oscillatory activity in the basal ganglia of · pdf fileoscillatory activity in the basal...

266
OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE by Moran Weinberger A thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy Graduate Department of Physiology University of Toronto © Copyright by Moran Weinberger 2009

Upload: lamnga

Post on 28-Mar-2018

217 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE

by

Moran Weinberger

A thesis submitted in conformity with the requirements

for the Degree of Doctor of Philosophy

Graduate Department of Physiology

University of Toronto

© Copyright by Moran Weinberger 2009

Page 2: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH

PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department of

Physiology, University of Toronto.

ABSTRACT

Parkinson’s disease (PD) is a movement disorder that is of basal ganglia origin. It is

characterized by a severe loss of dopaminergic input to the striatum and symptoms such

as bradykinesia, rigidity and tremor. There is growing evidence that PD is associated with

pathological synchronous oscillatory activity in the basal ganglia, which primarily occurs

in the 11-30 Hz range, the so-called beta band. The aim of this project was to better

understand the oscillatory activity recorded from the basal ganglia of PD patients and to

elucidate the significance of this activity in PD. To do this, neuronal firing and local field

potentials (LFPs) were recorded from the subthalamic nucleus (STN) and globus pallidus

internus (GPi) of PD patients undergoing stereotactic neurosurgery for implantation of

therapeutic deep brain stimulation electrodes. Beta oscillatory LFP activity in the STN

and GPi was found to be coherent with, and reflect to a certain degree, rhythmic activity

in a population of local neurons. I have demonstrated for the first time that the degree of

beta oscillatory firing in the STN, which is maximal in the motor portion, correlates with

the patients’ benefit from dopaminergic medications, but not with baseline motor deficits.

My study has also established that beta oscillatory firing in the STN does not positively

correlate with the patients’ tremor scores and that during periods of tremor patients tend

to have less beta oscillatory firing and increased neuronal oscillatory firing at the tremor

frequency. Temporal examination of the LFPs recorded during periods of intermittent

resting tremor revealed that stronger tremor is associated with increased LFP power in

the low gamma range (35-55 Hz) and there is a decrease in the ratio of beta to gamma

coherence. Similarly, a change in balance between oscillatory activities was observed

during levodopa-induced dyskinesias. Finally, when the oscillatory activity in the GPi of

PD patients was compared to that in dystonia I found that in dystonia, oscillatory LFP

activity is less likely to reflect the neuronal firing. These findings indicate that beta

oscillatory activity in the basal ganglia might reflect the degree of dopamine deficiency in

ii

Page 3: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

the striatum and that the relative strength of oscillatory rhythms may play an important

role in mediating the pathological features in PD.

iii

Page 4: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

ACKNOWLEDGEMENTS

I am very grateful for having Dr. Jonathan Dostrovsky as my supervisor. Without his mentoring,

patience and reassurance, this work would not have been possible. As an international student, I

owe him special thanks for opening the door for me, and giving me the opportunity to learn and

work in his laboratory. Being part of this research group made me feel like home. I am also

indebted to Dr. William Hutchison for his invaluable assistance, guidance and friendship

throughout these years, and to Dr. William MacKay for his knowledge and support. Their help

was crucial for the progression of this work. I would like to express my gratitude to Julie

Weedmark for guiding me at the beginning of this journey.

In addition, other members of the research group have contributed to this work by providing their

help: Neil Mahant, Clement Hamani, Andres Lozano, Mojgan Hodaie, Elena Moro and Anthony

Lang. Special thanks to the patients who participated in these studies. Their strength is also part

of this work.

I would also like to thank the following individuals for their precious friendship and

encouragement: Myriam Lafreniere-Roula, Nicole Carmichael, Michelle Olah, Yotam Blech-

Hermoni, Ravit Saada, Orit Zamir, Joyce Tang, Helen Belina, Adrian Fawcett, Ian Prescott,

Catherine Trepanier, Maria Brunello, Joseph Neimat, Kim Tanzer, Yafit Shimoni and Monique

Newton. Our friendship helped me through difficult times.

To the love of my life Lior Opek, I am fortunate to have you as my husband. You were always

there for me during this long process, understanding and supporting me. I couldn’t have done this

without you.

This work is dedicated to my parents, Dalia and Zeev who never wanted me to be so far away,

but I know they are proud.

iv

Page 5: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

Permission to reprint text, figures and tables from the following sources has been granted:

Weinberger M, Hutchison WD, Lozano AM, Hodaie M, Dostrovsky JO. Increased gamma oscillatory activity in the subthalamic nucleus during tremor in Parkinson’s disease patients. J Neurophysiol 101(2):789-802, 2009. Weinberger M, Hamani C, Hutchison WD, Moro E, Lozano AM, Dostrovsky JO. Pedunculopontine Nucleus Microelectrode Recordings in Movement Disorders Patients. Exp Brain Res 188(2):165-174, 2008. Weinberger M, Mahant N, Hutchison WD, Lozano AM, Moro E, Hodaie M, Lang AE, Dostrovsky JO. Beta oscillatory activity in the subthalamic nucleus and its relation to dopaminergic response in Parkinson’s disease. J Neurophysiol 96(6):3248-3256, 2006. Levy R, Lozano AM, Hutchison WD, Dostrovsky JO. Dual microelectrode technique for deep brain stereotactic surgery in human. Neurosurgery. 60(4 Suppl 2):277-83; discussion 283-4, 2007.

v

Page 6: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

TABLE OF CONTENTS: ABSTRACT………………………………………………………………………….…..ii ACKNOWLEDGEMENTS……………………………………………………...……..iv LIST OF TABLES……………………………………………………………………....x LIST OF FIGURES……………………………………………………………………..x ABBREVIATIONS……………………………………………………………………..xi CHAPTER 1 – GENERAL INTRODUCTION ............................................................. 1

1.1 - Parkinson’s disease .............................................................................................. 2 1.1.1 - Etiology of Parkinson’s disease ..................................................................... 2 1.1.2 - The clinical features of Parkinson’s disease .................................................. 5

1.1.2.a - Akinesia and bradykinesia ....................................................................... 5 1.1.2.b - Rigidity .................................................................................................... 6 1.1.2.c - Tremor...................................................................................................... 7 1.1.2.d - Postural instability ................................................................................... 8 1.1.2.e - Other motor abnormalities ....................................................................... 8 1.1.2.f - Non-motor features .................................................................................. 9

1.1.3 - Dopamine therapy in Parkinson’s disease ................................................... 10 1.1.4 - Deep brain stimulation for Parkinson’s disease........................................... 12

1.2 - The basal ganglia: anatomy and physiology...................................................... 15 1.2.1 - Functional organization and circuitry .......................................................... 16

1.2.1.a - Connectional architecture ...................................................................... 16 1.2.1.b - Differential effects of dopamine ............................................................ 18 1.2.1.c - Parallel organization .............................................................................. 19

1.2.2 - The striatum ................................................................................................. 19 1.2.3 - Subthalamic nucleus .................................................................................... 25 1.2.4 - Globus pallidus ............................................................................................ 32

1.2.4.a - Globus pallidus externus........................................................................ 33 1.2.4.b - Globus pallidus internus ........................................................................ 34

1.2.5 - Substantia nigra............................................................................................ 36 1.2.5.a - Substantia nigra pars compacta.............................................................. 36 1.2.5.b - Substantia nigra pars reticulata .............................................................. 37

1.3 - Models of basal ganglia function in relation to Parkinson’s disease ................. 39 1.3.1 - The rate model ............................................................................................. 39 1.3.2 - The centre-surround model .......................................................................... 42 1.3.3 - The oscillatory model .................................................................................. 44 1.3.4 - Tremor models ............................................................................................. 49

1.4 - Electrophysiological and behavioral findings in Parkinson’s disease patients and animal models of PD.......................................................................................... 52

1.4.1 - Single unit recordings and inactivation studies ........................................... 54 1.4.1.a - Striatum.................................................................................................. 54 1.4.1.b - Subthalamic nucleus .............................................................................. 55 1.4.1.c - Globus pallidus ...................................................................................... 58 1.4.1.d - Substantia nigra...................................................................................... 60 1.4.1.e - Changes in basal ganglia-related structures........................................... 62

vi

Page 7: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

1.4.2 - Local field potential recordings ................................................................... 65 1.4.3 - Oscillatory activity in relation to voluntary movements and dopaminergic

medications .................................................................................................. 67 1.4.4 - Oscillatory activity in relation to dyskinesias.............................................. 70 1.4.5 - Oscillatory activity in relation to tremor...................................................... 71

1.5 - Aim of present studies ....................................................................................... 73 CHAPTER 2 - GENERAL METHODS ....................................................................... 75

2.1 - Patients and consent........................................................................................... 75 2.2 - Intraoperative neuronal recordings .................................................................... 75

2.2.1 - Operative procedures ................................................................................... 75 2.2.2 - Physiological targeting................................................................................. 76

2.2.2.a - Subthalamic nucleus .............................................................................. 76 2.2.2.b - Globus pallidus internus ........................................................................ 78 2.2.2.c - Pedunculopontine nucleus ..................................................................... 79

2.2.3 - Microelectrode setup.................................................................................... 80 2.2.4 - Single unit and local field potential recordings ........................................... 81

2.3 - Data analysis ...................................................................................................... 82 2.3.1 - Spectral analysis of neuronal discharges and local field potentials............. 82

2.3.1.a - Power spectrum...................................................................................... 82 2.3.1.b - Coherence and cross-correlation............................................................ 84 2.3.1.c - Time frequency analysis ........................................................................ 85

2.3.2 - Statistical comparisons................................................................................. 86 CHAPTER 3 - BETA OSCILLATORY ACTIVITY IN THE SUBTHALAMIC

NUCLEUS AND ITS RELATION TO DOPAMINERGIC RESPONSE ........ 87 3.1 - Introduction........................................................................................................ 88 3.2 - Methods.............................................................................................................. 89 3.3 - Results................................................................................................................ 92

3.3.1 - Coherence between neuronal and beta LFP oscillatory activity in the STN 92 3.3.2 - Incidence of neuronal beta oscillatory activity is higher in the dorsal STN 93 3.3.3 - LFP beta oscillations are greatest in the dorsal STN ................................... 93 3.3.4 - Neuronal oscillations correlate with levodopa response.............................. 95

3.4 - Discussion........................................................................................................ 101 CHAPTER 4 - THE RELATIONSHIP BETWEEN SUBTHALAMIC

OSCILLATORY ACTIVITY AND TREMOR ................................................ 107 4.1 - Introduction...................................................................................................... 108 4.2 - Methods............................................................................................................ 110 4.3 - Results.............................................................................................................. 112

4.3.1 - Incidence of neurons with tremor and/or beta oscillatory firing is higher in the dorsal STN ........................................................................................... 112

4.3.2 - Neurons with tremor-frequency oscillations are coherent with the LFP at the beta frequencies ......................................................................................... 113

4.3.3 - Coherence between neuronal firing and LFP varies over time.................. 114 4.3.4 - Neurons with tremor-frequency oscillations are not constantly correlated

with tremor................................................................................................. 115 4.3.5 - LFP oscillatory activity in the 35-55 Hz gamma frequency band is increased

during tremor ............................................................................................. 116

vii

Page 8: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

4.3.6 - Altered balance between beta and gamma rhythms during periods of stronger tremor........................................................................................... 117

4.4 - Discussion........................................................................................................ 126 4.4.1 - Tremor-frequency neuronal activity: relation to tremor ............................ 126 4.4.2 - Tremor-frequency neuronal activity: relation to beta oscillations............. 127 4.4.3 - Low-gamma LFP oscillatory activity: relation to tremor .......................... 129

CHAPTER 5 - OSCILLATORY ACTIVITY IN THE GLOBUS PALLIDUS: COMPARISON BETWEEN PARKINSON’S DISEASE AND DYSTONIA 133

5.1 - Introduction...................................................................................................... 134 5.2 - Methods............................................................................................................ 136 5.3 - Results.............................................................................................................. 139

5.3.1 - Firing rates of GPi neurons ........................................................................ 140 5.3.2 - Coherence between neuronal firing and oscillatory LFPs in PD patients . 140 5.3.3 - The averaged beta LFP power in the GPi is closely related to the percentage

of correlated neurons.................................................................................. 141 5.3.4 - Neurons are less correlated to oscillatory LFPs in dystonia patients......... 141 5.3.5 - Changes in GPi oscillatory activity during levodopa-induced dyskinesias142

5.4 - Discussion........................................................................................................ 148 5.4.1 - Methodological constraints........................................................................ 148 5.4.2 - GPi firing rates........................................................................................... 148 5.4.3 - GPi LFP oscillatory activity and its relationship to neuronal discharges in

PD .............................................................................................................. 150 5.4.4 - Modulation of GPi oscillatory activity during levodopa-induced dyskinesias

.................................................................................................................... 152 5.4.5 - GPi oscillatory activity in dystonia and its comparison to PD .................. 154

CHAPTER 6 - GENERAL SUMMARY AND SIGNIFICANCE ............................ 156 6.1 - Oscillatory firing in the basal ganglia and its relationship to the local field

potential and dopamine .................................................................................... 156 6.2 - Spatial distribution of beta oscillations............................................................ 157 6.3 - The oscillatory neuronal population ................................................................ 158 6.4 - The relation of oscillatory activity to clinical aspects in PD patients.............. 159 6.5 - Implications to basal ganglia models............................................................... 161

APPENDIX - RECORDINGS FROM THE PEDUNCULOPONTINE NUCLEUS................................................................................................................................ 166

A.1 - Literature overview.......................................................................................... 167 A.1.1 - Functional anatomy of the PPN.................................................................. 167A.1.2 - Regulation of locomotion and gait in the PPN........................................... 169 A.1.3 - PPN and Parkinson’s disease...................................................................... 171 A.1.4 - PPN as a surgical targted for PD ................................................................ 173

A.2 - Introduction to the study .................................................................................. 175 A.3 - Methods............................................................................................................ 176 A.4 - Results.............................................................................................................. 180

A.4.1 - Action potentials and firing properties ....................................................... 180 A.4.2 - Movement responses .................................................................................. 182 A.4.3 - Microstimulation in the PPN region........................................................... 185 A.4.4 - LFP oscillatory activity in the PPN region................................................. 185

viii

Page 9: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

A.4.5 - Comparison of PD vs. PSP......................................................................... 187 A.5 - Discussion........................................................................................................ 187

REFERENCE LIST.......................................................................................................191

ix

Page 10: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

LIST OF TABLES Table 3.T1……………………………………………………………………………..…90 Table 4.T1……………………………………………………………………………....111 Table 4.T2………………………………………………………………………………114 Table 5.T1………………………………………………………………………………137 Table A.T1……………………………………………………………………………...177 Table A.T2……………………………………………………………………………...183 LIST OF FIGURES Figure 1.F1………………………………………………………………………………18 Figure 1.F2………………………………………………………………………………26 Figure 1.F3………………………………………………………………………………43 Figure 1.F4………………………………………………………………………………46 Figure 2.F1………………………………………………………………………………78 Figure 2.F2………………………………………………………………………………79 Figure 2.F3………………………………………………………………………………81 Figure 3.F1………………………………………………………………………………96 Figure 3.F2………………………………………………………………………………97 Figure 3.F3………………………………………………………………………………98 Figure 3.F4………………………………………………………………………………99 Figure 3.F5……………………………………………………………………………..100 Figure 3.F6……………………………………………………………………………..101 Figure 4.F1……………………………………………………………………………..119 Figure 4.F2……………………………………………………………………………..120 Figure 4.F3……………………………………………………………………………..121 Figure 4.F4……………………………………………………………………………..122 Figure 4.F5……………………………………………………………………………..123 Figure 4.F6……………………………………………………………………………..123 Figure 4.F7……………………………………………………………………………..124 Figure 4.F8……………………………………………………………………………..125 Figure 5.F1……………………………………………………………………………..144 Figure 5.F2……………………………………………………………………………..145 Figure 5.F3……………………………………………………………………………..146 Figure 5.F4……………………………………………………………………………..146 Figure 5.F5……………………………………………………………………………..147 Figure 6.F1……………………………………………………………………………..165 Figure A.F1…………………………………………………………………………….179 Figure A.F2…………………………………………………………………………….181 Figure A.F3…………………………………………………………………………….182 Figure A.F4…………………………………………………………………………….184 Figure A.F5…………………………………………………………………………….186

x

Page 11: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

ABBREVIATIONS Anterior commissure AC

2-aminomethyl-phenylacetic acid AMPA

Basal ganglia BG

Centromedian nucleus of the thalamus CM

Cyclic adenosine monophospate cAMP

Deep brain stimulation DBS

Electromyography EMG

Entopeduncular nucleus EPN

Excitatory postsynaptic potential EPSP

Gamma-aminobutyric acid GABA

Globus pallidus GP

Globus pallidus externus GPe

Globus pallidus internus GPi

6-hydroxydopamine 6-OHDA

Inhibitory postsynaptic potential IPSP

Internal capsule IC

Interspike interval ISI

Local field potential LFP

Long-term potentiation LTP

Long-term depression LTD

Magnetic resonance imaging MRI

Magnetoencephalography MEG

Medium spiny neurons MSNs

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

1-methyl-4-phenylpyridinium MPP+

Nicotinamide Adenine Dinucleotide Phosphate-oxidase NADPH

N-methyl-d-aspartate NMDA

Nucleus ventralis intermedius Vim

Nucleus ventro-oralis posterior Vop

Nucleus ventro-oralis anterior Voa

xi

Page 12: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

Optic tract OT

Posterior commissure PC

Parkinson’s disease PD

Positron emission tomography PET

Parafascicular nucleus of the thalamus Pf

Pedunculopontine nucleus PPN

Pedunculopontine nucleus pars compacta PPNc

Pedunculopontine nucleus pars dissipatus PPNd

Progressive supranuclear palsy PSP

Rapid eye movement sleep behaviour disorder RBD

Superior colliculus SC

Substantia nigra SN

Substantia nigra pars compacta SNc

Substantia nigra pars reticulata SNr

Subthalamic nucleus STN

Supplementary motor area SMA

Tonically active neurons TANs

Toronto Western Spasmodic Torticollis Rating Scale TWSTRS

Transcranial magnetic stimulation TMS

Ventral tegmenta area VTA

Unified Parkinson’s Disease Rating Scale UPDRS

xii

Page 13: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

1

1 - CHAPTER 1 – GENERAL INTRODUCTION

The main focus of the work presented in this thesis is a study of the oscillatory activity

recorded with microelectrodes from the basal ganglia in patients with Parkinson‟s disease

(PD) and its significance in mediating the motor symptoms of the disease. The data is

mostly descriptive, aiming to characterize the nature of synchronized oscillations in PD

and to relate it to some clinical aspects. The data from microelectrode recordings in

human enable to examine both cell firing and local field potentials (LFPs) and can

contributes critical new insight into this area of research, which is almost entirely derived

from LFP recordings taken from the much larger contacts used for deep brain stimulation.

The general introduction, which follows, is intended to provide an extensive literature

overview of Parkinson‟s disease (including etiology, clinical features and treatments) and

the anatomy and physiology of the basal ganglia. Some models of the basal ganglia

function in relation of PD will be presented, followed by an overview of

electrophysiological and behavioral findings in PD patients and animal models of PD.

The specific aims of the studies presented in the thesis will be described at the end of this

chapter.

Chapter 2 provides a detailed description of the methods used in this research work. Each

study is then presented in a separate chapter:

Chapter 3: Beta oscillatory activity in the subthalamic nucleus and its relation to

dopaminergic response.

Chapter 4: The relationship between subthalamic oscillatory activity and tremor.

Chapter 5: Oscillatory activity in the globus pallidus: comparison between Parkinson‟s

disease and dystonia.

A general discussion on the results and their significance is given in Chapter 6.

An additional study describing the physiological properties of the pedunculopontine

nucleus (as well as a literature review about this structure) is included in the Appendix.

Page 14: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

2

1.1 - Parkinson’s disease

Parkinson‟s disease is a progressive age-related neurodegenerative movement disorder

that was first described by James Parkinson in 1817. The pathologic hallmark of the

disease is progressive degeneration of dopaminergic neurons in the substantia nigra pars

compacta (SNc) that project to the striatum (Hornykiewicz, 1966). The dopaminergic loss

in the striatum leads to alterations in the activity of the neural circuits within the basal

ganglia that regulate movement (Lang and Lozano, 1998) resulting in a difficulty to

perform both voluntary and involuntary movements. Parkinson‟s disease also involves

non-motor symptoms such as depression, dementia and sleep disturbances (Ziemssen and

Reichmann, 2007;Poewe, 2008) and is therefore associated with significant disability and

substantially decreased quality of life.

PD is the second most common neurodegenerative disorder after Alzheimer‟s disease

(Lew, 2007) affecting approximately 0.3% of the world‟s general population and 3% of

the people over the age 65 (Zhang and Roman, 1993). In North America alone, over one

million people have been diagnosed with PD (Lang and Lozano, 1998). Prevalence of

PD increases with age. The mean age of onset is approximately 60 years of age although

5 to 10% of patients are diagnosed before the age of 40 and are considered to have

“young-onset Parkinson‟s disease” (Bennett et al., 1996;Lang and Lozano, 1998). The

mortality among affected individuals is two to five times higher than among aged

matched controls (Louis et al., 1997).

1.1.1 - Etiology of Parkinson’s disease

A severe loss (~90%) of dopamine-secreting neurons in the SNc of the basal ganglia is

the pathological hallmark of PD (Hornykiewicz and Kish, 1987;Hornykiewicz, 1966).

Noradrenergic, serotonergic and cholinergic neuron loss also occurs in other brain

structures (Jellinger, 1991). The exact cause of neural degeneration remains unclear, but

may be due to a combination of factors including excitotoxic cell death (Przedborski and

Jackson-Lewis, 1998) genetics (Gasser, 2007) and environmental factors (Priyadarshi et

Page 15: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

3

al., 2001).

Lewy bodies

Lewy body formation is suggested to be pathogenic in PD. Lewy bodies are eosinophilic

hyaline inclusions that are consistently found in vulnerable neuronal populations in

parkinsonian brains (Lang and Lozano, 1998). In particular, they can be found in the

brain stem, basal forebrain and cortex. Lewy bodies are not specific to Parkinson's

disease and are found in other neurodegenerative disorders (McKeith, 2000) as well as in

elderly people without clinically evident PD (Gibb and Lees, 1988), suggesting they are

unrelated to the symptoms of PD.

Mitochondrial failure and oxidative stress

1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) is a potent neurotoxin with

selective effects on nigral dopaminergic neurons (Marsden and Jenner, 1987). In the early

1980‟s, severe parkinsonism developed in some subjects exposed to MPTP (Langston et

al., 1983). As a lipophilic compound, MPTP can cross the blood brain barrier. Once in

the brain, MPTP is converted within serotonergic neurons and astrocytes by monoamine

oxidase into the neurotoxin 1-methyl-4-phenylpyridinium (MPP+) (Chiba et al.,

1984;Brooks et al., 1989). Thereafter MPP+ is released into the extracellular space and is

then taken up into dopaminergic neurons by the plasma membrane dopamine transporter

(Mayer et al., 1986). Subsequently, MPP+ rapidly accumulates in the mitochondrial

matrix and inhibits Complex 1 of the electron transport chain, leading to energy failure

and cell death (Chan et al., 1991). In addition, the energy failure of the cell further

promotes other toxic and genetic insults, thereby increasing its susceptibility

to apoptosis

(Rossetti et al., 1988). MPTP is widely used today to reproduce parkinsonian symptoms

in nonhuman primate models of PD.

It has been suggested that in Parkinson's disease, there is an excess of reactive oxygen

species and increased oxidative stress (Lang and Lozano, 1998). Elevation of iron levels

detected in the SNc of PD patients is believed to be an important factor in causing

oxidative stress (Hirsch and Faucheux, 1998).

Page 16: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

4

Excitotoxic cell death

Excitotoxic mechanisms are implicated to play a role in neuronal degeneration in the SNc.

Over-excitation of the glutamatergic N-methyl-d-aspartate (NMDA) receptor leads to an

increase in intracellular calcium levels that could potentially lead to mitochondrial

damage and cell death (Nicholls and Budd, 1998). The subthalamic nucleus (STN),

which sends glutamatergic projections to the SNc, is known to be hyperactive in PD.

Therefore, the excessive glutamatergic drive from STN could be the source of excitotoxic

death of dopaminergic neurons in SNc (Lang and Lozano, 1998).

Genetics

There is increasing evidence that genetic factors have an important role in Parkinson's

disease. Studies of families with inherited forms of PD suggest a few genes that are

associated with the disease. Mutations in the gene alpha-synuclein in families with a rare

autosomal inherited form of dominant PD lead to aggregation of alpha-synuclein protein,

a major constituent of Lewy body fibrils (Wakabayashi et al., 2007). Another cause for

dominant PD is mutations in the gene for LRRK2 that might be acting as a primer for

neurodegeneration (Singleton, 2005). Mutations in the parkin gene, DJ-1 and PINK1 all

cause autosomal recessive parkinsonism of early onset. Parkin, DJ-1 and PINK1 proteins

protect against mitochondrial damage and oxidative stress. Loss of function mutations in

one of those genes render dopamine neurons more susceptible to mitochondrial

dysfunction and oxidative stress (Dodson and Guo, 2007;Gasser, 2007). It should be

noted, however, that known genetic forms of PD can account for only a small percentage

of all cases of sporadic PD, suggesting further genetic etiology awaits discovery (Marras

and Lang, 2008).

Epidemiology

The discovery of the selective ability of various exogenous toxins (e.g. MPTP) to induce

nigral cell death attracted great interest in potential environmental factors capable of

causing Parkinson's disease. Epidemiological studies have reported that people living in a

rural area are at significantly increased risk of getting PD. This might be related to

farming and the exposure to potential neurotoxins present in pesticides (Priyadarshi et al.,

Page 17: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

5

2001).

1.1.2 - The clinical features of Parkinson’s disease

The cardinal motor symptoms of PD include bradykinesia (a reduction in amplitude and

velocity of voluntary movement), often with akinesia (initiation difficulty and absence of

spontaneous movement), rigidity (increased resistance to passive movement), tremor at

rest and postural instability (Jankovic, 2008). PD symptoms worsen with time (Calne

and Langston, 1983;Koller and Montgomery, 1997) with one side of the body initially

affected and bilateral involvement as the disease progresses (Poewe and Wenning, 1998).

The initial diagnosis is made on the basis of the presence and asymmetry of these

symptoms and a good response to levodopa. Levodopa (also called L-Dopa) is an oral

dopamine precursor and is currently the most effective therapy in the management of PD

(Lang and Lozano, 1998). A number of rating scales are used for the evaluation of

impairment and disability in patients with PD although the most well established is the

Unified Parkinson‟s Disease Rating scale (UPDRS) (Ramaker et al., 2002). The UPDRS

has four components: part I, mentation, behavior and mood; part II, activities of daily

living; part III, motor; part IV, therapy complications.

Investigation of the clinical symptoms expressed by PD patients is important because it

provides valuable clues regarding the physiology and pathophysiology of the brain

regions involved, thereby helping in the design of care for PD patients today and leading

to new directions in future research.

1.1.2.a - Akinesia and bradykinesia

Akinesia refers to the lack or paucity of spontaneous movement whereas bradykinesia

refers to slowness of movement. It has been proposed that akinesia might result from an

inability to integrate cognitive and motor processes (Brown and Jahanshahi, 1996) while

bradykinesia results from an inability to generate the appropriate amount of muscle

activity (to provide enough force) as measured with electromyography (EMG) (Hallett

and Khoshbin, 1980). Because patients with PD have decreased electromyographic

Page 18: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

6

activity (Berardelli et al., 2001), they need a series of multiple agonist bursts to

accomplish larger movements. Bradykinesia can be also measured by testing reaction

time or movement speed. Reaction times are significantly delayed in PD patients

(Heilman et al., 1976;Evarts et al., 1981) and are even slower with increasing task

complexity (Kutukcu et al., 1999). Similarly, increasing task complexity slows movement

speed and increases the length of pauses between the elements in sequential task

(Benecke et al., 1986;Benecke et al., 1987).

Bradykinesia is considered the most characteristic clinical feature of PD (Marsden, 1984)

and it appears to correlate best with the degree of dopamine deficiency in the striatum

(Vingerhoets et al., 1997). It has been hypothesized that bradykinesia results from altered

motor cortex activity that is mediated by dopamine depletion. Reduced firing rates of

cortical neurons following dopamine receptor blockade in the rat have been shown to

correlate with bradykinesia (Parr-Brownlie and Hyland, 2005). Functional neuroimaging

studies also suggest impairment in the recruitment of cortical and subcortical structures

that regulate kinematic movement parameters such as velocity (Turner et al., 2003). The

deficit appears to be localized in the putamen and globus pallidus (Lozza et al., 2002),

resulting in the reduction in the muscle force produced at the initiation of movement.

Interestingly, bradykinesia is dependent on the cognitive state of the patient. A patient

who becomes excited, for example, may be able to make quick movements such as

catching a ball or may be able to suddenly run in case of a danger. This phenomenon

(kinesia paradoxica) suggests that PD patients have intact motor programs

but have

difficulties accessing them without an external trigger, such as a loud noise or a visual

cue (Jankovic, 2008).

1.1.2.b - Rigidity

Rigidity in PD manifests as an increase in resistance of a joint to passive movement. The

subjective correlate of rigidity is a feeling of stiffness and a reduced ability to relax limb

muscles (Klockgether, 2004). Rigidity is the consequence of an enhanced response of

muscles to stretch (Lee, 1989). However, it cannot be explained by an exaggeration of the

Page 19: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

7

short-latency spinal component of the stretch reflex, which is mediated by a

monosynaptic spinal reflex (Klockgether, 2004). Since surgical interventions, such as

lesion or deep brain stimulation, on central structures (e.g. basal ganglia) improve rigidity

(Krack et al., 2003;Fine et al., 2000;Lang et al., 1999), it is likely that central mechanisms

are involved. Indeed, the stretch reflex has a long-latency component which, at least in

part, represents the output of a transcortical reflex loop (Marsden et al., 1983;Deuschl and

Lucking, 1990). Several studies have shown that the amplitude and duration of the long-

latency response of the stretch reflex are increased in PD patients (Lee and Tatton,

1975;Tatton and Lee, 1975;Cody et al., 1986;Rothwell et al., 1983;Berardelli et al., 1983).

Since the cerebral cortex is not primarily affected in PD, it is plausible that the over-

excitability of the transcortical reflex loop is due to an abnormal output of the basal

ganglia (Klockgether, 2004). This view is supported by the fact that the long-latency

response of the stretch reflex is also abnormal in another disorder affecting the basal

ganglia (e.g. Huntington‟s disease) (Noth et al., 1985).

Another mechanism involved in rigidity is an abnormal response to the shortening in the

antagonist muscle. In normal subjects, antagonist muscle activity is suppressed during

stretching. In PD patients, an exaggerated EMG response can be recorded in antagonist

muscles resulting in an almost constant joint torque independent of joint position (Xia

and Rymer, 2004). All these mechanisms appear to be under the control of the

dopaminergic system, as dopamine replacement medications effectively reduce rigidity

and normalize the abnormal stretch responses (Klockgether, 2004).

1.1.2.c - Tremor

Tremor is the most easily recognized symptom of PD. It occurs in approximately 75% of

patients with PD (Hughes et al., 1993), typically during rest (Deuschl et al., 1998).

Parkinson‟s tremor is defined as 4-6 Hz rhythmic activation of antagonist muscles,

usually in the distal portions of limbs (Shahani and Young, 1976). Resting tremor is

diminished during active movements of the limb and may be enhanced during periods of

mental stress (Deuschl et al., 1998). Unlike rigidity and bradykinesia, tremor does not

necessarily get worse with disease progression and the severity of tremor does not

Page 20: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

8

correlate with dopamine deficiency in the striatum (Stebbins et al., 1999;Deuschl et al.,

2000). Post-mortem and imaging studies suggest that the pathophysiology of rest tremor

may be distinct from that of rigidity and bradykinesia (Jellinger, 1999;Pavese et al., 2006).

A more detailed review on the pathophysiology of resting tremor in PD will be given in

sections 1.3.4 and 1.4.5.

Low amplitude 4-10 Hz postural and action tremors are also expressed in some PD

patients (Findley et al., 1981;Hadar and Rose, 1993;Palmer and Hutton, 1995). Postural

tremor appears when patients hold a posture while action tremor is expressed during

voluntary movement of the affected limb (Findley et al., 1981).

1.1.2.d - Postural instability

Postural instability is generally a manifestation of the late stages of PD (Jankovic, 2008)

and can become the most debilitating symptom to a patient‟s quality of life (Bloem,

1992). Postural instability is one of the most common cause of falls and contributes

significantly to the risk of hip fractures (Williams et al., 2006). The balance impairment is

due to abnormal modulation of postural reflexes in the lower extremities (Beckley et al.,

1991). Although the timing of associated postural adjustments is normal in PD, their size

may decrease (Dick et al., 1986) due to inability to modify the size of posturally

stabilizing long-latency reflexes (Beckley et al., 1993). Postural instability may also be

due to an increase in muscle tone and tremor in the lower extremities (Burleigh et al.,

1995), orthostatic hypotension (a sudden fall in blood pressure) and age related sensory

changes (Bloem, 1992).

1.1.2.e - Other motor abnormalities

Patients with PD may exhibit a number of secondary motor symptoms that may impact

on their functioning (Jankovic, 2008). Some patients display a re-emergence of primitive

reflexes such as the glabellar reflex (subjects blink in response to repetitive tapping on

their forehead) due to a breakdown of the frontal lobe inhibitory control (Vreeling et al.,

1993;Thomas, 1994). Speech disorders are also frequently observed in PD patients and

are characterized by monotonous, soft and breathy speech with variable rate and word

Page 21: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

9

finding difficulties (Critchley, 1981). A number of neuro-ophthalmological abnormalities

may be seen in patients with PD including decreased blink rate, ocular

surface irritation,

altered tear film, visual hallucinations and decreased convergence (Biousse et al., 2004).

In addition, abnormalities in ocular pursuit and saccades (including antisaccades) are also

observed (Rascol et al., 1989). PD patients may also suffer from respiratory disturbances.

Upper airway obstruction and chest wall restriction are both common (Shill and Stacy,

2002;Sabate et al., 1996) and are associated with significant morbidity and mortality. The

obstructive pattern may be related to rigidity, cervical arthrosis or restricted

range of

motion in the neck, and the restrictive pattern may be related to chest wall rigidity (Shill

and Stacy, 1998).

1.1.2.f - Non-motor features

PD patients also suffer from non-motor symptoms which impair their quality of life quite

considerably. Non-motor symptoms consist of autonomic dysfunction,

cognitive/neurobehavioral disorders, sleep disturbances and sensory abnormalities

(Ziemssen and Reichmann, 2007;Poewe, 2008). Autonomic failures include

cardiovascular dysfunction (e.g. orthostatic hypotension), as well as gastrointestinal,

urogenital and thermoregulatory dysfunctions (Jankovic, 2008).

Greater than 80% of patients at 15 years after their initial assessment are reported to show

cognitive decline and ~50% meet the diagnostic criteria for dementia (Hely et al., 2005).

PD related dementia is also associated with a number of other neuropsychiatric

conditions.

Among patients with dementia, nearly 60% suffer from depression, and many suffer from

apathy (54%), anxiety (49%) or hallucinations (44%) (Aarsland et al., 2007). The

prevalence rates of depression in patients with PD have been reported to be as high as

40% (Cummings, 1992;Cummings and Masterman, 1999). Depression is predominantly

caused by fronto-cortical dysfunction and by degeneration of monoaminergic

neurotransmitter systems (Ziemssen and Reichmann, 2007). Cognitive impairment in PD

might also result from the dysfunction of non-dopaminergic neuronal systems (Pillon et

al., 1989).

Page 22: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

10

Sleep disturbances such as insomnia, daytime sleepiness and rapid eye movement sleep

behavior disorder (RBD) are very frequent in PD (Comella, 2003). RBD occurs in

approximately one-third of PD patients and is characterized by an increase in violent

dream content accompanied by dramatic, violent and potentially injurious motor activities

(e.g. yelling, swearing, grabbing, punching, kicking, jumping and other) which may also

involve the bed partner (Gagnon et al., 2006). Insomnia and daytime sleepiness are also

frequent with about 50% prevalence (Gjerstad et al., 2007;Arnulf et al., 2002). The sleep

abnormalities observed in PD patients are possibly the result of a severe loss of

hypocretin (orexin) neurons (Fronczek et al., 2007;Thannickal et al., 2007).

Sensory symptoms such as olfactory dysfunction (Doty et al., 1992) and pain (Djaldetti et

al., 2004;Tinazzi et al., 2006) are also frequent but are often not recognized as PD

symptoms (Jankovic, 2008).

1.1.3 - Dopamine therapy in Parkinson’s disease

As mentioned above, by the time PD is diagnosed most (~90%) of the dopaminergic SNc

neurons have degenerated (Hornykiewicz, 1966;Hornykiewicz and Kish, 1987). Positron

emission tomography (PET) with the levodopa analogue 6-fluorodopa has demonstrated

that in humans exposed to MPTP, dopaminergic loss can exist in the absence of clinical

signs (Calne et al., 1985). The late appearance of the clinical symptoms is due to

compensatory mechanisms (Bezard et al., 1997b). Previous to the appearance of

symptoms, remaining SNc neurons produce larger amounts of dopamine and striatal

neurons become more sensitive to dopamine. There is also increased glutamatergic

inputs to the SNc during the presymptomatic period (Bezard et al., 1997b;Bezard et al.,

1997a;Bezard et al., 1999). When enough SNc neurons degenerate, compensatory

mechanisms can no longer balance the loss of dopamine and PD symptoms appear

(Bezard and Gross, 1998).

The dopamine precursor levodopa was discovered in the 1960‟s (Cotzias et al., 1967) and

is the most potent drug for controlling PD symptoms. When administered orally,

Page 23: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

11

levodopa is absorbed mainly in the duodenum and the proximal bowel. Unlike dopamine

levodopa is able to cross the blood brain barrier. Following absorption into brain,

levodopa is taken up by residual dopaminergic neurons in striatal terminals and is

converted into dopamine by the enzyme aromatic L-amino-acid decarboxylase dopamine.

Dopamine is then stored into vesicles of the presynaptic neuron from which it is released

into the synaptic cleft to stimulate dopamine receptors on post-synaptic striatal cells

(Thanvi and Lo, 2004). At the pre-synaptic level, dopamine acts to increase dopamine

synthesis and transmission. Early studies suggested that the relatively intact serotonergic

input to the basal ganglia from the dorsal raphe nucleus may help conversion of levodopa

to dopamine (Tanaka et al., 1999).

Metabolism of dopamine involves its reuptake by the presynaptic neuron

and the

enzymatic conversion to 3, 4, dihydroxyphenyl acetic acid by monoamine oxidase

inhibitor and to 3-methoxytyramine by the enzyme catecholamine-o-methyl transferase.

Levodopa can be broken down in the periphery as well. Thus, it is routinely administered

in combination with peripheral decarboxylase inhibitor (such as benserazide or carbidopa)

to reduce its conversion to dopamine at the periphery (and associated side effects such as

nausea and vomiting) and to increase dopamine availability in the central nervous system.

Levodopa significantly improves bradykinesia and akinesia. PET studies have

demonstrated that the degree of nigrostriatal dopamine deficiency in PD correlates most

with bradykinesia compared to other symptoms (Vingerhoets et al., 1997). Levodopa

also improves rigidity, tremor (Yuill, 1976), hypometria (small amplitude movements)

(Beckley et al., 1995), the performance of complex tasks (Benecke et al., 1987), and the

generation of internally cued movements (Burleigh-Jacobs et al., 1997). However, not all

symptoms of PD are responsive to dopaminergic medication. For instance, dopaminergic

medications have no significant effect on postural instability, suggesting that

nondopaminergic mechanisms are involved in this symptom (Bloem et al., 1996).

Unfortunately, patients with longstanding levodopa treatment often develop motor

fluctuations, where the effects of levodopa switch rapidly and unpredictably from no

Page 24: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

12

effect (“off”), to a positive effect (“on”), to uncontrollable involuntary movements

(dyskinesias) (Obeso et al., 2000a;Obeso et al., 2000b;Thanvi and Lo, 2004). Levodopa-

induced dyskinesias are observed in the majority of patients who have been treated for

5–

10 years with levodopa (Schrag and Quinn, 2000). These medication-related motor

fluctuations are difficult to treat and become a major contributor to disability in some

patients. The neurological basis of levodopa-induced dyskinesias is unclear but might be

related to the discontinuous stimulation of dopamine receptors that accompanies

intermittent levodopa dosing, leading to dopamine receptor hypersensitivity (Chase,

1998;Bezard et al., 2001). A recent PET study found reduced dopamine transporter

expression on presynaptic sites in PD patients with dyskinesias (Troiano et al., 2009).

Decrease in metabolism and/or decrease in the mean firing rate of basal ganglia neurons

might also be involved in levodopa-induced dyskinesias as well as overactivity of cortical

motor areas (Bezard et al., 2001). An interesting fact is that the emergence of dyskinesias

in PD patients correlates with the levodopa-induced reduction in bradykinesia, but not

rigidity (Caligiuri and Peterson, 1993).

One strategy to reduce the risk for motor complications is the use of dopamine receptor

agonists which have been proved safe and effective as initial therapy in early stages of

Parkinson's disease. However, it is still controversial whether they must be started early,

as opposed to initiation only after the levodopa complications develop (Ahlskog, 2003).

Dopamine agonists such as bromocriptine, pergolide and apomorphine bind at post-

synaptic receptor sites independently of the dopamine terminal. They have a longer half

life than levodopa (half life of levodopa is approximately 90 minutes), thereby reducing

receptor sensitivity and the development of motor complications (Junghanns et al., 2004).

Unfortunately, dopamine agonists have adverse effects on their own including nausea,

hypotension, hallucinations, daytime sleepiness, pathological gambling, cardiac

valvulopathy and edema.

1.1.4 - Deep brain stimulation for Parkinson’s disease

Stereotactic neurosurgery to treat PD in humans started with the initial work of Cooper

Page 25: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

13

who showed that chemical lesions of the pallidum in patients with PD could effectively

treat parkinsonian symptoms (COOPER, 1954). Starting in the 1950s, thalamotomy

(surgical destruction of the motor thalamus) and pallidotomy (surgical destruction of the

internal segment of the globus pallidus) were regularly performed to alleviate the

symptoms of PD. In these procedures, an electrode is placed in the target area of the

brain, using stereotactic neurosurgical techniques, and the exposed electrode tip is heated

with radiofrequency current to create the lesion. In 1969, with the introduction of

levodopa, the number of lesioning operations declined dramatically (Speelman and Bosch,

1998;Wichmann and DeLong, 2006b). But in the mid 1970‟s, as a result of the

shortcomings of the levodopa therapy in the long-term treatment, thalamotomy gradually

regained its place (Siegfried, 1980). Later in the early 1990‟s, with the better

understanding of the important role of the globus pallidus and the subthalamic nucleus in

the organization of normal motor control and PD (Alexander et al., 1990), there was a

resurgence of pallidotomy, which was also effective in alleviating levodopa-induced

dyskinesias (Laitinen et al., 1992a;Laitinen et al., 1992b;Lozano et al., 1995;Baron et al.,

1996;Kishore et al., 1997;Narabayashi et al., 1997).

The modern era of deep brain stimulation (DBS) to treat PD symptoms began in 1987

when Benabid and colleagues reported their experience with high-frequency (>100 Hz)

stimulation in the ventralis intermedius (Vim) nucleus of the thalamus for treating PD

tremor (Benabid et al., 1991;Benabid et al., 1987). Subsequently, DBS was applied to

basal ganglia structures, namely the internal segment of the globus pallidus (GPi) and the

subthalamic nucleus (STN), to treat all symptoms of PD (Guridi et al., 1993;Siegfried and

Lippitz, 1994;Limousin et al., 1995;Obeso et al., 1997). This procedure involves

implantation of an electrode into the target region using stereotactic neurosurgical

techniques (Lemaire et al., 2007). The electrode lead is then connected with an extension

wire to a programmable pulse generator that is implanted below the clavicle. The

stimulation parameters are programmed using noninvasive radio-telemetry to achieve

maximal clinical benefit. The clear advantage of DBS over lesioning is that there is

minimal destruction of brain tissue and the electrode can be potentially removed without

creating permanent damage (Lozano and Mahant, 2004).

Page 26: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

14

DBS is recommended to PD patients who are levodopa-responsive but suffer from

levodopa-related complications and motor fluctuations that cause significant disability

(Lozano and Mahant, 2004). The choice of target is largely dependent on the

neurosurgeon. Although thalamic stimulation has been well established as an effective

treatment for PD tremor (Benabid et al., 1996), it is rarely indicated since it does not

alleviate the other major features of PD (Lozano, 2000). Stimulation of the GPi or STN,

on the other hand, have been shown to exert beneficial effects for most of the main

symptoms of PD (Pahwa et al., 1997;Gross et al., 1997;Tronnier et al., 1997;Krack et al.,

1997;Ghika et al., 1998;Krack et al., 1998;Volkmann et al., 1998;Kumar et al.,

1998;Limousin et al., 1998;Kumar et al., 1999c;Moro et al., 1999). The most remarkable

effect of both GPi and STN surgery is a marked reduction in levodopa-induced

dyskinesias. While GPi stimulation directly suppresses levodopa-induced dyskinesias,

STN DBS allows patients to reduce their levodopa intake, leading to a reduction in

dyskinesias (Ghika et al., 1998;Krack et al., 1998;Moro et al., 1999). Since STN allows

the reduction of anti-parkinsonian medication, most centers prefer the STN to GPi

surgery (Lozano and Mahant, 2004). GPi, on the other hand, has became the favored

target in the treatment of patients with dystonia (Kumar et al., 1999b;Yianni et al.,

2003;Vidailhet et al., 2005), a syndrome characterized by sustained concurrent

contractions of the agonist and antagonist muscles, thereby producing abnormal, and

sometimes painful, movements or postures. It includes a vast array of diseases with

different etiologies and presentations.

Nowadays, DBS in different brain regions is remarkably effective in some brain areas (i.e.

STN, thalamus and GPi) to treat a range of neurological disorders, and the efficacy for

neuropsychiatric disease is awaiting blinded trials. It should be emphasized, however,

that despite the therapeutic benefits, the mechanisms of action of DBS are still unclear.

Whether the therapeutic effects are local or system-wide or whether the effects are related

to inhibition or excitation are still a matter of debate. Differences in techniques, anatomy,

cell type, and experimental setting limit the ability to make direct comparisons across the

different studies (Perlmutter and Mink, 2006). Multiple reports have documented the

Page 27: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

15

modulation of both discharge rates and firing patterns during and following high-

frequency stimulation throughout the cortico-basal ganglia loop. Early studies reported

inhibition of neuronal activity within the stimulated nucleus during high-frequency

stimulation in the GPi (Dostrovsky et al., 2000) and STN (Filali et al., 2004) of humans,

and GPi in primates (Boraud et al., 1996). Although these findings are in line with the

similar clinical effects of lesions and DBS, there are some evidence for only minor

effects of DBS on mean firing rates (McCairn and Turner, 2009). In studies that have

examined the effect of high-frequency stimulation on downstream targets, the findings

suggest an activation of efferent axons either directly or through activation of local cell

bodies to axon initial segments (Anderson et al., 2003;Hashimoto et al., 2003). Continous

high-frequency stimulation in the STN (Hashimoto et al., 2003) or pallidum (Bar-Gad et

al., 2004;McCairn and Turner, 2009) resulted in time-locked changes in firing in the

primate globus pallidus. Time-locked changes in firing were also found in the STN

following cortical (Nambu et al., 2000) and pallidal (Nambu et al., 2000;Kita et al., 2005)

stimulation, and in the thalamus following STN stimulation (Xu et al., 2008). High-

frequency stimulation was also shown to affect other properties of firing patterns such as

oscillatory and burst activity (Brown et al., 2004;Wingeier et al., 2006;Meissner et al.,

2005;Kuhn et al., 2008;Montgomery, 2006;Hahn et al., 2008;Dorval et al., 2008;Xu et al.,

2008;McCairn and Turner, 2009) suggesting that it suppresses the pathological patterns

of activity in the basal ganglia outputs. A recent study in MPTP-treated monkeys has

suggested that short-term depression of synaptic transmission may contribute to the

mechanism underlying the effects of high-frequency stimulation (Erez et al., 2009).

Importantly, cortical high-frequency stimulation may also have antiparkinsonian effects

as shown in MPTP monkeys (Drouot et al., 2004).

1.2 - The basal ganglia: anatomy and physiology

The basal ganglia (BG) are a group of subcortical nuclei that are involved in motor and

cognitive functions. The principle components of the BG are the striatum, the STN, the

globus pallidus (GP) and the substantia nigra (SN). The striatum is the main input

structure and is divided into the caudate and putamen which are separated by the internal

Page 28: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

16

capsule. The STN can also be considered an input nucleus because, like the striatum, it

receives direct input from the cerebral cortex (Kitai and Deniau, 1981;Monakow et al.,

1978). The GP is divided by the internal medullary lamina into the internal segment (GPi)

and external segment (GPe). The two divisions of the GP have different inputs and

outputs and are functionally distinct. Similarly, the SN is divided into the pars compacta

(SNc) and pars reticulata (SNr). These two parts of the substantia nigra share similar

inputs but have different outputs and are composed of neurochemically distinct neuron

types. In post-mortem tissue the SNc is defined by the black staining of neuromelanin

which is present in dopaminergic neurons. The GPi and SNr are the major output nuclei

of the BG through which the BG project to the thalamus, the superior colliculus and the

pedunculopontine nucleus (PPN) (Tepper et al., 2007;Herrero et al., 2002). There are no

direct outputs from the basal ganglia to spinal or brainstem motor neurons.

Through the thalamus, the BG influence motor, sensory, and cognitive cortical

information processing (Middleton and Strick, 1994;Middleton and Strick,

1996;Middleton and Strick, 2002;Hoover and Strick, 1993). The BG also influence

movements of the head and eyes through the superior colliculus (Hikosaka et al., 2000),

and influence spinal cord processing and aspects of locomotion and postural control

through the PPN (Takakusaki et al., 2004;Garcia-Rill et al., 1983).

Basal ganglia dysfunction is associated with a range of debilitating clinical conditions

whose most obvious manifestations are disturbances in movement. These movement

abnormalities can be found not only in well defined movement disorders such as

Parkinson‟s disease, Huntington‟s disease and dystonia but also in diseases such as

obsessive-compulsive disorder, schizophrenia and various addictive behaviors. Many of

these disorders, including Parkinson's disease, have associated cognitive and affective

components as well.

1.2.1 - Functional organization and circuitry

1.2.1.a - Connectional architecture

Page 29: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

17

The vast majority of neurons in the BG release gamma-aminobutyric acid (GABA) and

most are projection neurons. The striatum, both segments of the GP and the SNr are each

composed of GABAergic projection neurons, whereas the STN contains glutamatergic

projection neurons and the SNc contains mainly dopaminergic projection neurons. The

striatum also contains clearly defined populations of interneurons, all but one (the

cholinergic interneurons) of which are GABAergic (Tepper et al., 2007).

The principle pathways of information flow throughout the basal ganglia and their

associated regions are illustrated in Figure 1.F1. The BG receive major excitatory input

from the cerebral cortex (particularly supplementary motor area, premotor cortex,

precentral motor, and postcentral sensory areas) and the intralaminar thalamic cell nuclei

(e.g. centromedian-parafascicular complex). These afferents are glutamatergic and

terminate in the striatum and STN. The BG (GPi and SNr) send output back to the cortex

via the ventrolateral thalamus. For the last two decades, basal ganglia circuitry has been

considered to be dominated by two parallel pathways by which cortical information is

transmitted to the output nuclei of the basal ganglia, the GPi and SNr (Alexander and

Crutcher, 1990;Smith et al., 1998;Albin et al., 1989;DeLong, 1990). In the so-called

direct pathway, striatal GABAergic neurons monosynaptically inhibit the GABAergic

output neurons in GPi and SNr. In the indirect pathway, striatal neurons inhibit the

GABAergic neurons of the GPe. GPe neurons innervate GPi/SNr via direct projections

and indirectly by innervating the glutamatergic neurons of the STN that excite GPi/SNr.

As opposed to the direct pathway, the indirect pathway excites GPi/SNr. It is important to

note, however, that the separation between these two pathways is not absolute.

Anatomical evidence indicate that individual striatal neurons can be involved in both the

direct and indirect pathways (Kawaguchi et al., 1990;Parent et al., 1995;Parent et al.,

2000;Wu et al., 2000;Levesque and Parent, 2005b). In addition to these two pathways, a

cortical glutamatergic „hyperdirect‟ pathway to STN also exists (Monakow et al.,

1978;Kitai and Deniau, 1981). Activation of this pathway excites GPi/SNr neurons.

Page 30: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

18

1.2.1.b - Differential effects of dopamine

The balance between direct and indirect pathways is regulated by the differential actions

of dopamine, from the SNc, on striatal neurons. The striatal neurons giving rise to the

direct and indirect pathways are similar in their electrophysiological and morphological

properties, but are distinguished in their neurochemical features. The direct pathway

neurons express the dopamine D1 receptor together with substance P and dynorphin

whereas the indirect pathway neurons express the dopamine D2 receptor and enkephalin

(Aubert et al., 2000;Le Moine and Bloch, 1995;Gerfen et al., 1995). D1 receptors are

Page 31: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

19

coupled to G proteins that activate adenylate cyclase to raise cyclic adenosine

monophospate (cAMP) levels, thereby producing an excitatory effect. D2 receptors are

coupled to different types of G proteins that activate phosphodiesterase which break

down cAMP, thereby producing an inhibitory effect (Girault and Greengard, 2004). Thus,

striatal dopamine enhances transmission along the direct pathway (acting on D1 receptors)

and reduces transmission along the indirect pathway (acting on D2 receptors). Together

these actions result in a net reduction in GPi and SNr activity (DeLong and Wichmann,

2007).

1.2.1.c - Parallel organization

In addition to the serial connections between basal ganglia nuclei, there are many distinct

circuits which connect the BG, thalamus, and vast regions of the cortex.

Electrophysiological and anatomical studies suggest that the main circuits passing

through the BG remain “segregated” under normal conditions and that the functional

architecture of the BG is essentially parallel in nature (Alexander et al., 1986;Hoover and

Strick, 1993). It has been suggested that there are five closed loops within the basal

ganglia-thalamocortical architecture, each of which centered upon a separate part of the

frontal lobe (Alexander et al., 1986). The “motor” circuit connects the primary motor

cortex, supplementary motor area and ventral premotor cortex. The “oculomotor” circuit

connects the frontal and supplementary eye fields. There are two “associative” circuits,

one involving the dorsolateral prefrontal cortex and the other involving the lateral

orbitofrontal cortex. Finally, a “limbic” circuit connects to the anterior cingulate and the

medial orbitofrontal cortex (Alexander et al., 1986;Alexander et al., 1990). More

recently, however, it has been proposed that these circuits are rather interconnected (at

the level of their output to different cortical regions) (Joel and Weiner, 1994). This

organization may allow the BG to concomitantly participate in both the motor and

complex non-motor functions of the cerebral cortex (Alexander et al., 1990;DeLong et al.,

1984b;Middleton and Strick, 1994).

1.2.2 - The striatum

Page 32: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

20

Subdivisions of the striatum

The striatum, the major input structure of the BG, is comprised of two main nuclei, the

caudate and putamen. At the rostral end of the striatum, the two nuclei appear as one

structure. Further caudally, the caudate and putamen are separated by the internal capsule

(Utter and Basso, 2008). In primates, the putamen receives cortical projections mainly

from somatosensory and motor areas, and has been therefore related to pure motor

aspects such as the execution of movements or the processing of proprioceptive

information (Crutcher and DeLong, 1984a;Crutcher and DeLong, 1984b;Kimura,

1986;Alexander et al., 1986). Putaminal neurons are somatotopically organized

throughout the rostrocaudal extent of the nucleus (Alexander and DeLong, 1985b).

Microstimulation in the putamen results in movements of individual body parts, perhaps

due to the activation of grouped putamen output neurons known as striatal microexcitable

zones (Alexander and DeLong, 1985a). It has been shown that single putamen neurons

change their firing in response to a conditioned flexion movement in cats (Cheruel et al.,

1994) and, in rats, have been shown to have integrative receptive fields for combined

somatosensory and auditory stimuli, which may be used to facilitate motor responses

(Chudler et al., 1995).

The caudate nucleus, on the other hand, receives extensive input from prefrontal cortex

and parietal association area and is therefore believed to play a role in associative

functions such as planning and anticipation of upcoming events (Apicella et al., 1991).

Neurons in the caudate nucleus of monkeys integrate sensory, motor, and non-motor

responses driven by the frontal cortex to produce coordinated behavior (Aldridge et al.,

1980;Nishino et al., 1984). In freely moving cats, caudate neurons show changes of

activity related to conditioned stimuli, to initiation of movement or to reinforcement

during a reaction time task (Amalric et al., 1984).

Finally, the ventral extension of the striatum (also called the nucleus accumbens) serves

limbic (emotional aspects) functions and receives input from the anterior cingulate and

medial orbitofrontal cortex (Alexander et al., 1986).

Page 33: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

21

Projection neurons of the striatum

The major output neurons of the striatum are GABAergic, so-called medium spiny

neurons, that account for more than 90% of striatal neurons (Bolam et al., 1981). These

neurons project primarily to the globus pallidus and SNr but also send local axon

collaterals to other striatal MSNs (Gerfen, 1988). Although all medium spiny neurons are

GABAergic, they can be further classified on the basis of their projection targets and

other neurotransmitters they contain. Medium spiny neurons projecting directly to the

SNr and GPi comprise the „direct pathway‟ and express GABA, substance P and

dynorphin (an opioid peptide), whereas medium spiny neurons that project to the GPe

and comprise the „indirect pathway‟ express GABA and enkephalin (Albin et al., 1989).

Anatomical studies demonstrate that the striatopallidal projection system in primates is

highly ordered and displays a high degree of specificity with respect to its target sites in

the pallidum (Hazrati and Parent, 1992c). Electrical stimulation of the striatum produces

a short latency inhibition of both GPi and GPe neurons and is followed by a longer period

of excitation, especially in the GPe (Tremblay and Filion, 1989). The early inhibition

followed by excitation was always displayed by neurons located in the center of the

pallidal zone of influence of each striatal stimulation site. At the periphery of the zone,

however, only excitation was observed. This topological arrangement suggests that

excitation is used, temporally, to control the magnitude of the central striatopallidal

inhibitory signal and, spatially, to focus and contrast it onto a restricted number of

pallidal neurons (Tremblay and Filion, 1989).

Interneurons of the striatum

In addition to the projection neurons of the striatum, there are many types of interneurons

within the striatum. The striatal interneurons are classified into three groups. The first are

the cholinergic large non-spiny neurons (Kemp and Powell, 1971;Bolam et al., 1984),

also known as tonically active neurons (TANs) because of their physiological phenotype.

The two other groups are the parvalbumin/GABAergic medium non-spiny neurons, and

the somatostatin, neuropeptide Y, and NADPH (nicotinamide adenine dinucleotide

phosphate-oxidase) diaphorase containing non-spiny neurons (Gerfen and Wilson, 1996).

Page 34: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

22

TANs are the most well studied interneurons of the striatum (Pisani et al., 2001). They

are activated by presentation of sensory stimuli of behavioural significant or stimuli

linked to reward, as they become synchronized in firing or develop pauses (Aosaki et al.,

1995;Raz et al., 1996). Recently, it has been shown that TANs respond to instruction

stimuli associated with motivational outcomes but not to unassociated ones in monkeys

(Kimura et al., 2003;Yamada et al., 2004). In addition, TANs in the caudate nucleus tend

to respond more to stimuli associated with motivational outcomes, whereas in the

putamen, TANs tend to respond to „go‟ signals especially for an action anticipating a

reward. These findings suggest a distinct, pivotal role played by TANs in the caudate

nucleus and putamen for motivation and reward (Kimura et al., 2003;Yamada et al.,

2004).

Afferent projections of the striatum

The striatum receives excitatory input from many cortical areas (Kemp and Powell, 1970)

and thalamic nuclei (Cowan and Powell, 1956). Anatomical and physiological studies

have shown that different cortical areas project to distinct regions of the caudate and

putamen. Cortical areas send somatotopically organized glutamatergic projections onto

the dendritic tips of medium spiny neurons (Kemp and Powell, 1970). Although cortico-

striatal inputs are distributed, there is also massive convergence of inputs to individual

striatal neurons (Flaherty and Graybiel, 1991;Flaherty and Graybiel, 1993;Flaherty and

Graybiel, 1994). This convergence of corticostriatal projections might play a role in the

integration of cortical information from diverse regions of the cortex (Graybiel et al.,

1994). Electrical stimulation of motor cortex, together with intracellular recordings in the

ipsilateral striatal projection field in rats, demonstrated long-term potentiation of cortico-

striatal synaptic transmission (Charpier and Deniau, 1997). This suggests the existence of

activity-dependent plasticity at glutamatergic cortico-striatal synapses. The physiological

long-term potentiation at corticostriatal synapses can be also induced by low frequency (5

Hz) synchronization of cortical afferents (Charpier et al., 1999), suggesting that the

striatal output neuron may operate as a coincidence detector of converging cortical

information.

Page 35: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

23

In addition to the cerebral cortex, the striatum receives excitatory inputs from the

thalamus. The centromedian (CM) and parafascicular (Pf) nuclei are an important source

of thalamostriatal projections. High resolution anterograde tracer studies have shown that

the CM and Pf provide distinct inputs to different parts of the striatum (Sadikot et al.,

1992a;Sadikot et al., 1992b). The CM projects to the entire sensorimotor territory of the

striatum, whereas the Pf provides complementary input to the entire associative region

(Sadikot and Rymar, 2008).

The striatum also receives a major input from dopaminergic neurons located in the

ventral midbrain. Dopaminergic innervation of the striatum arises from two nuclei, the

SNc and the ventral tegmental area (VTA) (Beckstead et al., 1979). Dopamine acts on

post-synaptic dopamine receptors that are located at the base of the dendritic spines of

medium spiny neurons, and is therefore in a position to modulate the effect of the

corticostriatal excitatory input to these neurons (Gerfen, 1988). In addition to dopamine

receptors on medium spiny neurons, presynaptic dopamine receptors are also present on

corticostriatal terminals (Arbuthnott et al., 1998), where they act to dampen striatal

excitation (Bamford et al., 2004). Dopamine also modulates striatal aspiny neurons

(cholinergic interneurons) (Aosaki et al., 1998). It has been shown that long-term

depression of synaptic transmission in the corticostriatal pathway is blocked by dopamine

receptor antagonists or lesions of the nigrostriatal dopaminergic pathway and can be

restored by the application of exogenous dopamine (Calabresi et al., 1992).

D1 and D2 dopamine receptors

The D1 and D2 dopamine receptors are the major subtypes expressed in the striatum. Both

D1 and D2 receptors are G-protein coupled receptors. Binding of dopamine to the D1

receptor results in depolarization of striatal neurons whereas binding to the D2 receptor

cause neuronal hyperpolarization (Sealfon and Olanow, 2000). Thus, the action of D1

receptor is to enhance cotricostriatal influence whereas the action of D2 receptor is to

reduced cotricostriatal influence. It was originally suggested that D1 and D2 receptors are

differentially expressed on the dendrites of striatal neurons. Striatal neurons of the direct

pathway express D1 receptors whereas those neurons projecting indirectly express D2

Page 36: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

24

receptors. Later studies have argued that a substantial subpopulation of medium spiny

neurons co-express D1 and D2 receptors (Surmeier et al., 1996;Aizman et al., 2000). This

controversy was recently resolved due to the availability of transgenic mice expressing

enhanced green fluorescent protein selectively for D1 and D2 receptors. It was confirmed

that dopamine receptor subtypes are differentially expressed on separate populations of

spiny neurons. These studies also show that the role of dopamine in synaptic plasticity is

different for D1 and D2 MSNs (activation of D1/D2 leads to spike-timing dependent

LTP/LTD respectively) (Shen et al., 2008), and that D1 and D2 MSNs have different

morphology (D2 MSNs have smaller dendritic threes than D1 MSNs) (Gertler et al., 2008)

and electrophysiological properties (D2 MSNs dendrites are more excitable) (Day et al.,

2008).

The original dichotomy between the two population of MSNs led to the classic notion of

the indirect pathway inhibition of movement and the direct pathway facilitation of

movement (Gerfen et al., 1990). Although this classification has had tremendous value,

anatomical studies indicate that it is oversimplified and that several modifications need to

be made. For example, neurons projecting to the GPi and SNr can also send axon

collaterals to the GPe, arguing that direct and indirect pathways are not completely

segregated (Kawaguchi et al., 1990).

