and [18F]setoperone are similar in the prefrontal cortex and in all other regions. The close concurrence of the data obtained with the two ligands validates the use of in vitro phosphor imaging with positron-emitting ligands to corroborate in vivo PET studies, instead of the more traditional, but also more time consuming, tritiated ligand autoradiography. 56 Figure 3.7: Decreased 5-HT2 binding after chronic DMI treatment. Data are mean'± SD. There is a significant effect of treatment and region, and a significant interaction between the two, a reflection of the relatively high density of 5-HT 2 receptors in the prefrontal cortex, and relatively low density in the occipital cortex (see Discussion). 57 A ) B ) Figure 3.8: Autoradiographic detection of decreased 5-HT2 binding using phosphor imaging and two different radioligands. A) [18F] setoperone and B) [3H]ketanserin binding in the frontal cortex in a DMI- (left) and vehicle-treated (right) animal. Sections were taken at approximately +1.7 mm from Bregma, according to Paxinos and Watson (1997). Previous reports of decreased 5 - H T 2 binding after chronic DMI treatment of rats are limited to either whole cortex or frontal cortex homogenates (Bergstrom and Kellar, 1979; Goodnough and Baker, 1994). However, because autoradiography has higher spatial resolution compared to binding studies using tissue homogenates, we were able to observe, for the first time, that chronic DMI treatment decreases binding to 5 - H T 2 receptors throughout the cerebral cortex, and that this effect also follows a rostro-caudal gradient. In general, 5 - H T 2 binding was decreased by chronic DMI treatment in rats to a greater extent in anterior cortical regions, confirming in vivo observations by PET in humans (Yatham et al., 1998). The region/treatment interaction (Fig. 3.7), resulted from major differences in the normal distribution of the receptors throughout the rostro-caudal extent of the cortex, so that the percent decrease in binding was greater in anterior regions of the cortex because the density of receptors is higher, and therefore the effect of chronic DMI treatment is more substantial. Although we did not perform Scatchard analysis to determine whether changes in 5-HT 2 binding were due to altered affinity or density of the receptors, most reports indicate that chronic DMI treatment decreases the number of receptors, and not their affinity (Bergstrom and Kellar, 1979; Dumbrille-Ross et al., 1982; Cross and Horton, 1988). However, all previous investigations into the effects of chronic DMI on 5 -HT 2 receptors have been performed on membrane homogenates. There are discrepancies in the literature between ligand binding data obtained in homogenates compared to slices, and in general, the affinity and/or density of the receptors is greater in tissue sections than in homogenate preparations (Dohanich et al., 1986; Revila et al., 2000; Pavey et al., 2002). This effect may be related to slice thickness and the subsequent slower equilibration rate with thicker slices, as apparent Kd decreases with decreasing slice thickness (Dohanich et al., 1986), and must be determined for each ligand. In any case, quantitative receptor binding data obtained in homogenates and slices is probably not fully comparable. It is possible to perform Scatchard analysis using the phosphor imaging autoradiographic method with either tritiated or positron-emitting compounds (e.g. Gatley et a l , 1998; Pavey et al., 2002, Nikolaus et al., 2003), and the higher spatial resolution of this technique compared to homogenate binding allows more precise localization of changes. Tritium-sensitive phosphor screens One of the limitations of phosphor imaging with tritiated ligands has been screen contamination, which leads to poor reusability and increased costs. To allow detection of the low energy radiation of tritium (maximum 18.6 keV, mean 5.7 keV) these screens do not have a protective coating. To reduce contamination, we have adopted the technique of Liberatore and colleagues (1999) and post-fix the slides in paraformaldehyde vapour in a vacuum for two days before apposition to the phosphor screens. We have also found that removal of any dirt or particulate matter on the glass slides (e.g. grease pencil used to contain the radioligand during incubation) before exposure prolongs the usable life of the tritium-sensitive phosphor screens. Since adopting these precautions, we have not observed any screen contamination in the form of latent (or "ghost") images identified in subsequent scans. Another issue to consider is the loss of sensitivity of the tritium-sensitive screens over time. Without a protective coating, the bare phosphor crystals accumulate moisture, leading to increased background and reduced sensitivity. We tested the response of the screens over time, and found that the response of the screens remains linear, and that the loss of sensitivity due to accumulated moisture can easily be corrected by background subtraction up to three months after first opening. However, we still recommend that tritium-sensitive phosphor screens opened at or around the same time be used as a set for subsequent exposures. We have found that with the post-fixation of the tissue, careful handling of the screens, and correction for loss of sensitivity over time, tritium-sensitive phosphor screens can be reused 59 to obtain quantitative data at least seven times, and that they retain a high level of sensitivity for at least three months. Our experience also suggests that i f it remains uncontaminated, the useable life of a tritium-sensitive phosphor screen could be even longer. Phosphor Imaging with PET Ligands Quantitative phosphor imaging with tritiated ligands has been well described (Liberatore et al., 1999; Pavey et al., 2002). However, because of the short half-lives of positron-emitting ligands, quantification of phosphor autoradiography with PET tracers is difficult because standards have to be made daily. There are reports of the use of brain paste standards incubated with positron-emitting ligands (Nikolaus et al., 2001; 2003), but these are not always possible to create, especially when using ligands labelled with a short half-life nuclide such as [ n C ] (T1/2 = 20.4 min). We have adopted the technique of Gatley et al. (1998), the use of T L C plates with a set of serial dilutions of 5 uL drops of known activity. These standards are not only easy to produce, but also consistently and reproducibly provide linear standard curves with excellent fit. The correction for spillover that we describe above, whereby the image is displayed relative to the area in question, improves two characteristics of the standard curves. The goodness of fit of the curve is enhanced, and the slopes of the standard curves created on a single day and across days are more comparable, thereby decreasing the variability of receptor binding data from different screens and different days (Figs. 3.3 & 3.4). The spillover effect may also impact the receptor binding data directly. This could occur if, in the image, regions of particularly high binding are surrounded by regions of relatively low binding. As in the measurements of the points in the standard curve, we suggest that when placing ROIs, the investigator display the image relative to the area in question to minimize the effects of spillover. The spatial resolution of phosphor imaging autoradiography is a function of three main factors: the optics of the imaging system, the characteristics of the phosphor screen, and the energy of the isotope used. Each phosphor imaging system produces an image with a specific pixel size, based on the properties of the scanning laser and associated electronics. According to the manufacturer, the Cyclone phosphor imager uses a laser that is focused to less than 50 um, and scanning the screens at the highest resolution setting (600 dpi, as used in this report), displays an image with 42 um pixels, only slightly larger than that of more expensive phosphor imaging systems (the Fuji BAS-5000 imaging system has a minimum image pixel size of 25 urn). In contrast, phosphor screens of all types have an intrinsic resolution of approximately 50 um, based on the size of the photosensitive phosphor crystals (Yanai et al., 1992). This is fairly high resolution, but is still many times lower than film or emulsions, which can generally resolve on the order of a fraction of a micron. Finally, the energy of the isotope used to label the ligand contributes substantially to the resolution of the final images. Low energy particles (such as the (3 emission of tritium) interact only with the silver halide or phosphor crystals directly overlying them in the film or screen, giving superb resolution. In contrast, isotopes with higher energy, particularly the positron and gamma emission of positron-emitting ligands, are expected to have 1 R much lower resolution. Both our estimated resolution for [ F]-phosphor imaging (470 um) and our observation that the resolution was slightly better in the horizontal plane than vertically are similar to observations of the resolution of [nC]-phosphor imaging by Sihver and colleagues (1997), who found F W H M values of 566 n m horizontally and 650 um vertically. Sihver and colleagues (1997) attributed the horizontal/vertical resolution discrepancy to the mechanics of scanning the screen in a vertical direction, and our data support this hypothesis. We did not measure the resolution of the tritium-sensitive phosphor screens (for fear of long-lasting radioactive contamination of the Inkjet printer). In general, the spatial resolution achieved with phosphor imaging is well beyond that required for quantitative receptor autoradiography. 60 Due to the high cost of synthesizing the tracers, we perform in vitro studies with PET radioligands with the leftover doses from the patient scans. If there is not enough activity to perform the binding assay on all of the slides from one study, or there are too many slides to physically process in one day, the data may have to be collected over several days. Differences in specific activity lead to differences in the amount of radioactivity added to the incubation dishes in order to maintain a constant incubation concentration of the ligand, which must be corrected for by dividing the volume corrected data (nCi/cc of tissue) by the specific activity (Ci/mmol) to obtain a unit of pmol/cc. In this experiment, we have shown that even when the specific activities are widely different (265 Ci/mmol vs. 1214 Ci/mmol), correcting for the specific activity produces data that are statistically equivalent (table 3.3). Chronic D M I treatment, has dramatic effects on 5-HT2 binding, ranging from 24-38% decreases (tables 3.3 and 3.4). Changes that are less dramatic, but yet still biologically meaningful may be more difficult to detect by quantitative autoradiography with PET tracers. Power analysis indicates that due to the high variance of the data of some regions, the ability to detect a significant difference in the means of two groups as little as 10% with 80% power at the 0.05 level of significance, would require 30 or more animals per group. This number of subjects is not always feasible, and lack of sensitivity due to high interindividual variability is the major limitation of this technique. However, since we performed three separate binding studies, the amount of tissue we had available for each was limited, and binding measures were taken on only two sections per animal in each study. In the future, we recommend using four or more sections per animal to decrease the variability. Multisensitive phosphor screens are quite versatile and can be used to perform autoradiographic studies with many different nuclides. We have found no loss of sensitivity in the screens over three years of use, and although very hot sources (in the mCi range) can leave behind latent ("ghost") images after scanning, these ghost images fade over time, and are generally lower than background noise in an [1 8F] experiment (Strome et al., unpublished observations). Besides [18F]-labelled radioligands, we have also performed quantitative receptor binding studies with [ uC]-labelled PET radioligands such as the Di and D2 antagonists SCH 23390 and raclopride, and the vesicular monoamine transporter-2 ligand dihydrotetrabenazine (unpublished). Phosphor imaging is certainly not limited to autoradiography with PET or tritiated ligands. An especially common application of involves the use of [125I] or [35S] for quantitative receptor or in situ hybridization autoradiographic studies, (Key et al, 1991; Ito et al., 1995; Tang et al., 1995; Viz i et al., 2001). The energy of [3 5S] is similar enough to [ 1 4C] that commercially available [ l 4 C] standards can be employed for quantification, while [125I]microscales themselves are also commercially available. In fact, phosphor imaging systems can be used for almost any assay that involves detection of a 2-D radioactive source, making them a valuable addition to the array of standard laboratory equipment. In conclusion, we have described several ways to exploit the inherently high resolution and sensitivity of storage phosphor screens to obtain quantitative receptor binding data that show good reliability and reproducibility with either tritiated or positron-emitting ligands. The temporal, economical, and practical advantages of phosphor imaging make it an excellent technique to replace traditional x-ray film tritiated ligand autoradiography. In particular, quantitative in vitro phosphor imaging with positron-emitting ligands provides an important tool to corroborate in vivo PET data with higher resolution in vitro autoradiographic data. 61 References: Adam MJ , Lu S, Jivan S, Huser J (1997) Simplified purification and synthesis of F-18 setoperone. J Labelled Compd Radiopharm 40: 258. Appel N M , Mitchell W M , Garlick RK, Glennon R A , Teitler M , De Souza EB (1990) Autoradiographic characterization of (+-)-l-(2,5-dimethoxy-4-[125I]iodophenyl)-2-aminopropane ([1 2 5I]DOI) binding to 5-HT 2 and 5 - H T i c receptors in rat brain. J Pharmacol Exp Ther 255: 843-857. Bergstrom DA, Kellar K J (1979) Adrenergic and serotonergic receptor binding in rat brain after chronic desmethylimipramine treatment. J Pharmacol Exp Ther 209: 256-261. Bergstrom M , Awad R, Estrada S, Malman J, Lu L, Lendvai G, et al. (2004) Autoradiography with positron emitting isotopes in positron emission tomography tracer discovery. Mol Imag Biol 5: 390-396. Besret L, Dauphin F, Guillouet S, Dhilly M , Gourand F, Blaizot X , et al. (1998) [nC]S21007, a putative partial agonist for 5 -HT3 receptors PET studies. Rat and primate in vivo biological evaluation. Life Sci 62:115-129. Cross JA, Horton R (1988) Effects of chronic oral administration of the antidepressants desmethylimipramine and zimelidine on rat cortical G A B A B binding sites: a comparison with 5-H T 2 binding site changes. J Pharmacol93: 331-336. Dohanich GP, Halpain S, Lambdin LT, McEwen BS (1986) Features of ligand binding in homogenate and section preparations. Brain Res 366: 338-342. Doze P, van Waarde A , Tewson TJ, Vaalburg W, Elsinga P H (2002) Synthesis and evaluation of (^-[^FJ-fluoroethylcarazolol for in vivo (3-adrenoceptor imaging in the brain. Neurochem Int 41: 17-27. Dumbrille-Ross A , Tang WW, Coscina D V (1982) Lack of effect of raphe lesions on serotonin S 2 receptor changes induced by amitriptyline and desmethylimipramine. Psychiatry Res 7: 145-151. Gatley SJ, Ding Y-S, Brady LS, Gifford A N , Dewey SL, Carroll F l , et al. (1998) In vitro and ex vivo autoradiographic studies of nicotinic acetylcholine receptors using [18F]fluoronorchloroepibatidine in rodent and human brain. Nucl Med Biol 25: 449-454. Goodnough DB, Baker GB (1994) 5-hydroxytryptamine2 and P-adrenergic receptor regulation in rat brain following chronic treatment with desipramine and fluoxetine alone and in combination. J Neurochem ,62: 2262-2268. Hoyer D, Pazos A , Probst A , Palacios J M (1986) Serotonin receptors in the human brain. II. Characterization and autoradiographic localization of 5-HTi C and 5-HT 2 recognition sites. Brain Res 376:97-107. 62 Ito R, Suzuki T, Lim D K , Wellman SE, Ho IK (1995) A novel quantitative receptor autoradiography and in situ hybridization histochemistry technique using storage phosphor screen imaging. J Neurosci Methods 59:265-271. Johnston RF, Pickett SC, Barker D L (1990) Autoradiography using storage phosphor technology. Electrophoresis 11: 355-360. Key M , Wirick B, Cool D, Morris M (2001) Quantitative in situ hybridization for peptide mRNAs in mouse brain. Brain Res Protoc 8: 8-15. Liberatore GT, Wong JYF, Krenus D, Jeffreys BJ, Porritt M , Howells DW (1999) Tissue fixation prevents contamination of tritium-sensitive storage phosphor imaging plates. BioTechniques 6: 432-434. Lidow MS, Goldman-Rakic PS, Gallager DW, Rakic P (1989) Quantitative autoradiographic mapping of serotonin 5-HTi and 5 - H T 2 receptors and uptake sites in the neocortex of the rhesus monkey. J Comp Neurol 280: 7-42. Lopez-Gimenez JF, Mengod G, Palacios JN, Vilaro M T (1997) Selective visualization of rat brain 5 - H T 2 A receptors by autoradiography with [ 3 H]MDL 100,907. Naunyn-Schmiedeberg's Arch Pharmacol 356: 446-454. Matsumura K, Bergstrom M , Onoe H, Takechi H, Westerberg G, Anton G, et al. (1995) In vitro positron emission tomography (PET): use of positron emission tracer in functional imaging in living brain slices. Neurosci Res 22: 219-229. Nikolaus S, Larisch R, Beu M , Vosberg H , Muller-Gartner H W (2001) Imaging of striatal dopamine D 2 receptors with a PET system for small laboratory animals in comparison with storage phosphor autoradiography: a validation study with 18F-(N-Methyl)Benperidol. J Nucl Med 42: 1691-1696. Nikolaus S, Larisch R, Beu M , Hamacher K , Forutan F, Vosberg H , Muller HW (2003) In vivo measurement of D 2 receptor density and affinity for 18F-(3-N-methyl) benperidol in the rat striatum with a PET system for small laboratory animals. J Nucl Med 44: 618-624. Passchier J, van Waarde A , Doze P, Elsinga PH, Vaalburg W (2000) Influence of P-glycoprotein on brain uptake of [ 1 8F]MPPF in rats. Eur J Pharmacol 407: 273-280. Pavey G M , Copolov DL, Dean B (2002) High-resolution phosphor imaging: validation for use with human brain tissue sections to determine the affinity and density of radioligand binding. J Neurosci Methods 116: 157-163. Paxinos G, Watson C (1997) The Rat Brain in Stereotaxic Coordinates, compact third ed. Academic Press: San Diego. Pazos A , Cortes R, Palacios J M (1985) Quantitative autoradiographic mapping of serotonin receptors in the rat brain. II. Serotonin-2 receptors. Brain Res 346: 231-249. 63 Revilla R, Revilla V , Fernandez-Lopez A (2000) Autoradiographic characterisation of beta-adrenoceptors in chick brain using [ 3H]CGP 12177. Brain Res Pro toe 5: 140-145. Sasaki T, Ishiwata K , Murata T, Senda M (2002) Demonstration of competition between endogenous dopamine and [nC]raclopride binding in in vitro brain slices using a dynamic autoradiography technique. Synapse 44: 42-50. Sihver W, Sihver S, Bergstrom M , Murata T, Matsumura K , Onoe H, et al. (1997) Methodological aspects for in vitro characterization of receptor binding using nC-labeled receptor ligands: a detailed study with the benzodiazepine receptor antagonist [ n C]Ro 15-1788. Nucl Med Biol 24: 723-731. Tang C, Biemond I, Lamers CB (1995) Localization and quantification of cholecystokinin receptors in rat brain with storage phosphor autoradiography. Biotechniques 8: 886-889. Upham L V , Englert DF (1998) Radionuclide imaging. In L'Annunziata M F , editor. Handbook of radioactivity analysis. Academic Press: San Diego, 647-692. Viz i S, Palfi A , Hatvani L , Gulya K (2001) Methods for quantification of in situ hybridization signals obtained by film autoradiography and phosphorimaging applied for estimation of regional levels of calmodulin mRNA classes in the rat brain. Brain Res Protoc 8: 32-44. Yanai K , Ryu JH, Watanabe T, Iwata R, Ido T (1992) Receptor autoradiography with n C and [3H]-labelled ligands visualized by imaging plates. NeuroReport 3: 961-964. Yatham L N , Liddle PF, Dennie J, Shiah IS, Adam MJ , Lane CJ, et al. (1999) Decrease in brain serotonin 2 receptor binding in patients with major depression following desipramine treatment: a positron emission tomography study with fluorine-18-labeled setoperone. Arch. Gen Psychiatry 56: 705-711. 64 Chapter 4 EVALUATION OF THE INTEGRITY OF THE DOPAMINE SYSTEM IN A RODENT MODEL OF PARKINSON'S DISEASE: SMALL ANIMAL PET COMPARED TO BEHAVIOURAL ASSESMENT AND AUTORADIOGRAPHY1 Preamble: In the previous chapter we gave a detailed description of a method for using PET tracers for quantitative in vitro autoradiography. The technique was developed to provide us with a tool to use to address our specific hypotheses regarding the effects of repeated ECS treatment on D A receptors in 6-OHDA-lesioned rats. To address our hypotheses about the effects of repeated ECS treatment on the motor function of 6-OHDA-lesioned rats we decided to move away from the traditional drug-induced behavioural testing that is so commonly used in this animal model of PD (rotational behaviour after apomorphine or amphetamine treatment) because of the potential confounds that these DAergic drugs add to our primary neurochemical outcome measures. For example, even a single treatment with apomorphine sensitizes Di receptors (Morelli et al., 1999), and apomorphine and amphetamine can induce GDNF, BDNF, and FGF-2 (Meredith et al., 2002; Guo et al., 2002; Mueller et al., 2006). Instead, we adopted two non-pharmacological behavioural tests; the Cylinder Test, which measures forelimb function, has been shown to correlate with striatal D A levels (Tillerson et al., 2001), and is widely used to evaluate therapeutic interventions in the unilateral 6-OHDA-lesion rat model of PD (Cohen et al., 2003; Shi et al., 2004), and the tapered/ledged beam-walking (TB) test (Schallert and Woodlee, 2005), which has been less thoroughly investigated. In this chapter, we examined the relationship between performance on the TB test and striatal D A terminal integrity in animals with mild or severe unilateral 6-OHDA lesions, as measured by our in vitro autoradiographic technique described in Chapter 3, and by in vivo microPET imaging, using the small animal PET scanner that we had gained access to in the meantime. These PET studies also allowed us to investigate the possibility of using microPET to investigate our primary hypotheses regarding the mechanism of action of ECT in PD. In the end, we decided against using microPET for our preliminary studies for several reasons, including the poor resolution of the technique, and the difficulty in gaining access to the scanner and required tracers at specific time-points before, during, and after the course of ECS treatment. Introduction: In recent years, our understanding of the role that genes and/or the environment play in disease, and the development of therapeutic options have greatly improved as a result of better animal models, achieved through advances in genetic engineering or the use of specific toxins. Traditionally, behavioural analysis, with or without pharmacological challenge, was the method of choice for non-invasive longitudinal assessment of animal models of disease. There are several problems with relying solely on behavioural observation to evaluate animal models, however, including the fact that it does not provide specific information about underlying biochemical processes, and individual variation may be very large. A classical example of the evaluation of an animal model heavily dependent on behavioural analysis is the unilateral 6-1 A version of this chapter has been accepted for publication. Strome EM, Cepeda I, Sossi V, Doudet DJ (2006) Evaluation of the integrity of the dopamine system in a rodent model of Parkinson's disease: small animal PET compared to behavioral assessment and autoradiography. Molecular Imaging and Biology. 65 hydroxydopamine (6-OHDA, a selective neurotoxin of the nigrostriatal pathway) lesion rat model of Parkinson's disease (PD; Ungerstedt, 1968). In the past, longitudinal assessment of this model depended primarily on repeated pharmacological challenges (Costall et al., 1976), which in themselves can modify the system. The best of example of this is the robust observation of sensitization of both the behavioural response and the Di receptor after repeated apomorphine (Morelli et al., 1989; Klug and Norman, 1993). The 6-OHDA rat model is therefore a good example of an animal model of a human disease that can greatly benefit from advances in our ability to reliably, reproducibly, and sensitively evaluate longitudinally, in vivo, the disease process or the effects of therapeutics. Positron emission tomography (PET) is a powerful tool for in vivo imaging of biochemical processes. The low spatial resolution of the technique (maximum ~ 4 mm), however, necessitates the use of large animals as subjects (humans and non-human primates). Unfortunately, most biomedical research is done using small animals, particularly rodents, birds, fish or amphibians. In response, small animal PET (microPET) is an emerging field, and scanners are now commercially available with smaller fields of view, sensitive scintillation detectors, and higher spatial resolution (Tai et al., 2001; Knoess et al., 2003; Tai et al., 2003; Yang et al., 2004; Tai et al., 2005). These microPET scanners show great promise for use in life sciences research, particularly in the fields of drug development, oncology and cardiology. Although still fairly expensive and requiring the typical infrastructure of a full-size PET program, microPET offers many advantages over traditional post-mortem analysis in animal research, including limiting the number of animals used (by allowing longitudinal studies in the same animal, particularly important in studies of genetically modified mice), and direct translation to the clinic (through the use of tracers that can be directly applied to patients in clinical PET settings). Nevertheless, few studies have rigorously evaluated the performance of in vivo microPET compared to traditional instruments such as behaviour or post-mortem measures. In this study, we endeavoured to compare estimates of the severity of the lesion induced by intracerebral 6-OHDA using two in vivo measures, behaviour and microPET imaging, and one in vitro measure, autoradiography. The first challenge was to decide which PET tracer to use because changes in the availability of synaptic dopamine (DA) often lead to compensatory responses by pre- and/or post-synaptic aspects of D A neurotransmission. For example, D A receptors can be upregulated by the application of D A antagonists (Huang et al., 1997) or D A . terminal degeneration as in PD (Rinne et al., 1995) or experimental lesion models (Narang and Wamsley, 1995; Nikolaus et al., 2001; Doudet et al., 2002), while the D A transporter (DAT) is upregulated by chronic cocaine administration (Wilson et a l , 1994), but downregulated by pharmacological D A depletion (Kilbourn et al., 1992) or in early parkinsonism (Lee et al., 2000; Doudet, 2001). The vesicular monoamine transporter ( V M A T 2 , which packages dopamine into vesicles), however, appears to be insensitive to changes in synaptic DA, and compensatory responses to D A enhancing and depleting stimuli generally do not occur (Vander Borght et al., 1995; Wilson and Kish, 1996). The V M A T 2 is therefore widely regarded as a gold standard for measuring the integrity of striatal D A terminals. The drug most widely used to evaluate the V M A T 2 is tetrabenazine and its derivatives, methoxytetrabenazine and dihydrotetrabenazine (DTBZ). Both of the latter two have been radiolabeled with [3H] for in vitro binding studies and with [ U C] for in vivo PET studies (Frey et al., 1998). Thus, in this pilot validation study, we selected [ nC](±)dihydrotetrabenazine ( [ n C]DTBZ), and compared in vivo data from a microPET scanner to both in vitro autoradiography with the same tracer, and to behavioural analysis after a mild or severe unilateral 6-OHDA lesion of the nigrostriatal pathway. 