Mosaic organization of the striatum

The second level of functional compartmental organization in the striatum is the

segregation of medium spiny neurons into striosomes and matrix compartments, which

differ in several cytochemical markers, input-output connections, and time of

neurogenesis. The striosomes are identified by patches of dense opiate receptor binding,

and are enriched in enkephalin- and substance P-like immunoreactivity. The matrix has a

high acetylcholinesterase activity, and a dense plexus of fibres displaying somatostatin-

like immunoreactivity (Pert et al., 1976;Graybiel and Ragsdale, Jr., 1978;Gerfen,

1985;Herkenham and Pert, 1981;Graybiel and Chesselet, 1984;Chesselet and Graybiel,

1986). The matrix comprises the majority of the striatum and receives input from

somatosensory and associational areas of the cerebral cortex, while the striosomes receive

Page 37: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

25

input from limbic and the prefrontal cortex. Striosomes project mostly to dopaminergic

neurons of the SNc, whereas the matrix gives rise to the parallel striatonigral and

striatopallidal pathways (Gerfen, 1984). These anatomical features suggest that

striosomes and matrix play different roles in the processing of cortical inputs. Indeed,

during relatively neutral behavioral conditions, the highest metabolic activity in the

striatum was observed in the matrix rather than in the striosomes (Brown et al., 2002). On

the other hand, in vivo electrical stimulation through electrodes centered in or around the

striosomes, but not the matrix, led to rapid acquisition and maintenance of bar-pressing,

self-stimulation behaviors associated with reward (White and Hiroi, 1998). It is therefore

currently believed that reward-related, limbic forebrain circuits are centered in the

striosomes, while sensorimotor and associative circuits are present in the matrix.

1.2.3 - Subthalamic nucleus

The subthalamic nucleus is a biconvex-shaped structure (Yelnik and Percheron, 1979)

that is densely populated by glutamatergic projection neurons (Rafols and Fox,

1976;Iwahori, 1978;Afsharpour, 1985). In the STN there are no intra-nuclear interactions

as a result of recurrent collaterals (Kita et al., 1983b), but there are some evidence for the

existence of interneurons (Rafols and Fox, 1976;Yelnik and Percheron, 1979;van der

Kooy and Hattori, 1980;Levesque and Parent, 2005a). The STN plays a central role in

BG circuitry since it receives input from, and projects to, many diverse nuclei (Alexander

and Crutcher, 1990).

Intrinsic organization of the STN

The STN in primates can be separated into three distinct functional subdivisions. It is

subdivided into two rostral thirds and a caudal third. The rostral two-thirds can be further

divided into medial third and lateral two-thirds (Figure 1.F2). The medial potion of the

rostral two-thirds is thought to comprise the limbic and part of the associative territories.

The ventral aspect of the lateral portion of the rostral two-thirds comprises the additional

portion of the associative territory. The dorsal part of the lateral portion of the rostral

two-thirds and the caudal third are related to motor circuits (Parent and Hazrati,

Page 38: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

26

1995a;Hamani et al., 2004). In more general terms, the STN is largely divided into a

dorsolateral sensorimotor portion and a ventromedial associative portion. In primates, the

dorsolateral region was shown to be somatotopically organized. Within this region, the

lateral fraction contains neurons that respond to arm movements and the more medial

fraction contains neurons that respond to leg movement (DeLong et al., 1985;Wichmann

et al., 1994a;Rodriguez-Oroz et al., 2001). The ventromedial portion of the STN is

involved in oculomotor and associative aspects of motor behaviour and the neurons are

activated during visual and oculomotor tasks (Matsumura et al., 1992).

Subthalamic nucleus afferents

The STN receives excitatory glutamatergic projections from diverse regions of the

ipsilateral cortex. In primates, most of the cortical afferents to the STN arise from the

primary motor cortex, supplementary motor area (SMA), pre-SMA, and the dorsal and

ventral premotor cortices (Hamani et al., 2004). These projections, as part of the motor

Page 39: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

27

loop, innervate predominantly the dorsal aspect of the STN and are somatotopically

organized. Primary motor cortex fibres related to the leg, arm and face are represented

from medial to lateral in the lateral portion of the STN. The medial potion receives fibers

from the SMA and premotor cortex in an inverse somatotopic distribution with leg, arm

and face represented from lateral to medial respectively (Nambu et al., 1996;Nambu et al.,

1997;Nambu et al., 2002). The ventromedial STN, that is involved in circuits related to

eye movements, receives afferents from the frontal eye field and the supplementary

frontal eye field (Monakow et al., 1978). In rodents, prelimbic-medial orbital areas of the

prefrontal cortex project to the medial STN as part of the limbic loop (Groenewegen and

Berendse, 1990). Cortical afferents from the primary motor cortex to the STN in rodents

and cats originate mainly in layer V and are composed of collaterals of the pyramidal

tract or cortical fibres that also innervate the striatum (Kitai and Deniau, 1981;Giuffrida

et al., 1985). The cortico-subthalamic projections terminate primarily on distal dendrites

and spines of STN neurons (Moriizumi et al., 1987).

The STN receives major GABAergic inhibitory input from the GPe. The topographic

distribution of these afferents varies from species to species. In rodents, the lateral

portion of the pallidum innervates the lateral STN, while the medial and ventral portions

of the pallidum innervate the medial STN (Parent and Hazrati, 1995a). In primates the

picture is more complex. Generally, motor and limbic portions of the GPe innervate their

corresponding territories in the STN. However, the associative GPe afferents innervate

not only the associative portion of the STN but also the motor territory (Carpenter et al.,

1981;Shink et al., 1996;Joel and Weiner, 1997). Pallidal afferents have been shown to

terminate more commonly on proximal dendrites and cell bodies of STN neurons (Parent

and Hazrati, 1995a). The inhibitory drive of the GPe on the STN is supported by

electrophysiological studies in monkeys showing that muscimol (a GABA receptor

agonist) injections into the STN decrease the activity of STN neurons, while bicuculline

(a GABA receptor antagonist) injections slightly increase the activity of STN neurons

(Wichmann et al., 1994b). It seems possible that pallidal inputs interact with cortical

inputs to shape STN responses (Ryan and Clark, 1991). Cortical stimulation results in a

short latency excitation of the STN followed by an inhibitory period that is likely

Page 40: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

28

mediated by the GPe (Fujimoto and Kita, 1993).

The STN also receives direct projections from the thalamus and these originate mainly

from the CM and Pf nuclei (Sugimoto et al., 1983;Sugimoto et al., 1983;Sadikot et al.,

1992a). In primates, the Pf nucleus is the predominant thalamic source of input to the

STN, innervating mainly the associative and limbic STN territories. On the other hand,

there are only small number of fibres from the CM nucleus, which projects mainly to the

sensorimotor STN (Sadikot et al., 1992a;Parent and Hazrati, 1995a). These afferents are

glutamatergic and terminate mainly on dendrites of STN cells (Mouroux and Feger,

1993).

Brainstem afferents to the STN have also been described. The STN receive direct

dopaminergic projections from the SNc (Brown et al., 1979;Lavoie et al., 1989;Francois

et al., 2000). These projections are known to modulate the cortical and pallidal inputs to

the STN. In addition, the pedunclulopontine and laterodorsal tegmental nuclei send

cholinergic inputs to the STN (Jackson and Crossman, 1983;Scarnati et al., 1987;Lee et

al., 1988;Lavoie and Parent, 1994b), although non-cholinergic inputs from the

pedunculopontine nucleus also exist (Rye et al., 1987). Another source of afferent

innervation to the STN in rodents is the dorsal raphe nucleus (Woolf and Butcher,

1986;Canteras et al., 1990). This pathway is serotoninergic and may also be involved in

the modulation of STN activity.

Subthalamic nucleus efferents

The STN is the only nucleus of the BG that exerts an excitatory glutamatergic drive to its

target nuclei. Its major efferent projections are directed to both segments of the globus

pallidus. These projections arborize uniformly and affect an extensive number of cells

(Smith et al., 1990;Hazrati and Parent, 1992b;Parent and Hazrati, 1995a). The excitatory

STN drive to the globus pallidus is supported by electrophysiological studies in intact

monkeys in which a reduced firing rate in the globus pallidus is observed following STN

lesion (Hamada and DeLong, 1992). In addition to the pallidum, the STN projects to both

component of the substantia nigra, the SNc and the SNr. Although most of these

Page 41: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

29

projections innervate the SNr, some axons innervate the SNc perhaps regulating

dopamine release (Groenewegen and Berendse, 1990;Smith et al., 1990;Parent and

Hazrati, 1995a). The striatum receives scant projections from the STN in non-human

primates (Kita and Kitai, 1987;Smith et al., 1990). These projections provide en passant

excitatory influence over striatal cells (Parent and Hazrati, 1995a). In summary, the

ventromedial associative and limbic regions of the STN innervate mostly the caudate

nucleus and the SNr, while the dorsolateral sensorimotor portion innervate mostly the

putamen and the GP (Parent and Smith, 1987;Parent, 1990).

Aside from the main efferent projections described above, the STN also projects to the

pedunculopontine nucleus and ventral tegmental area (Jackson and Crossman,

1981;Granata and Kitai, 1989;Smith et al., 1990).

Pharmacological properties of STN cells

GABA plays a major role in STN physiology, modulating the firing rates and patterns of

its neurons (Bevan et al., 2007). The GABAergic activity within the STN occurs mainly

through the activation of post-synaptic GABAA receptors (Bevan et al., 2000). Recent

evidence, however, has shown that GABAB receptors also play a role, but their effects are

mostly pre-synaptic (Shen and Johnson, 2001). STN neurons also express glutamate

receptors. Both N-methyl-D-aspartate (NMDA) and 2-aminomethyl-phenylacetic acid

(AMPA) have been described in STN neurons, although AMPA receptors are more

prevalent (Klockgether et al., 1991;Shen and Johnson, 2000).

STN neurons express dopamine D5 receptors, and also D1 and D2 (Baufreton et al.,

2005b), which are also present in the pre-synaptic afferent terminals (Mansour et al.,

1991;Svenningsson and Le Moine, 2002). While some studies suggest that dopamine

agonists have an excitatory effect in STN (Mintz et al., 1986;Kreiss et al., 1996), others

imply that it reduces STN activity (Campbell et al., 1985;Hassani and Feger, 1999). In

STN slices, dopamine application suppresses both glutamatergic EPSPs (excitatory

postsynaptic potentials) and GABAergic IPSPs (inhibitory postsynaptic potentials)

(Hassani and Feger, 1999;Shen and Johnson, 2000). In rodents, although the systemic

Page 42: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

30

effects of selective D1 and D2 agonists are not consistent, it has been shown that systemic

administration of apomorphine, a non-selective dopamine agonist, increases STN activity.

Generally, it seems that co-activation of D1 and D2 receptors excite STN neurons (Kreiss

et al., 1997;Ni et al., 2001). It is important to consider, however, that most of the

structures that innervate the STN are modulated by dopamine. Thus, systemic application

of dopaminergic agents might produce a complex cascade of responses (Hamani et al.,

2004).

Cholinergic agonists applied topically in rodents, increase the activity of STN neurons

(Feger et al., 1979). Application of muscarinic agonists in slices reduce both EPSPs and

IPSPs in the STN, perhaps through M3 receptors (Flores et al., 1996;Shen and Johnson,

2000). Since the effects of the later potentials are greater, the end result is excitation of

STN neurons (Rosales et al., 1994;Shen and Johnson, 2000). Serotonin also increases the

spontaneous activity of STN neurons in slice preparations (Flores et al., 1995). Opioids

have been shown to inhibit both GABAergic and glutamatergic synaptic inputs to the

STN through pre-synaptic and δ receptors (Shen and Johnson, 2002).

Physiological properties of STN cells

In vitro, STN neurons are capable of firing independently of synaptic input due to their

pacemaker voltage-gated sodium channels. In the absence of GABAergic and

glutamatergic synaptic inputs, STN neurons rhythmically fire action potentials at 5-15 Hz

(Bevan and Wilson, 1999). This property may, in part, underlie the tonic activity of STN

neurons in vivo. STN neurons fire 13-18 spikes/sec in normal rodents, and 18-25

spikes/sec in non-human primates (Georgopoulos et al., 1983;DeLong et al.,

1985;Matsumura et al., 1992;Wichmann et al., 1994a;Overton and Greenfield,

1995;Hassani et al., 1996;Bergman et al., 1994).

Most of the neurons in the STN respond to cortical stimulation, usually with a triphasic

response. This response consists of initial excitation, followed by inhibition and a second

excitation (Kitai and Deniau, 1981;Fujimoto and Kita, 1993;Nambu et al., 2000). The

first excitation occurs ~2 ms following cortical stimulation and is believed to be directly

Page 43: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

31

related to the activation of cortico-subthalamic pathways. The following inhibition is

thought to be the result of activation of GPe neurons. The second excitatory peak, which

takes place ~15 ms after cortical stimulation, might be due to the activation of the

cortico-striatal pathway and the subsequent inhibition of GPe neurons, resulting in

disinhibition of the STN (Fujimoto and Kita, 1993;Feger et al., 1997;Nambu et al., 2000).

In agreement with this notion, lesions of the globus pallidus do not interfere with the first

excitation of the STN, but substantially decrease the inhibitory component of the

response and increase the second excitatory peak. Striatal lesions, on the other hand,

induce the opposite effects (Ryan and Clark, 1992).

Nearly all STN neurons respond to pallidal stimulation (Kita et al., 1983a). The pattern of

response depends on the amount and/or amplitude of the IPSPs that are generated

following pallidal stimulation. For instance, large IPSPs are strong enough to completely

reset the autonomous activity of STN neurons and can promote synchronization because

they completely deactivate the voltage-gated sodium channels that underlie autonomous

activity (Bevan et al., 2002a;Bevan et al., 2002b;Baufreton et al., 2005a;Bevan et al.,

2007). On the other hand, individual IPSPs with small amplitudes promote only partial

resetting of STN neurons, which results in phase-dependent delays between spikes,

leading to desynchronization among STN neurons. Importantly, multiple IPSPs do not

only reset STN firing but occasionally promote a rebound depolarization and a burst of

action potentials by bringing the membrane potential closer to the equilibrium potential

of GABA current (to a more hyperpolarized state) (Bevan et al., 2002a;Bevan et al.,

2002b;Bevan et al., 2000). Therefore, synchronous activity of GPe neurons has the

potential to synchronize the activity of STN neurons (Bevan et al., 2000). It has been

suggested that during normal movement, GABAergic activation is asynchronous and

provides only a small contribution to the synchronization of STN. It is possible that

asynchronous feedback inhibition by the GPe acts to limit STN synchronization and

provide the basis for de-correlated, functionally segregated, activity (Bevan et al., 2000).

In normal monkeys, between 30 and 50% of the neurons in STN are related to passive

and/or active movement, and their majority are activated solely by manipulation around a

Page 44: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

32

single contralateral joint (DeLong et al., 1985;Wichmann et al., 1994a). As mentioned

above, most of the movement-related neurons are localized in the dorsal portion of the

STN and are somatotopically organized (Wichmann et al., 1994a). In addition, about

20% of STN neurons are responsive to eye fixation, saccadic eye movements or visual

stimuli. These neurons are located primarily in the ventral portion of the STN and

participate in the visuooculomotor circuits (Matsumura et al., 1992).

1.2.4 - Globus pallidus

In primates, the GP is divided by the internal medullary lamina into the internal (GPi) and

external (GPe) segments. In rodents, the homologue of the GPi is the entopeduncular

nucleus (EPN) that is surrounded by the fibres of the internal capsule, and the GPe is

referred to as the GP (Nambu, 2007). Both GPi and GPe share similar morphology and

their neurons use GABA as their neurotransmitter (Smith et al., 1987). The GPe has

strong interconnections with the STN, while the GPi projects outside the basal ganglia

(Nambu, 2007).

Globus pallidus afferents

The GABAergic striatal afferents are the main input to the GP and they terminate along

the dendrites (Shink and Smith, 1995). As mentioned above, the substance P-containing

MSN neurons mainly project to the GPi, and the enkephalin-containing MSN neurons

mainly project to the GPe. In addition to inhibitory striatal input, the STN sends

excitatory glutamatergic input to both segments of the GP (Carpenter et al., 1981;Joel and

Weiner, 1997). Anatomical tracing techniques demonstrate that subthalamic and

neostriatal axonal terminals make convergent synaptic contact with individual GPi

neurons that project to the thalamus (Bevan et al., 1994b). It was suggested that the STN

uniformly excites a large number of GPi neurons, whereas the striatum exerts a more

specific inhibitory control upon selected subsets of STN-driven pallidal neurons to give a

more localized effect (Parent and Hazrati, 1993). The GPi also receives afferent input

from the GPe that terminates close to the cell body (Shink and Smith, 1995). Other inputs

to the GP include glutamatergic inputs from the intralaminar thalamic nuclei,

Page 45: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

33

serotonergic inputs from the dorsal raphe nucleus, glutamatergic and cholinergic inputs

from the pedunculopontine nucleus, and dopaminergic inputs from the SNc. These axons

arborize more profusely in the GPi than in the GPe (Lavoie et al., 1989;Lavoie and Parent,

1990;Lavoie and Parent, 1994b).

1.2.4.a - Globus pallidus externus

Globus pallidus externus efferents

The GPe sends inhibitory GABAergic projections to many BG nuclei including the STN

GPi, SNr and striatum (Carpenter et al., 1981;Bolam and Smith, 1992;Parent and Hazrati,

1995a). GPe axons in STN, GP and SNr terminate on the soma and proximal dendrites. In

the striatum, GPe axons terminate on aspiny interneurons and the dendritic shafts of spiny

projection neurons (Sato et al., 2000). In addition, GPe local collateral axons form

synapses on the somata and proximal dendrites of other GPe neurons (Kita, 2007). A

small number of GPe neurons also projects to the dorsal thalamus, reticular thalamic

nucleus, inferior colliculus and the PPN (Kita, 2007).

Physiology of the globus pallidus externus

The GPe is considered to be an intrinsic nucleus of the BG because it receives most of its

input from structures of the BG (i.e. the striatum and STN) and sends output to other BG

nuclei (Parent and Hazrati, 1995a). As mentioned above, the GPe receives input from the

enkephalin-containing MSN neurons that primarily express the D2 receptor, and is

viewed to be part of the indirect pathway. Electrophysiological studies demonstrate that

the majority of GPe neurons exhibit high frequency firing interspersed with spontaneous

pauses and have a mean firing rate of about 55 Hz. The rest of the neurons exhibit low-

frequency firing and bursts (DeLong, 1971;Gardiner and Kitai, 1992;Ni et al., 2000a). It

has been reported that 11% of the cells recorded in the GPe display an oscillatory

correlogram in the 5-20 Hz range, and fewer than 5% of recorded pairs are correlated

(Raz et al., 2000). GPe neurons respond to somatosensory input and are involved in the

control of movement parameters (DeLong, 1971;Georgopoulos et al., 1983;Dormont et

al., 1997). Ninety-one percent of GPe neurons have been shown to respond to the passive

movement of only one joint (Filion et al., 1988). The GPe can strongly inhibit the

Page 46: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

34

neurons of the output structures of the BG directly through its connections to GPi/SNr

and indirectly via the STN (Alexander and Crutcher, 1990). It has been shown that

injection of bicuculline into the GPe of intact monkeys, causes increased GPe activity and

elicits dyskinesias (Crossman et al., 1988;Matsumura et al., 1995). Electrical stimulation

of the GPe has been shown to increase movement times of the contralateral arm (Horak

and Anderson, 1984a).

1.2.4.b - Globus pallidus internus

Globus pallidus internus efferents

The GPi projects to the ventral anterior/ventral lateral thalamic complex (i.e. parvicellular

part of the ventral anterior nucleus and oral part of the ventral lateral nucleus), and gives

off axon collaterals to the centromedian nucleus (Kuo and Carpenter, 1973;Kim et al.,

1976). These fibres terminate predominantly on the soma and proximal dendrites of

thalamic projection neurons. Pallidal fibres also contact thalamic GABAergic inhibitory

interneurons, suggesting that pallidal inputs not only directly inhibit thalamic projection

neurons but also disinhibit projection neurons via interneurons (Ilinsky et al., 1997). GPi

neurons also project to the habenular nucleus and PPN (Parent et al., 1981;Parent and De

Bellefeuille, 1983).

Physiology of the globus pallidus internus

The GPi plays a key role in the BG network as it integrates information from the striatum,

GPe and STN. The GPi is part of the indirect pathway since it receives excitatory input

from the STN, as well as part of the direct pathway since it receives inhibitory input from

the substance P-containing striatal neurons (Parent and Hazrati, 1995b). Unlike GPe

neurons that exhibit pauses of activity, GPi neurons fire spontaneously with a mean rate

of 60-80 Hz without pauses (DeLong, 1971;DeLong et al., 1985). This activity is

believed to exert a tonic inhibitory effect on the target nuclei of the GPi (i.e. the thalamus

and brainstem). Little oscillatory firing or synchronized pairing has been demonstrated in

the GPi of normal animals (Nini et al., 1995;Raz et al., 2000). Neurons of the GPi are

relatively large (20-60 µm) (Yelnik et al., 1984) and have impressive dendritic trees that

are believed to play a role in the convergence of incoming information from the striatum

Page 47: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

35

(Parent and Hazrati, 1995b). In addition to the main group of GABAergic GPi neurons,

there is another group of neurons named “border cells” that are located near the perimeter

of pallidal segments, and adjacent to or in the external or internal medullary lamina.

These cells exhibit a regular discharge pattern with a mean frequency of 20-50 Hz, and

are considered to be cholinergic (DeLong, 1971).

Through the hyperdirect, direct and indirect pathways, the GPi receives topographical

inputs from the cerebral cortex. In the motor territories of the GPi (ventral two-thirds of

caudal GPi), the hindlimb, forelimb, and facial regions are somatotopically represented

from the dorsal part to the ventral part (DeLong, 1971;DeLong et al., 1985;Yoshida et al.,

1993;Strick et al., 1995). In normal animals, sensory-motor fields have been shown to

correlate mainly with one joint of the contralateral limb (Filion et al., 1988;Boraud et al.,

2000a). About 70% of responding neurons are activated and 30% inhibited in response to

movement of a corresponding body part (Anderson and Horak, 1985;Mink and Thach,

1991a;Turner and Anderson, 1997;Boraud et al., 2000a). It is important to note that

during voluntary limb/arm movement, a strong majority of GPi neurons display an

increase rather than decrease in firing rate (Georgopoulos et al., 1983;Anderson and

Horak, 1985;DeLong et al., 1985;Mitchell et al., 1987a;Hamada et al., 1990;Nambu et al.,

1990;Mink and Thach, 1991a;Turner and Anderson, 1997). Therefore, it was suggested

that excitatory inputs from the STN to the GPi contribute to voluntary movement (Nambu,

2007). Border cells show similar response pattern as GPi and GPe neurons, suggesting

that they share similar inputs (Mitchell et al., 1987b).

Importantly, movement-related activity of neurons in the GPi occurs not only later than

motor cortex (DeLong et al., 1984a) but also about 80-100 ms after the onset of muscle

contraction, indicating that they are not involved in the preparation or the initiation of

movement (DeLong, 1971;Anderson and Horak, 1985;Mink and Thach, 1987). Studies

on parameter encoding are more divergent. Some studies show a certain correlation

between GPi firing rate and kinematic parameters such as velocity, direction or strength

(Georgopoulos et al., 1983;Anderson and Turner, 1991), while in other studies this

correlation depends on the context (Mink and Thach, 1991a;Mink and Thach,

Page 48: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

36

1991b;Brotchie et al., 1991a). In monkeys performing sequential wrist movements,

pallidal neurons code more prominently for predictable and easy movements and signal

the end of a movement sequence (Brotchie et al., 1991b), suggesting that BG function is

more related to the performance of automatic than voluntary movements. In addition,

pallidal neurons encode the serial position of an event in remembered sequential pointing

movements in monkeys, suggesting that the GPi may be involved in encoding the

spatiotemporal characteristics of a sequence of actions (Mushiake and Strick, 1995).

Injections of bicuculline into the GPi of intact monkeys induces dose-dependent

hypokinesia with dystonic attitudes in contralateral limbs, whereas muscimol injections

elicit choreiform movements (Burbaud et al., 1998). Neurotoxic blockade or lesions of

the GPi lead to poor braking of movement (Horak and Anderson, 1984b;Mink and Thach,

1991b) and inaccuracy of control of the final adjustment of the limb to the target (Kato

and Kimura, 1992). It has been shown that electrical stimulation in the ventrolateral GPi

reduces movement times for the contralateral arm of normal monkeys, whereas

stimulation in the dorsal GPi increases movement times (Horak and Anderson, 1984a).

1.2.5 - Substantia nigra

The substantia nigra is located dorsal to the cerebral peduncle in the ventral midbrain,

and plays an important role in reward, addiction and movement. Although the substantia

nigra appears as a continuous band in brain sections, it consists of two anatomically

distinct parts with very different connections and functions, the pars compacta and pars

reticulata (Rinvik and Grofova, 1970;Gulley and Wood, 1971;Juraska et al., 1977;Yelnik

et al., 1987;Braak and Braak, 1986).

1.2.5.a - Substantia nigra pars compacta

The pars compacta of the substantia nigra receives GABAergic input from striatal

neurons and sends dopaminergic input back to the striatum (Graybiel, 1990). In addition,

collaterals of the nigrostriatal axons provide extrastriatal dopaminergic innervation of the

pallidum and STN (Cossette et al., 1999). The dopaminergic neurons of the SNc

Page 49: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

37

modulate the flow of information from the cortex through the BG (Bevan et al., 1996). In

intact monkeys, dopaminergic neurons have a relatively low discharge rate (0.5-8 Hz)

and long impulse duration (2.05 ms on average) (Aebischer and Schultz, 1984). It would

appear that these neurons are not related to movement itself, but change their activity in

response to rewards, during procedural learning and in response to conditioned stimuli

(DeLong et al., 1983;Schultz, 1986b). Their activity can also be depressed by systemic

injection of the dopamine agonist apomorphine (Aebischer and Schultz, 1984).

In addition to striatal input, SNc dopaminergic neurons are innervated by several other

BG nuclei. In the rat, SNc neurons receive multiple GABAergic inputs from neurons in

the globus pallidus and from axon collaterals of SNr projection neurons (Smith and

Bolam, 1990;Celada et al., 1999). Electrical stimulation of GABAergic pathways

originating in the striatum, GP or SNr produces inhibition of SNc dopaminergic neurons

in vivo (Paladini et al., 1999). This inhibition appears to be mediated, at least in part, by

the GABAA receptor subtype. Interestingly, it has been shown that the discharge pattern

of SNc dopaminergic neurons in vivo can be differentially modulated by local application

of GABAA and GABAB receptor antagonists. The GABAA antagonists (i.e. bicuculline,

gabazine and picrotoxin) strongly induce burst firing in dopaminergic neurons, whereas

GABAB antagonists cause a modest shift to a more regular firing in 50% of the cases

(Paladini and Tepper, 1999). This suggests that dopaminergic neurons are under tonic

GABAergic inhibition mediated by GABAA receptors. On the other hand, STN

stimulation has been shown to promote burst discharge in a subpopulation of SNc

neurons, most likely via NMDA receptors (Chergui et al., 1993;Chergui et al., 1994),

suggesting that excitatory inputs to midbrain dopaminergic neurons may participate in the

control of the dopamine release in target areas (Chergui et al., 1993).

1.2.5.b - Substantia nigra pars reticulata

Together with the GPi, the pars reticulata of the substantia nigra provides the main output

nucleus of the BG. The SNr receives GABAergic inputs from the striatum (Hedreen and

DeLong, 1991;Mendez et al., 1993;Lynd-Balta and Haber, 1994) and GPe (Mendez et al.,

1993;Bevan et al., 1996), and glutamatergic inputs from the STN (van der Kooy and

Page 50: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

38

Hattori, 1980;Smith et al., 1990). These inputs from striatum, GPe and STN converge

onto single SNr neurons (Smith and Bolam, 1991;Bolam et al., 1993;Bevan et al., 1994a).

The lateral half of the SNr is innervated by striatal subterritories that process inputs from

sensory and motor cortical areas, whereas the more medial region is innervated by striatal

subterritories that are related to prefrontal and limbic cortical areas (Deniau et al., 2007).

The SNr innervates the ventral medial and parafascicular thalamic nuclei, the superior

colliculus and the PPN (Beckstead et al., 1979;Di et al., 1979;Gerfen et al., 1982;Appell

and Behan, 1990;Bickford and Hall, 1992).

Compared to the overlying cell-rich pars compacta, the SNr is characterized by a low

neuronal density with smaller cell bodies (Francois et al., 1985;Yelnik et al., 1987).

Although the SNr is mainly composed of GABAergic projection neurons (Oertel and

Mugnaini, 1984;Yelnik et al., 1987;Smith et al., 1987), dopaminergic and cholinergic

neurons also exist (Hattori, 1993;Deutch et al., 1986;Decavel et al., 1987;Campbell and

Takada, 1989;Gould and Butcher, 1986;Martinez-Murillo et al., 1989). In rats,

dopaminergic neurons are clustered in the caudal and ventral SNr (Deutch et al., 1986),

while the cholinergic neurons are mainly located caudally (Gould and Butcher,

1986;Martinez-Murillo et al., 1989). The dopaminergic neurons are distinct from and not

considered part of the SNc. In addition to projection neurons, putative local circuit

neurons have been also described in the rat (Gulley and Wood, 1971;Juraska et al., 1977)

and monkey SNr (Schwyn and Fox, 1974). These neurons, however, appear to be very

few in number and their functional role is still unknown (Deniau et al., 2007).

Functional properties of the SNr

As opposed to the dopaminergic nigrostriatal neurons, the GABAergic output neurons of

the SNr are characterized by short-duration action potentials and a spontaneous repetitive

high-frequency firing rate reaching 40-80 spikes/sec in vivo (Wilson et al., 1977;Deniau

et al., 1978). These electrophysiological properties of SNr neurons are attributed to their

strong voltage-dependent K+ conductance, which allows for a short duration action

potential and prevents the membrane potential from reaching the Na+ inactivation level

(Nakanishi et al., 1987). No synchronized oscillations have been shown in the SNr of

Page 51: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

39

normal animals (Wichmann and DeLong, 1999). SNr neurons show phasic responses

during motor tasks and with somatosensory examination in monkeys (Mora et al.,

1977;DeLong et al., 1983;Schultz, 1986a;Lestienne and Caillier, 1986), cats (Joseph et al.,

1985) and rats (Gulley et al., 1999;Gulley et al., 2002). The SNr is also involved in

oculomotor functions (Hikosaka and Wurtz, 1983c;Hikosaka and Wurtz, 1983a;Joseph

and Boussaoud, 1985). For example, it has been shown that prior to rapid eye movements,

SNr neurons display a pause in activity, while superior colliculus neurons show a burst in

activity (Hikosaka and Wurtz, 1989). Almost all task-related SNr neurons are inhibited

during saccadic eye movement, suggesting that the pause in SNr provides a transient

disinhibition of the superior colliculus, leading to an eye movement command (Hikosaka

et al., 2000). It has been also shown that an injection of the GABA agonist muscimol into

the SNr disrupts guided and reflex eye movements in the cat (Boussaoud and Joseph,

1985). From more recent work, it is becoming clear that the SNr is involved in higher

motor functions and cognitive processes in addition to its role in movement (Hikosaka

and Wurtz, 1983b;Basso and Wurtz, 2002;Bayer et al., 2004;Wichmann and Kliem,

2004).

1.3 - Models of basal ganglia function in relation to Parkinson’s disease

Models of basal ganglia circuitry have provided some clues into the pathophysiological

mechanisms of movement disorders and the reasons that disruption of parts of the circuit

by lesion or deep brain stimulation might improve motor function. The main three models

will be discussed here: the classical rate model proposed over 15 yeas ago, the centre-

surround model and the relatively recent oscillatory model. A more comprehensive

review of the electrophysiological findings in PD patients and animal models of PD will

be given in section 1.4.

1.3.1 - The rate model

The rate model was first proposed by Albin (Albin et al., 1989) and Delong (DeLong,

1990). According to this model, the basal ganglia facilitate and inhibit movement through

Page 52: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

40

changes in firing rates in its nuclei. This model, which is largely based on the serial

connections between BG nuclei (see Figure 1.F1), provides a coherent explanation of

opposing movement disorders (i.e. hypo- and hyperkinetic) (DeLong, 1990). The rate

model proposes that activation of the direct pathway causes decreased activity in GPi/SNr

and facilitation of movements through a process of disinhibition of thalamocortical relay

neurons (Chevalier and Deniau, 1990). On the other hand, activation of the indirect

pathway disinhibits the STN and causes increased activity in GPi/SNr, therefore leading

to suppression of movements. In other words, the rate model predicts that decreased

GPi/SNr activity would result in movements that are fast and large and that increased

GPi/SNr activity would result in movements that are slow and small (Alexander and

Crutcher, 1990). A balance between the direct and indirect pathways is believed to exist

at the level of the output nuclei of the BG whereby the inhibitory effect of the direct

striatal output counteracts the excitatory effect of the STN in the indirect pathway.

The rate model in relation to Parkinson’s disease

As mentioned earlier, the neurotransmitter dopamine acts in the striatum on D1 receptors

to stimulate cells of the direct pathway, and on D2 receptors to inhibit cells of the indirect

pathway. In Parkinson‟s disease, the result of the dopamine deficiency is reduced activity

in the direct pathway from striatum to GPi, and increased activity in the indirect pathway

via GPe and STN. Together, these changes give rise to increased activity in the GPi/SNr,

inhibiting thalamic and cortical activation and impairing voluntary movement.

The rate model of the basal ganglia predicted that ablation or inactivation of GPi or STN

would lead to the alleviation of parkinsonian symptoms. Indeed, STN lesions have been

shown to reverse parkinsonian symptoms in MPTP monkeys (Bergman et al., 1990;Aziz

et al., 1991;Guridi et al., 1994) and set the stage for the introduction of the STN as a

neurosurgical target to treat PD patients (Benabid et al., 1994;Gill and Heywood,

1997;Krack et al., 2003). Even pallidotomy which had been largely abandoned since the

advent of levodopa therapy, was revisited and proven to be very successful at alleviating

PD symptoms in patients (Laitinen et al., 1992b;Lozano et al., 1995;Vitek and Bakay,

1997).

Page 53: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

41

Shortcomings of the rate model

If the rate model is correct, one would expect that BG-related hypo- and hyperkinetic

disorders would show the opposite pathological changes in BG. Indeed, increased

indirect pathway activity has been reported in PD patients (Hutchison et al.,

1994;Hutchison et al., 1998) and in MPTP-treated monkeys (a model of PD) (Bergman et

al., 1994). Increased activity in the direct pathway has also been demonstrated in

hyperkinetic pathophysiology such as in dystonia (Vitek, 2002;Starr et al., 2005).

However, a lesion in the GPi (pallidotomy), which reduces the inhibitory BG output to

the thalamus, improves symptoms for both hypokinetic and hyperkinetic disorders

(including levodopa-induced dyskinesias in PD) (Marsden and Obeso, 1994a). This is

inconsistent with the rate model that holds that the mechanisms underlying hypo- and

hyperkinesia are reciprocal. Moreover, recent studies of firing rates in GPi of patients

with various movement disorders fail to confirm the predictions of the rate model

(Hutchison et al., 2003;Tang et al., 2005). In addition, several studies in dopamine

depleted primates have shown no change in spontaneous pallidal (Boraud et al., 2002),

thalamic (Pessiglione et al., 2005) or motor cortical firing rates (Doudet et al.,

1990;Watts and Mandir, 1992;Goldberg et al., 2002).

The evolving picture of the basal ganglia anatomy is much more complex than the

simplified view of the direct and indirect pathways. As previously mentioned in section

1.2.2, striatal neurons projecting to the GPi and SNr can also send axon collaterals to the

GPe, suggesting that the direct and indirect pathways are not completely segregated

(Kawaguchi et al., 1990;Parent et al., 1995;Parent et al., 2000;Wu et al., 2000;Levesque

and Parent, 2005b). In addition, the rate model does not take into consideration the direct

projections from the motor cortex to the STN (Nambu et al., 2000), the reciprocal

connections between the STN and the GPe (Parent and Hazrati, 1995a), the feedback

projections from the GPe to the striatum, the feed-forward projections from the cortex to

the GABAergic interneurons of the striatum (Bolam et al., 2000), and the output to / input

from the PPN (Lavoie and Parent, 1994b). Finally, the rate model does not take into

account the direct effect of dopamine on STN and GPi (Parent and Cossette, 2001).

Page 54: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

42

The rate model is a static model that attempts to explain PD akinesia/bradykinesia in

terms of changes in tonic firing rates. However, the model fails to account for the

dynamic symptoms of PD, rigidity and resting tremor. Yet, we should keep in mind that

despite the limitations discussed above, the rate model has been the predominant and

pervading model over the last 20 years, as it shaped general thought on how the basal

ganglia function.

1.3.2 - The centre-surround model

The centre-surround model, proposed by Mink (Mink, 1996;Mink, 2003), suggests that

the basal ganglia functions through centre-surround mechanisms. Anatomical studies

have shown that afferents from the STN arborize more widely and excite a large number

of neurons within the GP, while the striatum exerts a more focused inhibition on a subset

of these neurons (Hazrati and Parent, 1992a;Hazrati and Parent, 1992b). Based on these

observations, the centre-surround model suggests that the striatal input to the GPi and

SNr provides a specific, focused, context-dependent inhibition, whereas the input from

STN provides a less specific, divergent excitation (see Figure 1.F3). Since the output

from GPi/SNr is inhibitory, it is converted into focused facilitation and surround

inhibition of motor programs in thalamocortical and brainstem circuits. This organization

enables selective facilitation of desired movements while preventing potentially

competing movements from interfering with the one selected.

This model is essentially different from the rate model which views the role of the

inhibitory and excitatory pathways to GPi/SNr to be the scaling of parameters for the

intended movement in a “push-pull” manner. In contrast to the rate model, the scheme

proposed in the centre-surround model views the role of the inhibitory input to GPi/SNr

to be the selection of the desired motor programs and the role of the excitatory input to

GPi/SNr to be the inhibition of competing motor programs with neither being involved in

the scaling of specific movement parameters (Mink, 1996).

Page 55: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

43

The centre-surround model in relation to Parkinson’s disease

According to the centre-surround model, abnormally increased GPi/SNr activity in

Parkinson‟s disease should lead to excessive inhibition of all motor programs and

inability to completely turn on those involved in the desired movement. The result would

be slowing of movement, decreased amplitude of movement, and in some cases an

inability to move at all. In addition to changes in tonic firing rates, abnormal phasic

activity of GPi neurons (Tremblay et al., 1989;Bergman et al., 1994) will result in an

inability to increase the GPi output appropriately during movement to inhibit competing

movements. This would explain the abnormal postures, co-contraction rigidity, and

inability to suppress unwanted postural reflexes in PD. Thus, according to the present

model, the abnormality in PD is two-fold. First, there is an inability to remove inhibition

from desired motor programs and second, there is an inability to fully inhibit competing

Page 56: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

44

motor programs (Mink, 1996).

Shortcomings of the centre-surround model

Although the centre-surround model has some appealing features, there is no direct

physiological proof of it. The model predicts a positive correlation between the discharge

of pallidal neurons that participate in the same action, and a negative correlation between

pallidal neurons that participate in competing actions. Physiological studies, however,

have failed to reveal such relationships between simultaneously recorded basal ganglia

neurons (Jaeger et al., 1994;Nini et al., 1995;Raz et al., 2000;Bar-Gad et al., 2003).

Moreover, changes in discharge in pallidal neurons often lag behind movement initiation

(DeLong, 1971;Anderson and Horak, 1985;Turner and Anderson, 1997).

1.3.3 - The oscillatory model

Oscillatory activity in the BG has attracted a great deal of interest in the past decade as it

is thought to be important in both the normal functioning of the system as well as in the

pathophysiology of PD. Findings in PD patients and animal models of PD have suggested

that loss of dopamine in the striatum leads to excessive synchronized oscillatory activity

in the BG which may underlie the clinical features of PD (Raz et al., 1996;Raz et al.,

2000;Nini et al., 1995;Bevan et al., 2000;Boraud et al., 2002;Brown, 2003;Levy et al.,

2002b) such as akinesia/bradykinesia (Chen et al., 2007;Kuhn et al., 2006b;Brown, 2006)

and limb tremor (Bergman et al., 1998b;Deuschl et al., 2000;Rivlin-Etzion et al.,

2006a;Levy et al., 2000).

Oscillations of single neurons, as well as the synchronized oscillations within populations

of neurons as indicated by oscillations in local field potentials (LFPs), have been studied

in the GP and STN of PD patients. Recordings in patients withdrawn from their

antiparkinsonian medication have consistently revealed prominent oscillations between

11 Hz and 30 Hz, so-called the „beta band‟ (Brown et al., 2001;Levy et al., 2000;Levy et

al., 2002b;Kuhn et al., 2005). (This frequency range is broader than that termed beta in

clinical studies of scalp electroencephalographic activity.) Pathological oscillatory

Page 57: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

45

activity in the tremor frequency (3-7 Hz) has also been observed in the basal ganglia

network after MPTP treatment in some monkeys (Bergman et al., 1994;Bergman et al.,

1998b) and in PD patients with tremor at rest (Hutchison et al., 1997;Levy et al.,

2002b;Hurtado et al., 1999;Lemstra et al., 1999).

Beta oscillations in LFPs recorded from DBS electrodes implanted in the STN of PD

patients were shown to be coherent with cortical electroencephalographic (EEG) activity

(Marsden et al., 2001;Williams et al., 2002) and to be intimately related to behavior. The

power of beta oscillations decreases with the preparation/execution of voluntary

movements (Cassidy et al., 2002;Williams et al., 2005;Kuhn et al., 2004;Amirnovin et al.,

2004;Williams et al., 2003;Doyle et al., 2005a;Foffani et al., 2005c;Priori et al., 2002)

and following dopamine replacement therapies such as levodopa administration (Brown

et al., 2001;Williams et al., 2002;Priori et al., 2004;Doyle et al., 2005a;Levy et al., 2002a).

It has been shown that the timing of the movement-related beta desynchronization in

LFPs correlates with the timing of voluntary movement (Williams et al., 2005). This,

together with the relative lack of beta LFP activity in the STN and GP in healthy animals,

led to the hypothesis that beta activity in untreated patients is pathologically exaggerated

and might be related to motor impairment. In contrast to beta oscillations, the power of

LFP oscillations in the gamma range (35-90 Hz) has been shown to increase following

dopaminergic medications and also before and during movement (Cassidy et al., 2002).

Moreover, the increase in gamma activity before movement onset is enhanced by

levodopa treatment (Androulidakis et al., 2007).

The oscillatory model for basal ganglia action proposed by Brown (Brown, 2003)

suggests that basal ganglia activity may be synchronized in multiple frequency bands,

each with different functional significance. In this model, beta activity plays an „anti-

kinetic‟ role that originates from the cortex while gamma activity, originating from the

BG, is „pro-kinetic‟ by virtue of its facilitation of motor cortical interaction in the gamma

band (Figure 1.F4). This theory resolved the paradox of pallidotomy (an effective

procedure for both hypo- and hyperkinetic movement disorders) by suggesting that

pallidotomy removes abnormally high BG output in the beta and gamma bands and

Page 58: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

46

therefore ameliorates motor symptoms in both types of disorders.

The directions of connectivity in the different frequency bands, shown in Figure 1.F4,

were initially speculated by Brown (2003) according to studies of coherence and phase

relationships between basal ganglia and cortex. In this regard, it is important to note that

phase estimates may be ambiguous in systems with bidirectional coupling, such as basal

ganglia-cortical loops (Cassidy and Brown, 2003). Indeed, later studies in anaesthetized

rats (Sharott et al., 2005a) and PD patients (Lalo et al., 2008) have confirmed

bidirectional pattern of cortico-basal ganglia communication with a net cortical driving of

STN activity at frequencies below 60 Hz. The flow of direction was found to be

symmetrical for frequencies between 65 to 90 Hz (Lalo et al., 2008).

How might prominent synchronization in the beta band impair movement?

Generally, the information encoded by a correlated network is smaller than the sum of

information encoded by its single elements because of the mutual information shared

between the neurons (Brown, 2007;Hammond et al., 2007). It is therefore possible that in

PD, the exaggerated beta synchronization limits the ability of neurons to code

Page 59: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

47

information in time and space, as both adjacent and spatially distributed neurons are

preferentially locked to the beta rhythm (Brown, 2007;Hammond et al., 2007). Only

when released from this rhythm, these neurons might more effectively engage in the

dynamic assembly formation and rate coding underlying the processing of movement.

Recent recordings in primates confirm an inverse relationship between oscillatory LFP

activity in the beta band and focal neuronal task-related activity, so that oscillations are

preferentially suppressed in the local area of the striatum showing task related increases

in discharge rate (Courtemanche et al., 2003). Parallel observations have been made in

the motor cortex, as exemplified by the „clamping‟ of cortical single unit firing rates to a

relatively narrow range during periods of 20-40 Hz oscillatory synchrony (Murthy and

Fetz, 1996) and the tendency of firing rate modulation to occur as oscillations decrease in

motor cortical LFP (Donoghue et al., 1998). In line with this, a number of studies have

demonstrated the capability of synchrony to limit the coding capacity of neuronal

populations in the cerebral cortex under certain circumstances (Stevens and Zador,

1998;Salinas and Sejnowski, 2000;Svirskis and Rinzel, 2000;Mazurek and Shadlen,

2002). Thus, it is possible that beta oscillatory activity in the basal ganglia might

similarly limit the information coding and therefore impair movement (Brown and

Williams, 2005). Indeed, it has been recently shown that STN neurons in PD demonstrate

reduced response variability when compared to STN neurons in the normal monkey brain

(Gale et al., 2009). Moreover, in dopamine depleted rats, there is a significant decrease in

GPe network entropy (Cruz et al., 2009). These data suggest that PD is associated with a

reduction in the physiological degrees of freedom of BG neurons with diminished

information carrying capacity. Because in PD the baseline level of synchrony is elevated

and is relatively resistant to suppression, the “reading” of cortical information can be

impaired. The difficulty to initiate movements might therefore be the result of the

inability of motor commands for initiation to override the enhanced oscillatory state.

As opposed to beta synchronization, the synchronization in the gamma band may share a

similar role to that posited for gamma band synchronization in the visual cortex. Gamma

synchronization in the visual cortex is believed to promote precise spike timing of single-

Page 60: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

48

cell responses for information processing, even among spatially remote parts of an

assembly (Freiwald et al., 2001). Given the reciprocal relationship between beta and

gamma activities in STN LFPs (Fogelson et al., 2005a), and with respect to dopaminergic

medications and movement, it is possible that extensive synchronization in the beta band

interferes with the involvement of neuronal assemblies in a different pattern of

synchronization (i.e. gamma) that is directly involved in information transfer (Brown and

Williams, 2005). Indeed, it has been recently shown that elevations of STN gamma LFP

activity do not only influence the relative timing of spikes in spike trains (Trottenberg et

al., 2006), but also have a major effect on information carrying capacity (Pogosyan et al.,

2006). It has been therefore suggested that subthalamic gamma oscillations can increase

the information that can be transmitted and decoded by neurons downstream of STN

(Foffani and Priori, 2007).

Shortcomings of the oscillatory model

Although the oscillatory model is based on physiological observations, it fails to provide

physiological mechanisms as for why beta synchronization could be detrimental and

gamma synchronization beneficial. One possible explanation can be based on the fact

that the properties of oscillatory networks are the result of the physical architecture of the

network and the limited speed of neuronal communication (due to axon conduction and

synaptic delays) (Buzsaki and Draguhn, 2004). Because the period of oscillations is

constrained by the size of the neuronal pool engaged in a given cycle, higher frequency

oscillations are confined to a small neuronal space, whereas very large networks can be

potentially recruited during slow oscillations. It is therefore possible that the relatively

slower beta cycle can be expressed over a larger neuronal space and can thus limit the

information carrying capacity of the network. Gamma oscillations, on the other hand, are

fast and can only be exerted locally to increase the responsiveness to selective inputs.

Recent studies in animal models of PD present a serious challenge to the oscillatory

model. In monkeys that underwent a gradual MPTP-treatment protocol that slowly

induces parkinsonism, synchronous oscillatory firing in the GPi occurs after the first

appearance of bradykinesia and akinesia (Leblois et al., 2007), indicating that oscillatory

Page 61: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

49

firing does not contribute to early parkinsonism. Similarly, beta oscillations in STN and

cerebral cortex were not exaggerated until several days after 6-OHDA injections (Mallet

et al., 2008) and the appearance of akinesia (Degos et al., 2008) in rats. Moreover, it has

been shown that while motor symptoms (i.e. rigidity and tremor) are improved in PD

patients following administration of anticholinergic drugs, beta oscillatory activity is

actually increased (Priori et al., 2004).

Considering the rate, centre-surround and oscillatory models, it appears that none of these

models is truly exclusive. Another approach to modeling the BG puts forward that the

normal dopaminergic system supports segregation of the functional subcircuits within the

BG, and that breakdown of this independent processing (by increased synchronization

between neurons) could play an important role in the pathophysiology of PD (Bergman et

al., 1998a). Indeed, many studies have demonstrated that PD is associated with dramatic

loss of spatial selectivity in various regions within the BG-thalamocortical circuit

including the STN (Bergman et al., 1994;Abosch et al., 2002), globus pallidus (Filion et

al., 1988;Boraud et al., 2000a), thalamus (Pessiglione et al., 2005) and cortex (Goldberg

et al., 2002).

Taken together, it is likely that parkinsonism arises from a combination of abnormalities

in neuronal activity. Increased firing rate in BG output together with increased oscillatory

activity and loss of selectivity can give rise to the pathophysiology of PD.

1.3.4 - Tremor models

Although the critical role of nigrostriatal dopamine depletion in generation of PD

symptoms is well accepted, the pathophysiological origin of parkinsonian rest tremor is

still unclear. Two possible mechanisms have been proposed so far, the peripheral and

central mechanisms (Bergman and Deuschl, 2002;Deuschl et al., 2000).

The peripheral mechanism hypothesis states that parkinsonian tremor results from an

unstable long-loop transcortical reflex arc (Oguztoreli and Stein, 1976). Imposing

Page 62: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

50

sinusoidal mechanical movements on the trembling limb can entrain the frequency of

parkinsonian tremor with an EMG discharge that is indistinguishable from a reflex

response (Rack, 1987;Rack and Ross, 1986) suggesting that peripheral reflexes are

important in parkinsonian tremor. Additionally, parkinsonian tremor has been

investigated in patients with the joints fixed in a cast (Burne, 1987). This has been shown

to reduce tremor amplitude and change tremor frequency and, when the limbs were

rigidly fixed, to stop tremor completely. It has been therefore suggested that PD resting

tremor may result from oscillatory “flip-flop” inhibitions between antagonist muscles due

to reduced imbalance between their spinal stretch reflexes. The supraspinal contribution

to the tremor was suggested to result from non-oscillatory descending facilitation of

spinal reflex pathways (Burne, 1987).

However, it becomes more and more clear that spinal reflexes play only a minor role for

the generation of PD tremor (Bergman and Deuschl, 2002;Deuschl et al., 2000). Early

clinical studies have demonstrated that the removal of the dorsal roots in a patient with

parkinsonian tremor reduces the tremor amplitude but only slightly changes its frequency

and cannot stop the tremor (Pollock and Davis, 1930). Several more recent studies show

no frequency reduction after loading of the trembling limb in PD patients (Homberg et al.,

1987;Timmer et al., 1993;Timmer et al., 1996). Other series of studies have dealt with the

modulation of parkinsonian tremor using different stimuli such as mechanical

perturbations (Lee and Stein, 1981;Britton et al., 1992), electrical stimulation of the

median nerve (Britton et al., 1993b) or magnetic stimulation of the motor cortex (Britton

et al., 1993a;Pascual-Leone et al., 1994). These studies failed to show a consistent

resetting of the tremor when stimulating the periphery but a complete resetting when the

cortex was stimulated. Thus, although PD tremor may be modulated by peripheral

manipulations, it is not likely to be generated by peripheral mechanisms. Furthermore, it

has been suggested that peripheral inputs can alter the tremor by modulating central

oscillators (Elble et al., 1992).

It seems that the role of central generators is more important for the generation of PD

tremor (Bergman and Deuschl, 2002;Deuschl et al., 2000). The central oscillator

Page 63: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

51

hypothesis proposes that tremor is caused by oscillations in central structures in the

absence of sensory input. The observation that deafferentation changes the frequency of

parkinsonian tremor but does not suppress it (Pollock and Davis, 1930) strongly supports

a central origin of PD tremor. In PD patients with tremor, PET analysis demonstrates

increased metabolic activity in a network comprising the thalamus, pons, and premotor

cortical regions, indicating increases in the functional activity of thalamo-motor cortical

projections (Antonini et al., 1998). Consistent with this hypothesis, it has been

demonstrated that magnetic brain stimulation over the motor cortex can modulate limb

tremor in patients with PD (Britton et al., 1993a;Pascual-Leone et al., 1994).

Studies linking brain and muscle rhythms in PD provide supportive evidence for the role

of central mechanisms in tremor. In PD patients, the normal neuromagnetic 10-Hz

rhythm is suppressed during periods of tremor (Makela et al., 1993), and 3 to 6 Hz

tremor-related magnetic activity can be observed over wide cortical areas (Volkmann et

al., 1996). Many studies have revealed a correlation between electrical activity of neurons

in the central nervous system and tremor in PD (see section 1.4). These correlation

studies, however, cannot prove causal relationship and central nervous system activation

could be the result of abnormal feedback from the periphery.

Different lesions within the central nervous system can suppress PD tremor. Lesioning of

the motor cortex or the internal capsule have been successful in suppressing tremor but

have produced other drastic side effects like paralysis (Putnam, 1940). The thalamus or

the zona incerta have been successful targets during stereotactic procedures (HASSLER

et al., 1960), and more recently it has been shown that chronic stimulation of these areas

and also of the STN and GPi are able to efficiently suppress PD tremor (Benabid et al.,

1991;Benabid et al., 1994;Hubble et al., 1997;Koller et al., 1997;Pahwa et al.,

1997;Limousin et al., 1998;Plaha et al., 2008;Plaha et al., 2006;Diamond et al., 2007).

Interestingly, the Vim is an effective surgical site for treating Parkinson tremor (Benabid

et al., 1991;Bakay et al., 1992;Kumar et al., 1999a;Lozano, 2000), although it receives

inputs from cerebellum and possibly from ascending spinal tracts, but not from the basal

ganglia (Inase and Tanji, 1995). Furthermore, the suppression of parkinsonian tremor

Page 64: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

52

with Vim stimulation is associated with reduced blood flow in the cerebellum (Deiber et

al., 1993), suggesting that transcerebellar pathways are entrained by the Parkinson tremor

oscillator. Removal of the cerebellum, however, does not prevent PD tremor (Deuschl et

al., 1999).

The motor portion of the GPi projects to ventralis oralis posterior (Vop), which is

adjacent to Vim (Inase and Tanji, 1995). Stereotactic destruction of Vop does not

suppress tremor but is more effective in the treatment of rigidity and bradykinesia (Bakay

et al., 1992). Nevertheless, like the Vim, the Vop contains many neurons that fire in

correlation with tremor (Lenz et al., 1994).

It seems that the preservation of some loops within the nervous system is critical for the

occurrence of parkinsonian tremor. The question where this abnormal activity is

generated is still open. The most likely location for such oscillating neuronal processes is

within the basal ganglia loop (Bergman and Deuschl, 2002;Deuschl et al., 2000) (also see

section 1.4.5). It is also possible that the origin of tremor oscillations might involve both

peripheral and central mechanisms (Oguztoreli and Stein, 1979;Elble et al., 1992;Elble,

1996). Peripheral input might resonate with central structures and affect the strength or

phase of oscillation (Lee and Stein, 1981;Rack and Ross, 1986;Lenz et al., 1993).

1.4 - Electrophysiological and behavioral findings in Parkinson’s disease patients

and animal models of PD

Over the years, a broad variety of experimental models of PD were developed and

applied in diverse species. The most common "classical" toxin-induced PD models are

the 6-hydroxy-dopamine (6-OHDA) rat model and the MPTP monkey model. The rat

model is produced by injection of the toxin 6-OHDA into the SNc and/or the medial

forebrain bundle, which produces a massive lesion of nigral dopaminergic cell bodies

within 2-3 days (Blandini et al., 2008). The neuronal damage induced by 6-OHDA is

mainly due to the massive oxidative stress caused by the toxin. Being similar to

dopamine, 6-OHDA enters the neurons via the dopamine reuptake transporters and

Page 65: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

53

accumulates in the cytosol (where it promotes free radical formation) and mitochondria

(where it inhibits the activity of the electron transport chain) (Schober, 2004). The

injection of the toxin is commonly carried out unilaterally, with the contralateral

hemisphere serving as control. The administration of dopamine agonists (i.e.

apomorhine) produces a rotational behaviour contralateral to the lesioned side.

The MPTP-treated non-human primate model of PD is the most reliable model of PD

discovered (Burns et al., 1983;Langston et al., 1984;Kopin and Markey, 1988;DeLong,

1990;Bloem et al., 1990). As previously mentioned in section 1.1, administration of

MPTP leads to the selective loss of dopaminergic neurons in the SNc since MPTP is

metabolized by monoamine oxidase B to the toxin MPP+, which is then taken up by

dopaminergic neurons in SNc. The most commonly used administration modes in

monkeys are multiple intraperitoneal or intramuscular injections, as well as intracarotid

infusions (Petzinger and Langston, 1998). Similar to humans, a decrease of greater than

95% of striatal dopamine in monkeys with SNc neurodegeneration is required to produce

parkinsonian motor abnormalities (Elsworth et al., 2000). These parkinsonian monkeys

display bilateral limb dystonic postures, rigidity, and bradykinesia. Unilateral intracarotid

infusion is also possible and causes mostly symptoms in the contralateral side. These

animals circle towards the injured side, whereas treatment with dopaminergic

medications induces circling towards the intact side (Bankiewicz et al., 1986).

The MPTP model produces symptoms that closely resemble those found in patients with

PD including levodopa-induced dyskinesia and response fluctuations that are commonly

observed in the later stages of PD (Clarke et al., 1987;Clarke et al., 1989). However, not

all primate species develop similar parkinsonian symptoms. This is true especially with

regard to the presentation of rest tremor (Bergman et al., 1998b). Vervet (African green)

monkeys develop 3-6 Hz rest tremor that can be aggravated by stress (Bergman et al.,

1990) thereby closely resembling that found in patients with PD. Rhesus monkeys, on

the other hand, do not develop low frequency resting tremor but instead exhibit 10-16 Hz

action/postural tremor (Burns et al., 1983).

Page 66: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

54

Although most research has been done on experimental models of PD, the relatively

recent renewal of interest in the surgical treatment of PD (Obeso et al., 1997;Quinn and

Bhatia, 1998;Gross et al., 1999) has allowed an access to the human BG, since most

surgical teams use microelectrode recordings to determine the exact location of the target.

In addition, the advent of non-invasive techniques, such as PET, EEG, magnetic

resonance imaging (MRI), magnetoencephalography (MEG), and transcranial magnetic

stimulation (TMS), have also contributed to our knowledge of neurophysiology of PD in

humans (Berendse and Stam, 2007;Rothwell, 2007;Nandhagopal et al., 2008;Brooks,

2008).

1.4.1 - Single unit recordings and inactivation studies

1.4.1.a - Striatum

A few studies were carried out on striatal activity in animal models of PD. One study has

reported a decrease in the firing rate of striatal tonically and phasically active neurons

(from ~6 to ~4 spikes/sec) following the development of parkinsonian symptoms in

MPTP-treated monkeys (Yoshida et al., 1991). However, more recent reports have shown

no change in the firing rate of TANs. Instead, there was an emergence of 3-19 Hz

oscillatory activity in TANs that was correlated with neuronal activity in the GPi, and an

increase in the number of synchronized pairs of TANs (Raz et al., 1996;Raz et al., 2001).

These changes have been proposed to result from enhanced electrotonic coupling

between neighbouring striatal neurons in the dopamine-depleted state (Onn and Grace,

1999). In addition, it has been shown that unilateral dopamine depletion in MPTP-treated

monkeys substantially reduces the acquired sensory responsiveness of striatal neurons,

and this can be reversed by administration of apomorphine (Aosaki et al., 1994).

In freely moving unilateral 6-OHDA-lesioned rats, the firing rates of medium spiny-like

neurons is significantly higher on the lesioned side compared to the intact hemisphere or

to normal controls (Kish et al., 1999). On the other hand, recent recordings in

anesthetized 6-OHDA rats have shown decreased activity in direct pathway MSNs

following dopamine depletion, while the activity of indirect pathway MSNs is increased

Page 67: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

55

(Mallet et al., 2006). It was demonstrated that striatal MSNs become more depolarized

after dopamine depletion and fire more frequently in conjunction with cortical inputs

(Tseng et al., 2001), suggesting that dopamine loss facilitate the passage of synchronized

activity through the basal ganglia by disrupting striatal “filtering” of cortical oscillatory

inputs. In addition, the normal arrangement of striatal neurons into clusters that respond

to sensory inputs from single body parts appears to be fragmented in the dopamine-

depleted state (Cho et al., 2002). Clusters were smaller following 6-OHDA lesion, and

more neurons related to single body parts were observed in isolation, outside of clusters.

In addition, more body parts were represented per unit volume, suggesting that following

dopamine depletion striatal neurons may respond to multiple body parts (Cho et al.,

2002). Since the physiological properties of striatal neurons are mediated by

corticostriatal inputs (Liles and Updyke, 1985), it is possible that dopamine depletion

resulted in a reorganization of corticostriatal connections.

The changes in dopaminergic transmission are known to affect corticostriatal function.

Long-term potentiation (LTP) and long-term depression (LTD), the two main forms of

synaptic plasticity, are both represented at corticostriatal synapses and strongly depend on

the activation of dopamine receptors. It has been shown that in 6-OHDA-lesioned rats,

corticostriatal LTP is impaired (Centonze et al., 1999), but can be restored with chronic

treatment with levodopa (Picconi et al., 2003). Interestingly, with chronic levodopa

treatment, a consistent number of animals develop involuntary movements, resembling

human dyskinesias. In these animals, as opposed to non-dyskinetic rats, corticostriatal

LTD was lost (Picconi et al., 2003). Thus, the inability of corticostriatal synapses to

depotentiate might represent the cellular basis of levodopa-induced dyskinesias, with

MSNs being unable to filter the redundant motor information that would normally be

eliminated (Picconi et al., 2003;Pisani et al., 2005).

1.4.1.b - Subthalamic nucleus

Several lines of evidence indicate that abnormal neuronal activity of the subthalamic

nucleus plays a pivotal role in the pathophysiology of parkinsonian motor symptoms.

Recordings from the STN in monkeys have shown that the mean firing rate of STN

Page 68: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

56

neurons increases dramatically (from 15-25 to 25-40 spikes/sec) after MPTP intoxication

(Miller and DeLong, 1987;Bergman et al., 1994), and lesion or deep brain stimulation of

this structure reverses parkinsonian symptoms in MPTP-treated monkeys (Bergman et al.,

1990;Benazzouz et al., 1993;Guridi et al., 1994;Guridi et al., 1996). Since STN activity is

increased after local injection of bicuculline (a GABAA receptor antagonist) and

decreased after local injection of muscimol (a GABAA receptor agonist) (Bergman et al.,

1994), the pathological increase in STN firing is thought to result from a decrease in the

GABAergic tone arising from the GPe. However, in 6-OHDA treated rats, the increase in

the firing rates of STN neurons is not solely dependent on GPe (Hassani et al., 1996)

suggesting that excitatory inputs to the STN, such as the corticosubthalamic pathway,

play a role in the pathology of the parkinsonian STN. In addition to increased firing rates

of STN neurons, there is an increased proportion of neurons responsive to somatosensory

stimulation and an increase in movement related excitability (duration and magnitude)

after MPTP-treatment (Bergman et al., 1994).

In awake and unmedicated PD patients, the mean firing rates of STN neurons range

between 30 and 65 Hz (Hutchison et al., 1998;Levy et al., 2000;Magarinos-Ascone et al.,

2000;Magnin et al., 2000;Steigerwald et al., 2008). In has been shown that STN neurons

discharge at significantly higher rates (~40 Hz) in PD patients compared to patients with

essential tremor (~20 Hz), a movement disorder without any known basal ganglia

pathology (Steigerwald et al., 2008). Some studies in PD patients have reported the

existence of a somatotopic organization of sensorimotor fields within the STN

(Rodriguez-Oroz et al., 2001;Theodosopoulos et al., 2003). Another study, however, has

failed to reveal a consistent somatotopic organization, and found that STN neurons

respond to more than one joint (Abosch et al., 2002). Therefore, it is possible that in PD

the somatotopic organization in STN is somewhat distorted and there is a loss of spatial

selectivity (Abosch et al., 2002). In PD patients, about 20% of STN neurons respond to

eye movements. The majority of these neurons are located in the ventral STN, while most

neurons showing somatic responses are located in the dorsal half of the nucleus (Fawcett

et al., 2005).

Page 69: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

57

In both animal models of PD and PD patients, the overactivity in STN is associated with

a significant increase in the number of bursting cells and the occurrences of synchronized

oscillatory activity among STN neurons (Bergman et al., 1994;Vila et al.,

2000;Steigerwald et al., 2008). In part, this may occur because of an enhanced interaction

between STN and cortical areas (Magill et al., 2000;Magill et al., 2001), but other

pathological phenomena might also be involved. For example, dopamine acts locally to

reduce inhibitory synaptic inputs to the STN (Shen and Johnson, 2000;Tofighy et al.,

2003;Zhu et al., 2002b;Zhu et al., 2002a), and its absence may enhance the impact of

synchronous GABAergic inputs on STN activity, resulting in rebound bursting (Plenz

and Kital, 1999;Bevan et al., 2002b;Shen and Johnson, 2005). In parkinsonian monkeys

and PD patients, abnormal oscillatory firing typically emerges in two frequency bands.

The first is tremor-locked, 3-7 Hz, bursting activity (Bergman et al., 1994;Hutchison et

al., 1998;Rodriguez-Oroz et al., 2001;Zhuang and Li, 2003;Amtage et al., 2008). In PD

patients, the percentage of STN neurons oscillating at tremor frequencies decreases

significantly following the administration of apomorphine, parallel with a reduction in

limb tremor (Levy et al., 2001a). In addition to tremor-related oscillations, higher-

frequency oscillations in the beta range can be observed in

the STN of PD patients and

parkinsonian animals (Bergman et al., 1994;Levy et al., 2002b;Moran et al., 2008;Mallet

et al., 2008). It is interesting to note that, for reasons yet unknown, the frequency of beta

synchronization tends to be higher in parkinsonian rodents and PD patients than in MPTP

monkeys (11-30 Hz vs. 8-20 Hz respectively) (Hammond et al., 2007).

Dopamine replacement therapy in PD patients significantly reduces beta oscillatory firing

in the STN (Levy et al., 2002a), suggesting a role of oscillations in the pathophysiology

of PD symptoms. In support for this hypothesis, it has been shown in patients that

stimulation of the STN at 10 Hz induces significant worsening of motor symptoms,

especially akinesia (Timmermann et al., 2004). Similarly, stimulation of the subthalamic

region at 20 Hz impairs performance in a simple tapping task (Chen et al., 2007). In these

studies, however, the worsening effect is often very small compared with the beneficial

effect of high frequency stimulation. Interestingly, STN stimulation at currents higher

than the amount needed to produce clinical effects induces dyskinesias (Benabid et al.,

Page 70: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

58

2000).

Intentional lesioning of the STN in PD patients was usually avoided because of the

possible risk of causing severe dyskinesias or ballism as was observed in normal

individuals with lesions of the STN. However, local inactivation of neuronal activity by

microinjections of the local anesthetic lidocaine and the GABAA agonist muscimol into

the STN was shown to ameliorate parkinsonian symptoms in PD patients (Levy et al.,

2001b). In addition, several reports have demonstrated good therapeutic effects of

subthalamotomy in patients with PD without adverse motor effects (Gill and Heywood,

1997;Barlas et al., 2001;Alvarez et al., 2001). This anti-parkinsonian effect of STN

lesions is consistent with the notion of STN hyperactivity in parkinsonism (Miller and

DeLong, 1987;Bergman et al., 1994). In parkinsonian monkeys, 2-deoxyglucose

metabolic mapping studies indicate that STN hyperactivity is responsible for the

excessive GPi activity (Mitchell et al., 1989) (see below). Indeed, STN lesions in

parkinsonian animals lead to the reduction of GPi activity (Wichmann et al.,

1994b;Guridi et al., 1996;Blandini et al., 1997).

1.4.1.c - Globus pallidus

Most electrophysiological studies carried out in MPTP monkeys have shown that

dopamine depletion induces an increase in the spontaneous activity of GPi neurons from

60-80 Hz to 80-100 Hz (Miller and DeLong, 1987;Filion and Tremblay, 1991;Boraud et

al., 1996;Boraud et al., 1998;Bezard et al., 1999;Boraud et al., 2000b;Leblois et al.,

2006b). A few more recent studies, however, have failed to find a significant increase in

GPi firing rates (Wichmann and DeLong, 1999;Raz et al., 2000;Leblois et al., 2007).

Electrophysiological recordings in the GPe have also furnished controversial results.

While some studies have shown a decrease in firing rate of GPe neurons from 50-70 Hz

to 30-50 Hz (Miller and DeLong, 1987;Filion and Tremblay, 1991;Boraud et al.,

1998;Raz et al., 2000), others have reported no change (Bezard et al., 1999;Boraud et al.,

2001). Recordings from PD patients have reported firing rates of 60-70 Hz in the GPe

and 80-90 Hz in the GPi (Hutchison et al., 1994;Hutchison et al., 1997;Lozano et al.,

1996). It is important to note that some studies have shown that the average spontaneous

Page 71: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

59

activity of GPi neurons does not differ significantly between patients suffering from PD,

dystonia and Huntington‟s disease (Hutchison et al., 2003;Tang et al., 2005), suggesting

that GPi firing rates do not change after dopamine depletion.

On the other hand, the firing patterns of both GPi and GPe are affected by nigral lesion.

In GPe, there is an increase in the number of bursting neurons and the occurrence of

pauses. Neurons that normally display low-frequency discharge bursting behavior have

more frequent and longer bursts after MPTP intoxication (DeLong et al., 1985;Miller and

DeLong, 1987;Filion and Tremblay, 1991;Boraud et al., 1998;Boraud et al., 2001). A

similar increase in bursting activity was also observed in the GPi (Filion and Tremblay,

1991;Bergman et al., 1994;Wichmann et al., 1999;Leblois et al., 2006b) and was

correlated with the degree of parkinsonism in MPTP monkeys (Filion and Tremblay,

1991). Likewise, 6-OHDA rats show a significant increase in the number of bursting cells

both in the globus pallidus (Ni et al., 2000a) and entopeduncular nucleus (Ruskin et al.,

2002). A high degree of bursting activity was also found in GPe and GPi neurons in PD

patients (Hutchison et al., 1994).

In addition to the increased bursting firing, the proportion of oscillatory cells and of

synchronized pairs has been shown to increase dramatically after nigrostriatal lesion in

rats (Ruskin et al., 2002) and monkeys (Raz et al., 2000;Leblois et al., 2007). Tremor-

locked oscillations of GPi neurons have been reported both in MPTP monkeys (Nini et al.,

1995;Bergman et al., 1998b;Raz et al., 2000) and PD patients (Hutchison et al.,

1997;Hurtado et al., 1999;Lemstra et al., 1999). Similar to STN, synchronized neuronal

oscillations at beta frequencies are also observed in the GPi and GPe (Raz et al.,

2000;Levy et al., 2002b;Leblois et al., 2007).

Studies that have compared GP sensorimotor fields in normal and MPTP monkeys show

that parkinsonism is associated with a dramatic loss of spatial focal selectivity. In

monkeys treated with MPTP, more GP neurons are activated by limb movements, and

there is an increase in the duration and magnitude of the response (Filion et al.,

1988;Boraud et al., 2000a;Leblois et al., 2007). In addition, the movement responses can

Page 72: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

60

be linked to as many as 9 joints and 4 limbs, compared to only one joint in normal

monkeys (DeLong, 1971;Filion et al., 1988;Boraud et al., 2000b;Leblois et al., 2006b).

This seems also to be the case in patients with PD (Sterio et al., 1994;Taha et al., 1996).

It is interesting to note that lesions of the STN in parkinsonian monkeys decrease the

magnitude of GPi responses to somatatosensory stimulations (Wichmann et al., 1994b).

Furthermore, it has been shown that contrary to the normal situation where GPi responses

to voluntary movement occur after the onset of muscle activity (Mink and Thach, 1991a),

after MPTP treatment the average response starts 80 ms before the onset of muscle

activity (Boraud et al., 2000b). This reversal of temporal relationship between GPi

neurons and motor output may be corollary to the enhancement of inhibition exerted by

the GPi neurons on their target neurons, with a consequent delay in movement initiation

(Leblois et al., 2006b).

Systemic injection of levodopa or dopaminergic agonists brings down the excessive

firing rate of GPi neurons by at least 50%, and increases GPe firing by up to 200% in

both MPTP monkeys (Filion et al., 1991;Boraud et al., 1998;Boraud et al., 2001;Papa et

al., 1999) and PD patients (Hutchinson et al., 1997;Stefani et al., 1997;Merello et al.,

1999b;Merello et al., 1999a;Lozano et al., 2000;Levy et al., 2001a). Dopamine agonists

also decrease the number of bursting neurons in the pallidum of MPTP monkeys (Filion

et al., 1991;Boraud et al., 1998;Boraud et al., 2001) and 6-OHDA rats (Ruskin et al.,

2002). Another study, in which bicuculline was injected into the GPe of intact monkeys

to elicit dyskinesias, showed that dyskinesias are associated with either an increase or a

decrease in the firing rate of GPi neurons, but there was a consistent modification of the

firing pattern to long pauses (Matsumura et al., 1995). This suggests that firing patterns in

the globus pallidus, rather than firing rates, are linked to dyskinesias.

1.4.1.d - Substantia nigra

Substantia nigra pars reticulate

Following MPTP treatment in monkeys, an increase in firing rates and bursting patterns

of SNr neurons is reported (Wichmann et al., 1999). Injections of muscimol into the SNr

of MPTP monkeys can reverse parkinsonian symptoms (Wichmann et al., 2001),

Page 73: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

61

suggesting that the SNr is involved in the pathophysiology of PD. In PD patients,

although SNr neurons fire at rates comparable to GPi firing rates (i.e. 60-90 Hz), the

pattern of SNr discharges is very regular and no oscillatory activity is observed

(Hutchison et al., 1998). These differences between humans and monkeys may be due to

differences in parkinsonism or in the location of the sampled neurons. Rats with

unilateral 6-OHDA lesion show an increase in firing rates and movement-related

responses in the ipsilateral SNr (Chang et al., 2006) together with the development of

bursting activity (Sanderson et al., 1986;Burbaud et al., 1995). STN lesions have been

shown to decrease both firing rates and the occurrence of bursting patterns in the SNr

(Burbaud et al., 1995;Tseng et al., 2000). Similar to observations in the GPi, intrastriatal

injections of dopaminergic agonists decrease the firing rate and regularize the discharge

pattern of SNr neurons (Akkal et al., 1996). Interestingly, it has been shown that after

repeated administration of apomorphine in 6-OHDA-lesioned rats, the SNr has an

increased percentage of units with bursting activity concurrently with the enhancement of

behavioral responses that resemble drug-induced dyskinesias in PD patients (Lee et al.,

2001). Very recently, it has been demonstrated that activity-dependent plasticity in the

SNr of PD patients is remarkably enhanced with low doses of levodopa and importantly,

the level of synaptic plasticity in the SNr negatively correlates with the patients‟ “off”

UPDRS motor scores (Prescott et al., 2009).

Substantia nigra pars compacta

Since the dopaminergic neurons of the SNc degenerate in PD, their activity in the course

of the establishment of the disease has not yet been investigated in behaving animals.

One study has examined the spontaneous and stimulation-induced overflow of

endogenous dopamine from striatal slices prepared from adult rats (Snyder et al., 1990).

This study has demonstrated that after lesioning with 6-OHDA, dopamine release from

residual terminals is increased relative to control. This increase in dopamine release

might serve a compensatory function, maintaining the control over striatal function

despite the extensive loss of dopaminergic neurons (Snyder et al., 1990). It is important

to note, however, that although the lesion increased dopamine release at moderate

stimulation frequencies (2-8 Hz), it had reduced the effective range of frequencies over

Page 74: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

62

which the dopaminergic terminals could operate.

1.4.1.e - Changes in basal ganglia-related structures

Thalamus

According to predictions of the rate model, the characteristic loss of dopaminergic

innervation of the striatum is associated with an excessive tonic inhibition of thalamic

motor nuclei that are part of the basal ganglia-thalamocortical circuitry, namely the

ventral anterior and ventral lateral nuclei (Miller and DeLong, 1987;DeLong, 1990;Filion

et al., 1991;Sterio et al., 1994;Boraud et al., 1998). This is supported by metabolic studies

in MPTP-treated monkeys that show a marked increase in the metabolism of the ventral

anterior and ventral lateral thalamic nuclei in parkinsonism (Mitchell et al., 1989;Rolland

et al., 2007), perhaps reflecting increased basal ganglia input to this area. In PD patients,

there is a correlation between increased GPi firing rates and ipsilateral ventral thalamic

glucose metabolism (Eidelberg et al., 1997). Indeed, electrophysiological studies in PD

patients have demonstrated a decrease of discharge rates in pallidal-receiving areas of the

thalamus (Voa and Vop) (Molnar et al., 2005). Nevertheless, studies in MPTP monkeys

have failed to show simlar results (Pessiglione et al., 2005). There is, however,

consistency among studies showing changes in thalamic firing patterns that resemble

those found in basal ganglia nuclei. In MPTP-treated monkeys, the incidences of burst

discharges and oscillatory firing are increased in the ventral lateral thalamus (Guehl et al.,

2003;Pessiglione et al., 2005). A high level of bursting and oscillatory firing in this area

was also observed in PD patients (Lenz et al., 1994;Lenz et al., 1988;Zirh et al.,

1998;Raeva et al., 1999;Magnin et al., 2000;Molnar et al., 2005). Tremor-related

oscillatory neuronal activity has been identified in thalamic nuclei of PD patients during

parkinsonian limb tremor (Lenz et al., 1988;Zirh et al., 1998). These tremor-related

neurons can also respond to sensory stimulation and voluntary movement (Lenz et al.,

1994). In addition, there is an increase in the correlation of the spiking activities of

neighboring neurons in the ventral anterior and ventral lateral thalamus of MPTP

monkeys, and it has been demonstrated that parkinsonism is associated with reduced

specificity of the sensory responses of the neurons in these areas (Pessiglione et al., 2005).

Page 75: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

63

Activity changes were also observed in the CM/Pf nuclei of the thalamus, which

participate in circuits by which basal ganglia output is fed back into the putamen.

According to the rate model, dopaminergic loss may result in a reduction in CM/Pf

activity, which may, in turn, result in reduced driving of MSNs that belong to the „direct‟

pathway (Sidibe and Smith, 1996). In line with this prediction, there is implicit evidence

that GABAergic basal ganglia output to these nuclei is increased in parkinsonism. In 6-

OHDA-lesioned rats, GABAA receptor subunit expression in Pf is decreased (Chadha et

al., 2000), and the mean firing rates of Pf neurons is transiently reduced (Ni et al., 2000b).

It should be noted here that the CM/Pf thalamus may also be affected by the

neurodegenerative process in PD. Studies have shown more than 50% of CM/Pf neurons

degenerate in parkinsonian patients and rodents (Xuereb et al., 1991;Henderson et al.,

2000;Aymerich et al., 2006).

Interestingly, many of the electrophysiological changes in the thalamus do not seem to be

specific to the nuclei that receive inputs from the basal ganglia. It has been shown that

burst firing, oscillations and abnormal sensory processing also occur in thalamic areas

that receive cerebellar input (i.e. the ventral intermediate nucleus) (Guehl et al.,

2003;Pessiglione et al., 2005). Indeed, DBS in the ventral intermediate thalamus can

effectively ameliorate parkinsonian tremor (Benabid et al., 1991;Blond et al.,

1992;Benabid et al., 1996;Koller et al., 1997;Kumar et al., 1999a), although gait, akinesia,

or the activities of daily living cannot be improved (Defebvre et al., 1996;Koller et al.,

1997).

Cortex

Early metabolic studies suggest that cerebral activation at rest is globally reduced in

MPTP-treated monkeys (Schwartzman and Alexander, 1985). Later electrophysiological

studies in MPTP-monkeys have reported a reduction in the activation of the motor cortex

and the supplementary motor area (Watts and Mandir, 1992), and a loss of the reciprocal

pattern of response of movement-related cells (Doudet et al., 1990). Similar to basal

ganglia structures, the synchrony between cortical neurons is enhanced following MPTP

treatment (Goldberg et al., 2002), and the specificity of receptive fields is markedly

Page 76: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

64

decreased both in the primary motor cortex (Goldberg et al., 2002) and the supplementary

motor area (Escola et al., 2002). It has been recently demonstrated that the functional

connectivity between the primary motor cortex and the globus pallidus is greatly

enhanced after MPTP treatment (Rivlin-Etzion et al., 2008). Microelectrode recordings in

the cortex of PD patients are currently not available. Nevertheless, EEG studies in

patients have demonstrated that parkinsonism may be associated with abnormal beta-

band synchronization of cortical networks (see below, section 1.4.2) and a failure to

modulate frontal and central beta-band activity with movement (Brown and Marsden,

1998;Brown, 2003). This is supported by studies showing significant increase in beta

oscillatory activity over the sensorimotor cortex in 6-OHDA rats (Sharott et al.,

2005b;Mallet et al., 2008).

Imaging studies have shown that cortical activation in PD patients, specifically in the

supplementary motor area and the anterior cingulate cortex, is reduced during motor tasks,

often with the activation of brain areas that are not activated in non-parkinsonian subjects

(e.g. premotor cortex, cerebellum, posterior parietal cortex and occipital lobe) (Jenkins et

al., 1992;Playford et al., 1992;Jahanshahi et al., 1995;Samuel et al., 1997;Brooks,

1997;Thobois et al., 2000;Haslinger et al., 2001;Turner et al., 2003). The decreased

activation of the supplementary motor area and the anterior cingulate cortex in PD

patients can be reversed with apomorphine (Jenkins et al., 1992) and DBS of the STN or

the GPi (Limousin et al., 1997;Davis et al., 1997). These studies suggest that dopamine

depletion may not only disturb frontal areas of the cerebral cortex that receive input from

the basal ganglia (via the thalamus), but may also result in compensatory shifts in

activation towards other areas of cortex.

Pedunculopontine nucleus

There is evidence that abnormalities in brainstem regions, specifically the PPN, may be

involved in the development of parkinsonism. The PPN is tightly connected to the basal

ganglia (Mena-Segovia et al., 2004). Studies in 6-OHDA-lesioned rats have suggested

that PPN activity is increased in the dopamine depleted state (Breit et al., 2001), and that

a lesion of the PPN can reduce some of the discharge abnormalities in STN and SNr

Page 77: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

65

(Breit et al., 2006). On the other hand, in MPTP-monkeys, injection of the GABA

receptor antagonist bicuculline into the PPN has been shown to reduce akinesia (Nandi et

al., 2002a). Neuronal degeneration in the PPN itself may cause dopamine-resistant

parkinsonian deficits, including gait disorders, postural instability and sleep disturbances

(Zweig et al., 1989;Pahapill and Lozano, 2000). A detailed review on the PPN and its

role in PD is given in the Appendix.

1.4.2 - Local field potential recordings

The physiological basis of local field potentials

Local field potentials (LFPs) are believed to be generated by the sum of numerous

overlapping postsynaptic potentials distributed among many cells (ECCLES, 1951). The

contribution of action potentials to the LFP is thought to be negligible for a number of

reasons. First, postsynaptic potentials can propagate much further in the extracellular

space compared to spikes due to the low-pass filter properties of the passive neuron.

Secondly, EPSPs and IPSPs have a longer duration than the very brief action potentials,

so they have a much higher chance of occuring in a temporally overlapping manner.

Finally, EPSPs and IPSPs are displayed by many more neurons than spikes (because not

all neurons reach the spike threshold at any given time). Thus, the extracellular field

potentials are generated because the slow postsynaptic potentials allow the temporal

summation of currents of relatively synchronously activated neurons (Buzsaki, 2006). It

should be noted, however, than non-synaptic events such as subthreshold oscillations,

calcium spikes and other intrinsic events which produce relatively long-lasting

transmembrane events, can also contribute to the local field potentials. The volume of

neurons that contribute to the LFP signal depends on the size and placement of the

extracellular electrode. With a very fine electrode, the LFP is likely to reflect a weighted

average of input signals on the dendrites and cell bodies of tens to thousands of neurons

in the vicinity of the electrode (Buzsaki, 2006).

LFPs recoded from the basal ganglia

The emergence of abnormal oscillatory activity in PD occurs not only at the single-cell

Page 78: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

66

level but is also reflected as changes in the LFPs. In patients, LFPs are more readily

recorded from the BG than the activity of single neurons because the former recordings

can be made not only intra-operatively, but also in the few days between surgical

implantation of the DBS electrode and the connection of the electrode to a subcutaneous

stimulator a few days later. As mentioned earlier (section 1.3.3), LFP recordings from the

GPi and STN of PD patients withdrawn from their anti-parkinsonian medication have

consistently revealed prominent oscillations in the 11-30 Hz range (Brown et al.,

2001;Marsden et al., 2001;Cassidy et al., 2002;Levy et al., 2002a;Williams et al.,

2002;Williams et al., 2002;Williams et al., 2003;Priori et al., 2002;Silberstein et al.,

2003;Brown, 2003;Kuhn et al., 2004;Brown and Williams, 2005;Foffani et al.,

2005c;Androulidakis et al., 2008a). These oscillatory LFPs are thought to reflect

synchronized oscillatory synaptic or neuronal activity that is generated by large

populations of local neural elements (Galvan and Wichmann, 2008). This notion has been

strongly supported by intra-operative recordings that demonstrate locking of neuronal

discharges in the STN to beta oscillatory LFPs (Kuhn et al., 2005). It is important to note

that although LFP oscillations are locally generated, they might index synchronization of

the whole basal ganglia-cortical loop (Brown and Williams, 2005;Hammond et al., 2007)

as beta oscillatory activities in the STN, GPi and cortex are largely coherent (Brown et al.,

2001;Williams et al., 2002;Marsden et al., 2001;Cassidy et al., 2002;Sharott et al.,

2005b;Fogelson et al., 2005b).

Given the relation between the basal ganglia and cortex, it is not surprising that

oscillatory activity in the beta band has also been observed in areas of cortex that are

related to the basal ganglia in PD patients and animal models of PD (Marsden et al.,

2001;Williams et al., 2002;Goldberg et al., 2002;Fogelson et al., 2005b;Sharott et al.,

2005b;Silberstein et al., 2005b;Goldberg et al., 2004;Mallet et al., 2008;Costa et al.,

2006). From studies of LFPs and spiking activity in the cortex and basal ganglia of

parkinsonian animals (Goldberg et al., 2004;Sharott et al., 2005b;Mallet et al.,

2008;Costa et al., 2006), it was concluded that in the parkinsonian conditions, cortex-

basal ganglia networks are tightly locked to the beta rhythms that are echoed by the

cortex and basal ganglia LFP (Hammond et al., 2007). Global engagement of the BG-

Page 79: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

67

thalamocortical circuits in synchronized oscillatory activity may severely disrupt

processing at all levels of the circuitry. This may reflect activities such as movement-

related modulation of beta-band synchronization (see section 1.4.3), or functions such as

motor planning or sequence learning (Wichmann and DeLong, 2006a). In support of this

concept, 10- or 20-Hz stimulation of the subthalamic nucleus of PD patients resulted in a

moderate increase in akinesia (Timmermann et al., 2004;Chen et al., 2007).

It has been speculated that DBS may act to desynchronize pathological oscillations in the

basal ganglia (Brown et al., 2004). Indeed, it has been recently shown that short-train

high-frequency stimulation in the STN of PD patients suppresses subthalamic LFP beta

activity and decreases motor cortical-STN coherence in the beta band for up to 25

seconds after the end of stimulation (Wingeier et al., 2006;Kuhn et al., 2008). However,

other studies showed that following or during STN-DBS, the power of beta oscillations

remained unchanged (Foffani et al., 2006;Rossi et al., 2008), suggesting that beta

oscillations in the STN do not necessarily reflect the clinical state. Investigation of the

effect of TMS on beta oscillatory activity in patients revealed that stimulation of the

motor cortex or SMA reduces beta activity in STN-LFPs (Gaynor et al., 2008).

1.4.3 - Oscillatory activity in relation to voluntary movements and dopaminergic

medications

In PD patients and animal models of PD, the oscillatory activity in the beta-band is

suppressed by treatment with dopaminergic medications in tandem with clinical

improvement (Brown et al., 2001;Levy et al., 2002a;Cassidy et al., 2002;Silberstein et al.,

2003;Priori et al., 2004;Sharott et al., 2005b;Kuhn et al., 2006b). Furthermore, the degree

of reduction in subthalamic beta oscillatory activity following dopaminergic medications

correlates with clinical improvement in bradykinesia and rigidity (Kuhn et al., 2006b).

This, together with the little or no beta activity in the STN and GP of healthy animals, is

the basis for presuming that beta activity might be of particular importance in the

generation of motor deficits in PD. In contrast to beta suppression, gamma activity at 35-

90 Hz has been reported in recordings from many PD patients following treatment with

Page 80: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

68

dopaminergic medication (Brown et al., 2001;Cassidy et al., 2002;Williams et al.,

2002;Androulidakis et al., 2007). There is also evidence for LFP activity at ~300 Hz in

the STN of PD patients especially following treatment with levodopa (Foffani et al.,

2003;Foffani et al., 2005b).

Other evidence in PD patients supports the view that beta synchronization is related to

motor impairment. One of the earliest behavioral observations was that beta band

oscillations picked up in STN and GPi are reduced prior to and during self- and

externally-paced voluntary movements (Levy et al., 2002a;Cassidy et al., 2002;Priori et

al., 2002;Williams et al., 2003;Kuhn et al., 2004;Doyle et al., 2005a;Devos et al.,

2006;Kempf et al., 2007). The suppression of beta oscillation prior to movement is clear

evidence that this effect is not simply the consequence of peripheral re-afference (Brown

and Williams, 2005). This beta suppression does not seem related to the parkinsonian

state per se, as it is also manifest in the striatum of healthy monkey (Courtemanche et al.,

2003) and in healthy human putamen (Sochurkova and Rektor, 2003) and cortex

(Pfurtscheller and Neuper, 1992;Toro et al., 1994;Leocani et al., 1997;Ohara et al.,

2000;Alegre et al., 2002;Doyle et al., 2005b). In contrast to beta oscillatory activity,

oscillations in the gamma frequencies significantly increase during voluntary movement

(Cassidy et al., 2002;Kempf et al., 2007), and tend to be negatively correlated with the

UPDRS motor score in PD patients (Kuhn et al., 2006b). Moreover, the increase in

gamma activity before movement onset is enhanced by levodopa treatment

(Androulidakis et al., 2007). Thus, synchronization at the gamma frequencies seems to

have a prokinetic effect.

LFP recordings from the subthalamic nucleus of PD patients have demonstrated that

during predictive trials, where a cue to move („go‟ signal) was preceded by a warning cue,

there was an obvious drop in beta power after the warning signal and an even more

marked drop following the „go‟ signal (Cassidy et al., 2002;Williams et al., 2003;Kuhn et

al., 2004). This drop in beta oscillations was followed by a late post-movement increase

in beta power. Interestingly, the mean timing of the drop in beta activity following a „go‟

signal precedes and positively correlates with the mean reaction time across PD patients

Page 81: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

69

(Kuhn et al., 2004) and even across single trials within individual subjects (Williams et

al., 2005). In other words, earlier reductions in beta power have been observed in trials

with faster motor responses. Furthermore, the power suppression prior to movement was

longer in duration when greater amounts of motor processing were required after a „go‟

cue (Williams et al., 2005). As opposed to predictive trials, in non-predictive trials, where

a warning cue was followed by a cue not to move („nogo‟ signal), the beta suppression

following the „nogo‟ signal was prematurely terminated and reversed into an early beta

power increase (Kuhn et al., 2004), suggesting that increased beta activity is needed to

inhibit movement. These data suggest that the degree of beta synchronization in the STN

may be an important determinant of whether motor programming and movement

initiation is favored or suppressed.

A drop in beta power preceding self-paced movements is also evident in STN LFP

activity. Such studies demonstrate that the relative degree of pre-movement beta

suppression is greater following levodopa treatment in PD patients (Doyle et al.,

2005a;Devos et al., 2006). In addition, beta desynchronization and synchronization

patterns during self-paced movements were very similar in the STN and primary

sensorimotor cortex with post-movement beta synchronization being more prominent

over contralateral structures (Devos et al., 2006). It has been shown that post-movement

beta synchronization is less intense over the sensorimotor cortex in PD patients

(Pfurtscheller et al., 1998), but can be increased with high frequency stimulation of the

STN or treatment with dopaminergic medications (Devos et al., 2003;Devos et al., 2006).

Interestingly, desynchronization of beta oscillations in the STN has also been observed

during motor imagery in PD patients (Kuhn et al., 2006a). The degree of event-related

desynchronization during motor imagery was correlated with the degree of event-related

desynchronization in trials of motor execution across patients and, like motor execution,

was accompanied by a decrease in cortico-STN coherence. As opposed to

desynchronization, event-related beta synchronization was significantly smaller in trials

of motor imagery than during motor execution. It was therefore concluded that the beta

desynchronization in STN during movement is related to the feedforward organization of

Page 82: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

70

movement and is relatively independent of peripheral feedback. In contrast, sensorimotor

feedback is an important factor in the post-movement synchronization occurring in the

STN (Kuhn et al., 2006a).

It is important to keep in mind that it is still unknown whether levodopa or DBS

treatment improve parkinsonism because they reduce beta activity, or whether the

reduction of beta oscillations is simply correlated with, but not causal to, the beneficial

motor effect (Galvan and Wichmann, 2008). Recent studies in animal models of PD shed

some light on this issue. As mentioned above, monkeys that underwent a gradual MPTP-

treatment, show synchronous oscillatory firing in the GPi only after the first appearance

of bradykinesia and akinesia (Leblois et al., 2007), indicating that oscillatory firing does

not contribute to early parkinsonism. Similarly, beta oscillations in STN and cerebral

cortex were not exaggerated until several days after 6-OHDA injections in rats (Mallet et

al., 2008), suggesting that abnormally amplified beta oscillations in cortico-basal ganglia

circuits do not result simply from an acute absence of dopamine receptor stimulation, but

are delayed consequence of chronic dopamine depletion. This is further supported by the

finding that excessive cortical beta synchronization in rats requires a prolonged

interruption in dopamine transmission and is detected only several days after 6-OHDA

injection and the appearance of akinesia (Degos et al., 2008).

1.4.4 - Oscillatory activity in relation to dyskinesias

Several reports have linked synchronization in the BG at low frequencies (4-10 Hz) with

dystonic unwanted movements. Such activity was initially observed in the GPi of both

generalized dystonia patients and medicated PD patients (Silberstein et al., 2003). LFP

power in the 11-30 Hz band was decreased and that in the 4-10 Hz band increased in

medicated compared with unmedicated PD patients. Moreover, dystonia patients had less

11-30 Hz power and greater 4-10 Hz power compared with unmedicated or medicated

PD patients (Silberstein et al., 2003). It therefore seems likely that synchronization in the

4-10 Hz range is associated with the symptoms of dystonia rather than with the disease

itself. Indeed, it has been recently shown that 4-10 Hz oscillations in the STN of patients

Page 83: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

71

with PD are increased during levodopa-induced dyskinesias (Alonso-Frech et al., 2006).

In contrast, beta oscillations in the GPi of PD patients have been shown to inversely

correlate with levodopa-induced dyskinesias in (Silberstein et al., 2005a). Furthermore, it

has been shown that low frequency oscillations in the GPi of patients with primary

dystonia are particularly associated with the mobile, rather than the postural, components

of this condition (Liu et al., 2006). In agreement with this notion, mobile dystonia could

be provoked by intra-operative DBS of the STN at 5 Hz (Liu et al., 2002).

It should be noted that a recent study on the relationship between neuronal firing and LFP

activity recorded from the GPi in patients with dystonia has demonstrated that neurons

can be synchronized to the LFP activity over 3-12 Hz (Chen et al., 2006b).

1.4.5 - Oscillatory activity in relation to tremor

As mentioned earlier, some experimental studies suggest that parkinsonian tremor is most

likely to be caused by abnormal synchronous oscillating neuronal activity within the

central nervous system, and peripheral factors play only a minor role in its generation,

maintenance, and modulation (Bergman and Deuschl, 2002;Deuschl et al., 2000).

However, the source of oscillation in PD tremor is still uncertain. In patients and animal

models, tremor-related neurons were found in the STN, GPi and the thalamus (Lenz et al.,

1988;Hutchison et al., 1997;Rodriguez et al., 1998;Bergman et al., 1994;Bergman and

Deuschl, 2002;Levy et al., 2000) and either nucleus could be the origin of parkinsonian

tremor. They could either be the generators themselves or may be an integral part of an

oscillating network.

Studies of the correlation or coherence between tremor and BG oscillations have not been

conclusive (Lemstra et al., 1999;Hurtado et al., 1999;Raz et al., 2000;Hurtado et al.,

2005). One possible explanation is that PD tremor is generated by multiple segregated

circuits, each involving a different limb (Alberts et al., 1965). Indeed, the frequency of

tremor is often dissimilar between different sides of the body and it is generally observed

that tremulous limbs oscillate independently of each other whereas muscles within a limb

Page 84: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

72

are more likely to be coupled (Hurtado et al., 2000;Raethjen et al., 2000;Ben-Pazi et al.,

2001). These observations support the original hypothesis by Alberts et al. (1965), that

different oscillators underlie PD tremor in the different extremities. A recent study has

further confirmed this hypothesis by showing that tremor-related activity in the globus

pallidus can be coupled with one tremulous limb but not with the other (Hurtado et al.,

2005).

Tremor-related oscillations, although common in single-unit recordings (Hutchison et al.,

1997;Levy et al., 2000;Levy et al., 2002b) are not a consistent or strong feature in LFP

signals, probably due to the variable phase relationships between neurons oscillating at

tremor frequencies (Lenz et al., 1994;Hurtado et al., 1999;Hurtado et al., 2005). As

previously mentioned, oscillatory field potential activity in PD patients „off‟ medications

is particularly prominent in the beta band. Beta oscillatory activity in the basal ganglia

has been previously associated with the pathology that gives rise to tremor in PD (Levy et

al., 2000). However, more recent studies in PD patients failed to find a clear positive

correlation between tremor severity and the degree of beta oscillations (Silberstein et al.,

2003;Amirnovin et al., 2004;Wang et al., 2005;Kuhn et al., 2006b;Ray et al., 2008).

It has been recently demonstrated that the cortex-BG-periphery loops have low-pass filter

properties to microstimulation patterns that contained bursts delivered at different

frequencies between 1 to 15 Hz (Rivlin-Etzion et al., 2008). It has been shown that the

motor cortex does not follow GP stimulation at frequencies above 5 Hz, and that

stimulation at frequencies higher than 5 Hz in the motor cortex did not evoke movement.

Thus, it was suggested that parkinsonian tremor is not directly driven by the BG

oscillations despite their similar frequencies. Rather, these BG oscillations should be

considered as disrupting the normal motor processing (Rivlin-Etzion et al., 2008). In fact,

oscillatory neuronal activity can occur without overt tremor (Levy et al., 2000;Wichmann

et al., 1999;Raz et al., 2001;Heimer et al., 2002;Soares et al., 2004). It is therefore

possible that additional anatomic or physiological factors (poorly defined at this moment)

may be necessary for tremor to occur (Wichmann and DeLong, 2006a).

Page 85: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

73

1.5 - Aim of present studies

The overall aim of this project was to obtain a better understanding of the characteristics

of the oscillatory activity recorded from the basal ganglia of PD patients and to elucidate

the significance of this activity in PD. To date, local field potentials recordings from DBS

electrodes implanted in the basal ganglia of PD patients are widely used as a tool to infer

subthalamic and pallidal function in various different studies and these have given rise to

several important claims. Thus, an important question is whether or not the LFPs

recorded from DBS electrodes reflect the neuronal firing in the recorded nuclei. By

recording neuronal firing and LFP activity with microelectrodes during stereotactic

neurosurgery for movement disorders (Chapter 2), we examined the hypothesis that LFP

oscillations recorded from the STN of PD patients are locally generated (Chapter 3). In

addition, we elucidated to what extent the oscillatory LFPs reflect neuronal firing in a

population of local STN neurons. The relationship between LFP activity and neuronal

firing was also investigated in the GPi of PD patients (Chapter 5) and compared to that in

dystonia patients in order to shed more light on the possible changes in oscillatory

patterns and neuronal firing that are involved in Parkinson‟s disease.

Behavioral studies in PD patients have suggested that increased beta activity is involved

in movement inhibition (see section 1.4.3). This led to the hypothesis that excessive

neuronal synchronization in the basal ganglia might contribute to the motor impairment

in PD and especially to bradykinesia and rigidity. However, at the time, there was no

clear evidence that the degree of synchronization in the beta band is related to the motor

deficits in PD. We therefore considered the incidence of oscillatory neurons in the STN

in relation to the patients‟ motor symptoms and their benefit from dopaminergic

medications (Chapter 3) in order to provide further clues into the role of beta oscillations

in PD. Another common motor symptom of PD is tremor at rest. Unlike bradykinesia and

rigidity, PD tremor does not correlate well with the degree of dopamine deficiency in the

striatum and with the progression of the disease, and previous studies regarding the

relationship between oscillatory activity in the BG and tremor have yield conflicting

results (see section 1.4.5). Since tremor is considered a nonstationary phenomenon,

Page 86: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

74

assessments of the interactions between basal ganglia oscillatory activities and tremor

should be obtained by analyzing these signals over time. Thus, we went on to explore the

temporal dynamics of oscillatory activity in the STN in relation to PD tremor (Chapter 4)

in order to gain a better understanding of the role of STN and oscillatory activity in

mediating parkinsonian tremor.

Page 87: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

75

2 - CHAPTER 2 - GENERAL METHODS

This chapter provides a detailed description of the methods used in the studies presented

in the thesis. Patients were operated by Dr. Andress Lozano and Dr. Mojgan Hodaie.

Clinical assessments of the patients were performed by Dr. Elena Moro and Dr. Anthony

Lang. Setup of microelectrode equipment and intraoperative microelectrode recordings

were carried out by me in conjunction with Dr. William Hutchisona and Dr. Jonathan

Dostrovsky. Some of the MATLAB programs used for data analysis were written by Dr.

Neil Mahant. All data analyses were performed by me.

2.1 - Patients and consent

All patients underwent functional stereotactic neurosurgery for the implantation of DBS

electrodes into the STN, GPi or PPN. The patients selected for STN surgery were

generally PD patients treated for the cardinal signs of PD (i.e. akinesia/bradykinesia,

rigidity and occasionally tremor). These patients were sensitive to levodopa and, in most

cases, suffered from levodopa-induced dyskinesias. On the other hand, patients selected

for DBS in the GPi were generally treated for dystonia, except for rare cases in which PD

patients who greatly suffer from dyskinesias were selected for GPi surgery. Patients

selected for PPN surgery were either PD patients who suffered from severe gait and

postural impairment or patients with progressive supranuclear palsy (PSP) that were

unable to walk.

Each patient gave his/her free and informed consent to participate in the study and all

procedures were reviewed and approved by the University Health Network Research

Ethics Board.

2.2 - Intraoperative neuronal recordings

2.2.1 - Operative procedures

Page 88: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

76

Detailed descriptions of operative procedures are given elsewhere (Hutchison et al.,

1994;Lozano et al., 1996;Hutchison et al., 1998) . Briefly, on the morning of the surgery,

a stereotactic frame (Leksell G, Elekta, Inc, Atlanta, Ga) was affixed to the patients‟

heads and preoperative magnetic resonance (MR) images were obtained (Signa, 1.5 T,

General Electric, Milwaukee, Wis). Then, a target within the nucleus was chosen using

one of the two following methods. In the first method, MRI was used to locate the frame

coordinates for the anterior and posterior commissures (AC, PC). These coordinates were

then entered into a computer program that shrinks or expands the standard stereotactic

human brain atlas map (Schaltenbrand and Wahren, 1977) to fit the patient's AC-PC

length and plots the electrode trajectories. In the second method, a target can be chosen

directly based on anatomical MRI visualization (Machado et al., 2006). In this method,

called direct targeting, triplanar reconstruction of the MR images with the axial series

parallel to the AC-PC plane is carried out on a surgical neuronavigation workstation

(Mach 4.1, StealthStation, Medtronic, SNT). This method was used in most of the

STN/GPi cases and in all the PPN cases.

After the initial target has been determined, the patient is brought to the operating room

where one (in unilateral cases) or two (in bilateral cases) burr holes (25 mm) were made

in the skull under local anesthesia. Physiological mapping of neuronal activity using

microelectrode recordings was then performed through this opening. Microelectrode

recording trajectories started 15 or 10 mm above the intended target and extended up to 5

mm below. A manual microdrive was used to drive the microelectrodes down through the

brain. All patients were awake during the microelectrode recordings. Microstimulations

were performed to elicit sensory or motor effects using an isolated stimulator (Axon

system GS3000) which delivered trains of square wave pulses (pulse width 150 s or 200

s) at 100 to 300 Hz and up to a maximum of 100 A. Recordings and stimulation

mapping allowed the identification of physiological landmarks.

2.2.2 - Physiological targeting

2.2.2.a - Subthalamic nucleus

Page 89: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

77

The use of microelectrode recordings to localize DBS electrode placement in the STN is

described in detail elsewhere (Hutchison et al., 1998). Briefly, parasagittal trajectories

were oriented at 12 mm from the midline. A typical microelectrode trajectory starts 10

mm above target and passes through the thalamic reticular nucleus and/or anterior

thalamus, zona incerta, STN, and the SNr. Exploration of the neuronal activity was

carried out along the entire dorsal/ventral extent of STN. The dorsal border of STN was

noted by increase in background activity and high-frequency neuronal discharge. The

main markers that were identified in order to localize the motor portion of the STN were

neurons with tremor-related activity and neurons that responded to passive or active

movements. As the electrodes were advanced past the ventral border of STN the

background noise decreased until the electrode reached the SNr,

which was characterized

by higher-frequency, lower-amplitude and more regular discharges compared with STN.

A representative example is shown in Figure 2.F1.

In order to determine the approximate locations of the recordings within the STN (as a

function of depth within the nucleus), distances of the recordings to the dorsal border of

the nucleus were calculated.

Page 90: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

78

2.2.2.b - Globus pallidus internus

The recording procedures to localize the posteroventral GPi during stereotactic

neurosurgery have been previously described (Hutchison et al., 1994;Lozano et al., 1996).

Briefly, parasagittal trajectories were oriented at 20 mm from the midline. In addition to

identifying GPi cells, the optic tract (located ventral to the GPi) and the internal capsule

(located posterior to the GPi) were identified in order to prevent misplacement of the

electrodes into these areas, which might result in reduced efficacy or side effects.

Microelectrode trajectories usually start 15 mm above target so that the recordings often

begin within the GPe. This would then be followed by the internal medullary lamina,

which lies between the GPe and GPi. This in an area of white matter and is therefore

identified by a significant decrease in overall activity. Occasionally, peripallidal „border‟

cells, characterized by a highly regular firing pattern and slower rate, are encountered

within the lamina and borders of GPi. Entrance into the GPi is marked by an overall

Page 91: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

79

increase in background noise and high frequency discharges. The optic tract could be

identified by microstimulation to elicit visual percepts and/or by using a flashing strobe

light to elicit a visual evoked response in the axonal activity of the optic tract. Sometimes

it can be identified by axonal electrical activity with narrow spikes or a subtle increase in

high frequency noise. Another important anatomical landmark is the internal capsule that

could be identified by stimulation-evoked tetanic contractions of contralateral muscles. A

representative example of a microelectrode trajectory is shown in Figure 2.F2.

2.2.2.c - Pedunculopontine nucleus

The PPN as a surgical target for deep brain stimulation was first introduced in 2005

(Plaha and Gill, 2005;Mazzone et al., 2005). Thus, physiological landmarks for the

microelectrode mapping of this region have not been yet described. Part of our studies

was to characterize the neurophysiology of the human PPN region and identify

neurophysiological landmarks that may aid in the localization of this target (see

Appendix).

Page 92: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

80

2.2.3 - Microelectrode setup

Microelectrodes were made by inserting a parylene-C coated tungsten microelectrode

with an exposed tip length of 15-25 m (Microprobe Inc.) into a 30-gauge stainless steel

tube (HTX-30 tubing; Small Parts Inc., Miami Lakes, FL) that was then covered by

Kapton tubing (Micro ML Tubing, #23 polyimide tubing). These parts were glued

together with epoxy. Microelectrode tips were plated with gold and platinum to reduce

the impedance to about 0.2-1 M at 1 kHz.

A dual microelectrode set-up was employed to allow simultaneous recordings to be made

from two microelectrodes (Levy et al., 2007). The two microelectrodes were

independently inserted into two adjacent stainless steel guide tubes that were constructed

by soldering two 23-gauge, thin walled stainless steel tubes (HTX-23; Small Parts Inc.).

These tubes were positioned medial-lateral to one another relative to the patient. The dual

inner guide tubes fit easily into the standard stereotactic frame outer guide tube, which, in

our setup, was constructed from 17-gauge stainless steel tubing (HTX-17; Small Parts

Inc.). The microelectrodes were parallel and were separated by a mediolateral distance of

600 µm. Each electrode was driven into the brain independently by a manually operated

hydraulic microdrive. Therefore, it was possible to record from pairs of neurons and/or

LFPs separated by about 600 m mediolaterally and at variable distances axially (but

usually at about the same height). A photograph of the dual microdrive head stage

assembly is shown in Figure 2.F3.

Page 93: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

81

2.2.4 - Single unit and local field potential recordings

During surgery, single unit activity and local field potentials were recorded from the

microelectrodes. The monopolar recordings from both microelectrodes shared a common

ground consisting of the stainless steel guide tube and frame attachments to the head.

The amplifier ground was also connected to the frame. Recordings were amplified 5,000-

10,000 times and filtered at 10 to 5,000 Hz (analog Butterworth filters: high-pass, one

pole; low-pass, two pole; at 5 Hz amplitude decreased by roughly 50%) using two

Guideline System GS3000 (Axon Instruments, Union City, CA) amplifiers. During the

recordings, signals were monitored on a loudspeaker and displayed on a computer screen.

Recorded signals were digitized at 10 kHz and directly stored onto a computer hard drive

with a CED 1401 data acquisition system running Spike2 (Cambridge Electronic Design,

Page 94: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

82

Cambridge, UK). All neuronal data were recorded simultaneously with wrist

flexor/extensor EMG and accelerometers (sampled at 1 kHz) to monitor muscle activity

and limb movement, respectively.

2.3 - Data analysis

2.3.1 - Spectral analysis of neuronal discharges and local field potentials

Only periods without voluntary movements or artifacts were analyzed. Single and

multiunit activity was discriminated using the wavemark template-matching tool in

Spike2 (Cambridge Electronic Design, Cambridge, UK). Only sites where good unit

recordings were obtained were analyzed. Spike times and unfiltered LFP data were

imported into MATLAB (version 6.5 or 7.1, The Math Works, Natick, MA) for further

analysis.

2.3.1.a - Power spectrum

Local field potentials

The main statistical tool for analyzig LFP recordings was the discrete Fourier transform

and its derivations calculated according to Halliday et al., (1995). After signals were

down-sampled to 1 kHz, spectra of LFP power were estimated by dividing the waveform

signal into a number of sections of equal duration of 1.024 seconds (1024 data points,

512 point overlap), each section was windowed (Hanning window) and the magnitudes of

the 1024 discrete Fourier transform of each section were squared and averaged to form

the power spectrum, yielding a frequency resolution of 0.97 Hz. The power was

transformed to a logarithmic scale and shown in decibels (dB). Because the estimated

power spectrum has a distribution analogous to a

2 distribution, the 95% confidence

intervals were given on the basis of the 2 distribution (Jarvis and Mitra, 2001), whereas

degrees of freedom values are based on the number of windowed sections.

In the case of time series, the Fourier transform can be thought of as performing a Fourier

decomposition of the sampled waveform into constituent frequency components, which

Page 95: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

83

should highlight any distinct periodic components in the data (Halliday et al., 1995).

Since the signal was sampled at 1000 Hz, the Nyquist frequency was 500 Hz.

Neuronal discharges

Similarly to the LFP, spectral analysis of spike trains was performed using the Fourier

transform according to Halliday et al. (1995). Before the power spectral density was

estimated, spike times were converted to a logical function (zeros and ones) where each

neuronal spike is stored as logical pulse together with a vector of their relative times. The

spectral analysis, in case of point process, can be thought of as performing a correlation

between the sinusoids and co-sinusoids of the complex Fourier exponential with the times

of occurrence of the events. The presence/absence of particular periodicities in the spike

timings will lead to increase/decrease of correlation at the frequency of periodicity from

that expected by chance alone, which would highlight periodic components in the

discharge (Halliday et al., 1995).

Significant oscillations in the power spectra were detected using shuffling of the

interspike intervals (ISIs) of the spike trains (Rivlin-Etzion et al., 2006b). Interspike

interval shuffling generates a new spike train by using the time differences between

adjacent spikes (first-order ISIs). Thus the spectrum of the new spike

train is determined

solely by the first-order ISIs of the original spike train, whereas higher-order effects (i.e.,

the time difference between spikes that are separated by one spike or more) are

abolished

by the shuffling process. Comparing the original spectrum to the new one enables one to

detect patterns such as oscillations that are generated by higher-order ISIs. To obtain an

accurate and less-noisy estimate, we repeated the shuffling process 100

times and

averaged the results. Subtraction of the new estimated spectrum from the original

spectrum resulted in a corrected spectrum, in which peaks were considered significant

when they exceeded the upper 95% confidence limit. The confidence limit was estimated

from the mean spectrum of the shuffled ISIs based on the 2 distribution as described

above. Because the the variance for the shuffled ISIs spectrum is the same for all

frequencies, the confidence interval depends solely on the degrees of freedom and the

term is a constant that does not depend on the frequency.

Page 96: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

84

2.3.1.b - Coherence and cross-correlation

Coherence analyses (Rosenberg et al., 1989;Halliday et al., 1995) were used to evaluate

the relationship between simultaneously recorded data from separate electrodes and

between LFP and spike data recorded from the same microelectrode. The

coherence

function provides a frequency domain bounded measure of association, taking on values

between 0 and 1, with 0 in the case of independence and 1 in the case of a perfect linear

relationship. Coherence can be estimated by direct substitution of the appropriate spectra

as: |fxy|2/fxxfyy with 95% confidence

level of 1–(0.05)

1/(L–1); where fxx and fyy are the

autospectra of the component processes, fxy is the cross-spectral density, and L is the

number of windowed sections.

It should be noted that since the LFP signals (sampled at 1000 Hz) did not undergo

additional low-pass filtering prior to the analyses, they were still likely to contain a bit of

information about the neuronal firing recorded from the same microelectrode. This was

often noticeable as significant coherence at frequencies above 40 Hz between LFP and

spike data that were recorded from the same microelectrode. Thus, significant coherence

at higher frequencies between neuronal firing and LFP from the same electrode might be

the result of spike contamination. However, in these studies the coherence between

neuronal firing and LFP data were only assessed for frequencies up to 30 Hz.

Cross-correlation provides another means of calculating the relationship between

simultaneously recorded data as it provides time-domain measure of dependency between

random processes. Cross-correlation is defined by the inverse Fourier transform of the

cross-spectrum fxy, but can also be estimated directly in the time domain. The estimation

throughout the frequency domain, however, facilitates the construction of confidence

limits (Halliday et al., 1995). To estimate the 95% confidence limits, the variance of the

cross-correlation (varcross-corr) was approximated based on the autospectra of the

component processes (fxx and fyy). Under the hypothesis that in the case of two

independent processes the value of the cross-correlation function is zero, the upper

and

lower 95% confidence limits for the estimated cross-correlation are given by 0 ±

Page 97: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

85

1.96(varcross-corr)1/2

. In these studies correlation histograms were plotted

for delays of 500

ms (1-ms bin width) and their main purpose was to provide as an additional way to

evaluate the correlation between signals at specific frequency bands, as LFP data were

band-pass filtered prior to the calculation. However, only the coherence estimates were

used to determine whether a relationship was significant or not.

It should be noted that phase relationships were not considered in this thesis. Although

information about the phase shifts between simultaneously recorded data could have

added to our results, we feel that such analyses will not contribute much to our overall

conclusions since we only used two microelectrodes that were very close to each other

(~1 mm) and the phase shifts are expected to be centered around zero (see Levy et al.

2000).

2.3.1.c - Time frequency analysis

We used the continuous wavelet transform to analyze changes in LFP power in the time-

frequency domain (Wang et al., 2005). The wavelet function is convolved with the

observed data at each time point and across a range of scales, allowing the identification

of specific frequency components (ω) over time (t). In the present study we used the

Morlet wavelet function, which is a sine wave (eiωt

= cos(ωt) + i sin(ωt)) that is

windowed by a Gaussian bell curve (e–t2

) to give a brief oscillation localized in both time

and frequency. At higher frequencies, time resolution increases where frequency

resolution decreases. A constant, the Morlet parameter, determines the number of sine

waves within the wave packet, and higher values give greater frequency resolution at the

expense of time resolution. The absolute value of the transform is then squared to give a

continuous time-frequency representation of the power content of the signal (Torrence

and Compo, 1998).

Wavelet-based coherence was calculated to characterize the temporal relationship

between neuronal firing and the simultaneously recorded LFP or tremor. The classical

coherence definition was modified utilizing the smoothed Morlet wavelets to yield time-

frequency coherence, defined as the ratio of the wavelet cross spectrum to the product of

Page 98: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

86

the wavelet autospectra of the two signals (Gurley et al., 2003). Areas of significant

coherence were detected with 95% confidence level of 1-(0.05)1/(df–1)

, where df is the

number of degrees of freedom required in order to detect coherence values above 0.1.

The percent time during the overall record when the two signals were significantly

coherent with each other for a given frequency band was calculated. A Morlet parameter

of 16 was used to investigate the coherence between a tremor-frequency neuron and the

simultaneously recorded LFP or between two LFPs, whereas a parameter of 8 was used

to investigate coherence between a tremor-frequency firing neuron and limb tremor

(which required a better time resolution). These parameters were chosen because they

were determined to give the most optimal time and frequency resolution over the

frequency range inspected.

2.3.2 - Statistical comparisons

All standard statistical tests were preformed using SigmaStat software (version 3.1, Systat

Software, Richmond, CA). Comparisons of firing rates, power of oscillations, or

recording locations (along the microelectrode track) between two groups were subjected

to the student t-test if distributions of data were normal; otherwise, Mann-Whitney rank

sum tests were performed. Paired t-tests / Wilcoxon signed-rank tests were also used to

compare neuronal and LFP measurements in differing circumstances (e.g. without/during

tremor). Chi-square comparisons were performed to compare proportions of observations

of different categories, but if the sample size was small Fisher‟s exact test was used

instead. In all statistical tests, a P value less than 0.05 was considered to be a significant

difference. Values are expressed in either mean ± standard error (SE) or mean ± standard

deviations (SD) as appropriate.

Page 99: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

87

3 - CHAPTER 3 - BETA OSCILLATORY ACTIVITY IN THE SUBTHALAMIC

NUCLEUS AND ITS RELATION TO DOPAMINERGIC RESPONSE

The work described in this chapter was published by Weinberger et al. (2006) J

Neurophysiol. 96(6):3248-56, and is used with permission from the American

Physiological Society.

Abstract

Recent studies suggest that beta (15-30 Hz) oscillatory activity in the subthalamic nucleus

(STN) is dramatically increased in Parkinson's disease (PD) and may interfere with

movement execution. Dopaminergic medications decrease beta activity and deep brain

stimulation (DBS) in the STN may alleviate PD symptoms by disrupting this oscillatory

activity. Depth recordings from PD patients have demonstrated beta oscillatory neuronal

and local field potential (LFP) activity in STN, although its prevalence and relationship to

neuronal activity are unclear. In this study, we recorded both LFP and neuronal spike

activity from the STN in 14 PD patients during functional neurosurgery. Of 200 single-

and multiunit recordings 56 showed significant oscillatory activity at about 26 Hz and

89% of these were coherent with the simultaneously recorded LFP. The incidence of

neuronal beta oscillatory activity was significantly higher in the dorsal STN (P = 0.01)

and corresponds to the significantly increased LFP beta power recorded in the same

region. Of particular interest was a significant positive correlation between the incidence

of oscillatory neurons and the patient's benefit from dopaminergic medications, but not

with baseline motor deficits off medication. These findings suggest that the degree of

neuronal beta oscillatory activity is related to the magnitude of the response of the basal

ganglia to dopaminergic agents rather than directly to the motor symptoms of PD. The

study also suggests that LFP beta oscillatory activity is generated largely within the

dorsal portion of the STN and can produce synchronous oscillatory activity of the local

neuronal population.

Page 100: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

88

3.1 - Introduction

Damage to the basal ganglia (BG) is well known to result in a variety of movement

disorders highlighting the important although still poorly understood function of these

nuclei in the motor system. The loss of nigral dopaminergic input to the striatum

in

Parkinson's disease (PD) leads to a poverty and slowness of movements, rigidity, and

frequently also postural instability and tremor, but the mechanisms underlying these

abnormalities remain unclear. Studies in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine

(MPTP)–treated monkeys and in PD patients suggest that one of the consequences of loss

of dopaminergic inputs to the BG is increased synchronized oscillatory activity in the

subthalamic nucleus (STN) and globus pallidus interna (GPi) (Bergman et al., 1994;Levy

et al., 2000;Levy et al., 2002a;Marceglia et al., 2006;Nini et al., 1995).

Depth recordings of local field potentials (LFPs) in STN of PD patients have

demonstrated prominent oscillatory activity in the beta frequency band (15–30 Hz)

(Brown et al., 2001;Cassidy et al., 2002;Kuhn et al., 2004;Levy et al., 2002a), which is

consonant with cortical EEG activity (Marsden et al., 2001;Williams et al., 2002). The

STN beta LFPs are decreased by dopaminergic medication and active movements (Alegre

et al., 2005;Alonso-Frech et al., 2006;Cassidy et al., 2002;Doyle et al., 2005a;Foffani et

al., 2005c;Kuhn et al., 2004;Levy et al., 2002a;Priori et al., 2004;Priori et al.,

2002;Williams et al., 2003). It was previously suggested that the therapeutic effects of

deep brain stimulation (DBS) arise from disrupting this pathologically increased

synchronized beta activity (Brown et al., 2004;Filali et al., 2004;Jahanshahi et al.,

2000;Lozano et al., 2002;Wingeier et al., 2006).

The origin of the beta LFP activity is still poorly understood, although studies by Levy et

al. (2000) demonstrated the existence of neurons in STN that fired rhythmically at

frequencies within the beta band and many pairs of neurons were shown to fire

synchronously in the beta range. Moreover, the STN oscillatory neurons fire

in synchrony

with the simultaneously recorded beta LFP (Kuhn et al., 2005;Levy et al., 2002a) and the

beta LFP activity is maximal in the dorsal part of STN (Kuhn et al., 2005).

Page 101: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

89

To further elucidate the relationship of the STN oscillatory neuronal activity and LFPs

and their possible role in mediating the motor symptoms of PD, we simultaneously

recorded neuronal and LFP activity from pairs of microelectrodes in the STN of

PD

patients at rest and off medication. Moreover, investigation of the changes in the neuronal

and LFP beta activity with depth along the electrode trajectories, above and within the

STN, allowed analysis of the distribution of these activities within

the nucleus. Also of

interest was whether there is a relationship between the incidence of beta oscillatory cells

and the patient's motor impairment and/or the effectiveness of dopaminergic medication

in alleviating the patient's parkinsonian symptoms. Thus we examined the relationship

between the percentage of STN oscillatory cells and the degree of motor disability on and

off dopaminergic medication. Clarification of the relationship between the recorded

LFP,

local neuronal discharge, and clinical symptoms will provide further insight into the

possible roles of these beta rhythms and their relationship to the therapeutic effectiveness

of dopaminergic medication.

3.2 - Methods

Patients

We studied 14 patients with advanced PD who were undergoing stereotactic surgery for

implantation of bilateral STN DBS electrodes. The group consisted of five women and

nine men who, at the time of operation, had a mean age of 57 yr (range 46–68) and

a

mean duration of PD of 14.1 ± 5.2 (mean ± SD) years. All patients were assessed

preoperatively using the Unified Parkinson's Disease Rating Scale (UPDRS) (Fahn et al.,

1987) before and after an acute levodopa challenge (Moro et al., 2002). During surgery

the patients were awake and off dopaminergic medications for at least 12 hours from the

last oral dose of antiparkinsonian medications. Demographic details of the patients

are

given in Table 3.T1.

Page 102: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

90

Recordings

The electrophysiological mapping procedure used to obtain physiological data for

localizing the target for the DBS electrode placement in the STN is described in detail in

the General methods (section 2.2.2.a). Extracellular recordings of neuronal firing and

LFPs were obtained simultaneously from two independently driven microelectrodes

(about 25 µm tip length, axes 600 µm apart, about 0.2-MΩ impedance at 1 kHz) as

described in detail in the General Methods (sections 2.2.3/4). All recordings were

performed in the resting state and under local anesthesia. Patients were not asked to

perform any task and epochs with movements were excluded.

Pairs of simultaneous recordings of neurons and/or LFPs were generally at roughly

the

same depth. The mean linear distance between the pairs of recording sites was 0.83 ± 0.3

mm (mean ± SD, n = 195). An example of simultaneously recorded LFP and neuronal

discharge is shown in Figure 3.F1A.

Data analysis

Single- and multiunit activity was discriminated using the wavemark template matching

Page 103: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

91

tool in Spike2 (Cambridge Electronic Design, Cambridge, UK). Only sites where good

unit recordings were obtained were analyzed. At 24% of the sites the action potentials

were well discriminated and deemed to be from a single unit. At the remainder of the sites

we could not rule out the contribution from an additional one or more

units and these

were then termed multiunit recordings. However, the difference between the mean (±SD)

firing rate of the single units and the multiunit sites was relatively small

(39.1 ± 18.6 vs.

49.1 ± 24.5 Hz), suggesting that even in the multiunit cases, most of the discriminated

spikes were probably from a single cell. Throughout this chapter the

use of the term

"cells" refers to both single units and multiunit recording sites. Spike times and unfiltered

LFP data were imported into MATLAB (version 6.5, The MathWorks, Natick, MA) for

further analysis.

Recordings from 21 sides in 14 patients were analyzed. In seven patients both right and

left sides were analyzed; only left STN recordings were analyzed in four patients and

right STN in three patients. Only data of ≥17-s duration during periods

without

movements (based on EMG recordings and observations) or artifacts were analyzed

(range: 17–214 s, mean ± SD: 36.5 ± 22.9 s). Recording depths were realigned to

the top

of STN in each track, where 0 is the dorsal border of STN and negative values are ventral

to this border.

Spectral analysis of spike trains and LFPs were performed using the discrete Fourier

transform and its derivations calculated according to Halliday et al. (1995) and Rivlin-

Etzion et al. (2006). These calculations included power spectra of recorded activity (spike

trains and LFPs) as well as coherence and cross-correlations between simultaneously

recorded data, as previously described in the General Methods (section 2.3.1.a/b). To

estimate the relative LFP beta power according to the distance above or below the top of

STN, we identified the frequency of the beta peak of each STN trajectory, and the mean

power across a 10-Hz window centered on the peak frequency was calculated

at each

recording site. LFP power at each recording site was then expressed as the percentage of

the maximum power observed in the trajectory. Percentages of maximum power were

averaged across subjects to give mean percentage of LFP beta power (Kuhn et al., 2005).

Page 104: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

92

Relative beta power according to depth for neuronal firing was

estimated from the

original power spectra in the same manner.

Changes in LFP power were evaluated by change-point analysis and control limits

(Change-Point Analyzer 2.0 shareware program; Taylor Enterprises, Libertyville, IL).

Change-point analysis iteratively uses a combination of cumulative sum (cusum) and

bootstrapping to detect changes and is more sensitive to change than control limits, based

on plots of serial deviations from the mean (Taylor, 2000). Cusums were determined by

plotting the sequentially summed deviation from the averaged power. Significant

changes

were determined by control limits that give the maximum range over which values are

expected to vary (with 95% probability). Application of this technique to measure power

changes in the basal ganglia was used in previous studies (Cassidy et al., 2002;Kuhn et al.,

2004;Williams et al., 2003).