66 Materials and Methods: Subjects Adult male Sprague-Dawley rats (n = 13, 300-350 g at the time of lesion) were housed in dyads with food and water available ad libitum. Their housing room was kept at constant humidity (55%) and temperature (21°C), and had a 12:12 light:dark cycle, with lights off at 12:00 p.m. 6-OHDA Lesioning A l l animals were given desipramine hydrochloride (Sigma, 25 mg/kg i.p.) 30-60 min before 6-OHDA infusion to protect noradrenergic terminals. Animals were anaesthetized with isoflurane in O2 (4% for induction, 1% for maintenance), given atropine sulfate (0.05 mg/kg s.e), and placed into a stereotaxic head holder (Kopf). Two different lesion protocols were used, meant to create either a severe or a mild unilateral lesion. The animals received either a total of 20 ug of 6-OHDA hydrobromide (Sigma; n = 7, rats A-G) delivered in 2 sites along the medial forebrain bundle (10 ug in 4 ul of saline plus 0.05% ascorbic acid per site; Coordinates: site 1: AP -2.8 mm, M L -1.8 mm, D V -8.0 mm (all from Bregma); site 2: AP -4.7 mm (Bregma), M L -1.5 mm (midline), D V -7.9 mm (hole), according to (Paxinos and Watson, 1997)) or 2 ug of 6-OHDA at a single site along the M F B (the 2 n d infusion site listed above; n = 6, rats H - M). The infusion rate was 1 uL per min, and the cannula was left in place an additional 4 min to allow diffusion of the 6-OHDA solution. After surgery, the animals received subcutaneous saline, antibiotics (Duplocillin 0.1 ml/kg i.m.), and analgesia (Anafen 2 mg/kg s.e), and were kept warm in an incubator until fully recovered from anaesthesia. The animals were allowed two weeks to recover from surgery, and their weight and condition were monitored daily during the first week of recovery. Two of the study animals in the severe lesion group were recruited from a different study, where they were excluded after lesioning based on their behavioural findings (few errors made with the hindlimb contralateral to the lesion on the tapered/ledged beam-walking test - see below). Tapered/Ledged Beam-Walking (TB) Test Prior to lesioning, animals were trained on a tapered/ledged beam-walking (TB) test, adapted with slight modifications from Schallert and colleagues (Schallert and Woodlee, 2005; Zhao et al., 2005). This test is similar to traditional rat beam-walking tasks most commonly used in models of stroke, but also sensitive to DA function (Feeney et al., 1982; Walsh and Wagner, 1992; Bowenkamp et al., 2000). The TB test apparatus is shown in Figure 4.1. This task requires the animals to walk along a 165 cm long, progressively narrowing (6.5 cm wide at the wide end, 1.5 cm at the narrow end) Plexiglas beam, elevated above the floor on an incline of 15 degrees, to reach their darkened home cage. The main surface of the beam is covered in rubber matting to provide traction. Two cm below the beam, there is a 2.5 cm wide Plexiglas ledge, which provides a platform to step on when there is a motor deficit. It is this ledge that makes this particular beam-walking test unique, in that the ledge allows the animals to express their motor deficit, and removes the need for postural compensation to prevent falling off the beam (Schallert and Woodlee, 2005). Taking a step with only 1 or 2 toes on the main surface of the beam (and the other 4 or 3 toes overhanging the ledge) is scored as a half footfault, while stepping with the entire foot on the ledge rather than on the main surface of the beam is scored as a full footfault. Training on the TB test involves 10 trials, with the animal starting further and further away from the home cage goal (3 trials starting at the beginning of the narrow end, 3 starting at the beginning of the medium end, and 4 starting at the wide end). In between trials, the animal remains in its home cage with the lights in the room turned off for one minute for reinforcement. The animals learn this task quickly, and training need only occur once in the 67 lifetime of the animal. Normal animals make very few errors (footfaults), and those only occur on the narrowest section of the beam. H 45 cm 45 cm 45 cm 1 6.5 cm A V 3 D IT 1.5 cm 165 cm Figure 4.1: The tapered/ledged beam-walking (TB) test apparatus. A l l animals were trained prior to lesioning and tested 2-3 weeks post-lesion, a period which corresponded to 1 -2 weeks prior to the microPET scan. On the testing day, the animals were brought to the behavioural room during the dark cycle and allowed to habituate for 15 - 30 min. Before testing, each animal is allowed one "refresher" trial, which is not videotaped. One TB test is made up of 5 consecutive trials, and each trial was videotaped from the rear to allow a clear observation of the hindlimbs, and scored at a later date by an investigator blind to the animal's lesion condition. For each hindlimb, the number of steps taken in each of the three bins (wide, medium and narrow portions of the beam) and the number of full and half footfaults in the 5 trials was determined. Because the scores represent mainly large or small proportions (# errors per steps), transforming the data via an arcsine transformation (Eq. 4.1) normalizes the data (Zar, 1999). Equation 4.1:p'• ^[(arcsin VxTn+T) + (arcsin Vx + 1/n + 1)] Where X = # errors, and n = # steps. MicroPET Imaging On the day of the PET scan, the animal was brought to the scanning room and anaesthetized with ketamine:xylazine (90:10 mg/kg i.m.), and given atropine sulfate (0.05 mg/kg s.e). After the induction of anaesthesia, the animal was positioned prone in a plastic head holder attached to the bed of the scanner (Rubins et al., 2001) to ensure that all of the study animals were positioned with a flat skull and in the same plane, and an i.v. line was placed in a lateral tail vein. A transmission scan with a rotating Ge-68 point source supplied by the scanner 68 manufacturer was performed prior to radiotracer injection and the data were used for attenuation correction. The tracer [ U C]DTBZ was synthesized as described previously (Jewett et al., 1997). Animals were injected with 0.1 mCi/lOOg body weight [ U C]DTBZ (278-2738 Ci/mmol at injection; median 1603 Ci/mmol), and scanning with a microPET R4 scanner (Concorde Microsystems, Knoxville TN) occurred for the 30-60 min post-injection. Data were reconstructed by filter back projection after attenuation and scatter correction (Alexoff et al., 2003), and the images from the 16 time frames were summed to produce a higher statistic image of the rat brain. Regions of interest (ROI) analysis was performed using the system's software (ASIPro 6.0). Small, circular ROIs (area = 6 mm2) were placed bilaterally on the 3 slices containing striatum, and oval ROIs (area = 20 mm2) were placed in two slices containing cerebellum. Binding data are presented as the average of the left or right striatal ROI values divided by the average of the cerebellar ROI values, minus one, to give an estimate of the left or right striatal binding potential (BP). Autoradiography After fully recovering from the anaesthesia used for the microPET scan, the animals were sacrificed by decapitation, and the brain quickly removed and frozen in isopentane cooled to -70°C with dry Ice. The brains were stored in a freezer at -80°C until sectioning on a cryostat (Leica) at -17°C. Twenty micron sections were taken through the striatum and thaw mounted onto glass slides (Superfrost Plus, Fisher Scientific). The tissue was then stored in the -80°C freezer until autoradiography was performed The details of the quantitative autoradiographic procedure have been described in detail (Strome et al., 2005). Specifically, for [ U C]DTBZ binding, the buffers were taken from (Vander Borght et al., 1995). The slides were removed from the freezer and allowed to warm,up to room temperature. Following pre-washing for 5 min at 25°C in sucrose buffer (300 m M sucrose, 50 mM Tris-HCl, 1 m M EDTA, pH 8.0), the slides were incubated for 30 min in the same buffer with the addition of 5 nM [ n C]DTBZ. Non-specific binding was performed simultaneously on an adjacent set of sections with the addition of 1 u M unlabelled tetrabenazine. After incubation, the slides were washed twice for 3 min each in the same sucrose buffer at 25°C, and then briefly dipped in deionized distilled water at 4°C to remove excess sucrose. The slides were then allowed to dry for 20 min in a fumehood, and then placed against multisensitive storage phosphor screens (Perkin-Elmer) along with a set of standards of known activity (Strome et al., 2005) in standard film cassettes. The cassettes were stored behind lead for the 2 h exposure time, at which point, the screens were removed and scanned in a Cyclone phosphor imaging system (Perkin-Elmer) at 600 dpi. The binding data were analyzed as described (Strome et al., 2005) using the Cyclone's inherent software (Optiquant v.4.0). Small circular ROIs (area = 5 mm ) were placed in the striatum. Total binding data were averaged over 6 sections per animal from at least 3 levels of the striatum (anterior, mid and posterior, chosen to duplicate the three striatal frames on which ROIs were placed in the microPET data), and non-specific binding was subtracted. The data were converted from D L U (digital light units) to pmol/cc tissue using the standard curves and the specific activity of the tracer at incubation (Strome et al., 2005). Statistical Analysis Because the animals primarily made mistakes in the narrow section of the beam, each animal's TB test score was calculated based on the number of errors made per steps taken on the narrow section of the beam in all 5 trials with both hindlimbs. The distribution of the autoradiographic, microPET and behavioural data was tested for normality (Kolmogorov-69 Smironov test), and Pearson product moment correlation analysis was performed with GraphPad Prism v.3.0 for Windows (San Diego, CA). Results: TB test data were not obtained for 2 animals (animals E and H in Table 1 and Fig. 4.3) due to a technical mishap (footage accidentally taped over). Figure 4.2 shows correlation analyses of the three measures of striatal D A integrity. There was a significant correlation between in vivo and in vitro measures of striatal [ U C]DTBZ binding (Pearson product-moment correlation, r = 0.64 p < 0.005; Fig. 4.2a). There was also a significant correlation between hindlimb errors on the narrow section of the TB test and striatal [' CJDTBZ binding as measured in vitro with autoradiography (r = -0.43, p < 0.05; Fig. 4.2b). TB test scores were not significantly correlated with striatal [ n C ] D T B Z binding as measured in vivo with microPET (r = -0.05, p > 0.05; Fig. 4.2c). Table 4.1 shows the lesion severity as measured in vivo by PET and at post-mortem with autoradiography. Five out of seven that received the high dose of 6-OHDA (animals A-E) showed severe unilateral depletion of striatal [ n C ] D T B Z binding as measured with autoradiography. The remaining two animals in this lesion protocol (animals E-F) did not show any marked asymmetry in striatal [ n C ] D T B Z binding as measured with autoradiography, and, in fact, were the two who had been recruited from a different study in our laboratory where they had been excluded for having mild lesions based on a lack of errors with the hindlimb contralateral to the lesion on the TB test. Animals that received a low dose of 6-OHDA experienced mild unilateral 6-OHDA lesions, ranging from 30-65% according to the autoradiographic results. Similarly, these animals displayed little or no behavioural impairments (Fig. 4.3). Table 1 also shows that lesion severity measured with PET is much lower than lesion severity measured with autoradiography in the five severely lesioned animals (A-E), whereas in the animals with milder lesions, while lesion severity was generally lower when measured by PET, the two techniques gave much more comparable results. Figure 4.3 shows TB test scores for both hindlimbs in the narrow section of the beam. Only animals that had severe unilateral 6-OHDA lesions as measured by autoradiography (> 90%> depletion of [ n C ] D T B Z binding in the lesioned striatum), the gold standard for measuring striatal D A terminal integrity after 6-OHDA lesion, also showed a severe impairment in their hindlimb contralateral to the lesion on the TB test (animals A - D ; Fig. 4.3). Figure 4.4 shows examples of microPET and autoradiographic images of striatal [ n C ] D T B Z binding in three representative animals, showing A) an intact animal, B) an animal with a mild unilateral lesion, and C) an animal with a severe unilateral lesion . In the intact striatum, in vitro [ n C ] D T B Z binding was 176.50 i 14.31 pmol/cc (mean ± SD), and in vivo, the mean BP was 2.07 ± 0.68. 70 Table 4.1: Comparison of percent lesion after unilateral infusion of 20 jig (High Dose) or 2 Hg (Low Dose) 6-OHDA, as measured by striatal [nC]DTBZ binding in vivo by microPET and in vitro with autoradiography. Animal ID % Lesion in vivo" % Lesion in vitrob High Dose A 60.84 92.59 6-OHDA B 59.17 98.40 C 49.67 96.49 D 41.73 92.92 E 58.23 91.94 F 47.06 44.07 G 62.90 74.42 Low Dose H 26.36 29.92 6-OHDA I 40.94 56.25 J 36.36 30.76 K 31.82 45.82 L 36.00 64.49 M 30.12 30.80 a. (100 - % ratio ['*C]DTBZ BP lesioned/non-lesioned striatum) b. (100 - % ratio [ n C ] D T B Z binding lesioned/non-lesioned striatum) 71 DTBZ Binding (pmol/cc) C) r = -0.05, p > 0.05 7 5 - | •D d) I M p a t o * 5 0 -g 2 8 " 2 5 " DTBZ BP Figure 4.2: Relationships between the three measures of striatal DA integrity. A) There is a significant correlation between autoradiography and microPET. B) There is a significant negative correlation between autoradiography and the TB test. C) There is no correlation between microPET and the TB test. 72 E™3 Ipsilateral • • Contralateral 75n High Dose 6-OHDA Low Dose 6-OHDA F i g u r e 4.3: T B t e s t s c o r e s f o r e a c h h i n d l i m b i n i n d i v i d u a l a n i m a l s a f t e r s e v e r e ( A - G ) o r m i l d ( H - M ) u n i l a t e r a l 6 - O H D A l e s i o n i n g . Only data from the narrow section of the beam are shown. Black bars represent the hindlimb contralateral to the lesion. Figure 4.4: Coronal images of striatal [ n C ] D T B Z binding through the head and brain with microPET (left) and in brain only with autoradiography (right). A) A normal control animal for comparison. B) A n animal with a mild right unilateral lesion. C) A n animal with a severe right unilateral lesion. For the purposes of publication, microPET images were maximum a posteriori (MAP) reconstructed. 74 Discussion: The results of this study show that there is a strong correspondence between striatal D A terminal integrity as measured in vivo by microPET and in vitro by autoradiography using the same tracer, [ n C ] D T B Z . In addition, TB test scores were significantly correlated with our autoradiographic determination of striatal D A terminal integrity, but were not significantly correlated with striatal D A terminal integrity as measured by microPET. The TB test appears to be a good behavioural measure of DA-dependent motor function only in severely unilateral 6-OHDA-lesioned rats, suggesting that a lack of impairment on this task could be used as an exclusion criterion for animals with poor lesions prior to further interventions. Our in vitro autoradiographic data are in good agreement with literature values of DTBZ binding in the intact and 6-OHDA-lesioned rat striatum. Although this is the first autoradiographic study with [ U C]DTBZ, [ 3H]DTBZ has been widely employed to study the V M A T 2 , and our quantitative data correspond well to previous studies (Darchen et al., 1989; Masuo et al., 1990; Vander Borght et al., 1995). Our data are also in good agreement with previous studies showing a strong positive correlation between small animal PET imaging of the brain and autoradiographic analysis with the same tracer. However, whereas previously it has been shown that this is the case for ex vivo 1 R autoradiography with the D2 receptor ligand [ F](7V-mefhyl)benperidol (Nikolaus et al., 2003) and 2-[18F]fluoro-2-deoxy-D-glucose (Moore et al., 2000; Matsumura et al., 2003), our data are the first to show the strong correspondence with a C - l l labeled tracer and in vitro autoradiography in the same animal. It is notable that although the correlation between [ n C ] D T B Z binding in the same animals in vivo compared to in vitro is very high, as shown in Fig. 2a, this correlation is not a perfect 1 to 1 relationship and the slope has a positive intercept (i.e. it does not go through the origin). This phenomenon arises from an overestimation of the low BP values in the lesioned striatum with microPET, and may have several explanations, including the existence of anaesthesia-related effects in the in vivo studies, and/or a bias in the in vivo data due to the intrinsic properties of the scanner, in particular, the limited resolution.. We chose to perform our microPET studies under ketamine/xylazine anaesthesia. Although ketamine, a non-competitive N M D A antagonist, is known to affect the binding of other DAergic radiotracers, namely those for Di (Momosaki et al., 2004) and D2 (Kobayashi et al., 1995; Vollenweider et al., 2000) receptors, it is not known how ketamine or ketamine/xylazine may influence the binding of [ n C ] D T B Z to the V M A T 2 . An ongoing study of the effects of anaesthesia on microPET data in our lab indicates that, while, in comparison to isoflurane anaesthesia, ketamine/xylazine substantially decreases [ n C ] D T B Z binding in vivo, the effect is uniform in the striatum and cerebellum, and does not influence the striatumxerebellum ratio (unpublished observations). At this point, we cannot exclude the possibility that ketamine/xylazine (or any anaesthesia) may have a differential influence on the binding of [ n C ] D T B Z in the lesioned vs. intact striatum. If the binding was affected to a greater extent in the intact striatum, this would decrease the apparent lesion severity, a phenomenon observed in this study, but which we attribute mainly to the partial volume effect (see next paragraph). As shown in Table 1, while we were able to detect severe unilateral lesions autoradiographically, we were unable to detect large asymmetries in striatal [ n C ] D T B Z binding with microPET. In animals with severe unilateral lesions (>90%) as measured by autoradiography, microPET showed only 42-61% lesions, whereas in the animals that did not suffer severe lesions, the striatal asymmetry in [ n C ] D T B Z binding is comparable when measured either in vivo or in vitro. The inability to measure large asymmetries in striatal [ U C]DTBZ binding in vivo in unilateral 6-OHDA lesioned rats most likely results from the 75 partial volume effect. Because the size of the rat striatum (width -2-3 mm; (Paxinos and Watson, 1997) is generally less than twice the resolution of the scanner (~ 2 mm F W H M at the center of the field of view in each direction), it is susceptible to partial volume effects (Mazziotta et al., 1981). The partial volume effect would be particularly strong in the lesioned striatum, due to spillover from the adjacent intact striatum, which would lead to an overestimation of the true radioactivity concentration in the lesioned striatum. The partial volume effect, therefore, apparently decreases the quantitative accuracy of the microPET determination of [ n C ] D T B Z binding in the lesioned striatum, suggesting that post-mortem measures should be included in any study examining unilateral striatal lesions and microPET imaging. More recent, better resolution scanners, as well as the use of other reconstruction algorithms allowing higher resolution and the use of co-registration of microPET with CT or MRI may also improve the identification of specific brain regions and ROI placement, and thus reduce the apparent effect of partial voluming. In addition to the partial volume effect, the microPET data in this experiment may also be subject to a "floor effect", i.e. low sensitivity of the technique in measuring low striatal [ U C]DTBZ binding. We are currently performing a number of follow-up studies to further investigate some of these issues. The TB test appears to be most sensitive in animals that have severe unilateral lesions, as these were the only animals in our study to show strong asymmetries between TB test scores with the hindlimbs ipsilateral and contralateral to the lesion (animals A - E ; Fig. 3). Two of the animals in this study (animals F and G) had originally been assigned to another study, but were excluded based on their TB test scores (few errors made with the hindlimb contralateral to the lesion), and transferred to the current investigation to assess the status of the D A nigrostriatal system. Not surprisingly, these two animals also showed poor lesions when evaluated with both microPET and autoradiography. Indeed, none of the animals that had mild lesions according to our autoradiographic results showed a clear deficit with the hindlimb contralateral to the lesion on the TB task, and, in fact, some animals made more errors with the hindlimb ipsilateral to the lesion (Fig. 3). Since the severe unilateral 6-OHDA-lesion model is the most commonly employed rat model of PD, our data suggests that the TB test can be used to screen animals with poor lesions and remove them from further analysis early in a study, thereby reducing the waste of time, energy, and materials that occurs when an animal is excluded at the end of a study due to a poor lesion. The same inherent issues with quantification of microPET data in unilateral 6-OHDA-lesioned rats discussed above may have contributed to the fact that TB test scores and [ n C ] D T B Z binding measured with microPET were not significantly correlated, even though the autoradiographic results did correspond well to the TB test scores. Taken together, the limitations of microPET in measuring strong asymmetries in striatal D A innervation, and of the TB test in measuring mild asymmetries in striatal D A innervation, likely led to a poor correlation between these two measures. Many beam-walking tasks designed to evaluate motor function have been reported in the literature, but the task that we report here is unique in its design. The ledge positioned below allows the animal to make slips off the main surface of the beam without a fear of falling, and thus allows them to express any motor deficit without masking it with postural compensations (Schallert and Woodlee, 2005). We have found this task to be useful in evaluating motor deficits in both severely unilaterally and bilaterally 6-OHDA-lesioned rats, and have been able to detect significant improvement in lesioned animals after anti-parkinsonian interventions (unpublished). Non-pharmacological behavioural testing in models of PD is an emerging trend, as evidenced by the popularity of many new tasks including the Forelimb Use Asymmetry Test ("Cylinder Test"), the Forelimb Placing Task, and the Sensorimotor Asymmetry Test ("Sticker Test") (Schallert et al., 2000; Schallert and Woodlee, 2005), and the TB test is a valuable addition to the 76 battery of tools available to evaluate motor behaviour without the use of drugs, thereby removing an additional potential confound. In conclusion, we have shown in this study, that while measurement of the integrity of the striatal D A system using microPET correlates significantly with in vitro autoradiography with the same tracer, autoradiography is more sensitive at detecting lesion severity, likely due to the substantial partial volume effect that can occur when imaging with the microPET R4 in the unilaterally 6-OHDA-lesioned striatum. Autoradiographic measurement of striatal [ n C ] D T B Z binding was also significantly correlated with behavioural measures of DA-dependent motor function. This study highlights the limitations of both behavioural testing and microPET imaging in the unilateral 6-OHDA-lesion rat model of PD, and emphasizes the strengths of autoradiographic analysis of striatal D A innervation. We also suggest that the TB test can be used in studies employing the severe unilateral 6-OHDA-lesion model as a means of quickly and easily evaluating lesion severity to exclude animals with poor lesions prior to further intervention. 77 References: Alexoff D, Vaska P, Marsteller D, Gerasimov T, L i J, Logan J, et al.(2003) Reproducibility of "C-raclopride binding in the rat brain measured with microPET R4: effects of scatter correction and tracer specific activity. J Nucl Med 44: 815-822. Bowenkamp K E , Ujhelyi L, Cline EJ, Bickford PC (2000) Effects of intra-striatal GDNF on motor coordination and striatal electrophysiology in aged F344 rats. Neurobiol Aging 21: 117-124. Cohen A D , Tillerson JL, Smith A D , Schallert T, Zigmond M J (2003) Neuroprotective effects of prior limb use in 6-hydroxydopamine-treated rats: possible role of GDNF. J Neurochem 85: 299-305. Costall B, Naylor RJ, Pycock C (1976) Non-specific supersensitivity of striatal dopamine receptors after 6-hydroxydopamine lesion of the nigrostriatal pathway. Eur J Pharmacol 35: 276-283. Darchen F, Masuo Y , Vial M , Rostene W, Scherman D (1989) Quantitative autoradiography of the rat brain vesicular monoamine transporter using the binding of [3H]dihydrotetrabenazine and 7-amino-8-[125I]iodoketanserin. Neuroscience 33: 341-349. Doudet DJ (2001) PET studies in the MPTP model of Parkinson's disease. Adv Neurol 86: 187-195. Doudet DJ, Jivan S, Ruth TJ, Holden JE (2002) Density and affinity of the dopamine D 2 receptors in aged symptomatic and asymptomatic MPTP-treated monkeys: PET studies with [nC]raclopride. Synapse 44: 198-202. . Feeney D M , Gonzalez A , Law W A (1982) Amphetamine, haloperidol, and experience interact to affect rate of recovery after motor cortex injury. Science 217: 855-857. Frey K A , Wieland D M , Kilbourn M R (1998) Imaging of monoaminergic and cholinergic vesicular transporters in the brain. Adv Pharmacol 42: 269-272. Guo H, Tang Z, Y u Y , X u L, Jin G, Zhou J (2002) Apomorphine induces trophic factors that support fetal rat mesencephalic dopaminergic neurons in cultures. Eur J Neurosci 16: 1861-1870. Huang N , Ase AR, Hebert C, van Gelder N M , Reader T A (1997) Effects of chronic neuroleptic treatments on dopamine Di and D 2 receptors: homogenate binding and autoradiographic studies. Neurochem Int 30: 277-290. Jewett D M , Kilbourn MR, Lee L C (1997) A simple synthesis of [nC]dihydrotetrabenazine (DTBZ). Nucl Med Biol 24: 197-199. Kilbourn MR, Sherman PS, Pisani T (1992) Repeated reserpine administration reduces in vivo [ 1 8 F]GBR 13119 binding to the dopamine uptake site. Eur J Pharmacol 216: 109-112. Klug JM, Norman A B (1993) Long-term sensitization of apomorphine-induced rotation behavior in rats with dopamine deafferentation or excitotoxin lesions of the striatum. Pharmacol Biochem Behav 46: 397-403. 78 Knoess C, Siegel S, Smith A , Newport D, Richerzhagen N , Winkeler A , et al. (2003) Performance evaluation of the microPET R4 PET scanner for rodents. Eur J Nucl Med Mol Imaging 30: 737-747. Kobayashi K , Inoue O, Watanabe Y , Onoe H, Langstrom B (1995) Difference in response of D 2 receptor binding between 1 lC-N-methylspiperone and nC-raclopride against anesthetics in rhesus monkey brain. J Neural Transm Gen Sect 100: 147-151. Lee CS, Samii A , Sossi V , Ruth TJ, Schulzer M , Holden JE, et al. (2000) In vivo positron emission tomographic evidence for compensatory changes in presynaptic dopaminergic nerve terminals in Parkinson's disease. Ann Neurol 47: 493-503. Masuo Y , Pelaprat D, Scherman D, Rostene W (1990) [3H]Dihydrotetrabenazine, a new marker for the visualization of dopaminergic denervation in the rat striatum. Neurosci Lett 114: 45-50. Matsumura A , Mizokawa S, Tanaka M , Wada Y , Nozaki S, Nakamura F, et al. (2003) Assessment of microPET performance in analyzing the rat brain under different types of anesthesia: comparison between quantitative data obtained with microPET and ex vivo autoradiography. Neuroimage 20: 2040-2050. Mazziotta JC, Phelps M E , Plummer D, Kuhl DE (1981) Quantitation in positron emission computed tomography: 5. Physical—anatomical effects. J Comput Assist Tomogr 5: 734-743. Meredith GE, Callen S, Scheuer D A (2002) Brain-derived neurotrophic factor expression is increased in the rat amygdala, piriform cortex and hypothalamus following repeated amphetamine administration. Brain Res 949: 218-227. Momosaki S, Hatano K, Kawasumi Y , Kato T, Hosoi R, Kobayashi K, et al. (2004) Rat-PET study without anesthesia: anesthetics modify the dopamine Di receptor binding in rat brain. Synapse 54: 207-213. Moore A H , Osteen CL, Chatziioannou A , Hovda DA, Cherry SR (2000) Quantitative assessment of longitudinal metabolic changes in vivo after traumatic brain injury in the adult rat using FDG-microPET. J Cereb Blood Flow Metab 20: 1492-1501. Morelli M , Fenu S, Garau L, Di Chiara G (1989) Time and dose dependence of the 'priming' of the expression of dopamine receptor supersensitivity. Eur J Pharmacol 162: 329-335. Mueller D, Chapman C A , Stewart J (2006) Amphetamine induces dendritic growth in ventral tegmental area dopaminergic neurons in vivo via basic fibroblast growth factor. Neuroscience 137: 727-735. Narang N , Wamsley JK (1995) Time dependent changes in D A uptake sites, Di and D 2 receptor binding and mRNA after 6-OHDA lesions of the medial forebrain bundle in the rat brain. J Chem Neuroanat 9: 41-53. Nikolaus S, Larisch R, Beu M , Hamacher K, Forutan F, Vosberg H , et al. (2003) In vivo measurement of D 2 receptor density and affinity for 18F-(3-N-methyl) benperidol in the rat striatum with a PET system for small laboratory animals. J Nucl Med 44: 618-624. 79 Nikolaus S, Larisch R, Beu M , Vosberg H, Muller-Gartner H W (2001) Imaging of striatal dopamine D 2 receptors with a PET system for small laboratory animals in comparison with storage phosphor autoradiography: a validation study with 18F-(N-Methyl)Benperidol. J Nucl Med 42: 1691-1696. Paxinos G, Watson C (1997) The Rat Brain in Stereotaxic Coordinates. San Diego: Academic Press. Rinne JO, Laihinen A , Ruottinen H, Ruotsalainen U , Nagren K , Lehikoinen P, et al. (1995) Increased density of dopamine D 2 receptors in the putamen, but not in the caudate nucleus in early Parkinson's disease: a PET study with [nC]raclopride. J Neurol Sci 132: 156-161. Rubins DJ, Meadors A K , Yee S, Melega WP, Cherry SR (2001) Evaluation of a stereotactic frame for repositioning of the rat brain in serial positron emission tomography imaging studies. J Neurosci Methods 107: 63-70. Schallert T, Fleming SM, Leasure JL, Tillerson JL, Bland ST (2000) CNS plasticity and assessment of forelimb sensorimotor outcome in unilateral rat models of stroke, cortical ablation, parkinsonism and spinal cord injury. Neuropharmacology 39: 777-787. Schallert T, Woodlee M T (2005) Orienting and Placing. In: The behavior of the laboratory rat: a handbook with tests. (Whishaw IQ, Kolb B, eds), pp 129-140. New York: Oxford University Press. Shi L - H , Woodward DJ, Luo F, Anstrom K, Schallert T, Chang J-Y (2004) High-frequency stimulation of the subthalamic nucleus reverses limb-use asymmetry in rats with unilateral 6-hydroxydopamine lesions. Brain Res 1013: 98-106. Strome E M , Jivan S, Doudet DJ (2005) Quantitative in vitro phosphor imaging using [3H] and [1 8F] radioligands: the effects of chronic desipramine treatment on serotonin 5-HT2 receptors. J Neurosci Methods 141: 143-154. Tai C, Chatziioannou A , Siegel S, Young J, Newport D, Goble R N , et al. (2001) Performance evaluation of the microPET P4: a PET system dedicated to animal imaging. Phys Med Biol 46: 1845-1862. Tai Y C , Chatziioannou A F , Yang Y , Silverman RW, Meadors K , Siegel S, et al. (2003) MicroPET II: design, development and initial performance of an improved microPET scanner for small-animal imaging. Phys Med Biol 48: 1519-1537. Tai Y C , Ruangma A , Rowland D, Siegel S, Newport DF, Chow PL, Laforest R (2005) Performance Evaluation of the microPET Focus: A Third-Generation microPET Scanner Dedicated to Animal Imaging. J Nucl Med 46: 455-463. Ungerstedt U (1968) 6-hydroxy-dopamine induced degeneration of central monoamine neurons. Eur J Pharmacol 5: 107-110. Vander Borght T, Kilbourn M , Desmond T, Kuhl D, Frey K (1995) The vesicular monoamine transporter is not regulated by dopaminergic treatments. Eur J Pharmacol 294: 577-583. 