To examine the relationship between incidence of oscillatory cells and clinical motor

symptoms and effectiveness of medications we used part III (motor) of

the UPDRS "on"

and "off" total scores as well as subscores as evaluated 8.1 ± 7.8 days (mean ± SD) before

surgery. The effectiveness of the antiparkinsonian medications was calculated

as the

difference between the "on" and "off" scores, divided by the "off" score to give the

percentage of benefit. Nonlinear regression statistics in addition to linear regression were

used to better describe the correlations using SigmaStat (version 3.1, Systat Software,

Richmond, CA).

3.3 - Results

3.3.1 - Coherence between neuronal and beta LFP oscillatory activity in the STN

Of 200 single- and multiunit recordings in 14 patients, 56 (28%) displayed significant

beta (15–30 Hz) oscillatory activity at 25.8 ± 3.7 Hz (mean ± SD). There was no

significant difference in the fraction of oscillatory cells between the

single- and the

multiunit recordings (31.3% vs. 27%, respectively, χ

2 test). However, oscillatory cells had

Page 105: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

93

higher firing rates than the non-oscillatory cells (medians: 52.8 vs. 40.9 Hz respectively,

P = 0.005). In 50 (89.3%) of the oscillatory cells, the oscillatory activity was coherent

with the LFP recorded from one or both electrodes. Neuronal oscillatory activity at other

frequencies was only rarely encountered. We observed significant coherence

in the beta

range between the LFPs at all the STN sites where LFP activity was recorded from both

electrodes, even in cases where no peak in the beta band was observed in the LFP power

spectrum. In contrast, the firing of only 17 of the 67 (25.4%) pairs of cells recorded from

the two electrodes was significantly coherent in the beta range. Coherence was present in

nine of ten cases where beta oscillations were detected in both cells

and in six of 15 cases

where beta oscillations were detected in just one cell; the firing activity in two pairs of

cells was coherent even though no beta oscillations were detected

in either cell. Figure

3.F1B shows an example of coherence plots and correlograms from simultaneous

recordings of neuronal and LFP activity where there was significant neuronal and LFP

beta activity.

3.3.2 - Incidence of neuronal beta oscillatory activity is higher in the dorsal STN

The majority of the beta oscillatory cells was localized in the dorsal STN. The mean

depths from the top of STN of the oscillatory (n = 56) and non-oscillatory (n = 144) cells

(±SD) were –1.5 ± 1.1 and –2.1 ± 1.3 mm, respectively

(P = 0.001, t-test). Figure 3.F2A

shows the distributions of the locations of the oscillatory and non-oscillatory cells at

successive 0.3-mm intervals. The two distributions are illustrated by box

plots in Figure

3.F2B (median locations: –1.3 and –2.0 mm for oscillatory and non-oscillatory cells,

respectively). Seventy-five percent of the oscillatory cells were found in

the dorsal STN,

whereas 25% were in the ventral STN (P = 0.01, χ

2 test). On the other hand, there was no

significant difference in the incidence of non-oscillatory cells in the dorsal and ventral

STN (54% vs. 46%, respectively). There was a significant relationship between the

dorsal/ventral location and the presence/absence of neuronal beta activity (P = 0.003, χ

2

test).

3.3.3 - LFP beta oscillations are greatest in the dorsal STN

Page 106: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

94

The power of the beta activity recorded from the pair of microelectrodes in the 14 patients

was calculated for 351 sites located from 5 mm dorsal to 5 mm ventral to the dorsal

border of STN. Figure 3.F3A plots the changes in the percentage maximum LFP power

from each individual patient, and Figure 3.F3B shows the average percentage

maximum

LFP power for all 14 subjects as a function of depth in 0.5-mm intervals. Note that LFP

power exceeds the upper 99% confidence limit (mean + 2.58 SD) for the intervals

between –0.5 and –2.5 mm in STN. A similar distribution

of the beta LFP power was also

observed when recording depths were aligned to the bottom instead of the top of STN.

Averaging the percentage of beta LFP power every 2.5 mm reveals significantly

greater

power in the dorsal part (P < 0.001, Mann–Whitney rank-sum test) (Figure 3.F3C). The

power in the ventral STN was significantly higher than the power above STN (P = 0.002).

No significant change was observed between the 2 segments above STN. These results

were confirmed by control charts and change-point analysis. Analyses

were performed on

the median LFP beta power in each site (Figure 3.F3D). Two significant changes were

detected: the first change was an increase in power in the 0.5-mm interval between –0.5

and –1.0 mm and the second change was a decrease in power at the –2.5- to –3.0-mm

interval. A similar analysis of the neuronal oscillatory power according to depth revealed

a gradual reduction in the mean beta power from the dorsal to the ventral border of STN

(data not shown) resembling that found for LFP power.

The distributions of oscillatory cells and mean LFP beta power by depth within the STN

reveal the same pattern of greater beta activity dorsally. Figure 3.F4A shows the

variations in mean percentage LFP beta power and percentage of oscillatory cells with

depth within the STN in 0.5-mm intervals and the corresponding regression

lines (linear

regression, R2 = 0.63 and R

2 = 0.71, respectively).

The relationship between depth within

the STN and amount of beta activity is not significantly different for cells and LFPs

(t-test

for difference in slope, P = 0.23). Figure 3.F4B shows the strong linear correlation

between the percentage of oscillatory cells and the mean LFP beta power at different

depths within the STN (linear regression, R

2 = 0.63).

Page 107: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

95

3.3.4 - Neuronal oscillations correlate with levodopa response

The percentage of beta oscillatory cells in STN (including both sides) was found to vary

substantially between patients (see Figure 3.F5), but interestingly was negatively

correlated with the "on" drug UPDRS score (Figure 3.F5B) and positively correlated with

the magnitude of the preoperative levodopa response (Figure 3.F5C) (nonlinear

regression, R2 = 0.49, P < 0.05 and R

2 = 0.62,

P < 0.005, respectively; linear regression,

R2 = 0.37, P

= 0.02 and R

2 = 0.46, P < 0.01, respectively). Similar significant

correlations

occurred between the percentage of oscillatory cells and medication response on tremor

and on the other non-tremor subscores of the UPDRS (data not shown). However, no

significant correlation was observed between the percentage of oscillatory

cells and the

"off" drug UPDRS score (Figure 3.F5a). Moreover, no significant correlations were

observed with the different "off" drug UPDRS subscores such as tremor, rigidity, and

postural instability. Figure 3.F6 demonstrates the lack of association between

the

percentage of oscillatory cells and "off" total tremor scores.

We did not use the LFP beta power for these analyses because we did not have a valid

way to normalize the LFP power in a way that enables to compare between patients. In

each patient, the LFP was recorded in different sites within the STN and its power over

the beta frequencies differ according to the location of the microelectrode within the

nucleus (see previous section) and also according to the microelectrode impedance.

Page 108: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

96

Page 109: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

97

Page 110: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

98

Page 111: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

99

Page 112: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

100

Page 113: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

101

3.4 - Discussion

This study provides new data documenting the relationship of neuronal beta oscillatory

activity to local field potential activity at the recording site and at a distance of about 1

mm. In addition, we describe the distribution of neurons with oscillatory activity within

STN. Most interestingly, our data show a significant positive correlation of neuronal

oscillatory activity with the patient's benefit from dopaminergic medication.

The study

also confirmed previous reports that oscillatory LFP activity in the beta range is

consistently observed in all PD patients off dopaminergic medications (Alonso-Frech et

al., 2006;Brown et al., 2001;Cassidy et al., 2002;Kuhn et al., 2004;Levy et al.,

2002a;Marceglia et al., 2006;Priori et al., 2004). Unexpectedly, however, the incidence of

cells with oscillatory firing in the beta range varied considerably between patients.

Our results demonstrating coherence between neuronal discharge and LFPs in the beta

range confirm and extend the findings of two previous reports (Kuhn et al., 2005;Levy et

al., 2002a). The study by Levy et al. reported data from only a single case showing

statistically significant coherence between the LFP recorded from a macroelectrode

and

the discharge of a cell recorded about 1 mm away. The study by Kuhn et al. used spike-

triggered averaging (equivalent to cross-correlation) of the LFP recorded from a different

Page 114: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

102

contact at the microelectrode tip <30 µm away in six patients, to show that the activity of

some of the neurons was time locked to the beta oscillations in the

LFP. Our study is the

first to examine the coherence between unit activity and LFPs recorded from both the

recording microelectrode and another microelectrode located about 1 mm away, showing

that neuronal activity was frequently coherent with beta oscillatory activity recorded ~1

mm away. Furthermore, the firing of 90% of the cells with beta oscillatory activity was

coherent with the LFP in the beta range. Interestingly, the LFPs at all pairs

of recording

sites within STN showed significant coherence in the beta range even in cases where no

peak in the beta band was observed in the LFP power spectrum. These observations

suggest that the generators of the beta LFP oscillations are distributed

and synchronized

over a large region (at least several millimeters) of the STN.

The present study confirmed the observation in the Levy et al. (2000) study that pairs of

cells in STN can fire rhythmically and synchronously at beta frequencies. In total we

found that 25% of the pairs fired coherently at beta frequencies, which is comparable to

the 30% in the Levy et al. study. However, the findings of the Levy et al. study suggested

that beta oscillatory activity was present primarily in patients with tremor because all but

one of the coherent pairs were from the patients with tremor where 47% of the pairs were

found to fire coherently. In contrast, the current study, which

included a larger and

different population of patients, failed to find such a relationship. The tremor patients in

the Levy study were identified on the basis of their having tremor during

the operation,

whereas in our study patients were defined as having tremor on the basis of the

preoperatively evaluated UPDRS tremor subscores. However, this is unlikely to have had

a major impact on the conclusions of the two studies and we ascribe the difference

to the

lower number of patients in the Levy study where only three patients were in the non-

tremor group.

Unexpectedly, the percentage of cells with beta oscillatory activity varied greatly and

ranged from a low of no cells to a high of 90% of STN cells per patient, even though in

all cases the local field potentials revealed the presence of beta oscillatory

activity. It is

possible that the variability in the number of oscillatory cells is a confound resulting from

Page 115: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

103

limited sampling. However, this would imply that only a relatively small part

of the STN

contains cells firing rhythmically at beta frequencies, which would seem unlikely and

inconsistent with the data and conclusions discussed above. A more likely possibility is

that STN neuron membrane potential oscillations that generate the

beta LFPs are largely

confined to the dendrites and do not necessarily strongly influence the probability of the

neuron firing. Thus neurons with non-significant beta oscillatory firing may nonetheless

contribute to the generation of beta oscillatory LFPs. With increased oscillatory synaptic

inputs and/or somatodendritic coupling there would be increased probability and power of

oscillatory spike activity. Even in the cases where there was only a small

number of

oscillatory neurons, their activity could still produce a significant effect on STN target

nuclei because a weak correlation among very many neurons could add up to become

prominent at the population level (Schneidman et al., 2006).

Because only some of the cells in the dorsal STN fired in synchrony with LFP beta

oscillations and in some patients with clear beta LFPs there were only very few

oscillatory firing cells, it would appear that the oscillatory firing in STN is not the source

of beta LFPs. However, we cannot rule out the possibility that some of the non-

synchronous cells did have a weak synchronicity but that the coherence failed to reach

significance. Nevertheless, we feel that a more likely source is oscillatory afferent inputs,

which are generated by oscillatory firing cells outside of the STN, possibly in cortex. For

example, in vivo recordings from the rat STN demonstrate a close correspondence

between synchronized neuronal and LFP activity after cortical stimulation (Magill et al.,

2004). Phase estimates between the subthalamic area and cortical EEG suggest that

cortical inputs drive STN LFP beta oscillatory activity (Fogelson et al., 2006;Marsden et

al., 2001;Williams et al., 2002) by two possible routes, either indirectly by the

striatum/globus pallidus externus (GPe) or by a direct projection to the subthalamic

nucleus (Parent and Hazrati, 1995a). Moreover, previous studies in a rodent model of

parkinsonism established that the cerebral cortex can induce pathological

patterns of

neuronal activity in the STN, perhaps as the result of greater sensitivity of the STN to

rhythmic cortical inputs (Magill et al., 2001). It was recently suggested by that feedback

GABAergic inhibition from the reciprocally connected GPe can prime STN neurons to

Page 116: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

104

respond more efficiently to excitatory input by increasing the availability of the

postsynaptic voltage-dependent Na+ channels

and is critical for the emergence of coherent

beta oscillations between the cortex and STN in PD (Baufreton et al., 2005a).

Our analysis of the distribution of local field potential beta oscillations revealed that they

are maximal within the dorsal portion of the subthalamic nucleus and confirm similar

recent findings by Kuhn et al. (2005). That study, however, did

not find a significant

difference in power between the dorsolateral and the ventromedial STN, probably as a

result of the limited number of recordings in the ventral part (Kuhn et al., 2005).

This

finding is also supported by our data showing that most of the beta oscillatory cells were

observed within the dorsal part of the nucleus. Moreover, we were able to demonstrate a

significant positive relationship between the percentage of oscillatory cells and the LFP

beta power within the STN. This suggests that there is a close association between the

power of the LFP beta oscillations and the neuronal beta oscillations.

The LFP is likely to

be generated by the postsynaptic potentials produced by rhythmic activity in STN

afferents that terminate preferentially on neurons in the dorsal STN. The dorsal region

is

known to receive inputs from the primary motor cortex in the monkey (Monakow et al.,

1978;Nambu et al., 1996) and to contain neurons responsive to passive and active

movements (Abosch et al., 2002;DeLong et al., 1985;Rodriguez-Oroz et al., 2001). On

the other hand, neurons in the more ventromedial STN are associated with the limbic and

associative cortical regions and have no sensorimotor response (Maurice et al.,

1998;Parent and Hazrati, 1995a). Thus these results, showing the largest amount of

oscillatory cells and the greatest LFP oscillations in the dorsal part,

provide further

evidence for the association of this region with motor functions and the target for DBS

stimulation to alleviate PD motor symptoms.

The oscillatory activity is thought to interfere with initiation and regulation of movements

and, if this were the case, one might expect to find a positive correlation between the

incidence of oscillatory neurons and the baseline UPDRS score off medication.

Interestingly, we did not observe a significant relationship between the "off" UPDRS

scores and degree of oscillatory activity. This observation is somehow inconsistent with

Page 117: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

105

the oscillatory model, which holds that exaggerated beta activity plays an anti-kinetic role.

The clinical features remaining after the levodopa response are considered "non-

dopaminergic" (particularly those involving axial structures) and in part may be generated

from pathology in other brain (particularly brain stem) regions that are affected in

parallel

to the nigrostraiatal pathology (Schapira, 2005). This may account for the correlation

between the beta oscillatory activity and levodopa response (i.e., dopaminergic features)

and not with the “off” motor scores, which combine both dopaminergic

and non-

dopaminergic features. In this study, however, no correlation with symptom severity was

found even when considering each of the motor sub-scores alone (see section 3.3.4).

Dopaminergic medication and active movements are known to decrease STN beta

synchronization (Cassidy et al., 2002;Alonso-Frech et al., 2006;Alegre et al., 2005;Brown

et al., 2001;Foffani et al., 2005c;Kuhn et al., 2004;Levy et al., 2002a;Marsden et al.,

2001;Priori et al., 2002;Priori et al., 2004;Williams et al., 2003) and deep brain

stimulation in the STN may alleviate PD symptoms by disrupting this oscillatory

activity

(Brown et al., 2004;Filali et al., 2004;Jahanshahi et al., 2000;Lozano et al.,

2002;Wingeier et al., 2006). It was previously hypothesized that dopamine action on the

striatum acts as a filter for cortical input to the STN (Doyle et al., 2005a;Magill et al.,

2001;Magill et al., 2004;Sharott et al., 2005b). Thus the increase in number of STN cells

with oscillatory beta activity might reflect the degree of nigrostriatal dopamine deficiency.

The positive correlation between the magnitude of the levodopa response and the

oscillatory activity we observed may be related to the increase in the magnitude of the

levodopa response with progression of PD (Zappia et al., 1997). This is

consistent with

the idea that abolishing the excessive beta activity by levodopa or DBS produces an

improvement in parkinsonian symptoms. Our findings are also consistent with the

recently reported findings that levodopa-induced reduction in subthalamic

LFP power in

the 8- to 35-Hz band as recorded postoperatively from DBS electrodes correlates with the

simultaneously observed clinical improvement in PD patients (Kuhn et al., 2006b).

However, the fact that severity of the patients' motor symptoms did not relate to the

percentage of oscillatory cells suggests that beta oscillatory neuronal activity, alone, may

Page 118: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

106

not reflect the clinical state of the patient and other mechanisms must

also be involved in

the pathophysiology and the contribution of each to the patient's symptoms can vary. A

dissociation between beta oscillatory LFPs and the clinical symptoms was also observed

by Priori et al. (2004), who showed that the anticholinergic drug orphenadrine, in contrast

to levodopa, increases rather than decreases STN beta oscillations while decreasing

tremor and rigidity.

It has been suggested by Bevan et al. (2002b) that both the pattern of inhibitory input

from the GPe and the polarization level of STN neurons are crucial in determining

whether STN neurons fire in a single-spiking or oscillatory pattern. This implies

that

neuromodulators, such as dopamine, that influence the membrane potential of STN

neurons will have a profound effect on the activity in the GPe–STN network (Bevan et al.,

2002b). Based on this, we hypothesize that in the subgroup of patients who

have a greater

amount of neuronal oscillatory activity, the dopamine can act both by suppressing the

overactivity of the indirect pathway at the level of the striatum and by changing

the

polarization level of STN neurons and blocking the oscillatory firing. In these patients,

the STN might be more susceptible to rhythmic cortical inputs and therefore show a

better response to levodopa in relation to those patients with smaller numbers

of

oscillating cells.

It should be noted that some later studies are consistent with our findings. In a recent

study, the LFP power over the 8-35 Hz band measured “off” medication positively

correlated with the improvements in motor symptoms after dopamine intake. In addition,

similarly to our findings, no correlation with baseline motor symptoms was found (Ray et

al., 2008). This was true even when considering only dopamine-responsive motor

features (i.e. bradykinesia+rigidity). Interestingly, both the baseline power and reduction

of 8-35 Hz LFP after dopamine therapy correlated with the degree of improvement in

bradykinesia/rigidity but not tremor (Kuhn et al., 2006b;Ray et al., 2008;Kuhn et al.,

2009).

Page 119: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

107

4 - CHAPTER 4 - THE RELATIONSHIP BETWEEN SUBTHALAMIC

OSCILLATORY ACTIVITY AND TREMOR

The work described in this chapter was published by Weinberger et al. (2009) J

Neurophysiol. 101(2):789-802, and is used with permission from the American

Physiological Society.

Abstract

Rest tremor is one of the main symptoms in Parkinson's disease (PD), although in

contrast to rigidity and akinesia, the severity of the tremor does not correlate well with

the degree of dopamine deficiency or the progression of the disease. Studies suggest that

akinesia in PD patients is related to abnormal increased beta (15-30 Hz) and decreased

gamma (35-90 Hz) synchronous oscillatory activity in the basal ganglia. Here we

investigated the dynamics of oscillatory activity in the subthalamic nucleus (STN) during

tremor. We used two adjacent microelectrodes to simultaneously record neuronal firing

and local field potential (LFP) activity in 9 PD patients who exhibited resting tremor

during functional neurosurgery. We found that neurons exhibiting oscillatory activity at

tremor-frequency are located in the dorsal region of STN, where neurons with beta

oscillatory activity are observed, and that their activity is coherent with LFP oscillations

in the beta frequency range. Interestingly, in 85% of the 58 sites examined, the LFP

exhibited increased oscillatory activity in the low gamma frequency range (35-55 Hz)

during periods with stronger tremor. Furthermore, in 17 of 26 cases where two LFPs were

recorded simultaneously, their coherence in the gamma range increased with increased

tremor. When averaged across subjects, the ratio of the beta to gamma coherence was

significantly lower in periods with stronger tremor compared to periods of no or weak

tremor (P = 0.003). These results suggest that resting tremor in PD is associated with an

altered balance between beta and gamma oscillations in the motor circuits of STN.

Page 120: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

108

4.1 - Introduction

A severe loss of midbrain dopaminergic neurons is the pathological hallmark of

Parkinson‟s disease (PD). Despite significant advances in our understanding of the

anatomy and cellular physiology of the basal ganglia, the casual link between the

pathology and the symptoms is still unclear. Tremor at rest is a well recognized cardinal

symptom of Parkinson‟s disease affecting roughly 70% of PD patients. It is classically

defined as a 4-6 Hz tremor, with or without postural/kinetic tremor. Unlike rigidity and

akinesia, tremor does not necessarily get worse with disease progression and the severity

of tremor does not correlate with dopamine deficiency in the striatum (Deuschl et al.,

2000;Deuschl et al., 2001). Post-mortem and imaging studies suggest that the

pathophysiology of rest tremor may be distinct from that of rigidity and akinesia

(Jellinger, 1999;Pavese et al., 2006). However, the mechanisms underlying parkinsonian

tremor remain to be elucidated.

Most recently, hypotheses regarding the pathophysiology of PD have emphasized

dynamic changes in the basal ganglia network. Changes in firing patterns and in

particular oscillatory activity, in addition to mean firing rates, are considered critical in

PD pathophysiology (Bevan et al., 2002b;Brown, 2003;Hammond et al., 2007;Brown,

2006). Studies in monkeys with 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-

induced parkinsonism suggest that one of the consequences of loss of dopaminergic

inputs to the basal ganglia is increased oscillatory firing and synchronization in basal

ganglia nuclei, i.e. the subthalamic nucleus (STN) and globus pallidus (Bergman et al.,

1994;Heimer et al., 2006;Nini et al., 1995;Raz et al., 2000). Oscillatory neuronal firing at

the frequency of tremor and the beta frequency band (15-30 Hz) in these structures has

also been observed in PD patients undergoing functional neurosurgery and presumed to

be related to the pathophysiological changes (Hurtado et al., 1999;Hutchison et al.,

1997;Hutchison et al., 1998;Levy et al., 2000;Levy et al., 2002b;Magnin et al.,

2000;Weinberger et al., 2006).

Beta oscillatory activity in the basal ganglia has been previously associated with the

Page 121: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

109

pathology that gives rise to tremor in PD (Levy et al., 2000). However, more recent

studies in PD patients failed to find a clear positive correlation between tremor severity

and the degree of beta oscillations (Amirnovin et al., 2004;Kuhn et al., 2006b;Ray et al.,

2008;Silberstein et al., 2003;Wang et al., 2005;Weinberger et al., 2006). Instead, there is

growing evidence suggesting that beta oscillations play an anti-kinetic role and contribute

to bradykinesia and rigidity in PD (Brown, 2003;Chen et al., 2007;Kuhn et al.,

2006b;Kuhn et al., 2008;Ray et al., 2008). On the other hand, oscillatory firing in the

tremor frequency range is frequently coherent with the tremor of a particular limb,

although uncorrelated oscillations are commonplace as well (Hurtado et al.,

1999;Hurtado et al., 2005;Lemstra et al., 1999;Raz et al., 2000). It was originally

suggested by Alberts et al. (1969) that PD tremor is generated by segregated parallel

networks, each involving a different limb. This idea was supported by studies showing

that different limbs oscillate independently of each other (Ben-Pazi et al., 2001;Hurtado

et al., 2000;Raethjen et al., 2000). A recent study has further confirmed this hypothesis

by showing that tremor-related activity in the globus pallidus can be coherent with one

tremulous limb but not with the other (Hurtado et al., 2005).

In spectral analysis terms, tremor is considered a nonstationary phenomenon since it

comes and goes over time (Hurtado et al., 2005). Therefore, assessments of the

interactions between basal ganglia oscillatory activities and tremor should be obtained by

analyzing these signals over time. One means of studying changes in the pattern of local

neuronal synchrony is through a frequency-based analysis of local field potential (LFP)

signals. The LFP is believed to reflect synchronized dendritic currents in a group of

neurons. Tremor-related oscillations, although common in single-unit recordings

(Hutchison et al., 1997;Levy et al., 2000;Levy et al., 2002a;Levy et al., 2002b) are not a

consistent or strong feature in LFP signals, probably due to the variable phase

relationships between neurons oscillating at tremor frequencies (Hurtado et al.,

1999;Hurtado et al., 2005;Levy et al., 2000). Instead, oscillatory field potential activity in

PD patients off medications is particularly prominent in the beta band (Brown et al.,

2001;Kuhn et al., 2005;Levy et al., 2002a;Weinberger et al., 2006) and, to a lesser extent,

in the gamma band (35-90 Hz) (Fogelson et al., 2005a;Pogosyan et al., 2006;Trottenberg

Page 122: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

110

et al., 2006). In contrast to the beta activity, gamma band oscillations in the basal ganglia

have been hypothesized to play a pro-kinetic role and to contribute to movement

generation (Brown, 2003;Brown and Williams, 2005), as these oscillations are increased

during movement and by treatment with dopaminergic medications, in tandem with

clinical improvement (Alonso-Frech et al., 2006;Androulidakis et al., 2007;Cassidy et al.,

2002;Williams et al., 2002). However, their possible involvement in PD tremor has not

been reported.

In order to gain a better understanding of the role of STN and oscillatory activity in

mediating parkinsonian tremor, we studied the relationship between tremor and

oscillatory activities in the STN, at both gamma and beta frequencies, and how these

change with time and in relation to variations in tremor intensity. In addition, we

examined the coherence of STN neurons displaying oscillatory firing at the tremor

frequencies with the simultaneously recorded limb tremor and how the coherence varied

over time. Also, we examined and compared the spatial distribution of these neurons and

beta oscillatory neurons along the dorso-ventral STN axis and their relation to the

simultaneously recorded LFP.

4.2 - Methods

Patients

We studied 9 patients with advanced PD who exhibited intermittent periods of resting

tremor during stereotactic surgery for the implantation of deep brain stimulating

electrodes in the STN. The group consisted of 2 women and 7 men who, at the time

of

operation, had a mean age of 59.5 years (range 55–67) and a mean duration of PD of 10.7

± 3.7 (mean ± SD) years. All patients were assessed preoperatively using the Unified

Parkinson's Disease Rating Scale (UPDRS) (Fahn et al., 1987) before and after an acute

levodopa challenge (Moro et al., 2002). During surgery the patients were awake and off

dopaminergic medications for at least 12 hours from the last oral dose of antiparkinsonian

medications. Demographic details of the patients are given in Table 4.T1.

Page 123: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

111

Recordings

Neuronal activity and LFPs were recorded simultaneously from each of two

independently driven microelectrodes during the electrophysiological

mapping procedure

used to obtain physiological data for localizing the STN for the deep brain stimulation

electrode placement. The localization procedure of the STN using microelectrode

recording is described in detail in the General methods (section 2.2.2.a), as well as the

methods for microelectrode recordings of neuronal firing and LFP (sections 2.2.3/4).

Limb tremor was measured with EMG and/or triaxial accelerometers (with summated x-

y-z signals) that were recorded simultaneously with neuronal activity and were sampled

at 1000 Hz. In all cases the tremor was contralateral to the side of the recordings.

Data analysis

Neuronal discharges were discriminated using spike sorting algorithms in Spike2

(Cambridge Electronic Design, Cambridge, UK). Spike times, unfiltered LFP and

accelerometer/EMG data were imported into MATLAB (version 7.1, The MathWorks,

Natick, MA) for further analysis. Recordings from 13 sides in 9 patients were analyzed.

In four patients, both right and left STN were analyzed; only left STN was analyzed in 3

patients and right STN in two patients. Only periods without voluntary movements or

artifacts were analyzed. Recording depths were realigned to the top of STN in each track,

where 0 is the dorsal border of STN and negative values are ventral to this border.

Page 124: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

112

Tremor amplitude was quantified off-line by calculating the root mean square (RMS)

value of the sampled accelerometer signal. For frequency domain analysis of spike trains

and LFP signals we calculated power spectra and coherence according to Halliday et al.

(1995) and Rivlin-Etzion et al. (2006). These techniques are described in the General

Methods (sections 2.3.1.a/b). To analyze changes in LFP power in the time-frequency

domain, we used the continuous wavelet transform (Morlet wavelet). The wavelet

technique has been previously used to analyze LFP signals recorded from the STN

(Wang et al., 2005) and is described in the General Methods (section 2.3.1.c). In addition,

wavelet-based coherence was calculated to characterize the temporal relationship

between neuronal firing and the simultaneously recorded LFP and tremor (see General

Methods, section 2.3.1.c). A Morlet parameter of 16 was used to investigate the

coherence between a tremor-frequency neuron and the simultaneously recorded LFP or

between two LFPs, whereas a parameter of 8 was used to investigate coherence between

tremor-frequency neuron and limb tremor. These parameters were determined to give

good frequency resolution over the 0.5-35 Hz without sacrificing time resolution. The

percent time during the overall record when the two signals were significantly coherent

with each other for a given frequency band was calculated.

4.3 - Results

4.3.1 - Incidence of neurons with tremor and/or beta oscillatory firing is higher in

the dorsal STN

A total of 102 neurons was recorded from the STN during periods of relatively constant

tremor amplitude. The mean (± SD) duration of recordings was 34.6 ± 20.5 sec (range:

14-129 sec). Of these 102 neurons, 15 neurons exhibited significant oscillatory firing at a

frequency in the range of parkinsonian tremor (tremor-frequency activity) (mean

oscillation frequency ± SD: 4.1 ± 0.6 Hz). In addition, 12 neurons exhibited significant

oscillatory firing in the beta frequency range (mean frequency ± SD: 22.1 ± 8 Hz). Two

of these neurons had both beta and tremor-frequency activity. Examples of power spectra

Page 125: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

113

of neurons firing at tremor and/or beta frequencies are shown in Figures 4.F1A and 4.F2.

There was no significant difference between the mean firing rate of neurons with tremor-

frequency activity and neurons with beta activity (mean ± SD: 55.7 ± 30 and 67.8 ± 27

Hz respectively, t-test), however the oscillatory neurons (n = 25) had a significantly

higher mean firing rate than non-oscillatory neurons (n = 77) (medians: 60.5 and 36 Hz

respectively, P = 0.007, Mann-Whitney rank-sum test). The distribution of the observed

neurons among patients is indicated in Table 4.T1.

The majority (13/15) of the neurons with tremor-frequency activity was localized in the

dorsal 2.5 mm of STN. This distribution was similar to that of the beta oscillatory

neurons. The mean location within the STN (where 0 indicates the dorsal border) of

neurons with tremor-frequency activity was not significantly different than that of

neurons with beta activity (mean ± SD: –1.3 ± 1.0 and –1.7 ± 1.0 mm respectively, t-test).

Figure 4.F1B shows the distributions of the oscillatory and non-oscillatory neurons

within the STN at successive 0.5-mm intervals.

4.3.2 - Neurons with tremor-frequency oscillations are coherent with the LFP at the

beta frequencies

Seven of the 15 neurons (46.6%) with tremor-frequency oscillatory firing showed

significant coherence with the simultaneously recorded LFP in the tremor frequency

range (Figure 4.F2A and B). Interestingly however, 13 tremor-frequency neurons (86.6%)

showed significant coherence with the LFP in the beta frequency range. Eight of these

neurons were coherent with the LFPs recorded from both microelectrodes (including the

two neurons with both tremor and beta activity) (Figure 4.F2). The electrophysiological

characteristics of the neurons with tremor-frequency activity are summarized in Table

4.T2.

Page 126: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

114

Of the 12 neurons with beta oscillations, 11 (92%) showed significant coherence with the

LFP, and 10 of these neurons were coherent with the LFPs from both microelectrodes. It

is important to note that the values of coherence, at the beta frequencies, between the

tremor-frequency neurons and the LFP were significantly lower than the values of

coherence between the beta oscillatory neurons and the LFP (median: 0.17 and 0.35

respectively, P 0.001, Mann-Whitney rank-sum test, excluding neurons with both

tremor and beta activity). In contrast to oscillatory neurons, only 19 out of the 77 non-

oscillatory neurons (25%) showed coherent firing with the LFP at the beta frequencies.

These neurons might exhibit weak beta oscillatory firing that failed to reach significance

but nonetheless had significant coherence when compared to the LFP.

4.3.3 - Coherence between neuronal firing and LFP varies over time

To further characterize the coherence between neurons and the simultaneously recorded

LFP, we used wavelet-based coherence, which allowed us to estimate the percentage of

time during which the two signals were significantly related. Examples of temporal

coherence between tremor / beta oscillatory neurons and the simultaneously recorded

LFP are shown in Figures 4.F3A and 4.F3B respectively. During these periods, tremor

Page 127: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

115

amplitude was relatively constant. Only 8 out of the 15 tremor-frequency neurons

displayed periods of significant coherence with the LFP recorded from one or both

microelectrodes at the tremor frequencies. On average, this coherence lasted for 68 ±

23% (mean ± SD) of the time. In addition, 14 tremor-frequency neurons were coherent

with the LFP at the beta frequencies for an average of 50 ± 26% of the time. It is

important to consider that although more neurons displayed coherence with the LFP at

the beta frequencies, there was no significant difference between the durations of

coherence in the two frequency bands (P = 0.13, t-test). Beta oscillatory neurons were

coherent with the LFP for 77 ± 33% of the time and only one neuron was not coherent

with the LFP. At the beta frequencies, beta oscillatory neurons were coherent with the

LFP for a significantly longer duration relative to tremor-frequency neurons (P = 0.03, t-

test). The data are summarized in Figure 4.F3C. Note that neurons displaying only beta

oscillatory activity were not coherent with the LFP at the tremor frequencies.

4.3.4 - Neurons with tremor-frequency oscillations are not constantly correlated

with tremor

Of the 15 neurons with tremor-frequency oscillations, only 7 were coherent with the

simultaneously recorded tremor (EMG and/or accelerometer) (Table 4.T2, Figure 4.F2).

However, the frequency of oscillations of 12 out of the 15 neurons was similar to the

frequency of tremor, ranging between 4 to 4.9 Hz. In the remaining three neurons, the

frequency of oscillations was 3 Hz and was lower than the frequency of tremor (about 4.8

Hz). Nevertheless, the activity of one of these neurons was coherent with the EMG at 4.8

Hz (see Figure 4.F2C) suggesting that some of its activity was nevertheless related to the

tremor.

Since the relationship between tremor-frequency neuronal activity in the STN and tremor

is most likely a nonstationary process, we examined this relationship over periods of

relatively constant tremor amplitude. Consistent with our conventional coherence

analyses, only 7 out of the 15 tremor-frequency neurons were significantly related to the

tremor over time. These neurons were coherent with the tremor for 86 ± 23 % (mean ±

Page 128: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

116

SD) of the time. Figure 4.F4A shows an example of temporal coherence between a

tremor-frequency neuron and the simultaneously recorded tremor.

Six tremor-frequency neurons were recorded during periods of altering tremor amplitude.

Three of these neurons were oscillating significantly only during periods of stronger

tremor, whereas 2 neurons were oscillating significantly only when tremor ceased. One

neuron was oscillating during both periods of tremor and non-tremor. Figure 4.F4B

shows an example of the activity of a tremor-frequency neuron over time. This neuron

exhibited significant tremor-frequency oscillations only during episodes of simultaneous

limb tremor (although, in this case, there was no significant coherence with the recorded

tremor). The tremor-frequency activity could be observed together with significant beta

oscillations.

4.3.5 - LFP oscillatory activity in the 35-55 Hz gamma frequency band is increased

during tremor

LFP signals during periods with altering tremor amplitude were recorded from 58 sites

within the STN (mean recording duration ± SD: 91.1 ± 69.7 sec, range: 34-436 sec).

Interestingly, in 49 out of the 58 sites (85%), we observed an increase in the LFP power

in the low gamma-frequency range (35-55 Hz) during periods of stronger tremor

compared to periods of weak or no tremor. Examples of increased LFP gamma power

with increased tremor amplitude are shown in Figures 4.F5 and 4.F6. The majority of

these sites (40/49) were located in the dorsal 3.0 millimeters of the STN. In contrast, 7

out of the 9 sites where the LFP gamma power did not increase during periods of stronger

tremor were in the ventral 3.0 millimeters. The distributions of sites where gamma

activity increased and sites with no increase were significantly different (mean depths ±

SD: -2.0 ± 1.4 vs. -3.3 ± 1.5 mm respectively, P = 0.013, t-test). The increase in the LFP

power in the low gamma-frequency range was observed in each of the 9 patients (Figure

4.F7A).

In 37 out of the 58 sites, neuronal spiking activity was recorded simultaneously with the

Page 129: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

117

LFP. The mean firing rate of these neurons was significantly higher during periods of

stronger tremor amplitude relative to periods of weak/no tremor (mean ± SD: 48.7 ± 27

and 38.3 ± 23.9 Hz respectively, P 0.001, paired t-test).

The LFP power over the gamma frequencies was then averaged across patients for the

periods of weak tremor (mean tremor RMS = 0.14) and periods of stronger tremor (mean

tremor RMS = 1.12). On average, the gamma power was significantly higher during the

periods of stronger tremor (median power: 0.014 and 0.026 for weak and stronger tremor

respectively; n = 58, P 0.001, Wilcoxon singed rank test) (Figure 4.F7B). Figure 4.F7C

shows an example of time frequency analysis revealing increased low gamma activity

during tremor.

An increase in the LFP power in the tremor frequencies was also observed, however, we

cannot rule out the possibility that this increase is due to tremor-movement related

artifacts.

4.3.6 - Altered balance between beta and gamma rhythms during periods of

stronger tremor

In addition to the increased power of individual LFPs, the coherence between two LFPs

in the low gamma frequency range was also increased during periods of stronger tremor

(see Figures 4.F5 and 4.F6: B3, C3). This was observed in 17 (65%) of the 26 sites where

the two LFPs were recorded simultaneously. In comparison, the coherence between LFPs

in the beta frequency range was increased in 46% of the cases (12/26 sites). When

averaging across sites for the periods of weak tremor (mean RMS = 0.13) and stronger

tremor (mean RMS = 1.12), we observed that the gamma coherence was significantly

increased with increasing tremor (medians: 0.01 and 0.23 for weaker and stronger tremor

respectively; n = 26, P 0.001, Wilcoxon singed rank test). On the other hand, beta

coherence did not change significantly (means: 0.38 and 0.33; P = 0.28, paired t-test).

Figure 8A shows the relative changes in beta and gamma coherence that were calculated

by normalizing the coherence values during periods of strong tremor to the values that

Page 130: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

118

were measured during weak tremor. On average, the coherence in the beta frequencies

decreased only by 0.04 ± 0.2, whereas the coherence in the gamma frequencies increased

by 0.14 ± 0.2 (mean ± SD). In addition, the averaged ratio of the coherence value in the

beta frequencies and the coherence value in the gamma frequencies was significantly

lower during periods of strong tremor versus weak tremor (medians: 3.95 and 1.28; P =

0.003, Wilcoxon singed rank test) (Figure 4.F8B).

To provide further support for the relative increase in gamma coherence compared to beta

coherence, we used temporal coherence to calculate the percentage of time during which

the simultaneously recorded LFPs were coherent with each other. We found that during

the periods of stronger tremor, gamma coherence significantly increased in duration and

lasted for 58.7 ± 33.5% of the time (mean ± SD) compared to 34.5 ± 32.8% during

periods of weaker tremor (P 0.001, paired t-test). The duration of beta coherence, on

the other hand, did not alter significantly (mean ± SD: 83.5 ± 34.2 and 93.1 ± 13.1% for

weak and stronger tremor respectively) (Figure 4.F8C). An example of an increase in

gamma coherence with increasing tremor amplitude is shown in Figure 4.F8D.

Page 131: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

119

Page 132: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

120

Page 133: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

121

Page 134: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

122

Page 135: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

123

Page 136: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

124

Page 137: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

125

Page 138: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

126

4.4 - Discussion

This study provides data documenting the relationship of STN neuronal and LFP

oscillatory activity to rest tremor in PD. This is the first study to show that STN neurons

oscillating at PD tremor frequencies are not only located in the same region as beta

oscillatory neurons (in dorsal STN) but are also coherent with the simultaneously

recorded local field potentials in the beta frequency band. In addition, we have shown

that the oscillatory LFP activity in the low-gamma frequency band is enhanced during

periods when patients exhibit tremor at rest.

4.4.1 - Tremor-frequency neuronal activity: relation to tremor

In the present study we found that in tremulous PD patients, nearly 15% of STN neurons

exhibit significant oscillations at the tremor frequency and that in 80% of the cases, the

frequency of oscillations was similar to the frequency of limb tremor suggesting that this

oscillatory activity might be tremor-related. However, significant coherence between

tremor-frequency oscillation and limb tremor was only observed in 47% of the cases

(Table 4.T2). Tremor characteristically varies in location and timing, and practical

limitations preclude simultaneous monitoring of all somatic muscles that may be

tremulous. Also, in our study in many cases accelerometer signals were used to

determine coherence with firing and may not detect low intensity tremors. This may

account at least in some of the cases for the lack of correlation between the STN tremor-

frequency neuronal activity and the tremor recorded in the sampled muscles. Our

observation is consistent with previous reports showing that neuronal oscillations in the

globus pallidus are not always coherent with the tremulous limb (Hurtado et al.,

1999;Lemstra et al., 1999;Raz et al., 2000), although their oscillation frequency is similar

to the frequency of the limb tremor (Hutchison et al., 1997). Several studies of multilimb

EMG recordings in PD patients also indicated that tremor in different limbs is largely

uncorrelated (Ben-Pazi et al., 2001;Hurtado et al., 2000;Raethjen et al., 2000). These

studies, together with the fact that only half of the tremor-frequency neurons fire

coherently with the local field potential, support the earlier hypothesis that independent

Page 139: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

127

oscillatory circuits may underlie parkinsonian tremor in different extremities (Alberts et

al., 1969). This idea was further strengthened by Hurtado et al. (2005) who showed that

tremor-related activity in the globus pallidus can be coherent with one tremulous limb but

not with the other. A multichannel recording method might be able to reveal neuronal

sharing and network movement of oscillations amongst different cells.

Interestingly, we found that oscillatory activity in single cells, as well as their coherence

with limb tremor, occurs intermittently, and in some cases independently from the

fluctuations in tremor amplitude. This finding suggests that tremor oscillatory circuits are

not only independent but that their oscillatory activity and synchrony with tremor can

fluctuate over time. These findings are similar to those of Hurtado et al. (1999, 2005)

who reported transient synchronization between limb tremor and neuronal oscillations in

globus pallidus of PD patients. This data suggest that tremor might be a property of the

population activity in an oscillatory network, as oppose to a property of the individual

cells (Hurtado et al., 2005).

4.4.2 - Tremor-frequency neuronal activity: relation to beta oscillations

We have demonstrated that the vast majority (~87%) of the tremor-frequency neurons are

located in the dorsal part of STN where neurons firing with beta oscillatory rhythms are

observed. This observation is consistent with an earlier report that 84% of the tremor-

related neurons are located in the dorsolateral STN (Rodriguez-Oroz et al., 2001) and

with our previous study showing that most of the beta oscillatory neurons are located in

the dorsal portion of STN (Weinberger et al., 2006) where also the beta LFP oscillatory

activity is maximal (Kuhn et al., 2005;Weinberger et al., 2006). Anatomical studies

indicate that the dorsolateral region receives inputs from the primary motor cortex

(Monakow et al., 1978;Nambu et al., 1996), whereas inputs from the premotor and

supplementary motor areas project mainly to the medial region of the STN (Nambu et al.,

1997). The dorsolateral STN contains neurons responsive to passive and active

movements in both monkey and human (Abosch et al., 2002;DeLong et al.,

1985;Rodriguez-Oroz et al., 2001). In contrast, neurons located in the ventral and medial

Page 140: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

128

region of the STN have relatively weak or no sensorimotor responses and are reciprocally

connected with the associative and limbic cortical regions (Maurice et al., 1998;Parent

and Hazrati, 1995a). It is therefore not surprising that neurons with tremor-related activity

are clustered in the dorsal STN, which is most likely to be the STN region involved in

mediating the cardinal motor features of Parkinson's disease. It has been shown that

microstimulation in the dorsolateral STN can induce tremor arrest with a short latency of

<200 ms (Rodriguez-Oroz et al., 2001), and that local block of this region by lidocaine or

muscimol microinjections reduces tremor (Levy et al., 2001b). Our results provide further

support for the association of this region with tremor and as a clinically effective target

for deep brain stimulation to alleviate PD tremor (Herzog et al., 2004).

Interestingly, apart from the similar distribution of tremor-frequency and beta oscillatory

neurons within the STN, our results demonstrate that while only half of the tremor-

frequency neurons oscillate coherently with the local field potential at the tremor

frequencies, the majority fire in coherence with the LFP at the beta frequencies. This is

surprising since only 2/25 showed both tremor and beta oscillatory activity and also in

light of the growing evidence for the lack of relationship between beta oscillations and

tremor (Amirnovin et al., 2004;Kuhn et al., 2006b;Ray et al., 2008;Silberstein et al.,

2003;Wang et al., 2005;Weinberger et al., 2006). Instead, beta oscillations have been

suggested to contribute to bradykinesia and rigidity (Brown, 2003;Chen et al., 2007;Kuhn

et al., 2006b;Kuhn et al., 2008;Ray et al., 2008). The low incidence of neurons with both

tremor and beta oscillation could indicate the existence of two separate but interdigitated

populations of neurons with different functions and connectivity within the same region.

However, the fact that tremor-activity neurons are coherent with LFP beta oscillatory

activity suggests the opposite – i.e. that most if not all of the neurons in the dorsal region

receive oscillatory inputs in both frequencies. The tremor-frequency neurons might

therefore exhibit weak beta oscillations that failed to reach significance. The coherence

analysis used in this study is more sensitive and reveals activity that may not reach

significance in the power spectrum.

In light of these results, we hypothesize that those neurons in STN that are likely to

Page 141: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

129

oscillate at tremor frequencies, are also those prone to oscillate at beta frequencies. In

other words, the same STN neuron that oscillates at tremor frequencies during tremulous

behavior can oscillate at beta frequencies during akinetic/rigid behavior. When

examining the distribution of the neurons among our tremulous group of patients, it can

be observed that tremor oscillatory activity was detected in most patients (7 out of 9),

whereas beta oscillatory neurons were only observed in 4 patients (see Table 4.T1). It is

possible that in tremulous patients, neurons in the dorsal STN are more likely to oscillate

at the tremor frequencies since cortical oscillations at these same frequencies are also

increased and this region receives direct projections from the corresponding cortical

regions (Hellwig et al., 2000). In our previous study (Weinberger et al., 2006) we found

that 28% of the neurons in STN display significant beta oscillatory activity (also see

Chapter 3). On the other hand, in this study, which included a different population of

patients, only 12% of the neurons were oscillating at beta frequencies. However, overall

24.5% of the neurons were oscillatory at either tremor or beta frequencies. These neurons

are perhaps those which, in the dopamine depleted state, become more sensitive to

rhythmic cortical inputs (Bevan et al., 2007;Magill et al., 2001;Paz et al., 2005;Sharott et

al., 2005b). Supporting this hypothesis and consistent with previous finding by Levy et al.

(2000), there was no significant difference between the mean firing rate of neurons with

tremor activity and neurons with beta activity. These neurons fire at a faster rate relative

to the non-oscillatory neurons. The higher mean firing rates of the oscillatory neurons is

perhaps due to a greater efficacy of the excitatory cortico-subthalamic inputs on those

neurons following dopamine depletion (Bevan et al., 2007;Magill et al., 2001). The

chronic loss of dopamine in the STN (in addition to its loss in the striatum) might further

promote the capability of the cortex to drive rhythmic STN activity through actions both

on the excitability of STN neurons and on the activity-dependent plasticity at synapses in

the STN (Bevan et al., 2006).

4.4.3 - Low-gamma LFP oscillatory activity: relation to tremor

Increased LFP power in the low gamma frequency range (35-55 Hz) was observed during

periods of stronger tremor. We have shown that the increase in gamma activity during

Page 142: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

130

tremor is most likely to occur at sites in the dorsal STN suggesting it is related to

sensorimotor processing. It is yet to be elucidated, however, whether this increase is

associated with the mechanism of generation of resting tremor in PD patients, or is

simply a result of increased motor drive to, or output from, STN during tremor. In

addition to the increased low gamma power at individual sites within STN, we also

observed increased coherence between nearby sites, suggesting that the low gamma LFP

oscillations are distributed and synchronized over at least a few millimeters within the

STN. It has been previously demonstrated that subthalamic gamma LFP oscillations

recorded from PD patients during rest are greater within the upper STN and bordering

zona incerta, where they are consistently synchronized to neuronal discharge and might

therefore reflect synchronized population activity of local neurons (Trottenberg et al.,

2006).

In our study, the mean firing rate of local neurons was significantly higher during

stronger tremor, suggesting that the elevated gamma LFP activity is associated with

changes in STN neuronal firing rates during tremor. Indeed, it has been recently shown

that elevations of STN gamma LFP activity do not only influence the relative timing of

spikes in spike trains (Trottenberg et al., 2006), but also have a major effect on

information carrying capacity (Pogosyan et al., 2006). It has been argued that STN

gamma oscillations can increase the information that can be transmitted and decoded by

neurons downstream of STN (Foffani and Priori, 2007). Thus, exaggerated increase in

overall information flow in downstream neurons may indirectly contribute to the

generation of rest tremor in PD. In line with our findings, it has been recently shown that

deep brain stimulation in the zona incerta, where gamma oscillations are maximal

(Trottenberg et al., 2006), improves PD resting tremor by 95% (Plaha et al., 2008).

Moreover, an earlier study showed that electrical stimulation of primary motor cortex

during neurosurgery at a frequency of 60 Hz evoked 5 Hz tremor, whereas 1-8 Hz

stimulation resulted in movements of the same frequency as the stimulation (Alberts,

1972), suggesting that abnormal high frequency synchronization may indirectly cause

tremulous movements.

Page 143: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

131

We found that the ratio of the beta to gamma coherence was significantly lower during

stronger tremor, suggesting that alteration of the balance between beta and gamma

synchronization in the STN is related to tremor and/or its magnitude. STN beta

oscillations are prominent in PD patients withdrawn from dopaminergic therapy (Brown

et al., 2001;Levy et al., 2002a;Priori et al., 2004;Williams et al., 2002) and are thought to

play an anti-kinetic role (Brown, 2003;Kuhn et al., 2004;Williams et al., 2005). As

mentioned above, increasing evidence suggests that beta oscillations contribute to

bradykinesia and rigidity rather than to tremor in PD (Chen et al., 2007;Kuhn et al.,

2006b;Kuhn et al., 2008;Ray et al., 2008;Kuhn et al., 2009). In contrast, STN oscillations

in the gamma range are increased following dopaminergic therapy and during movement,

and have been related to specific coding of movement related parameters (Androulidakis

et al., 2007;Brown, 2003;Cassidy et al., 2002;Williams et al., 2002). Negative correlation

between beta and gamma oscillations in STN has been shown to occur after treatment

with dopaminergic medications (Alonso-Frech et al., 2006;Fogelson et al., 2005a)

providing further evidence for the reciprocal interactions between these two functionally

different oscillatory patterns. It has been argued that the balance between beta and

gamma oscillations determines the effects of basal ganglia-thalamocortical projections on

the motor areas of the cortex (Brown, 2003). Our results may therefore suggest that

altered balance between these two modes of activity may also be related to resting tremor

in PD. It is important to mention that the baseline levels of gamma LFP in our study may

be relatively low as the recordings were done during off medication state. The elevated

gamma levels together with the relatively high beta activity (that did not decrease during

tremor episodes) may characterize the STN state during resting tremor in PD.

It is important to note, however, that the gamma oscillatory activity we observed was in

the low gamma range (35-55 Hz) and we did not see activity at the higher range of

gamma frequencies (up to 90 Hz) usually observed in STN following levodopa treatment

and with movement. In fact, parkinsonian rest tremor is known to be suppressed with

onset of voluntary movement of the same limb. Thus, the low-gamma activity observed

during tremor might not necessarily be movement-related. Parkinsonian rest tremor is

diagnosed only if its amplitude increases during mental stress (Deuschl et al., 1998).

Page 144: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

132

Therefore, an alternative hypothesis is that subthalamic low-gamma oscillations are

related to mental processes that lead to fluctuations in STN state during tremor. Gamma-

band synchrony in the 40-60 Hz range has been implicated in intense mental activity and

in various cognitive functions such as memory and attention (Fitzgibbon et al.,

2004;Kaiser and Lutzenberger, 2005;Kissler et al., 2000;Tallon-Baudry et al.,

1998;Tiitinen et al., 1993). Attention-related gamma power modulations were reported in

humans especially over the frontal and central cortices (Tiitinen et al., 1993;Ward, 2003).

The cerebral cortex is known to influence the activity of the STN directly via

monosynaptic projections (Kitai and Deniau, 1981). As mentioned above, in the

dopamine depleted state, there is an increased coupling between cortex and STN

activities (Magill et al., 2001) perhaps due to a greater efficacy of the excitatory synaptic

inputs (Bevan et al., 2007). It has been previously suggested that basal ganglia beta- and

gamma-band oscillatory activity may arise in the cortex (Hammond et al., 2007). Thus,

the low-gamma synchrony observed in STN during tremor might reflect an enhanced

response to rhythmic cortical input during periods when there is increased low-gamma

cortical activity.

The significance of increased low-gamma activity during tremor requires future studies,

but we believe that this might represent an important step to better understand tremor in

PD. One way to test the hypothesis that low-gamma activity is related to attentional

processes is to analyze the latency of the appearance of low-gamma oscillations in STN,

and possibly in the cortex, while patients are asked to perform mental arithmetics.

Page 145: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

133

5 - CHAPTER 5 - OSCILLATORY ACTIVITY IN THE GLOBUS PALLIDUS:

COMPARISON BETWEEN PARKINSON’S DISEASE AND DYSTONIA

Abstract

Deep brain stimulation in the globus pallidus internus (GPi) is used to treat the motor

symptoms of both Parkinson's disease (PD) and dystonia. It has been suggested that PD

and dystonia are characterized by different temporal patterns of synchronized oscillatory

activity in the GPi. We hypothesize that distinct patterns of oscillations can be observed

in the neuronal firing and local field potentials (LFPs) in GPi of PD compared to dystonia

patients. To test this hypothesis, extracellular recordings of GPi neuronal firing and LFPs

were simultaneously obtained using two microelectrodes separated by about 1 mm during

stereotactic surgery for the implantation of DBS electrodes in two PD patients and six

dystonia patients. In the PD patients, beta (11-30 Hz) oscillations were observed in the

LFPs and these were coherent between the 2 electrodes. Furthermore, of the 73 neurons

recorded in these patients, the firing activity of 21 (~29%) was significantly coherent

with the LFP recorded from one or both microelectrodes. In addition, the average beta

LFP power in the dorsal and ventral GPi of each patient was closely related to the

percentage of correlated neurons within each region. In contrast, in the dystonia patients,

the peak frequency of LFP oscillations was lower (8-20 Hz) and the activity of only 27

(10.5%) of the 257 neurons recorded in these patients, was coherent with the LFP

oscillations. This proportion was significantly lower than in PD (P 0.001). These

findings suggest that the oscillatory LFPs recorded from the GPi in PD patients are

closely associated with beta oscillatory synchronous activity in populations of GPi

neurons. In dystonia, however, the frequencies of LFP oscillations are lower than in PD

and are not as closely related to the neuronal firing. It is possible that in PD synchronized

oscillatory inputs to the GPi are more likely to influence the neuronal firing than in

dystonia perhaps due to a stronger influence of dendritic potentials on the soma in the

dopamine depleted state and/or a differential source of oscillatory input.

Page 146: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

134

5.1 - Introduction

Parkinson‟s disease (PD) and dystonia are two movement disorders that are known to be

of basal ganglia (BG) origin. PD is a neurodegenerative disorder, which is characterized

by a severe loss of dopaminergic neurons in the substantia nigra pars compacta and motor

symptoms such as rigidity and akinesia. Classical models of basal ganglia function are

based on discharge rates in the BG structures and predict that in PD, the dramatic

decrease in dopamine concentration in the striatum results in an increased activity in the

output nuclei, the globus pallidus internus (GPi) and substantia nigra pars reticulata, and

over-inhibition of the thalamocortical drive (Albin et al., 1989;DeLong, 1990).

Conversely, dystonia, which is characterized by sustained co-contractions of agonist and

antagonist muscles that lead to abnormal posture and involuntary movements, is believed

to result from decreased basal ganglia output that, in turn, leads to decreased inhibition of

thalamic activity and consequently to increased excitability of the motor cortex (Vitek,

2002). These predictions of the so-called „rate model‟ have been generally confirmed in

animal models of PD (Miller and DeLong, 1987;Filion and Tremblay, 1991;Boraud et al.,

1998) and are also supported by recording studies in humans undergoing functional

neurosurgery as a treatment for PD or dystonia (Lenz et al., 1998;Vitek et al., 1998;Vitek

et al., 1999;Merello et al., 2004;Zhuang et al., 2004;Starr et al., 2005;Tang et al., 2007).

It is noteworthy, however, that the predicted rate changes have not been found in all

studies (Bergman et al., 1994;Hutchison et al., 2003;Boraud et al., 2002;Lenz et al.,

1999;Pessiglione et al., 2005;Raz et al., 2000;Wichmann et al., 1999). Instead, there is

increasing evidence for changes in the activity patterns of basal ganglia neurons. In

parkinsonism, the incidence of bursts in the globus pallidus and subthalamic nucleus

(STN) is increased (Filion and Tremblay, 1991;Boraud et al., 1998;Hutchison et al.,

1994), most often in the context of synchronous oscillatory activity, which is

pathologically high in both the STN and GPi (Bergman et al., 1994;Nini et al.,

1995;Bergman et al., 1998b;Raz et al., 2000;Levy et al., 2002b;Magnin et al., 2000)

(Brown et al., 2001;Silberstein et al., 2003;Levy et al., 2002a). This might be, in part, due

to an enhanced interaction between cortical areas and basal ganglia structures (Magill et

Page 147: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

135

al., 2000;Magill et al., 2001;Sharott et al., 2005b;Baufreton et al., 2005a;Walters et al.,

2007), although other pathological phenomena such as the STN-GPe oscillatory network

(Plenz and Kital, 1999) may also play a role. The oscillatory discharges are rapidly

reversed by treatment with dopaminergic medications (Levy et al., 2002a), and are

positively correlated with the patients‟ response to the medications (Weinberger et al.,

2006), suggesting that these oscillations might be a direct effect of dopamine deficiency.

Importantly, abnormal oscillatory firing in the GPi has also been observed in patients

with dystonia (Starr et al., 2005). It is known that a lesion or high-frequency stimulation

in the GPi ameliorates the motor disabilities of both PD and dystonia (Marsden and

Obeso, 1994b;Lang et al., 1997;Lozano et al., 1997;Vidailhet et al., 2005;Lozano, 2001).

This paradoxical improvement, which is inconsistent with the rate model, might be

explained if the different symptoms are dependent on abnormal, but different, patterns of

synchronized oscillations in the basal ganglia circuits, so that eliminating this activity by

lesion or high frequency stimulation results in decreased disability.

One means of studying changes in the pattern of local neuronal synchrony is through a

frequency-based analysis of local field potentials (LFPs). These signals are believed to

reflect synchronized dendritic currents averaged over a larger neuronal population. In a

relatively recent study carried out by Silberstein et al. (2003), it has been shown that PD

and dystonia are indeed characterized by different oscillatory patterns. The LFPs

recorded from the deep brain stimulation (DBS) electrodes implanted in the GPi reveal

that in PD the dominant frequency of oscillation is between 11 to 30 Hz, known as the

beta range. In dystonia patients, the power of LFP beta oscillations are weaker and there

is increased oscillatory activity in the lower frequencies (4-10 Hz) (Silberstein et al.,

2003). In contrast to beta oscillatory activity which is related to rigidity and bradykinesia

(Kuhn et al., 2006b;Kuhn et al., 2008;Chen et al., 2007;Ray et al., 2008;Kuhn et al.,

2009), the 4-10 Hz oscillations have been implicated in dyskinesias (Silberstein et al.,

2003;Silberstein et al., 2005a).

In PD patients, beta activity in the GPi is mainly evident in the local field potentials

Page 148: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

136

recorded from the macroelectrodes used for therapeutic high-frequency stimulation of

this region (Brown et al., 2001;Priori et al., 2002;Silberstein et al., 2003;Silberstein et al.,

2005a). Although synchronized beta oscillatory neuronal firing in the GPi has been

previously reported (Levy et al., 2002b;Tang et al., 2007), the relationship between

neuronal firing and the LFP in GPi remains to be elucidated. Previous studies in STN of

PD patients have demonstrated a close association between oscillatory neuronal firing

and the LFP in the beta frequencies (Kuhn et al., 2005;Weinberger et al., 2006). This

phenomenon was hypothesized to result from a greater sensitivity of basal ganglia

neurons to oscillatory inputs in the dopamine depleted state (Magill et al., 2001;Sharott et

al., 2005b;Baufreton et al., 2005a;Walters et al., 2007). This hypothesis can be examined

in the GPi by comparing the relationship between neuronal firing and LFP in PD and

dystonia. This comparison is possible since the GPi is widely targeted to treat dystonia

(Vidailhet et al., 2005) and in some cases, levodopa-induced dyskinesias in PD patients

(Krack et al., 1998).

In the present study, we investigated the relationship between neuronal discharges and

the oscillatory local field potentials in the GPi in both PD and dystonia patients. The

degree of association between neuronal firing and LFP, as well as the predominant

frequency of oscillations, was compared between PD and dystonia. In addition, we

investigated the changes in GPi oscillatory activity during levodopa-induced dyskinesias

in PD.

5.2 - Methods

Patients

We studied two PD patients undergoing stereotactic neurosurgery for the implantation of

DBS electrodes in the left GPi, mainly to treat their severe levodopa-induced dyskinesias

and motor fluctuations. The second patient (Table 5.T1) also suffered from off-period

dystonic cramping in the lower limbs, and had previously undergone bilateral

implantation of DBS electrodes in the STN. These two patients were the only PD cases

selected for GPi surgery recently in our center. Both patients were women who, at the

Page 149: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

137

time of operation, reached an age of 70 and 40 years and PD duration of 27 and 8 years

respectively. For comparison purposes, we also studied one male and five female

dystonia patients with a mean age of 55.8 ± 9 (±SD) years and disease duration of 16.2 ±

14 years. No sedatives or anesthetics (e.g. propofol) were administered during or before

surgery. Pre-surgical clinical assessments of all patients were performed by movement

disorder specialists at the Toronto Western Hospital. PD patients were assessed according

to the Unified Parkinson's Disease Rating Scale (UPDRS) (Fahn et al., 1987) before and

after acute levodopa challenge (Moro et al., 2002), whereas dystonia patients were

evaluated according to the Toronto Western Spasmodic Torticollis Rating Scale

(TWSTRS) (Comella et al., 1997) and the Burke-Fahn-Marsden (BFM) rating scale

(Burke et al., 1985). Clinical and demographic details of the patients are given in Table

5.T1.

Recordings

The methods of microelectrode-guided stereotactic neurosurgery for the implantation of

DBS electrodes into the GPi are described in detail in the General Methods (Section

2.2.2.b). Extracellular recordings of neuronal firing and LFPs were obtained

simultaneously using two independently driven microelectrodes (about 25 µm tip length,

axes 600 µm apart, ~0.2-MΩ impedance at 1 kHz) as described in the General methods

(sections 2.2.3 and 2.2.4). The distinction between the external and internal segments of

the pallidum was determined based on characteristics of neuronal activity such as firing

patterns and background noise, and the existence of border cells and white matter. The

Page 150: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

138

bottom of the GPi was determined based on white matter and identification of the optic

tract, which lies close to the ventral border of the GPi. An example of simultaneous GPi

recordings of LFP and neuronal discharge is shown in Figure 5.F1A. All recordings were

performed in the resting state while the patients were awake and with local anesthesia.

Patients were not asked to perform any task and epochs with voluntary movements were

excluded. Limb movements were measured using electromyography (EMG) and/or

triaxial accelerometers (with summated x-y-z signals) that were recorded simultaneously

with the neuronal activity.

The first PD patient (#1) was given 1 tablet (100 mg) of levodopa (Sinemet,

levodopa/carbidopa 100/25) in the morning prior to the surgery as it was deemed

medically necessary for the patient. Microelectrode recordings started approximately two

hours later so that during the first trajectory, which lasted for about 1.5 hours, the patient

was generally “on” and displayed levodopa-induced dyskinesias almost continuously

during the recording session. During the second microelectrode trajectory, which was

preformed at least 3.5 hours after the last dose of levodopa, there was an eradication of

dyskinesias together with recurrence of rest tremor in the left foot. This track was

therefore considered to be in the “semi-off” state. The second PD patient (#2) was given

1 tablet of Sinemet (100/25) approximately 4 hours before surgery. In this patient, only

one microelectrode trajectory was performed (see Table 5.T1), during which the patient

was very stiff and was considered to be “semi-off”. The term “semi-off” was chosen in

order to distinguish the patients‟ state in the current study from the classical “off”, in

which patients are withdrawn from their anti-parkinsonian medication for at least 12

hours before surgery (also see Discussion). Dyskinesias were not observed during the

“semi-off” state in either of the PD patients, and importantly, dystonic activity was not

observed during the recordings in any of the dystonia patients.

Data analysis

Action potentials arising from single- and multiunits were discriminated using template

matching tools in Spike2 (Cambridge Electronic Design, Cambridge, UK). Only

recordings ≥17 seconds and during periods without voluntary movements (based on EMG

Page 151: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

139

recordings and observations) or artifacts were analyzed. Border cells were identified and

excluded from the analyses (see (Lozano et al., 1998)). Spike times and unfiltered LFP

data were imported into MATLAB (version 7.1, The MathWorks, Natick, MA) and

spectral analyses were performed using the discrete Fourier transform and its derivations

calculated according to Halliday et al. (1995) and Rivlin-Etzion et al. (2006). These

calculations included power spectra of recorded activity (spike trains and LFPs) as well

as coherence and cross-correlations between simultaneous recordings, as previously

described in the General methods (section 2.3.1.a/b).

Locations of recording sites were determined retrospectively according to the

physiological landmarks observed intraoperatively and a reconstruction of the electrode

trajectory using the stereotactic atlas of Schaltenbrand and Wahren (1977). From these

reconstructions, recording sites were approximated to be either in the dorsal or ventral

part of the GPi. In order to compare the relative LFP beta power in the dorsal and ventral

GPi in each PD patient, the mean power across the 10-Hz band centered on the frequency

of the peak power was calculated at each recording site. The LFP power at each recording

site was then expressed as the percentage of the maximum power observed

in the

trajectory. Finally, the percentage values were averaged in each half of the GPi to give

mean percentage of LFP beta power in the dorsal and ventral GPi. This method of

measuring LFP power changes in basal ganglia nuclei was used previously (Kuhn et al.,

2005;Weinberger et al., 2006;Chen et al., 2006b).

5.3 - Results

We analyzed recordings from 73 GPi neurons along 2 tracks in two PD patients “semi-

off” medications, and from 257 neurons along 12 tracks in six dystonia patients. The

mean recording durations (SD) were 43.7 18.1 and 41.9 26.8 respectively (no

statistical difference). In addition, 17 neurons were recorded in track 1 of PD patient #1

when the patient was “on” and the findings are reported separately in section 5.3.5. The

number of microelectrode trajectories that were analyzed in each patient and the number

of neurons obtained are given in Table 5.T1.

Page 152: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

140

5.3.1 - Firing rates of GPi neurons

The mean firing rate (SD) of GPi neurons recorded from PD patients “semi-off”

medication was 79.3 39.7 Hz. In the dystonia group, the mean firing rate was 60.4

30.8 Hz and was significantly lower than that from PD patients (P 0.001, Mann-

Whitney rank-sum test on the medians: 55.1 and 71.8 Hz respectively).

5.3.2 - Coherence between neuronal firing and oscillatory LFPs in PD patients

In the PD patients, beta (11-30 Hz) oscillatory activity was observed in the LFPs

recorded from the GPi, and there was significant coherence between pairs of LFPs in 35

out of 42 sites (~83%) where LFPs were recorded simultaneously from the two

microelectrodes. Of the 73 neurons recorded in the PD patients, the firing of 21 (~29%)

neurons was significantly coherent with the simultaneously recorded LFP at the beta

frequencies. The frequency of coherence ranged between 15 to 30 Hz with a mean of

24.3 ± 4.8 Hz. Seven of these neurons were coherent with the LFPs recorded from both

microelectrodes, whereas 11 neurons were coherent only with the LFP recorded from the

same electrode. The remaining 3 neurons were coherent only with the LFP recorded from

the other electrode. Figure 5.F1 shows an example of spectral analysis of simultaneous

recordings of neuronal firing and LFP activity, in which significant coherence in the beta

range was observed.

It should be noted, however, that although ~29% of the neurons displayed significant

coherence with the LFP, significant oscillatory firing was only observed in 11 (15%)

neurons, of which 9 were coherent with the LFP (see Figure 5.F1B for an example). The

mean frequency of neuronal oscillations was 24.4 ± 5.9 Hz, and this was similar to the

mean frequency of coherence with the LFP (P = 0.8, Mann-Whitney rank-sum test).

Coherence between simultaneously recorded neurons was observed in only 3 out of 38

pairs.

Page 153: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

141

5.3.3 - The averaged beta LFP power in the GPi is closely related to the percentage

of correlated neurons

Interestingly, we found that in each of the PD patients, the average beta LFP power in the

dorsal and ventral GPi is closely related to the percentage of oscillatory and correlated

neurons within that region. In patient #1, the firing of 13 out of 31 neurons was

oscillatory and/or correlated with the LFP. Seven of these neurons were located in the

dorsal GPi, where they comprised 37% of the total number of neurons, and 6 were in the

ventral GPi where they comprised 50% of the neurons. These proportions were not

significantly different (P > 0.05, Fisher exact test). Likewise, the average beta LFP power

was not significantly different between the dorsal and ventral GPi (medians: 38.2 and

45.1% respectively; P = 0.4, rank-sum test) (see Figure 5.F2). In patient #2, 10 out of 42

neurons were oscillatory and/or correlated with the LFP. But in this case, these neurons

were unevenly distributed within the GPi: 9 neurons were located in the ventral GPi

(37.5% of the neurons there) whereas only one neuron was found in the dorsal part (5.5%)

(P < 0.03, Fisher exact test). In this patient, the average LFP beta power was significantly

greater in the ventral part (medians: 4.7 and 11.4 respectively; P 0.001, rank-sum test)

(Figure 5.F2).

5.3.4 - Neurons are less correlated to oscillatory LFPs in dystonia patients

In the dystonia group, oscillatory activity was observed in the LFPs mostly in the

frequencies between 8 to 20 Hz. Of 200 sites where LFPs were recorded simultaneously

from the two microelectrodes, 177 (88.5%) showed significant coherence across this

frequency band. Examples of the coherence between the two simultaneously recorded

LFPs in each of the six dystonia patients are shown in Figure 5.F3. Of 257 neurons

recorded in these patients, the firing activity of 27 (10.5%) was significantly coherent

with the simultaneously recorded LFP. This proportion, however, was significantly lower

than that observed in PD (P 0.001, χ2 test). Of these neurons, 6 were coherent with the

LFPs recorded from both microelectrodes, whereas the vast majority (n = 21) were

Page 154: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

142

coherent only with the LFP recorded from the same electrode. In this group of patients,

the mean frequency of coherence was 18 ± 4.2 Hz, and was determined to be

significantly lower than the frequencies observed in PD (P 0.001, t-test).

The number of correlated neurons in each of the dystonia patients varied from 0 to 11

with an average of 4.5 ± 4 neurons per patient. In general, these neurons were evenly

distributed within the dorsal and ventral halves of the GPi. In the dorsal GPi, the activity

of 12 out of the 137 neurons (8.5%) was correlated with the LFP, whereas in the ventral

part, the activity of 15 of the 120 neurons (12.5%) was correlated with the LFP.

Significant oscillatory activity of neuronal discharges at the 8-20 Hz band was not

observed in any of the dystonia patients. Because of the relatively small number of

neurons that fired coherently with the LFP in each dystonia patient, we did not compare

their distribution to that of the LFP power within the GPi.

5.3.5 - Changes in GPi oscillatory activity during levodopa-induced dyskinesias

In one of the PD patients (patient #1), we had the opportunity to perform one

microelectrode trajectory through the GPi during levodopa-induced dyskineisas. In this

trajectory, the LFPs revealed oscillatory activity at frequencies around 9 and 70 Hz. In

addition, all sites (n = 16) in which LFPs were recorded simultaneously from the two

electrodes showed significant coherence at these frequencies, whereas beta coherence

was significantly reduced (from a median value of 0.5 to 0.2; P 0.001, rank-sum test).

Figure 5.F4 shows examples of power spectra of simultaneously recorded LFPs and their

corresponding coherence and cross-correlation functions, as were observed in each of the

two microelectrode trajectories performed in this patient. Importantly, we did not find

correlation between the recorded LFP and limb dyskinesias.

Two out of the 17 (~12%) neurons that were recorded in this track, showed significant

coherence with the LFPs recorded from both microelectrodes at a frequency around 9 Hz.

One of these two neurons also exhibited significant oscillatory firing at this frequency

(see Figure 5.F5). Examination of the firing rates of neurons recorded during levodopa-

Page 155: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

143

induced dyskinesias (n = 17) revealed no significant difference from the neurons recorded

during the “semi-off” state (n = 73) in the two PD patients (mean firing rate of 74.9

29.8 Hz compared to 79.3 36.7 Hz respectively; P = 0.66 t-test).

Page 156: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

144

Page 157: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

145

Page 158: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

146

Page 159: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

147

Page 160: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

148

5.4 - Discussion

5.4.1 - Methodological constraints

The number of Parkinson‟s disease patients (n = 2) that participated in this study is very

limited because of the low prevalence of patients with PD that are selected for GPi

surgery in our center. PD patients normally receive bilateral STN implantation as a

standard therapy. These exceptional two cases were selected for GPi surgery due to their

severe drug-induced dyskinesias. In addition, patient #2 also suffered from off-period

dystonic cramping in her lower limbs. Thus, the pathology in this patient might have

some similarities to the pathology in dystonia, which might influence our results.

Moreover, this patient has been previously treated with bilateral deep brain stimulation of

the STN. Therefore, the possibility that chronic stimulation of the STN have resulted in

some long-term physiological changes in the circuit should be considered. Another

limitation was that these data were not obtained after overnight withdrawal from

dopaminergic medication since full medication withdrawal was contraindicated in these

severely affected patients. The patients were considered to be either “on” (during

levodopa-induced dyskinesias) or “semi-off” in the absence of dyskinesias and

reappearance of symptoms such as rigidity and tremor, but the exact clinical state was not

evaluated. Despite these limitations, we feel that our results should be reported in view of

their uniqueness and since they provide interesting insights regarding the

pathophysiology of PD and dystonia.

5.4.2 - GPi firing rates

In the current study, the firing rates of GPi neurons were significantly higher in the PD

group compared to the dystonia group. This is consistent with the predictions of the rate

model of the basal ganglia for PD (Albin et al., 1989;DeLong, 1990) and dystonia (Vitek,

2002), although the absence of control data limits our ability to determine whether the

firing rates observed in PD/dystonia were respectively higher/lower than normal. In the

normal monkey, the mean GPi firing rate was found to be around 80 Hz (Filion and

Page 161: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

149

Tremblay, 1991;Starr et al., 2005), which is similar to the mean rate of 79 Hz found in

our PD patients. This might imply that the firing rates observed in our PD patients are

close to normal. Indeed, previous studies in PD patients that were completely off

dopaminergic medication have reported higher GPi firing of about 90-95 Hz (Starr et al.,

2005;Tang et al., 2005;Tang et al., 2007), which is comparable to the firing observed in

the GPi of MPTP monkeys (~95 Hz) (Filion and Tremblay, 1991). This suggests that the

“semi-off” state of our PD patients (about 4 hours after the last dose of levodopa) might

not be comparable to the practically-defined off state, in which patients are withdrawn

from their anti-parkinsonian medication for at least 12 hours prior to the recordings. In

fact, the mean firing rate recorded during the “on” levodopa state (~75 Hz) was not

significantly different than that recorded during the “semi-off” state.

Because in our study the patients‟ clinical motor scores were not evaluated

intraoperatively, it was not possible to directly compare their “semi-off” state to the

complete off state. According to pharmacokinetic and pharmacodynamics studies,

levodopa plasma levels peak 0.5 to 2 hours after its administration, when the patients‟

best motor status is reached, and may take 5-6 hours to wear off completely even though

motor scores decline to baseline after approximately 3 hours (Fabbrini et al., 1987;Deleu

et al., 2002). Therefore, it is likely that the “semi-off” state in this study was associated

with residual levels of dopamine in the system, which might explain the relatively low

firing rates observed in our PD patients. Indeed, it has been recently shown in PD patients

that basal ganglia output neurons are sensitive to even low doses of levodopa, as

observed in the form of synaptic plasticity (Prescott et al., 2009). Alternatively, the low

firing rates might be due to the fact that the two PD patients in this study suffered from

severe dyskinesias. Some studies have reported decreased firing rate in GPi neurons in

PD patients with off-period dystonia and levodopa-induced dyskinesias (Hallett,

2000;Hashimoto et al., 2001).

In the dystonia group, on the other hand, the mean GPi firing rate found (~60 Hz) is

similar to that reported previously (Vitek et al., 1999;Merello et al., 2004;Starr et al.,

2005) and is lower than the mean firing observed in normal monkeys (Filion and

Page 162: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

150

Tremblay, 1991;Starr et al., 2005), suggesting that it might be abnormally reduced (Vitek,

2002). It is important to note, however, that in some studies the firing rates in the GPi of

dystonia patients were closer to normal with a mean of about 75 Hz (Hutchison et al.,

2003;Tang et al., 2007).

5.4.3 - GPi LFP oscillatory activity and its relationship to neuronal discharges in PD

We found that in PD patients, oscillatory activity in the local field potentials recorded

from the GPi during the “semi-off” levodopa state is mostly prominent in the beta-

frequency band. This finding is consistent with previous studies that have reported beta

oscillatory activity in GPi of PD patients off medications, in whom local field potentials

were recorded from the large contact DBS macroelectrodes (Brown et al.,

2001;Silberstein et al., 2003;Silberstein et al., 2005a;Foffani et al., 2005a). The present

study is the first to demonstrate beta oscillations in GPi LFPs recorded from the more

focal, higher-impedance, microelectrodes. In addition, we were able to reveal significant

coherence between simultaneously recorded LFPs, suggesting that the beta LFP

oscillations are distributed and synchronized over at least a few millimeters within the

GPi. This finding is similar to our previous observation in the STN of PD patients off

dopaminergic medication (Weinberger et al., 2006), implying that STN and GPi share

similar oscillatory patterns in PD.

Our results demonstrate for the first time that in PD, about 30% of the neurons in the GPi

are significantly coherent with the simultaneously recorded LFP and there appears to be a

spatial association within the GPi between the percentage of correlated neurons and the

relative LFP power in the beta range. These data further confirm that the beta LFP

recorded from the basal ganglia reflects, at least in part, synchronized activity in a

population of local neurons (Kuhn et al., 2005;Chen et al., 2006b;Weinberger et al.,

2006). However, although 29% of the neurons fired coherently with the beta LFP

oscillations, only 15% displayed significant beta oscillatory firing. This can be attributed

to the fact that coherence analysis is generally more sensitive and reveals activity that

may not reach significance in the power spectrum. Thus, we cannot role out the

Page 163: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

151

possibility that the correlated, non-oscillatory, neurons might exhibit weak beta

oscillations that failed to reach significance. Nevertheless, in our previous study on the

relationship between neuronal firing and beta oscillatory LFPs in the STN (Weinberger et

al., 2006), we found that as many as 28% of the neurons displayed significant oscillatory

firing, and 25% of the simultaneously recorded pairs fired in synchrony (compared to 8%

in the present study). It is possible that the relatively small number of oscillatory neurons

found in the present study is due to the “semi-off” state of the patients, as opposed to the

overnight withdrawal from dopaminergic medication in our previous study.

Alternatively, the difference between these results and those observed in STN may reflect

anatomical and physiological differences between STN and GPi. This, however, seems

unlikely since a tendency toward neuronal synchronization has been demonstrated in the

monkey GPi following MPTP treatment (Nini et al., 1995;Raz et al., 2000) and in PD

patients (Levy et al., 2002b). Such synchronization has been suggested to result, at least

in part, from a greater influence of striatal tonically active neurons that, in the dopamine

depleted state, display oscillatory activity that is correlated with neuronal activity in the

GPi (Raz et al., 1996;Raz et al., 2001). These changes have been proposed to result from

enhanced electrotonic coupling between neighbouring striatal neurons in the absence of

dopamine (Onn and Grace, 1999), which has been shown to directly modulate the activity

of striatal tonically active neurons (Watanabe and Kimura, 1998). A significant increase

in the occurrences of synchronized beta oscillatory firing was also observed among STN

neurons (Bergman et al., 1994;Vila et al., 2000;Steigerwald et al., 2008), and coherence

studies have documented that the activity in STN and GPi of PD patients are well

synchronized at the beta frequencies (Brown et al., 2001;Cassidy et al., 2002;Foffani et

al., 2005a). It has been suggested that dopamine depletion at the level of the STN and

globus pallidus may enhance the sensitivity of the network to rhythmic activity

originating from the cortex (Magill et al., 2001;Sharott et al., 2005b;Baufreton et al.,

2005a). Since the nigro-striatal dopaminergic projection has axon collaterals to the STN

and pallidum (Lavoie et al., 1989;Hassani et al., 1997;Cossette et al., 1999;Hedreen,

1999;Francois et al., 2000), extrastriatal dopaminergic modulation may facilitate the

normal interaction of subthalamic and pallidal neurons and its loss in PD may contribute

Page 164: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

152

to the pathological synchronization within and between these structures (Bevan et al.,

2006).

Taken together, it is possible that in our study, the residual amount of dopamine that

might have remained in the system during the “semi-off” state (see above) resulted in

reduced oscillatory firing and synchrony among GPi neurons by means of striatal and/or

extrastriatal mechanisms. In addition, we speculate that in some of the neurons that

showed significant coherence with the LFP, the oscillatory membrane potentials were

largely confined to the dendrites and had only a relatively small influence on the

probability of the neuron firing. Therefore, even neurons with non-significant beta

oscillatory firing may nonetheless contribute to the generation of beta oscillatory LFPs.

This hypothesis is further supported by the association observed between the distribution

of correlated neurons within the GPi and the relative LFP power at the beta frequencies.

5.4.4 - Modulation of GPi oscillatory activity during levodopa-induced dyskinesias

Interestingly, we found that although the firing rates of GPi neurons were not

significantly different between the “semi-off” and the “on” states, the oscillatory patterns

were different. During periods of levodopa-induced dyskinesias, beta activity in the local

field potential was largely reduced and there was an emergence of new peaks around 9

and 70 Hz. High gamma activity around 70 Hz in the GPi (and also in STN) has been

shown to increase in PD patients following dopaminergic medication (Brown et al., 2001)

and during voluntary movement (Cassidy et al., 2002) in concurrence with a reduction in

beta activity. This led to the hypothesis that beta synchronization may be essentially anti-

kinetic perhaps by limiting the ability of neurons to code information in time and space,

as both adjacent and spatially distributed neurons are preferentially locked to the beta

rhythm (Brown, 2003;Williams et al., 2005;Brown, 2007;Hammond et al., 2007).

Gamma activity, on the other hand, has been considered to be pro-kinetic and related to

movement demands, as it may share similar role to that posited for gamma band

synchronization in the visual cortex (Freiwald et al., 2001). Our finding is therefore

consistent with previous studies demonstrating decreased beta activity and increased

Page 165: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

153

gamma during the “on” state (Brown et al., 2001;Cassidy et al., 2002), and suggest that

gamma activity could also be observed during epochs of levodopa-induced dyskinesias.

Furthermore, it has been shown by Silberstein et al. (2005) in two PD patients that

oscillatory pallidal LFP activity over 8-30 Hz is inversely correlated to the degree of

levodopa-induced dyskinesias, suggesting that the levodopa-induced suppression of

neuronal synchronization in the basal ganglia may play a role in the origin of dyskinesias

(Silberstein et al., 2005a). However, the reduction in 8-30 Hz synchronization alone

cannot account for dyskinesias, particularly since beta suppression occurs following

levodopa in PD patients without dyskinesias. Our results demonstrate that together with

beta suppression, there is increased synchronization at lower frequencies around 9 Hz

during periods of levodopa-induced dyskinesias. These data are consistent with a recent

LFP study in a PD patient exhibiting unilateral off-period dyskinesias (Foffani et al.,

2005a). That study demonstrated that contralateral to the dyskinetic side of the body,

there was a reduced coherence between GPi and STN in the high-beta frequencies (20-30

Hz) and an enhanced coherence at low frequencies (<10 Hz) compared to the GPi and

STN ipsilateral to dyskinesias. In addition, 4-10 Hz oscillations in the LFPs were also

recorded from the STN during levodopa-induced dyskinesias, and this was observed only

contralateral to the side of dyskinesias (Alonso-Frech et al., 2006). Here, we also

revealed that during levodopa-induced dyskinesias, neuronal firing can oscillate and be

coherent with the LFP at low frequencies. This observation, together with the lack of

correlation between the low-frequency LFP oscillatory activity and limb dyskinesias,

rules out the possibility that these LFPs reflect movement-related artifacts. In fact, it has

been previously shown that the spectral content of dyskinetic movements in PD patients

is generally between 1 to 3 Hz (Manson et al., 2000)

Taken together our findings suggest that parkinsonian dyskinesias are related to an

altered balance between oscillatory rhythms within the basal ganglia network rather than

to a further decrease in firing rates. These findings also support the hypothesis that the

balance between oscillatory rhythms in the basal ganglia determines the effects of basal

ganglia-thalamocortical projections on the motor areas of the cortex (Brown, 2003). We

Page 166: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

154

have previously demonstrated that altered balance between rhythms of oscillations can

also characterize the state of the subthalamic nucleus during rest tremor in PD patients

(Weinberger et al., 2009).

5.4.5 - GPi oscillatory activity in dystonia and its comparison to PD

Our spectral analysis of the local field potentials in the GPi of dystonia patients revealed

oscillatory activity at frequencies between 8 to 20 Hz, which are higher than those (3-15

Hz) previously reported in dystonia (Liu et al., 2002;Silberstein et al., 2003;Chen et al.,

2006b;Liu et al., 2006;Foncke et al., 2007). By using microelectrode recordings, Chen et

al. (2006b) showed that in dystonia patients, 3-12 Hz oscillatory LFPs are maximal in the

GPi and can be synchronized with the activity of the neurons. This activity was also

shown to be coherent with dystonic EMG activity in the contralateral muscles (Liu et al.,

2002;Liu et al., 2006;Chen et al., 2006a;Foncke et al., 2007). Furthermore, it has been

demonstrated by Silberstein et al. (2003) that the LFPs recorded from the DBS electrodes

implanted in the GPi in dystonia patients have higher power over the 4-10 Hz band and

less power over the 11-30 Hz band compared to PD patients, suggesting that the two

diseases are characterized by different patterns of pallidal activity. It should be noted,

however, that although the percentage of LFP power in the 4-10 Hz band was higher in

patients with dystonia, a discrete peak was found consistently only in the 11-30 Hz band

and not in the 4-10 Hz band (Silberstein et al., 2003). Studies that have investigated

movement-related changes in LFPs recorded from the GPi of dystonia patients revealed

that movements cause modulation of oscillatory activity in the range of 8-20 Hz, which is

equivalent to the frequency range observed in our dystonia patients. In comparison with

rest conditions, a significant decrease in the oscillatory power of pallidal LFPs at 8-20 Hz

was induced by active and passive movements as well as by vibrations (Liu et al.,

2006;Liu et al., 2008;Brucke et al., 2008).

Here, not only did we observe 8-20 Hz LFP oscillations but we also found that the firing

of 10.5% of the neurons was correlated with the LFP at those frequencies. In Chen et al.

(2006b), spike triggered averages with significant activity were observed in 28% of the

Page 167: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

155

sites recorded in the GPi and the synchronization was mainly evident in the 3-12 Hz band.

These differences might be related to the fact that some of the patients in their study

experienced spontaneous dystonic movements during the recordings (Chen et al., 2006b).

In our study, patients exhibited neither voluntary nor dystonic movements, so that the

level of 8-20 Hz synchronization within the GPi might have remained relatively high

compared to the lower frequencies.

Importantly, we found that the percentage of neurons that were correlated with the LFP in

the dystonia group was significantly lower than that observed in PD. A possible

explanation is that synchronized activity in dystonia tends to not be oscillatory and thus,

could not be detected by our coherence analysis. Whether this is the case or not, our

findings provide further evidence for the hypothesis that dopaminergic loss at the level of

the STN and globus pallidus in Parkinson‟s disease may increase the sensitivity of the

network to rhythmic, perhaps cortical, oscillatory inputs (Magill et al., 2001;Sharott et al.,

2005b;Baufreton et al., 2005a;Walters et al., 2007), and might contribute to the

pathological synchronized oscillations within and between these structures (Bevan et al.,

2006).

In summary, our findings suggest that the oscillatory LFPs recorded from the GPi in PD

patients are closely associated with both significant and subthreshold beta oscillatory

activity in populations of GPi neurons. The relatively low firing rate and the small

number of oscillatory neurons observed in these patients may be attributed to their “semi-

off” state during the recordings. In dystonia, however, the firing rates of GPi neurons and

the frequencies of LFP oscillations are significantly lower than in PD, and the oscillatory

LFPs are not as closely related to the neuronal firing. It is possible that in PD

synchronized oscillatory inputs to the GPi are more likely to influence the neuronal firing

than in dystonia, perhaps due to a stronger influence of dendritic potentials on the soma

in the dopamine depleted state.

Page 168: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

156

6 - CHAPTER 6 - GENERAL SUMMARY AND SIGNIFICANCE

6.1 - Oscillatory firing in the basal ganglia and its relationship to the local field

potential and dopamine

The results of these studies provide important data documenting the nature of

synchronized oscillations in PD. These data add new insights on the relationship of

neuronal beta oscillatory discharges in the basal ganglia to local field potential activity in

Parkinson‟s disease patients. The coherence between neuronal discharges and LFPs in the

beta range suggests that LFP beta oscillations recorded from the basal ganglia reflect, to

some extent, rhythmic activity in a population of local neurons. This conception was

strengthened even further by our results showing topographical association between the

power of LFP beta oscillations and the locations of oscillatory and/or coherent neurons

within the STN or GPi. These findings are particularly important because they suggest

that LFP recordings from the basal ganglia, which are often used in various recent studies,

essentially reflect the activity within its nuclei. Importantly, however, it seems from our

studies that there are certain conditions during which beta oscillatory LFPs are likely to

reflect only the oscillatory synaptic inputs but not neuronal firing.

When patients were studied off medications (and only 3 out of 14 exhibited tremor during

the recordings) we found that approximately 30% of the neurons in STN display

significant beta oscillatory firing and there is a significant correlation between the

locations of oscillatory neurons within the STN and the LFP power (Chapter 3). In

comparison, we found that the proportion of beta oscillatory neurons in the STN is

somewhat lower in tremulous PD patients, and perhaps there are more neurons that

oscillate at the frequency of tremor (Chapter 4). Our data show for the first time that even

the tremor-frequency neurons can be coherent with the LFP oscillations over the beta

frequency band, suggesting that they may still contribute to the generation of beta

oscillatory LFPs. We speculate that during periods of rest tremor, beta oscillatory LFPs

are more confined to the dendrites relative to periods without tremor. Similarly, our GPi

analyses revealed that although ~30% of the neurons were coherent with the LFP, only

Page 169: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

157

half of these displayed significant beta oscillatory firing (Chapter 5), leading to the

conclusion that even non-oscillatory neurons might contribute to the generation of beta

LFPs. In this case, the relatively small number of oscillatory neurons in the GPi was

attributed to the dopaminergic state of the patients that were considered to be “semi-off”.

Taken together, our data suggest that beta oscillatory LFPs are likely to reflect oscillatory

synaptic inputs to the neurons in the recorded area and these may or may not influence

the probability of firing depending on the clinical state.

If the “semi-off” state (Chapter 5) was indeed associated with residual amounts of

dopamine in the basal ganglia (see section 5.4.2), we may conclude that the oscillatory

firing in the basal ganglia is a consequence of dopaminergic loss. This conclusion is also

supported by the finding that the proportion of oscillatory neurons in the STN of PD

patients can predict the patients‟ response to dopaminergic medication (Chapter 3). Thus,

oscillatory firing in the basal ganglia might, at least in part, reflect the degree of

dopamine deficiency in the system. Indeed, when the oscillatory activity in the GPi of PD

patients was compared to that of dystonia patients (Chapter 5), we found that in dystonia

a significantly smaller number of neurons are coherent with the LFP. Taken together, our

data add to the growing evidence (emerging from animal models and in vitro studies) that

dopaminergic loss in Parkinson‟s disease may increase the sensitivity of the network to

rhythmic, perhaps cortical, oscillatory inputs (Tseng et al., 2001;Magill et al.,

2001;Sharott et al., 2005b;Baufreton et al., 2005a;Walters et al., 2007), and might

contribute to the pathological synchronization within and between these structures

(Bevan et al., 2006). Figure 6.F1 summarizes some possible mechanisms involved in the

generation of oscillatory activity in the BG network following dopamine depletion.

6.2 - Spatial distribution of beta oscillations

We found that beta oscillations in both neuronal firing and LFPs are maximal in the

motor portion of STN (i.e. the dorsal part) (see Chapters 3 and 4). In addition, we also

obtained data suggesting that beta oscillatory activity tends to be greater in the motor

portion of GPi (i.e. the ventral part) (Chapter 5). In the GPi, however, although a trend

Page 170: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

158

was observed in the two PD patients studied, it was only significant in one of the patients.

Future studies that include larger number of patients will have to confirm that in the GPi,

beta oscillatory activity is maximal in the ventral part. It has been recently reported that

the ventral GPi is the region that shows differential changes (i.e. in firing rates and

patterns) in various movement disorders including PD (Tang et al., 2007). Overall, our

results provide important evidence that beta oscillatory activity in the basal ganglia might

be related to motor functions, therefore supporting its potential role in mediating some of

the motor symptoms in Parkinson‟s disease (Brown and Williams, 2005).

6.3 - The oscillatory neuronal population

In general, oscillatory firing in the basal ganglia of PD patients has been shown to occur

in two major frequency bands. Neurons can fire at either the beta or the tremor

frequencies (Levy et al., 2000;Levy et al., 2002b), and some neurons can display

significant oscillatory firing at both frequencies (see for example Figures 3.F1, 4.F1 and

4.F2). Our findings show that STN neurons that exhibit oscillatory firing in either the

tremor or the beta band are located in the same region within the STN and more

interestingly, neurons with tremor-frequency activity are coherent with the LFP over the

beta band. These observations suggest that tremor-frequency neurons might also exhibit

weak beta oscillations and lead us to hypothesize that the neurons that are likely to

oscillate at tremor frequencies are also those prone to oscillate at the beta frequencies and

vice versa. This hypothesis is supported by the fact that although in tremulous patients

only 12% of the neurons in STN have beta oscillatory activity, a total of 24.5% of the

neurons were oscillatory at either tremor or beta frequencies, and this percentage is

comparable to the 28% oscillatory neurons observed in Chapter 3 in which the majority

of the patients did not exhibit tremor during the recordings.

Our data therefore suggest that irrespective of the frequency of oscillations, between a

quarter to two thirds of the neurons in the STN are capable of oscillatory firing and these

neurons are more likely to be found in the motor portion of STN. We believe that these

neurons reflect the neuronal population that, in the dopamine depleted state, become

Page 171: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

159

more sensitive to rhythmic cortical inputs (Bevan et al., 2007;Magill et al., 2001;Paz et

al., 2005;Sharott et al., 2005b), which then determine their frequency of oscillations. In

agreement with this hypothesis, the firing rates of STN neurons with tremor-frequency

activity are similar to those with beta activity, and are significantly higher than those of

the non-oscillatory neurons (see Chapters 3 and 4). The higher mean firing rates of the

oscillatory neurons is perhaps due to a greater efficacy of the excitatory cortico-

subthalamic inputs on those neurons following dopamine depletion (Bevan et al.,

2007;Magill et al., 2001). Thus, our data suggest that in addition to the increased

inhibition of the GPe, the increased sensitivity of STN neurons to cortical inputs (see

Figure 6.F1) might also contribute to the established STN hyperactivity in PD (Miller and

DeLong, 1987;Bergman et al., 1994).

6.4 - The relation of oscillatory activity to clinical aspects in PD patients

In these studies, the oscillatory activity in the basal ganglia was considered in relation to

various clinical aspects. In Chapter 3, the incidence of beta oscillatory neurons in the

STN was studied in relation to the patients‟ motor symptoms and their benefit from

dopaminergic medications. Interestingly, the percentage of neurons exhibiting oscillatory

firing in the beta range correlated well with the degree by which PD motor symptoms

improved after dopamine replacement therapy, whereas the incidence of oscillatory

neurons did not correlate with the severity of symptoms before treatment. This suggests

that beta oscillatory activity does not characterize motor impairment per se. Instead, it

might be related to the magnitude of the response of the basal ganglia, or at least STN, to

dopaminergic agents. These novel findings were later supported by newer studies

showing that the LFP power over the 8-35 Hz band (measured “off” medication) is

positively correlated with the improvements in motor symptoms after levodopa intake

(Ray et al., 2008) and that levodopa-induced reduction in subthalamic LFP beta power

correlates with the simultaneously observed clinical improvement in bradykinesia and

rigidity, but not tremor (Kuhn et al., 2006b;Ray et al., 2008;Kuhn et al., 2009). Similar to

our findings, no correlation between beta power and baseline motor symptoms was found.

Page 172: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

160

The lack of correlation between beta oscillations and motor impairment is also in line

with recent studies in animal models of PD. In monkeys that underwent a gradual MPTP-

treatment protocol that slowly induces parkinsonism, except for a slight increase in the

very beginning, robust oscillatory firing in the GPi appears only after first motor

symptoms have already developed (Leblois et al., 2007), suggesting that oscillatory firing

is not directly related to the early manifestations of parkinsonian motor symptoms.

Similarly, beta oscillations in STN and cerebral cortex were not exaggerated until several

days after 6-OHDA injections in rats (Mallet et al., 2008) and are detected only several

days after the appearance of akinesia (Degos et al., 2008). These results suggest that

abnormally amplified beta oscillations in cortico-basal ganglia circuits do not result

simply from an acute absence of dopamine receptor stimulation, but are a delayed

consequence of chronic dopamine depletion.

In Chapter 4, the temporal dynamics of oscillatory activity in the subthalamic nucleus

was investigated in relation to parkinsonian resting tremor. After establishing that beta

oscillatory firing in the STN does not correlate with tremor severity (Chapter 3) we

observed that in fact, during periods of tremor patients tend to have less beta oscillatory

firing and in addition, there are neurons firing at the frequency of limb tremor.

Interestingly, however, their oscillations, as well as their coherence with limb tremor,

occur intermittently and in some cases independently from the fluctuations in tremor

amplitude. This observation is in line with previous studies showing transient

synchronization between limb tremor and neuronal oscillations in globus pallidus of PD

patients (Hurtado et al., 1999;Hurtado et al., 2005) and supports the hypothesis that

tremor might be a property of the population activity in an oscillatory network, as

opposed to a property of the individual cells (Hurtado et al., 2005). In addition, it has

been suggested that independent oscillatory circuits may underlie parkinsonian tremor in

different extremities (Alberts et al., 1969). Indeed, tremor-frequency neurons tend to

oscillate in different phases (Levy et al., 2000;Moran et al., 2008), which is perhaps why

tremor-frequency oscillations are not easily detected by LFP recordings in the STN.

Temporal examination of the LFPs recorded in our tremulous patients revealed, for the

Page 173: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

161

first time, that during periods of stronger tremor the oscillatory power in the LFP is

preferentially increased in the low gamma range (35-55 Hz). This is an intriguing finding

in light of the fact that parkinsonian rest tremor is known to increase in amplitude during

mental stress (Deuschl et al., 1998), which has been previously associated with increased

low-gamma synchrony across cortical areas (Tiitinen et al., 1993;Ward, 2003). Thus, it is

possible that the low-gamma synchrony observed in STN during tremor reflects rhythmic,

perhaps attentional-related, cortical input to the STN.

In addition to the increased LFP power in the low-gamma range, the ratio of the beta to

gamma coherence was significantly lower during stronger tremor, suggesting that

alteration of the balance between beta and gamma oscillations in the STN might be

related to tremor and/or its magnitude. These data are consistent with the notion that the

balance between oscillatory rhythms within the basal ganglia determines its effects on the

motor areas of the cortex (Brown, 2003). Similarly, overt changes in oscillatory activities

within the basal ganglia were observed during levodopa-induced dyskinesias (Chapter 5).

Beta activity in the local field potential was largely reduced and there was an emergence

of new peaks around 9 and 70 Hz. Importantly, however, there was no significant change

in firing rates, suggesting that parkinsonian dyskinesias are related to an altered balance

between oscillatory rhythms within the basal ganglia network rather than to a further

decrease in firing rates.

6.5 - Implications to basal ganglia models

Although these studies were not aimed to directly test the predictions of the rate model of

basal ganglia function (Albin et al., 1989;DeLong, 1990), results from Chapter 5 seem to

confirm that hyperkinetic movement disorders, such as dystonia, are associated with low

firing rates in GPi output neurons. Similar to hyperkinetic movement disorders,

levodopa-induced dyskinesias in PD were predicted to result from a substantial reduction

in GPi firing rates. This, however, was not supported by our data showing that GPi firing

rates remain unchanged during periods of levodopa-induced dyskinesias relative to

periods without dyskinesias. Thus, in contrast to the predictions of the model,

Page 174: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

162

hyperkinetic motor symptoms in PD might not necessarily result from reduced neuronal

firing rates in the GPi and other mechanisms must be involved.

Our results from PD patients that were off dopaminergic medications for at least 12 hours

(Chapters 3 and 4) have established that neurons in the dorsal STN tend to exhibit

significantly elevated firing rates together with pathological oscillatory activity. This

observation is not only consistent with the rate model, which predicts hyperactivity of the

subthalamic nucleus in PD, but also suggests that oscillatory neurons tend to be

hyperactive perhaps due to their greater sensitivity to excitatory cortical inputs (see

above).

The overt beta band synchronization observed in the STN and GPi of PD patients is in

concurrence with the oscillatory model of basal ganglia (Brown, 2003), which

hypothesizes that in PD the beta oscillations are enhanced and that dopaminergic

medications decrease the pathological beta band oscillations and enhance the gamma

band oscillations. Our results confirm this prediction as a reduction in beta activity was

observed in the GPi right after levodopa administration together with an increase in high

gamma activity. The oscillatory model also proposes that STN and GPi activity in the

beta band is driven from the motor areas of the cortex. This is also supported by our

findings showing that beta oscillatory activity was observed in the LFPs (likely to reflect

oscillatory inputs) even in the absence of oscillatory firing. Our data expand on the

oscillatory model by suggesting that beta oscillatory inputs to STN do not always

influence the neuronal firing depending on the clinical state (see above) and in particular,

might reflect the amount of dopamine deficiency in the system. As opposed to beta

oscillations, the oscillatory model suggests that oscillations at tremor frequencies arise

within the basal ganglia and spread to the cortex. Our data do not directly support this

prediction as they suggest instead that neurons oscillating at either tremor or beta

frequencies might be influenced by cortical input. Indeed, recent findings in

anaesthetized rats (Sharott et al., 2005a) and PD patients (Lalo et al., 2008) have

confirmed a net cortical driving of STN activity at frequencies below 60 Hz.

Page 175: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

163

According to the oscillatory model, in PD beta oscillations are enhanced to such an extent

that the desynchronization required to initiate movement cannot "break through" the

elevated threshold, leading to the poverty of movement (Brown, 2003). This idea is

challenged by our data showing lack of correlation between symptom severity and

prevalence of oscillatory neurons (presented in Chapter 3). Indeed, some evidence

suggests that pathological oscillations may occur after the appearance of first motor

symptoms (Leblois et al., 2007;Degos et al., 2008). In this regard, it should be mentioned

that the occurrence of motor impairment before the appearance of oscillations has been

previously predicted by a theoretical model that suggests that pathological

synchronization may be driven at the level of the BG-cortical loops (Leblois et al., 2006a).

According to this model, oscillatory activity emerges in the pathological condition as a

dynamic imbalance between two feedback BG–cortical loops (i.e. the direct and

hyperdirect loops), but it is not related to parkinsonian motor symptoms that may be

expressed in the absence of oscillations.

In 1998, Harris and Wolpert have suggested that signal-dependent noise plays a

fundamental role in motor planning. According to their theory, biological noise, which is

present in the neural control signals and whose variance increases with the size of the

control signal, shapes the trajectory planning of a movement to minimize the variance of

the final eye or arm position (Harris and Wolpert, 1998). Based on this theory, it is

tempting to hypothesize that elevated levels of synchonization within the basal ganglia

may interfere with motor planning by reducing noise levels (but this is not to suggest that

the basal gnglia is simply devoted to a noise-making function). The idea that under

normal conditions the basal ganglia uses noise for motor planing is similar to the role

proposed for the avian homologe of the basal ganglia, in which residual “noisy”

variability in well learned skills reflects meaningful motor exploration that can support

optimization of performance (Tumer and Brainard, 2007). Indeed, PD patients become

even slower with increased task complexity which requires more motor planning.

Alternatively, it has been proposed that segregation of the functional subcircuits of the

basal ganglia (i.e. de-correlations) is an important aspect of normal BG function, and that

Page 176: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

164

a breakdown of this independent processing could play an important role in the

pathophysiology of PD (Bergman et al., 1998a). As mentioned in the General

Introduction, the specificity of receptive fields in the BG-thalamocortical network is

markedly reduced in PD. The loss of functional segregation within the motor circuit can

be partly due to the increased sensitivity of the BG to oscillatory inputs in the beta and

tremor frequencies. The excess synchronized oscillations might result in co-selection of

antagonist motor programs resulting in rigidity or tremor, depending on the oscillation

frequency. These frequencies (i.e. beta and tremor) are relatively low compared to the

gamma band, and are therefore capable of synchronizing information channels over

larger distances because their synchronization is less susceptible to long conduction

delays (Buzsaki and Draguhn, 2004).

To summarize, these studies indicate that beta oscillatory activity in the STN and GPi

might be involved in the pathological functioning of the basal ganglia in Parkinson‟s

disease and may reflect the degree of striatal and/or extra-striatal dopamine deficiency.

The beta oscillatory LFPs recorded from the STN and GPi may or may not reflect the

degree of synchronized firing among population of local neurons depending on factors

such as the existence of tremor and the amounts of dopamine in the system. In addition,

the balance between oscillatory rhythms may play an important role in the pathological

functioning of the basal ganglia in PD. It seems likely that alterations in both serial and

parallel information processing are involved in the pathological functioning of the BG in

PD.

Page 177: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

165

Page 178: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

166

APPENDIX - RECORDINGS FROM THE PEDUNCULOPONTINE NUCLEUS

The work described in this appendix was published by Weinberger et al. (2008), and is

used with kind permission from Springer Science+Business Media: Experimental Brian

Research, Pedunculopontine Nucleus Microelectrode Recordings in Movement Disorders

Patients, 188(2):165-74, 2008. Weinberger M, Hamani C, Hutchison WD, Moro E,

Lozano AM, Dostrovsky JO, Figures 1-5, Tables 1-2.

Abstract

The pedunculopontine nucleus (PPN) lies within the brainstem reticular formation and is

involved in the motor control of gait and posture. Interest has focused recently on the

PPN as a target for implantation of chronic deep brain stimulation (DBS) electrodes for

Parkinson's disease (PD) and progressive supranuclear palsy (PSP) therapy. The aim of

this study was to examine the neurophysiology of the human PPN region and to identify

neurophysiological landmarks that may aid the proper placement of DBS electrodes in

the nucleus for the treatment of PD and PSP. Neuronal firing and local field potentials

were recorded simultaneously from two independently driven microelectrodes during

stereotactic neurosurgery for implantation of a unilateral DBS electrode in the PPN in

five PD patients and two PSP patients. Within the PPN region, the majority (57%) of the

neurons fired randomly while about 21% of the neurons exhibited 'bursty' firing. In

addition, 21% of the neurons had a long action potential duration and significantly lower

firing rate suggesting they were cholinergic neurons. A change in firing rate produced by

passive and/or active contralateral limb movement was observed in 38% of the neurons

that were tested in the PPN region. Interestingly, oscillatory local field potential activity

in the beta frequency range (at approximately 25 Hz) was also observed in the PPN

region. These electrophysiological characteristics of the PPN region provide further

support for the proposed role of this region in motor control. It remains to be seen to what

extent the physiological characteristics of the neurons and the stimulation-evoked effects

will permit reliable identification of PPN and determination of the optimal target for DBS

therapy.

Page 179: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

167

A.1 - Literature overview

A.1.1 - Functional anatomy of the PPN

The PPN is located in the mesopontine tegmentum and can be divided into two territories

based on cell density and neurochemical characteristics: the pars compacta (PPNc) and

the pars dissipatus (PPNd) (Matsumura, 2005;Pahapill and Lozano, 2000). The PPNc is

located within the caudal half of the nucleus in its dorsolateral aspect and mainly contains

cholinergic neurons (Mesulam et al., 1989). However, some non-cholinergic cells are

also present. The PPNd contains a nearly equal number of cholinergic and non-

cholinergic neurons that are distributed sparsely within the rostral half of the nucleus

(Kus et al., 2003). Non-cholinergic PPN neurons are mostly glutamatergic (Lavoie and

Parent, 1994c), but also noradrenergic, dopaminergic, GABAergic and peptidergic

(Pahapill and Lozano, 2000).

Inputs to the PPN

Inhibitory GABAergic projections from the GPi and the SNr to the PPN are well

established (Noda and Oka, 1986;Granata and Kitai, 1991) and are the most widely

studied projection to the primate PPN. In monkeys, pallidal afferents to the PPN

terminate preferentially on the non-cholinergic neurons of the PPNd (Shink et al., 1997)

and their majority also project to the ventolateral thalamus (Harnois and Filion,

1982;Parent et al., 2001). Projections from the SNr also appear to terminate mainly but

not exclusively on the non-cholinergic neurons of the PPNd in rats (Kang and Kitai,

1990;Spann and Grofova, 1991). It is not known, however, to what extent GPi and SNr

afferents are segregated within the PPN or converge onto the same neurons (Pahapill and

Lozano, 2000).

Cortical inputs from the primary motor cortex innervate the dorsal aspect of the PPN in

non-human primates, with representations of orofacial, forelimb and hindlimb structures

arranged somatotopically from medial to lateral respectively (Matsumura et al., 2000).

Terminals from the supplementary motor areas, as well as dorsal and ventral premotor

Page 180: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

168

cortices overlap this mediolateral arrangement in the middle portion of the nucleus.

Frontal eye field projections are scattered through the PPN (Matsumura et al., 2000). The

cortical afferents to the PPN are likely glutamatergic.

Glutamatergic inputs from the STN to the PPN have been described in rats (Kita and

Kitai, 1987;Hammond et al., 1983;Steininger et al., 1992;Granata and Kitai, 1989) and, to

a small degree, in non-human primates (Smith et al., 1990). However, the different

subpopulations of PPN neurons that receive input from STN have not been established.

Brainstem afferents to the PPN have been described in cats and rodents but not in non-

human primates (Inglis and Winn, 1995).

Outputs from the PPN

The efferent projections of the PPN can be divided into descending and ascending, with

cholinergic and non-cholinergic neurons contributing to both (Lavoie and Parent, 1994b).

Although some PPN neurons have specific ascending projections and others specific

descending projections, some neurons collateralize and project in both directions

(Pahapill and Lozano, 2000).

Ascending cholinergic PPN neurons project to all thalamic nuclei in the rat (Hallanger et

al., 1987;Rye et al., 1987;Sofroniew et al., 1985), cat (Steriade et al., 1988) and monkey

(Lavoie and Parent, 1994b), and in particular to the associative and non-specific midline

and intralaminar nuclei. In addition, ascending PPN projection provide dense innervation

to basal ganglia structures via the ventral tegmental bundle (Lavoie and Parent,

1994a;Lavoie and Parent, 1994b). The most densely innervated structures are the SNc,

the STN and the globus pallidus. PPN projections to the SNc in non-human primates and

rats have been shown to consist of both cholinergic and glutamatergic neurons. Only a

minor contingent of PPN cholinergic neurons project to SNr (Lavoie and Parent,

1994a;Charara et al., 1996;Takakusaki et al., 1996). The STN receives bilateral

projections from the PPN in the rat (Woolf and Butcher, 1986), cat (Edley and Graybiel,

1983) and monkey (Lavoie and Parent, 1994b). In the rat, these connections have been

shown to be both cholinergic (Woolf and Butcher, 1986) and non-cholinergic (Lee et al.,

Page 181: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

169

1988). Other studies in the rat suggest that PPN efferents to the STN and adjacent zona

incerta are collaterals of PPN projections to the globus pallidus (Hammond et al.,

1983;Hallanger and Wainer, 1988). In the monkey, the PPN projections to the pallidal

complex are less dense than these of the STN and SNc, and appear to arborize more

profusely in the GPi than the GPe (Lavoie and Parent, 1994b). The neurotransmitter

involved in the PPN-pallidal connections has not yet been identified with certainty

(Pahapill and Lozano, 2000). Other ascending targets thought to exist in non-human

primates include the striatum, superior colliculus and a number of limbic structures such

as the basal forebrain, hypothalamus and amygdala (Inglis and Winn, 1995;Pahapill and

Lozano, 2000;Reese et al., 1995b).

Although not as numerous as ascending projections, descending PPN efferents travel to

several midbrain, pontine, medullary and spinal cord areas, including several nuclei of

the reticular formation and deep cerebellar nuclei (Inglis and Winn, 1995;Reese et al.,

1995b). These descending projections are thought to collateralize extensively to caudal

structures (Rye et al., 1988) as well as to the thalamus (Semba et al., 1990). In rodents, it

has been shown that most of the descending cholinergic PPN neurons travel a shorter

distance to the medullary reticular formation, which in turn provides bilateral outputs to

the spinal cord. The direct PPN projections to the spinal cord are thought to be mainly

non-cholinergic, although a small number of cholinergic neurons may project as far as the

spinal cord (Rye et al., 1988;Skinner et al., 1990a;Skinner et al., 1990b;Goldsmith and

van der Kooy, 1988).

A.1.2 - Regulation of locomotion and gait in the PPN

The role of the PPN in locomotion is bound up with the fact that the PPN is part of the

so-called mesencephalic locomotor region, a functionally defined area of the brainstem

within which it is possible to elicit controlled locomotion in decerebrate animals by

electrical stimulation (Shik et al., 1969). The effects of stimulation on locomotion and

muscle tone seem to be dependent on the parameters employed. For instance, continuous

stimulation of the mesencephalic locomotor region at frequencies between 20-60 Hz for

Page 182: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

170

several seconds is effective in eliciting locomotor activity, whereas stimulation at

frequencies higher than 100 Hz suppress muscle tone (Garcia-Rill and Skinner,

1987a;Garcia-Rill and Skinner, 1987b;Lai and Siegel, 1990). In the area of the PPN,

there are three separate populations of neurons displaying rhythmic activity in relation to

locomotion in the decerebrate cat (Garcia-Rill and Skinner, 1988). The first group of

neurons displays tonic firing during locomotion, which decreases in frequency or stops

completely with the termination of the locomotor episode. These neurons have been

termed „on‟ cells. The second group of neurons also displays tonic firing, but the firing

rate decreases as the locomotion frequency increases. These neurons are called „off‟ cells.

The third group of neurons displays a bursting pattern of firing during locomotion. It has

been suggested that the „on‟/‟off‟ cells (though to be cholinergic) might modulate the

duration of the stepping episode, while the bursting cells (though to be non-cholinergic)

may be involved in modulating the frequency (and possible initiation) of stepping

(Garcia-Rill and Skinner, 1991).

Although the optimal sites for the stimulus induction of locomotion appear to be within

the cholinergic neuron mass of the PPNc (Garcia-Rill et al., 1987), there are several

brainstem regions, including prominent sensory nuclei, which can be stimulated to

initiate locomotion. Each of these areas project to the spinal cord and none of these can

be considered specific to locomotion. Indeed, bilateral excitotoxic lesions of the whole

PPN have repeatedly failed to affect locomotor activity in rodents (Inglis et al.,

1994a;Inglis et al., 1994b;Olmstead and Franklin, 1994;Steiniger and Kretschmer,

2004;Swerdlow and Koob, 1987). However, unilateral microinjections of the cholinergic

antagonist cabachol into the PPN decreased spontaneous locomotor activity of rats

(Brudzynski et al., 1988). In non human primates, excititoxic lesions of the PPN (with

kainic acid injections) have been shown to produce contralateral hemiparkinsonism

characterized by flexed posture and hypokinesia (Kojima et al., 1997;Munro-Davies et al.,

1999). In addition, radiofrequency lesions in the PPN significantly reduced motor activity

in monkeys, as reflected by generalized bradykinesia that resembled Parkinson‟s disease

(Aziz et al., 1998). Although animals recovered to a near normal condition after

approximately one week following unilateral lesions, bilateral lesions induced

Page 183: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

171

significantly long-lasting effects (Munro-Davies et al., 1999).

In the normal monkey, nearly half of the recorded PPN neurons changed their firing rates

in response to voluntary movement of the ipsilateral and/or contralateral arm (Matsumura

et al., 1997), as well as in response to voluntary guided saccades (Kobayashi et al., 2002).

In the cat it has been shown that PPN neurons change their firing activity during a

conditioned movement of the forelimb (Dormont et al., 1998) and that pharmacological

inactivation of the PPN by injection of muscimol induces prolonged arrest in the

performance of a conditioned motor task (Conde et al., 1998). These findings suggest that

the PPN could influence more widespread somatic movements (Matsumura, 2005).

Interestingly, unilateral electrical stimulation of the PPN in a normal monkey results in

frequency dependent motor effects: low frequencies (<30 Hz) elicit contralateral

proximal

limb tremor while high frequencies (>45 Hz) cause loss of postural control and severe

akinesia (Nandi et al., 2002b).

Neurons in the PPN were also shown to respond to somatosensory stimuli (especially in

contralateral trigeminal areas) (Grunwerg et al., 1992) and auditory inputs (Reese et al.,

1995a). The responsiveness of PPN neurons to somatosensory stimuli, together with the

proposed inputs to the PPN from lamina I of the cat spinal cord (Hylden et al., 1985),

suggests that the PPN may also be involved in the modulation of sensory information to

thalamic nuclei. The prominent cholinergic PPN projection to the thalamus and its

connections with deep cerebellar nuclei further facilitate the potential role of the PPN as a

relay station, providing feedback information important for the modulation of posture and

gait initiation (Pahapill and Lozano, 2000).

Apart from this, the PPN have been considered to play a role in the control of sleep and

to provide a link between the limbic motivation system and the selection of motor output

by the basal ganglia (Winn, 2008).

A.1.3 - PPN and Parkinson’s disease

Page 184: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

172

Neuropathological studies on humans have reported nearly 50% degeneration of the

cholinergic neurons in the PPNc in akinetic disorders such as PD and progressive

supranuclear palsy (PSP) (Hirsch et al., 1987;Jellinger, 1988;Zweig et al., 1989).

Moreover, the degree of akinesia in PD has been linked to the extent of loss of the large

cholinergic PPN neurons (Zweig et al., 1989). The fact that the neuronal loss within the

PPNc is similar in magnitude to that of the SNc suggests that PPN neurons might be

susceptible to similar pathogenetic mechanisms as nigral dopaminergic neurons (Pahapill

and Lozano, 2000).

According to the direct and indirect pathways model, one would predict that in

Parkinson‟s disease there would be both increased inhibitory input to PPN from basal

ganglia output nuclei and increased excitatory input from STN. Indeed, experimental

studies support these predictions. Increased inhibition from the GPi and SNr would lead

to suppression of activity in the PPN, which, because the PPN is thought to facilitate

movement, would in turn produce a state of motor hypoactivity. In agreement with this,

pharmacological blockade of the inhibitory input to PPN with the GABAergic antagonist

bicuculline markedly attenuates akinesia in the MPTP monkeys (Nandi et al., 2002a). By

contrast, PPN neurons also exhibit increased activity in animal models of PD (Orieux et

al., 2000;Breit et al., 2001), which depends on STN (Breit et al., 2001) and thus

presumably relates to the hyperactivity of STN neurons after chronic loss of dopamine. In

anesthetized rats, 6-OHDA lesion of the SNc not only induces a significant increase in

firing rates of PPN neurons (from ~10 Hz to ~18 Hz), but also results in an irregular

firing pattern (6% vs. 27% of the neurons) (Breit et al., 2001).

Of relevance to the clinical practice, it was further shown that some of the

electrophysiological changes recorded in basal ganglia structures after 6-OHDA nigral

injections could be reversed by PPN lesions. These include a decrease in the firing rate of

the STN (from 17 to 11 Hz) and SNr (from 26 to 17 Hz), with no remarkable change in

the firing patterns (Breit et al., 2006). The reversal of STN and SNr hyperactivity in 6-

OHDA-lesioned rats by additional PPN lesion suggests an important modulation of STN

activity by the PPN. Another study in anesthetized rats has investigated the response of

Page 185: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

173

PPN cells to STN stimulation (Florio et al., 2007). When applied in single pulses, STN

stimulation induced excitatory responses in 20% of the PPN neurons recorded, but when

stimulation was provided in 1-5 second trains at 130 Hz, approximately 40% of the

recorded PPN neurons responded. Of those, 85% were inhibited and 15% excited by the

STN stimulation. This pattern of responsiveness (i.e, approximately 20% to single pulses,

and 40% to stimulation trains) was maintained after nigral 6-OHDA injections. In

animals that had entopeduncular lesions however, 75% of PPN neurons became

responsive to STN stimulation. In addition, a higher proportion of cells responded with

excitation (85%) vs. inhibition (15%) following entopeduncular lesions in animals that

had nigral saline injections or 6-OHDA. These results suggest that the inhibition recorded

in PPN cells after trains of STN stimulation is likely mediated by entopeduncular neurons

(Florio et al., 2007).

In MPTP monkeys, microinjection of the GABAergic antagonist bicuculline (but not

saline or muscimol) into the PPN, was able to improve parkinsonian symptoms (Nandi et

al., 2002a). In addition, when MPTP was systematically injected in monkeys with prior

PPN lesioning, the monkeys developed no or very mild parkinsonism relative to MPTP

monkeys without a previous PPN lesion (Matsumura and Kojima, 2001). In a later study,

Jenkinson and colleagues have studied the effects of unilateral PPN DBS prior to and

after inducing MPTP parkinsonism in a macaque monkey (Jenkinson et al., 2004). In the

normal monkey, PPN stimulation at 5 Hz increased movement counts (measured per

hour), whereas 100-Hz stimulation decreased the same measures. After making the

monkey parkinsonian with MPTP, low-frequency PPN stimulation was as effective as

levodopa, but high-frequency DBS was not.

Taken together, these findings have suggested an important role for the PPN in

generating parkinsonian symptoms and opened up new possibilities in the management of

advanced PD.

A.1.4 - PPN as a surgical targted for PD

Page 186: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

174

The first report on deep brain stimulation in the human PPN involved two PD patients

who displayed dominant symptoms of freezing of gait and postural instability before

surgery (Plaha and Gill, 2005). Postoperatively short-term clinical improvement was

observed with stimulation at 20–25 Hz. Total UPDRS improved 53% and the motor

subscore improved 57% when averaged between the two patients. High-frequency

stimulation actually worsened gait in one patient. In a different report, Mazzone and

colleagues described intraoperative PPN DBS in two PD patients (Mazzone et al., 2005).

In this study, low-frequency stimulation at 10 Hz was associated with a feeling of

„wellbeing‟ that correlated with modest improvements in motor function (bradykinesia

and rigidity). The effects of posture and gait, however, were not reported in this initial

investigation.

In a later study by the same group, the outcomes of bilateral PPN DBS, with or without

simultaneous stimulation in the STN, were investigated in 6 patients with severe

symptoms including disabling dyskinesia and postural instability (Stefani et al., 2007).

PPN DBS was given at 25Hz (60 µsec pulse width) and 1.5-2 volts in a bipolar

configuration. At 6 months post-surgery, UPDRS motor scores in the „off‟ medication

condition were improved by 33% with PPN DBS. This improvement, however, was

significantly less than the benefits obtained with STN DBS (54%) or STN + PPN DBS

(56%). In contrast, improvements in axial symptoms such as rising from a chair, posture,

gait, and postural stability were similar with PPN, STN and PPN +STN DBS (60-70%).

In the „on‟ medication condition, UPDRS motor scores were significantly improved by

either PPN (44.3%) or STN DBS (51%), as compared to the „on‟ medication „off‟

stimulation condition. A significantly better outcome has been achieved with STN + PPN

DBS (66.4%), showing that the activation of both electrodes was of value when the

patients were taking their medicine. When only the axial symptoms were taken into

account, the use of medication and either PPN DBS or PPN+ STN DBS led to a 50-60%

improvement, as compared to the „on‟ medication „off‟ stimulation condition. STN

stimulation alone did not improve axial symptoms when the patients were taking their

medication. Importantly, no major complications were described in either series. The

only inconvenient side effect reported with PPN stimulation at high frequency was

Page 187: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

175

paresthesia, likely due to the proximity of the target to the medial lemniscus (Stefani et

al., 2007).

A.2 - Introduction to the study

Although STN deep brain stimulation (DBS) to treat advanced Parkinson‟s disease (PD)

is a well established procedure and produces striking clinical improvement in motor

symptoms such as tremor, rigidity and bradykinesia, it has only a moderate and a short

term effect in improving axial symptoms like gait akinesia and postural instability

(Kleiner-Fisman et al., 2003). Recently, the pedunculopontine tegmental nucleus (PPN)

has been proposed (Pahapill and Lozano, 2000) and examined as an alternative

therapeutic target for PD, especially for patients with severe gait and postural impairment

(Mazzone et al., 2005;Plaha and Gill, 2005;Stefani et al., 2007).

The PPN is part of the brainstem reticular formation and projects to the thalamus and

spinal cord. It is also reciprocally connected to the basal ganglia and is believed to

regulate the control of complex central nervous system functions including posture and

gait (for review, see Pahapill and Lozano, 2000). In primates, the PPN receives a major

projection from the two output structures of the basal ganglia (substantia nigra pars

reticulata and internal globus pallidus - GPi) and the subthalamic nucleus (STN). In turn,

the PPN projects to the substantia nigra pars compacta and STN (Garcia-Rill,

1991;Pahapill and Lozano, 2000;Winn, 2006). This organization emphasizes the potential

role of the PPN in controlling basal ganglia output through subsidiary loops.

It has been shown that neurons in the PPN region degenerate in akinetic disorders such as

PD and PSP (Hirsch et al., 1987;Jellinger, 1988;Zweig et al., 1989), and lesions in the

PPN can lead to decreased movements (Aziz et al., 1998;Kojima et al., 1997). Therefore,

it has been suggested that the PPN may be an alternative target for DBS for the treatment

of PD (Jenkinson et al., 2004;Matsumura, 2005;Pahapill and Lozano, 2000;Plaha and Gill,

2005). Indeed, recent studies have shown that low frequency stimulation (<25 Hz) of the

PPN increases motor activity in a monkey model of PD (Jenkinson et al., 2004) and

Page 188: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

176

improves clinical motor scores in PD patients (Mazzone et al., 2005;Plaha and Gill,

2005;Stefani et al., 2007).

The PPN occupies an area in the mesopontine tegmentum that contains two populations

of neurons, non-cholinergic (likely glutamatergic) and cholinergic (Mesulam et al., 1989).

The cholinergic neurons have larger somata, but are less numerous than the non-

cholinergic neurons (Honda and Semba, 1995;Rye et al., 1987;Spann and Grofova, 1992).

Electrophysiological studies of the PPN in non-human primates have rarely appeared in

the literature (Matsumura, 2005). In the PPN of the normal monkey, nearly half of the

neurons recorded changed their firing rates in response to voluntary movement of the

ipsilateral and/or contralateral arm (Matsumura et al., 1997). In the cat it has been shown

that PPN neurons change their firing activity during a conditioned movement of the

forelimb (Dormont et al., 1998) and that inactivation of the PPN by injection of muscimol

(a γ-aminobutyric acid (GABA) agonist) induces prolonged arrest in the performance of

the task (Conde et al., 1998). These findings suggest that the PPN is involved in the

initiation and modulation of movements (for review, see Matsumura, 2005).

The aim of the present study was to describe the neurophysiology of the human PPN

region and identify neurophysiological landmarks and characteristics that may help in

localizing the target for insertion of DBS electrodes in the nucleus for the treatment of

PD and PSP.