80 Vollenweider F X , Vontobel P, Oye I, Hell D, Leenders K L (2000) Effects of (S)-ketamine on striatal dopamine: a [uC]raclopride PET study of a model psychosis in humans. J Psychiatry Res 34:35-43. Walsh SL, Wagner GC (1992) Motor impairments after methamphetamine-induced neurotoxicity in the rat. J Pharmacol Exp Ther 263: 617-626. • Wilson JM, Kish SJ (1996) The vesicular monoamine transporter, in contrast to the dopamine transporter, is not altered by chronic cocaine self-administration in the rat. J Neurosci 16: 3507-3510. Wilson JM, Nobrega JN, Carroll M E , Niznik HB, Shannak K , Lac ST, et al. (1994) Heterogeneous subregional binding patterns of 3 H-WIN 35,428 and 3 H - G B R 12,935 are differentially regulated by chronic cocaine self-administration. J Neurosci 14: 2966-2979. Yang Y , Tai Y C , Siegel S, Newport DF, Bai B, L i Q, et al. (2004) Optimization and performance evaluation of the microPET II scanner for in vivo small-animal imaging. Phys Med Biol 49: 2527-2545. Zar JH (1999) Biostatistical Analysis. Upper Saddle River: Prentice-Hall. Zhao CS, Puurunen K, Schallert T, Sivenius J, Jolkkonen J (2005) Behavioral and histological effects of chronic antipsychotic and antidepressant drug treatment in aged rats with focal ischemic brain injury. Behav Brain Res 158: 211-220. 81 Chapter 5 ELECTROCONVULSIVE SHOCK ENHANCES STRIATAL D! AND D3 RECEPTOR BINDING AND IMPROVES MOTOR PERFORMANCE IN 6-OHDA-LESIONED RATS1 Preamble: The previous three chapters of this thesis describe our preliminary studies evaluating both the animal models and the tools that we wanted to use to address our primary hypotheses regarding the mechanism of action of ECT in PD. These studies gave us a great deal of insight into the most appropriate methods and experimental design for the investigation of our main research question. First of all, it became clear that, although non-human primates would be a better model in which to investigate the effects of ECT in PD, their use was not financially, technically, or ethically feasible. For our goals, a rat model had many advantages, including the fact that the 6-OHDA rat model of PD is well-characterized, easily produced, and reproducible. Secondly, although the use of microPET imaging allows longitudinal studies to be performed, the difficulty of obtaining all the required tracers at specific time-points, as well as the poor resolution, precluded its consistent use for these preliminary studies. In addition, even though several groups are actively working at developing specific tracers for D 3 receptors, none is currently available for PET. Quantitative autoradiography, therefore, emerged as the technique of choice to investigate our hypotheses regarding ECS effects on D A receptors, and we have demonstrated several uses of the technique in these preliminary studies. Finally, several non-pharmacological behavioural tests have been described for the evaluation of motor deficits in the unilateral 6-OHDA rat model of PD that both allow simple evaluation of lesion severity, and are sensitive to therapeutic interventions. In the previous chapter, we evaluated the use of the TB test for characterizing hindlimb motor deficits in the unilateral 6-OHDA model, and showed that impairment on the task correlates with the degree of striatal D A terminal integrity as measured by autoradiography. Not included in the chapters, but reported in Appendix A , we also performed a small validation and feasibility study on the use of anaesthesia for rodent ECS, in an effort to design a rodent study as clinically relevant as possible. The outcome of this study led us to choose ketamine as the anaesthetic of choice for all future studies, based on its ease of use and relevance to the clinical situation, as well as the fact that we were able to reproduce the increase in 5-HT2 receptor binding widely reported after treating unanaesthetized rodents with ECS. At this point, we were ready to undertake our first investigations of the effects of repeated ECS treatment in a rat model of PD. This chapter describes our first study to directly address the working hypothesis (Fig. 1.1, page 10), and in it we examine Hypothesis 1, by evaluating forelimb and hindlimb motor function, as measured by the Cylinder and TB tests respectively, and Hypothesis 2, by measuring D A receptor binding using quantitative phosphor imaging autoradiography. ' A version of this chapter has been submitted for publication. Strome E M , Zis AP, Doudet DJ (2006) Electroconvulsive shock enhances striatal D, and D 3 receptor binding and improves motor performance in 6-OHDA-lesioned rats. 82 Introduction: Parkinson's disease (PD) is a neurodegenerative movement disorder characterized by tremor, bradykinesia, rigidity, and postural disturbances. The primary pathological features of PD are the loss of dopamine (DA)-producing cell bodies within the substantia nigra pars compacta of the midbrain, and the concomitant loss of striatal D A (Bernheimer et al., 1973). Therapeutic options for PD attempt to either replace the lost D A tone, for example by treating patients with L-DOPA (the biochemical precursor to DA) and/or D A 'agonists, or to inhibit the basal ganglia output structures that become overactive in PD via surgical intervention, as in the case of pallidotomy or deep .brain stimulation (DBS). Most patients, however, experience negative side effects or a loss of efficacy after prolonged L-DOPA treatment (reviewed in (Jankovic, 2005), and not all patients are good candidates for brain surgery. The development of adjunctive or alternative therapeutic options for PD is essential in order to improve the quality of life of patients living with this disease. Electroconvulsive therapy (ECT) is a widely used treatment for psychiatric disorders, and several recent meta-analyses have shown that it is the most effective antidepressant treatment available (Kho et al., 2003; Pagnin et al., 2004). Not only is ECT an effective treatment for depression, but it also appears to have positive effects on the motor symptoms of PD patients, regardless of whether or not they are depressed. Over 200 PD patients treated with ECT have been reported in the literature (Kennedy et al., 2003), with the majority showing dramatic improvement in their motor symptoms. Case reports, open trials, and double-blind placebo-controlled studies have all shown that ECT treatment significantly improves a wide range of motor symptoms, including rigidity, bradykinesia, and "on-off phenomenon, and the improvements last from several weeks to months after the last treatment (Andersen et al., 1987; Stern, 1991; Zervas and Fink, 1991; Fall et a l , 1995; Pridmore and Pollard, 1996). While there is still a need for further, well-designed clinical trials, with large numbers of subjects to determine the most appropriate parameters needed to achieve optimal impact on the motor symptoms of PD, this treatment holds great promise as a potential adjunct treatment. In animals, repeated treatment with electroconvulsive shock (ECS) has been shown to have specific effects in limbic brain regions such as the frontal cortex and hippocampus. One of the most consistent effects of repeated ECS on the normal rodent limbic system is an enhancement of serotonin (5-HT) neurotransmission, as evidenced by increased 5-HT-mediated behaviours (Green et al., 1983a; Goodwin et al., 1984; Moorman et al., 1996), increased interstitial 5-HT metabolites (Yoshida et a l , 1998), and upregulation of the 5-HT2 receptor (Green et al., 1983b; Butler et al., 1993). Taken together with increases in neurotrophic factors (Nibuya et al., 1995; Angelucci et al., 2002) and cell growth (Vaidya et al., 1999; Malberg et a l , 2000), the enhancement of 5-HT neurotransmission after repeated ECS is part of a cascade of cellular events thought to underlie the mechanism of action of ECT in mood disorders (Duman et al., 1997). We suggest that the effects of repeated ECS treatment on the D A system of the parkinsonian striatum may be similar to those observed on the 5-HT system of the hippocampus, and that these changes may partially underlie the positive effects of ECT on the motor symptoms of PD. In this study, we hypothesize that repeated ECS treatment in unilateral 6-hydroxydopamine (6-OHDA)-lesioned rats, a rat model of PD, will improve motor function and increase binding to D A receptors of the direct pathway of the basal ganglia. 83 Materials and Methods: Subjects Adult male Sprague-Dawley rats [bred at the University of British Columbia animal facility from Charles River Canada (Montreal, PQ) stock], weighing 250 g at the start of the experiment were housed on a 12:12 light:dark schedule (with lights off at 12:00 p.m.), at constant temperature and humidity (21°C, 55%). The animals had access to food and water ad libitum, and were housed in pairs. Twenty of the animals were used for Di receptor binding (n = 10 sham and 10 ECS) and 20 were used for D 2 and D 3 receptor binding. Cylinder Test data were pooled from the entire group of animals, while data for the TB test also include data from animals in a parallel study (Strome et al., 2006b). A l l procedures were approved by the University of British Columbia Committee on Animal Care. 6-OHDA Lesioning The animals were allowed to habituate and were handled for at least three days before receiving a right unilateral 6-OHDA-induced lesion of the D A nigrostriatal pathway. Desipramine hydrochloride (25 mg/kg i.p.; Sigma-Aldrich Canada, Oakville, ON) was administered 30-60 min before 6-OHDA infusion to protect noradrenergic terminals. Animals were anaesthetized with isoflurane in 0 2 (4% for induction, 1 % for maintenance), given atropine sulfate (0.05 mg/kg s.e), and placed into a stereotaxic frame (Kopf). With the skull flat between lambda and Bregma, a 2% solution of 6-OHDA hydrobromide (8 u.g in 4 uL 0.05% ascorbic acid in saline; Sigma) was infused at two sties along the medial forebrain bundle [site 1: AP -2.8 mm, M L -1.8 mm, D V -8.0 (all from Bregma); site 2: AP -4.7 mm (Bregma), M L -1.5 mm (midline), DV -7.9 mm (hole) according to (Paxinos and Watson, 1997)]. The infusion rate was 1 uL per min, and the cannula was left in place an additional 4 min to allow diffusion of the 6-OHDA solution. After surgery, the animals received subcutaneous'saline, antibiotics (Duplocillin 0.1 ml/kg i.m.), and analgesia (Anafen 2 mg/kg s.e), and were kept warm in an incubator until fully recovered from anaesthesia. Animals were allowed to recover for at least 2 weeks after surgery before being treated with ECS. Forelimb Use Asymmetry Test (Cylinder Test) Animals were evaluated in the forelimb use asymmetry test (Cylinder Test; Schallert et al., 2000; Schallert and Woodlee, 2005) at three time points: before lesion, after lesion, and after ECS or sham treatment. At each time point, the animals were placed in a Plexiglas cylinder (20 cm diameter x 30 cm high) elevated on a glass plate for a 3 min period on two consecutive days. Testing was done during the dark phase of the cycle and under red lighting. The trials were videotaped from below, and scored at a later date by an investigator blind to the animals' treatment. Forelimb placements on the walls of the cylinder were categorized as left independent, right independent, or simultaneous movements, and a forelimb use asymmetry score was calculated as: Equation 5.1: [(ipsi + Vi both) divided by (ipsi + contra + both)] x 100 (Schallert and Woodlee, 2005) 84 where ipsi and contra refer to the forelimbs ipsilateral and contralateral to the 6-OHDA-induced lesion respectively. Animals had to make greater than 20 movements at any given time point for their data to be included in the analysis, and no animals were excluded based on this criterion. Tapered/LedgedBeam-Walking (TB) Test Animals were also evaluated before lesion, after lesion, and after ECS or sham treatment on a tapered/ledged beam-walking (TB) test, adapted with slight modifications from Schallert and colleagues (Schallert and Woodlee, 2005; Zhao et al., 2005a). The details of the test have been described in detail (Strome et al., 2006a). Briefly, the animals are trained to walk across a 165 cm long beam which progressively narrows as they approach the goal (their darkened home cage). Beneath the surface of the beam is a ledge, which allows the animals to make slips with their feet off the main surface of the beam without falling off, and therefore prevents postural compensation in lesioned animals. Taking a step with only 1 or 2 toes on the main surface of the beam (and the other 4 or 3 toes overhanging the ledge) is scored as a half footfault, while stepping with the entire foot on the ledge rather than on the main surface of the beam is scored as a full footfault. In between trials, the animal remains in its home cage with the lights in the room turned off for one minute for reinforcement. Normal animals make very few errors (footfaults), and those only occur on the narrowest section of the beam. On the testing day, the animals were brought to the behavioural room at the start of the dark cycle (12:00 p.m.) and allowed to habituate for 15 - 30 min. Before testing, each animal is allowed one "refresher" trial, which is not videotaped. One TB test was made up of 5 consecutive trials, and each trial was videotaped from the rear to allow a clear observation of the hindlimbs, and scored at a later date by an investigator blind to the animal's condition. For each hindlimb, the number of steps taken in each of the three sections (wide, medium and narrow portions of the beam) and the number of full and half footfaults in the 5 trials was determined and summed to obtain a single composite score per TB test. Because the scores represent mainly large or small proportions (# errors per steps), transforming the data via an arcsine transformation (Eq. 2) normalizes the data (Zar, 1999). Equation 5.2: p' = ^[(arcsin Vx/n + 1) + (arcsin Vx + 1/n + 1)] Where X = # errors, and n = # steps. Electroconvulsive Shock Treatment Animals were assigned randomly to the ECS or sham treatment groups, and were treated every day for 10 days between 08:00 and 11:00. Atropine sulfate (0.2 mg/kg s.e.) was administered, followed 30 min later by ketamine hydrochloride (80 mg/kg i.p.). After induction of ketamine anaesthesia, animals were given either sham treatment (electrodes placed, but no current administered), or bilateral ECS (80-99 mA, 5-9.9 s, 70 pulse/s, 0.5 ms pulse width) via earclip electrodes coated with electroconductive gel using a small animal ECS machine (Model 57800, Ugo Basile, Italy). A l l animals received the same initial current dose, based on our previous experience with ECS in rats under ketamine anaesthesia, and current doses during subsequent treatments were modified based on the nature of the previous seizure. A l l ECS-treated animals experienced seizures of 13-19 s duration. A l l animals in this study consistently showed tonic hind limb extension. One sham-treated animal from each of the Di and D2/D3 binding groups was lost to ketamine anaesthesia. 85 Forty-eight hours after the last ECS, animals were decapitated and the brains were removed and quickly frozen in isopentane cooled with dry ice, and stored at -80°C until sectioning. Twenty micron coronal sections were cut at -18°C on a cryostat (Leica) and thaw-mounted onto glass microscope slides (Superfrost Plus, Fisher Scientific, Canada). The slides were stored at -80°C until the receptor binding assays were performed. Vesicular Monoamine Transporter-2 Binding For verification of the extent of lesion, coronal sections through the striatum were incubated with [ nC](±)dihydrotetrabenazine, a marker for D A terminals, which binds to the vesicular monoamine transporter-2 (DaSilva et al., 1994). The details of the autoradiographic technique have been described in detail in Chapter 4. Di Receptor Binding The slides were warmed up to room temperature, and pre-washed for 15 min in Tris-HCl buffer (50 mM Tris-HCl, 120 m M NaCl, 5 mM KC1, 2 m M CaCl 2 , 1 m M MgCl 2 , room temperature, pH 7.4, all from Sigma). Incubation was in 2 nM [ 3H]SCH 23390 (Perkin Elmer, Canada; specific activity 81 Ci/mmol) plus 30 nM ritanserin (to block 5-HT 2 receptors; Sigma), in the same buffer at 20°C for 45 min. Non-specific binding, was determined by incubating adjacent slices with an additional 10 u M (+)-butaclamol (Sigma). At the end of the incubation, the slides were washed for 2 x 3 min in fresh buffer at 4°C, dipped briefly in cold distilled water, and allowed to dry on the bench top overnight. After post-fixation in paraformaldehyde vapor under vacuum in a dessicator for 24 h (Liberatore et al., 1999), the slides were apposed to pre-erased tritium-sensitive phosphor screens (Fuji Medical Systems Inc., Stamford, CT) in standard film cassettes with [3H] microscales (Amersham, UK) for 3 days. On the third day, the screens were removed from the cassettes and scanned in the Cyclone phosphor imager at 600 dpi resolution. D2 Receptor Binding The slides were warmed up to room temperature, and pre-washed for 15 min in Tris-HCl buffer (50 m M Tris-HCl, 120 m M NaCl, 5 mM KC1, 2 m M CaCl 2 , 1 m M M g C l 2 , room temperature, pH 7.4, all from Sigma). The slides were incubated for 45 min in 3 nM [uC]raclopride (specific activity 959 or 1975 Ci/mmol at the start of incubation) in the same buffer at room temperature. Non-specific binding was determined by incubating adjacent slices with an additional 10 u M (+)-butaclamol. At the end of the incubation, the slides were washed for 3 x 1 min in fresh buffer at 4°C, dipped briefly in cold distilled water, and allowed to dry in a fumehood for 20 min. The slides were then apposed to Multisensitive storage phosphor screens (Perkin-Elmer) along with n C standards prepared as previously described (Strome et al., 2005). The screens were scanned as above after 2 h of exposure. Receptor Binding D 3 receptor binding was performed with R-(+)-7-Hydroxy-[3H]di-n-propyl-2-aminotetralin ([ HJ7-OH-DPAT) as described by (Levesque et al., 1992) with minor modifications. The slides were warmed up to room temperature, and pre-washed for 3 x 5 min in a HEPES buffer (50 m M HEPES, 1 mM EDTA, 0.1% bovine serum albumen, 120 mM NaCl, all from Sigma). Incubation was in 1 nM [ 3H]7-OH-DPAT (Amersham, specific activity 115 Ci/mmol) in the same buffer at room temperature for 90 min. Non-specific binding, was 86 determined by incubating adjacent slices with an additional 10 u M (+)-butaclamol. At the end of the incubation, the slides were washed for 3 x 1 min in fresh buffer at 4°C, dipped briefly in cold distilled water, and allowed to dry on the bench top overnight. Post-fixation, exposure and plate scanning were identical to that described for Di binding, except exposure time was 7 d. Data Analysis Optical density analysis was performed using the inherent software on the phosphor imager (Optiquant v4.00, Perkin-Elmer). D A receptor binding was measured in the nucleus accumbens (NAcc) shell and in the dorsal striatum (approximately + 1.70 mm from Bregma according to (Paxinos and Watson, 1997). Small regions of interest were placed bilaterally in at least four total binding and two adjacent non-specific binding sections for each animal in each region. The optical density data were converted to nCi/mg tissue using a standard curve derived from the [3H] or [ n C ] microscales. For each animal, non-specific binding was subtracted from total binding to get a measure of specific radiotracer binding. Statistical Analysis Repeated measures analysis of variance (ANOVA; treatment x time) was used for the analysis of the behavioural data because multiple measurements were made in the same animals. The effects of ECS treatment on D A receptor binding were evaluated using two-way (treatment x hemisphere) A N O V A . Post-hoc testing of significant main effects was performed using Tukey's honest significant difference test for unequal n. A l i statistical analyses were performed using the software program StatSoft Statistica '98 v5.1 (Tulsa, OK). Results: Cylinder Test Figure 5.1 shows asymmetry scores from the Cylinder Test. There was no significant effect of ECS treatment on forelimb use asymmetry. Tapered/Ledged Beam-Walking Test Figure 5.2 shows the results from the TB test. The animals mainly made mistakes in the narrow section of the beam with their hindlimb contralateral to the lesion, so only these data are shown. Repeated measures A N O V A indicates that there was a significant interaction effect between treatment and time ( F i i 4 8 = 7.96, p < 0.007), and visual inspection of the data indicates that the ECS-treated group had lower scores after treatment than the sham-treated group (Fig. 5.2). Vesicular MonoamineTtransporter-2 Binding A l l animals showed > 90% depletion of vesicular monoamine transporter-2 binding in the lesioned compared to intact dorsal striatum (mean ± S E M = 94.05 ±1 .89 %, data not shown, but Fig. 4.4C, in Chapter 4, page 74 is representative). 87 100-i Sham ECS (n = 17) (n = 16) Figure 5.1: Forelimb use asymmetry scores before and 4 8 h after repeated ECS or sham treatment. There was no significant effect of repeated ECS treatment. Values are mean ± SEM. Figure 5.2: T B test data for the hindlimb contralateral to the lesion on the narrow section of the beam in ECS- and sham-treated rats before and 48 h after treatment. There is a significant treatment x time interaction (*p < 0.007), with ECS-treated rats showing lower scores after treatment, compared to sham-treated controls. 88 D j Receptor Bin ding There was a significant effect of treatment on Di binding in the dorsal striatum (Two-way A N O V A : Fi,34 = 5.20, p < 0.03; Figure 5.3a), with post-hoc testing indicating that Di receptor binding was significantly increased after ECS treatment (p < 0.04). There was also a significant effect of treatment on Di binding in the NAcc shell (F132 = 5.75, p < 0.03; Figure 5.3a), and again, post-hoc testing indicates that Di receptor binding was significantly increased after ECS treatment (p < 0.03). There were no significant main effects of hemisphere in these analyses, indicating that ECS treatment increased Di binding, regardless of the 6-OHDA lesion. D2 Receptor Binding There were no significant main effects of treatment, indicating that ECS-treatment had no effect on D2 receptor binding in either the dorsal striatum or NAcc shell. There was a significant effect of hemisphere on D2 binding in the dorsal striatum ( F i ^ = 47.45, p < 0.0001; Figure 5.3b), with post-hoc testing indicating that D2 receptor binding was significantly increased in the lesioned hemisphere (p < 0.0002). D3 Receptor Binding There was a significant effect of treatment on D 3 binding in the dorsal striatum (Two-way A N O V A : F134 = 4.55, p < 0.05; Figure 5.3c), with post-hoc testing indicating that D 3 receptor binding was significantly increased after ECS treatment (p < 0.05). There was also a significant effect of treatment on D 3 binding in the NAcc shell (F134 = 7.62, p < 0.01; Figure 5.3c), and again, post-hoc testing indicates that D 3 receptor binding was significantly increased after ECS treatment (p < 0.01). There was also a significant effect of hemisphere on D 3 binding in the NAcc shell (F [ ;34 = 4.55, p < 0.002; Figure 5.3c), with post-hoc testing indicating that D 3 receptor binding was significantly decreased in the lesioned hemisphere (p < 0.002). 89 D1 Binding A) SHAM ECS Dorsal Striatum Ventral Striatum D 2 Binding Dorsal Striatum Ventral Striatum D 3 B i n d i n g F i g u r e 5 . 3 : D A r e c e p t o r b i n d i n g v a l u e s i n the d o r s a l a n d v e n t r a l s t r i a t u m i n E C S - a n d s h a m - t r e a t e d r a t s . A ) Di binding was significantly increased after repeated ECS treatment in both the dorsal and ventral striatum (*p < 0.04). B) There was no effect of repeated ECS treatment on D 2 binding, but it was significantly increased in the lesioned dorsal striatum (#p < 0.001). C) D 3 receptor binding was significantly increased in both the dorsal and ventral striatum after repeated ECS treatment (*p < 0.04), and significantly decreased in the lesioned ventral striatum (#p < 0.002). N = 8-10 animals per group; values are mean ± SEM. 91 Discussion: This study shows that repeated ECS treatment in unilaterally 6-OHDA-lesioned animals increases dorsal striatal and NAcc shell Di and D 3 receptor binding, without affecting D2 binding, and improves hindlimb motor performance on a beam-walking task. The primary observation of the effects of ECT on PD patients is a fairly immediate and long-lasting improvement in their motor symptoms (Andersen et al., 1987; Zervas and Fink, 1991; Fall et al., 1995; Pridmore and Pollard, 1996). Very few studies, however, have examined motor behaviour after ECS treatment in rodents. Those reports looked at drug-induced behaviours (Wielosz, 1981; Green et al., 1983b; Smith and Sharp, 1997), with only one report in unilateral 6-OHDA-lesioned rats (Green et al., 1977), and all of these studies showed significant increases in DA-mediated behaviours after repeated ECS treatment. We hypothesized that non-pharmacological motor behaviour of unilateral 6-OHDA-lesioned rats would also improve after a course of ECS treatment. Our results indicate no significant effect of repeated ECS treatment on forelimb use asymmetry in the Cylinder Test (Fig. 5.1), but a significant improvement in hindlimb motor performance on the TB test (Fig. 5.2). Considering the severity of the unilateral lesion (> 90% depletion of striatal D A terminals), the absence of an effect of ECS on the use of the forelimb contralateral to the lesion in the Cylinder Test was not surprising. After unilateral 6-OHDA lesioning, it is not necessary for the animals to use the impaired forelimb for many tasks, including exploration, and preferential use of the unimpaired forelimb has been shown to develop very rapidly and persist for a long period of time after severe unilateral 6-OHDA lesion (Evenden and Robbins, 1984; Dunnett et al., 1987; Miklyaeva et al., 1994). This preference for using the forelimb ipsilateral to the lesion makes it more difficult for the animal to "relearn" the use of the impaired forelimb, even when therapeutic interventions occur, especially in the case of mild or short-term therapeutic effects. Some recovery of impaired forelimb function in unilateral 6-OHDA-lesioned rats has been shown to occur after very specific interventions, such as DBS (Shi et al., 2004), lentivector delivery of GDNF to the striatum and SN (Dowd et al., 2005), or through forced use of the impaired forelimb (Tillerson et al., 2001), likely as a result of increased striatal GDNF (Cohen et al., 2003), interventions that either bypass the striatal D A deficit, and act to inhibit the overactive basal ganglia output structures to normalize motor cortex activity, or provide strong DAergic trophic support to the SN and striatum. We did, however, see a significant improvement in TB test scores after repeated ECS treatment. The TB test however measures different aspects of locomotion and a lack of performance errors as opposed to an increase in use/performance as the cylinder test. Because the rat never has the choice to "ignore" its impaired hindlimb, it is likely that the degree of asymmetry between hindlimbs remains milder than for specialized movements of the forelimbs. Thus, a smaller contralateral improvement may be more likely to improve scores on this test. We have shown the relationship between performance on the task and the integrity of the striatal D A system (Chapter 4), and have used this task to measure improved motor performance after striatal cell transplantation in the unilateral 6-OHDA-lesioned rat (Cepeda et al., 2006). In addition, in a genetic mouse model of PD, L-DOPA treatment significantly improves performance on an adaptation of the test for mice (Hwang et al., 2005). The task is also widely used in models of stroke, where it has been shown to be sensitive to ischemic brain injury (Zhao et al., 2005a; Zhao et al., 2005b). In short, the TB test is a simple and valid test of gross motor function, and we have shown that repeated ECS treatment in unilateral 6-OHDA-lesioned rats improves performance on this task. 92 In this study, we have also shown that repeated ECS treatment in unilateral 6-OHDA-lesioned rats increases binding to specific D A receptor subtypes in both the dorsal striatum and NAcc shell. Binding to the Di and D 3 receptors was increased in both the lesioned and non-lesioned hemispheres after repeated ECS treatment (although the increase in Di binding was much smaller in the unlesioned versus lesioned striatum), whereas D2 receptor binding was unchanged