A.3 - Methods

Patients

We studied seven patients, five with advanced PD and two with PSP, who were

undergoing stereotactic surgery for implantation of unilateral PPN-DBS electrodes under

local anesthesia. The group consisted of two women and five men who, at the time of

operation, had a mean age (±SD) of 64 ± 6.7 years and mean disease duration of

12.4 ± 5.1 years. During surgery the patients were awake and „off‟ dopaminergic

medications for at least 12 h from the last oral dose of medication. Patient #7 had

Page 189: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

177

bilateral STN-DBS electrodes that were implanted one year before the surgery and were

switched off during the recordings. Demographic details of the patients are given in

Table A.T1.

Surgical procedure

On the morning of the surgery, a stereotactic frame (Leksell G, Elekta, Inc, Atlanta, Ga)

was affixed to the patients‟ heads and preoperative magnetic resonance (MR) images

were obtained (Signa, 1.5 T, General Electric, Milwaukee, Wis). Once the MRI based

coordinates of the target were calculated, the patient was brought to the operating room

and a burr hole was drilled 2 cm from the midline in front of the coronal suture.

Microelectrode recordings were then conducted from 15 mm above the target and

extended 2–5 mm below. In each patient except one, only one dual electrode track,

consisting of two microelectrodes, was performed (see General Methods, section 2.2.3

for details). In patient #4 a second electrode track was performed 2 mm medial to the first

since the thresholds for evoking paresthesia in the first track were very low. The

coordinates chosen for electrophysiological recordings were similar across patients.

Microelectrode recordings and stimulation

Neuronal spontaneous firing activity and local field potentials (LFPs) were recorded

simultaneously from two independently driven microelectrodes. For further details of the

procedure see General Methods (section 2.2.3/4). In all patients, we tested the effects of a

1 s microstimulation train (200 Hz, 150 μs pulse width at intensities up to 100 μA) along

Page 190: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

178

the microelectrode trajectory at intervals of 2 mm or less.

Localization of electrode trajectory

The approximate location of the microelectrode tracks was estimated post-operatively (by

Dr. Clement Hamani) as follows: Preoperative axial stereotactic 3D inversion recovery

and T2-weighted MR images were transferred to a StealthStation workstation. Using the

FrameLink 4.1 software (Mach 4.1, StealthStation, Medtronic, SNT) these two images

were merged, the fiducials of the frame were recognized, and the mean rod marking error

was calculated and registered. Coronal and sagittal planes were reconstructed based on

axial images. The anterior and posterior commissures (AC and PC) were then targeted in

the axial plane and three additional points were plotted in the midline. Thereafter, the

images were reformatted parallel to the AC–PC plane and orthogonal to the midline.

Pitch, roll, and yaw were corrected in the StealthStation. The coordinate of the

microelectrode track and the angles of each trajectory were entered into the

neuronavigation system so that they could be reconstructed. As the PPN cannot be clearly

visualized on imaging studies, we estimated the location of the region that most likely

encompassed the nucleus. We defined this PPN region as the anterolateral portion of the

pontine/mesencephalic tegmentum that extended from 2 mm below the inferior colliculus

to the transition between the inferior and superior colliculi, according to the Paxinos and

Huang atlas (Paxinos G and Huang X.F., 1995).

Data analysis

Only sites with good signal-to-noise and stable single unit recordings were analyzed. To

characterize the firing activity of the recorded neurons the signals were bandpass filtered

at 425–5,000 Hz and single unit activity was discriminated using the template matching

tool in Spike2 (CED). As a measure of action potential width we used the duration of the

negative phase of the averaged spike waveform (see Figure A.F1A). A similar method

was previously used to measure spike duration of PPN neurons (Matsumura et al.,

1997;Takakusaki et al., 1997).

Page 191: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

179

Mean firing rates and measurement of firing patterns were performed in MATLAB

(version 6.5, The Math Works, Natick, MA) using the Kaneoke and Vitek method

(Kaneoke and Vitek, 1996), which uses discharge density to categorize firing patterns

into random, bursty or regular. In this method, the discharge density histogram of a

neuron is determined by calculating the number of spikes in an interval equal to the mean

ISI. The number of occurrences of no spikes, one spike, two spikes, and so on in each

time interval is then counted and a discharge density histogram is constructed. This

discharge density histogram represents the probability distribution of the neuron‟s

discharge density and can be compared to a discharge density of a Poisson process with a

mean of 1 by using a 2 goodness-of-fit test. If the neuronal discharge pattern was

random, its discharge density distribution would be statistically similar to that of a

Poisson distribution. If the neuron‟s discharge pattern was significantly non-random it

could be either regular or bursty. In cases of regular firing patterns, the probability of

finding one spike per time segment is high. On the other hand, in cases of bursting firing

patterns, the probability of finding no spikes or many spikes per time segment is high.

Since a Poisson process with a mean of one has a variance equal to one, the former case

represents a significantly non-Poisson discharge density distribution with a variance of

less than one and the latter case represents a significantly non-Poisson discharge density

distribution with a variance of greater than one.

Page 192: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

180

For examining neuronal responses to movement, movement onset was determined from

accelerometer recordings and/or electromyograms and a peri-stimulus histogram

constructed. Movement-related changes in neuronal firing were initially identified by

visual inspection of the histograms. Then, the identified responses were evaluated

statistically using nonparametric tests comparing the firing rate during a 200 or 300 ms

control period and during 200/300 ms of the visually determined response (P < 0.05,

Wilcoxon signed rank test).

For spectral analysis, recordings of ≥11 second duration (mean ± SD: 37.9 ± 19.7 s, range:

11–112 s) were imported into MATLAB. Power spectra and coherence of neuronal

discharges and LFP recordings was performed according to Halliday et al. (1995) and

Rivlin-Etzion et al. (2006), and is described in detail in the General Methods (section

2.3.1.a/b).

A.4 - Results

A.4.1 - Action potentials and firing properties

A total of 244 neurons was recorded in seven patients. Of these, 235 were recorded while

the patient was at rest and were used to characterize the baseline firing activity. Neurons

were classified according to the width and polarity of their action potentials. The

incidence histogram of spike durations did not reveal a clear bimodal distribution.

Nevertheless, we observed a marked drop in the number of neurons that had action

potential durations ≥0.65 ms (see Figure A.F1B) which suggests they may have

originated from a different population of neurons that are likely to be composed of

cholinergic neurons (see Discussion). This population consisted of 38 neurons that

comprised 15.6% of the total number of neurons. These neurons were found in all the

patients except one (patient #3, PSP) in whom a total of only nine neurons were recorded

(see Table A.T1). Their mean (±SD) spike duration was 0.84 ± 0.2 ms compared with

0.47 ± 0.1 ms for the neurons with brief spikes. In addition, these neurons had a

significantly lower mean firing rate (medians: 9.3 vs. 17.3 spikes/s, P ≤ 0.001, Mann–

Page 193: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

181

Whitney rank sum test) (Figure A.F2B, D). This provides further support for the

assumption that these cells represent a different neuronal population. The majority of

these neurons (86.8%) was found within and below the PPN region. In the region of the

PPN, these neurons comprised 21% of the neurons (18/85) compared to 19.5% below that

region (15/77) and 6% above (5/82).

Interestingly, 16 units had positive-going action potentials and significantly higher mean

firing rates (median 67.6 spikes/s, P ≤ 0.001) (Figure A.F2C, D). These units were found

in all seven patients. Six of these neurons were located in the PPN region and nine were

located below that region. In addition, most (88%) of these neurons displayed a regular

firing pattern as determined by the Kaneoke and Vitek method (Figure A.F3A). This was

significantly different from the patterns of activity exhibited by the two types of neurons

described above which were primarily random (65 and 68% for the neurons with short

and long spike durations respectively) (P ≤ 0.001, χ2-test). Comparison of the proportions

of both brief-and broad-spike duration neurons exhibiting regular, random or bursty firing

Page 194: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

182

patterns above, within and below the PPN region, revealed a statistically significant

difference (P = 0.016, χ2-test) (Figure A.F3B). In the PPN region, the majority (57%) of

the neurons fired randomly while about 21% of the neurons exhibited bursty firing.

Above and below the PPN region, the proportion of bursting neurons was smaller (11 and

19% respectively) while more neurons exhibited random firing (65 and 75% respectively).

A.4.2 - Movement responses

The activity of 103 neurons was recorded during contralateral limb movements. Tasks

varied between passive and/or voluntary movements of the wrist, hand, foot, elbow and

sometimes shoulder. 36% (n = 37) of the neurons responded to at least one type of

passive or voluntary movement of the limb. Out of the 41 neurons that were tested only

for passive movements, 16 responded (39%) and out of the 50 neurons that were tested

only for voluntary movements, 12 responded (24%). Twelve neurons were tested for both

passive and voluntary movement, out of which, five neurons responded to passive

movement only, two neurons responded to voluntary movement only and two neurons

responded to both. Table A.T2 summarizes the results for each of the neuron types.

Page 195: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

183

The changes in firing rates were mostly excitatory (about 80%) and in some cases, were

followed by inhibition (see Figure A.4A, C). Within the PPN region 38% of the neurons

responded to movement and similar percentages were recorded above and below the PPN

region where 37 and 33% of the neurons responded respectively. Movement responsive

neurons included all three neuron types. Figure A.F4 shows an example of responses to

contralateral movement from each neuron type.

Page 196: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

184

Page 197: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

185

A.4.3 - Microstimulation in the PPN region

In all patients, stimulus trains of 1 second evoked paresthesia at some sites along the

microelectrode trajectory. In three patients, the threshold intensities for stimulation-

evoked paresthesia varied between 50 and 100 μA (with three exceptions of 20, 25 and

25 μA) (mean ± SD: 79.2 ± 29.6 μA, n = 24). In the remaining four patients, the threshold

intensities were lower and varied between 1 and 50 μA (with two exceptions of 100 and

75 μA) (mean ± SD: 18.8 ± 19.9 μA, n = 55). The thresholds were not significantly

different when stimulating within or close to the PPN region. These responses were

assumed to be due to activation of ascending axons in the medial lemniscus. In addition,

eye movements were evoked by microstimulation in two patients and occurred only at the

maximum stimulation intensity of 100 μA, and only within and up to 2 mm above the

PPN region

A.4.4 - LFP oscillatory activity in the PPN region

Spectral analyses of the local field potentials revealed oscillatory activity in the beta

frequency range (15–30 Hz) in the PPN region of three PD patients (#1, 4 and 7). In all

three patients, a significant coherence in the beta range between the two microelectrodes

was also observed. It was not possible to adequately assess the existence of LFP

oscillatory activity in the other four patients due to suboptimal slow wave recordings in

those cases. Figure A.F5A demonstrates LFP beta oscillations recorded from the two

microelectrodes (in patient #1), and their significant coherence. In these patients, the LFP

power in the beta frequency range varied with depth along the microelectrode track. Beta

oscillations were most prominent in the region of the PPN (Figure A.F5B) and up to

4 mm below. Frequencies between 35 and 50 Hz were sometimes encountered in the LFP

but the coherence in these frequencies rarely reached significance (this is notable in

Figure A.F5A). Similarly, oscillations at frequencies lower than 15 Hz, although

occasionally observed, were not a consistent feature in the PPN region. Spectral analyses

of the neuronal firing in the PPN region failed to show significant oscillatory activity in

the beta frequencies. Moreover, coherence and cross-correlation analyses did not reveal a

Page 198: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

186

significant relationship to the simultaneously recorded LFP.

Page 199: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

187

A.4.5 - Comparison of PD vs. PSP

No obvious differences were noted in any of the findings between the two groups of

patients other than a lower number of neurons encountered in the PSP patients compared

to all but one of the PD patients.

A.5 - Discussion

The present study provides new data documenting various electrophysiological

characteristics of the PPN region in PD and PSP patients. The lower number of cells

recorded from the PPN region in the two PSP patients (see Table A.T1) suggests a lower

density of neurons which might reflect the large degree of neuronal degeneration and

brainstem atrophy that occurs in PSP (Dickson et al., 2007;Hirsch et al., 1987). We did

not observe any obvious neurophysiological differences between the two patient groups.

However, in view of the small number of neurons in the 2 PSP patients this comparison

may not be very meaningful.

Analysis of spike duration and shape revealed three types of neurons. The existence of

short and long duration negative going action potentials in this region agrees with the

findings of previous in vitro and in vivo animal studies. In vivo recordings from the PPN

in rats (Takakusaki et al., 1997), cats (Dormont et al., 1998) and monkeys (Matsumura et

al., 1997) revealed the existence of two types of neurons: ones which fire at low rate and

exhibit a long duration spike, and ones which fire at higher rates and exhibit a short

duration spike. Yet, the incidence histogram of spike durations in our study did not reveal

a clear bimodal distribution as previously demonstrated in the rat PPN (Takakusaki et al.,

1997). That study, however, utilized intracellular recordings and therefore cannot be

directly compared to our extracellular recordings. Moreover, the incidence histogram in

our study is consistent with data from extracellular recordings in the monkey PPN that

failed to exhibit a bimodal distribution (Matsumura et al., 1997).

It has been proposed on the basis of in vitro studies that the neurons with broad spikes are

Page 200: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

188

cholinergic (Dormont et al., 1998;Takakusaki et al., 1997). Although the exact borders of

the PPN remain indistinct, the cholinergic neurons are a very obvious component of the

PPN (Winn, 2006). Indeed, the neurons with broad spikes in our study were found

primarily within the region of the PPN and below it. The region below the PPN is likely

to be the adjacent laterodorsal tegmental nucleus which is known to contain cholinergic

neurons that are interconnected with the cholinergic population in the PPN (Winn, 2006).

These findings, together with the low rate of firing, provide further support for the

assumption that these neurons are cholinergic. In our study, the majority of the neurons in

the PPN region exhibited random firing while about 21% of the neurons had a bursty

firing pattern. This finding is supported by earlier studies in rats (Scarnati et al., 1987),

cats (Garcia-Rill et al., 2004) and monkeys (Matsumura et al., 1997) which describe a

small population of bursty PPN neurons, whereas the majority of the neurons have been

reported to fire in an irregular pattern.

The positive-going action potentials may possibly have originated from large diameter

axons of passage as extracellularly recorded positive going potentials are generally

assumed to be axonal. They had a significantly higher mean firing rate and a regular

firing pattern. These electrophysiological properties are different than those of the PPN

neurons previously described and are likely to arise from afferents to the PPN region,

possibly ascending sensory axons of the medial lemniscus which lies laterally to the PPN

(Olszewsky and Baxter, 1982).

Previous electrophysiological studies in cats and monkeys have identified PPN neurons

that exhibit changes in firing rate in response to limb movement and include neurons with

both brief and broad spikes (Dormont et al., 1998;Matsumura et al., 1997). In monkey,

PPN neurons have also been found to respond with either an increase or a decrease in

firing rate to voluntary saccades (Kobayashi et al., 2002). Mazzone et al. (2005) reported

finding 6 neurons (of a total of 27 studied) in the PPN of two PD patients that responded

with a “slight inhibition or inhibition followed by excitation” to a single-joint movement,

but no further details were provided. It is important to note that the PPN region targeted

in the current study (PPTg according to Paxinos and Huang 1995) may be different from

Page 201: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

189

the region studied by Mazzone et al. (see also (Yelnik, 2007;Zrinzo et al., 2007a;Zrinzo

et al., 2007b)). Our study clearly demonstrates the existence of movement-responsive

neurons in the human PPN, consistent with the findings of animal studies. Moreover, we

found that both short and long spike duration neurons responded to movements. It has

been reported that PPN inputs to the basal ganglia are mediated by both cholinergic and

glutamatergic synapses (Lavoie and Parent, 1994a). Taken together, these findings

suggest that both groups of neurons may contribute to basal ganglia-related motor

functions.

In the three PD patients analyzed, the LFPs in the PPN region exhibited oscillatory

activity in the beta range (15-30 Hz). Similar oscillations have been previously observed

in the STN and GPi of PD patients (Brown et al., 2001;Kuhn et al., 2005;Levy et al.,

2002a;Weinberger et al., 2006) and are suggested to play an „antikinetic‟ role in PD

(Brown, 2003;Brown and Williams, 2005). It is important to note, however, that the

neuronal firing did not show oscillations in these frequencies. The absence of oscillatory

firing may be due to limited sampling or may reflect a lack of coupling between dendritic

membrane potential oscillations and the soma. Since the beta oscillations were maximal

in the PPN region they are unlikely to reflect far field potentials from distant structures

(e.g. STN, cortex). The elevation of the beta frequency band LFP activity around the

region of the PPN suggests that this activity may be useful in determining the final

positioning of the DBS electrode intra-operatively, but further studies are required to

examine the consistency of this feature. In a recent study by Androulidakis et al. (2008),

LFP recordings from DBS electrodes in the PPN region in six PD patients revealed 7–

11 Hz oscillations that were prominent only after treatment with levodopa, suggesting

that these oscillations may represent a physiological pattern of activity in this region

(Androulidakis et al., 2008b;Androulidakis et al., 2008c). However, inconsistent with our

finding, that study did not reveal beta oscillations in the „off‟ dopaminergic state. This

discrepancy might be due to the difference in the region targeted in that study, which is

similar to that reported by Mazzone et al. (see above).

In summary, the electrophysiological characteristics of the PPN region described in this

Page 202: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

190

paper provide further support for the proposed role of this region in motor control.

However, in contrast to microelectrode recordings in thalamus, pallidum and STN, we

did not find distinctive and definitive neurophysiological characteristics and clear nuclear

boundaries for PPN. This is consistent with the absence of precise anatomical borders to

the PPN (Winn, 2006), which might reflect its anatomical complexity. Thus

microelectrode recordings and microstimulation-induced effects in PPN, although helpful,

do not provide the same degree of confirmation of the target as they do in these other

regions although future studies may reveal additional features distinctive for the PPN

which would be useful in localizing the region and identifying an optimal target for DBS

therapy.

Page 203: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

191

Reference List

Aarsland D, Bronnick K, Ehrt U, De Deyn PP, Tekin S, Emre M, Cummings JL (2007)

Neuropsychiatric symptoms in patients with Parkinson's disease and dementia: frequency,

profile and associated care giver stress. J Neurol Neurosurg Psychiatry 78:36-42.

Abosch A, Hutchison WD, Saint-Cyr JA, Dostrovsky JO, Lozano AM (2002) Movement-

related neurons of the subthalamic nucleus in patients with Parkinson disease. J

Neurosurg 97:1167-1172.

Aebischer P, Schultz W (1984) The activity of pars compacta neurons of the monkey

substantia nigra is depressed by apomorphine. Neurosci Lett 50:25-29.

Afsharpour S (1985) Light microscopic analysis of Golgi-impregnated rat subthalamic

neurons. J Comp Neurol 236:1-13.

Ahlskog JE (2003) Parkinson's disease: is the initial treatment established? Curr Neurol

Neurosci Rep 3:289-295.

Aizman O, Brismar H, Uhlen P, Zettergren E, Levey AI, Forssberg H, Greengard P,

Aperia A (2000) Anatomical and physiological evidence for D1 and D2 dopamine

receptor colocalization in neostriatal neurons. Nat Neurosci 3:226-230.

Akkal D, Burbaud P, Audin J, Bioulac B (1996) Responses of substantia nigra pars

reticulata neurons to intrastriatal D1 and D2 dopaminergic agonist injections in the rat.

Neurosci Lett 213:66-70.

Alberts WW (1972) A simple view of Parkinsonian tremor. Electrical stimulation of

cortex adjacent to the Rolandic fissure in awake man. Brain Res 44:357-369.

Alberts WW, Libet B, Wright EW, Jr., Feinstein B (1965) Physiological mechanisms of

tremor and rigidity in parkinsonism. Confin Neurol 26:318-327.

Alberts WW, Wright EW, Jr., Feinstein B (1969) Cortical potentials and Parkinsonian

tremor. Nature 221:670-672.

Albin RL, Young AB, Penney JB (1989) The functional anatomy of basal ganglia

disorders. Trends Neurosci 12:366-375.

Aldridge JW, Anderson RJ, Murphy JT (1980) Sensory-motor processing in the caudate

nucleus and globus pallidus: a single-unit study in behaving primates. Can J Physiol

Pharmacol 58:1192-1201.

Alegre M, Labarga A, Gurtubay IG, Iriarte J, Malanda A, Artieda J (2002) Beta

electroencephalograph changes during passive movements: sensory afferences contribute

to beta event-related desynchronization in humans. Neurosci Lett 331:29-32.

Page 204: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

192

Alegre M, onso-Frech F, Rodriguez-Oroz MC, Guridi J, Zamarbide I, Valencia M,

Manrique M, Obeso JA, Artieda J (2005) Movement-related changes in oscillatory

activity in the human subthalamic nucleus: ipsilateral vs. contralateral movements. Eur J

Neurosci 22:2315-2324.

Alexander GE, Crutcher MD (1990) Functional architecture of basal ganglia circuits:

neural substrates of parallel processing. Trends Neurosci 13:266-271.

Alexander GE, Crutcher MD, DeLong MR (1990) Basal ganglia-thalamocortical circuits:

parallel substrates for motor, oculomotor, "prefrontal" and "limbic" functions. Prog Brain

Res 85:119-146.

Alexander GE, DeLong MR (1985a) Microstimulation of the primate neostriatum. I.

Physiological properties of striatal microexcitable zones. J Neurophysiol 53:1401-1416.

Alexander GE, DeLong MR (1985b) Microstimulation of the primate neostriatum. II.

Somatotopic organization of striatal microexcitable zones and their relation to neuronal

response properties. J Neurophysiol 53:1417-1430.

Alexander GE, DeLong MR, Strick PL (1986) Parallel organization of functionally

segregated circuits linking basal ganglia and cortex. Annu Rev Neurosci 9:357-381.

Alonso-Frech F, Zamarbide I, Alegre M, Rodriguez-Oroz MC, Guridi J, Manrique M,

Valencia M, Artieda J, Obeso JA (2006) Slow oscillatory activity and levodopa-induced

dyskinesias in Parkinson's disease. Brain 129:1748-1757.

Alvarez L, Macias R, Guridi J, Lopez G, Alvarez E, Maragoto C, Teijeiro J, Torres A,

Pavon N, Rodriguez-Oroz MC, Ochoa L, Hetherington H, Juncos J, DeLong MR, Obeso

JA (2001) Dorsal subthalamotomy for Parkinson's disease. Mov Disord 16:72-78.

Amalric M, Conde H, Dormont JF, Farin D, Schmied A (1984) Activity of caudate

nucleus neurons in cat performing a reaction time task. Neurosci Lett 49:253-258.

Amirnovin R, Williams ZM, Cosgrove GR, Eskandar EN (2004) Visually guided

movements suppress subthalamic oscillations in Parkinson's disease patients. J Neurosci

24:11302-11306.

Amtage F, Henschel K, Schelter B, Vesper J, Timmer J, Lucking CH, Hellwig B (2008)

Tremor-correlated neuronal activity in the subthalamic nucleus of Parkinsonian patients.

Neurosci Lett 442:195-199.

Anderson ME, Horak FB (1985) Influence of the globus pallidus on arm movements in

monkeys. III. Timing of movement-related information. J Neurophysiol 54:433-448.

Anderson ME, Postupna N, Ruffo M (2003) Effects of high-frequency stimulation in the

internal globus pallidus on the activity of thalamic neurons in the awake monkey. J

Neurophysiol 89:1150-1160.

Page 205: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

193

Anderson ME, Turner RS (1991) A quantitative analysis of pallidal discharge during

targeted reaching movement in the monkey. Exp Brain Res 86:623-632.

Androulidakis AG, Brucke C, Kempf F, Kupsch A, Aziz T, Ashkan K, Kuhn AA, Brown

P (2008a) Amplitude modulation of oscillatory activity in the subthalamic nucleus during

movement. Eur J Neurosci.

Androulidakis AG, Khan S, Litvak V, Pleydell-Pearce CW, Brown P, Gill SS (2008b)

Local field potential recordings from the pedunculopontine nucleus in a Parkinsonian

patient. Neuroreport 19:59-62.

Androulidakis AG, Kuhn AA, Chen CC, Blomstedt P, Kempf F, Kupsch A, Schneider

GH, Doyle L, Dowsey-Limousin P, Hariz MI, Brown P (2007) Dopaminergic therapy

promotes lateralized motor activity in the subthalamic area in Parkinson's disease. Brain

130:457-468.

Androulidakis AG, Mazzone P, Litvak V, Penny W, Dileone M, Gaynor LM, Tisch S, Di

L, V, Brown P (2008c) Oscillatory activity in the pedunculopontine area of patients with

Parkinson's disease. Exp Neurol 211:59-66.

Antonini A, Moeller JR, Nakamura T, Spetsieris P, Dhawan V, Eidelberg D (1998) The

metabolic anatomy of tremor in Parkinson's disease. Neurology 51:803-810.

Aosaki T, Graybiel AM, Kimura M (1994) Effect of the nigrostriatal dopamine system on

acquired neural responses in the striatum of behaving monkeys. Science 265:412-415.

Aosaki T, Kimura M, Graybiel AM (1995) Temporal and spatial characteristics of

tonically active neurons of the primate's striatum. J Neurophysiol 73:1234-1252.

Aosaki T, Kiuchi K, Kawaguchi Y (1998) Dopamine D1-like receptor activation excites

rat striatal large aspiny neurons in vitro. J Neurosci 18:5180-5190.

Apicella P, Ljungberg T, Scarnati E, Schultz W (1991) Responses to reward in monkey

dorsal and ventral striatum. Exp Brain Res 85:491-500.

Appell PP, Behan M (1990) Sources of subcortical GABAergic projections to the

superior colliculus in the cat. J Comp Neurol 302:143-158.

Arbuthnott GW, Ingham CA, Wickens JR (1998) Modulation by dopamine of rat

corticostriatal input. Adv Pharmacol 42:733-736.

Arnulf I, Konofal E, Merino-Andreu M, Houeto JL, Mesnage V, Welter ML, Lacomblez

L, Golmard JL, Derenne JP, Agid Y (2002) Parkinson's disease and sleepiness: an

integral part of PD. Neurology 58:1019-1024.

Aubert I, Ghorayeb I, Normand E, Bloch B (2000) Phenotypical characterization of the

neurons expressing the D1 and D2 dopamine receptors in the monkey striatum. J Comp

Neurol 418:22-32.

Page 206: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

194

Aymerich MS, Barroso-Chinea P, Perez-Manso M, Munoz-Patino AM, Moreno-Igoa M,

Gonzalez-Hernandez T, Lanciego JL (2006) Consequences of unilateral nigrostriatal

denervation on the thalamostriatal pathway in rats. Eur J Neurosci 23:2099-2108.

Aziz TZ, Davies L, Stein J, France S (1998) The role of descending basal ganglia

connections to the brain stem in parkinsonian akinesia. Br J Neurosurg 12:245-249.

Aziz TZ, Peggs D, Sambrook MA, Crossman AR (1991) Lesion of the subthalamic

nucleus for the alleviation of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-

induced parkinsonism in the primate. Mov Disord 6:288-292.

Bakay RA, DeLong MR, Vitek JL (1992) Posteroventral pallidotomy for Parkinson's

disease. J Neurosurg 77:487-488.

Bamford NS, Zhang H, Schmitz Y, Wu NP, Cepeda C, Levine MS, Schmauss C,

Zakharenko SS, Zablow L, Sulzer D (2004) Heterosynaptic dopamine neurotransmission

selects sets of corticostriatal terminals. Neuron 42:653-663.

Bankiewicz KS, Oldfield EH, Chiueh CC, Doppman JL, Jacobowitz DM, Kopin IJ (1986)

Hemiparkinsonism in monkeys after unilateral internal carotid artery infusion of 1-

methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Life Sci 39:7-16.

Bar-Gad I, Elias S, Vaadia E, Bergman H (2004) Complex locking rather than complete

cessation of neuronal activity in the globus pallidus of a 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine-treated primate in response to pallidal microstimulation. J Neurosci

24:7410-7419.

Bar-Gad I, Heimer G, Ritov Y, Bergman H (2003) Functional correlations between

neighboring neurons in the primate globus pallidus are weak or nonexistent. J Neurosci

23:4012-4016.

Barlas O, Hanagasi HA, Imer M, Sahin HA, Sencer S, Emre M (2001) Do unilateral

ablative lesions of the subthalamic nucleu in parkinsonian patients lead to hemiballism?

Mov Disord 16:306-310.

Baron MS, Vitek JL, Bakay RA, Green J, Kaneoke Y, Hashimoto T, Turner RS,

Woodard JL, Cole SA, McDonald WM, DeLong MR (1996) Treatment of advanced

Parkinson's disease by posterior GPi pallidotomy: 1-year results of a pilot study. Ann

Neurol 40:355-366.

Basso MA, Wurtz RH (2002) Neuronal activity in substantia nigra pars reticulata during

target selection. J Neurosci 22:1883-1894.

Baufreton J, Atherton JF, Surmeier DJ, Bevan MD (2005a) Enhancement of excitatory

synaptic integration by GABAergic inhibition in the subthalamic nucleus. J Neurosci

25:8505-8517.

Baufreton J, Zhu ZT, Garret M, Bioulac B, Johnson SW, Taupignon AI (2005b)

Page 207: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

195

Dopamine receptors set the pattern of activity generated in subthalamic neurons. FASEB

J 19:1771-1777.

Bayer HM, Handel A, Glimcher PW (2004) Eye position and memory saccade related

responses in substantia nigra pars reticulata. Exp Brain Res 154:428-441.

Beckley DJ, Bloem BR, Remler MP (1993) Impaired scaling of long latency postural

reflexes in patients with Parkinson's disease. Electroencephalogr Clin Neurophysiol

89:22-28.

Beckley DJ, Bloem BR, Van Dijk JG, Roos RA, Remler MP (1991) Electrophysiological

correlates of postural instability in Parkinson's disease. Electroencephalogr Clin

Neurophysiol 81:263-268.

Beckley DJ, Panzer VP, Remler MP, Ilog LB, Bloem BR (1995) Clinical correlates of

motor performance during paced postural tasks in Parkinson's disease. J Neurol Sci

132:133-138.

Beckstead RM, Domesick VB, Nauta WJ (1979) Efferent connections of the substantia

nigra and ventral tegmental area in the rat. Brain Res 175:191-217.

Ben-Pazi H, Bergman H, Goldberg JA, Giladi N, Hansel D, Reches A, Simon ES (2001)

Synchrony of rest tremor in multiple limbs in parkinson's disease: evidence for multiple

oscillators. J Neural Transm 108:287-296.

Benabid AL, Koudsie A, Benazzouz A, Fraix V, Ashraf A, Le Bas JF, Chabardes S,

Pollak P (2000) Subthalamic stimulation for Parkinson's disease. Arch Med Res 31:282-

289.

Benabid AL, Pollak P, Gao D, Hoffmann D, Limousin P, Gay E, Payen I, Benazzouz A

(1996) Chronic electrical stimulation of the ventralis intermedius nucleus of the thalamus

as a treatment of movement disorders. J Neurosurg 84:203-214.

Benabid AL, Pollak P, Gervason C, Hoffmann D, Gao DM, Hommel M, Perret JE, de RJ

(1991) Long-term suppression of tremor by chronic stimulation of the ventral

intermediate thalamic nucleus. Lancet 337:403-406.

Benabid AL, Pollak P, Gross C, Hoffmann D, Benazzouz A, Gao DM, Laurent A, Gentil

M, Perret J (1994) Acute and long-term effects of subthalamic nucleus stimulation in

Parkinson's disease. Stereotact Funct Neurosurg 62:76-84.

Benabid AL, Pollak P, Louveau A, Henry S, de RJ (1987) Combined (thalamotomy and

stimulation) stereotactic surgery of the VIM thalamic nucleus for bilateral Parkinson

disease. Appl Neurophysiol 50:344-346.

Benazzouz A, Gross C, Feger J, Boraud T, Bioulac B (1993) Reversal of rigidity and

improvement in motor performance by subthalamic high-frequency stimulation in MPTP-

treated monkeys. Eur J Neurosci 5:382-389.

Page 208: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

196

Benecke R, Rothwell JC, Dick JP, Day BL, Marsden CD (1987) Simple and complex

movements off and on treatment in patients with Parkinson's disease. J Neurol Neurosurg

Psychiatry 50:296-303.

Benecke R, Rothwell JC, Dick JP, Day BL, Marsden CD (1986) Performance of

simultaneous movements in patients with Parkinson's disease. Brain 109 ( Pt 4):739-757.

Bennett DA, Beckett LA, Murray AM, Shannon KM, Goetz CG, Pilgrim DM, Evans DA

(1996) Prevalence of parkinsonian signs and associated mortality in a community

population of older people. N Engl J Med 334:71-76.

Berardelli A, Rothwell JC, Thompson PD, Hallett M (2001) Pathophysiology of

bradykinesia in Parkinson's disease. Brain 124:2131-2146.

Berardelli A, Sabra AF, Hallett M (1983) Physiological mechanisms of rigidity in

Parkinson's disease. J Neurol Neurosurg Psychiatry 46:45-53.

Berendse HW, Stam CJ (2007) Stage-dependent patterns of disturbed neural synchrony in

Parkinson's disease. Parkinsonism Relat Disord 13 Suppl 3:S440-S445.

Bergman H, Deuschl G (2002) Pathophysiology of Parkinson's disease: from clinical

neurology to basic neuroscience and back. Mov Disord 17 Suppl 3:S28-S40.

Bergman H, Feingold A, Nini A, Raz A, Slovin H, Abeles M, Vaadia E (1998a)

Physiological aspects of information processing in the basal ganglia of normal and

parkinsonian primates. Trends Neurosci 21:32-38.

Bergman H, Raz A, Feingold A, Nini A, Nelken I, Hansel D, Ben-Pazi H, Reches A

(1998b) Physiology of MPTP tremor. Mov Disord 13 Suppl 3:29-34.

Bergman H, Wichmann T, DeLong MR (1990) Reversal of experimental parkinsonism

by lesions of the subthalamic nucleus. Science 249:1436-1438.

Bergman H, Wichmann T, Karmon B, DeLong MR (1994) The primate subthalamic

nucleus. II. Neuronal activity in the MPTP model of parkinsonism. J Neurophysiol

72:507-520.

Bevan MD, Atherton JF, Baufreton J (2006) Cellular principles underlying normal and

pathological activity in the subthalamic nucleus. Curr Opin Neurobiol 16:621-628.

Bevan MD, Bolam JP, Crossman AR (1994a) Convergent synaptic input from the

neostriatum and the subthalamus onto identified nigrothalamic neurons in the rat. Eur J

Neurosci 6:320-334.

Bevan MD, Crossman AR, Bolam JP (1994b) Neurons projecting from the

entopeduncular nucleus to the thalamus receive convergent synaptic inputs from the

subthalamic nucleus and the neostriatum in the rat. Brain Res 659:99-109.

Page 209: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

197

Bevan MD, Hallworth NE, Baufreton J (2007) GABAergic control of the subthalamic

nucleus. Prog Brain Res 160:173-188.

Bevan MD, Magill PJ, Hallworth NE, Bolam JP, Wilson CJ (2002a) Regulation of the

timing and pattern of action potential generation in rat subthalamic neurons in vitro by

GABA-A IPSPs. J Neurophysiol 87:1348-1362.

Bevan MD, Magill PJ, Terman D, Bolam JP, Wilson CJ (2002b) Move to the rhythm:

oscillations in the subthalamic nucleus-external globus pallidus network. Trends Neurosci

25:525-531.

Bevan MD, Smith AD, Bolam JP (1996) The substantia nigra as a site of synaptic

integration of functionally diverse information arising from the ventral pallidum and the

globus pallidus in the rat. Neuroscience 75:5-12.

Bevan MD, Wilson CJ (1999) Mechanisms underlying spontaneous oscillation and

rhythmic firing in rat subthalamic neurons. J Neurosci 19:7617-7628.

Bevan MD, Wilson CJ, Bolam JP, Magill PJ (2000) Equilibrium potential of GABA(A)

current and implications for rebound burst firing in rat subthalamic neurons in vitro. J

Neurophysiol 83:3169-3172.

Bezard E, Boraud T, Bioulac B, Gross CE (1999) Involvement of the subthalamic

nucleus in glutamatergic compensatory mechanisms. Eur J Neurosci 11:2167-2170.

Bezard E, Boraud T, Bioulac B, Gross CE (1997a) Compensatory effects of

glutamatergic inputs to the substantia nigra pars compacta in experimental parkinsonism.

Neuroscience 81:399-404.

Bezard E, Boraud T, Bioulac B, Gross CE (1997b) Presymptomatic revelation of

experimental parkinsonism. Neuroreport 8:435-438.

Bezard E, Brotchie JM, Gross CE (2001) Pathophysiology of levodopa-induced

dyskinesia: potential for new therapies. Nat Rev Neurosci 2:577-588.

Bezard E, Gross CE (1998) Compensatory mechanisms in experimental and human

parkinsonism: towards a dynamic approach. Prog Neurobiol 55:93-116.

Bickford ME, Hall WC (1992) The nigral projection to predorsal bundle cells in the

superior colliculus of the rat. J Comp Neurol 319:11-33.

Biousse V, Skibell BC, Watts RL, Loupe DN, Drews-Botsch C, Newman NJ (2004)

Ophthalmologic features of Parkinson's disease. Neurology 62:177-180.

Blandini F, Armentero MT, Martignoni E (2008) The 6-hydroxydopamine model: news

from the past. Parkinsonism Relat Disord 14 Suppl 2:S124-S129.

Blandini F, Garcia-Osuna M, Greenamyre JT (1997) Subthalamic ablation reverses

Page 210: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

198

changes in basal ganglia oxidative metabolism and motor response to apomorphine

induced by nigrostriatal lesion in rats. Eur J Neurosci 9:1407-1413.

Bloem BR (1992) Postural instability in Parkinson's disease. Clin Neurol Neurosurg 94

Suppl:S41-S45.

Bloem BR, Beckley DJ, Van Dijk JG, Zwinderman AH, Remler MP, Roos RA (1996)

Influence of dopaminergic medication on automatic postural responses and balance

impairment in Parkinson's disease. Mov Disord 11:509-521.

Bloem BR, Irwin I, Buruma OJ, Haan J, Roos RA, Tetrud JW, Langston JW (1990) The

MPTP model: versatile contributions to the treatment of idiopathic Parkinson's disease. J

Neurol Sci 97:273-293.

Blond S, Caparros-Lefebvre D, Parker F, Assaker R, Petit H, Guieu JD, Christiaens JL

(1992) Control of tremor and involuntary movement disorders by chronic stereotactic

stimulation of the ventral intermediate thalamic nucleus. J Neurosurg 77:62-68.

Bolam JP, Hanley JJ, Booth PA, Bevan MD (2000) Synaptic organisation of the basal

ganglia. J Anat 196 ( Pt 4):527-542.

Bolam JP, Powell JF, Totterdell S, Smith AD (1981) The proportion of neurons in the rat

neostriatum that project to the substantia nigra demonstrated using horseradish

peroxidase conjugated with wheatgerm agglutinin. Brain Res 220:339-343.

Bolam JP, Smith Y (1992) The striatum and the globus pallidus send convergent synaptic

inputs onto single cells in the entopeduncular nucleus of the rat: a double anterograde

labelling study combined with postembedding immunocytochemistry for GABA. J Comp

Neurol 321:456-476.

Bolam JP, Smith Y, Ingham CA, von KM, Smith AD (1993) Convergence of synaptic

terminals from the striatum and the globus pallidus onto single neurones in the substantia

nigra and the entopeduncular nucleus. Prog Brain Res 99:73-88.

Bolam JP, Wainer BH, Smith AD (1984) Characterization of cholinergic neurons in the

rat neostriatum. A combination of choline acetyltransferase immunocytochemistry,

Golgi-impregnation and electron microscopy. Neuroscience 12:711-718.

Boraud T, Bezard E, Bioulac B, Gross C (1996) High frequency stimulation of the

internal Globus Pallidus (GPi) simultaneously improves parkinsonian symptoms and

reduces the firing frequency of GPi neurons in the MPTP-treated monkey. Neurosci Lett

215:17-20.

Boraud T, Bezard E, Bioulac B, Gross CE (2000a) Ratio of inhibited-to-activated pallidal

neurons decreases dramatically during passive limb movement in the MPTP-treated

monkey. J Neurophysiol 83:1760-1763.

Boraud T, Bezard E, Bioulac B, Gross CE (2002) From single extracellular unit recording

Page 211: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

199

in experimental and human Parkinsonism to the development of a functional concept of

the role played by the basal ganglia in motor control. Prog Neurobiol 66:265-283.

Boraud T, Bezard E, Bioulac B, Gross CE (2001) Dopamine agonist-induced dyskinesias

are correlated to both firing pattern and frequency alterations of pallidal neurones in the

MPTP-treated monkey. Brain 124:546-557.

Boraud T, Bezard E, Guehl D, Bioulac B, Gross C (1998) Effects of L-DOPA on

neuronal activity of the globus pallidus externalis (GPe) and globus pallidus internalis

(GPi) in the MPTP-treated monkey. Brain Res 787:157-160.

Boraud T, Bezard E, Stutzmann JM, Bioulac B, Gross CE (2000b) Effects of riluzole on

the electrophysiological activity of pallidal neurons in the 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine-treated monkey. Neurosci Lett 281:75-78.

Boussaoud D, Joseph JP (1985) Role of the cat substantia nigra pars reticulata in eye and

head movements. II. Effects of local pharmacological injections. Exp Brain Res 57:297-

304.

Braak H, Braak E (1986) Nuclear configuration and neuronal types of the nucleus niger

in the brain of the human adult. Hum Neurobiol 5:71-82.

Breit S, Bouali-Benazzouz R, Benabid AL, Benazzouz A (2001) Unilateral lesion of the

nigrostriatal pathway induces an increase of neuronal activity of the pedunculopontine

nucleus, which is reversed by the lesion of the subthalamic nucleus in the rat. Eur J

Neurosci 14:1833-1842.

Breit S, Lessmann L, Unterbrink D, Popa RC, Gasser T, Schulz JB (2006) Lesion of the

pedunculopontine nucleus reverses hyperactivity of the subthalamic nucleus and

substantia nigra pars reticulata in a 6-hydroxydopamine rat model. Eur J Neurosci

24:2275-2282.

Britton TC, Thompson PD, Day BL, Rothwell JC, Findley LJ, Marsden CD (1992)

"Resetting" of postural tremors at the wrist with mechanical stretches in Parkinson's

disease, essential tremor, and normal subjects mimicking tremor. Ann Neurol 31:507-514.

Britton TC, Thompson PD, Day BL, Rothwell JC, Findley LJ, Marsden CD (1993a)

Modulation of postural wrist tremors by magnetic stimulation of the motor cortex in

patients with Parkinson's disease or essential tremor and in normal subjects mimicking

tremor. Ann Neurol 33:473-479.

Britton TC, Thompson PD, Day BL, Rothwell JC, Findley LJ, Marsden CD (1993b)

Modulation of postural tremors at the wrist by supramaximal electrical median nerve

shocks in essential tremor, Parkinson's disease and normal subjects mimicking tremor. J

Neurol Neurosurg Psychiatry 56:1085-1089.

Brooks DJ (2008) The role of structural and functional imaging in parkinsonian states

with a description of PET technology. Semin Neurol 28:435-445.

Page 212: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

200

Brooks DJ (1997) PET and SPECT studies in Parkinson's disease. Baillieres Clin Neurol

6:69-87.

Brooks WJ, Jarvis MF, Wagner GC (1989) Astrocytes as a primary locus for the

conversion MPTP into MPP+. J Neural Transm 76:1-12.

Brotchie P, Iansek R, Horne MK (1991a) Motor function of the monkey globus pallidus.

1. Neuronal discharge and parameters of movement. Brain 114 ( Pt 4):1667-1683.

Brotchie P, Iansek R, Horne MK (1991b) Motor function of the monkey globus pallidus.

2. Cognitive aspects of movement and phasic neuronal activity. Brain 114 ( Pt 4):1685-

1702.

Brown LL, Feldman SM, Smith DM, Cavanaugh JR, Ackermann RF, Graybiel AM

(2002) Differential metabolic activity in the striosome and matrix compartments of the rat

striatum during natural behaviors. J Neurosci 22:305-314.

Brown LL, Markman MH, Wolfson LI, Dvorkin B, Warner C, Katzman R (1979) A

direct role of dopamine in the rat subthalamic nucleus and an adjacent intrapeduncular

area. Science 206:1416-1418.

Brown P (2007) Abnormal oscillatory synchronisation in the motor system leads to

impaired movement. Curr Opin Neurobiol 17:656-664.

Brown P (2003) Oscillatory nature of human basal ganglia activity: relationship to the

pathophysiology of Parkinson's disease. Mov Disord 18:357-363.

Brown P (2006) Bad oscillations in Parkinson's disease. J Neural Transm Suppl27-30.

Brown P, Marsden CD (1998) What do the basal ganglia do? Lancet 351:1801-1804.

Brown P, Mazzone P, Oliviero A, Altibrandi MG, Pilato F, Tonali PA, Di Lazzaro V

(2004) Effects of stimulation of the subthalamic area on oscillatory pallidal activity in

Parkinson's disease. Exp Neurol 188:480-490.

Brown P, Oliviero A, Mazzone P, Insola A, Tonali P, Di Lazzaro V (2001) Dopamine

dependency of oscillations between subthalamic nucleus and pallidum in Parkinson's

disease. J Neurosci 21:1033-1038.

Brown P, Williams D (2005) Basal ganglia local field potential activity: Character and

functional significance in the human. Clin Neurophysiol 116:2510-2519.

Brown RG, Jahanshahi M (1996) Cognitive-motor dysfunction in Parkinson's disease.

Eur Neurol 36 Suppl 1:24-31.

Brucke C, Kempf F, Kupsch A, Schneider GH, Krauss JK, Aziz T, Yarrow K, Pogosyan

A, Brown P, Kuhn AA (2008) Movement-related synchronization of gamma activity is

lateralized in patients with dystonia. Eur J Neurosci 27:2322-2329.

Page 213: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

201

Brudzynski SM, Wu M, Mogenson GJ (1988) Modulation of locomotor activity induced

by injections of carbachol into the tegmental pedunculopontine nucleus and adjacent

areas in the rat. Brain Res 451:119-125.

Burbaud P, Bonnet B, Guehl D, Lagueny A, Bioulac B (1998) Movement disorders

induced by gamma-aminobutyric agonist and antagonist injections into the internal

globus pallidus and substantia nigra pars reticulata of the monkey. Brain Res 780:102-

107.

Burbaud P, Gross C, Benazzouz A, Coussemacq M, Bioulac B (1995) Reduction of

apomorphine-induced rotational behaviour by subthalamic lesion in 6-OHDA lesioned

rats is associated with a normalization of firing rate and discharge pattern of pars

reticulata neurons. Exp Brain Res 105:48-58.

Burke RE, Fahn S, Marsden CD, Bressman SB, Moskowitz C, Friedman J (1985)

Validity and reliability of a rating scale for the primary torsion dystonias. Neurology

35:73-77.

Burleigh A, Horak F, Nutt J, Frank J (1995) Levodopa reduces muscle tone and lower

extremity tremor in Parkinson's disease. Can J Neurol Sci 22:280-285.

Burleigh-Jacobs A, Horak FB, Nutt JG, Obeso JA (1997) Step initiation in Parkinson's

disease: influence of levodopa and external sensory triggers. Mov Disord 12:206-215.

Burne JA (1987) Reflex origin of parkinsonian tremor. Exp Neurol 97:327-339.

Burns RS, Chiueh CC, Markey SP, Ebert MH, Jacobowitz DM, Kopin IJ (1983) A

primate model of parkinsonism: selective destruction of dopaminergic neurons in the pars

compacta of the substantia nigra by N-methyl-4-phenyl-1,2,3,6-tetrahydropyridine. Proc

Natl Acad Sci U S A 80:4546-4550.

Buzsaki G (2006) Rhythms of the Brain. New York: Oxford University Press.

Buzsaki G, Draguhn A (2004) Neuronal oscillations in cortical networks. Science

304:1926-1929.

Calabresi P, Maj R, Pisani A, Mercuri NB, Bernardi G (1992) Long-term synaptic

depression in the striatum: physiological and pharmacological characterization. J

Neurosci 12:4224-4233.

Caligiuri MP, Peterson S (1993) A quantitative study of levodopa-induced dyskinesia in

Parkinson's disease. J Neural Transm Park Dis Dement Sect 6:89-98.

Calne DB, Langston JW (1983) Aetiology of Parkinson's disease. Lancet 2:1457-1459.

Calne DB, Langston JW, Martin WR, Stoessl AJ, Ruth TJ, Adam MJ, Pate BD, Schulzer

M (1985) Positron emission tomography after MPTP: observations relating to the cause

of Parkinson's disease. Nature 317:246-248.

Page 214: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

202

Campbell GA, Eckardt MJ, Weight FF (1985) Dopaminergic mechanisms in subthalamic

nucleus of rat: analysis using horseradish peroxidase and microiontophoresis. Brain Res

333:261-270.

Campbell KJ, Takada M (1989) Bilateral tectal projection of single nigrostriatal

dopamine cells in the rat. Neuroscience 33:311-321.

Canteras NS, Shammah-Lagnado SJ, Silva BA, Ricardo JA (1990) Afferent connections

of the subthalamic nucleus: a combined retrograde and anterograde horseradish

peroxidase study in the rat. Brain Res 513:43-59.

Carpenter MB, Baton RR, III, Carleton SC, Keller JT (1981) Interconnections and

organization of pallidal and subthalamic nucleus neurons in the monkey. J Comp Neurol

197:579-603.

Cassidy M, Brown P (2003) Spectral phase estimates in the setting of multidirectional

coupling. J Neurosci Methods 127:95-103.

Cassidy M, Mazzone P, Oliviero A, Insola A, Tonali P, Di L, V, Brown P (2002)

Movement-related changes in synchronization in the human basal ganglia. Brain

125:1235-1246.

Celada P, Paladini CA, Tepper JM (1999) GABAergic control of rat substantia nigra

dopaminergic neurons: role of globus pallidus and substantia nigra pars reticulata.

Neuroscience 89:813-825.

Centonze D, Gubellini P, Picconi B, Calabresi P, Giacomini P, Bernardi G (1999)

Unilateral dopamine denervation blocks corticostriatal LTP. J Neurophysiol 82:3575-

3579.

Chadha A, Dawson LG, Jenner PG, Duty S (2000) Effect of unilateral 6-

hydroxydopamine lesions of the nigrostriatal pathway on GABA(A) receptor subunit

gene expression in the rodent basal ganglia and thalamus. Neuroscience 95:119-126.

Chan P, DeLanney LE, Irwin I, Langston JW, Di MD (1991) Rapid ATP loss caused by

1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine in mouse brain. J Neurochem 57:348-351.

Chang JY, Shi LH, Luo F, Woodward DJ (2006) Neural responses in multiple basal

ganglia regions following unilateral dopamine depletion in behaving rats performing a

treadmill locomotion task. Exp Brain Res 172:193-207.

Charara A, Smith Y, Parent A (1996) Glutamatergic inputs from the pedunculopontine

nucleus to midbrain dopaminergic neurons in primates: Phaseolus vulgaris-

leucoagglutinin anterograde labeling combined with postembedding glutamate and

GABA immunohistochemistry. J Comp Neurol 364:254-266.

Charpier S, Deniau JM (1997) In vivo activity-dependent plasticity at cortico-striatal

connections: evidence for physiological long-term potentiation. Proc Natl Acad Sci U S

Page 215: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

203

A 94:7036-7040.

Charpier S, Mahon S, Deniau JM (1999) In vivo induction of striatal long-term

potentiation by low-frequency stimulation of the cerebral cortex. Neuroscience 91:1209-

1222.

Chase TN (1998) The significance of continuous dopaminergic stimulation in the

treatment of Parkinson's disease. Drugs 55 Suppl 1:1-9.

Chen CC, Kuhn AA, Hoffmann KT, Kupsch A, Schneider GH, Trottenberg T, Krauss JK,

Wohrle JC, Bardinet E, Yelnik J, Brown P (2006a) Oscillatory pallidal local field

potential activity correlates with involuntary EMG in dystonia. Neurology 66:418-420.

Chen CC, Kuhn AA, Trottenberg T, Kupsch A, Schneider GH, Brown P (2006b)

Neuronal activity in globus pallidus interna can be synchronized to local field potential

activity over 3-12 Hz in patients with dystonia. Exp Neurol 202:480-486.

Chen CC, Litvak V, Gilbertson T, Kuhn A, Lu CS, Lee ST, Tsai CH, Tisch S, Limousin

P, Hariz M, Brown P (2007) Excessive synchronization of basal ganglia neurons at 20 Hz

slows movement in Parkinson's disease. Exp Neurol 205:214-221.

Chergui K, Akaoka H, Charlety PJ, Saunier CF, Buda M, Chouvet G (1994) Subthalamic

nucleus modulates burst firing of nigral dopamine neurones via NMDA receptors.

Neuroreport 5:1185-1188.

Chergui K, Charlety PJ, Akaoka H, Saunier CF, Brunet JL, Buda M, Svensson TH,

Chouvet G (1993) Tonic activation of NMDA receptors causes spontaneous burst

discharge of rat midbrain dopamine neurons in vivo. Eur J Neurosci 5:137-144.

Cheruel F, Dormont JF, Amalric M, Schmied A, Farin D (1994) The role of putamen and

pallidum in motor initiation in the cat. I. Timing of movement-related single-unit activity.

Exp Brain Res 100:250-266.

Chesselet MF, Graybiel AM (1986) Striatal neurons expressing somatostatin-like

immunoreactivity: evidence for a peptidergic interneuronal system in the cat.

Neuroscience 17:547-571.

Chevalier G, Deniau JM (1990) Disinhibition as a basic process in the expression of

striatal functions. Trends Neurosci 13:277-280.

Chiba K, Trevor A, Castagnoli N, Jr. (1984) Metabolism of the neurotoxic tertiary amine,

MPTP, by brain monoamine oxidase. Biochem Biophys Res Commun 120:574-578.

Cho J, Duke D, Manzino L, Sonsalla PK, West MO (2002) Dopamine depletion causes

fragmented clustering of neurons in the sensorimotor striatum: evidence of lasting

reorganization of corticostriatal input. J Comp Neurol 452:24-37.

Chudler EH, Sugiyama K, Dong WK (1995) Multisensory convergence and integration in

Page 216: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

204

the neostriatum and globus pallidus of the rat. Brain Res 674:33-45.

Clarke CE, Boyce S, Robertson RG, Sambrook MA, Crossman AR (1989) Drug-induced

dyskinesia in primates rendered hemiparkinsonian by intracarotid administration of 1-

methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). J Neurol Sci 90:307-314.

Clarke CE, Sambrook MA, Mitchell IJ, Crossman AR (1987) Levodopa-induced

dyskinesia and response fluctuations in primates rendered parkinsonian with 1-methyl-4-

phenyl-1,2,3,6-tetrahydropyridine (MPTP). J Neurol Sci 78:273-280.

Cody FW, MacDermott N, Matthews PB, Richardson HC (1986) Observations on the

genesis of the stretch reflex in Parkinson's disease. Brain 109 ( Pt 2):229-249.

Comella CL (2003) Sleep disturbances in Parkinson's disease. Curr Neurol Neurosci Rep

3:173-180.

Comella CL, Stebbins GT, Goetz CG, Chmura TA, Bressman SB, Lang AE (1997)

Teaching tape for the motor section of the Toronto Western Spasmodic Torticollis Scale.

Mov Disord 12:570-575.

Conde H, Dormont JF, Farin D (1998) The role of the pedunculopontine tegmental

nucleus in relation to conditioned motor performance in the cat. II. Effects of reversible

inactivation by intracerebral microinjections. Exp Brain Res 121:411-418.

COOPER IS (1954) Procaine injection of the globus pallidus in Parkinsonism. Psychiatr

Q 28:22-23.

Cossette M, Levesque M, Parent A (1999) Extrastriatal dopaminergic innervation of

human basal ganglia. Neurosci Res 34:51-54.

Costa RM, Lin SC, Sotnikova TD, Cyr M, Gainetdinov RR, Caron MG, Nicolelis MA

(2006) Rapid alterations in corticostriatal ensemble coordination during acute dopamine-

dependent motor dysfunction. Neuron 52:359-369.

Cotzias GC, Van Woert MH, Schiffer LM (1967) Aromatic amino acids and modification

of parkinsonism. N Engl J Med 276:374-379.

Courtemanche R, Fujii N, Graybiel AM (2003) Synchronous, focally modulated beta-

band oscillations characterize local field potential activity in the striatum of awake

behaving monkeys. J Neurosci 23:11741-11752.

Cowan WM, Powell TP (1956) A study of thalamo-striate relations in the monkey. Brain

79:364-390.

Critchley EM (1981) Speech disorders of Parkinsonism: a review. J Neurol Neurosurg

Psychiatry 44:751-758.

Crossman AR, Mitchell IJ, Sambrook MA, Jackson A (1988) Chorea and myoclonus in

Page 217: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

205

the monkey induced by gamma-aminobutyric acid antagonism in the lentiform complex.

The site of drug action and a hypothesis for the neural mechanisms of chorea. Brain 111

( Pt 5):1211-1233.

Crutcher MD, DeLong MR (1984a) Single cell studies of the primate putamen. II.

Relations to direction of movement and pattern of muscular activity. Exp Brain Res

53:244-258.

Crutcher MD, DeLong MR (1984b) Single cell studies of the primate putamen. I.

Functional organization. Exp Brain Res 53:233-243.

Cruz AV, Mallet N, Magill PJ, Brown P, Averbeck BB (2009) The effects of dopamine

depletion on network entropy in the external globus pallidus. J Neurophysiol.

Cummings JL (1992) Depression and Parkinson's disease: a review. Am J Psychiatry

149:443-454.

Cummings JL, Masterman DL (1999) Depression in patients with Parkinson's disease. Int

J Geriatr Psychiatry 14:711-718.

Davis KD, Taub E, Houle S, Lang AE, Dostrovsky JO, Tasker RR, Lozano AM (1997)

Globus pallidus stimulation activates the cortical motor system during alleviation of

parkinsonian symptoms. Nat Med 3:671-674.

Day M, Wokosin D, Plotkin JL, Tian X, Surmeier DJ (2008) Differential excitability and

modulation of striatal medium spiny neuron dendrites. J Neurosci 28:11603-11614.

Decavel C, Lescaudron L, Mons N, Calas A (1987) First visualization of dopaminergic

neurons with a monoclonal antibody to dopamine: a light and electron microscopic study.

J Histochem Cytochem 35:1245-1251.

Defebvre L, Blatt JL, Blond S, Bourriez JL, Gieu JD, Destee A (1996) Effect of thalamic

stimulation on gait in Parkinson disease. Arch Neurol 53:898-903.

Degos B, Deniau JM, Chavez M, Maurice N (2008) Chronic but not Acute Dopaminergic

Transmission Interruption Promotes a Progressive Increase in Cortical Beta Frequency

Synchronization: Relationships to Vigilance State and Akinesia. Cereb Cortex.

Deiber MP, Pollak P, Passingham R, Landais P, Gervason C, Cinotti L, Friston K,

Frackowiak R, Mauguiere F, Benabid AL (1993) Thalamic stimulation and suppression

of parkinsonian tremor. Evidence of a cerebellar deactivation using positron emission

tomography. Brain 116 ( Pt 1):267-279.

Deleu D, Northway MG, Hanssens Y (2002) Clinical pharmacokinetic and

pharmacodynamic properties of drugs used in the treatment of Parkinson's disease. Clin

Pharmacokinet 41:261-309.

DeLong MR (1990) Primate models of movement disorders of basal ganglia origin.

Page 218: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

206

Trends Neurosci 13:281-285.

DeLong MR (1971) Activity of pallidal neurons during movement. J Neurophysiol

34:414-427.

DeLong MR, Alexander GE, Georgopoulos AP, Crutcher MD, Mitchell SJ, Richardson

RT (1984a) Role of basal ganglia in limb movements. Hum Neurobiol 2:235-244.

DeLong MR, Crutcher MD, Georgopoulos AP (1983) Relations between movement and

single cell discharge in the substantia nigra of the behaving monkey. J Neurosci 3:1599-

1606.

DeLong MR, Crutcher MD, Georgopoulos AP (1985) Primate globus pallidus and

subthalamic nucleus: functional organization. J Neurophysiol 53:530-543.

DeLong MR, Georgopoulos AP, Crutcher MD, Mitchell SJ, Richardson RT, Alexander

GE (1984b) Functional organization of the basal ganglia: contributions of single-cell

recording studies. Ciba Found Symp 107:64-82.

DeLong MR, Wichmann T (2007) Circuits and circuit disorders of the basal ganglia.

Arch Neurol 64:20-24.

Deniau JM, Hammond C, Riszk A, Feger J (1978) Electrophysiological properties of

identified output neurons of the rat substantia nigra (pars compacta and pars reticulata):

evidences for the existence of branched neurons. Exp Brain Res 32:409-422.

Deniau JM, Mailly P, Maurice N, Charpier S (2007) The pars reticulata of the substantia

nigra: a window to basal ganglia output. Prog Brain Res 160:151-172.

Deuschl G, Bain P, Brin M (1998) Consensus statement of the Movement Disorder

Society on Tremor. Ad Hoc Scientific Committee. Mov Disord 13 Suppl 3:2-23.

Deuschl G, Lucking CH (1990) Physiology and clinical applications of hand muscle

reflexes. Electroencephalogr Clin Neurophysiol Suppl 41:84-101.

Deuschl G, Raethjen J, Baron R, Lindemann M, Wilms H, Krack P (2000) The

pathophysiology of parkinsonian tremor: a review. J Neurol 247 Suppl 5:V33-V48.

Deuschl G, Raethjen J, Lindemann M, Krack P (2001) The pathophysiology of tremor.

Muscle Nerve 24:716-735.

Deuschl G, Wilms H, Krack P, Wurker M, Heiss WD (1999) Function of the cerebellum

in Parkinsonian rest tremor and Holmes' tremor. Ann Neurol 46:126-128.

Deutch AY, Goldstein M, Roth RH (1986) The ascending projections of the

dopaminergic neurons of the substantia nigra, zona reticulata: a combined retrograde

tracer-immunohistochemical study. Neurosci Lett 71:257-263.

Page 219: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

207

Devos D, Labyt E, Cassim F, Bourriez JL, Reyns N, Touzet G, Blond S, Guieu JD,

Derambure P, Destee A, Defebvre L (2003) Subthalamic stimulation influences

postmovement cortical somatosensory processing in Parkinson's disease. Eur J Neurosci

18:1884-1888.

Devos D, Szurhaj W, Reyns N, Labyt E, Houdayer E, Bourriez JL, Cassim F,

Krystkowiak P, Blond S, Destee A, Derambure P, Defebvre L (2006) Predominance of

the contralateral movement-related activity in the subthalamo-cortical loop. Clin

Neurophysiol 117:2315-2327.

Di CG, Porceddu ML, Morelli M, Mulas ML, Gessa GL (1979) Evidence for a

GABAergic projection from the substantia nigra to the ventromedial thalamus and to the

superior colliculus of the rat. Brain Res 176:273-284.

Diamond A, Shahed J, Jankovic J (2007) The effects of subthalamic nucleus deep brain

stimulation on parkinsonian tremor. J Neurol Sci 260:199-203.

Dick JP, Rothwell JC, Berardelli A, Thompson PD, Gioux M, Benecke R, Day BL,

Marsden CD (1986) Associated postural adjustments in Parkinson's disease. J Neurol

Neurosurg Psychiatry 49:1378-1385.

Dickson DW, Rademakers R, Hutton ML (2007) Progressive supranuclear palsy:

pathology and genetics. Brain Pathol 17:74-82.

Djaldetti R, Shifrin A, Rogowski Z, Sprecher E, Melamed E, Yarnitsky D (2004)

Quantitative measurement of pain sensation in patients with Parkinson disease.

Neurology 62:2171-2175.

Dodson MW, Guo M (2007) Pink1, Parkin, DJ-1 and mitochondrial dysfunction in

Parkinson's disease. Curr Opin Neurobiol 17:331-337.

Donoghue JP, Sanes JN, Hatsopoulos NG, Gaal G (1998) Neural discharge and local

field potential oscillations in primate motor cortex during voluntary movements. J

Neurophysiol 79:159-173.

Dormont JF, Conde H, Cheruel F, Farin D (1997) Correlations between activity of

pallidal neurons and motor parameters. Somatosens Mot Res 14:281-294.

Dormont JF, Conde H, Farin D (1998) The role of the pedunculopontine tegmental

nucleus in relation to conditioned motor performance in the cat. I. Context-dependent and

reinforcement-related single unit activity. Exp Brain Res 121:401-410.