by ECS treatment. There were also specific effects of the 6-OHDA lesion on D 2 and D 3 receptor binding, with D 2 binding increased in the lesioned dorsal striatum, and D 3 binding decreased in the lesioned NAcc shell. D 2 receptor upregulation has been widely reported early after D A depletion in rodents (Graham et al., 1990; Narang and Wamsley, 1995), non-human primates (Doudet et al., 2000), and PD patients (Kaasinen et al., 2000), and the 6-OHDA-induced decrease in D 3 binding in the NAcc is also widely recognized (Levesque et al., 1995; Stanwood et al., 2000; van Kampen and Stoessl, 2003). Our results on the effects of repeated ECS treatment on D A receptor binding in 6-OHDA-lesioned rats are consistent with the previous literature in normal rodents. Using both homogenate and autoradiographic receptor binding techniques, upregulation of Di receptors is a common finding in the normal striatum after a course of ECS (Fochtmann et al., 1989; Nowak and Zak, 1989; Barkai et al., 1990), while D 2 receptors are typically unchanged in the dorsal striatum (Bergstrom and Kellar, 1979; Reches et al., 1984; Martin et al., 1995), and D 2 and D 3 receptors have been reported to be upregulated in the NAcc (Barkai et al., 1990; Lammers et al., 2000). Our observations, then, of increased Di and D 3 binding, without concomitant changes in D 2 binding after repeated ECS treatment in 6-OHDA-lesioned rats are in good agreement with literature on the effects of repeated ECS on these receptors in normal animals. The D 3 receptor is most abundant in the Islands of Calleja and NAcc, and is expressed at very low levels in the dorsal striatum and the rest of the rat brain under normal circumstances (Levesque et al., 1992; Diaz et al., 1995). Chronic treatment with L-DOPA in unilateral 6-OHDA-lesioned rats leads to behavioural sensitization (the rodent homolog of L-DOPA-induced dyskinesia) and a dramatic increase in the expression of the D 3 receptor in the lesioned dorsal striatum (Bordet et a l , 1997; van Kampen and Stoessl, 2003). It appears that three conditions must be met to increase the dorsal striatal expression of the D 3 receptor in the rat brain after chronic pulsatile L-DOPA treatment: 1) severe 6-OHDA depletion of striatal DA, 2) activation of the Di receptor, and 3) elevated brain-derived neurotrophic factor (BDNF) levels (Bordet et al., 1997; Guillin et al., 2001; van Kampen and Stoessl, 2003). Although the increased D 3 receptor binding that we observed in the dorsal striatum after repeated ECS treatment was less pronounced than in L-DOPA-induced behavioural sensitization (12-15% in this study, vs. 130-680%; Bordet et a l , 1997; van Kampen and Stoessl, 2003), it may follow a similar mechanism. There is strong evidence that BDNF activity is increased after repeated ECS treatment, not only in the hippocampus (Nibuya et al., 1995; Altar et al., 2003; Jacobsen and Mork, 2004), but also in the striatum of both normal (Angelucci et al., 2002) and 6-OHDA-lesioned rats (Strome et al., 2006b). BDNF expression is regulated by cyclic adenosine monophosphate (cAMP) response element-binding protein (CREB) (Shieh et al., 1998), and Di receptor activation leads to phosphorylation of CREB and the transcription of BDNF (Fang et al., 2003). CREB itself is upregulated in the hippocampus by repeated ECS treatment (Nibuya et al., 1996). Taken together, it seems reasonable to hypothesize that the ECS-induced increase in the Dj receptor in the striatum could lead to increased CREB activity, leading to the transcription of BDNF, and increased expression of the D 3 receptor. The fact that the D 3 receptor was only moderately upregulated by repeated ECS treatment compared to behavioural sensitization to L-DOPA treatment may, in fact, be advantageous. In 93 the rat, the induction of the D 3 receptor in the dorsal striatum after chronic L-DOPA treatment occurs mainly in dynorphin/substance P (and Di) expressing neurons of the direct striatonigral pathway (Bordet et al., 2000). Co-expression of Di and D 3 receptors has been shown to have both opposite and synergistic effects on cAMP and on gene expression (Ridray et al., 1998). When the two receptor subtypes are in a synergistic relationship, the relative abundance of the receptors may dictate the functional outcome. For example, synergy between Di and D 3 occurs in L-DOPA-induced behavioural sensitization, but in this case, the D 3 receptor is expressed at high levels, leading to overactivity of the direct pathway of the basal ganglia, and the development of sensitization (Bordet et a l , 1997; Bordet et a l , 2000). If D 3 receptors are expressed in a low to moderate ratio compared to the Di receptor, however, the synergy between them may enhance the activity in the direct pathway, without causing excessive stimulation. The nature of the synergistic relationship between Di and D 3 therefore may depend on the relative expression of the two receptor subtypes, with moderate levels of D 3 being advantageous, and high levels detrimental. If ECT treatment enhances D 3 expression only moderately, then activity in the direct pathway will be enhanced, but not excessive. In conclusion, this is the first study to show improvements in non-pharmacological motor performance and increases in specific D A receptor subtypes after repeated ECS-treatment in 6-OHDA-lesioned rats. ECT is a non-invasive and safe treatment, and is widely used to treat psychiatric disorders. The clinical evidence suggests that in some patients, ECT can provide almost immediate and fairly long-lasting relief of the motor symptoms of PD. ECT should be considered in PD patients with poor response to medication, prior to surgical intervention in patients with severe motor symptoms, and, given its potential neurotrophic effects, perhaps even in patients early in the course of the disease. While the mechanism of action is not completely known, and further research is necessary, this study increases our understanding of the effects of ECT on the brain, and provides support for the continued use and study of ECT as a potential adjunct treatment for PD. 94 References: Altar CA, Whitehead RE, Chen R, Wortwein G, Madsen T M (2003) Effects of electroconvulsive seizures and antidepressant drugs on brain-derived neurotrophic factor protein in rat brain. Biol Psychiatry 54: 703-709. Andersen K, Balldin J, Gottfries CG, Granerus A K , Modigh K , Svennerholm L, et al. (1987) A double-blind evaluation of electroconvulsive therapy in Parkinson's disease with "on-off phenomena. Acta Neurol Scand 76: 191-199. Angelucci F, Aloe L , Jimenez-Vasquez P, Mathe A A (2002) Electroconvulsive stimuli alter the regional concentrations of nerve growth factor, brain-derived neurotrophic factor, and glial cell line-derived neurotrophic factor in adult rat brain. J ECT 18: 138-143. Barkai AI, Durkin M , Nelson HD (1990) Localized alterations of dopamine receptor binding in rat brain by repeated electroconvulsive shock: an autoradiographic study. Brain Res 529: 208-213. Bergstrom DA, Kellar K J (1979) Effect of electroconvulsive shock on monoaminergic receptor binding sites in rat brain. Nature 278: 464-466. Bernheimer H, Birkmayer W, Hornykiewicz O, Jellinger K , Seitelberger F (1973) Brain dopamine and the syndromes of Parkinson and Huntington. Clinical, morphological and neurochemical correlations. J Neurol Sci 20: 415-455. Bordet R, Ridray S, Carboni S, Diaz J, Sokoloff P, Schwartz JC (1997) Induction of dopamine D 3 receptor expression as a mechanism of behavioral sensitization to levodopa. Proc Natl Acad Sci U S A 94: 3363-3367. Bordet R, Ridray S, Schwartz JC, Sokoloff P (2000) Involvement of the direct striatonigral pathway in levodopa-induced sensitization in 6-hydroxydopamine-lesioned rats. Eur J Neurosci 12: 2117-2123. Butler MO, Morinobu S, Duman RS (1993) Chronic electroconvulsive seizures increase the expression of serotonin2 receptor mRNA in rat frontal cortex. J Neurochem 61: 1270-1276. Cepeda IL, Cornfeldt M L , Doudet DJ (2006) Human retinal pigment epithelial cell implants ameliorate motor deficits in two rat models of Parkinson's disease. Submitted. Cohen A D , Tillerson JL, Smith A D , Schallert T, Zigmond M J (2003) Neuroprotective effects of prior limb use in 6-hydroxydopamine-treated rats: possible role of GDNF. J Neurochem 85: 299-305. DaSilva JN, Carey JE, Sherman PS, Pisani TJ, Kilbourn M R (1994) Characterization of [nC]tetrabenazine as an in vivo radioligand for the vesicular monoamine transporter. Nucl Med Biol 21: 151-156. 95 Diaz J, Levesque D, Lammers C H , Griffon N , Martres MP, Schwartz JC, et al. (1995) Phenotypical characterization of neurons expressing the dopamine D 3 receptor in the rat brain. Neuroscience 65: 731-745. Doudet DJ, Holden JE, Jivan S, McGeer E, Wyatt RJ (2000) In vivo PET studies of the dopamine D 2 receptors in rhesus monkeys with long-term MPTP-induced parkinsonism. Synapse 38: 1 OS-I D . Dowd E, Monville C, Torres E M , Wong LF, Azzouz M , Mazarakis ND, et al. (2005) Lentivector-mediated delivery of GDNF protects complex motor functions relevant to human Parkinsonism in a rat lesion model. Eur J Neurosci 22: 2587-2595. Duman RS, Heninger GR, Nestler EJ (1997) A molecular and cellular theory of depression. Arch Gen Psychiatry 54: 597-606. Dunnett SB, Whishaw IQ, Rogers DC, Jones G H (1987) Dopamine-rich grafts ameliorate whole body motor asymmetry and sensory neglect but not independent limb use in rats with 6-hydroxydopamine lesions. Brain Res 415: 63-78. Evenden JL, Robbins TW (1984) Effects of unilateral 6-hydroxydopamine lesions of the caudate putamen on skilled forepaw use in the rat. Behav Brain Res 14: 61-68. Fall PA, Ekman R, Granerus A K , Thorell L H , Walinder J (1995) ECT in Parkinson's disease. Changes in motor symptoms, monoamine metabolites and neuropeptides. J Neural Transm Park Dis Dement Sect 10: 129-140. Fang H, Chartier J, Sodja C, Desbois A , Ribecco-Lutkiewicz M , Walker PR, et al. (2003) Transcriptional Activation of the Human Brain-derived Neurotrophic Factor Gene Promoter III by Dopamine Signaling in NT2/N Neurons. J Biol Chem 278: 26401-26409. Fochtmann LJ, Cruciani R, Aiso M , Potter WZ (1989) Chronic electroconvulsive shock increases D - l receptor binding in rat substantia nigra. Eur J Pharmacol 167: 305-306. Goodwin G M , Green A R , Johnson P (1984) 5-HT2 receptor characteristics in frontal cortex and 5-HT2 receptor-mediated head-twitch behaviour following antidepressant treatment to mice. Br J Pharmacol 83: 235-242. Graham WC, Crossman AR, Woodruff G N (1990) Autoradiographic studies in animal models of hemi-parkinsonism reveal dopamine D 2 but not Di receptor supersensitivity. I. 6-OHDA lesions of ascending mesencephalic dopaminergic pathways in the rat. Brain Res 514: 93-102. Green AR, Heal DJ, Grahame-Smith D G (1977) Further observations on the effect of repeated electroconvulsive shock on the behavioural responses of rats produced by increases in the functional ability of brain 5-hydroxytryptamine and dopamine. Psychopharmacology 52: 195-200. 96 Green AR, Heal DJ, Johnson P, Laurence BE, Nimgaonkar V (1983a) Antidepressant treatments: effects in rodents on dose-response curves of 5-hydroxytryptamine- and dopamine-mediated behaviours and 5 - H T 2 receptor number in frontal cortex. Br J Pharmacol 80: 377-385. Green AR, Johnson P, Nimgaonkar V (1983b) Increased 5 - H T 2 receptor number in brain as a probable explanation for the enhanced 5-hydroxytryptamine-mediated behaviour following repeated electroconvulsive shock administration to rats. Br J Pharmacol 80: 173-177. Guillin O, Diaz J, Carrol P, Griffon N , Schwartz JC, Sokoloff P (2001) BDNF controls dopamine D 3 receptor expression and triggers behavioural sensitization. Nature 411: 86-89. Hwang DY, Fleming S M , Ardayfio P, Moran-Gates T, K im H, Tarazi F l , et al. (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. Jacobsen JP, Mork A (2004) The effect of escitalopram, desipramine, electroconvulsive seizures and lithium on brain-derived neurotrophic factor mRNA and protein expression in the rat brain and the correlation to 5-HT and 5-HIAA levels. Brain Res 1024: 183-192. Jankovic J (2005) Motor fluctuations and dyskinesias in Parkinson's disease: clinical manifestations. Mov Disord 20 Suppl 11: SI 1-S16. Kaasinen V, Ruottinen H M , Nagren K , Lehikoinen P, Oikonen V, Rinne JO (2000) Upregulation of putaminal dopamine D 2 receptors in early Parkinson's disease: a comparative PET study with [ n C] raclopride and [nC]N-methylspiperone. J Nucl Med 41: 65-70. Kennedy R, Mittal D, O'Jile J (2003) Electroconvulsive therapy in movement disorders: an update. J Neuropsychiatry Clin Neurosci 15: 407-421. Kho K H , van Vreeswijk M F , Simpson S, Zwinderman A H (2003) A meta-analysis of electroconvulsive therapy efficacy in depression. J ECT 19: 139-147. Lammers C H , Diaz J, Schwartz JC, Sokoloff P (2000) Selective increase of dopamine D 3 receptor gene expression as a common effect of chronic antidepressant treatments. Mol Psychiatry 5(4): 378-388. Levesque D, Diaz J, Pilon C, Martres MP, Giros B, Souil E, et al. (1992) Identification, characterization, and localization of the dopamine D 3 receptor in rat brain using 7-[3H]hydroxy-N,N-di-n-propyl-2-aminotetralin. Proc Natl Acad Sci U S A 89: 8155-8159. Levesque D, Martres MP, Diaz J, Griffon N , Lammers C H , Sokoloff P, et al. (1995) A paradoxical regulation of the dopamine D 3 receptor expression suggests the involvement of an anterograde factor from dopamine neurons. Proc Natl Acad Sci U S A 92: 1719-1723. Liberatore GT, Wong JYF, Krenus D, Jeffreys BJ, Porritt M , Howells D W (1999) Tissue-fixation prevents contamination of tritium-sensitive storage phosphor imaging plates. BioTechniques 26: 432-434. 97 Malberg J, Eisch A J , Nestler EJ, Duman RS (2000) Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 20: 9104-9110. Martin KF, Phillips I, Cheetham SC, Heal DJ (1995) Dopamine D 2 receptors: a potential pharmacological target for nomifensine and tranylcypromine but not other antidepressant treatments. Pharmacol Biochem Behav 51: 565-569. Moorman JM, Grahame-Smith DG, Smith SE, Leslie R (1996) Chronic electroconvulsive shock enhances 5-HT 2 receptor-mediated head shakes but not brain C-fos induction. Neuropharmacology 35: 303-313. Miklyaeva EI, Castaneda E, Whishaw IQ (1994) Skilled reaching deficits in unilateral dopamine-depleted rats: impairments in movement and posture and compensatory adjustments. J Neurosci 14:7148-7158. Narang N , Wamsley JK (1995) Time dependent changes in D A uptake sites, Di and D 2 receptor binding and mRNA after 6-OHDA lesions of the medial forebrain bundle in the rat brain. J Chem Neuroanat 9: 41-53. Nibuya M , Morinobu S, Duman RS (1995) Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J Neurosci 15: 7539-7547. Nibuya M , Nestler EJ, Duman RS (1996) Chronic antidepressant administration increases the expression of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci 16: 2365-2372. Nowak G, Zak J (1989) Repeated electroconvulsive shock (ECS) enhances striatal D - l dopamine receptor turnover in rats. Eur J Pharmacol 167: 307-308. Pagnin D, de Q, V , Pini S, Cassano GB (2004) Efficacy of ECT in depression: a meta-analytic review. J ECT 20: 13-20. Paxinos G, Watson C (1997) The Rat Brain in Stereotaxic Coordinates. San Diego: Academic Press. Pridmore S, Pollard C (1996) Electroconvulsive therapy in Parkinson's disease: 30 month follow up. J Neurol Neurosurg Psychiatry 61: 693-700. Reches A , Wagner HR, Barkai AI, Jackson V , Yablonskaya-Alter E, Fahn S (1984) Electroconvulsive treatment and haloperidol: effects on pre- and postsynaptic dopamine receptors in rat brain. Psychopharmacology 83: 155-158. Ridray S, Griffon N , Mignon V , Souil E, Carboni S, Diaz J, et al. (1998) Coexpression of dopamine D\ and D 3 receptors in islands of Calleja and shell of nucleus accumbens of the rat: opposite and synergistic functional interactions. Eur J Neurosci 10: 1676-1686. 98 Schallert T, Fleming S M , Leasure JL, Tillerson JL, Bland ST (2000) CNS plasticity and assessment of forelimb sensorimotor outcome in unilateral rat models of stroke, cortical ablation, parkinsonism and spinal cord injury. Neuropharmacology 39: 777-787. Schallert T, Woodlee M T (2005) Orienting and Placing. In: The behavior of the laboratory rat: a handbook with tests. (Whishaw IQ, Kolb B, eds), pp 129-140. New York: Oxford University Press. Shi L - H , Woodward DJ, Luo F, Anstrom K, Schallert T, Chang J-Y (2004) High-frequency stimulation of the subthalamic nucleus reverses limb-use asymmetry in rats with unilateral 6-hydroxydopamine lesions. Brain Res 1013: 98-106. Shieh PB, Hu SC, Bobb K, Timmusk T, Ghosh A (1998) Identification of a signaling pathway involved in calcium regulation of BDNF expression. Neuron 20: 727-740. Smith SE, Sharp T (1997) Evidence that the enhancement of dopamine function by repeated electroconvulsive shock requires concomitant activation of Di-like and D2-like dopamine receptors. Psychopharmacology 133: 77-84. Stanwood GD, Artymyshyn RP, Kung MP, Kung HF, Lucki I, McGonigle P (2000) Quantitative autoradiographic mapping of rat brain dopamine D 3 binding with [ 1 2 5I]7-OH-PIPAT: evidence for the presence of D 3 receptors on dopaminergic and nondopaminergic cell bodies and terminals. J Pharmacol Exp Ther 295: 1223-1231. Stern M B (1991) Electroconvulsive therapy in untreated Parkinson's disease. Mov Disord 6: 265. Strome E M , Jivan S, Doudet DJ (2005) Quantitative in vitro phosphor imaging using [3H] and [1 8F] radioligands: the effects of chronic desipramine treatment on serotonin 5 - H T 2 receptors. J Neurosci Methods 141: 143-154. Strome E M , Cepeda IL, Sossi S, Doudet DJ (2006a) Evaluation of the integrity of the dopamine system in a rodent model of Parkinson's disease: small animal PET compared to behavioural assessment and autoradiography. Submitted. Strome E M , Zis AP , Doudet DJ (2006b) The effects of electroconvulsive shock on GDNF, BDNF, and FGF-2 concentrations in the 6-OHDA-lesioned rat brain. Submitted. Tillerson JL, Cohen A D , Philhower J, Miller G, Zigmond MJ , Schallert T (2001) Forced limb-use effects on the behavioral and neurochemical effects of 6-hydroxydopamine. J Neurosci 21: 4427-4435. Vaidya V A , Siuciak JA, Du F, Duman RS (1999) Hippocampal mossy fiber sprouting induced by chronic electroconvulsive seizures. Neuroscience 89: 157-166. van Kampen JM, Stoessl A J (2003) Effects of oligonucleotide antisense to dopamine D 3 receptor mRNA in a rodent model of behavioural sensitization to levodopa. Neuroscience 116: 307-314. 99 Wielosz M (1981) Increased sensitivity to dopaminergic agonists after repeated electroconvulsive shock (ECS) in rats. Neuropharmacology 20: 941-945. Yoshida K, Higuchi H , Kamata M , Yoshimoto M , Shimizu T, Hishikawa Y (1998) Single and repeated electroconvulsive shocks activate dopaminergic and 5-hydroxytryptaminergic neurotransmission in the frontal cortex of rats. Prog Neuropsychopharmacol Biol Psychiatry 22: 435-444. Zar JH (1999) Biostatistical Analysis. Upper Saddle River: Prentice-Hall. Zervas IM, Fink M (1991) ECT for refractory Parkinson's disease. Convulsive Ther 7: 222-223. Zhao CS, Puurunen K , Schallert T, Sivenius J, Jolkkonen J (2005a) Behavioral and histological effects of chronic antipsychotic and antidepressant drug treatment in aged rats with focal ischemic brain injury. Behav Brain Res 158: 211-220. Zhao CS, Puurunen K, Schallert T, Sivenius J, Jolkkonen J (2005b) Effect of cholinergic medication, before and after focal photothrombotic ischemic cortical injury, on histological and functional outcome in aged and young adult rats. Behav Brain Res 156: 85-94. 100 Chapter 6 THE EFFECTS OF ELECTROCONVULSIVE SHOCK ON GDNF, BDNF, AND FGF-2 CONCENTRATIONS IN THE 6-OHDA-LESIONED RAT BRAIN 1 Preamble: In the previous chapter, we addressed two key aspects of our working hypothesis (Figure 1.1, page 10); the effects of repeated ECS treatment in unilateral 6-OHDA-lesioned rats on 1) motor function, and 2) D A receptors. In this chapter, we address another key feature of our working hypothesis, Hypothesis 3, that repeated ECS treatment in 6-OHDA-lesioned rats enhances striatal protein levels of the neurotrophic factors GDNF, BDNF, and FGF-2. We originally performed a small pilot study looking at the effects of repeated ECS on BDNF expression using in situ hybridization, and although we replicated the previously reported ECS-induced increase in piriform cortex BDNF mRNA, we were unable to detect BDNF mRNA within the striatum (see Appendix B for more detail). We therefore decided to move away from mRNA, and instead investigate neurotrophic factor protein levels directly in our regions of interest. Introduction: Parkinson's disease (PD) is a progressive and debilitating neurodegenerative disorder affecting millions of adults worldwide. The hallmark neuropathological finding is the degeneration of midbrain dopamine (DA)-producing neurons, the loss of which leads to the characteristic motor symptoms of bradykinesia, tremor, rigidity and postural imbalance (Hornykiewicz, 1972). Various drug therapies exist, but these can cause debilitating side effects and/or lose their effectiveness after extended use (Jankovic, 2005). Surgical interventions, such as targeted lesions of the basal ganglia output structures that become overactive in PD (pallidotomy or subthalamotomy), or the implantation of stimulating electrodes to overexcite those same brain regions, and thereby inhibit their activity (deep brain stimulation), are becoming more common in advanced patients, but these are invasive treatments and carry inherent risks. As such, there is currently a need for the development of alternative treatments for PD. One of the most effective and widely used treatments for depression, electroconvulsive therapy (ECT), appears to also improve the motor symptoms of PD patients, regardless of whether or not they are depressed. While the majority of reports on this phenomenon in the literature have been case reports, open trials, or retrospective studies, ECT treatment has consistently been shown to improve the main motor symptoms of PD, including rigidity and bradykinesia, and the improvements last from several weeks to months (Stern, 1991; Zervas and Fink, 1991; Fall et al., 1995; Pridmore and Pollard, 1996). To date, only one placebo-controlled double-blind trial on the effects of ECT in PD has been performed, but it showed that patients treated with ECT had prolonged "on" phases and decreased severity of their motor symptoms (Andersen e ta l , 1987). ECT has mainly been investigated as an antidepressant treatment, and there is a large body of literature showing that it has very specific effects in cortical and hippocampal areas, limbic brain regions that are involved in mood disorders. The most consistent effects are: 1) 1 A version of this manuscript has been submitted for publication. Strome EM, Zis AP, Doudet DJ (2006) The effects of electroconvulsive shock on GDNF, BDNF, and FGF-2 concentrations in the 6-OHDA-lesioned brain. 101 Increased activity in the serotonin and D A systems via changes in the concentration of these neurotransmitters and/or their receptors (Nowak and Zak, 1989; Zis et al., 1992; Yoshida et al., 1998a; Yoshida et al., 1998b; Burnet et al., 1999); 2) Increased neurotrophic factor levels (Nibuya et al., 1995; Angelucci et al., 2002; Gwinn et al., 2002); and 3) Enhanced neuronal growth (through either neuronal sprouting or neurogenesis; Gombos et al., 1999; Vaidya et al., 1999; Malberg et a l , 2000; Scott et al., 2000; Lamont et al., 2001). We hypothesize that these effects of electroconvulsive shock (ECS - the animal model of ECT) are not limited to the limbic system, and that they may also occur in other brain regions, including areas involved in motor control. Although these effects may not lead to noticeable behavioural changes in subjects with an intact nigrostriatal system, they may account for some of the clinical improvements seen after ECT in parkinsonian patients. More specifically, in this study, we hypothesize that increased neurotrophic factor activity in the striatum may partially underlie the positive effects of ECT on the motor symptoms of PD. To investigate this hypothesis, we studied unilaterally 6-hydroxydopamine (6-OHDA)-lesioned rats after repeated ECS or sham treatments. We then examined protein levels of glial cell line-derived neurotrophic factor (GDNF), brain-derived neurotrophic factor (BDNF), and basic fibroblast growth factor (FGF-2), three neurotrophic factors that are reduced in the brains of PD patients (Tooyama et al., 1993.; Mogi et al., 1999; Chauhan et a l , 2001) and that are all protective against striatal D A loss in animal models of PD (Altar et al., 1994; Tomac et al., 1995; Shults et al., 1995; Bowenkamp et al., 1997; Shults et a l , 2000). We chose to evaluate trophic factor levels in two limbic regions, the hippocampus and prefrontal cortex (PFC) as a control for the effectiveness of our ECS and immunohistochemical techniques, and in the striatum, the major target of the D A neurons that degenerate in PD, using the sensitive and specific immunohistochemical technique enzyme-linked immunosorbent assay (ELISA). Materials and Methods: Subjects Adult male Sprague-Dawley rats [bred at the University of British Columbia animal facility from Charles River Canada (Montreal, PQ) stock], weighing 250 g at the start of the experiment were housed on a 12:12 light:dark schedule (with lights off at 12:00 p.m.), at constant temperature and humidity (21°C, 55%). Initial pilot studies were performed in 10 animals (n = 5 per group), and an additional 20 animals were added to the study later (n = 10 per group). Of the 30 animals, 3 were excluded due to poor lesions, 6 were excluded because they did not show consistent seizure activity, and 3 were lost to ketamine anaesthesia. Eighteen animals completed the study (n = 10 sham, 8 ECS). The animals had access to food and water ad libitum, and were housed in pairs. A l l procedures were approved by the University of British Columbia Committee on Animal Care. 6-OHDA Lesioning After at least 5 days of habituation and handling, allowing them to reach a weight of at least 270 g, the rats received a right unilateral 6-OHDA-induced lesion of the D A nigrostriatal pathway. Desipramine hydrochloride (25 mg/kg i.p.; Sigma-Aldrich Canada, Oakville, ON) was administered 30-60 min before 6-OHDA infusion to protect noradrenergic terminals. Animals were anaesthetized with isoflurane in 0 2 (4% for induction, 1% for maintenance), given atropine sulfate (0.05 mg/kg s.e), and placed into a stereotaxic frame (Kopf). With the skull flat between lambda and Bregma, a 2% solution of 6-OHDA hydrobromide (8 jag in 4 uL 0.05% ascorbic acid in saline; Sigma) was infused at two sties along the medial forebrain bundle [site 1: AP -2.8 mm, M L -1.8 mm, D V -8.0 (all from Bregma); site 2: AP -4.7 mm (Bregma), M L -1.5 mm (midline), D V -7.9 mm (hole) according to (Paxinos and Watson, 1997)]. The infusion rate was 102 1 uL per min, and the cannula was left in place an additional 4 min to allow diffusion of the 6-OHDA solution. After surgery, the animals received subcutaneous saline, antibiotics (Duplocillin 0.1 ml/kg i.m.), and analgesia (Anafen 2 mg/kg s.e), and were kept warm in an incubator until fully recovered from anaesthesia. Animals were allowed to recover for at least 2 weeks after surgery before being treated with ECS. Cylinder Test Animals were evaluated in the forelimb use asymmetry test (Cylinder Test; Schallert et al., 2000; Schallert and Woodlee, 2005) at three time points: before lesion, after lesion, and after ECS or sham treatment. At each time point, the animals were placed in a Plexiglas cylinder (20 cm diameter x 30 cm high) elevated on a glass plate for a 3 min period on two consecutive days. Testing was done during the dark phase of the cycle and under red lighting. The trials were videotaped from below, and scored at a later date by an investigator blind to the animals' treatment. Forelimb placements on the walls of the cylinder were categorized as left independent forelimb use, right independent forelimb use, or simultaneous use of both forelimbs, and a forelimb use asymmetry score was calculated as: Equation 6.1: [(ipsi + lA both) divided by (ipsi + contra + both)] X 100 (Schallert and Woodlee, 2005) where ipsi and contra refer to the forelimbs ipsilateral and contralateral to the 6-OHDA-induced lesion respectively. Animals showing less than 70% asymmetry after lesion were excluded from further analysis (2 animals in this study). Electroconvulsive Shock Treatment Animals were assigned randomly to the ECS or sham treatment groups, and were treated every day for 10 days between 08:00 and 11:00 a.m. Atropine sulfate (0.2 mg/kg s.e.) was administered, followed 30 min later by ketamine hydrochloride (80 mg/kg i.p.). After induction of ketamine-induced anaesthesia, animals were given either sham treatment (electrodes placed, but no current administered), or bilateral ECS (80-99 mA, 5-9.9 s, 70 pulse/s, 0.5 ms pulse width) via earclip electrodes coated with electroconductive gel using a small animal ECS machine (Model 57800, Ugo Basile, Italy). A l l animals received the same initial current dose, based on our previous experience with rats receiving ECS under ketamine anaesthesia, and current doses during subsequent treatments were modified based on the nature of the previous seizure. Mean seizure length was 15 ± 1 s (mean ± SEM), and did not change significantly over the course of ECS treatment (data not shown). Any animal that did not consistently show tonic hind limb extension after ECS treatment was excluded from the analysis (6 animals). Tissue Processing Forty-eight hours after the last ECS treatment, and immediately following the last behavioural session, animals were sacrificed by decapitation. Their brains were quickly removed and dissected on ice. The hemispheres were divided and the left and right PFC (anterior to approximately +3.0 mm from Bregma, but excluding the olfactory bulb and tract; Paxinos and Watson, 1997), hippocampus and striatum were stored individually in cold, labelled, pre-weighed microtubes. The tubes were immediately re-weighed, and stored at -80°C until further processing with ELISA. On the day of the ELISA, the microtubes were removed from the freezer and thawed on ice. The tissues were diluted lOOx w/v in ice-cold lysis buffer with freshly added protease inhibitors [100 m M Tris-HCl, pH 7.2, 400 m M NaCl, 4 m M EDTA, 0.2 m M 103 phenylmethylsulfonyl fluoride, 0.2 mM benzethonium chloride, 0.05% sodium azide, 2% bovine serum albumen, 2 m M benzamidine, 40 U/ml aprotinin (Calbiochem, La Jolla, CA), 0.2% Triton X-100 (all Sigma unless otherwise noted)], and homogenized on ice with ultrasonication (Sonic Dismembrator, 3x 5 s pulses @ 20 s intervals, power level 3; Fisher Scientific, Ottawa, ON). The tubes were then centrifuged at 10 000 g for 20 min at 4°C. The supernatants were removed, and total protein concentration in the supernatants was determined by a commercially available detergent-compatible kit (DC Protein Assay; Bio-Rad, Hercules, CA). Samples were further diluted 1:1 v/v in lysis buffer, kept on ice, and immediately used for ELISA. GDNF and BDNF ELISA A l l ELISAs were performed in 96-well microplates (Nunc-Immuno MaxiSorp Plates; Nalge-Nunc International, Rochester, NY) . GDNF and BDNF concentrations were determined following the directions in the appropriate Emax ELISA kit from Promega (Madison, WI). Briefly, the microplates were coated overnight (at 4°C) with the appropriate capture antibody diluted in carbonate coating buffer (pH 8.2 GDNF, pH 9.7 BDNF). On the second day, the plates were blocked, and diluted tissue samples were added to the plate in triplicate, and incubated at room temperature with shaking (6 h GDNF, 2 h BDNF), allowing the antigen (neurotrophic factor) to bind to the capture antibody. Following washing, the plates had an Anti-Human polyclonal antibody added, were incubated (16 h at 4°C GDNF, 2 h at room temperature BDNF), washed again, and then incubated with shaking at room temperature with a secondary anti-Ig-Y antibody conjugated to horseradish peroxidase (2 h GDNF, 1 h BDNF). Finally, the substrate (hydrogen peroxide plus tetramethylbenzidine solution, room temperature) was added to produce the color reaction. After 15 (GDNF) or 10 (BDNF) min, the reaction was stopped with 1 N hydrochloric acid, and the plates were read immediately in a microplate reader at 450 nm (Bio-Tek Instruments Inc., Winooski, VT). Standard curves were created for each microplate (range 1000 - 15.6 pg/mL), based on pilot studies and the literature (Angelucci et al., 2002). FGF-2 ELISA For the determination of FGF-2 concentrations, an antibody pair from R&D Systems Inc. (Human FGF basic Duo-Set; Minneapolis, MN) was used, following the manufacturer's instructions. Briefly, the microplates were coated with capture antibody diluted in PBS buffer, and incubated overnight at room temperature. The following day, plates were washed and blocked with PBS buffer containing 1% bovine serum albumen (Sigma). Following washing, the tissue samples were added in triplicate and incubated for 2 h at room temperature. The plates were washed again, the polyclonal detection antibody was added, and allowed to incubate for 2 h at room temperature, followed by plate washing. Streptavidin-horseradish peroxidase was added to the plate and incubated for 20 min. Following washing, the substrate solution (hydrogen peroxide plus tetramethylbenzidine solution, room temperature; R & D Systems Inc.) was added, and the color reaction was stopped after 20 min with 2 N sulphuric acid (R&D Systems Inc.). The plates were read immediately at 450 nm. A standard curve was included with each plate (4000-62.5 pg/mL). Data Analysis Cylinder Test data were analyzed with repeated measures analysis of variance (ANOVA; treatment x time) and the software program Statistica '98 (StatSoft, Tulsa, OK). The best 2 out of 3 optical densities for each sample were used for further analysis (the point furthest from the mean was discarded). Some assays for some animals showed low reproducibility and the data were discarded, so although 18 animals completed the study, the N 104 values for each assay are 7-10 per group for the BDNF and FGF-2 ELISA. The GDNF ELISA was only performed on the pilot group of animals (n = 10) because the results were clear even with a small number of subjects. Because many of the data points for the GDNF ELISA were below the standard curve, we were unable to fully quantify the data. Instead, because all the samples from a particular brain region were analyzed on the same microplate, the optical density values (corrected for total protein content), were compared, and the data are presented as mean ± S E M percent of control. Fully quantified data for BDNF and FGF-2 were corrected for total protein content, and are presented as mean ± S E M percent of control (sham-treated animals). The effects of ECS on neurotrophic factor protein concentrations were evaluated using two-way A N O V A (treatment x hemisphere). Examination of all main effects was performed with post-hoc testing using Tukey's test for unequal N . R e s u l t s : Cylinder Test The mean asymmetry score for the entire group before lesioning (n = 18) was 52.17 ± 1.58 % (mean ± SEM). A l l animals except for two (and these were excluded from the study) showed at least 70% asymmetry in forelimb use when tested 2 weeks post-lesion (mean ± SEM = 83.72 ± 2.15%; Fig. 6.1). Two-way A N O V A indicates no significant difference between the two treatment groups, but a significant effect of time (F 2 j32 = 124.01, p < 0.0001), with post-hoc analysis showing the asymmetry scores are significantly lower before lesioning (p < 0.001), but no significant difference before or after treatment (Fig. 6.1). GDNF ELISA There was a significant effect of treatment on GDNF in the PFC ( F 1 J 4 = 12.97, p < 0.003), with post-hoc testing showing ECS treatment significantly decreased the amount of GDNF protein in the PFC (p < 0.004.; Fig. 6.2a). There was no effect of ECS in either the hippocampus or striatum (Fig. 6.2b, c). There was a significant effect of hemisphere in the striatum (F i j5 = 7.11, p < 0.02), with post-hoc testing indicating that the lesioned striatum had significantly increased GDNF in both the ECS and sham groups (p < 0.02; Fig. 6.2c). BDNF ELISA There was a significant effect of treatment on BDNF protein in all three brain regions examined (PFC: F131 = 4.94, p < 0.04, hippocampus: F 1 3 0 = 6.61, p < 0.02; striatum: F132 = 5.86, p < 0.03), with post-hoc testing showing a significant decrease in the PFC (p < 0.03), and significant increases in the hippocampus (p < 0.02) and striatum (p < 0.03) (Fig. 6.3). There was also a significant effect of hemisphere in the PFC (Fi^i = 5.20, p < 0.03), with post-hoc testing showing the lesioned hemisphere having lower BDNF (p < 0.03; Fig. 6.3a). FGF-2 ELISA There was no effect of treatment on FGF-2 levels in the PFC, and a slight, but not significant (p < 0.09) effect in the hippocampus (Fig. 6.4). There was a significant effect of treatment on FGF-2 levels in the striatum (F130 = 12.99, p < 0.002; Fig. 6.4c). Post-hoc testing shows ECS treatment significantly increased FGF-2 protein in the striatum (p < 0.002). 105 1 0 ( H Sham ECS (n = 10) (n = 8) Figure 6.1: Forelimb use asymmetry scores before lesioning, and before and 48 h after repeated ECS or sham treatment. A l l animals showed > 70% asymmetry after lesion, but there was no effect of treatment. (* pre-lesion scores are significantly lower than post-lesion or post-ECS, p< 0.001). 106 a) I50n b) 15 increases in the sham and ECS groups) was more substantial. The evidence for 6-OHDA-induced changes in GDNF, BDNF or FGF-2 protein is minimal, and the existing reports are conflicting. While our results are in good agreement with one study showing increased GDNF, but no changes in BDNF and FGF-2 in the lesioned striatum of young rats (Nakajima et al., 2001), they do not agree with another showing increased striatal BDNF and GDNF in young, but decreased striatal BDNF in old rats after 6-OHDA lesions (Yurek and Fletcher-Turner, 2001). Either of the lesion effects that we observed may be compensatory reactions to the unilateral 6-OHDA lesion, as the system attempts to maintain homeostasis in the face of severe D A depletion. For example, the substantial decrease in the PFC ipsilateral to the 6-OHDA lesion may have resulted from the fact that most striatal BDNF is anterogradely transported from the frontal cortex (Altar et al., 1997), and with a loss of negative feedback inhibition from striatal D A terminals, the release of BDNF from corticostriatal projections may have been upregulated. In recent years, various techniques have identified changes in neurotrophic factors as one of the most consistent and robust effects of repeated ECS treatment. Gene microarray studies have shown that, while the expression of many genes is altered by ECS, some specific pathways are affected. In particular, in the hippocampus and cerebral cortex, growth-promoting genes appear to be enhanced by ECS, including genes for angiogenesis, neurogenesis, immediate-early genes, BDNF and its downstream signaling molecules, and FGF-2 (Newton et al., 2003; Altar et a l , 2004). Indeed, one of the most widely reported effects of ECS is increased expression of BDNF (both mRNA and protein), particularly in the hippocampus, cortical regions, and the striatum (Nibuya et al., 1995; Zetterstrom et al., 1998; Angelucci et al., 2002; Altar et al., 2003; Jacobsen and Mork, 2004). Although BDNF has been the most thoroughly investigated neurotrophin with respect to the effects of ECS, mRNA and protein expression of FGF-2 have also been shown to increase in the cortex and hippocampus after minimal ECS (Follesa et al., 1994; Gwinn et al , 2002; Kondratyev et al., 2002). In contrast, while ECS treatment increases mRNA for the GDNF receptors GFRa-1 and GFRa-2 in the hippocampus and parietal cortex, mRNA for the protein itself appears to be unaffected by ECS treatment (Chen et al , 2001a). So far, only one study has looked at GDNF protein after ECS (Angelucci et al., 2002), and it showed significant decreases in both the striatum and hippocampus, but no changes in the frontal or occipital cortices. Although there are some technical differences between our study and most of the literature (e.g. we used ketamine anaesthesia for ECS and sacrificed the animals 48 hrs after the last ECS or sham treatment, and all of the studies cited above did not use anaesthesia and sacrificed the animals 24 hrs after the last treatment, our data on the effects of repeated ECS treatment in 6-OHDA-lesioned animals are generally consistent with the literature on the effects of ECS on neurotrophic factor expression in the normal rodent brain. Two observations, however, the significant decrease of BDNF in the PFC, and the effects of repeated ECS on GDNF protein in the brain regions examined in this study, have not been reported previously, and deserve further attention. I l l Our finding that repeated ECS decreased BDNF in the PFC was unexpected since the literature shows increases (Altar et al., 2003; Jacobsen and Mork, 2004) or no change (Angelucci et al., 2002) in frontal cortex BDNF protein after ECS treatment in normal rats. Whereas most previous studies looked at a large area of the frontal cortex, however, we examined neurotrophic factor concentrations in a very discrete region, the most anterior 2-3 mm of the cortex. We chose this specific region because in the rat, it receives fairly extensive D A innervation from the midbrain (Berger et al., 1991) and because neurons in this region also make direct projections to the striatum (Berendse et al., 1992). The significant decrease in BDNF in the PFC that we found in this study may arise from the fact that BDNF can be transported anterogradely from the PFC to the striatum (Altar et al., 1997). BDNF is released in an activity-dependent manner (Kojima et al., 2001; Balkowiec and Katz, 2002; Goggi et al., 2003), so the stimulation of the cortex by ECS treatment may have caused BDNF stores to be depleted in the PFC and released into the striatum, and our observation of a reciprocal change in BDNF concentrations in the PFC and striatum after repeated ECS treatment. The effects of ECS on GDNF concentrations in the brain have only been examined in one previous report in normal animals (Angelucci et al., 2002), and this study showed decreases in the hippocampus and striatum, and no changes in the frontal and occipital cortices. The fact that we saw a significant decrease in GDNF in the PFC after ECS treatment, while Angelucci and colleagues saw no change in the frontal cortex, may, again, be due to differences in the regions being investigated. The other discrepancies between these two reports may be explained by the nature of GDNF trafficking, and the fact that we examined GDNF levels 48 hrs after the last ECS treatment, while Angelucci and colleagues measured it at 24 hrs post-ECS. Studies with radioactive GDNF show that it redistributes quickly after intraventricular injection in the rat brain (within 1 hr) and that by 7 days post-injection, it accumulates in specific brain regions, including the substantia nigra (SN), the hypothalamus, the fimbria and the cerebellum (Lapchak et al., 1997). So while we saw decreases or no changes in our regions of interest, there may have been increases in GDNF in areas receiving input from these regions (such as the hypothalamus, cerebellum, fimbria, or SN). In this study, we were not able to measure neurotrophic factor concentrations in the SN because of the small size of this region. While we may have been able to dissect a whole midbrain region, in our animals one SN was lesioned while the other was intact, and it was not possible to clearly separate the two hemispheres at that level during our dissections. It will be very important in future studies to measure neurotrophic factor changes in the SN of parkinsonian animals after repeated ECS treatment. In recent years, evidence of the beneficial effects of trophic factors on neuroprotection and survival of D A neurons in cell culture, as well as in animal models, has led to their investigation as potential treatments PD. Several different delivery methods have been investigated, including gene delivery via viral vectors (Bjorklund et al., 2000; Wang et al., 2002; Eslamboli et a l , 2005), and direct infusion of the protein into the target area [e.g. intraputamenal infusion of GDNF (Gill et al., 2003; Patel et al., 2005; Slevin et al., 2005; Lang et al., 2006)]. However, this line of investigation faces significant barriers because of the large size of the proteins, their inability to cross the blood-brain-barrier, and hence the necessary invasiveness of the treatment. Currently, safety concerns are prohibiting the use of viral vectors in human patients, and one Phase II GDNF direct infusion study was halted due to a lack of statistically significant improvement at the dose used and the appearance of neutralizing antibodies in several patients (Lang et al., 2006). While the therapeutic potential of neurotrophic factors in treating PD and other diseases remains of great interest, clearly much more work must be done before their use is a clinical reality. In contrast, ECT treatment may be a non-invasive method of delivering endogenously produced neurotrophic factors to the brain regions that lose their normal trophic support in 112 parkinsonism. Taking into account the published literature on trophic factors, our data suggest that part of the mechanism of action of ECT in PD may be through the enhancement of striatal BDNF and FGF-2 concentrations by either limiting cell death or promoting cell growth. Future studies are required to investigate the true relationship between increased trophic expression and behavioural improvement. As mentioned earlier, the choice of behavioural tests to use in animals with severe 95% lesion may be crucial and needs to be carefully considered. Future studies should examine the neurotrophic effects of ECS in the parkinsonian brain in more detail to determine specifically i f and how increased BDNF and/or FGF-2 may enhance motor and striatal D A terminal function, and what the effects of ECS are on neurotrophic factor concentrations in the SN. We hypothesize that increased neurotrophic activity may have direct effects on striatal D A neurotransmission. BDNF activity plays a direct role in the expression of D 3 receptors in the striatum (Guillin et al., 2001), and may also influence the expression of the Di receptor (Do and Kuzhikandathil, 2005). We have recently shown that repeated ECS in 6-ODA-lesioned rats enhances striatal Di and D 3 receptor binding (Strome et al., 2006b), but further studies are required to determine i f this is a result of enhanced BDNF concentrations in the striatum. In conclusion, this is the first study to investigate the effects of repeated ECS treatment on the concentrations of GDNF, BDNF, and FGF-2 in the brains of 6-OHDA-lesioned rats. We have shown that repeated ECS has region- and protein-specific effects, decreasing GDNF in PFC, increasing BDNF in the striatum and hippocampus, but decreasing BDNF in the PFC, and increasing FGF-2 in the striatum. ECT is widely available as a treatment for psychiatric disorders, and is non-invasive, and safe to use. The clinical evidence strongly suggests that in some patients, ECT can provide almost immediate and fairly long-lasting relief of the motor symptoms of PD. While the mechanism of action is not completely known, and further research is necessary, this study, showing that ECS enhances striatal neurotrophic factor levels, is the first step to the clinical realization of ECT as an effective adjunctive therapy for PD. 113 References: Altar CA, Boylan CB, Fritsche M , Jones BE, Jackson C, Wiegand SJ, et al. (1994) Efficacy of brain-derived neurotrophic factor and neurotrophin-3 on neurochemical and behavioral deficits associated with partial nigrostriatal dopamine lesions. J Neurochem 63: 1021-1032. Altar CA, Cai N , Bilven T, Juhasz M , Conner JM, Acheson A , et al. (1997) Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 389: 856-860. Altar CA, Laeng P, Jurata LW, Brockman JA, Lemire A , Bullard J, et al. (2004) Electroconvulsive seizures regulate gene expression of distinct neurotrophic signaling pathways. J Neurosci 24: 2667-2677. Altar CA, Whitehead RE, Chen R, Wortwein G, Madsen T M (2003) Effects of electroconvulsive seizures and antidepressant drugs on brain-derived neurotrophic factor protein in rat brain. Biol Psychiatry 54: 703-709. Andersen K, Balldin J, Gottfries CG, Granerus A K , Modigh K , Svennerholm L, et al. (1987) A double-blind evaluation of electroconvulsive therapy in Parkinson's disease with "on-off phenomena. Acta Neurol Scand 76: 191-199. Angelucci F, Aloe L , Jimenez-Vasquez P, Mathe A A (2002) Electroconvulsive stimuli alter the regional concentrations of nerve growth factor, brain-derived neurotrophic factor, and glial cell line-derived neurotrophic factor in adult rat brain. J ECT 18: 138-143. Balkowiec A , Katz D M (2002) Cellular mechanisms regulating activity-dependent release of native brain-derived neurotrophic factor from hippocampal neurons. J Neurosci 22: 10399-10407. Berendse HW, Galis-de Graaf Y , Groenewegen HJ (1992) Topographical organization and relationship with ventral striatal compartments of prefrontal corticostriatal projections in the rat. J Comp Neurol 316: 314-347. Berger B, Gaspar P, Verney C (1991) Dopaminergic innervation of the cerebral cortex: unexpected differences between rodents and primates. Trends Neurosci 14: 21-27. Bjorklund A , Kirik D, Rosenblad C, Georgievska B, Lundberg C, Mandel RJ (2000) Towards a neuroprotective gene therapy for Parkinson's disease: use of adenovirus, A A V and lentivirus vectors for gene transfer of GDNF to the nigrostriatal system in the rat Parkinson model. Brain Res 886: 82-98. Bowenkamp K E , Lapchak PA, Hoffer BJ, Miller PJ, Bickford PC (1997) Intracerebroventricular glial cell line-derived neurotrophic factor improves motor function and supports nigrostriatal dopamine neurons in bilaterally 6-hydroxydopamine lesioned rats. Exp Neurol 145: 104-117. Burnet PWJ, Sharp T, LeCorre SM, Harrison PJ (1999) Expression of 5-HT receptors and the 5-HT transporter in rat brain after electroconvulsive shock. Neurosci Lett 277: 79-82. Chauhan N B , Siegel GJ, Lee J M (2001) Depletion of glial cell line-derived neurotrophic factor in substantia nigra neurons of Parkinson's disease brain. J Chem Neuroanat 21: 277-288. 114 Chen A C , Eisch A J , Sakai N , Takahashi M , Nestler EJ, Duman RS (2001a) Regulation of GFRalpha-1 and GFRalpha-2 mRNAs in rat brain by electroconvulsive seizure. Synapse 39: 42-50. Chen A C H , Shin K - H , Duman RS, Sanacora G (2001b) ECS-induced mossy fiber sprouting and BDNF expression are attenuated by ketamine pretreatment. J ECT 17: 27-32. Datto C, Rai A K , Ilivicky HJ, Caroff SN (2002) Augmentation of seizure induction in electroconvulsive therapy: a clinical reappraisal. J ECT 18: 118-125. Do TN, Kuzhikandathil E V (2005) Neurotrophins regulate Di dopamine receptor expression. Soc Neurosci Abstr 33.2. Eslamboli A , Georgievska B, Ridley R M , Baker HF, Muzyczka N , Burger C, et al. (2005) Continuous low-level glial cell line-derived neurotrophic factor delivery using recombinant adeno-associated viral vectors provides neuroprotection and induces behavioral recovery in a primate model of Parkinson's disease. J Neurosci 25: 769-777. Fall PA, Ekman R, Granerus A K , Thorell L H , Walinder J (1995) ECT in Parkinson's disease. Changes in motor symptoms, monoamine metabolites and neuropeptides. J Neural Transm Park Dis Dement Sect 10: 129-140. Follesa P, Gale K , Mocchetti I (1994) Regional and temporal pattern of expression of nerve growth factor and basic fibroblast growth factor mRNA in rat brain following electroconvulsive shock. Exp Neurol 127: 37-44. Gil l SS, Patel N K , Hotton GR, O'Sullivan K, McCarter R, Bunnage M , et al. (2003) Direct brain infusion of glial cell line-derived neurotrophic factor in Parkinson disease. Nat Med 9: 589-595. Goggi J, Pullar IA, Carney SL, Bradford HF (2003) The control of [ 1 2 5 I]BDNF release from striatal rat brain slices. Brain Res 967: 201-209. Gombos Z, Spiller A , Cottrell GA, Racine RJ, Burnham W M (1999) Mossy fiber sprouting induced by repeated electroconvulsive shock seizures. Brain Res 844: 28-33. Green AR, Heal DJ, Grahame-Smith D G (1977) Further observations on the effect of repeated electroconvulsive shock on the behavioural responses of rats produced by increases in the functional ability of brain 5-hydroxytryptamine and dopamine. Psychopharmacology 52: 195-200. Guillin O, Diaz J, Carrol P, Griffon N , Schwartz JC, Sokoloff P (2001) BDNF controls dopamine D 3 receptor expression and triggers behavioural sensitization. Nature 411: 86-89. Guo H, Tang Z, Y u Y , X u L, Jin G, Zhou J (2002) Apomorphine induces trophic factors that support fetal rat mesencephalic dopaminergic neurons in cultures. Eur J Neurosci 16: 1861-1870. Gwinn RP, Kondratyev A , Gale K (2002) Time-dependent increase in basic fibroblast growth factor protein in limbic regions following electroshock seizures. Neuroscience 114: 403-409. Hornykiewicz O (1972) Dopamine and extrapyramidal motor function and dysfunction. Res Publ Assoc Res Nerv Ment Dis 50: 390-415. 115 Iancu R, Mohapel P, Brundin P, Paul G (2005) Behavioral characterization of a unilateral 6-OHDA-lesion model of Parkinson's disease in mice. Behav Brain Res 162: 1-10. Jacobsen JP, Mork A (2004) The effect of escitalopram, desipramine, electroconvulsive seizures and lithium on brain-derived neurotrophic factor mRNA and protein expression in the rat brain and the correlation to 5-HT and 5-HIAA levels. Brain Res 1024: 183-192. Jankovic J (2005) Motor fluctuations and dyskinesias in Parkinson's disease: clinical manifestations. Mov Disord 20 Suppl 11: SI 1-S16. Klein RL, Lewis M H , Muzyczka N , Meyer E M (1999) Prevention of 6-hydroxydopamine-induced rotational behavior by BDNF somatic gene transfer. Brain Res 847: 314-320. Kojima M , Takei N , Numakawa T, Ishikawa Y , Suzuki S, Matsumoto T, et al. (2001) Biological characterization and optical imaging of brain-derived neurotrophic factor-green fluorescent protein suggest an activity-dependent local release of brain-derived neurotrophic factor in neurites of cultured hippocampal neurons. J Neurosci Res 64: 1-10. Kondratyev A , Ved R, Gale K (2002) The effects of repeated minimal electroconvulsive shock exposure on levels of mRNA encoding fibroblast growth factor-2 and nerve growth factor in limbic regions. Neuroscience 114: 411-416. Lamont SR, Paulls A , Stewart C A (2001) Repeated electroconvulsive stimulation, but not antidepressant drugs, induced mossy fibre sprouting in the rat hippocampus. Brain Res 893: 53-58. Lang A E , Gi l l S, Patel N K , Lozano A, Nutt JG, Perm R, et al. (2006) Randomized controlled trial of intraputamenal glial cell line-derived neurotrophic factor infusion in Parkinson disease. Ann Neurol 59: 459-466. Lapchak PA, Jiao S, Collins F, Miller PJ (1997) Glial cell line-derived neurotrophic factor: distribution and pharmacology in the rat following a bolus intraventricular injection. Brain Res 747: 92-102. Malberg J, Eisch A J , Nestler EJ, Duman RS (2000) Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 20: 9104-9110. Meredith GE, Callen S, Scheuer D A (2002) Brain-derived neurotrophic factor expression is increased in the rat amygdala, piriform cortex and hypothalamus following repeated amphetamine administration. Brain Res 949: 218-227. Mogi M , Togari A , Kondo T, Mizuno Y , Komure O, Kuno S, et al. (1999) Brain-derived growth factor and nerve growth factor concentrations are decreased in the substantia nigra in Parkinson's disease. Neurosci Lett 270: 45-48. Mueller D, Chapman C A , Stewart J (2006) Amphetamine induces dendritic growth in ventral tegmental area dopaminergic neurons in vivo via basic fibroblast growth factor. Neuroscience 137: 727-735. 116 Nakajima K, Hida H, Shimano Y , Fujimoto I, Hashitani T, Kumazaki M , et al. (2001) GDNF is a major component of trophic activity in DA-depleted striatum for survival and neurite extension of DAergic neurons. Brain Res 916: 76-84. Newton SS, Collier EF, Hunsberger J, Adams D, Terwilliger RZ, Selvanayagam E, et al. (2003) Gene profile of electroconvulsive seizures: induction of neurotrophic and angiogenic factors. J Neurosci 23: 10841-10851. Nibuya M , Morinobu S, Duman RS (1995) Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J Neurosci 15: 7539-7547. Nowak G, Zak J (1989) Repeated electroconvulsive shock (ECS) enhances striatal D - l dopamine receptor turnover in rats. Eur J Pharmacol 167: 307-308. Patel N K , Bunnage M , Plaha P, Svendsen CN, Heywood P, Gi l l SS (2005) Intraputamenal infusion of glial cell line-derived neurotrophic factor in PD: a two-year outcome study. Ann Neurol 57: 298-302. Paxinos G, Watson C (1997) The Rat Brain in Stereotaxic Coordinates. San Diego: Academic Press. Pridmore S, Pollard C (1996) Electroconvulsive therapy in Parkinson's disease: 30 month follow up. J Neurol Neurosurg Psychiatry 61: 693-700. Schallert T, Fleming S M , Leasure JL, Tillerson JL, Bland ST (2000) CNS plasticity and assessment of forelimb sensorimotor outcome in unilateral rat models of stroke, cortical ablation, parkinsonism and spinal cord injury. Neuropharmacology 39: 777-787. Schallert T, Tillerson JL (2000) Intervention strategies for degeneration of dopamine neurons in Parkinsonism. In: Central Nervous System Diseases (Emerich DF, Dean RL, Sanberg PR, eds), pp 131-151. Totowa: Humana Press. Schallert T, Woodlee M T (2005) Orienting and Placing. In: The behavior of the laboratory rat: a handbook with tests. (Whishaw IQ, Kolb B, eds), pp 129-140. New York: Oxford University Press. Scott BW, Wojtowicz JM, Burnham W M (2000) Neurogenesis in the dentate gyrus of the rat following electroconvulsive shock seizures. Exp Neurol 165: 231-236. Shi L - H , Woodward DJ, Luo F, Anstrom K, Schallert T, Chang J-Y (2004) High-frequency stimulation of the subthalamic nucleus reverses limb-use asymmetry in rats with unilateral 6-hydroxydopamine lesions. Brain Res 1013: 98-106. Shults CW, Kimber T, Altar C A (1995) BDNF attenuates the effects of intrastriatal injection of 6-hydroxydopamine. NeuroReport 6: 1109-1112. Shults CW, Ray J, Tsuboi K, Gage FH (2000) Fibroblast growth factor-2-producing fibroblasts protect the nigrostriatal dopaminergic system from 6-hydroxydopamine. Brain Res 883: 192-204. 