Dorval AD, Russo GS, Hashimoto T, Xu W, Grill WM, Vitek JL (2008) Deep brain

stimulation reduces neuronal entropy in the MPTP-primate model of Parkinson's disease.

J Neurophysiol 100:2807-2818.

Dostrovsky JO, Levy R, Wu JP, Hutchison WD, Tasker RR, Lozano AM (2000)

Microstimulation-induced inhibition of neuronal firing in human globus pallidus. J

Page 220: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

208

Neurophysiol 84:570-574.

Doty RL, Stern MB, Pfeiffer C, Gollomp SM, Hurtig HI (1992) Bilateral olfactory

dysfunction in early stage treated and untreated idiopathic Parkinson's disease. J Neurol

Neurosurg Psychiatry 55:138-142.

Doudet DJ, Gross C, Arluison M, Bioulac B (1990) Modifications of precentral cortex

discharge and EMG activity in monkeys with MPTP-induced lesions of DA nigral

neurons. Exp Brain Res 80:177-188.

Doyle LM, Kuhn AA, Hariz M, Kupsch A, Schneider GH, Brown P (2005a) Levodopa-

induced modulation of subthalamic beta oscillations during self-paced movements in

patients with Parkinson's disease. Eur J Neurosci 21:1403-1412.

Doyle LM, Yarrow K, Brown P (2005b) Lateralization of event-related beta

desynchronization in the EEG during pre-cued reaction time tasks. Clin Neurophysiol

116:1879-1888.

Drouot X, Oshino S, Jarraya B, Besret L, Kishima H, Remy P, Dauguet J, Lefaucheur JP,

Dolle F, Conde F, Bottlaender M, Peschanski M, Keravel Y, Hantraye P, Palfi S (2004)

Functional recovery in a primate model of Parkinson's disease following motor cortex

stimulation. Neuron 44:769-778.

ECCLES JC (1951) Interpretation of action potentials evoked in the cerebral cortex.

Electroencephalogr Clin Neurophysiol 3:449-464.

Edley SM, Graybiel AM (1983) The afferent and efferent connections of the feline

nucleus tegmenti pedunculopontinus, pars compacta. J Comp Neurol 217:187-215.

Eidelberg D, Moeller JR, Kazumata K, Antonini A, Sterio D, Dhawan V, Spetsieris P,

Alterman R, Kelly PJ, Dogali M, Fazzini E, Beric A (1997) Metabolic correlates of

pallidal neuronal activity in Parkinson's disease. Brain 120 ( Pt 8):1315-1324.

Elble RJ (1996) Central mechanisms of tremor. J Clin Neurophysiol 13:133-144.

Elble RJ, Higgins C, Hughes L (1992) Phase resetting and frequency entrainment of

essential tremor. Exp Neurol 116:355-361.

Elsworth JD, Taylor JR, Sladek JR, Jr., Collier TJ, Redmond DE, Jr., Roth RH (2000)

Striatal dopaminergic correlates of stable parkinsonism and degree of recovery in old-

world primates one year after MPTP treatment. Neuroscience 95:399-408.

Erez Y, Czitron H, McCairn K, Belelovsky K, Bar-Gad I (2009) Short-term depression of

synaptic transmission during stimulation in the globus pallidus of 1-methyl-4-phenyl-

1,2,3,6-tetrahydropyridine-treated primates. J Neurosci 29:7797-7802.

Escola L, Michelet T, Douillard G, Guehl D, Bioulac B, Burbaud P (2002) Disruption of

the proprioceptive mapping in the medial wall of parkinsonian monkeys. Ann Neurol

Page 221: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

209

52:581-587.

Evarts EV, Teravainen H, Calne DB (1981) Reaction time in Parkinson's disease. Brain

104:167-186.

Fabbrini G, Juncos J, Mouradian MM, Serrati C, Chase TN (1987) Levodopa

pharmacokinetic mechanisms and motor fluctuations in Parkinson's disease. Ann Neurol

21:370-376.

Fahn S, Elton RL, members of the UPDRS Development Committe (1987) Unified

Parkinson's Disease Rating Scale. In: Recent developments in Parkinson's disease (Fahn

S, Marsden SD, Calne DB, Goldstein M, eds), pp 153-163.

Fawcett AP, Dostrovsky JO, Lozano AM, Hutchison WD (2005) Eye movement-related

responses of neurons in human subthalamic nucleus. Exp Brain Res 162:357-365.

Feger J, Hammond C, Rouzaire-Dubois B (1979) Pharmacological properties of

acetylcholine-induced excitation of subthalamic nucleus neurones. Br J Pharmacol

65:511-515.

Feger J, Hassani OK, Mouroux M (1997) The subthalamic nucleus and its connections.

New electrophysiological and pharmacological data. Adv Neurol 74:31-43.

Filali M, Hutchison WD, Palter VN, Lozano AM, Dostrovsky JO (2004) Stimulation-

induced inhibition of neuronal firing in human subthalamic nucleus. Exp Brain Res

156:274-281.

Filion M, Tremblay L (1991) Abnormal spontaneous activity of globus pallidus neurons

in monkeys with MPTP-induced parkinsonism. Brain Res 547:142-151.

Filion M, Tremblay L, Bedard PJ (1988) Abnormal influences of passive limb movement

on the activity of globus pallidus neurons in parkinsonian monkeys. Brain Res 444:165-

176.

Filion M, Tremblay L, Bedard PJ (1991) Effects of dopamine agonists on the

spontaneous activity of globus pallidus neurons in monkeys with MPTP-induced

parkinsonism. Brain Res 547:152-161.

Findley LJ, Gresty MA, Halmagyi GM (1981) Tremor, the cogwheel phenomenon and

clonus in Parkinson's disease. J Neurol Neurosurg Psychiatry 44:534-546.

Fine J, Duff J, Chen R, Chir B, Hutchison W, Lozano AM, Lang AE (2000) Long-term

follow-up of unilateral pallidotomy in advanced Parkinson's disease. N Engl J Med

342:1708-1714.

Fitzgibbon SP, Pope KJ, Mackenzie L, Clark CR, Willoughby JO (2004) Cognitive tasks

augment gamma EEG power. Clin Neurophysiol 115:1802-1809.

Page 222: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

210

Flaherty AW, Graybiel AM (1993) Two input systems for body representations in the

primate striatal matrix: experimental evidence in the squirrel monkey. J Neurosci

13:1120-1137.

Flaherty AW, Graybiel AM (1994) Input-output organization of the sensorimotor

striatum in the squirrel monkey. J Neurosci 14:599-610.

Flaherty AW, Graybiel AM (1991) Corticostriatal transformations in the primate

somatosensory system. Projections from physiologically mapped body-part

representations. J Neurophysiol 66:1249-1263.

Flores G, Hernandez S, Rosales MG, Sierra A, Martines-Fong D, Flores-Hernandez J,

Aceves J (1996) M3 muscarinic receptors mediate cholinergic excitation of the

spontaneous activity of subthalamic neurons in the rat. Neurosci Lett 203:203-206.

Flores G, Rosales MG, Hernandez S, Sierra A, Aceves J (1995) 5-Hydroxytryptamine

increases spontaneous activity of subthalamic neurons in the rat. Neurosci Lett 192:17-20.

Florio T, Scarnati E, Confalone G, Minchella D, Galati S, Stanzione P, Stefani A,

Mazzone P (2007) High-frequency stimulation of the subthalamic nucleus modulates the

activity of pedunculopontine neurons through direct activation of excitatory fibres as well

as through indirect activation of inhibitory pallidal fibres in the rat. Eur J Neurosci

25:1174-1186.

Foffani G, Ardolino G, Egidi M, Caputo E, Bossi B, Priori A (2006) Subthalamic

oscillatory activities at beta or higher frequency do not change after high-frequency DBS

in Parkinson's disease. Brain Res Bull 69:123-130.

Foffani G, Ardolino G, Meda B, Egidi M, Rampini P, Caputo E, Baselli G, Priori A

(2005a) Altered subthalamo-pallidal synchronisation in parkinsonian dyskinesias. J

Neurol Neurosurg Psychiatry 76:426-428.

Foffani G, Ardolino G, Rampini P, Tamma F, Caputo E, Egidi M, Cerutti S, Barbieri S,

Priori A (2005b) Physiological recordings from electrodes implanted in the basal ganglia

for deep brain stimulation in Parkinson's disease. the relevance of fast subthalamic

rhythms. Acta Neurochir Suppl 93:97-99.

Foffani G, Bianchi AM, Baselli G, Priori A (2005c) Movement-related frequency

modulation of beta oscillatory activity in the human subthalamic nucleus. J Physiol

568:699-711.

Foffani G, Priori A (2007) Information theory, single neurons and gamma oscillations in

the human subthalamic nucleus. Exp Neurol 205:292-293.

Foffani G, Priori A, Egidi M, Rampini P, Tamma F, Caputo E, Moxon KA, Cerutti S,

Barbieri S (2003) 300-Hz subthalamic oscillations in Parkinson's disease. Brain

126:2153-2163.

Page 223: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

211

Fogelson N, Pogosyan A, Kuhn AA, Kupsch A, van BG, Speelman H, Tijssen M,

Quartarone A, Insola A, Mazzone P, Di L, V, Limousin P, Brown P (2005a) Reciprocal

interactions between oscillatory activities of different frequencies in the subthalamic

region of patients with Parkinson's disease. Eur J Neurosci 22:257-266.

Fogelson N, Williams D, Tijssen M, van BG, Speelman H, Brown P (2006) Different

functional loops between cerebral cortex and the subthalmic area in Parkinson's disease.

Cereb Cortex 16:64-75.

Fogelson N, Williams D, Tijssen M, van BG, Speelman H, Brown P (2005b) Different

Functional Loops between Cerebral Cortex and the Subthalmic Area in Parkinson's

Disease. Cereb Cortex 16:64-75.

Foncke EM, Bour LJ, Speelman JD, Koelman JH, Tijssen MA (2007) Local field

potentials and oscillatory activity of the internal globus pallidus in myoclonus-dystonia.

Mov Disord 22:369-376.

Francois C, Percheron G, Yelnik J, Heyner S (1985) A histological atlas of the macaque

(Macaca mulatta) substantia nigra in ventricular coordinates. Brain Res Bull 14:349-367.

Francois C, Savy C, Jan C, Tande D, Hirsch EC, Yelnik J (2000) Dopaminergic

innervation of the subthalamic nucleus in the normal state, in MPTP-treated monkeys,

and in Parkinson's disease patients. J Comp Neurol 425:121-129.

Freiwald WA, Kreiter AK, Singer W (2001) Synchronization and assembly formation in

the visual cortex. Prog Brain Res 130:111-140.

Fronczek R, Overeem S, Lee SY, Hegeman IM, van PJ, van Duinen SG, Lammers GJ,

Swaab DF (2007) Hypocretin (orexin) loss in Parkinson's disease. Brain 130:1577-1585.

Fujimoto K, Kita H (1993) Response characteristics of subthalamic neurons to the

stimulation of the sensorimotor cortex in the rat. Brain Res 609:185-192.

Gagnon JF, Postuma RB, Mazza S, Doyon J, Montplaisir J (2006) Rapid-eye-movement

sleep behaviour disorder and neurodegenerative diseases. Lancet Neurol 5:424-432.

Gale JT, Shields DC, Jain FA, Amirnovin R, Eskandar EN (2009) Subthalamic nucleus

discharge patterns during movement in the normal monkey and Parkinsonian patient.

Brain Res 1260:15-23.

Galvan A, Wichmann T (2008) Pathophysiology of parkinsonism. Clin Neurophysiol

119:1459-1474.

Garcia-Rill E (1991) The pedunculopontine nucleus. Prog Neurobiol 36:363-389.

Garcia-Rill E, Homma Y, Skinner RD (2004) Arousal mechanisms related to posture and

locomotion: 1. Descending modulation. Prog Brain Res 143:283-290.

Page 224: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

212

Garcia-Rill E, Houser CR, Skinner RD, Smith W, Woodward DJ (1987) Locomotion-

inducing sites in the vicinity of the pedunculopontine nucleus. Brain Res Bull 18:731-738.

Garcia-Rill E, Skinner RD (1988) Modulation of rhythmic function in the posterior

midbrain. Neuroscience 27:639-654.

Garcia-Rill E, Skinner RD (1991) Modulation of rhythmic functions by the brainstem. In:

Neurobiological Basis of Human Locomotion (Shimamura M, Grillner S, Edgerton VR,

eds), pp 137-158. Tokyo: Japan Scientific Societies Press.

Garcia-Rill E, Skinner RD (1987a) The mesencephalic locomotor region. I. Activation of

a medullary projection site. Brain Res 411:1-12.

Garcia-Rill E, Skinner RD (1987b) The mesencephalic locomotor region. II. Projections

to reticulospinal neurons. Brain Res 411:13-20.

Garcia-Rill E, Skinner RD, Jackson MB, Smith MM (1983) Connections of the

mesencephalic locomotor region (MLR) I. Substantia nigra afferents. Brain Res Bull

10:57-62.

Gardiner TW, Kitai ST (1992) Single-unit activity in the globus pallidus and neostriatum

of the rat during performance of a trained head movement. Exp Brain Res 88:517-530.

Gasser T (2007) Update on the genetics of Parkinson's disease. Mov Disord 22:S343-

S350.

Gaynor LM, Kuhn AA, Dileone M, Litvak V, Eusebio A, Pogosyan A, Androulidakis

AG, Tisch S, Limousin P, Insola A, Mazzone P, Di L, V, Brown P (2008) Suppression of

beta oscillations in the subthalamic nucleus following cortical stimulation in humans. Eur

J Neurosci 28:1686-1695.

Georgopoulos AP, DeLong MR, Crutcher MD (1983) Relations between parameters of

step-tracking movements and single cell discharge in the globus pallidus and subthalamic

nucleus of the behaving monkey. J Neurosci 3:1586-1598.

Gerfen CR (1985) The neostriatal mosaic. I. Compartmental organization of projections

from the striatum to the substantia nigra in the rat. J Comp Neurol 236:454-476.

Gerfen CR (1984) The neostriatal mosaic: compartmentalization of corticostriatal input

and striatonigral output systems. Nature 311:461-464.

Gerfen CR (1988) Synaptic organization of the striatum. J Electron Microsc Tech

10:265-281.

Gerfen CR, Engber TM, Mahan LC, Susel Z, Chase TN, Monsma FJ, Jr., Sibley DR

(1990) D1 and D2 dopamine receptor-regulated gene expression of striatonigral and

striatopallidal neurons. Science 250:1429-1432.

Page 225: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

213

Gerfen CR, Keefe KA, Gauda EB (1995) D1 and D2 dopamine receptor function in the

striatum: coactivation of D1- and D2-dopamine receptors on separate populations of

neurons results in potentiated immediate early gene response in D1-containing neurons. J

Neurosci 15:8167-8176.

Gerfen CR, Staines WA, Arbuthnott GW, Fibiger HC (1982) Crossed connections of the

substantia nigra in the rat. J Comp Neurol 207:283-303.

Gerfen CR, Wilson CJ (1996) The Basal Ganglia. In: The Handbook of Chemical

Neuroanatomy, Integrated Systems of the CNS, Part III (Swanson LW, Bjorklund A,

Hokfelt T, eds), pp 371-468. Amsterdam: Elsevier.

Gertler TS, Chan CS, Surmeier DJ (2008) Dichotomous anatomical properties of adult

striatal medium spiny neurons. J Neurosci 28:10814-10824.

Ghika J, Villemure JG, Fankhauser H, Favre J, Assal G, Ghika-Schmid F (1998)

Efficiency and safety of bilateral contemporaneous pallidal stimulation (deep brain

stimulation) in levodopa-responsive patients with Parkinson's disease with severe motor

fluctuations: a 2-year follow-up review. J Neurosurg 89:713-718.

Gibb WR, Lees AJ (1988) The relevance of the Lewy body to the pathogenesis of

idiopathic Parkinson's disease. J Neurol Neurosurg Psychiatry 51:745-752.

Gill SS, Heywood P (1997) Bilateral dorsolateral subthalamotomy for advanced

Parkinson's disease. Lancet 350:1224.

Girault JA, Greengard P (2004) The neurobiology of dopamine signaling. Arch Neurol

61:641-644.

Giuffrida R, Li VG, Maugeri G, Perciavalle V (1985) Influences of pyramidal tract on the

subthalamic nucleus in the cat. Neurosci Lett 54:231-235.

Gjerstad MD, Wentzel-Larsen T, Aarsland D, Larsen JP (2007) Insomnia in Parkinson's

disease: frequency and progression over time. J Neurol Neurosurg Psychiatry 78:476-479.

Goldberg JA, Boraud T, Maraton S, Haber SN, Vaadia E, Bergman H (2002) Enhanced

synchrony among primary motor cortex neurons in the 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine primate model of Parkinson's disease. J Neurosci 22:4639-4653.

Goldberg JA, Rokni U, Boraud T, Vaadia E, Bergman H (2004) Spike synchronization in

the cortex/basal-ganglia networks of Parkinsonian primates reflects global dynamics of

the local field potentials. J Neurosci 24:6003-6010.

Goldsmith M, van der Kooy D (1988) Separate non-cholinergic descending projections

and cholinergic ascending projections from the nucleus tegmenti pedunculopontinus.

Brain Res 445:386-391.

Gould E, Butcher LL (1986) Cholinergic neurons in the rat substantia nigra. Neurosci

Page 226: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

214

Lett 63:315-319.

Granata AR, Kitai ST (1989) Intracellular analysis of excitatory subthalamic inputs to the

pedunculopontine neurons. Brain Res 488:57-72.

Granata AR, Kitai ST (1991) Inhibitory substantia nigra inputs to the pedunculopontine

neurons. Exp Brain Res 86:459-466.

Graybiel AM (1990) Neurotransmitters and neuromodulators in the basal ganglia. Trends

Neurosci 13:244-254.

Graybiel AM, Aosaki T, Flaherty AW, Kimura M (1994) The basal ganglia and adaptive

motor control. Science 265:1826-1831.

Graybiel AM, Chesselet MF (1984) Compartmental distribution of striatal cell bodies

expressing [Met]enkephalin-like immunoreactivity. Proc Natl Acad Sci U S A 81:7980-

7984.

Graybiel AM, Ragsdale CW, Jr. (1978) Histochemically distinct compartments in the

striatum of human, monkeys, and cat demonstrated by acetylthiocholinesterase staining.

Proc Natl Acad Sci U S A 75:5723-5726.

Groenewegen HJ, Berendse HW (1990) Connections of the subthalamic nucleus with

ventral striatopallidal parts of the basal ganglia in the rat. J Comp Neurol 294:607-622.

Gross C, Rougier A, Guehl D, Boraud T, Julien J, Bioulac B (1997) High-frequency

stimulation of the globus pallidus internalis in Parkinson's disease: a study of seven cases.

J Neurosurg 87:491-498.

Gross CE, Boraud T, Guehl D, Bioulac B, Bezard E (1999) From experimentation to the

surgical treatment of Parkinson's disease: prelude or suite in basal ganglia research? Prog

Neurobiol 59:509-532.

Grunwerg BS, Krein H, Krauthamer GM (1992) Somatosensory input and thalamic

projection of pedunculopontine tegmental neurons. Neuroreport 3:673-675.

Guehl D, Pessiglione M, Francois C, Yelnik J, Hirsch EC, Feger J, Tremblay L (2003)

Tremor-related activity of neurons in the 'motor' thalamus: changes in firing rate and

pattern in the MPTP vervet model of parkinsonism. Eur J Neurosci 17:2388-2400.

Gulley JM, Kosobud AE, Rebec GV (2002) Behavior-related modulation of substantia

nigra pars reticulata neurons in rats performing a conditioned reinforcement task.

Neuroscience 111:337-349.

Gulley JM, Kuwajima M, Mayhill E, Rebec GV (1999) Behavior-related changes in the

activity of substantia nigra pars reticulata neurons in freely moving rats. Brain Res

845:68-76.

Page 227: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

215

Gulley RL, Wood RL (1971) The fine structure of the neurons in the rat surstantia nigra.

Tissue Cell 3:675-690.

Guridi J, Herrero MT, Luquin MR, Guillen J, Ruberg M, Laguna J, Vila M, Javoy-Agid F,

Agid Y, Hirsch E, Obeso JA (1996) Subthalamotomy in parkinsonian monkeys.

Behavioural and biochemical analysis. Brain 119 ( Pt 5):1717-1727.

Guridi J, Herrero MT, Luquin R, Guillen J, Obeso JA (1994) Subthalamotomy improves

MPTP-induced parkinsonism in monkeys. Stereotact Funct Neurosurg 62:98-102.

Guridi J, Luquin MR, Herrero MT, Obeso JA (1993) The subthalamic nucleus: a possible

target for stereotaxic surgery in Parkinson's disease. Mov Disord 8:421-429.

Gurley K, Kijewski T, Kareem A (2003) First- and higher-order correlation detection

using wavelet transforms. J Eng Mechanics 129:188-201.

Hadar U, Rose FC (1993) Is parkinsonian arm tremor a resting tremor? Eur Neurol

33:221-228.

Hahn PJ, Russo GS, Hashimoto T, Miocinovic S, Xu W, McIntyre CC, Vitek JL (2008)

Pallidal burst activity during therapeutic deep brain stimulation. Exp Neurol 211:243-251.

Hallanger AE, Levey AI, Lee HJ, Rye DB, Wainer BH (1987) The origins of cholinergic

and other subcortical afferents to the thalamus in the rat. J Comp Neurol 262:105-124.

Hallanger AE, Wainer BH (1988) Ascending projections from the pedunculopontine

tegmental nucleus and the adjacent mesopontine tegmentum in the rat. J Comp Neurol

274:483-515.

Hallett M (2000) Clinical physiology of dopa dyskinesia. Ann Neurol 47:S147-S150.

Hallett M, Khoshbin S (1980) A physiological mechanism of bradykinesia. Brain

103:301-314.

Halliday DM, Rosenberg JR, Amjad AM, Breeze P, Conway BA, Farmer SF (1995) A

framework for the analysis of mixed time series/point process data--theory and

application to the study of physiological tremor, single motor unit discharges and

electromyograms. Prog Biophys Mol Biol 64:237-278.

Hamada I, DeLong MR (1992) Excitotoxic acid lesions of the primate subthalamic

nucleus result in reduced pallidal neuronal activity during active holding. J Neurophysiol

68:1859-1866.

Hamada I, DeLong MR, Mano N (1990) Activity of identified wrist-related pallidal

neurons during step and ramp wrist movements in the monkey. J Neurophysiol 64:1892-

1906.

Hamani C, Saint-Cyr JA, Fraser J, Kaplitt M, Lozano AM (2004) The subthalamic

Page 228: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

216

nucleus in the context of movement disorders. Brain 127:4-20.

Hammond C, Bergman H, Brown P (2007) Pathological synchronization in Parkinson's

disease: networks, models and treatments. Trends Neurosci 30:357-364.

Hammond C, Rouzaire-Dubois B, Feger J, Jackson A, Crossman AR (1983) Anatomical

and electrophysiological studies on the reciprocal projections between the subthalamic

nucleus and nucleus tegmenti pedunculopontinus in the rat. Neuroscience 9:41-52.

Harnois C, Filion M (1982) Pallidofugal projections to thalamus and midbrain: a

quantitative antidromic activation study in monkeys and cats. Exp Brain Res 47:277-285.

Harris CM, Wolpert DM (1998) Signal-dependent noise determines motor planning.

Nature 394:780-784.

Hashimoto T, Elder CM, Okun MS, Patrick SK, Vitek JL (2003) Stimulation of the

subthalamic nucleus changes the firing pattern of pallidal neurons. J Neurosci 23:1916-

1923.

Hashimoto T, Tada T, Nakazato F, Maruyama T, Katai S, Izumi Y, Yamada Y, Ikeda S

(2001) Abnormal activity in the globus pallidus in off-period dystonia. Ann Neurol

49:242-245.

Haslinger B, Erhard P, Kampfe N, Boecker H, Rummeny E, Schwaiger M, Conrad B,

Ceballos-Baumann AO (2001) Event-related functional magnetic resonance imaging in

Parkinson's disease before and after levodopa. Brain 124:558-570.

Hassani OK, Feger J (1999) Effects of intrasubthalamic injection of dopamine receptor

agonists on subthalamic neurons in normal and 6-hydroxydopamine-lesioned rats: an

electrophysiological and c-Fos study. Neuroscience 92:533-543.

Hassani OK, Francois C, Yelnik J, Feger J (1997) Evidence for a dopaminergic

innervation of the subthalamic nucleus in the rat. Brain Res 749:88-94.

Hassani OK, Mouroux M, Feger J (1996) Increased subthalamic neuronal activity after

nigral dopaminergic lesion independent of disinhibition via the globus pallidus.

Neuroscience 72:105-115.

HASSLER R, RIECHERT T, MUNDINGER F, UMBACH W, GANGLBERGER JA

(1960) Physiological observations in stereotaxic operations in extrapyramidal motor

disturbances. Brain 83:337-350.

Hattori T (1993) Conceptual history of the nigrostriatal dopamine system. Neurosci Res

16:239-262.

Hazrati LN, Parent A (1992a) Convergence of subthalamic and striatal efferents at

pallidal level in primates: an anterograde double-labeling study with biocytin and PHA-L.

Brain Res 569:336-340.

Page 229: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

217

Hazrati LN, Parent A (1992b) Differential patterns of arborization of striatal and

subthalamic fibers in the two pallidal segments in primates. Brain Res 598:311-315.

Hazrati LN, Parent A (1992c) The striatopallidal projection displays a high degree of

anatomical specificity in the primate. Brain Res 592:213-227.

Hedreen JC (1999) Tyrosine hydroxylase-immunoreactive elements in the human globus

pallidus and subthalamic nucleus. J Comp Neurol 409:400-410.

Hedreen JC, DeLong MR (1991) Organization of striatopallidal, striatonigral, and

nigrostriatal projections in the macaque. J Comp Neurol 304:569-595.

Heilman KM, Bowers D, Watson RT, Greer M (1976) Reaction times in Parkinson

disease. Arch Neurol 33:139-140.

Heimer G, Bar-Gad I, Goldberg JA, Bergman H (2002) Dopamine replacement therapy

reverses abnormal synchronization of pallidal neurons in the 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine primate model of parkinsonism. J Neurosci 22:7850-7855.

Heimer G, Rivlin M, Israel Z, Bergman H (2006) Synchronizing activity of basal ganglia

and pathophysiology of Parkinson's disease. J Neural Transm Suppl17-20.

Hellwig B, Haussler S, Lauk M, Guschlbauer B, Koster B, Kristeva-Feige R, Timmer J,

Lucking CH (2000) Tremor-correlated cortical activity detected by

electroencephalography. Clin Neurophysiol 111:806-809.

Hely MA, Morris JG, Reid WG, Trafficante R (2005) Sydney Multicenter Study of

Parkinson's disease: non-L-dopa-responsive problems dominate at 15 years. Mov Disord

20:190-199.

Henderson JM, Carpenter K, Cartwright H, Halliday GM (2000) Degeneration of the

centre median-parafascicular complex in Parkinson's disease. Ann Neurol 47:345-352.

Herkenham M, Pert CB (1981) Mosaic distribution of opiate receptors, parafascicular

projections and acetylcholinesterase in rat striatum. Nature 291:415-418.

Herrero MT, Barcia C, Navarro JM (2002) Functional anatomy of thalamus and basal

ganglia. Childs Nerv Syst 18:386-404.

Herzog J, Fietzek U, Hamel W, Morsnowski A, Steigerwald F, Schrader B, Weinert D,

Pfister G, Muller D, Mehdorn HM, Deuschl G, Volkmann J (2004) Most effective

stimulation site in subthalamic deep brain stimulation for Parkinson's disease. Mov

Disord 19:1050-1054.

Hikosaka O, Takikawa Y, Kawagoe R (2000) Role of the basal ganglia in the control of

purposive saccadic eye movements. Physiol Rev 80:953-978.

Hikosaka O, Wurtz RH (1983c) Visual and oculomotor functions of monkey substantia

Page 230: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

218

nigra pars reticulata. I. Relation of visual and auditory responses to saccades. J

Neurophysiol 49:1230-1253.

Hikosaka O, Wurtz RH (1983a) Visual and oculomotor functions of monkey substantia

nigra pars reticulata. II. Visual responses related to fixation of gaze. J Neurophysiol

49:1254-1267.

Hikosaka O, Wurtz RH (1983b) Visual and oculomotor functions of monkey substantia

nigra pars reticulata. III. Memory-contingent visual and saccade responses. J

Neurophysiol 49:1268-1284.

Hikosaka O, Wurtz RH (1989) The basal ganglia. Rev Oculomot Res 3:257-281.

Hirsch EC, Faucheux BA (1998) Iron metabolism and Parkinson's disease. Mov Disord

13 Suppl 1:39-45.

Hirsch EC, Graybiel AM, Duyckaerts C, Javoy-Agid F (1987) Neuronal loss in the

pedunculopontine tegmental nucleus in Parkinson disease and in progressive

supranuclear palsy. Proc Natl Acad Sci U S A 84:5976-5980.

Homberg V, Hefter H, Reiners K, Freund HJ (1987) Differential effects of changes in

mechanical limb properties on physiological and pathological tremor. J Neurol Neurosurg

Psychiatry 50:568-579.

Honda T, Semba K (1995) An ultrastructural study of cholinergic and non-cholinergic

neurons in the laterodorsal and pedunculopontine tegmental nuclei in the rat.

Neuroscience 68:837-853.

Hoover JE, Strick PL (1993) Multiple output channels in the basal ganglia. Science

259:819-821.

Horak FB, Anderson ME (1984a) Influence of globus pallidus on arm movements in

monkeys. II. Effects of stimulation. J Neurophysiol 52:305-322.

Horak FB, Anderson ME (1984b) Influence of globus pallidus on arm movements in

monkeys. I. Effects of kainic acid-induced lesions. J Neurophysiol 52:290-304.

Hornykiewicz O (1966) Dopamine (3-hydroxytyramine) and brain function. Pharmacol

Rev 18:925-964.

Hornykiewicz O, Kish SJ (1987) Biochemical pathophysiology of Parkinson's disease.

Adv Neurol 45:19-34.

Hubble JP, Busenbark KL, Wilkinson S, Pahwa R, Paulson GW, Lyons K, Koller WC

(1997) Effects of thalamic deep brain stimulation based on tremor type and diagnosis.

Mov Disord 12:337-341.

Hughes AJ, Daniel SE, Blankson S, Lees AJ (1993) A clinicopathologic study of 100

Page 231: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

219

cases of Parkinson's disease. Arch Neurol 50:140-148.

Hurtado JM, Gray CM, Tamas LB, Sigvardt KA (1999) Dynamics of tremor-related

oscillations in the human globus pallidus: a single case study. Proc Natl Acad Sci U S A

96:1674-1679.

Hurtado JM, Lachaux JP, Beckley DJ, Gray CM, Sigvardt KA (2000) Inter- and intralimb

oscillator coupling in parkinsonian tremor. Mov Disord 15:683-691.

Hurtado JM, Rubchinsky LL, Sigvardt KA, Wheelock VL, Pappas CT (2005) Temporal

evolution of oscillations and synchrony in GPi/muscle pairs in Parkinson's disease. J

Neurophysiol 93:1569-1584.

Hutchinson WD, Levy R, Dostrovsky JO, Lozano AM, Lang AE (1997) Effects of

apomorphine on globus pallidus neurons in parkinsonian patients. Ann Neurol 42:767-

775.

Hutchison WD, Allan RJ, Opitz H, Levy R, Dostrovsky JO, Lang AE, Lozano AM (1998)

Neurophysiological identification of the subthalamic nucleus in surgery for Parkinson's

disease. Ann Neurol 44:622-628.

Hutchison WD, Lang AE, Dostrovsky JO, Lozano AM (2003) Pallidal neuronal activity:

implications for models of dystonia. Ann Neurol 53:480-488.

Hutchison WD, Lozano AM, Davis KD, Saint-Cyr JA, Lang AE, Dostrovsky JO (1994)

Differential neuronal activity in segments of globus pallidus in Parkinson's disease

patients. Neuroreport 5:1533-1537.

Hutchison WD, Lozano AM, Tasker RR, Lang AE, Dostrovsky JO (1997) Identification

and characterization of neurons with tremor-frequency activity in human globus pallidus.

Exp Brain Res 113:557-563.

Hylden JL, Hayashi H, Bennett GJ, Dubner R (1985) Spinal lamina I neurons projecting

to the parabrachial area of the cat midbrain. Brain Res 336:195-198.

Ilinsky IA, Yi H, Kultas-Ilinsky K (1997) Mode of termination of pallidal afferents to the

thalamus: a light and electron microscopic study with anterograde tracers and

immunocytochemistry in Macaca mulatta. J Comp Neurol 386:601-612.

Inase M, Tanji J (1995) Thalamic distribution of projection neurons to the primary motor

cortex relative to afferent terminal fields from the globus pallidus in the macaque monkey.

J Comp Neurol 353:415-426.

Inglis WL, Allen LF, Whitelaw RB, Latimer MP, Brace HM, Winn P (1994a) An

investigation into the role of the pedunculopontine tegmental nucleus in the mediation of

locomotion and orofacial stereotypy induced by d-amphetamine and apomorphine in the

rat. Neuroscience 58:817-833.

Page 232: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

220

Inglis WL, Dunbar JS, Winn P (1994b) Outflow from the nucleus accumbens to the

pedunculopontine tegmental nucleus: a dissociation between locomotor activity and the

acquisition of responding for conditioned reinforcement stimulated by d-amphetamine.

Neuroscience 62:51-64.

Inglis WL, Winn P (1995) The pedunculopontine tegmental nucleus: where the striatum

meets the reticular formation. Prog Neurobiol 47:1-29.

Iwahori N (1978) A Golgi study on the subthalamic nucleus of the cat. J Comp Neurol

182:383-397.

Jackson A, Crossman AR (1983) Nucleus tegmenti pedunculopontinus: efferent

connections with special reference to the basal ganglia, studied in the rat by anterograde

and retrograde transport of horseradish peroxidase. Neuroscience 10:725-765.

Jackson A, Crossman AR (1981) Subthalamic projection to nucleus tegmenti

pedunculopontinus in the rat. Neurosci Lett 22:17-22.

Jaeger D, Kita H, Wilson CJ (1994) Surround inhibition among projection neurons is

weak or nonexistent in the rat neostriatum. J Neurophysiol 72:2555-2558.

Jahanshahi M, Ardouin CM, Brown RG, Rothwell JC, Obeso J, Albanese A, Rodriguez-

Oroz MC, Moro E, Benabid AL, Pollak P, Limousin-Dowsey P (2000) The impact of

deep brain stimulation on executive function in Parkinson's disease. Brain 123 ( Pt

6):1142-1154.

Jahanshahi M, Jenkins IH, Brown RG, Marsden CD, Passingham RE, Brooks DJ (1995)

Self-initiated versus externally triggered movements. I. An investigation using

measurement of regional cerebral blood flow with PET and movement-related potentials

in normal and Parkinson's disease subjects. Brain 118 ( Pt 4):913-933.

Jankovic J (2008) Parkinson's disease: clinical features and diagnosis. J Neurol

Neurosurg Psychiatry 79:368-376.

Jarvis MR, Mitra PP (2001) Sampling properties of the spectrum and coherency of

sequences of action potentials. Neural Comput 13:717-749.

Jellinger K (1988) The pedunculopontine nucleus in Parkinson's disease, progressive

supranuclear palsy and Alzheimer's disease. J Neurol Neurosurg Psychiatry 51:540-543.

Jellinger KA (1991) Pathology of Parkinson's disease. Changes other than the

nigrostriatal pathway. Mol Chem Neuropathol 14:153-197.

Jellinger KA (1999) Post mortem studies in Parkinson's disease--is it possible to detect

brain areas for specific symptoms? J Neural Transm Suppl 56:1-29.

Jenkins IH, Fernandez W, Playford ED, Lees AJ, Frackowiak RS, Passingham RE,

Brooks DJ (1992) Impaired activation of the supplementary motor area in Parkinson's

Page 233: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

221

disease is reversed when akinesia is treated with apomorphine. Ann Neurol 32:749-757.

Jenkinson N, Nandi D, Miall RC, Stein JF, Aziz TZ (2004) Pedunculopontine nucleus

stimulation improves akinesia in a Parkinsonian monkey. Neuroreport 15:2621-2624.

Joel D, Weiner I (1997) The connections of the primate subthalamic nucleus: indirect

pathways and the open-interconnected scheme of basal ganglia-thalamocortical circuitry.

Brain Res Brain Res Rev 23:62-78.

Joel D, Weiner I (1994) The organization of the basal ganglia-thalamocortical circuits:

open interconnected rather than closed segregated. Neuroscience 63:363-379.

Joseph JP, Boussaoud D (1985) Role of the cat substantia nigra pars reticulata in eye and

head movements. I. Neural activity. Exp Brain Res 57:286-296.

Joseph JP, Boussaoud D, Biguer B (1985) Activity of neurons in the cat substantia nigra

pars reticulata during drinking. Exp Brain Res 60:375-379.

Junghanns S, Glockler T, Reichmann H (2004) Switching and combining of dopamine

agonists. J Neurol 251 Suppl 6:VI/19-VI/23.

Juraska JM, Wilson CJ, Groves PM (1977) The substantia nigra of the rat: a Golgi study.

J Comp Neurol 172:585-600.

Kaiser J, Lutzenberger W (2005) Human gamma-band activity: a window to cognitive

processing. Neuroreport 16:207-211.

Kaneoke Y, Vitek JL (1996) Burst and oscillation as disparate neuronal properties. J

Neurosci Methods 68:211-223.

Kang Y, Kitai ST (1990) Electrophysiological properties of pedunculopontine neurons

and their postsynaptic responses following stimulation of substantia nigra reticulata.

Brain Res 535:79-95.

Kato M, Kimura M (1992) Effects of reversible blockade of basal ganglia on a voluntary

arm movement. J Neurophysiol 68:1516-1534.

Kawaguchi Y, Wilson CJ, Emson PC (1990) Projection subtypes of rat neostriatal matrix

cells revealed by intracellular injection of biocytin. J Neurosci 10:3421-3438.

Kemp JM, Powell TP (1970) The cortico-striate projection in the monkey. Brain 93:525-

546.

Kemp JM, Powell TP (1971) The structure of the caudate nucleus of the cat: light and

electron microscopy. Philos Trans R Soc Lond B Biol Sci 262:383-401.

Kempf F, Kuhn AA, Kupsch A, Brucke C, Weise L, Schneider GH, Brown P (2007)

Premovement activities in the subthalamic area of patients with Parkinson's disease and

Page 234: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

222

their dependence on task. Eur J Neurosci 25:3137-3145.

Kim R, Nakano K, Jayaraman A, Carpenter MB (1976) Projections of the globus pallidus

and adjacent structures: an autoradiographic study in the monkey. J Comp Neurol

169:263-290.

Kimura M (1986) The role of primate putamen neurons in the association of sensory

stimuli with movement. Neurosci Res 3:436-443.

Kimura M, Yamada H, Matsumoto N (2003) Tonically active neurons in the striatum

encode motivational contexts of action. Brain Dev 25 Suppl 1:S20-S23.

Kish LJ, Palmer MR, Gerhardt GA (1999) Multiple single-unit recordings in the striatum

of freely moving animals: effects of apomorphine and D-amphetamine in normal and

unilateral 6-hydroxydopamine-lesioned rats. Brain Res 833:58-70.

Kishore A, Turnbull IM, Snow BJ, de lF-F, Schulzer M, Mak E, Yardley S, Calne DB

(1997) Efficacy, stability and predictors of outcome of pallidotomy for Parkinson's

disease. Six-month follow-up with additional 1-year observations. Brain 120 ( Pt 5):729-

737.

Kissler J, Muller MM, Fehr T, Rockstroh B, Elbert T (2000) MEG gamma band activity

in schizophrenia patients and healthy subjects in a mental arithmetic task and at rest. Clin

Neurophysiol 111:2079-2087.

Kita H (2007) Globus pallidus external segment. Prog Brain Res 160:111-133.

Kita H, Chang HT, Kitai ST (1983a) Pallidal inputs to subthalamus: intracellular analysis.

Brain Res 264:255-265.

Kita H, Chang HT, Kitai ST (1983b) The morphology of intracellularly labeled rat

subthalamic neurons: a light microscopic analysis. J Comp Neurol 215:245-257.

Kita H, Kitai ST (1987) Efferent projections of the subthalamic nucleus in the rat: light

and electron microscopic analysis with the PHA-L method. J Comp Neurol 260:435-452.

Kita H, Tachibana Y, Nambu A, Chiken S (2005) Balance of monosynaptic excitatory

and disynaptic inhibitory responses of the globus pallidus induced after stimulation of the

subthalamic nucleus in the monkey. J Neurosci 25:8611-8619.

Kitai ST, Deniau JM (1981) Cortical inputs to the subthalamus: intracellular analysis.

Brain Res 214:411-415.

Kleiner-Fisman G, Fisman DN, Sime E, Saint-Cyr JA, Lozano AM, Lang AE (2003)

Long-term follow up of bilateral deep brain stimulation of the subthalamic nucleus in

patients with advanced Parkinson disease. J Neurosurg 99:489-495.

Klockgether T (2004) Parkinson's disease: clinical aspects. Cell Tissue Res 318:115-120.

Page 235: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

223

Klockgether T, Turski L, Honore T, Zhang ZM, Gash DM, Kurlan R, Greenamyre JT

(1991) The AMPA receptor antagonist NBQX has antiparkinsonian effects in

monoamine-depleted rats and MPTP-treated monkeys. Ann Neurol 30:717-723.

Kobayashi Y, Inoue Y, Yamamoto M, Isa T, Aizawa H (2002) Contribution of

pedunculopontine tegmental nucleus neurons to performance of visually guided saccade

tasks in monkeys. J Neurophysiol 88:715-731.

Kojima J, Yamaji Y, Matsumura M, Nambu A, Inase M, Tokuno H, Takada M, Imai H

(1997) Excitotoxic lesions of the pedunculopontine tegmental nucleus produce

contralateral hemiparkinsonism in the monkey. Neurosci Lett 226:111-114.

Koller W, Pahwa R, Busenbark K, Hubble J, Wilkinson S, Lang A, Tuite P, Sime E,

Lazano A, Hauser R, Malapira T, Smith D, Tarsy D, Miyawaki E, Norregaard T, Kormos

T, Olanow CW (1997) High-frequency unilateral thalamic stimulation in the treatment of

essential and parkinsonian tremor. Ann Neurol 42:292-299.

Koller WC, Montgomery EB (1997) Issues in the early diagnosis of Parkinson's disease.

Neurology 49:S10-S25.

Kopin IJ, Markey SP (1988) MPTP toxicity: implications for research in Parkinson's

disease. Annu Rev Neurosci 11:81-96.

Krack P, Batir A, Van BN, Chabardes S, Fraix V, Ardouin C, Koudsie A, Limousin PD,

Benazzouz A, LeBas JF, Benabid AL, Pollak P (2003) Five-year follow-up of bilateral

stimulation of the subthalamic nucleus in advanced Parkinson's disease. N Engl J Med

349:1925-1934.

Krack P, Pollak P, Limousin P, Benazzouz A, Benabid AL (1997) Stimulation of

subthalamic nucleus alleviates tremor in Parkinson's disease. Lancet 350:1675.

Krack P, Pollak P, Limousin P, Hoffmann D, Xie J, Benazzouz A, Benabid AL (1998)

Subthalamic nucleus or internal pallidal stimulation in young onset Parkinson's disease.

Brain 121 ( Pt 3):451-457.

Kreiss DS, Anderson LA, Walters JR (1996) Apomorphine and dopamine D(1) receptor

agonists increase the firing rates of subthalamic nucleus neurons. Neuroscience 72:863-

876.

Kreiss DS, Mastropietro CW, Rawji SS, Walters JR (1997) The response of subthalamic

nucleus neurons to dopamine receptor stimulation in a rodent model of Parkinson's

disease. J Neurosci 17:6807-6819.

Kuhn AA, Doyle L, Pogosyan A, Yarrow K, Kupsch A, Schneider GH, Hariz MI,

Trottenberg T, Brown P (2006a) Modulation of beta oscillations in the subthalamic area

during motor imagery in Parkinson's disease. Brain 129:695-706.

Kuhn AA, Kempf F, Brucke C, Gaynor DL, Martinez-Torres I, Pogosyan A, Trottenberg

Page 236: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

224

T, Kupsch A, Schneider GH, Hariz MI, Vandenberghe W, Nuttin B, Brown P (2008)

High-frequency stimulation of the subthalamic nucleus suppresses oscillatory beta

activity in patients with Parkinson's disease in parallel with improvement in motor

performance. J Neurosci 28:6165-6173.

Kuhn AA, Kupsch A, Schneider GH, Brown P (2006b) Reduction in subthalamic 8-35

Hz oscillatory activity correlates with clinical improvement in Parkinson's disease. Eur J

Neurosci 23:1956-1960.

Kuhn AA, Trottenberg T, Kivi A, Kupsch A, Schneider GH, Brown P (2005) The

relationship between local field potential and neuronal discharge in the subthalamic

nucleus of patients with Parkinson's disease. Exp Neurol 194:212-220.

Kuhn AA, Tsui A, Aziz T, Ray N, Brucke C, Kupsch A, Schneider GH, Brown P (2009)

Pathological synchronisation in the subthalamic nucleus of patients with Parkinson's

disease relates to both bradykinesia and rigidity. Exp Neurol 215:380-387.

Kuhn AA, Williams D, Kupsch A, Limousin P, Hariz M, Schneider GH, Yarrow K,

Brown P (2004) Event-related beta desynchronization in human subthalamic nucleus

correlates with motor performance. Brain 127:735-746.

Kumar K, Kelly M, Toth C (1999a) Deep brain stimulation of the ventral intermediate

nucleus of the thalamus for control of tremors in Parkinson's disease and essential tremor.

Stereotact Funct Neurosurg 72:47-61.

Kumar R, Dagher A, Hutchison WD, Lang AE, Lozano AM (1999b) Globus pallidus

deep brain stimulation for generalized dystonia: clinical and PET investigation.

Neurology 53:871-874.

Kumar R, Lozano AM, Montgomery E, Lang AE (1998) Pallidotomy and deep brain

stimulation of the pallidum and subthalamic nucleus in advanced Parkinson's disease.

Mov Disord 13 Suppl 1:73-82.

Kumar R, Lozano AM, Sime E, Halket E, Lang AE (1999c) Comparative effects of

unilateral and bilateral subthalamic nucleus deep brain stimulation. Neurology 53:561-

566.

Kuo JS, Carpenter MB (1973) Organization of pallidothalamic projections in the rhesus

monkey. J Comp Neurol 151:201-236.

Kus L, Borys E, Ping CY, Ferguson SM, Blakely RD, Emborg ME, Kordower JH, Levey

AI, Mufson EJ (2003) Distribution of high affinity choline transporter immunoreactivity

in the primate central nervous system. J Comp Neurol 463:341-357.

Kutukcu Y, Marks WJ, Jr., Goodin DS, Aminoff MJ (1999) Simple and choice reaction

time in Parkinson's disease. Brain Res 815:367-372.

Lai YY, Siegel JM (1990) Muscle tone suppression and stepping produced by stimulation

Page 237: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

225

of midbrain and rostral pontine reticular formation. J Neurosci 10:2727-2734.

Laitinen LV, Bergenheim AT, Hariz MI (1992b) Ventroposterolateral pallidotomy can

abolish all parkinsonian symptoms. Stereotact Funct Neurosurg 58:14-21.

Laitinen LV, Bergenheim AT, Hariz MI (1992a) Leksell's posteroventral pallidotomy in

the treatment of Parkinson's disease. J Neurosurg 76:53-61.

Lalo E, Thobois S, Sharott A, Polo G, Mertens P, Pogosyan A, Brown P (2008) Patterns

of bidirectional communication between cortex and basal ganglia during movement in

patients with Parkinson disease. J Neurosci 28:3008-3016.

Lang AE, Duff J, Saint-Cyr JA, Trepanier L, Gross RE, Lombardi W, Montgomery E,

Hutchinson W, Lozano AM (1999) Posteroventral medial pallidotomy in Parkinson's

disease. J Neurol 246 Suppl 2:II28-II41.

Lang AE, Lozano AM (1998) Parkinson's disease. First of two parts. N Engl J Med

339:1044-1053.

Lang AE, Lozano AM, Montgomery E, Duff J, Tasker R, Hutchinson W (1997)

Posteroventral medial pallidotomy in advanced Parkinson's disease. N Engl J Med

337:1036-1042.

Langston JW, Ballard P, Tetrud JW, Irwin I (1983) Chronic Parkinsonism in humans due

to a product of meperidine-analog synthesis. Science 219:979-980.

Langston JW, Forno LS, Rebert CS, Irwin I (1984) Selective nigral toxicity after

systemic administration of 1-methyl-4-phenyl-1,2,5,6-tetrahydropyrine (MPTP) in the

squirrel monkey. Brain Res 292:390-394.

Lavoie B, Parent A (1990) Immunohistochemical study of the serotoninergic innervation

of the basal ganglia in the squirrel monkey. J Comp Neurol 299:1-16.

Lavoie B, Parent A (1994a) Pedunculopontine nucleus in the squirrel monkey:

cholinergic and glutamatergic projections to the substantia nigra. J Comp Neurol

344:232-241.

Lavoie B, Parent A (1994b) Pedunculopontine nucleus in the squirrel monkey:

projections to the basal ganglia as revealed by anterograde tract-tracing methods. J Comp

Neurol 344:210-231.

Lavoie B, Parent A (1994c) Pedunculopontine nucleus in the squirrel monkey:

distribution of cholinergic and monoaminergic neurons in the mesopontine tegmentum

with evidence for the presence of glutamate in cholinergic neurons. J Comp Neurol

344:190-209.

Lavoie B, Smith Y, Parent A (1989) Dopaminergic innervation of the basal ganglia in the

squirrel monkey as revealed by tyrosine hydroxylase immunohistochemistry. J Comp

Page 238: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

226

Neurol 289:36-52.

Le Moine C, Bloch B (1995) D1 and D2 dopamine receptor gene expression in the rat

striatum: sensitive cRNA probes demonstrate prominent segregation of D1 and D2

mRNAs in distinct neuronal populations of the dorsal and ventral striatum. J Comp

Neurol 355:418-426.

Leblois A, Boraud T, Meissner W, Bergman H, Hansel D (2006a) Competition between

feedback loops underlies normal and pathological dynamics in the basal ganglia. J

Neurosci 26:3567-3583.

Leblois A, Meissner W, Bezard E, Bioulac B, Gross CE, Boraud T (2006b) Temporal and

spatial alterations in GPi neuronal encoding might contribute to slow down movement in

Parkinsonian monkeys. Eur J Neurosci 24:1201-1208.

Leblois A, Meissner W, Bioulac B, Gross CE, Hansel D, Boraud T (2007) Late

emergence of synchronized oscillatory activity in the pallidum during progressive

Parkinsonism. Eur J Neurosci 26:1701-1713.

Lee HJ, Rye DB, Hallanger AE, Levey AI, Wainer BH (1988) Cholinergic vs.

noncholinergic efferents from the mesopontine tegmentum to the extrapyramidal motor

system nuclei. J Comp Neurol 275:469-492.

Lee JI, Shin HJ, Nam DH, Kim JS, Hong SC, Shin HJ, Park K, Eoh W, Kim JH, Lee WY

(2001) Increased burst firing in substantia nigra pars reticulata neurons and enhanced

response to selective D2 agonist in hemiparkinsonian rats after repeated administration of

apomorphine. J Korean Med Sci 16:636-642.

Lee RG (1989) Pathophysiology of rigidity and akinesia in Parkinson's disease. Eur

Neurol 29 Suppl 1:13-18.

Lee RG, Stein RB (1981) Resetting of tremor by mechanical perturbations: a comparison

of essential tremor and parkinsonian tremor. Ann Neurol 10:523-531.

Lee RG, Tatton WG (1975) Motor responses to sudden limb displacements in primates

with specific CNS lesions and in human patients with motor system disorders. Can J

Neurol Sci 2:285-293.

Lemaire JJ, Coste J, Ouchchane L, Caire F, Nuti C, Derost P, Cristini V, Gabrillargues J,

Hemm S, Durif F, Chazal J (2007) Brain mapping in stereotactic surgery: a brief

overview from the probabilistic targeting to the patient-based anatomic mapping.

Neuroimage 37 Suppl 1:S109-S115.

Lemstra AW, Verhagen ML, Lee JI, Dougherty PM, Lenz FA (1999) Tremor-frequency

(3-6 Hz) activity in the sensorimotor arm representation of the internal segment of the

globus pallidus in patients with Parkinson's disease. Neurosci Lett 267:129-132.

Lenz FA, Jaeger CJ, Seike MS, Lin YC, Reich SG, DeLong MR, Vitek JL (1999)

Page 239: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

227

Thalamic single neuron activity in patients with dystonia: dystonia-related activity and

somatic sensory reorganization. J Neurophysiol 82:2372-2392.

Lenz FA, Kwan HC, Martin RL, Tasker RR, Dostrovsky JO, Lenz YE (1994) Single unit

analysis of the human ventral thalamic nuclear group. Tremor-related activity in

functionally identified cells. Brain 117 ( Pt 3):531-543.

Lenz FA, Suarez JI, Metman LV, Reich SG, Karp BI, Hallett M, Rowland LH,

Dougherty PM (1998) Pallidal activity during dystonia: somatosensory reorganisation

and changes with severity. J Neurol Neurosurg Psychiatry 65:767-770.

Lenz FA, Tasker RR, Kwan HC, Schnider S, Kwong R, Murayama Y, Dostrovsky JO,

Murphy JT (1988) Single unit analysis of the human ventral thalamic nuclear group:

correlation of thalamic "tremor cells" with the 3-6 Hz component of parkinsonian tremor.

J Neurosci 8:754-764.

Lenz FA, Vitek JL, DeLong MR (1993) Role of the thalamus in parkinsonian tremor:

evidence from studies in patients and primate models. Stereotact Funct Neurosurg 60:94-

103.

Leocani L, Toro C, Manganotti P, Zhuang P, Hallett M (1997) Event-related coherence

and event-related desynchronization/synchronization in the 10 Hz and 20 Hz EEG during

self-paced movements. Electroencephalogr Clin Neurophysiol 104:199-206.

Lestienne F, Caillier P (1986) Role of the monkey substantia nigra pars reticulata in

orienting behaviour and visually triggered arm movements. Neurosci Lett 64:109-115.

Levesque JC, Parent A (2005a) GABAergic interneurons in human subthalamic nucleus.

Mov Disord 20:574-584.

Levesque M, Parent A (2005b) The striatofugal fiber system in primates: a reevaluation

of its organization based on single-axon tracing studies. Proc Natl Acad Sci U S A

102:11888-11893.

Levy R, Ashby P, Hutchison WD, Lang AE, Lozano AM, Dostrovsky JO (2002a)

Dependence of subthalamic nucleus oscillations on movement and dopamine in

Parkinson's disease. Brain 125:1196-1209.

Levy R, Dostrovsky JO, Lang AE, Sime E, Hutchison WD, Lozano AM (2001a) Effects

of apomorphine on subthalamic nucleus and globus pallidus internus neurons in patients

with Parkinson's disease. J Neurophysiol 86:249-260.

Levy R, Hutchison WD, Lozano AM, Dostrovsky JO (2000) High-frequency

synchronization of neuronal activity in the subthalamic nucleus of parkinsonian patients

with limb tremor. J Neurosci 20:7766-7775.

Levy R, Hutchison WD, Lozano AM, Dostrovsky JO (2002b) Synchronized neuronal

discharge in the basal ganglia of parkinsonian patients is limited to oscillatory activity. J

Page 240: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

228

Neurosci 22:2855-2861.

Levy R, Lang AE, Dostrovsky JO, Pahapill P, Romas J, Saint-Cyr J, Hutchison WD,

Lozano AM (2001b) Lidocaine and muscimol microinjections in subthalamic nucleus

reverse Parkinsonian symptoms. Brain 124:2105-2118.

Levy R, Lozano AM, Hutchison WD, Dostrovsky JO (2007) Dual microelectrode

technique for deep brain stereotactic surgery in humans. Neurosurgery 60:277-283.

Lew M (2007) Overview of Parkinson's disease. Pharmacotherapy 27:155S-160S.

Liles SL, Updyke BV (1985) Projection of the digit and wrist area of precentral gyrus to

the putamen: relation between topography and physiological properties of neurons in the

putamen. Brain Res 339:245-255.

Limousin P, Greene J, Pollak P, Rothwell J, Benabid AL, Frackowiak R (1997) Changes

in cerebral activity pattern due to subthalamic nucleus or internal pallidum stimulation in

Parkinson's disease. Ann Neurol 42:283-291.

Limousin P, Krack P, Pollak P, Benazzouz A, Ardouin C, Hoffmann D, Benabid AL

(1998) Electrical stimulation of the subthalamic nucleus in advanced Parkinson's disease.

N Engl J Med 339:1105-1111.

Limousin P, Pollak P, Benazzouz A, Hoffmann D, Le Bas JF, Broussolle E, Perret JE,

Benabid AL (1995) Effect of parkinsonian signs and symptoms of bilateral subthalamic

nucleus stimulation. Lancet 345:91-95.

Liu X, Ford-Dunn HL, Hayward GN, Nandi D, Miall RC, Aziz TZ, Stein JF (2002) The

oscillatory activity in the Parkinsonian subthalamic nucleus investigated using the macro-

electrodes for deep brain stimulation. Clin Neurophysiol 113:1667-1672.

Liu X, Wang S, Yianni J, Nandi D, Bain PG, Gregory R, Stein JF, Aziz TZ (2008) The

sensory and motor representation of synchronized oscillations in the globus pallidus in

patients with primary dystonia. Brain 131:1562-1573.

Liu X, Yianni J, Wang S, Bain PG, Stein JF, Aziz TZ (2006) Different mechanisms may

generate sustained hypertonic and rhythmic bursting muscle activity in idiopathic

dystonia. Exp Neurol 198:204-213.

Louis ED, Marder K, Cote L, Tang M, Mayeux R (1997) Mortality from Parkinson

disease. Arch Neurol 54:260-264.

Lozano A, Hutchison W, Kiss Z, Tasker R, Davis K, Dostrovsky J (1996) Methods for

microelectrode-guided posteroventral pallidotomy. J Neurosurg 84:194-202.

Lozano AM (2001) Deep brain stimulation for Parkinson's disease. Parkinsonism Relat

Disord 7:199-203.

Page 241: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

229

Lozano AM (2000) Vim thalamic stimulation for tremor. Arch Med Res 31:266-269.

Lozano AM, Dostrovsky J, Chen R, Ashby P (2002) Deep brain stimulation for

Parkinson's disease: disrupting the disruption. Lancet Neurol 1:225-231.

Lozano AM, Hutchison WD, Tasker RR, Lang AE, Junn F, Dostrovsky JO (1998)

Microelectrode recordings define the ventral posteromedial pallidotomy target. Stereotact

Funct Neurosurg 71:153-163.

Lozano AM, Kumar R, Gross RE, Giladi N, Hutchison WD, Dostrovsky JO, Lang AE

(1997) Globus pallidus internus pallidotomy for generalized dystonia. Mov Disord

12:865-870.

Lozano AM, Lang AE, Galvez-Jimenez N, Miyasaki J, Duff J, Hutchinson WD,

Dostrovsky JO (1995) Effect of GPi pallidotomy on motor function in Parkinson's

disease. Lancet 346:1383-1387.

Lozano AM, Lang AE, Levy R, Hutchison W, Dostrovsky J (2000) Neuronal recordings

in Parkinson's disease patients with dyskinesias induced by apomorphine. Ann Neurol

47:S141-S146.

Lozano AM, Mahant N (2004) Deep brain stimulation surgery for Parkinson's disease:

mechanisms and consequences. Parkinsonism Relat Disord 10 Suppl 1:S49-S57.

Lozza C, Marie RM, Baron JC (2002) The metabolic substrates of bradykinesia and

tremor in uncomplicated Parkinson's disease. Neuroimage 17:688-699.

Lynd-Balta E, Haber SN (1994) Primate striatonigral projections: a comparison of the

sensorimotor-related striatum and the ventral striatum. J Comp Neurol 345:562-578.

Machado A, Rezai AR, Kopell BH, Gross RE, Sharan AD, Benabid AL (2006) Deep

brain stimulation for Parkinson's disease: surgical technique and perioperative

management. Mov Disord 21 Suppl 14:S247-S258.

Magarinos-Ascone CM, Figueiras-Mendez R, Riva-Meana C, Cordoba-Fernandez A

(2000) Subthalamic neuron activity related to tremor and movement in Parkinson's

disease. Eur J Neurosci 12:2597-2607.

Magill PJ, Bolam JP, Bevan MD (2000) Relationship of activity in the subthalamic

nucleus-globus pallidus network to cortical electroencephalogram. J Neurosci 20:820-833.

Magill PJ, Bolam JP, Bevan MD (2001) Dopamine regulates the impact of the cerebral

cortex on the subthalamic nucleus-globus pallidus network. Neuroscience 106:313-330.

Magill PJ, Sharott A, Bevan MD, Brown P, Bolam JP (2004) Synchronous unit activity

and local field potentials evoked in the subthalamic nucleus by cortical stimulation. J

Neurophysiol 92:700-714.

Page 242: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

230

Magnin M, Morel A, Jeanmonod D (2000) Single-unit analysis of the pallidum, thalamus

and subthalamic nucleus in parkinsonian patients. Neuroscience 96:549-564.

Makela JP, Hari P, Karhu J, Salmelin R, Teravainen H (1993) Suppression of magnetic

mu rhythm during parkinsonian tremor. Brain Res 617:189-193.

Mallet N, Ballion B, Le MC, Gonon F (2006) Cortical inputs and GABA interneurons

imbalance projection neurons in the striatum of parkinsonian rats. J Neurosci 26:3875-

3884.

Mallet N, Pogosyan A, Sharott A, Csicsvari J, Bolam JP, Brown P, Magill PJ (2008)

Disrupted dopamine transmission and the emergence of exaggerated beta oscillations in

subthalamic nucleus and cerebral cortex. J Neurosci 28:4795-4806.

Manson AJ, Brown P, O'Sullivan JD, Asselman P, Buckwell D, Lees AJ (2000) An

ambulatory dyskinesia monitor. J Neurol Neurosurg Psychiatry 68:196-201.

Mansour A, Meador-Woodruff JH, Zhou QY, Civelli O, Akil H, Watson SJ (1991) A

comparison of D1 receptor binding and mRNA in rat brain using receptor

autoradiographic and in situ hybridization techniques. Neuroscience 45:359-371.

Marceglia S, Foffani G, Bianchi AM, Baselli G, Tamma F, Egidi M, Priori A (2006)

Dopamine-dependent non-linear correlation between subthalamic rhythms in Parkinson's

disease. J Physiol 571:579-591.

Marras C, Lang A (2008) Invited article: changing concepts in Parkinson disease: moving

beyond the decade of the brain. Neurology 70:1996-2003.

Marsden CD (1984) The pathophysiology of movement disorders. Neurol Clin 2:435-459.

Marsden CD, Jenner PG (1987) The significance of 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine. Ciba Found Symp 126:239-256.

Marsden CD, Obeso JA (1994a) The functions of the basal ganglia and the paradox of

stereotaxic surgery in Parkinson's disease. Brain 117 ( Pt 4):877-897.

Marsden CD, Obeso JA (1994b) The functions of the basal ganglia and the paradox of

stereotaxic surgery in Parkinson's disease. Brain 117 ( Pt 4):877-897.

Marsden CD, Rothwell JC, Day BL (1983) Long-latency automatic responses to muscle

stretch in man: origin and function. Adv Neurol 39:509-539.

Marsden JF, Limousin-Dowsey P, Ashby P, Pollak P, Brown P (2001) Subthalamic

nucleus, sensorimotor cortex and muscle interrelationships in Parkinson's disease. Brain

124:378-388.

Martinez-Murillo R, Villalba R, Montero-Caballero MI, Rodrigo J (1989) Cholinergic

somata and terminals in the rat substantia nigra: an immunocytochemical study with

Page 243: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

231

optical and electron microscopic techniques. J Comp Neurol 281:397-415.

Matsumura M (2005) The pedunculopontine tegmental nucleus and experimental

parkinsonism. A review. J Neurol 252 Suppl 4:IV5-IV12.

Matsumura M, Kojima J (2001) The role of the pedunculopontine tegmental nucleus in

experimental parkinsonism in primates. Stereotact Funct Neurosurg 77:108-115.

Matsumura M, Kojima J, Gardiner TW, Hikosaka O (1992) Visual and oculomotor

functions of monkey subthalamic nucleus. J Neurophysiol 67:1615-1632.