117 Slevin JT, Gerhardt G A , Smith CD, Gash D M , Kryscio R, Young B (2005) Improvement of bilateral motor functions in patients with Parkinson disease through the unilateral intraputaminal infusion of glial cell line-derived neurotrophic factor. J Neurosurg 102: 216-222. Stern M B (1991) Electroconvulsive therapy in untreated Parkinson's disease. Mov Disord 6: 265. Strome E M , Clark C M , Zis AP, Doudet DJ (2005) Electroconvulsive shock decreases binding to 5 -HT2 receptors in nonhuman primates: an in vivo positron emission tomography study with [18F]setoperone. Biol Psychiatry 57: 1004-1010. Strome E M , Zis AP, Doudet DJ (2006) Electroconvulsive shock enhances Di and D 3 receptor binding and improves motor performance in 6-OHDA-lesioned rats. Submitted. Tillerson JL, Cohen A D , Philhower J, Miller G, Zigmond MJ , Schallert T (2001) Forced limb-use effects on the behavioral and neurochemical effects of 6-hydroxydopamine. J Neurosci 21: 4427-4435. Tomac A , Lindqvist E, Lin L-FH, Ogren SO, Young D, Hoffer BJ, et al. (1995) Protection and repair of the nigrostriatal dopaminergic system by GDNF in vivo. Nature 373: 335-339. Tooyama I, Kawamata T, Walker D, Yamada T, Hanai K , Kimura H , et al. (1993) Loss of basic fibroblast growth factor in substantia nigra neurons in Parkinson's disease. Neurology 43: 372-376. Vaidya V A , Siuciak JA, Du F, Duman RS (1999) Hippocampal mossy fiber sprouting induced by chronic electroconvulsive seizures. Neuroscience 89: 157-166. Wang L, Muramatsu S, Lu Y , Ikeguchi K, Fujimoto K, Okada T, et al. (2002) Delayed delivery of A A V - G D N F prevents nigral neurodegeneration and promotes functional recovery in a rat model of Parkinson's disease. Gene Ther 9: 381-389. Yoshida K, Higuchi H , Kamata M , Kishimoto M , Shimizu T, Ishikawa Y (1998a) Dopamine releasing response in rat striatum to single and repeated electroconvulsive shock treatment. Prog Neuropsychopharmacol Biol Psychiatry 21:707-715. Yoshida K, Higuchi H , Kamata M , Yoshimoto M , Shimizu T, Hishikawa Y (1998b) Single and repeated electroconvulsive shocks activate dopaminergic and 5-hydroxytryptaminergic neurotransmission in the frontal cortex of rats. Prog Neuropsychopharmacol Biol Psychiatry 22: 435-444. Yurek D M , Fletcher-Turner A (2001) Differential expression of GDNF, BDNF, and NT-3 in the aging nigrostriatal system following a neurotoxic lesion. Brain Res 891: 228-235. Zervas IM, Fink M (1991) ECT for refractory Parkinson's disease. Convuls Ther 7: 222-223. Zetterstrom TSC, Pei Q, Grahame-Smith D G (1998) Repeated electroconvulsive shock extends the duration of enhanced gene expression for BDNF in rat brain compared with a single administration. Mol Brain Res 57: 106-110. 118 Zis AP, Nomikos GG, Brown EE, Damsma G, Fibiger HC (1992) Neurochemical effects of electrically and chemically induced seizures: an in vivo microdialysis study in the rat hippocampus. Neuropsychopharmacology 7: 189-195. 119 Chapter 7 General Discussion Preamble: The ultimate goal of this body of research was to investigate the observation that repeated ECT treatments in patients with PD has a dramatic effect on their motor symptoms, improvements that usually take only a few treatments to emerge, and that last for several weeks. Based on the known neurochemical deficits in PD and the anatomy of the main system involved, the basal ganglia, and taking into account the literature on the effects of repeated ECS treatment in the normal brain, and the proposed mechanism of action of ECT in depression, we developed a working hypothesis to explain how ECT treatment can improve the motor symptoms of PD. In order to address the various aspects of our working hypothesis, however, it was necessary to first identify the appropriate animal models and tools to use. We performed several pilot and validation studies (Chapters 2-4 and Appendices A and B) to identify and develop those animal models and techniques. Only with the right tools at our disposal could we move forward to undertake studies directly pertaining to our working hypothesis regarding the mechanism of action of ECT in PD (Chapters 5 and 6). Taken together, this body of work tells the story of a journey of discovery, focused on examining the effects of ECS on the brain, but encompassing different animal models, different neurochemical systems, and a variety of tools, ranging from non-human primates to rats, 5-HT to D A to neurotrophic factors, PET to behavioural analysis, autoradiography to immunohistochemistry. This final chapter will summarize the findings, explore outstanding issues, discuss the limitations and significance of the research, and look towards the future. Model and Tool Development: Model Development: Antidepressant Treatments and the 5-HT2 Receptor In developing our models and tools, we relied heavily upon the observation that the 5-H T 2 receptor undergoes specific plastic changes after particular antidepressant interventions. In rodents and humans, chronic treatment with many classes of antidepressant drugs, including tricyclic antidepressants (TCAs), monoamine oxidase inhibitors, some SSRIs, and atypical antidepressants, downregulation of 5-HT 2 receptors appears to be the predominant effect (Bergstrom and Kellar, 1979a; Tang et al., 1981; Peroutka and Snyder, 1990; Todd et al., 1995; Attar-Levy et al., 1999; Yatham et a l , 1999; Mischoulon et al., 2002). In contrast, in rodents, increased 5-HT 2 receptor binding and mRNA expression are seen after chronic ECS (Bergstrom and Kellar, 1979b; Kellar and Stockmeier, 1986; Biegon and Israeli, 1987; Butler et al., 1993; Burnet et al., 1995). These specific and widely reproduced changes to the 5-HT 2 receptor gave us a context to work within as we examined the use of non-human primates to investigate ECS-induced changes in neurochemistry (Chapter 2), the use of anaesthesia for rodent ECS (Appendix A), and the use of PET tracers for in vitro autoradiography (Chapter 3). In Chapter 2, we show that, like antidepressant drug treatment in humans and rodents, repeated ECS in non-human primates decreases cortical 5-HT 2 binding. This study was particularly significant because it was the first to examine the effects of electroconvulsive stimuli on 5-HT2 receptors in an organism phylogenetically higher than a rodent. The upregulation of the 5-HT 2 receptor in the rodent brain after repeated ECS is somewhat of an anomaly in the antidepressant literature, and is counterintuitive. If antidepressant treatments act to enhance 5-HT, then the post-synaptic 5-HT 2 receptor should downregulate to compensate. We suggested that the reason ECS in rats was consistently reported to increase 5-HT 2 receptor binding, instead 120 of decreasing it as in primates, may have been partially due to the fact that ECT is administered under anaesthesia to humans and non-human primates, but typically without anaesthesia to rats. As part of our pilot study investigating the use of anaesthesia for rat ECS, we also investigated i f anaesthesia does have an impact on ECS-induced changes in the 5-HT2 receptor (Appendix A). That study confirmed that, even under anaesthesia, repeated electroconvulsive stimulation in rats increases cortical 5-HT2 receptor binding. For our purposes, this was encouraging, in that we could confidently administer ECS to rats under ketamine anaesthesia and still elicit one of the hallmark effects. In screening the literature more thoroughly, several other reports emerged where ECS had been given to rodents under halothane anaesthesia, and these all also show 5-HT 2 upregulation (Goodwin et al., 1984; Metz and Heal, 1986; Burnet et al., 1995; Burnet et a l , 1999). None of these reports, however, presented a satisfactory explanation for the unexpected upregulation of the 5-HT 2 receptor after ECS in rats. There are several possible explanations for the differential regulation of 5-HT2 receptors after repeated electroconvulsive stimulation in different species. First of all, the basal state of the system may strongly influence the effects exogenous treatments have on 5-HT neurotransmission. The state of cortical 5-HT2 receptors in depressed or remitted patients appears to be somewhat context-dependent, showing increases, decreases, or no changes in specific patient populations. Post-mortem studies of suicide victims or PET studies of patients with severely pessimistic attitudes show increased prefrontal cortex 5-HT 2 binding (Yates et al., 1990; Hrdina and Vu, 1993; Meyer et al., 2003). Antidepressant drug naive patients (Meltzer et al., 1999), and those studied in their first episode (Meyer et al., 1999) do not show changes in cortical 5-HT2 receptor binding, while those with chronic or recurrent depression show widespread decreases in cortical 5-HT2 binding (Yatham et al., 2000; Larisch et al., 2001). We did not investigate the effects of ECS on 5-HT2 receptors in an animal model of depression, nor, for that matter, did any of the previous investigators of ECS-induced changes in 5-HT 2 receptors in rodents. The normal state of the 5-HT system before ECS treatment in these animals could influence the plasticity of the system, and further investigations of the effects of antidepressants on 5-HT neurotransmission in animal models of depression are warranted. Secondly, species differences in cortical neuroanatomy and, in particular connectivity, are not insignificant (reviewed in Kaas, 1987). As a treatment, ECT is highly dependent on cortical circuitry to exert its effects, and the differences between the primate and rodent cortex may contribute to the unexpected upregulation of 5-HT2 receptors by ECS treatment. Finally, that TCAs and ECT exert opposite effects on 5-HT2 receptors in the rat brain may result from the fact that they have different mechanisms of action. Indeed, while the decreases in 5-HT 2 receptor binding induced by the TCAs in the frontal cortex as measured by [3H]ketanserin is in the order of 40% (Staton et al., 1986; Geretsegger et al., 1998; Strome et al., 2005), the effects of ECS are more subdued, 16% in our study (Fig. A.4) and typically 20% in the literature (Pandey et al., 1992; Butler et al., 1993). While ECT is thought to exert its effects through direct enhancement of synaptic 5-HT (Zis et al., 1992; Yoshida et al., 1998; Juckel et al., 1999), TCAs, along with their noradrenergic reuptake inhibition properties, may also be acting directly as 5-HT2 antagonists (Sanchez and Hyttel, 1999). Antidepressant treatments may then be considered to act by modulating different aspects of 5-HT neurotransmission, and either influence 5-HT 2 receptor expression directly, as in the case of TCAs (direct 5-HT2 antagonists), or indirectly, through ECS -induced changes in synaptic 5-HT (indirect 5-HT 2 agonists). From a clinical standpoint, however, it may not be important what effects these various antidepressant treatments have on 5-HT 2 receptors. We use 5-HT2 receptors in research as an indirect means to infer the effects of interventions on synaptic levels of their endogenous ligand, 5-HT. If the serotonin hypothesis of depression holds, then any treatment that increases the synaptic availability of 5-HT in target regions should have a beneficial effect. ECS or ECT may 121 do this directly by releasing 5-HT through widespread electrical stimulation of the brain, and TCAs may increase synaptic 5-HT indirectly, by inhibiting reuptake. If the net effect of diverse antidepressant treatments is the same, that is increased synaptic 5-HT, i f it occurs via different mechanisms, then the use of 5 - H T 2 receptors as an indirect marker for synaptic levels of 5-HT may be misleading. The fact that antidepressant treatments can have opposing effects on 5-HT2 receptors and still be effective treatments speaks to the incredible plasticity of monoamine neurotransmitter systems, and suggests that a variety of mechanisms have evolved to compensate for changes in neurochemical activity. Tool Development: Autoradiography, Behavioural Analysis, Measuring Neurotrophic Factors In our MPTP-treated non-human primates, the effects of repeated ECS treatment on motor behaviour were dramatic and fairly long-lasting (Doudet et al, unpublished). The effects on D A neurotransmission, however, were less clear, and were difficult to interpret in our small sample size. To pursue this line of investigation further, it was necessary to move on to a small animal rodent model, where larger numbers of subjects can be employed without the same ethical and financial limitations of non-human primate research. While we moved to a rodent model, we still wanted to employ the tracers for the D A system used in vivo in PET. At the time, we did not have access to a small animal PET scanner, so instead, we decided to use the tracers for in vitro autoradiography. One of the major technical projects of this doctoral research, then, was the development and validation of a quantitative in vitro technique for using PET tracers autoradiographically (whereas most previous autoradiographic work with these tracers had been qualitative). In Chapter 3, we describe, in detail, the procedure for performing the binding assay with PET tracers, the quantification and analysis of the data, and our determination of the resolution of the technique. To validate the technique, we treated rats chronically with the antidepressant desipramine, which downregulates cortical 5-HT 2 receptors (Bergstrom and Kellar, 1979a; Goodnough and Baker, 1994), to show that quantitative autoradiography with a PET tracer for 5-HT 2 receptors gives data that are not significantly different than the traditional tritiated ligand for this receptor. In the last several years, one of the major projects in our laboratory has been the development and validation of a battery of non-pharmacological behavioural tests for rat models of PD. The TB test is a new beam-walking task that is both designed to prevent the animals from making postural compensations and masking improvements in motor function, and is simple for the animals to learn (Schallert and Woodlee, 2005). The relationship between the integrity of the D A system and performance on the task, however, had not been examined previously. In Chapter 4, we show that, indeed, in unilaterally 6-OHDA-lesioned rats, TB test scores are correlated with striatal D A terminal integrity as measured by autoradiography. Among neurotrophic factors, BDNF has two very unique characteristics: 1) its expression is activity-dependent (Kojima et al., 2001; Balkowiec and Katz, 2002; Goggi et al., 2003), and 2) it can be transported retrogradely from cortical areas to subcortical targets (Altar et al., 1997). Both of these characteristics make it a prime target for investigators examining the effects of ECS on the brain (see Table 1.3 in Chapter 1), because ECS causes widespread neuronal activity and because the primary stimulus is applied to the cerebral cortex. In Appendix B, we used in situ hybridization in an attempt to measure BDNF mRNA in the rat brain after ECS treatment. While no prior reports on ECS-induced changes in striatal BDNF mRNA existed, we wanted to investigate this phenomenon ourselves, since striatal transcription of the neurotrophic factors was a key aspect to our early working hypothesis. Our inability to detect BDNF mRNA in the striatum, while at the same time reproducing the widely reported increase in piriform cortex 122 BDNF mRNA expression after repeated ECS treatment, led us to modify our working hypothesis, and to examine neurotrophic factor protein levels directly in our later study. Thus, in our early pilot and validation studies, we explored the use of non-human primates and PET imaging to examine the neurochemical effects of repeated ECS treatment, developed and validated a technique for in vitro quantitative autoradiography with PET tracers, investigated the use of anaesthesia for rat ECS, showed that the TB test is a DA-dependent motor task, and determined that in situ hybridization was not the best tool for investigating the effects or repeated ECS treatment on striatal levels of neurotrophic factors. These early studies allowed us to determine the most appropriate animal models and tools to use to address our specific hypotheses regarding the mechanism of action of ECT in PD. The Effects of Repeated ECS on the DA System in Parkinsonian Rats: Status of Working Hypothesis The major goal of this body of work was to perform several key studies examining the mechanism of action of ECT in PD. In particular, our working hypothesis identified three specific areas of investigation: motor behaviour, D A receptors, and neurotrophic factors. Our experiments addressing those hypotheses have shown that repeated ECS treatment in unilateral 6-OHDA-lesioned rats: 1) improves hindlimb, but not forelimb, motor function (Chapter 5); 2) increases D i and D 3 receptor binding in the dorsal and ventral striatum, without affecting T>2 receptor binding (Chapter 5); and 3) enhances BDNF and FGF-2, but decreases GDNF protein levels in the striatum (Chapter 6). Hypothesis 1): Repeated ECS improves motor performance in 6-OHDA-lesioned rats In Chapter 5, we showed that, while repeated ECS treatment in unilateral 6-OHDA-lesioned rats did not improve forelimb function in the Cylinder Test, hindlimb performance on the TB test was significantly improved compared to sham-treated controls. In hindsight, the lack of improvement on the Cylinder Test was not surprising, considering the extent of the lesion in this rat model of PD. In the unilateral 6-OHDA lesion model of PD, the loss of D A innervation to the ipsilateral striatum is extreme, and all animals in our studies showed greater than 90% depletion of striatal D A terminals. The Cylinder Test is widely used to evaluate forelimb motor function in the unilateral 6-OHDA-lesion model of PD, and while all reports agree that the lesion itself causes a severe impairment in the forelimb contralateral to the lesion on this task, recovery of that function only occurs after very specific interventions, such as lentivector delivery of GDNF to the striatum and SN (Dowd et al., 2005), through forced use of the impaired forelimb (Tillerson et al., 2001), likely as a result of increased striatal GDNF (Cohen et al., 2003), or DBS (Shi et al., 2004). In contrast, striatal transplantation of fetal D A neurons, which provides a focal replacement of striatal D A does not improve forelimb function in the Cylinder test, even though drug-induced behaviours are enhanced (Bjorklund et al., 1994; Dowd and Dunnett, 2004). It appears, then, that this test is only sensitive to antiparkinsonian interventions that either a) provide strong DAergic trophic support to the SN and striatum (and presumably restore some of the lost D A innervation), or b) bypass the striatal D A deficit, and act to inhibit the overactive basal ganglia output structures to normalize motor cortex activity. Synthesizing the literature, our inability to detect an improvement on the Cylinder Test may suggest that repeated ECS treatment in unilateral 6-OHDA-lesioned rats does not induce a major reinnervation of the lesioned striatum. We did, however, detect a significant improvement in test scores on the TB test in ECS-treated rats compared to sham-treated controls (Chapter 5). While the TB test measures the number of hindlimb footfaults off the main surface of the beam, this task evaluates more than just the placement of the limb, requiring a whole host of skills to perform, including balance, 123 coordination, and motivation. In that sense, it is more of a measure of gross motor function than of either hindlimb or fine motor function. Given both the extremity of unilateral striatal D A depletion in our studies, and the non-invasive and indirect nature of ECS treatment, without using pharmacological agents to elicit DA-dependent motor behaviours, gross improvements in motor function are a noteworthy finding. Hypothesis 2): Repeated ECS upregulates striatal Dj and D3 receptors We show in Chapter 5 that repeated ECS treatment in unilateral 6-OHDA-lesioned rats has specific effects on the expression of D A receptors, upregulating Di and D3 receptor binding in the dorsal striatum and NAcc shell, but having no effect on the D 2 receptor. Interestingly, the effects of repeated ECS were not different in the lesioned versus unlesioned striatum, which may result from the fact that the ECS treatment was applied bilaterally. We also observed specific changes in D A receptors as a result of 6-OHDA lesioning, with the D 2 receptor upregulated in the ipsilateral striatum, and the opposite effect on D3 receptor binding in the ipsilateral NAcc shell, both changes that are widely reported in the literature after D A depletion (Levesque et al., 1992; Narang and Wamsley, 1995; Doudet, 2001; van Kampen and Stoessl, 2003). Based on the previous literature of the effects of repeated ECS on D A receptors in the normal rat brain, we expected to find increased Di (Nowak and Zak, 1989; Sershen et al., 1991) and D3 (Lammers et al., 2000) receptor binding in the dorsal striatum and NAcc shell, respectively, but this is the first report of increased D 3 binding in the dorsal striatum after repeated ECS treatment. Behavioural studies show that enhanced locomotor activity after repeated ECS treatment is only elicited by either mixed agonists (such as apomorphine; Green et al., 1977; Green et al., 1983), or co-administration of Di-like and D 2-like agonists (Smith and Sharp, 1997). Since D 2 receptor binding is unchanged after repeated ECS (Bergstrom and Kellar, 1979b; Reches et a l , 1984; Martin et a l , 1995; Chapter 5), the motor activating effects of D A agonists may result from increased striatal Di and D3 receptor binding. The D 3 receptor is most abundant in the Islands of Calleja and NAcc, and is expressed at very low levels in the dorsal striatum and the rest of the rat brain under normal circumstances (Levesque et al., 1992; Diaz et al., 1995). Chronic treatment with L-DOPA in unilateral 6-OHDA-lesioned rats leads to behavioural sensitization and a dramatic increase in the expression of the D3 receptor in the lesioned dorsal striatum, via activation of the Di receptor and increased BDNF levels (Bordet et a l , 1997; Guillin et al., 2001; van Kampen and Stoessl, 2003). The increase in D3 receptor binding that we observed in the dorsal striatum after repeated ECS treatment was less pronounced than in L-DOPA-induced behavioural sensitization, and yet it may follow a similar mechanism, since we have also shown increased Di receptor binding (Chapter 5) and elevated striatal BDNF (Chapter 6). Although we have not explicitly shown co-expression of the Di and D 3 receptors in this study, co-localization of these two receptor subtypes appears to be the norm whenever the D 3 receptor is expressed, for example in the dorsal striatum after chronic L-DOPA treatment (Bordet et al., 2000), and in the Islands of Calleja and NAcc shell under normal circumstances (Ridray et al., 1998). One important question that arises from our working hypothesis is that, since Di and D 3 receptors have opposite effects on cAMP accumulation, how can their co-expression and co-activation enhance motor output? In fact, it appears that in some cases, the D 3 receptor can be coupled to another intracellular signal transduction pathway, that of mitogen-activated protein (MAP) kinase (Leigh et al., 1983; Cussac et al., 1999; Oldenhof et al., 2001). In addition, co-expression of Di and D 3 receptors has been shown to have both opposite and synergistic effects (Ridray et al., 1998). Ridray and colleagues (1998) have hypothesized that when the Di and D 3 receptors are acting synergistically, the D3 receptor is actually coupled to M A P kinase. 124 We suggest that when the two receptor subtypes are in a synergistic relationship, the relative abundance of the receptors may dictate the functional outcome. For example, synergy between Di and D 3 occurs in L-DOPA-induced behavioural sensitization, but in this case, the D 3 receptor is expressed at high levels, leading to overactivity of the direct pathway of the basal ganglia, and the development of sensitization (Bordet et al., 1997; Bordet et al., 2000). If D 3 receptors are expressed in a low to moderate ratio compared to the Di receptor, however, the synergy between them may enhance the activity in the direct pathway, without causing excessive stimulation. The nature of the synergistic relationship between Di and D 3 therefore may depend on the relative expression of the two receptor subtypes, with moderate levels of D 3 being advantageous, and high levels detrimental. If ECT treatment enhances D 3 expression only moderately, then activity in the direct pathway will be enhanced, but not excessive. Hypothesis 3): Repeated ECS induces the trophic factors BDNF, GDNF and/or FGF-2 in the striatum of 6-OHDA-lesioned rats In Chapter 6, we show that the protein levels of the neurotrophic factors GDNF, BDNF, and FGF-2 are modified by repeated ECS treatment in 6-OFfDA-lesioned rats. We originally hypothesized that all three factors would be increased in the brain regions we investigated (PFC, hippocampus, and striatum), but found that the effects were not as simple as that, and instead observed protein- and region-specific effects. In unilateral 6-OFfDA-lesioned rats, repeated ECS treatment increased BDNF protein in the hippocampus and striatum, and increased FGF-2 protein in the striatum. BDNF protein was significantly decreased in the PFC, and there were slight, but insignificant increases in FGF-2 in the hippocampus and PFC. GDNF was significantly decreased in the PFC, but was not changed elsewhere. We also detected specific effects of the 6-OHDA lesion on brain neurotrophic factor levels, with small increases in GDNF and FGF-2 in the striatum, and a decrease in BDNF in the lesioned PFC. In general, our findings were consistent with the literature on the effects of both 6-OHDA lesions and repeated ECS treatment in the normal rat brain, but there were a few exceptions. BDNF protein was decreased in the PFC, a finding that we suggest might be explained by striatal release of BDNF from prefrontal corticostriatal projections. The majority of striatal BDNF protein arrives from the frontal cortex (Altar et al., 1997), and BDNF is released in an activity-dependent manner (Balkowiec and Katz, 2002). Since electroconvulsive stimuli primarily activate cortical neurons, repeated ECS treatment could result in an accumulation of BDNF in the striatum, and depletion of frontal cortex BDNF stores, as observed in our study. This same pathway may also underlie the lesion-induced decrease that we detected in BDNF protein in the PFC. As a compensatory response to the death of striatal D A terminals, the main striatal BDNF afferent pathway from the frontal cortex may have increased its activity. We also found decreased GDNF protein levels in the PFC, but no changes in the hippocampus or striatum. GDNF is the most potent trophic factor for D A neurons, so our original hypothesis suggested that for GDNF to play a role in the mechanism of action of ECT in PD, its levels should be increased in the striatum after ECS treatment. What may be more important, however, is the concentration of GDNF in the SN. Studies with radioactive GDNF show that it redistributes quickly after intraventricular injection in the rat brain (within 1 hr; (Lapchak et al., 1997), and that by 7 days post-injection, it accumulates in specific brain regions, including the SN, the hypothalamus, the fimbria and the cerebellum (Lapchak et al., 1997). We examined GDNF levels 48 hrs after the last ECS treatment, which may have been enough time for the protein to redistribute. So while we saw decreased GDNF in the PFC, and no changes in the hippocampus, and striatum, there may have been increases in GDNF in areas receiving input from these regions (such as the hypothalamus, cerebellum, fimbria, or SN). 125 Synthesis of the Findings Our results have generally provided support for our working hypothesis (updated here as Figure 7.1, which summarizes the results of these studies). A n improvement in motor function is the primary observation after ECT treatment in patients with PD, and we have shown that 1) motor function is also improved in parkinsonian rodents after repeated ECS treatment. The basis for this improvement in motor function is presumably an enhancement of D A neurotransmission within the striatum, which may result from 2) increased binding to striatal Di and D 3 receptors. The neurotrophic factors BDNF and FGF-2 can both enhance the survival and growth of D A neurons, and we have shown that 3) striatal BDNF and FGF-2 levels are enhanced in the striatum after repeated ECS treatment in 6-OHDA-lesioned rats. Individually, either of our observations on the effects of repeated ECS on striatal D A receptors or neurotrophic factor protein levels could explain improved motor function in parkinsonian animals, but together, they may interact to enhance D A function even further. A schematic of this hypothetical synergism between neurotrophic factors and the D A system is shown in Figure 7.2. Repeated ECS treatment increases basal striatal D A release (Zis et al., 1991), and i f this D A release activates D] receptors and the direct output pathway of the basal ganglia, then motor output is facilitated (Gerfen and Young, 1988; Albin et a l , 1989; Gerfen et a l , 1990). Repeated ECS treatment also likely causes increased striatal BDNF release. BDNF activity plays a direct role in the expression of the D 3 receptor (Guillin et al., 2001), which is typically colocalized with the Di receptor (Ridray et al., 1998). Activating the D 3 receptor, then, also, leads to enhanced motor output. BDNF also enhances the survival of D A neurons, which could lead to even further increases in basal striatal D A release. Individually, ECS-induced striatal D A or BDNF release can facilitate motor output, but together, their combined effects may provide an even longer-lasting effect. This synergism may partially explain the fact that in PD patients, the effects of a course of ECT treatment on their motor function lasts for weeks to months (Friedman and Gordon, 1992; Fall et al., 1995; Pridmore and Pollard, 1996), and does not simply extinguish within hours or days after the end of treatment. Limitations of the Studies One of the primary limitations of this research is the use of the unilateral 6-OHDA-lesion rat model of PD. While this is the most commonly employed animal model of PD, some of the features are not representative of the clinical situation. In this model, the D A and motor deficits are induced in only one hemisphere of the brain and on one side of the body, and the unilateral loss of striatal D A terminals is extreme (> 90% depletion). In these ways, the model is a poor replication of the clinical reality, where patients suffer D A depletion and motor symptoms on both sides of their brain and body, and the disease is usually diagnosed around the time patients have lost approximately half of their nigrostriatal D A neurons and 80% of striatal D A (Bernheimer et al., 1973). A partial lesion model, using, for example, intrastriatal administration of 6-OHDA (Kirik et al., 1998), may have been a more appropriate choice for these studies. A major limitation of the neurotrophic factor study (Chapter 6) was the fact that we were unable to investigate changes in neurotrophic factor protein concentrations in the SN. While we may have been able to dissect a whole midbrain region, in our animals, one SN was lesioned, while the other was intact, and it was not possible to clearly and reproducibly separate the two hemispheres at that level during our dissections. Since all three of the neurotrophic factors that we investigated in this study are transported retrogradely from the striatum to SN (Mufson et al., 1999), where they help promote the survival of D A neurons, it will be very important in future studies to measure neurotrophic factor changes in the SN of parkinsonian animals after repeated ECS treatment. 126 Adapted from Duman et al, 1997. Figure 7.1: Summary of the results of this body of work in the context of the working hypothesis. We have shown that repeated ECS treatment in unilateral 6-OHDA-lesioned rats: (1) improves the performance of the lesioned hindlimb on the TB test; (2) increases binding to striatal Dl and D 3 receptors; and (3) enhances striatal protein levels of BDNF and FGF-2. ECS t M O T O R O U T P U T Figure 7.2: DA and BDNF synergism within the striatum after ECS treatment. Repeated ECS treatment elevates basal striatal DA release, which could act via D, receptors to facilitate motor output. Repeated ECS treatment also increases striatal BDNF protein levels, which 129 In our examinations of repeated ECS treatment in unilateral 6-OHDA-lesioned rats, we chose to examine the behavioural and neurochemical effects at the specific time point of 48 hours after the last ECS treatment. We chose this time point based on the literature and on our own unique situation. Almost all of previous studies in the literature examined the effects of repeated ECS treatment on the rodent brain at 24 hours after the last ECS treatment, but the majority of these studies also did not administer ECS under anaesthesia. We wanted to ensure that any effects of anaesthesia had been eliminated before we tested the animals on complex behavioural tasks and, just as importantly, measured D A receptor binding, which has been shown to be affected by anaesthesia (Momosaki et al., 2004), and so we chose the time point of 48 hours to ensure that the anaesthetic had washed out. Examining the animals' behaviour and neurochemistry at the 48 hour post-ECS time point, however, gave us only a very confined view of the effects of repeated ECS treatment. In contrast to our primate studies (Chapter 2), where we examined the effects of ECS longitudinally, giving the animals PET scans both early in the course of repeated ECS treatment and 24 hours, 1 week and 1 month after the end of treatment, we only investigated the behavioural and neurochemical effects of repeated ECS treatment in 6-OHDA-lesioned rats at a single time point. This limitation of our investigations means that we have gained no understanding of either the development or the duration of the effects of repeated ECS treatment in unilateral 6-OHDA-lesioned rats. In fact, there are very few publications examining the effects of ECS in animal models longitudinally. In rats, our own unpublished observations indicate that, in normal animals, improved performance on the TB test lasts up to 3 weeks after the end of treatment (Strome et al., unpublished). In agreement, Smith and Sharp (1997) found that the motor activating effects of simultaneous administration of Di-like and D 2 -like agonists lasts at least 3 weeks after the end of repeated ECS treatment. To provide further support for the use of ECT as an adjunct treatment for PD, it will be important to examine both the development and the duration of ECS-induced changes in parkinsonian animals in the future. Finally, it is important to note that the evidence that we provide here for ECS-induced changes in parkinsonian animals is purely correlative. That is, we show that ECS causes specific effects in parkinsonian animals, but not that these effects are either necessary or sufficient to improve motor performance. A causal link between the improvement in motor performance observed in parkinsonian animals after repeated ECS treatment and changes in striatal D A receptors or neurotrophic factor levels can only be demonstrated by removing the receptor or neurotrophic factor. Various methods are commonly in use to perform such studies, including genetically modified mice [that either do not express (knockout) or have low expression (heterozygote) of a gene of interest], the use of antisense oligodeoxynucleotides against a specific gene of interest, which binds to the mRNA of that gene and prevents its translation, or the use of pharmacological antagonists, or antibodies against specific proteins. How some of these technologies might be applied to our research question is outlined below in "Directions for Future Study". Directions for Future Study These studies are some of the first to look at the effects of ECS in an animal model of PD, and many questions remain unanswered regarding the mechanism of action of ECT in PD. Specifically, five important questions should be addressed in the future: 1) Do the Di and D 3 receptors underlie the improvement in motor function seen after ECS?, 2) What is the relationship between the increased D 3 receptor binding and striatal BDNF activity?, 3) What is the mechanism for the increased striatal BDNF and FGF-2 concentrations?, 4) What is the role of CREB in ECS-induced changes in 6-OHDA-lesioned rats?, and 5) Does ECS treatment in 6-OHDA-lesioned rats promote survival or growth of striatal D A terminals? 130 To investigate the roles that the Di and D 3 receptors play, antisense oligodeoxynucleotides could be used to knock down the expression of these genes. This approach was taken by van Kampen and Stoessl (2000, 2003) to show that the behavioural sensitization that develops after chronic treatment with L-DOPA depends on Di and D 3 receptors. In our research model, using antisense to block D i and/or D 3 receptors, and then examining the effects of that knock down on motor improvement after ECS treatment would provide strong evidence of the role that these receptor subtypes play in mediating the positive effects of ECT on the motor symptoms of PD. The relationship between the upregulation of D A receptors and increased BDNF activity after ECS could be addressed in the future using an antibody against the BDNF receptor, TrkB. This approach was used by Guillin and colleagues (2001) to show that the induction of the D 3 receptor in the dorsal striatum after behavioural sensitization to L-DOPA depended on BDNF signalling via TrkB. In our studies, we could infuse IgG-TrkB directly into the striatum during the course of repeated ECS treatment, and then examine the expression of Dj and D 3 receptor mRNA and protein, as well as behaviour. We have suggested that increased striatal neurotrophic factor concentrations after ECS treatment could arise from 2 sources; 1) anterograde release from afferent regions (such as the frontal cortex), or 2) the activation of striatal gene transcription and local release of trophic factors. At least for BDNF, we have shown that the latter alternative seems unlikely, since we were unable to detect BDNF mRNA in the striatum after repeated ECS treatment (Appendix B). One way to investigate the first theory is to perform aspiration lesions of the frontal cortex as a means of destroying the corticostriatal BDNF or FGF-2 projections. This approach has been used in two of the seminal studies of the relationship between D A neurotransmission, BDNF, and the D 3 receptor, to show that BDNF is transported anterogradely from the frontal cortex to the striatum (Altar et al., 1997), and that BDNF from the frontal cortex is necessary for L -DOPA-induction of the D 3 receptor in the dorsal striatum (Guillin et al., 2001). By destroying this pathway, and then examining the effects of repeated ECS on BDNF and FGF-2 protein levels in the striatum, the source of the proteins could be determined to be (or not to be) the frontal cortex. An important component to our working hypothesis, but one that we have not addressed in these studies, is CREB. CREB is a transcription factor for many genes, including the neurotrophic factors, and it has been shown to play a crucial role in mediating the effects of chronic antidepressant and ECS treatment in the hippocampus (Nibuya et al., 1996; Thome et al., 2000; Chen et al., 2001). We suggest that CREB may also be important in the ECS-induced changes in the striatum. To investigate the role of CREB in the effects of repeated ECS in the parkinsonian brain, it is important to first examine whether or not CREB phosphorylation is increased in the striatum after repeated ECS treatment, which could be done using a specific antibody against phospho-CREB and immunoblotting (Newton et al., 2002). Duman's group has used a tetracycline-regulated system to reversibly and selectively express the gene for CREB in specific brain regions of transgenic mice (Chen et al., 1998; Sakai et al., 2002), and this approach could also be used to investigate the role of CREB in the ECS-induced changes in the striatum of parkinsonian animals. One of the most robust effects of ECS in the hippocampus is cell growth, in the form of both axonal sprouting (Gombos et al., 1999; Vaidya et al., 1999; Madhav et a l , 2000; Lamont et al., 2001) and neurogenesis (Madsen et al., 2000; Malberg et al., 2000; Scott et al., 2000). In fact, while hippocampal neurogenesis can be stimulated by many antidepressant treatments (Malberg et al., 2000), mossy fibre sprouting appears to be unique to ECS (Lamont et a l , 2001). Striatal D A terminals undergo sprouting in response to very specific stimuli, including partial or severe 6-OHDA or MPTP lesions (Blanchard et al., 1996; Bezard et al., 2000), D 2 receptor 131 antagonism or knockout (Parish et al., 2002), or infusion of GDNF and/or BDNF (Hudson et al., 1995; Lucidi-Phillipi et al., 1995; Batchelor et al., 2000). There is evidence for low levels of cellular proliferation in the striatum and SN under very specific circumstances, including D 3 stimulation (van Kampen et al., 2004; van Kampen and Robertson, 2005). Considering that PD is characterized by a loss of a very specific subset of cells, patients suffering from this disease could greatly benefit from therapies that enhance cell survival or growth. As such, it will very important to address cell proliferation and axonal sprouting after ECS treatment in 6-OHDA-lesioned rats in future studies. Significance of the Findings to PD The most overt effect of ECS in PD patients is a dramatic improvement of their motor symptoms, so it was important to demonstrate improved motor function in our animal model. Moreover, D A receptors are important pharmacological targets in PD, with the standard drug treatments, L-DOPA or D A agonists, acting at these sites. Most PD patients who undergo ECT treatment must have their drug doses reduced or eliminated in order to avoid the development of negative side effects, such as dyskinesia (Zervas and Fink, 1992; Nymeyer and Grossberg, 2002), a phenomenon that could be explained by our observation of upregulated Di and D 3 receptors after repeated ECS in 6-OHDA-lesioned rats. Indeed, the fact that D A receptors within the direct pathway of the basal ganglia, the pathway that facilitates movement, were specifically upregulated, supports the use of ECT to treat the motor symptoms of PD. In recent years, the development of treatments for PD involving the administration of exogenous neurotrophic factors has been a major focus of research, from basic science to clinical trials. Several different delivery methods have been investigated, including gene delivery via viral vectors (Klein et al., 1999; Bjorklund et al., 2000; Wang et al., 2002; Eslamboli et a l , 2005) and direct infusion of the protein into the target area (e.g. intraputamenal infusion of GDNF; (Gill et al., 2003; Patel et a l , 2005; Slevin et al., 2005; Lang et al., 2006). However, this line of investigation faces significant barriers because of the large size of the proteins, and their inability to cross the blood-brain-barrier, and hence the necessary invasiveness of the treatment. Currently, safety concerns are prohibiting the use of viral vectors in human patients, and one Phase II GDNF direct infusion study was halted due to a lack of statistically significant improvement and the appearance of neutralizing antibodies in several patients (Lang et al., 2006) . While the therapeutic potential of neurotrophic factors in treating PD and other diseases is well recognized, clearly much more work must be done before their use is a clinical reality. In contrast, ECT treatment may be a non-invasive method of delivering endogenously produced neurotrophic factors to the brain regions that lose their normal trophic support in parkinsonism. If, as our data suggest, part of the mechanism of action of ECT in PD is the enhancement of striatal BDNF and FGF-2 concentrations, then ECT may provide symptomatic relief for PD patients by either limiting cell death or promoting cell growth. ECT may be a viable therapeutic option for several specific populations of PD patients. Since ECT is primarily used as a treatment for depression, and since 30-40% of PD patients are also depressed (Tandberg et a l , 1996; Slaughter et al., 2001), this seems like an ideal population to treat with ECT. If the neurotrophic effects of ECS are one of the prime outcomes, then patients in the early stages of the disease may benefit the most, since they will still have some remaining D A terminals. Neurotrophic factors could enhance the function of remaining D A terminals either by promoting their survival or by stimulating growth, perhaps in the form of sprouting, to allow them to compensate for those D A neurons that have already been lost. ECT should also be tried in patients who have severe motor symptoms and/or are experiencing a poor response to pharmacotherapy. The usual course of action in such patients is surgery for PD, in the form of pallidotomy or DBS. But brain surgery has inherent risks, and many older patients 132 are not good surgical candidates. Before undergoing surgery, these patients should consider a course of ECT. ECT may also be worth considering in patients with young onset PD, since this population typically experiences complications of L-DOPA treatment much earlier, and delaying L-DOPA treatment for as long as possible is recommended (Kostic et al., 1991). Conclusions: In conclusion, this body of work describes a series of studies undertaken to investigate the mechanism of action of ECT in PD. After carefully evaluating our animal models and developing a set of tools to use, we have shown that repeated ECS treatment in parkinsonian animals has specific effects on motor behaviour, and on striatal D A receptors and neurotrophic factor levels. The combined DAergic and neurotrophic effects of ECS treatment may account for the improved motor function after ECT treatment in PD patients. ECT is a safe, effective, and widely available treatment for depression, and our studies, showing similar effects in the striatum in the parkinsonian brain as the effects in limbic brain regions used to explain the mechanism of action of ECT in depression, provide support for its further use and study as a potential adjunct treatment for PD. 133 References: Albin RL, Young A B , Penney JB (1989) The functional anatomy of basal ganglia disorders. Trends Neurosci 12: 366-375. Altar CA, Cai N , Bilven T, Juhasz M , Conner JM, Acheson A , et al. (1997) Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 389: 856-860. Attar-Levy D, Martinot JL, Blin J, Dao-Castellana M H , Crouzel C, Mazoyer B, et al. (1999) The cortical serotonin2 receptors studied with positron-emission tomography and [18F]-setoperone during depressive illness and antidepressant treatment with clomipramine. Biol Psychiatry 45: 180-186. Balkowiec A , Katz D M (2002) Cellular mechanisms regulating activity-dependent release of native brain-derived neurotrophic factor from hippocampal neurons. J Neurosci 22: 10399-10407. Batchelor PE, Liberatore GT, Porritt M , Donnan GA, Howells D W (2000) Inhibition of brain-derived neurotrophic factor and glial cell line-derived neurotrophic factor expression reduces dopaminergic sprouting in the injured striatum. Eur J Neurosci 12: 3462-3468. Bergstrom DA, Kellar K J (1979a) Adrenergic and serotonergic receptor binding in rat brain after chronic desmethylimipramine treatment. J Pharmacol Exp Ther 209: 256-261. Bergstrom DA, Kellar K J (1979b) Effect of electroconvulsive shock on monoaminergic receptor binding sites in rat brain. Nature 278: 464-466. Bernheimer H , Birkmayer W, Hornykiewicz O, Jellinger K , Seitelberger F (1973) Brain dopamine and the syndromes of Parkinson and Huntington. Clinical, morphological and neurochemical correlations. J Neurol Sci 20: 415-455. Biegon A , Israeli M (1987) Quantitative autoradiographic analysis of the effects of electroconvulsive shock on serotonin-2 receptors in male and female rats. J Neurochem 48: 1386-1391. Bjorklund A , Dunnett SB, Nikkhah G (1994) Nigral transplants in the rat Parkinson model. Functional limitations and strategies to enhance nigrostriatal reconstruction. In: Functional neural transplantation. (Dunnett SB, Bjorkland A , eds), pp 47-68. New York: Raven. Bjorklund A , Kirik D, Rosenblad C, Georgievska B, Lundberg C, Mandel RJ (2000) Towards a neuroprotective gene therapy for Parkinson's disease: use of adenovirus, A A V and lentivirus vectors for gene transfer of GDNF to the nigrostriatal system in the rat Parkinson model. Brain Res 886: 82-98. Bordet R, Ridray S, Carboni S, Diaz J, Sokoloff P, Schwartz JC (1997) Induction of dopamine D 3 receptor expression as a mechanism of behavioral sensitization to levodopa. Proc Natl Acad Sci U S A 94: 3363-3367. Bordet R, Ridray S, Schwartz JC, Sokoloff P (2000) Involvement of the direct striatonigral pathway in levodopa-induced sensitization in 6-hydroxydopamine-lesioned rats. Eur J Neurosci 12: 2117-2123. 134 Bumet PWJ, Mead A , Eastwood SL, Lacey K, Harrison PJ, Sharp T (1995) Repeated ECS differentially affects rat brain 5-HTi A and 5 - H T 2 A receptor expression. NeuroReport 6: 901-904. Burnet PWJ, Sharp T, LeCorre SM, Harrison PJ (1999) Expression of 5-HT receptors and the 5-HT transporter in rat brain after electroconvulsive shock. Neurosci Lett 277: 79-82. Butler M O , Morinobu S, Duman RS (1993) Chronic electroconvulsive seizures increase the expression of serotonin2 receptor mRNA in rat frontal cortex. J Neurochem 61: 1270-1276. Chen A C , Shirayama Y , Shin K H , Neve RL, Duman RS (2001) Expression of the cAMP response element binding protein (CREB) in hippocampus produces an antidepressant effect. Biol Psychiatry 49: 753-762. Chen J, Kelz M B , Zeng G, Sakai N , Steffen C, Shockett PE, et al. (1998) Transgenic animals with inducible, targeted gene expression in brain. Mol Pharmacol 54: 495-503. Cohen A D , Tillerson JL, Smith A D , Schallert T, Zigmond M J (2003) Neuroprotective effects of prior limb use in 6-hydroxydopamine-treated rats: possible role of GDNF. J Neurochem 85: 299-305. Cussac D, Newman-Tancredi A , Pasteau V , Millan M J (1999) Human dopamine D(3) receptors mediate mitogen-activated protein kinase activation via a phosphatidylinositol 3-kinase and an atypical protein kinase C-dependent mechanism. Mol Pharmacol 56: 1025-1030. Diaz J, Levesque D, Lammers C H , Griffon N , Martres MP, Schwartz JC, Sokoloff P (1995) Phenotypical characterization of neurons expressing the dopamine D 3 receptor in the rat brain. Neuroscience 65: 731-745. Doudet DJ (2001) PET studies in the MPTP model of Parkinson's disease. Adv Neurol 86: 187-195. Dowd E, Monville C, Torres E M , Wong LF, Azzouz M , Mazarakis ND, Dunnett SB (2005) Lentivector-mediated delivery of GDNF protects complex motor functions relevant to human Parkinsonism in a rat lesion model. Eur J Neurosci 22: 2587-2595. Dowd E, Dunnett SB (2004) Deficits in a lateralized associative learning task in dopamine-depleted rats with functional recovery by dopamine-rich transplants. Eur J Neurosci 20: 1953-1959. Eslamboli A , Georgievska B, Ridley R M , Baker HF, Muzyczka N , Burger C, et al. (2005) Continuous low-level glial cell line-derived neurotrophic factor delivery using recombinant adeno-associated viral vectors provides neuroprotection and induces behavioral recovery in a primate model of Parkinson's disease. J Neurosci 25: 769-777. Fall PA, Ekman R, Granerus A K , Thorell L H , Walinder J (1995) ECT in Parkinson's disease. Changes in motor symptoms, monoamine metabolites and neuropeptides. J Neural Transm Park Dis Dement Sect 10: 129-140. Friedman J, Gordon N (1992) Electroconvulsive therapy in Parkinson's disease: a report on five cases. Convuls Ther 8: 204-210. 135 Geretsegger C, Rochowanski E, Kartnig, Unterrainer A F (1998) Propofol and methohexital as anesthetic agents for electroconvulsive therapy (ECT): a comparison of seizure-quality measures and vital signs. J ECT 14: 28-35. Gerfen CR, Engber T M , Mahan L C , Susel Z, Chase TN, Monsma FJ, Jr., et al. (1990) Di and D 2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science 250: 1429-1432. Gerfen CR, Young WS, III (1988) Distribution of striatonigral and striatopallidal peptidergic neurons in both patch and matrix compartments: an in situ hybridization histochemistry and fluorescent retrograde tracing study. Brain Res 460: 161-167. Gil l SS, Patel N K , Hotton GR, O'Sullivan K , McCarter R, Bunnage M , et al. (2003) Direct brain infusion of glial cell line-derived neurotrophic factor in Parkinson disease. Nat Med 9: 589-595. Goggi J, Pullar IA, Carney SL, Bradford HF (2003) The control of [ 1 2 5 I]BDNF release from striatal rat brain slices. Brain Res 967: 201-209. Gombos Z, Spiller A , Cottrell GA, Racine RJ, Burnham W M (1999) Mossy fiber sprouting induced by repeated electroconvulsive shock seizures. Brain Res 844: 28-33. Goodnough DB, Baker GB (1994) 5-hydroxytryptamine2 and (3-adrenergic receptor regulation in rat brain following chronic treatment with desipramine alone and in combination. J Neurochem 62: 2262-2268. Goodwin G M , Green A R , Johnson P (1984) 5-HT 2 receptor characteristics in frontal cortex and 5-HT2 receptor-mediated head-twitch behaviour following antidepressant treatment to mice. Br J Pharmacol 83: 235-242. Green AR, Heal DJ, Grahame-Smith D G (1977) Further observations on the effect of repeated electroconvulsive shock on the behavioural responses of rats produced by increases in the functional ability of brain 5-hydroxytryptamine and dopamine. Psychopharmacology 52: 195-200. Green AR, Heal DJ, Johnson P, Laurence BE, Nimgaonkar V (1983) Antidepressant treatments: effects in rodents on dose-response curves of 5-hydroxytryptamine- and dopamine-mediated behaviours and 5-HT 2 receptor number in frontal cortex. Br J Pharmacol 80: 377-385. Guillin O, Diaz J, Carrol P, Griffon N , Schwartz JC, Sokoloff P (2001) BDNF controls dopamine D 3 receptor expression and triggers behavioural sensitization. Nature 411: 86-89. Hrdina P, Vu TB (1993) Chronic fluoxetine treatment upregulates 5-HT uptake sites and 5-HT2 receptors in rat brain: an autoradiographic study. Synapse 14: 324-331. Hudson J, Granholm A C , Gerhardt GA, Henry M A , Hoffman A , Biddle P, et al. (1995) Glial cell line-derived neurotrophic factor augments midbrain dopaminergic circuits in vivo. Brain Res Bull 36: 425-432. Juckel G, Mendlin A , Jacobs B L (1999) Electrical stimulation of rat medial prefrontal cortex enhances forebrain serotonin output: implications for electroconvulsive therapy and transcranial magnetic stimulation in depression. Neuropsychopharmacology 21: 391-398. 136 Kaas JH (1987) The organization of neocortex in mammals: implications for theories of brain function. Annu Rev Psychol 38: 129-151. Kellar K J , Stockmeier C A (1986) Effects of electroconvulsive shock and serotonin axon lesions on beta-adrenergic and serotonin-2 receptors in rat brain. Ann N Y Acad Sci 76-90. Kirik D, Rosenblad C, Bjorkland A (1998) Characterization of behavioral and neurodegenerative changes following partial lesions of the nigrostriatal dopamine system induced by intrastriatal 6-hydroxydopamine in the rat. Exp Neurol 152: 259-277. Klein RL, Lewis M H , Muzyczka N , Meyer E M (1999) Prevention of 6-hydroxydopamine-induced rotational behavior by BDNF somatic gene transfer. Brain Res 847: 314-320. Kojima M , Takei N , Numakawa T, Ishikawa Y , Suzuki S, Matsumoto T, et al. (2001) Biological characterization and optical imaging of brain-derived neurotrophic factor-green fluorescent protein suggest an activity-dependent local release of brain-derived neurotrophic factor in neurites of cultured hippocampal neurons. J Neurosci Res 64: 1-10. Kostic V , Przedborski S, Flaster E, Sternic N (1991) Early development of levodopa-induced dyskinesias and response fluctuations in young-onset Parkinson's disease. Neurology 41: 202-205. Lammers C H , Diaz J, Schwartz JC, Sokoloff P (2000) Selective increase of dopamine D 3 receptor gene expression as a common effect of chronic antidepressant treatments. Mol Psychiatry 5(4): 378-388. Lamont SR, Paulls A , Stewart C A (2001) Repeated electroconvulsive stimulation, but not antidepressant drugs, induced mossy fibre sprouting in the rat hippocampus. Brain Res 893: 53-58. Lang A E , Gi l l S, Patel N K , Lozano A , Nutt JG, Penn R, et al. (2006) Randomized controlled trial of intraputamenal glial cell line-derived neurotrophic factor infusion in Parkinson disease. Ann Neurol 59: 459-466. Lapchak PA, Jiao S, Collins F, Miller PJ (1997) Glial cell line-derived neurotrophic factor: distribution and pharmacology in the rat following a bolus intraventricular injection. Brain Res 747:92-102. Larisch R, Klimke A , Mayoral F, Hamacher K , Herzog HR, Vosberg H , et al. (2001) Disturbance of serotonin 5HT 2 receptors in remitted patients suffering from hereditary depressive disorder. Nuklearmedizin 40: 129-134. Leigh PN, Reavill C, Jenner P, Marsden CD (1983) Basal ganglia outflow pathways and circling behaviour in the rat. J Neural Transm 58: 1-41. Levesque D, Diaz J, Pilon C, Martres MP, Giros B, Souil E, et al. (1992) Identification, characterization, and localization of the dopamine D 3 receptor in rat brain using 7-[3H]hydroxy-N,N-di-n-propyl-2-aminotetralin. Proc Natl Acad Sci U S A 89: 8155-8159. 137 Lucidi-Phillipi C A , Gage FH, Shults CW, Jones KR, Reichardt LF , Kang UJ (1995) Brain-derived neurotrophic factor-transduced fibroblasts: production of BDNF and effects of grafting to the adult rat brain. J Comp Neurol 354: 361-376. Madhav TR, Pei Q, Grahame-Smith DG, Zetterstrom TSC (2000) Repeated electroconvulsive shock promotes the sprouting of serotonergic axons in the lesioned rat hippocampus. Neuroscience 97: 677-683. Madsen T M , Treschow A , Bengzon J, Bolwig TG, Lindvall O (2000) Increased neurogenesis in a model of electroconvulsive therapy. Biol Psychiatry 47: 1043-1049. Malberg J, Eisch A J , Nestler EJ, Duman RS (2000) Chronic antidepressant treatment increases neurogenesis in adult rat hippocampus. J Neurosci 20: 9104-9110. Martin K F , Phillips I, Cheetham SC, Heal DJ (1995) Dopamine D 2 receptors: a potential pharmacological target for nomifensine and tranylcypromine but not other antidepressant treatments. Pharmacol Biochem Behav 51: 565-569. Meltzer CC, Price JC, Mathis C A , Greer PJ, Cantwell M N , Houck PR, et al. (1999) PET imaging of serotonin type 2A receptors in late-life neuropsychiatric disorders. Am J Psychiatry 156: 1871-1878. Metz A , Heal DJ (1986) In mice repeated administration of electroconvulsive shock or desmethylimipramine produces rapid alterations in 5-HT2-mediated head-twitch responses and cortical 5-HT 2 receptor number. Eur J Pharmacol 126: 159-162. Meyer J, Kapur S, Houle S, DaSilva J, Owczarek B, Brown G M , et al. (1999) Prefrontal cortex 5-HT2 receptors in depression: an [18F]setoperone PET imaging study. A m J Psychiatry 156: 1029-1034. Meyer JH, McMain S, Kennedy SH, Korman L, Brown G M , DaSilva JN, et al. (2003) Dysfunctional attitudes and 5-HT 2 receptors during depression and self-harm. Am J Psychiatry 160: 90-99. Mischoulon D, Dougherty DD, Bottonari K , Gresham RL, Sonawalla SB, Fischman AJ , et al. (2002) An open pilot study of nefazodone in depression with anger attacks: relationship between clinical response and receptor binding. Psychiatry Res 116: 151-161. Momosaki S, Hatano K , Kawasumi Y , Kato T, Hosoi R, Kobayashi K , et al. (2004) Rat-PET study without anesthesia: anesthetics modify the dopamine Di receptor binding in rat brain. Synapse 54: 207-213. Mufson EJ, Kroin JS, Sendera TJ, Sobreviela T (1999) Distribution and retrograde transport of trophic factors in the central nervous system: functional implications for the treatment of neurodegenerative diseases. Prog Neurobiol 57: 451-484. Narang N , Wamsley JK (1995) Time dependent changes in D A uptake sites, Di and D 2 receptor binding and mRNA after 6-OHDA lesions of the medial forebrain bundle in the rat brain. J Chem Neuroanat 9: 41-53. 138 Newton SS, Thome J, Wallace TL, Shirayama Y , Schlesinger L , Sakai N , et al. (2002) Inhibition of cAMP response element-binding protein or dynorphin in the nucleus accumbens produces an antidepressant-like effect. J Neurosci 22: 10883-10890. Nibuya M , Nestler EI, Duman RS (1996) Chronic antidepressant administration increases the expression of cAMP response element binding protein (CREB) in rat hippocampus. J Neurosci 16: 2365-2372. Nowak G, Zak J (1989) Repeated electroconvulsive shock (ECS) enhances striatal D - l dopamine receptor turnover in rats. Eur J Pharmacol 167: 307-308. Nymeyer L , Grossberg GT (2002) Delirium in a 75-year-old woman receiving ECT and levodopa. Convuls Ther 13: 114-116. Oldenhof J, Ray A , Vickery R, Van Tol H H (2001) SH3 ligands in the dopamine D 3 receptor. Cell Signal 13:411-416. Pandey GN, Pandey SC, Isaac L, Davis J (1992) Effect of electroconvulsive shock on 5-HT2 and ai-adrenoreceptors and phosphoinositide signalling system in rat brain. Eur J Pharmacol 226: 303-310. Patel N K , Bunnage M , Plaha P, Svendsen CN, Heywood P, Gi l l SS (2005) Intraputamenal infusion of glial cell line-derived neurotrophic factor in PD: a two-year outcome study. Ann Neurol 57: 298-302. Peroutka SJ, Snyder SH (1990) Long-term antidepressant treatment decreases spiroperidol-labeled serotonin receptor binding. Science 210: 88-90. Pridmore S, Pollard C (1996) Electroconvulsive therapy in Parkinson's disease: 30 month follow up. J Neurol Neurosurg Psychiatry 61: 693-700. Reches A , Wagner HR, Barkai AI, Jackson V , Yablonskaya-Alter E, Fahn S (1984) Electroconvulsive treatment and haloperidol: effects on pre- and postsynaptic dopamine receptors in rat brain. Psychopharmacology 83: 155-158. Ridray S, Griffon N , Mignon V , Souil E, Carboni S, Diaz J, et al. (1998) Coexpression of dopamine Di and D 3 receptors in islands of Calleja and shell of nucleus accumbens of the rat: opposite and synergistic functional interactions. Eur J Neurosci 10: 1676-1686. Sakai N , Thome J, Newton SS, Chen J, Kelz M B , Steffen C, et al. (2002) Inducible and brain region-specific CREB transgenic mice. Mol Pharmacol 61: 1453-1464. Sanchez C, Hyttel J (1999) Comparison of the effects of antidepressants and their metabolites on reuptake of biogenic amines and on receptor binding. Cell Mol Neurobiol 19: 467-489. Schallert T, Woodlee M T (2005) Orienting and Placing. In: The behavior of the laboratory rat: a handbook with tests. (Whishaw IQ, Kolb B, eds), pp 129-140. New York: Oxford University Press. Scott BW, Wojtowicz JM, Burnham W M (2000) Neurogenesis in the dentate gyrus of the rat following electroconvulsive shock seizures. Exp Neurol 165: 231-236. 139 Sershen H, Wolinsky T, Douyon R, Hashim A, Wiener HL, Lajtha A , et al. (1991) The effects of electroconvulsive shock on dopamine-1 and dopamine-2 receptor ligand binding activity in MPTP-treated mice. Neuroscience 3: 58-63. Shi L - H , Woodward DJ, Luo F, Anstrom K , Schallert T, Chang J-Y (2004) High-frequency stimulation of the subthalamic nucleus reverses limb-use asymmetry in rats with unilateral 6-hydroxydopamine lesions. Brain Res 1013: 98-106. Slaughter JR, Slaughter K A , Nichols D, Holmes SE, Martens MP (2001) Prevalence, clinical manifestations, etiology, and treatment of depression in Parkinson's disease. J Neuropsychiatry Clin Neurosci 13: 187-196. Slevin JT, Gerhardt GA, Smith CD, Gash D M , Kryscio R, Young B (2005) Improvement of bilateral motor functions in patients with Parkinson disease through the unilateral intraputaminal infusion of glial cell line-derived neurotrophic factor. J Neurosurg 102: 216-222. Smith SE, Sharp T (1997) Evidence that the enhancement of dopamine function by repeated electroconvulsive shock requires concomitant activation of Di-like and D2-like dopamine receptors. Psychopharmacology 133: 77-84. Staton RD, Enderle JD, Gerst JW (1986) The electroencephalographic pattern during electroconvulsive therapy IV. spectral energy distributions with methohexital, innovar and ketamine anesthesias. Clin E E G 17: 203-215. Strome E M , Jivan S, Doudet DJ (2005) Quantitative in vitro phosphor imaging using [3H] and [1 8F] radioligands: the effects of chronic desipramine treatment on serotonin 5-HT 2 receptors. J Neurosci Methods 141: 143-154. Tandberg E, Larsen JP, Aarsland D, Cummings JL (1996) The occurrence of depression in Parkinson's disease. A community-based study. Arch Neurol 53: 175-179. Tang SW, Seeman P, Kwan S (1981) Differential effect of chronic desipramine and amitriptyline treatment on rat brain adrenergic and serotonergic receptors. Psychiatry Research 4: 129-138. Thome J, Sakai N , Shin K - H , Steffen C, Zhang Y-J , Impey S, et al. (2000) cAMP response element-mediated gene transcription is upregulated by chronic antidepressant treatment. J Neuroscience 20: 4030-4036. Tillerson JL, Cohen A D , Philhower J, Miller G, Zigmond MJ , Schallert T (2001) Forced limb-use effects on the behavioral and neurochemical effects of 6-hydroxydopamine. J Neurosci 21: 4427-4435. Todd K G , McManus DJ, Baker GB (1995) Chronic administration of the antidepressants phenelzine, desipramine, clomipramine, or maprotiline decreases binding to 5-hydroxytryptamine2A receptors without affecting benzodiazepine binding sites in rat brain. Cellular and Molecular Neurobiology 15: 361-370. Vaidya V A , Siuciak JA, Du F, Duman RS (1999) Hippocampal mossy fiber sprouting induced by chronic electroconvulsive seizures. Neuroscience 89: 157-166. 140 van Kampen JM, Stoessl A J (2000) Effects of oligonucleotide antisense to dopamine D( l A) receptor messenger R N A in a rodent model of levodopa-induced dyskinesia. Neuroscience 98: 61-67. van Kampen JM, Stoessl A J (2003) Effects of oligonucleotide antisense to dopamine D 3 receptor mRNA in a rodent model of behavioural sensitization to levodopa. Neuroscience 116: 307-314. van Kampen JM, Hagg T, Robertson H A (2004) Induction of neurogenesis in the adult rat subventricular zone and neostriatum following dopamine D 3 receptor stimulation. Eur J Neurosci 19: 2377-2387. van Kampen JM, Robertson H A (2005) A possible role for dopamine D 3 receptor stimulation in the induction of neurogenesis in the adult rat substantia nigra. Neuroscience 136: 381-386. Wang L , Muramatsu S, Lu Y , Ikeguchi K , Fujimoto K, Okada T, et al. (2002) Delayed delivery of A A V - G D N F prevents nigral neurodegeneration and promotes functional recovery in a rat model of Parkinson's disease. Gene Ther 9: 381-389. Yates M , Leake A , Candy JM, Fairbairn AF , McKeith IG, Ferrier IN (1990) 5HT 2 receptor changes in major depression. Biol Psychiatry 27: 489-496. Yatham L N , Liddle PF, Dennie J, Shiah IS, Adam MJ , Lane CJ, et al. (1999) Decrease in brain serotonin 2 receptor binding in patients with major depression following desipramine treatment: a positron emission tomography study with fluorine-18-labeled setoperone. Arch Gen Psychiatry 56: 705-711. Yatham L N , Liddle PF, Shiah IS, Scarrow G, Lam RW, Adam M J , et al. (2000) Brain serotonin2 receptors in major depression. Arch Gen Psychiatry 57: 850-858. Yoshida K , Higuchi H , Kamata M , Yoshimoto M , Shimizu T, Hishikawa Y (1998) Single and repeated electroconvulsive shocks activate dopaminergic and 5-hydroxytryptaminergic neurotransmission in the frontal cortex of rats. Prog Neuropsychopharmacol Biol Psychiatry 22: 435-444. Zervas IM, Fink M (1992) ECT and delirium in Parkinson's disease. A m J Psychiatry 149: 1758. Zis AP, Nomikos GG, Brown EE, Damsma G, Fibiger HC (1992) Neurochemical effects of electrically and chemically induced seizures: an in vivo microdialysis study in the rat hippocampus. Neuropsychopharmacology 7: 189-195. Zis AP, Nomikos GG, Damsma G, Fibiger HC (1991) In vivo neurochemical effects of electroconvulsive shock studied by microdialysis in the rat striatum. Psychopharmacology 103: 343-350. 141 Appendix A THE USE OF ANAESTHESIA FOR REPEATED ECS TREATMENT IN THE RAT Introduction: Electroconvulsive shock (ECS) is a widely used animal model of the application of electroconvulsive therapy (ECT) to human patients as a treatment for psychiatric disorders. There is a major discrepancy, however, between the method of administration of ECS to rodents vs. ECT to humans, in that rodents usually receive ECS without anaesthesia, whereas in modern ECT, humans are always anaesthetized. In developing our protocol for the administration of ECS to rats, we wanted to model the clinical situation as closely as possible, and chose to administer ECS under anaesthesia. To investigate the use of anaesthesia for ECS, we administered a course of ECS treatment under three different anaesthetics, ketamine hydrochloride, a non-competitive N M D A antagonist dissociant, thiopental sodium, a short-acting barbiturate, and isoflurane gas, a volatile anaesthetic. We chose the three anaesthetics based on their applicability to clinical ECT (ketamine and thiopental are used routinely in humans), and their ease of use (isoflurane is a safe and widely used gas anaesthetic). Using a small number of subjects in each group, we examined the feasibility of using these anaesthetics, and their effects on seizure threshold and duration compared to ECS without anaesthesia. One of the most consistently reported effects of repeated ECS in rodents is increased 5-H T 2 receptor binding and mRNA in the frontal cortex (Kellar et al., 1981; Vetulani et al., 1981; Green et al., 1983; Vetulani et al., 1983; Stockmeier and Kellar, 1986; Pandey et al., 1992; Butler et al., 1993; Burnet et al., 1999). Interestingly, this effect appears to be opposite to the effects of successful antidepressant drug treatment in human patients (Attar-Levy et al., 1999; Yatham et al., 1999; Meyer et al., 2001; Mischoulon et al., 2002) and repeated ECS in non-human primates (Strome et al., 2005a). In rodents, too, chronic treatment with tricyclic antidepressants (TCAs) has repeatedly been shown to decrease cortical 5-HT 2 receptor binding (Cross and Horton, 1988; Newman et al., 1990; Peroutka and Snyder, 1990; Goodnough and Baker, 1994; Strome et al., 2005b). Although it has long been known that ECS in rodents has opposite effects on 5-HT 2 receptors, compared to the effects of most antidepressant drug treatments in rodents and humans, an explanation for this inconsistency has not been presented. One possible explanation may arise from the fact that rodents are typically not anaesthetized during ECS treatment. This lack of anaesthesia may impact the immediate effects of ECS, thereby influencing the downstream compensatory changes that accrue with repeated treatments. After evaluating the feasibility of using anaesthesia for ECS, we chose ketamine as our anaesthetic of choice for all further studies. To support our decision to use ketamine, and to further corroborate our in vitro autoradiographic technique described in Chapter 3, we also examined frontal cortex 5-HT 2 receptor binding in animals treated with ECS or sham under ketamine. Materials and Methods: Adult male Sprague-Dawley rats were the subjects of this experiment. Subjects (n = 4-10 per group) were given ECS as described in Chapters 5 and 6 every day for 10 days, under one of four conditions: 1) ketamine HCI (80 mg/kg i.p.), 2) thiopental sodium (20 mg/kg i.v.), 3) isoflurane (4% in 0 2 as needed), or 4) without anaesthesia, hereafter referred to as the KET, 142 THIO, ISO and NO groups, respectively. Control rats were given the same treatment and electrodes were placed, but no current was administered. Animals in the K E T group were sacrificed by decapitation 24 hours after the last treatment. The brains were removed and quickly frozen in isopentane cooled with dry ice, and stored at -80°C until sectioning. Twenty micron coronal sections were cut, thaw-mounted onto microscope slides, and the slides were stored at -80°C until the receptor binding assay was performed. Serotonin 5-HT 2 receptor binding with [18F]setoperone was performed on sections through the frontal cortex as described in (Strome et al., 2005b). Results: Effects of Anaesthesia and ECS on Body Weight The mean weight of each group was not significantly different, nor was there a significant effect of ECS on weight in any of the groups. A l l groups were significantly heavier at the end of the course of ECS compared to the beginning (Two-way repeated measures A N O V A : Main effect of time, F i , 6 6 = 8.16, p < 0.01; data not shown). Stimulus Properties and Seizure Characteristics Due to the difficulty of finding an adequate tail vein to inject into each day, the THIO group did not receive ECS every day for 10 days. Animals in both the sham and ECS conditions did, however, receive at least 8 treatments, with at least the last 4 treatments administered on consecutive days, and no animal ever went more than 2 days without treatment. Figure A . l shows the relationship between the stimulus intensity and the course of treatment under all four conditions. In order to continue to elicit similar length seizures in the anaesthetized groups, stimulus intensity (charge dose: #mC administered = stimulus duration x current x pulse width x frequency) was increased as the course of treatment progressed. This increase was significant in the KET, ISO and THIO groups (linear regression, F-test for slopes, p < 0.05 for all), but not in the NO group (p > 0.05) (Fig. A . l ) . The slopes of all four lines were significantly different, but the mean charge doses were not significantly different between the KE T and THIO groups (Tukey test, p > 0.05). Seizure length (mean 9-12 s for anaesthetic groups, 75 s for non-anaesthetized) did not change significantly as the course of ECS progressed (linear regression, F-test for slopes, p > 0.05 for all), but animals in the NO group had significantly longer seizures than all other groups (Fig. A.2; A N C O V A for intercepts: F 3 , 3 5 = 646.35, p < 0.01). A l l of the animals in this study experienced a motor seizure after receiving the ECS stimulus each day, however not every animal experienced a fully generalized seizure with tonic hind limb extension (THLE) after each ECS treatment. We have adopted the terminology of (Andrade et al., 2002)and colleagues (2002) to describe the types of motor seizures elicited after ECS in rats. In the K E T and THIO groups, the probability of eliciting a completely generalized seizure with THLE (Type 3 seizure) increased as the course of ECS progressed (Fig. A.3). In the NO group, the animals invariably exhibited THLE, whereas animals in the ISO group usually did not exhibit THLE as part of the motor seizure (Fig. A.3). A l l seizures without THLE were similar to the Type 2 seizures described by Andrade and colleagues (2002) i.e. tonic-clonic seizures involving mainly the head and forelimbs. Although we did not measure it directly, the recovery from the seizure (the time it took for the animals to move around spontaneously with good balance and coordination) was about the same for all three anaesthesia groups, but was much faster for the animals treated without anaesthesia. 143 1 2 3 4 5 6 7 8 9 10 E C S session Figure A.l: Mean charge dose (± SEM) required to elicit a seizure during the course of ECS under one of four conditions. - • - I S O —±— K E T - • - T H I O NO T-~l 1 1 1 I I 1 1 I 1 2 3 4 5 6 . 7 8 9 10 E C S session Figure A.2: Mean seizure length (± SEM) during the course of ECS under one of four conditions. 144 i i i — i — i — i — i — i — i — r 1 2 3 4 5 6 7 8 9 10 ECS session Figure A.3: Probability of eliciting a seizure with tonic hindlimb extension (THLE) under four different conditions as the course of E C S progressed. 5-HT2 Receptor Binding in Ketamine-Treated Animals Repeated ECS treatment under ketamine anaesthesia significantly increased 5-HT2 receptor binding in the frontal cortex (Student's t-test: t= 3.08, p < 0.008; Figs. A.4, A.5). O) 150-i SHAM ECS (n = 9) (n = 9) Figure A.4: Repeated E C S treatment under ketamine increases frontal cortex [18F]setoperone binding. (* p < 0.05). 145 ECS SHAM Figure A.5: Frontal cortex 5-HT 2 receptor binding in rats treated with ECS or sham under ketamine anaesthesia as measured by [18F]setoperone. 4^ Discussion: In this study, we evaluated the use of three different anaesthetics for repeated ECS treatment in rodents, compared to the more common method of administering ECS to these animals without anaesthesia. We found that, in general, the use of anaesthesia for ECS administration influences the seizure characteristics, requiring higher charge doses to continue to elicit similar length seizures as the course of treatment progressed (i.e. increased seizure threshold), and decreasing both seizure length, and the probability of eliciting a fully generalized seizure. In addition, there were feasibility and applicability issues of using anaesthesia for repeated ECS in rodents. Thiopental is a short-acting barbiturate, the class of anaesthetic most commonly used for human ECT (Wagner et al., 2005). However, these drugs require intravenous administration, which was difficult to accomplish non-invasively on a daily basis in the rat. Isoflurane is a widely used volatile anaesthetic, but it has anticonvulsant properties (Eger, 1985), and is generally not used for ECT, and the subjects in this group therefore had a low probability of a fully-generalized seizure (Fig. A.3). Based on these factors, its relatively mild anticonvulsant effects, and the fact that it is commonly used for human ECT, we chose ketamine as our anaesthetic of choice for all future studies. We also examined frontal cortex 5-HT 2 receptor binding in the K E T group, and, as in many other reports, found that repeated ECS treatment increases binding to this receptor. While this upregulation of 5-HT 2 receptors is unexpected based on the effects of antidepressant drugs in humans and animals and repeated ECS in non-human primates, it may be explained by technical differences in the administration of the treatment (differences in schedule, current, or electrode placement), or by species differences in metabolism, neuroanatomy, or neurochemistry. In conclusion, we have shown here the feasibility of using ketamine anaesthesia for repeated ECS in rats, and that, as in ECS in rodents without anaesthesia, ECS under ketamine upregulates frontal cortex 5-HT 2 receptor binding. 147 References: Andrade C, Kurinji S, Sudha S, Suresh J (2002) Effects of pulse amplitude, pulse frequency, and stimulus duration on seizure threshold: a laboratory investigation. J ECT 18: 144-148. Attar-Levy D, Martinot JL, Blin J, Dao-Castellana M H , Crouzel C, Mazoyer B, et al. (1999) The cortical serotonin2 receptors studied with positron-emission tomography and [18F]-setoperone during depressive illness and antidepressant treatment with clomipramine. Biol Psychiatry 45: 180-186. Burnet PWJ, Sharp T, LeCorre S M , Harrison PJ (1999) Expression of 5-HT receptors and the 5-HT transporter in rat brain after electroconvulsive shock. Neurosci Lett 277: 79-82. Butler M O , Morinobu S, Duman RS (1993) Chronic electroconvulsive seizures increase the expression of serotonin2 receptor mRNA in rat frontal cortex. J Neurochem 61: 1270-1276. Cross JA, Horton R (1988) Effects of chronic oral administration of the antidepressants desmethylimipramine and zimelidine on rat cortical G A B A B binding sites: a comparison with 5-HT 2 binding site changes. J Pharmacol 93: 331-336. Eger EI (1985) Isoflurane (Forane): a compendium and reference. Madison: Anaquest. Goodnough DB, Baker GB (1994) 5-hydroxytryptamine2 and (3-adrenergic receptor regulation in rat brain following chronic treatment with desipramine alone and in combination. J Neurochem 62: 2262-2268. Green AR, Heal DJ, Johnson P, Laurence BE, Nimgaonkar V (1983) Antidepressant treatments: effects in rodents on dose-response curves of 5-hydroxytryptamine- and dopamine-mediated behaviours and 5-HT 2 receptor number in frontal cortex. Br J Pharmacol 80: 377-385. Kellar K J , Cascio CS, Butler JA, Kurtzke R (1981) Differential effects of electroconvulsive shock and antidepressant drugs on serotonin-2 receptors in rat brain. Eur J Pharmacol 69: 515-518. Meyer J, Kapur S, Eisfeld B, Brown G M , HouleS, DaSilva J, et al. (2001) The effect of paroxetine on 5 -HT2A receptors in depression: an [ F]setoperone PET imaging study. Am J Psychiatry 158: 78-85. Mischoulon D, Dougherty DD, Bottonari K, Gresham RL, Spnawalla SB, Fischman AJ, et al. (2002) An open pilot study of nefazodone in depression with anger attacks: relationship between clinical response and receptor binding. Psychiatry Res 116: 151-161. Newman M E , Drummer D, Lerer B (1990) Single and combined effects of desimipramine and lithium on serotonergic receptor number and second messenger function in rat brain. J Pharmacol Exp Therap 252: 826-831. Pandey G N , Pandey SC, Isaac L , Davis J (1992) Effect of electroconvulsive shock on 5 -HT2 and ai-adrenoreceptors and phosphoinositide signalling system in rat brain. Eur J Pharmacol 226: 303-310. 148 Peroutka SJ, Snyder SH (1990) Long-term antidepressant treatment decreases spiroperidol-labeled serotonin receptor binding. Science 210: 88-90. Stockmeier CA, Kellar K J (1986) In vivo regulation of the serotonin-2 receptor in rat brain. Life Sci 38: 116-127. Strome E M , Clark C M , Zis AP, Doudet DJ (2005a) Electroconvulsive shock decreases binding to 5-HT2 receptors in nonhuman primates: an in vivo positron emission tomography study with [18F]setoperone. Biol Psychiatry 57: 1004-1010. Strome E M , Jivan S, Doudet DJ (2005b) Quantitative in vitro phosphor imaging using [3H] and [1 8F] radioligands: the effects of chronic desipramine treatment on serotonin 5-HT2 receptors. J Neurosci Methods 141: 143-154. Vetulani J, Lebrecht TJ, Pile A (1981) Enhancement of responsiveness of the central serotonergic system and serotonin-2 receptor density in rat frontal cortex by electroconvulsive treatment. Eur J Pharmacol 76: 81-85. Vetulani J, Szpak J, Pile A (1983) Spaced electroconvulsive treatment: effects on responses associated with oc2- and 5-HT2-receptors. J Pharm Pharmacol 35: 326-328. Wagner K J , Mollenberg O, Rentrop M , Werner C, Kochs EF (2005) Guide to anaesthetic selection for electroconvulsive therapy. CNS Drugs 19: 745-758. Yatham L N , Liddle PF, Dennie J, Shiah IS, Adam MJ , Lane CJ, et al. (1999) Decrease in brain serotonin 2 receptor binding in patients with major depression following desipramine treatment: a positron emission tomography study with fluorine- 18-labeled setoperone. Arch Gen Psychiatry 56: 705-711. 149 Appendix B THE EFFECTS OF REPEATED ECS UNDER KETAMINE ANAESTHESIA ON BDNF mRNA EXPRESSION Introduction: One of the most widely reported effects of repeated electroconvulsive shock (ECS) on the rodent brain is the upregulation of brain-derived neurotrophic factor (BDNF) mRNA expression in limbic regions (Nibuya et al., 1995; Smith et al., 1997; Zetterstrom et al., 1998; Dias et al., 2003; Jacobsen and Mork, 2004). However, most studies of the effects of repeated ECS treatment in rodents are performed without anaesthesia, whereas we have opted to use ketamine anaesthesia in our model. In addition, no reports have been made on the effects of repeated ECS treatment on BDNF mRNA expression in the striatum, our primary region of interest in examining the mechanism of action of electroconvulsive therapy in Parkinson's disease. In this study, we investigated the effects of repeated ECS treatment under ketamine anaesthesia on BDNF mRNA expression in the piriform cortex, a limbic brain region, and striatum of unilateral 6-hydroxydopamine (6-OHDA)-lesioned rats. Materials and Methods: Repeated ECS or sham treatment (n = 7 per group) was administered under ketamine anaesthesia to unilateral 6-OHDA-lesioned rats as described in Chapters 5 and 6. Twenty-four hours after the last treatment, the animals were sacrificed by decapitation, their brains frozen in isopentane cooled with dry ice, and stored in a -80°C freezer until sectioning. Twenty micron sections were thaw-mounted onto glass slides in pairs, with one section from a sham-treated and one section from an ECS-treated animal per slide, alternating the placement on the slide of sham-and ECS-treated animals (top or bottom). In situ hybridization for BDNF mRNA was performed as described by (Kobayashi et al., 1996). Results: We were unable to detect any hybridization signal in the striatum. In the piriform cortex, ECS treatment significantly increased BDNF mRNA expression compared to sham-treated controls (Unpaired T-test, tn = 2.82, p < 0.02; Fig. B . l ) . Discussion: As in previous reports (Hofer et al., 1990; Maisonpierre et al., 1990) no hybridization signal was detected in the striatum of either ECS- or sham-treated rats. Protein levels of BDNF in the striatum are detectable, and we (Strome et al., 2006) and others (Angelucci et al., 2002; Altar et al., 2003) have shown that striatal BDNF protein increases after repeated ECS treatment. The observation that BDNF protein is present in the striatum, while mRNA is absent has led to the suggestion that striatal BDNF is anterogradely transported from other brain regions, and Altar and colleagues have shown that, indeed, this is the case, with most striatal BDNF arising from frontal cortical projections, and a small amount arising from nigrostriatal projections (Altar etal., 1997). In this study, we also found that, as in animals administered ECS without anaesthesia, ECS under ketamine anaesthesia increases BDNF mRNA within the piriform cortex. This brain 150 region is part of the limbic system and while it has traditionally been thought of as a sensory region, based on its connectivity with the olfactory system, more recent studies suggest that it functions more as associative cortex, rather than primary sensory cortex (Johnson et al., 2000). One of the most interesting features of the piriform cortex is the fact that it is the only site in the rat brain where small amounts of y-aminobutyric acid (GABA) antagonists can elicit tonic-clonic seizures (Gale et al., 1992), and this discovery has led the area to be investigated thoroughly for its role in seizures, kindling, and ECS. In conclusion, this investigation showed that one of the typical effects of repeated ECS treatment on BDNF mRNA expression, its upregulation in the piriform cortex, is not negatively impacted by the use of ketamine as an anaesthetic for ECS. In addition, we identified that BDNF mRNA expression was not detectable in the striatum of unilateral 6-OHDA-lesioned rats, after either ECS or sham treatment. Figure B . l : The effects of repeated E C S treatment under ketamine anaesthesia on BDNF mRNA expression in the piriform cortex. ECS treatment significantly increased BDNF mRNA compared to sham (*p < 0.02) 151 References: Altar CA, Cai N , Bilven T, Juhasz M , Conner JM, Acheson A , et al. (1997) Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 389: 856-860. Altar CA, Whitehead RE, Chen R, Wortwein G, Madsen T M (2003) Effects of electroconvulsive seizures and antidepressant drugs on brain-derived neurotrophic factor protein in rat brain. Biol Psychiatry 54: 703-709. Angelucci F, Aloe L , Jimenez-Vasquez P, Mathe A A (2002) Electroconvulsive stimuli alter the regional concentrations of nerve growth factor, brain-derived neurotrophic factor, and glial cell line-derived neurotrophic factor in adult rat brain. J ECT 18: 138-143. Dias BG, Banerjee SB, Duman RS, Vaidya V A (2003) Differential regulation of brain derived neurotrophic factor transcripts by antidepressant treatments in the adult rat brain. Neuropharmacology 45: 553-563. Gale K , Zhong P, Miller LP, Murray TF (1992) Amino acid neurotransmitter interactions in 'area tempestas': an epileptogenic trigger zone in the deep prepiriform cortex. Epilepsy Res Suppl 8: 229-234. Hofer M , Pagliusi SR, Hohn A, Leibrock J, Barde Y A (1990) Regional distribution of brain-derived neurotrophic factor mRNA in the adult mouse brain. E M B O J 9: 2459-2464. Jacobsen JP, Mork A (2004) The effect of escitalopram, desipramine, electroconvulsive seizures and lithium on brain-derived neurotrophic factor mRNA and protein expression in the rat brain and the correlation to 5-HT and 5-HIAA levels. Brain Res 1024: 183-192. Johnson D M , Illig KR, Behan M , Haberly L B (2000) New features of connectivity in piriform cortex visualized by intracellular injection of pyramidal cells suggest that "primary" olfactory • cortex functions like "association" cortex in other sensory systems. J Neurosci 20: 6974-6982. Kobayashi NR, Bedard A M , Hincke MT, Tetzlaff W (1996) Increased expression of BDNF and trkB mRNA in rat facial motoneurons after axotomy. Eur J Neurosci 8: 1018-1029. Maisonpierre PC, Belluscio L , Friedman B, Alderson RF, Wiegand SJ, Furth M E , et al. (1990) NT-3, BDNF, and NGF in the developing rat nervous system: parallel as well as reciprocal patterns of expression. Neuron 5: 501-509. Nibuya M , Morinobu S, Duman RS (1995) Regulation of BDNF and trkB mRNA in rat brain by chronic electroconvulsive seizure and antidepressant drug treatments. J Neurosci 15: 7539-7547. Smith M A , Zhang L X , Lyons WE, Mamounas L A (1997) Anterograde transport of endogenous brain-derived neurotrophic factor in hippocampal mossy fibers. NeuroReport 8: 1829-1834. Strome E M , Zis AP, Doudet DJ (2006) The effects of electroconvulsive shock on GDNF, BDNF, and FGF-2 concentrations in the 6-OHDA-lesioned rat brain. Submitted. Zetterstrom TSC, Pei Q, Grahame-Smith D G (1998) Repeated electroconvulsive shock extends the duration of enhanced gene expression for BDNF in rat brain compared with a single administration. Molecular Brain Research 57: 106-110. 152