Matsumura M, Nambu A, Yamaji Y, Watanabe K, Imai H, Inase M, Tokuno H, Takada

M (2000) Organization of somatic motor inputs from the frontal lobe to the

pedunculopontine tegmental nucleus in the macaque monkey. Neuroscience 98:97-110.

Matsumura M, Tremblay L, Richard H, Filion M (1995) Activity of pallidal neurons in

the monkey during dyskinesia induced by injection of bicuculline in the external pallidum.

Neuroscience 65:59-70.

Matsumura M, Watanabe K, Ohye C (1997) Single-unit activity in the primate nucleus

tegmenti pedunculopontinus related to voluntary arm movement. Neurosci Res 28:155-

165.

Maurice N, Deniau JM, Menetrey A, Glowinski J, Thierry AM (1998) Prefrontal cortex-

basal ganglia circuits in the rat: involvement of ventral pallidum and subthalamic nucleus.

Synapse 29:363-370.

Mayer RA, Kindt MV, Heikkila RE (1986) Prevention of the nigrostriatal toxicity of 1-

methyl-4-phenyl-1,2,3,6-tetrahydropyridine by inhibitors of 3,4-

dihydroxyphenylethylamine transport. J Neurochem 47:1073-1079.

Mazurek ME, Shadlen MN (2002) Limits to the temporal fidelity of cortical spike rate

signals. Nat Neurosci 5:463-471.

Mazzone P, Lozano A, Stanzione P, Galati S, Scarnati E, Peppe A, Stefani A (2005)

Implantation of human pedunculopontine nucleus: a safe and clinically relevant target in

Parkinson's disease. Neuroreport 16:1877-1881.

McCairn KW, Turner RS (2009) Deep brain stimulation of the globus pallidus internus in

the parkinsonian primate: local entrainment and suppression of low-frequency

oscillations. J Neurophysiol 101:1941-1960.

McKeith IG (2000) Clinical Lewy body syndromes. Ann N Y Acad Sci 920:1-8.

Meissner W, Leblois A, Hansel D, Bioulac B, Gross CE, Benazzouz A, Boraud T (2005)

Subthalamic high frequency stimulation resets subthalamic firing and reduces abnormal

oscillations. Brain 128:2372-2382.

Page 244: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

232

Mena-Segovia J, Bolam JP, Magill PJ (2004) Pedunculopontine nucleus and basal

ganglia: distant relatives or part of the same family? Trends Neurosci 27:585-588.

Mendez I, Elisevich K, Flumerfelt BA (1993) GABAergic synaptic interactions in the

substantia nigra. Brain Res 617:274-284.

Merello M, Balej J, Delfino M, Cammarota A, Betti O, Leiguarda R (1999a)

Apomorphine induces changes in GPi spontaneous outflow in patients with Parkinson's

disease. Mov Disord 14:45-49.

Merello M, Cerquetti D, Cammarota A, Tenca E, Artes C, Antico J, Leiguarda R (2004)

Neuronal globus pallidus activity in patients with generalised dystonia. Mov Disord

19:548-554.

Merello M, Lees AJ, Balej J, Cammarota A, Leiguarda R (1999b) GPi firing rate

modification during beginning-of-dose motor deterioration following acute

administration of apomorphine. Mov Disord 14:481-483.

Mesulam MM, Geula C, Bothwell MA, Hersh LB (1989) Human reticular formation:

cholinergic neurons of the pedunculopontine and laterodorsal tegmental nuclei and some

cytochemical comparisons to forebrain cholinergic neurons. J Comp Neurol 283:611-633.

Middleton FA, Strick PL (1994) Anatomical evidence for cerebellar and basal ganglia

involvement in higher cognitive function. Science 266:458-461.

Middleton FA, Strick PL (1996) The temporal lobe is a target of output from the basal

ganglia. Proc Natl Acad Sci U S A 93:8683-8687.

Middleton FA, Strick PL (2002) Basal-ganglia 'projections' to the prefrontal cortex of the

primate. Cereb Cortex 12:926-935.

Miller WC, DeLong MR (1987) Altered tonic activity of neurons in the globus pallidus

and subthalamic nucleus in the primate MPTP model of Parkinsonism. In: Basal Ganglia

II (Carpenter MB, Jayaraman A, eds), pp 415-427. New York: Plenum Press.

Mink JW (1996) The basal ganglia: focused selection and inhibition of competing motor

programs. Prog Neurobiol 50:381-425.

Mink JW (2003) The Basal Ganglia and involuntary movements: impaired inhibition of

competing motor patterns. Arch Neurol 60:1365-1368.

Mink JW, Thach WT (1987) Preferential relation of pallidal neurons to ballistic

movements. Brain Res 417:393-398.

Mink JW, Thach WT (1991a) Basal ganglia motor control. II. Late pallidal timing

relative to movement onset and inconsistent pallidal coding of movement parameters. J

Neurophysiol 65:301-329.

Page 245: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

233

Mink JW, Thach WT (1991b) Basal ganglia motor control. III. Pallidal ablation: normal

reaction time, muscle cocontraction, and slow movement. J Neurophysiol 65:330-351.

Mintz I, Hammond C, Feger J (1986) Excitatory effect of iontophoretically applied

dopamine on identified neurons of the rat subthalamic nucleus. Brain Res 375:172-175.

Mitchell IJ, Clarke CE, Boyce S, Robertson RG, Peggs D, Sambrook MA, Crossman AR

(1989) Neural mechanisms underlying parkinsonian symptoms based upon regional

uptake of 2-deoxyglucose in monkeys exposed to 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine. Neuroscience 32:213-226.

Mitchell SJ, Richardson RT, Baker FH, DeLong MR (1987a) The primate globus pallidus:

neuronal activity related to direction of movement. Exp Brain Res 68:491-505.

Mitchell SJ, Richardson RT, Baker FH, DeLong MR (1987b) The primate nucleus basalis

of Meynert: neuronal activity related to a visuomotor tracking task. Exp Brain Res

68:506-515.

Molnar GF, Pilliar A, Lozano AM, Dostrovsky JO (2005) Differences in neuronal firing

rates in pallidal and cerebellar receiving areas of thalamus in patients with Parkinson's

disease, essential tremor, and pain. J Neurophysiol 93:3094-3101.

Monakow KH, Akert K, Kunzle H (1978) Projections of the precentral motor cortex and

other cortical areas of the frontal lobe to the subthalamic nucleus in the monkey. Exp

Brain Res 33:395-403.

Montgomery EBJr (2006) Effects of GPi stimulation on human thalamic neuronal

activity. Clin Neurophysiol 117:2691-2702.

Mora F, Mogenson GJ, Rolls ET (1977) Activity of neurons in the region of the

substantia nigra during feeding in the monkey. Brain Res 133:267-276.

Moran A, Bergman H, Israel Z, Bar-Gad I (2008) Subthalamic nucleus functional

organization revealed by parkinsonian neuronal oscillations and synchrony. Brain

131:3395-3409.

Moriizumi T, Nakamura Y, Kitao Y, Kudo M (1987) Ultrastructural analyses of afferent

terminals in the subthalamic nucleus of the cat with a combined degeneration and

horseradish peroxidase tracing method. J Comp Neurol 265:159-174.

Moro E, Esselink RJ, Benabid AL, Pollak P (2002) Response to levodopa in parkinsonian

patients with bilateral subthalamic nucleus stimulation. Brain 125:2408-2417.

Moro E, Scerrati M, Romito LM, Roselli R, Tonali P, Albanese A (1999) Chronic

subthalamic nucleus stimulation reduces medication requirements in Parkinson's disease.

Neurology 53:85-90.

Mouroux M, Feger J (1993) Evidence that the parafascicular projection to the

Page 246: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

234

subthalamic nucleus is glutamatergic. Neuroreport 4:613-615.

Munro-Davies LE, Winter J, Aziz TZ, Stein JF (1999) The role of the pedunculopontine

region in basal-ganglia mechanisms of akinesia. Exp Brain Res 129:511-517.

Murthy VN, Fetz EE (1996) Synchronization of neurons during local field potential

oscillations in sensorimotor cortex of awake monkeys. J Neurophysiol 76:3968-3982.

Mushiake H, Strick PL (1995) Pallidal neuron activity during sequential arm movements.

J Neurophysiol 74:2754-2758.

Nakanishi H, Kita H, Kitai ST (1987) Intracellular study of rat substantia nigra pars

reticulata neurons in an in vitro slice preparation: electrical membrane properties and

response characteristics to subthalamic stimulation. Brain Res 437:45-55.

Nambu A (2007) Globus pallidus internal segment. Prog Brain Res 160:135-150.

Nambu A, Takada M, Inase M, Tokuno H (1996) Dual somatotopical representations in

the primate subthalamic nucleus: evidence for ordered but reversed body-map

transformations from the primary motor cortex and the supplementary motor area. J

Neurosci 16:2671-2683.

Nambu A, Tokuno H, Hamada I, Kita H, Imanishi M, Akazawa T, Ikeuchi Y, Hasegawa

N (2000) Excitatory cortical inputs to pallidal neurons via the subthalamic nucleus in the

monkey. J Neurophysiol 84:289-300.

Nambu A, Tokuno H, Inase M, Takada M (1997) Corticosubthalamic input zones from

forelimb representations of the dorsal and ventral divisions of the premotor cortex in the

macaque monkey: comparison with the input zones from the primary motor cortex and

the supplementary motor area. Neurosci Lett 239:13-16.

Nambu A, Tokuno H, Takada M (2002) Functional significance of the cortico-

subthalamo-pallidal 'hyperdirect' pathway. Neurosci Res 43:111-117.

Nambu A, Yoshida S, Jinnai K (1990) Discharge patterns of pallidal neurons with input

from various cortical areas during movement in the monkey. Brain Res 519:183-191.

Nandhagopal R, McKeown MJ, Stoessl AJ (2008) Functional imaging in Parkinson

disease. Neurology 70:1478-1488.

Nandi D, Aziz TZ, Giladi N, Winter J, Stein JF (2002a) Reversal of akinesia in

experimental parkinsonism by GABA antagonist microinjections in the pedunculopontine

nucleus. Brain 125:2418-2430.

Nandi D, Liu X, Winter JL, Aziz TZ, Stein JF (2002b) Deep brain stimulation of the

pedunculopontine region in the normal non-human primate. J Clin Neurosci 9:170-174.

Narabayashi H, Miyashita N, Hattori Y, Saito K, Endo K (1997) Posteroventral

Page 247: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

235

pallidotomy: its effect on motor symptoms and scores of MMPI test in patients with

Parkinson's disease. Parkinsonism Relat Disord 3:7-20.

Ni Z, Bouali-Benazzouz R, Gao D, Benabid AL, Benazzouz A (2000a) Changes in the

firing pattern of globus pallidus neurons after the degeneration of nigrostriatal pathway

are mediated by the subthalamic nucleus in the rat. Eur J Neurosci 12:4338-4344.

Ni Z, Gao D, Bouali-Benazzouz R, Benabid AL, Benazzouz A (2001) Effect of

microiontophoretic application of dopamine on subthalamic nucleus neuronal activity in

normal rats and in rats with unilateral lesion of the nigrostriatal pathway. Eur J Neurosci

14:373-381.

Ni ZG, Gao DM, Benabid AL, Benazzouz A (2000b) Unilateral lesion of the nigrostriatal

pathway induces a transient decrease of firing rate with no change in the firing pattern of

neurons of the parafascicular nucleus in the rat. Neuroscience 101:993-999.

Nicholls DG, Budd SL (1998) Neuronal excitotoxicity: the role of mitochondria.

Biofactors 8:287-299.

Nini A, Feingold A, Slovin H, Bergman H (1995) Neurons in the globus pallidus do not

show correlated activity in the normal monkey, but phase-locked oscillations appear in

the MPTP model of parkinsonism. J Neurophysiol 74:1800-1805.

Nishino H, Ono T, Sasaki K, Fukuda M, Muramoto K (1984) Caudate unit activity during

operant feeding behavior in monkeys and modulation by cooling prefrontal cortex. Behav

Brain Res 11:21-33.

Noda T, Oka H (1986) Distribution and morphology of tegmental neurons receiving

nigral inhibitory inputs in the cat: an intracellular HRP study. J Comp Neurol 244:254-

266.

Noth J, Podoll K, Friedemann HH (1985) Long-loop reflexes in small hand muscles

studied in normal subjects and in patients with Huntington's disease. Brain 108 ( Pt 1):65-

80.

Obeso JA, Guridi J, Obeso JA, DeLong M (1997) Surgery for Parkinson's disease. J

Neurol Neurosurg Psychiatry 62:2-8.

Obeso JA, Olanow CW, Nutt JG (2000a) Levodopa motor complications in Parkinson's

disease. Trends Neurosci 23:S2-S7.

Obeso JA, Rodriguez-Oroz MC, Rodriguez M, DeLong MR, Olanow CW (2000b)

Pathophysiology of levodopa-induced dyskinesias in Parkinson's disease: problems with

the current model. Ann Neurol 47:S22-S32.

Oertel WH, Mugnaini E (1984) Immunocytochemical studies of GABAergic neurons in

rat basal ganglia and their relations to other neuronal systems. Neurosci Lett 47:233-238.

Page 248: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

236

Oguztoreli MN, Stein RB (1979) Interactions between centrally and peripherally

generated neuromuscular oscillations. J Math Biol 7:1-30.

Oguztoreli MN, Stein RB (1976) The effects of multiple reflex pathways on the

oscillations in neuro-muscular systems. J Math Biol 3:87-101.

Ohara S, Ikeda A, Kunieda T, Yazawa S, Baba K, Nagamine T, Taki W, Hashimoto N,

Mihara T, Shibasaki H (2000) Movement-related change of electrocorticographic activity

in human supplementary motor area proper. Brain 123 ( Pt 6):1203-1215.

Olmstead MC, Franklin KB (1994) Lesions of the pedunculopontine tegmental nucleus

block drug-induced reinforcement but not amphetamine-induced locomotion. Brain Res

638:29-35.

Olszewsky J, Baxter D (1982) Cytoarchitecture of the Human Brain Stem. Basale: Karger.

Onn SP, Grace AA (1999) Alterations in electrophysiological activity and dye coupling

of striatal spiny and aspiny neurons in dopamine-denervated rat striatum recorded in vivo.

Synapse 33:1-15.

Orieux G, Francois C, Feger J, Yelnik J, Vila M, Ruberg M, Agid Y, Hirsch EC (2000)

Metabolic activity of excitatory parafascicular and pedunculopontine inputs to the

subthalamic nucleus in a rat model of Parkinson's disease. Neuroscience 97:79-88.

Overton PG, Greenfield SA (1995) Determinants of neuronal firing pattern in the guinea-

pig subthalamic nucleus: an in vivo and in vitro comparison. J Neural Transm Park Dis

Dement Sect 10:41-54.

Pahapill PA, Lozano AM (2000) The pedunculopontine nucleus and Parkinson's disease.

Brain 123 ( Pt 9):1767-1783.

Pahwa R, Wilkinson S, Smith D, Lyons K, Miyawaki E, Koller WC (1997) High-

frequency stimulation of the globus pallidus for the treatment of Parkinson's disease.

Neurology 49:249-253.

Paladini CA, Celada P, Tepper JM (1999) Striatal, pallidal, and pars reticulata evoked

inhibition of nigrostriatal dopaminergic neurons is mediated by GABA(A) receptors in

vivo. Neuroscience 89:799-812.

Paladini CA, Tepper JM (1999) GABA(A) and GABA(B) antagonists differentially

affect the firing pattern of substantia nigra dopaminergic neurons in vivo. Synapse

32:165-176.

Palmer SS, Hutton JT (1995) Postural finger tremor exhibited by Parkinson patients and

age-matched subjects. Mov Disord 10:658-663.

Papa SM, Desimone R, Fiorani M, Oldfield EH (1999) Internal globus pallidus discharge

is nearly suppressed during levodopa-induced dyskinesias. Ann Neurol 46:732-738.

Page 249: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

237

Parent A (1990) Extrinsic connections of the basal ganglia. Trends Neurosci 13:254-258.

Parent A, Charara A, Pinault D (1995) Single striatofugal axons arborizing in both

pallidal segments and in the substantia nigra in primates. Brain Res 698:280-284.

Parent A, Cossette M (2001) Extrastriatal dopamine and Parkinson's disease. Adv Neurol

86:45-54.

Parent A, De Bellefeuille L (1983) The pallidointralaminar and pallidonigral projections

in primate as studied by retrograde double-labeling method. Brain Res 278:11-27.

Parent A, Gravel S, Boucher R (1981) The origin of forebrain afferents to the habenula in

rat, cat and monkey. Brain Res Bull 6:23-38.

Parent A, Hazrati LN (1993) Anatomical aspects of information processing in primate

basal ganglia. Trends Neurosci 16:111-116.

Parent A, Hazrati LN (1995a) Functional anatomy of the basal ganglia. II. The place of

subthalamic nucleus and external pallidum in basal ganglia circuitry. Brain Res Brain Res

Rev 20:128-154.

Parent A, Hazrati LN (1995b) Functional anatomy of the basal ganglia. I. The cortico-

basal ganglia-thalamo-cortical loop. Brain Res Brain Res Rev 20:91-127.

Parent A, Sato F, Wu Y, Gauthier J, Levesque M, Parent M (2000) Organization of the

basal ganglia: the importance of axonal collateralization. Trends Neurosci 23:S20-S27.

Parent A, Smith Y (1987) Organization of efferent projections of the subthalamic nucleus

in the squirrel monkey as revealed by retrograde labeling methods. Brain Res 436:296-

310.

Parent M, Levesque M, Parent A (2001) Two types of projection neurons in the internal

pallidum of primates: single-axon tracing and three-dimensional reconstruction. J Comp

Neurol 439:162-175.

Parr-Brownlie LC, Hyland BI (2005) Bradykinesia induced by dopamine D2 receptor

blockade is associated with reduced motor cortex activity in the rat. J Neurosci 25:5700-

5709.

Pascual-Leone A, Valls-Sole J, Toro C, Wassermann EM, Hallett M (1994) Resetting of

essential tremor and postural tremor in Parkinson's disease with transcranial magnetic

stimulation. Muscle Nerve 17:800-807.

Pavese N, Evans AH, Tai YF, Hotton G, Brooks DJ, Lees AJ, Piccini P (2006) Clinical

correlates of levodopa-induced dopamine release in Parkinson disease: a PET study.

Neurology 67:1612-1617.

Paxinos G, Huang X.F. (1995) Atlas of the Human Brainstem. San Diego: Academic

Page 250: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

238

Press.

Paz JT, Deniau JM, Charpier S (2005) Rhythmic bursting in the cortico-subthalamo-

pallidal network during spontaneous genetically determined spike and wave discharges. J

Neurosci 25:2092-2101.

Perlmutter JS, Mink JW (2006) Deep Brain Stimulation. Annu Rev Neurosci.

Pert CB, Kuhar MJ, Snyder SH (1976) Opiate receptor: autoradiographic localization in

rat brain. Proc Natl Acad Sci U S A 73:3729-3733.

Pessiglione M, Guehl D, Rolland AS, Francois C, Hirsch EC, Feger J, Tremblay L (2005)

Thalamic neuronal activity in dopamine-depleted primates: evidence for a loss of

functional segregation within basal ganglia circuits. J Neurosci 25:1523-1531.

Petzinger GM, Langston JW (1998) The MPTP-lesioned non-human primate: a model for

Parkinson's disease. In: Advances in Neurodegenerative Disease (Marwah J, Teitelbaum

H, eds), pp 113-148. Scottsdale, AZ: Prominent Press.

Pfurtscheller G, Neuper C (1992) Simultaneous EEG 10 Hz desynchronization and 40 Hz

synchronization during finger movements. Neuroreport 3:1057-1060.

Pfurtscheller G, Pichler-Zalaudek K, Ortmayr B, Diez J, Reisecker F (1998)

Postmovement beta synchronization in patients with Parkinson's disease. J Clin

Neurophysiol 15:243-250.

Picconi B, Centonze D, Hakansson K, Bernardi G, Greengard P, Fisone G, Cenci MA,

Calabresi P (2003) Loss of bidirectional striatal synaptic plasticity in L-DOPA-induced

dyskinesia. Nat Neurosci 6:501-506.

Pillon B, Dubois B, Cusimano G, Bonnet AM, Lhermitte F, Agid Y (1989) Does

cognitive impairment in Parkinson's disease result from non-dopaminergic lesions? J

Neurol Neurosurg Psychiatry 52:201-206.

Pisani A, Bonsi P, Picconi B, Tolu M, Giacomini P, Scarnati E (2001) Role of tonically-

active neurons in the control of striatal function: cellular mechanisms and behavioral

correlates. Prog Neuropsychopharmacol Biol Psychiatry 25:211-230.

Pisani A, Centonze D, Bernardi G, Calabresi P (2005) Striatal synaptic plasticity:

implications for motor learning and Parkinson's disease. Mov Disord 20:395-402.

Plaha P, Ben-Shlomo Y, Patel NK, Gill SS (2006) Stimulation of the caudal zona incerta

is superior to stimulation of the subthalamic nucleus in improving contralateral

parkinsonism. Brain 129:1732-1747.

Plaha P, Gill SS (2005) Bilateral deep brain stimulation of the pedunculopontine nucleus

for Parkinson's disease. Neuroreport 16:1883-1887.

Page 251: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

239

Plaha P, Khan S, Gill SS (2008) Bilateral stimulation of the caudal zona incerta nucleus

for tremor control. J Neurol Neurosurg Psychiatry 79:504-513.

Playford ED, Jenkins IH, Passingham RE, Nutt J, Frackowiak RS, Brooks DJ (1992)

Impaired mesial frontal and putamen activation in Parkinson's disease: a positron

emission tomography study. Ann Neurol 32:151-161.

Plenz D, Kital ST (1999) A basal ganglia pacemaker formed by the subthalamic nucleus

and external globus pallidus. Nature 400:677-682.

Poewe W (2008) Non-motor symptoms in Parkinson's disease. Eur J Neurol 15 Suppl

1:14-20.

Poewe WH, Wenning GK (1998) The natural history of Parkinson's disease. Ann Neurol

44:S1-S9.

Pogosyan A, Kuhn AA, Trottenberg T, Schneider GH, Kupsch A, Brown P (2006)

Elevations in local gamma activity are accompanied by changes in the firing rate and

information coding capacity of neurons in the region of the subthalamic nucleus in

Parkinson's disease. Exp Neurol 202:271-279.

Pollock LJ, Davis L (1930) Muslce tone in parkinsonian states. Arch Neurol Psychiat

23:303-309.

Prescott IA, Dostrovsky JO, Moro E, Hodaie M, Lozano AM, Hutchison WD (2009)

Levodopa enhances synaptic plasticity in the substantia nigra pars reticulata of

Parkinson's disease patients. Brain 132:309-318.

Priori A, Foffani G, Pesenti A, Bianchi A, Chiesa V, Baselli G, Caputo E, Tamma F,

Rampini P, Egidi M, Locatelli M, Barbieri S, Scarlato G (2002) Movement-related

modulation of neural activity in human basal ganglia and its L-DOPA dependency:

recordings from deep brain stimulation electrodes in patients with Parkinson's disease.

Neurol Sci 23 Suppl 2:S101-S102.

Priori A, Foffani G, Pesenti A, Tamma F, Bianchi AM, Pellegrini M, Locatelli M, Moxon

KA, Villani RM (2004) Rhythm-specific pharmacological modulation of subthalamic

activity in Parkinson's disease. Exp Neurol 189:369-379.

Priyadarshi A, Khuder SA, Schaub EA, Priyadarshi SS (2001) Environmental risk factors

and Parkinson's disease: a metaanalysis. Environ Res 86:122-127.

Przedborski S, Jackson-Lewis V (1998) Mechanisms of MPTP toxicity. Mov Disord 13

Suppl 1:35-38.

Putnam TJ (1940) The operative treatment of diseases characterized by involuntary

movements (tremor, athetosis). Assoc Res Nerv Ment Dis 21:666-696.

Quinn N, Bhatia K (1998) Functional neurosurgery for Parkinson's disease. Has come a

Page 252: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

240

long way, though much remains experimental. BMJ 316:1259-1260.

Rack PM (1987) Stretch reflexes and parkinsonian tremor. Acta Neurochir Suppl (Wien )

39:51-53.

Rack PM, Ross HF (1986) The role of reflexes in the resting tremor of Parkinson's

disease. Brain 109 ( Pt 1):115-141.

Raethjen J, Lindemann M, Schmaljohann H, Wenzelburger R, Pfister G, Deuschl G

(2000) Multiple oscillators are causing parkinsonian and essential tremor. Mov Disord

15:84-94.

Raeva S, Vainberg N, Dubinin V (1999) Analysis of spontaneous activity patterns of

human thalamic ventrolateral neurons and their modifications due to functional brain

changes. Neuroscience 88:365-376.

Rafols JA, Fox CA (1976) The neurons in the primate subthalamic nucleus: a Golgi and

electron microscopic study. J Comp Neurol 168:75-111.

Ramaker C, Marinus J, Stiggelbout AM, Van Hilten BJ (2002) Systematic evaluation of

rating scales for impairment and disability in Parkinson's disease. Mov Disord 17:867-

876.

Rascol O, Clanet M, Montastruc JL, Simonetta M, Soulier-Esteve MJ, Doyon B, Rascol

A (1989) Abnormal ocular movements in Parkinson's disease. Evidence for involvement

of dopaminergic systems. Brain 112 ( Pt 5):1193-1214.

Ray NJ, Jenkinson N, Wang S, Holland P, Brittain JS, Joint C, Stein JF, Aziz T (2008)

Local field potential beta activity in the subthalamic nucleus of patients with Parkinson's

disease is associated with improvements in bradykinesia after dopamine and deep brain

stimulation. Exp Neurol 213:108-113.

Raz A, Feingold A, Zelanskaya V, Vaadia E, Bergman H (1996) Neuronal

synchronization of tonically active neurons in the striatum of normal and parkinsonian

primates. J Neurophysiol 76:2083-2088.

Raz A, Frechter-Mazar V, Feingold A, Abeles M, Vaadia E, Bergman H (2001) Activity

of pallidal and striatal tonically active neurons is correlated in mptp-treated monkeys but

not in normal monkeys. J Neurosci 21:RC128.

Raz A, Vaadia E, Bergman H (2000) Firing patterns and correlations of spontaneous

discharge of pallidal neurons in the normal and the tremulous 1-methyl-4-phenyl-1,2,3,6-

tetrahydropyridine vervet model of parkinsonism. J Neurosci 20:8559-8571.

Reese NB, Garcia-Rill E, Skinner RD (1995a) Auditory input to the pedunculopontine

nucleus: II. Unit responses. Brain Res Bull 37:265-273.

Reese NB, Garcia-Rill E, Skinner RD (1995b) The pedunculopontine nucleus--auditory

Page 253: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

241

input, arousal and pathophysiology. Prog Neurobiol 47:105-133.

Rinvik E, Grofova I (1970) Observations on the fine structure of the substantia nigra in

the cat. Exp Brain Res 11:229-248.

Rivlin-Etzion M, Marmor O, Heimer G, Raz A, Nini A, Bergman H (2006a) Basal

ganglia oscillations and pathophysiology of movement disorders. Curr Opin Neurobiol

16:629-637.

Rivlin-Etzion M, Marmor O, Saban G, Rosin B, Haber SN, Vaadia E, Prut Y, Bergman H

(2008) Low-pass filter properties of basal ganglia cortical muscle loops in the normal and

MPTP primate model of parkinsonism. J Neurosci 28:633-649.

Rivlin-Etzion M, Ritov Y, Heimer G, Bergman H, Bar-Gad I (2006b) Local shuffling of

spike trains boosts the accuracy of spike train spectral analysis. J Neurophysiol 95:3245-

3256.

Rodriguez MC, Guridi OJ, Alvarez L, Mewes K, Macias R, Vitek J, DeLong MR, Obeso

JA (1998) The subthalamic nucleus and tremor in Parkinson's disease. Mov Disord 13

Suppl 3:111-118.

Rodriguez-Oroz MC, Rodriguez M, Guridi J, Mewes K, Chockkman V, Vitek J, DeLong

MR, Obeso JA (2001) The subthalamic nucleus in Parkinson's disease: somatotopic

organization and physiological characteristics. Brain 124:1777-1790.

Rolland AS, Herrero MT, Garcia-Martinez V, Ruberg M, Hirsch EC, Francois C (2007)

Metabolic activity of cerebellar and basal ganglia-thalamic neurons is reduced in

parkinsonism. Brain 130:265-275.

Rosales MG, Flores G, Hernandez S, Martinez-Fong D, Aceves J (1994) Activation of

subthalamic neurons produces NMDA receptor-mediated dendritic dopamine release in

substantia nigra pars reticulata: a microdialysis study in the rat. Brain Res 645:335-337.

Rosenberg JR, Amjad AM, Breeze P, Brillinger DR, Halliday DM (1989) The Fourier

approach to the identification of functional coupling between neuronal spike trains. Prog

Biophys Mol Biol 53:1-31.

Rossetti ZL, Sotgiu A, Sharp DE, Hadjiconstantinou M, Neff NH (1988) 1-Methyl-4-

phenyl-1,2,3,6-tetrahydropyridine (MPTP) and free radicals in vitro. Biochem Pharmacol

37:4573-4574.

Rossi L, Marceglia S, Foffani G, Cogiamanian F, Tamma F, Rampini P, Barbieri S,

Bracchi F, Priori A (2008) Subthalamic local field potential oscillations during ongoing

deep brain stimulation in Parkinson's disease. Brain Res Bull 76:512-521.

Rothwell J (2007) Transcranial magnetic stimulation as a method for investigating the

plasticity of the brain in Parkinson's disease and dystonia. Parkinsonism Relat Disord 13

Suppl 3:S417-S420.

Page 254: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

242

Rothwell JC, Obeso JA, Traub MM, Marsden CD (1983) The behaviour of the long-

latency stretch reflex in patients with Parkinson's disease. J Neurol Neurosurg Psychiatry

46:35-44.

Ruskin DN, Bergstrom DA, Walters JR (2002) Nigrostriatal lesion and dopamine

agonists affect firing patterns of rodent entopeduncular nucleus neurons. J Neurophysiol

88:487-496.

Ryan LJ, Clark KB (1991) The role of the subthalamic nucleus in the response of globus

pallidus neurons to stimulation of the prelimbic and agranular frontal cortices in rats. Exp

Brain Res 86:641-651.

Ryan LJ, Clark KB (1992) Alteration of neuronal responses in the subthalamic nucleus

following globus pallidus and neostriatal lesions in rats. Brain Res Bull 29:319-327.

Rye DB, Lee HJ, Saper CB, Wainer BH (1988) Medullary and spinal efferents of the

pedunculopontine tegmental nucleus and adjacent mesopontine tegmentum in the rat. J

Comp Neurol 269:315-341.

Rye DB, Saper CB, Lee HJ, Wainer BH (1987) Pedunculopontine tegmental nucleus of

the rat: cytoarchitecture, cytochemistry, and some extrapyramidal connections of the

mesopontine tegmentum. J Comp Neurol 259:483-528.

Sabate M, Rodriguez M, Mendez E, Enriquez E, Gonzalez I (1996) Obstructive and

restrictive pulmonary dysfunction increases disability in Parkinson disease. Arch Phys

Med Rehabil 77:29-34.

Sadikot AF, Parent A, Francois C (1992a) Efferent connections of the centromedian and

parafascicular thalamic nuclei in the squirrel monkey: a PHA-L study of subcortical

projections. J Comp Neurol 315:137-159.

Sadikot AF, Parent A, Smith Y, Bolam JP (1992b) Efferent connections of the

centromedian and parafascicular thalamic nuclei in the squirrel monkey: a light and

electron microscopic study of the thalamostriatal projection in relation to striatal

heterogeneity. J Comp Neurol 320:228-242.

Sadikot AF, Rymar VV (2008) The primate centromedian-parafascicular complex:

Anatomical organization with a note on neuromodulation. Brain Res Bull.

Salinas E, Sejnowski TJ (2000) Impact of correlated synaptic input on output firing rate

and variability in simple neuronal models. J Neurosci 20:6193-6209.

Samuel M, Ceballos-Baumann AO, Blin J, Uema T, Boecker H, Passingham RE, Brooks

DJ (1997) Evidence for lateral premotor and parietal overactivity in Parkinson's disease

during sequential and bimanual movements. A PET study. Brain 120 ( Pt 6):963-976.

Sanderson P, Mavoungou R, be-Fessard D (1986) Changes in substantia nigra pars

reticulata activity following lesions of the substantia nigra pars compacta. Neurosci Lett

Page 255: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

243

67:25-30.

Sato F, Lavallee P, Levesque M, Parent A (2000) Single-axon tracing study of neurons of

the external segment of the globus pallidus in primate. J Comp Neurol 417:17-31.

Scarnati E, Proia A, Di LS, Pacitti C (1987) The reciprocal electrophysiological influence

between the nucleus tegmenti pedunculopontinus and the substantia nigra in normal and

decorticated rats. Brain Res 423:116-124.

Schaltenbrand G, Wahren W (1977) Atlas for Stereotaxy of the Human Brain. Stuttgart:

Georg Thieme.

Schapira AH (2005) Present and future drug treatment for Parkinson's disease. J Neurol

Neurosurg Psychiatry 76:1472-1478.

Schneidman E, Berry MJ, Segev R, Bialek W (2006) Weak pairwise correlations imply

strongly correlated network states in a neural population. Nature 440:1007-1012.

Schober A (2004) Classic toxin-induced animal models of Parkinson's disease: 6-OHDA

and MPTP. Cell Tissue Res 318:215-224.

Schrag A, Quinn N (2000) Dyskinesias and motor fluctuations in Parkinson's disease. A

community-based study. Brain 123 ( Pt 11):2297-2305.

Schultz W (1986a) Activity of pars reticulata neurons of monkey substantia nigra in

relation to motor, sensory, and complex events. J Neurophysiol 55:660-677.

Schultz W (1986b) Responses of midbrain dopamine neurons to behavioral trigger

stimuli in the monkey. J Neurophysiol 56:1439-1461.

Schwartzman RJ, Alexander GM (1985) Changes in the local cerebral metabolic rate for

glucose in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) primate model of

Parkinson's disease. Brain Res 358:137-143.

Schwyn RC, Fox CA (1974) The primate substantia nigra: a Golgi and electron

microscopic study. J Hirnforsch 15:95-126.

Sealfon SC, Olanow CW (2000) Dopamine receptors: from structure to behavior. Trends

Neurosci 23:S34-S40.

Semba K, Reiner PB, Fibiger HC (1990) Single cholinergic mesopontine tegmental

neurons project to both the pontine reticular formation and the thalamus in the rat.

Neuroscience 38:643-654.

Shahani BT, Young RR (1976) Physiological and pharmacological aids in the differential

diagnosis of tremor. J Neurol Neurosurg Psychiatry 39:772-783.

Sharott A, Magill PJ, Bolam JP, Brown P (2005a) Directional analysis of coherent

Page 256: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

244

oscillatory field potentials in the cerebral cortex and basal ganglia of the rat. J Physiol

562:951-963.

Sharott A, Magill PJ, Harnack D, Kupsch A, Meissner W, Brown P (2005b) Dopamine

depletion increases the power and coherence of beta-oscillations in the cerebral cortex

and subthalamic nucleus of the awake rat. Eur J Neurosci 21:1413-1422.

Shen KZ, Johnson SW (2000) Presynaptic dopamine D2 and muscarine M3 receptors

inhibit excitatory and inhibitory transmission to rat subthalamic neurones in vitro. J

Physiol 525 Pt 2:331-341.

Shen KZ, Johnson SW (2001) Presynaptic GABA(B) receptors inhibit synaptic inputs to

rat subthalamic neurons. Neuroscience 108:431-436.

Shen KZ, Johnson SW (2005) Dopamine depletion alters responses to glutamate and

GABA in the rat subthalamic nucleus. Neuroreport 16:171-174.

Shen KZ, Johnson SW (2002) Presynaptic modulation of synaptic transmission by opioid

receptor in rat subthalamic nucleus in vitro. J Physiol 541:219-230.

Shen W, Flajolet M, Greengard P, Surmeier DJ (2008) Dichotomous dopaminergic

control of striatal synaptic plasticity. Science 321:848-851.

Shik ML, Severin FV, Orlovsky GN (1969) Control of walking and running by means of

electrical stimulation of the mesencephalon. Electroencephalogr Clin Neurophysiol

26:549.

Shill H, Stacy M (1998) Respiratory function in Parkinson's disease. Clin Neurosci

5:131-135.

Shill H, Stacy M (2002) Respiratory complications of Parkinson's disease. Semin Respir

Crit Care Med 23:261-265.

Shink E, Bevan MD, Bolam JP, Smith Y (1996) The subthalamic nucleus and the

external pallidum: two tightly interconnected structures that control the output of the

basal ganglia in the monkey. Neuroscience 73:335-357.

Shink E, Sidibe M, Smith Y (1997) Efferent connections of the internal globus pallidus in

the squirrel monkey: II. Topography and synaptic organization of pallidal efferents to the

pedunculopontine nucleus. J Comp Neurol 382:348-363.

Shink E, Smith Y (1995) Differential synaptic innervation of neurons in the internal and

external segments of the globus pallidus by the GABA- and glutamate-containing

terminals in the squirrel monkey. J Comp Neurol 358:119-141.

Sidibe M, Smith Y (1996) Differential synaptic innervation of striatofugal neurones

projecting to the internal or external segments of the globus pallidus by thalamic afferents

in the squirrel monkey. J Comp Neurol 365:445-465.

Page 257: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

245

Siegfried J (1980) Is the neurosurgical treatment of Parkinson's disease still indicated? J

Neural Transm Suppl195-198.

Siegfried J, Lippitz B (1994) Bilateral chronic electrostimulation of ventroposterolateral

pallidum: a new therapeutic approach for alleviating all parkinsonian symptoms.

Neurosurgery 35:1126-1129.

Silberstein P, Kuhn AA, Kupsch A, Trottenberg T, Krauss JK, Wohrle JC, Mazzone P,

Insola A, Di L, V, Oliviero A, Aziz T, Brown P (2003) Patterning of globus pallidus local

field potentials differs between Parkinson's disease and dystonia. Brain 126:2597-2608.

Silberstein P, Oliviero A, Di L, V, Insola A, Mazzone P, Brown P (2005a) Oscillatory

pallidal local field potential activity inversely correlates with limb dyskinesias in

Parkinson's disease. Exp Neurol 194:523-529.

Silberstein P, Pogosyan A, Kuhn AA, Hotton G, Tisch S, Kupsch A, Dowsey-Limousin P,

Hariz MI, Brown P (2005b) Cortico-cortical coupling in Parkinson's disease and its

modulation by therapy. Brain 128:1277-1291.

Singleton AB (2005) Altered alpha-synuclein homeostasis causing Parkinson's disease:

the potential roles of dardarin. Trends Neurosci 28:416-421.

Skinner RD, Kinjo N, Henderson V, Garcia-Rill E (1990a) Locomotor projections from

the pedunculopontine nucleus to the spinal cord. Neuroreport 1:183-186.

Skinner RD, Kinjo N, Ishikawa Y, Biedermann JA, Garcia-Rill E (1990b) Locomotor

projections from the pedunculopontine nucleus to the medioventral medulla. Neuroreport

1:207-210.

Smith Y, Bolam JP (1990) The output neurones and the dopaminergic neurones of the

substantia nigra receive a GABA-containing input from the globus pallidus in the rat. J

Comp Neurol 296:47-64.

Smith Y, Bolam JP (1991) Convergence of synaptic inputs from the striatum and the

globus pallidus onto identified nigrocollicular cells in the rat: a double anterograde

labelling study. Neuroscience 44:45-73.

Smith Y, Hazrati LN, Parent A (1990) Efferent projections of the subthalamic nucleus in

the squirrel monkey as studied by the PHA-L anterograde tracing method. J Comp Neurol

294:306-323.

Smith Y, Parent A, Seguela P, Descarries L (1987) Distribution of GABA-

immunoreactive neurons in the basal ganglia of the squirrel monkey (Saimiri sciureus). J

Comp Neurol 259:50-64.

Smith Y, Shink E, Sidibe M (1998) Neuronal circuitry and synaptic connectivity of the

basal ganglia. Neurosurg Clin N Am 9:203-222.

Page 258: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

246

Snyder GL, Keller RW, Jr., Zigmond MJ (1990) Dopamine efflux from striatal slices

after intracerebral 6-hydroxydopamine: evidence for compensatory hyperactivity of

residual terminals. J Pharmacol Exp Ther 253:867-876.

Soares J, Kliem MA, Betarbet R, Greenamyre JT, Yamamoto B, Wichmann T (2004)

Role of external pallidal segment in primate parkinsonism: comparison of the effects of

1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-induced parkinsonism and lesions of the

external pallidal segment. J Neurosci 24:6417-6426.

Sochurkova D, Rektor I (2003) Event-related desynchronization/synchronization in the

putamen. An SEEG case study. Exp Brain Res 149:401-404.

Sofroniew MV, Priestley JV, Consolazione A, Eckenstein F, Cuello AC (1985)

Cholinergic projections from the midbrain and pons to the thalamus in the rat, identified

by combined retrograde tracing and choline acetyltransferase immunohistochemistry.

Brain Res 329:213-223.

Spann BM, Grofova I (1992) Cholinergic and non-cholinergic neurons in the rat

pedunculopontine tegmental nucleus. Anat Embryol (Berl) 186:215-227.

Spann BM, Grofova I (1991) Nigropedunculopontine projection in the rat: an anterograde

tracing study with phaseolus vulgaris-leucoagglutinin (PHA-L). J Comp Neurol 311:375-

388.

Speelman JD, Bosch DA (1998) Resurgence of functional neurosurgery for Parkinson's

disease: a historical perspective. Mov Disord 13:582-588.

Starr PA, Rau GM, Davis V, Marks WJ, Jr., Ostrem JL, Simmons D, Lindsey N, Turner

RS (2005) Spontaneous pallidal neuronal activity in human dystonia: comparison with

Parkinson's disease and normal macaque. J Neurophysiol 93:3165-3176.

Stebbins GT, Goetz CG, Lang AE, Cubo E (1999) Factor analysis of the motor section of

the unified Parkinson's disease rating scale during the off-state. Mov Disord 14:585-589.

Stefani A, Lozano AM, Peppe A, Stanzione P, Galati S, Tropepi D, Pierantozzi M, Brusa

L, Scarnati E, Mazzone P (2007) Bilateral deep brain stimulation of the pedunculopontine

and subthalamic nuclei in severe Parkinson's disease. Brain 130:1596-1607.

Stefani A, Stanzione P, Bassi A, Mazzone P, Vangelista T, Bernardi G (1997) Effects of

increasing doses of apomorphine during stereotaxic neurosurgery in Parkinson's disease:

clinical score and internal globus pallidus activity. Short communication. J Neural

Transm 104:895-904.

Steigerwald F, Potter M, Herzog J, Pinsker M, Kopper F, Mehdorn H, Deuschl G,

Volkmann J (2008) Neuronal activity of the human subthalamic nucleus in the

parkinsonian and nonparkinsonian state. J Neurophysiol 100:2515-2524.

Steiniger B, Kretschmer BD (2004) Effects of ibotenate pedunculopontine tegmental

Page 259: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

247

nucleus lesions on exploratory behaviour in the open field. Behav Brain Res 151:17-23.

Steininger TL, Rye DB, Wainer BH (1992) Afferent projections to the cholinergic

pedunculopontine tegmental nucleus and adjacent midbrain extrapyramidal area in the

albino rat. I. Retrograde tracing studies. J Comp Neurol 321:515-543.

Steriade M, Pare D, Parent A, Smith Y (1988) Projections of cholinergic and non-

cholinergic neurons of the brainstem core to relay and associational thalamic nuclei in the

cat and macaque monkey. Neuroscience 25:47-67.

Sterio D, Beric A, Dogali M, Fazzini E, Alfaro G, Devinsky O (1994)

Neurophysiological properties of pallidal neurons in Parkinson's disease. Ann Neurol

35:586-591.

Stevens CF, Zador AM (1998) Input synchrony and the irregular firing of cortical

neurons. Nat Neurosci 1:210-217.

Strick PL, Dum RP, Picard N (1995) Macro-organization of the circuits connecting the

basal ganglia with the cortical motor areas. In: Models of Information Processing in the

Basal Ganglia (Houk JC, Davis JL, Beiser DG, eds), pp 117-130. Cambridge: MIT.

Sugimoto T, Hattori T, Mizuno N, Itoh K, Sato M (1983) Direct projections from the

centre median-parafascicular complex to the subthalamic nucleus in the cat and rat. J

Comp Neurol 214:209-216.

Surmeier DJ, Song WJ, Yan Z (1996) Coordinated expression of dopamine receptors in

neostriatal medium spiny neurons. J Neurosci 16:6579-6591.

Svenningsson P, Le Moine C (2002) Dopamine D1/5 receptor stimulation induces c-fos

expression in the subthalamic nucleus: possible involvement of local D5 receptors. Eur J

Neurosci 15:133-142.

Svirskis G, Rinzel J (2000) Influence of temporal correlation of synaptic input on the rate

and variability of firing in neurons. Biophys J 79:629-637.

Swerdlow NR, Koob GF (1987) Lesions of the dorsomedial nucleus of the thalamus,

medial prefrontal cortex and pedunculopontine nucleus: effects on locomotor activity

mediated by nucleus accumbens-ventral pallidal circuitry. Brain Res 412:233-243.

Taha JM, Favre J, Baumann TK, Burchiel KJ (1996) Characteristics and somatotopic

organization of kinesthetic cells in the globus pallidus of patients with Parkinson's

disease. J Neurosurg 85:1005-1012.

Takakusaki K, Oohinata-Sugimoto J, Saitoh K, Habaguchi T (2004) Role of basal

ganglia-brainstem systems in the control of postural muscle tone and locomotion. Prog

Brain Res 143:231-237.

Takakusaki K, Shiroyama T, Kitai ST (1997) Two types of cholinergic neurons in the rat

Page 260: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

248

tegmental pedunculopontine nucleus: electrophysiological and morphological

characterization. Neuroscience 79:1089-1109.

Takakusaki K, Shiroyama T, Yamamoto T, Kitai ST (1996) Cholinergic and

noncholinergic tegmental pedunculopontine projection neurons in rats revealed by

intracellular labeling. J Comp Neurol 371:345-361.

Tallon-Baudry C, Bertrand O, Peronnet F, Pernier J (1998) Induced gamma-band activity

during the delay of a visual short-term memory task in humans. J Neurosci 18:4244-4254.

Tanaka H, Kannari K, Maeda T, Tomiyama M, Suda T, Matsunaga M (1999) Role of

serotonergic neurons in L-DOPA-derived extracellular dopamine in the striatum of 6-

OHDA-lesioned rats. Neuroreport 10:631-634.

Tang JK, Moro E, Lozano AM, Lang AE, Hutchison WD, Mahant N, Dostrovsky JO

(2005) Firing rates of pallidal neurons are similar in Huntington's and Parkinson's disease

patients. Exp Brain Res 166:230-236.

Tang JK, Moro E, Mahant N, Hutchison WD, Lang AE, Lozano AM, Dostrovsky JO

(2007) Neuronal firing rates and patterns in the globus pallidus internus of patients with

cervical dystonia differ from those with Parkinson's disease. J Neurophysiol 98:720-729.

Tatton WG, Lee RG (1975) Evidence for abnormal long-loop reflexes in rigid

Parkinsonian patients. Brain Res 100:671-676.

Taylor WA (2000) Change-point analysis: a powerful new tool for detecting changes.

Tepper JM, Abercrombie ED, Bolam JP (2007) Basal ganglia macrocircuits. Prog Brain

Res 160:3-7.

Thannickal TC, Lai YY, Siegel JM (2007) Hypocretin (orexin) cell loss in Parkinson's

disease. Brain 130:1586-1595.

Thanvi BR, Lo TC (2004) Long term motor complications of levodopa: clinical features,

mechanisms, and management strategies. Postgrad Med J 80:452-458.

Theodosopoulos PV, Marks WJ, Jr., Christine C, Starr PA (2003) Locations of

movement-related cells in the human subthalamic nucleus in Parkinson's disease. Mov

Disord 18:791-798.

Thobois S, Dominey P, Decety J, Pollak P, Gregoire MC, Broussolle E (2000)

Overactivation of primary motor cortex is asymmetrical in hemiparkinsonian patients.

Neuroreport 11:785-789.

Thomas RJ (1994) Blinking and the release reflexes: are they clinically useful? J Am

Geriatr Soc 42:609-613.

Tiitinen H, Sinkkonen J, Reinikainen K, Alho K, Lavikainen J, Naatanen R (1993)

Page 261: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

249

Selective attention enhances the auditory 40-Hz transient response in humans. Nature

364:59-60.

Timmer J, Gantert C, Deuschl G, Honerkamp J (1993) Characteristics of hand tremor

time series. Biol Cybern 70:75-80.

Timmer J, Lauk M, Deuschl G (1996) Quantitative analysis of tremor time series.

Electroencephalogr Clin Neurophysiol 101:461-468.

Timmermann L, Wojtecki L, Gross J, Lehrke R, Voges J, Maarouf M, Treuer H, Sturm V,

Schnitzler A (2004) Ten-Hertz stimulation of subthalamic nucleus deteriorates motor

symptoms in Parkinson's disease. Mov Disord 19:1328-1333.

Tinazzi M, Del VC, Fincati E, Ottaviani S, Smania N, Moretto G, Fiaschi A, Martino D,

Defazio G (2006) Pain and motor complications in Parkinson's disease. J Neurol

Neurosurg Psychiatry 77:822-825.

Tofighy A, Abbott A, Centonze D, Cooper AJ, Noor E, Pearce SM, Puntis M, Stanford

IM, Wigmore MA, Lacey MG (2003) Excitation by dopamine of rat subthalamic nucleus

neurones in vitro-a direct action with unconventional pharmacology. Neuroscience

116:157-166.

Toro C, Deuschl G, Thatcher R, Sato S, Kufta C, Hallett M (1994) Event-related

desynchronization and movement-related cortical potentials on the ECoG and EEG.

Electroencephalogr Clin Neurophysiol 93:380-389.

Torrence C, Compo G (1998) A practical guide to wavelet analysis. Bull Am Meteor Soc

79:61-78.

Tremblay L, Filion M (1989) Responses of pallidal neurons to striatal stimulation in

intact waking monkeys. Brain Res 498:1-16.

Tremblay L, Filion M, Bedard PJ (1989) Responses of pallidal neurons to striatal

stimulation in monkeys with MPTP-induced parkinsonism. Brain Res 498:17-33.

Troiano AR, de lF-F, Sossi V, Schulzer M, Mak E, Ruth TJ, Stoessl AJ (2009) PET

demonstrates reduced dopamine transporter expression in PD with dyskinesias.

Neurology 72:1211-1216.

Tronnier VM, Fogel W, Kronenbuerger M, Steinvorth S (1997) Pallidal stimulation: an

alternative to pallidotomy? J Neurosurg 87:700-705.

Trottenberg T, Fogelson N, Kuhn AA, Kivi A, Kupsch A, Schneider GH, Brown P (2006)

Subthalamic gamma activity in patients with Parkinson's disease. Exp Neurol 200:56-65.

Tseng KY, Kasanetz F, Kargieman L, Riquelme LA, Murer MG (2001) Cortical slow

oscillatory activity is reflected in the membrane potential and spike trains of striatal

neurons in rats with chronic nigrostriatal lesions. J Neurosci 21:6430-6439.

Page 262: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

250

Tseng KY, Riquelme LA, Belforte JE, Pazo JH, Murer MG (2000) Substantia nigra pars

reticulata units in 6-hydroxydopamine-lesioned rats: responses to striatal D2 dopamine

receptor stimulation and subthalamic lesions. Eur J Neurosci 12:247-256.

Tumer EC, Brainard MS (2007) Performance variability enables adaptive plasticity of

'crystallized' adult birdsong. Nature 450:1240-1244.

Turner RS, Anderson ME (1997) Pallidal discharge related to the kinematics of reaching

movements in two dimensions. J Neurophysiol 77:1051-1074.

Turner RS, Grafton ST, McIntosh AR, DeLong MR, Hoffman JM (2003) The functional

anatomy of parkinsonian bradykinesia. Neuroimage 19:163-179.

Utter AA, Basso MA (2008) The basal ganglia: An overview of circuits and function.

Neurosci Biobehav Rev.

van der Kooy D, Hattori T (1980) Single subthalamic nucleus neurons project to both the

globus pallidus and substantia nigra in rat. J Comp Neurol 192:751-768.

Vidailhet M, Vercueil L, Houeto JL, Krystkowiak P, Benabid AL, Cornu P, Lagrange C,

Tezenas du MS, Dormont D, Grand S, Blond S, Detante O, Pillon B, Ardouin C, Agid Y,

Destee A, Pollak P (2005) Bilateral deep-brain stimulation of the globus pallidus in

primary generalized dystonia. N Engl J Med 352:459-467.

Vila M, Perier C, Feger J, Yelnik J, Faucheux B, Ruberg M, Raisman-Vozari R, Agid Y,

Hirsch EC (2000) Evolution of changes in neuronal activity in the subthalamic nucleus of

rats with unilateral lesion of the substantia nigra assessed by metabolic and

electrophysiological measurements. Eur J Neurosci 12:337-344.

Vingerhoets FJ, Schulzer M, Calne DB, Snow BJ (1997) Which clinical sign of

Parkinson's disease best reflects the nigrostriatal lesion? Ann Neurol 41:58-64.

Vitek JL (2002) Pathophysiology of dystonia: a neuronal model. Mov Disord 17 Suppl

3:S49-S62.

Vitek JL, Bakay RA (1997) The role of pallidotomy in Parkinson's disease and dystonia.

Curr Opin Neurol 10:332-339.

Vitek JL, Chockkan V, Zhang JY, Kaneoke Y, Evatt M, DeLong MR, Triche S, Mewes

K, Hashimoto T, Bakay RA (1999) Neuronal activity in the basal ganglia in patients with

generalized dystonia and hemiballismus. Ann Neurol 46:22-35.

Vitek JL, Zhang J, Evatt M, Mewes K, DeLong MR, Hashimoto T, Triche S, Bakay RA

(1998) GPi pallidotomy for dystonia: clinical outcome and neuronal activity. Adv Neurol

78:211-219.

Volkmann J, Joliot M, Mogilner A, Ioannides AA, Lado F, Fazzini E, Ribary U, Llinas R

(1996) Central motor loop oscillations in parkinsonian resting tremor revealed by

Page 263: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

251

magnetoencephalography. Neurology 46:1359-1370.

Volkmann J, Sturm V, Weiss P, Kappler J, Voges J, Koulousakis A, Lehrke R, Hefter H,

Freund HJ (1998) Bilateral high-frequency stimulation of the internal globus pallidus in

advanced Parkinson's disease. Ann Neurol 44:953-961.

Vreeling FW, Verhey FR, Houx PJ, Jolles J (1993) Primitive reflexes in Parkinson's

disease. J Neurol Neurosurg Psychiatry 56:1323-1326.

Wakabayashi K, Tanji K, Mori F, Takahashi H (2007) The Lewy body in Parkinson's

disease: molecules implicated in the formation and degradation of alpha-synuclein

aggregates. Neuropathology 27:494-506.

Walters JR, Hu D, Itoga CA, Parr-Brownlie LC, Bergstrom DA (2007) Phase

relationships support a role for coordinated activity in the indirect pathway in organizing

slow oscillations in basal ganglia output after loss of dopamine. Neuroscience 144:762-

776.

Wang SY, Aziz TZ, Stein JF, Liu X (2005) Time-frequency analysis of transient

neuromuscular events: dynamic changes in activity of the subthalamic nucleus and

forearm muscles related to the intermittent resting tremor. J Neurosci Methods 145:151-

158.

Ward LM (2003) Synchronous neural oscillations and cognitive processes. Trends Cogn

Sci 7:553-559.

Watanabe K, Kimura M (1998) Dopamine receptor-mediated mechanisms involved in the

expression of learned activity of primate striatal neurons. J Neurophysiol 79:2568-2580.

Watts RL, Mandir AS (1992) The role of motor cortex in the pathophysiology of

voluntary movement deficits associated with parkinsonism. Neurol Clin 10:451-469.

Weinberger M, Hutchison WD, Lozano AM, Hodaie M, Dostrovsky JO (2009) Increased

gamma oscillatory activity in the subthalamic nucleus during tremor in Parkinson's

disease patients. J Neurophysiol 101:789-802.

Weinberger M, Mahant N, Hutchison WD, Lozano AM, Moro E, Hodaie M, Lang AE,

Dostrovsky JO (2006) Beta oscillatory activity in the subthalamic nucleus and its relation

to dopaminergic response in Parkinson's disease. J Neurophysiol 96:3248-3256.

White NM, Hiroi N (1998) Preferential localization of self-stimulation sites in

striosomes/patches in the rat striatum. Proc Natl Acad Sci U S A 95:6486-6491.

Wichmann T, Bergman H, DeLong MR (1994a) The primate subthalamic nucleus. I.

Functional properties in intact animals. J Neurophysiol 72:494-506.

Wichmann T, Bergman H, DeLong MR (1994b) The primate subthalamic nucleus. III.

Changes in motor behavior and neuronal activity in the internal pallidum induced by

Page 264: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

252

subthalamic inactivation in the MPTP model of parkinsonism. J Neurophysiol 72:521-

530.

Wichmann T, Bergman H, Starr PA, Subramanian T, Watts RL, DeLong MR (1999)

Comparison of MPTP-induced changes in spontaneous neuronal discharge in the internal

pallidal segment and in the substantia nigra pars reticulata in primates. Exp Brain Res

125:397-409.

Wichmann T, DeLong MR (2006a) Basal ganglia discharge abnormalities in Parkinson's

disease. J Neural Transm Suppl21-25.

Wichmann T, DeLong MR (2006b) Deep brain stimulation for neurologic and

neuropsychiatric disorders. Neuron 52:197-204.

Wichmann T, DeLong MR (1999) Oscillations in the basal ganglia. Nature 400:621-622.

Wichmann T, Kliem MA (2004) Neuronal activity in the primate substantia nigra pars

reticulata during the performance of simple and memory-guided elbow movements. J

Neurophysiol 91:815-827.

Wichmann T, Kliem MA, DeLong MR (2001) Antiparkinsonian and behavioral effects of

inactivation of the substantia nigra pars reticulata in hemiparkinsonian primates. Exp

Neurol 167:410-424.

Williams D, Kuhn A, Kupsch A, Tijssen M, van BG, Speelman H, Hotton G, Loukas C,

Brown P (2005) The relationship between oscillatory activity and motor reaction time in

the parkinsonian subthalamic nucleus. Eur J Neurosci 21:249-258.

Williams D, Kuhn A, Kupsch A, Tijssen M, van BG, Speelman H, Hotton G, Yarrow K,

Brown P (2003) Behavioural cues are associated with modulations of synchronous

oscillations in the human subthalamic nucleus. Brain 126:1975-1985.

Williams D, Tijssen M, van BG, Bosch A, Insola A, Di L, V, Mazzone P, Oliviero A,

Quartarone A, Speelman H, Brown P (2002) Dopamine-dependent changes in the

functional connectivity between basal ganglia and cerebral cortex in humans. Brain

125:1558-1569.

Williams DR, Watt HC, Lees AJ (2006) Predictors of falls and fractures in bradykinetic

rigid syndromes: a retrospective study. J Neurol Neurosurg Psychiatry 77:468-473.

Wilson CJ, Young SJ, Groves PM (1977) Statistical properties of neuronal spike trains in

the substantia nigra: cell types and their interactions. Brain Res 136:243-260.

Wingeier B, Tcheng T, Koop MM, Hill BC, Heit G, Bronte-Stewart HM (2006) Intra-

operative STN DBS attenuates the prominent beta rhythm in the STN in Parkinson's

disease. Exp Neurol 197:244-251.

Winn P (2008) Experimental studies of pedunculopontine functions: are they motor,

Page 265: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

253

sensory or integrative? Parkinsonism Relat Disord 14 Suppl 2:S194-S198.

Winn P (2006) How best to consider the structure and function of the pedunculopontine

tegmental nucleus: evidence from animal studies. J Neurol Sci 248:234-250.

Woolf NJ, Butcher LL (1986) Cholinergic systems in the rat brain: III. Projections from

the pontomesencephalic tegmentum to the thalamus, tectum, basal ganglia, and basal

forebrain. Brain Res Bull 16:603-637.

Wu Y, Richard S, Parent A (2000) The organization of the striatal output system: a

single-cell juxtacellular labeling study in the rat. Neurosci Res 38:49-62.

Xia R, Rymer WZ (2004) The role of shortening reaction in mediating rigidity in

Parkinson's disease. Exp Brain Res 156:524-528.

Xu W, Russo GS, Hashimoto T, Zhang J, Vitek JL (2008) Subthalamic nucleus

stimulation modulates thalamic neuronal activity. J Neurosci 28:11916-11924.

Xuereb JH, Perry RH, Candy JM, Perry EK, Marshall E, Bonham JR (1991) Nerve cell

loss in the thalamus in Alzheimer's disease and Parkinson's disease. Brain 114 ( Pt

3):1363-1379.

Yamada H, Matsumoto N, Kimura M (2004) Tonically active neurons in the primate

caudate nucleus and putamen differentially encode instructed motivational outcomes of

action. J Neurosci 24:3500-3510.

Yelnik J (2007) PPN or PPD, what is the target for deep brain stimulation in Parkinson's

disease? Brain 130:e79.

Yelnik J, Francois C, Percheron G, Heyner S (1987) Golgi study of the primate substantia

nigra. I. Quantitative morphology and typology of nigral neurons. J Comp Neurol

265:455-472.

Yelnik J, Percheron G (1979) Subthalamic neurons in primates: a quantitative and

comparative analysis. Neuroscience 4:1717-1743.

Yelnik J, Percheron G, Francois C (1984) A Golgi analysis of the primate globus pallidus.

II. Quantitative morphology and spatial orientation of dendritic arborizations. J Comp

Neurol 227:200-213.

Yianni J, Bain P, Giladi N, Auca M, Gregory R, Joint C, Nandi D, Stein J, Scott R, Aziz

T (2003) Globus pallidus internus deep brain stimulation for dystonic conditions: a

prospective audit. Mov Disord 18:436-442.

Yoshida M, Nagatsuka Y, Muramatsu S, Niijima K (1991) Differential roles of the

caudate nucleus and putamen in motor behavior of the cat as investigated by local

injection of GABA antagonists. Neurosci Res 10:34-51.

Page 266: OSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF · PDF fileOSCILLATORY ACTIVITY IN THE BASAL GANGLIA OF PATIENTS WITH PARKINSON’S DISEASE, doctoral program 2009, Moran Weinberger, Department

254

Yoshida S, Nambu A, Jinnai K (1993) The distribution of the globus pallidus neurons

with input from various cortical areas in the monkeys. Brain Res 611:170-174.

Yuill GM (1976) Letter: Effect of levodopa on Parkinsonian tremor. Br Med J 1:283-284.

Zappia M, Colao R, Montesanti R, Rizzo M, Aguglia U, Gambardella A, Oliveri RL,

Quattrone A (1997) Long-duration response to levodopa influences the

pharmacodynamics of short-duration response in Parkinson's disease. Ann Neurol

42:245-248.

Zhang ZX, Roman GC (1993) Worldwide occurrence of Parkinson's disease: an updated

review. Neuroepidemiology 12:195-208.

Zhu Z, Bartol M, Shen K, Johnson SW (2002a) Excitatory effects of dopamine on

subthalamic nucleus neurons: in vitro study of rats pretreated with 6-hydroxydopamine

and levodopa. Brain Res 945:31-40.

Zhu ZT, Shen KZ, Johnson SW (2002b) Pharmacological identification of inward current

evoked by dopamine in rat subthalamic neurons in vitro. Neuropharmacology 42:772-781.

Zhuang P, Li Y, Hallett M (2004) Neuronal activity in the basal ganglia and thalamus in

patients with dystonia. Clin Neurophysiol 115:2542-2557.

Zhuang P, Li YJ (2003) Characteristics of subthalamic neuronal activities in Parkinson s

disease. Sheng Li Xue Bao 55:435-441.

Ziemssen T, Reichmann H (2007) Non-motor dysfunction in Parkinson's disease.

Parkinsonism Relat Disord 13:323-332.

Zirh TA, Lenz FA, Reich SG, Dougherty PM (1998) Patterns of bursting occurring in

thalamic cells during parkinsonian tremor. Neuroscience 83:107-121.

Zrinzo L, Zrinzo LV, Hariz M (2007a) The pedunculopontine and peripeduncular nuclei:

a tale of two structures. Brain 130:e73.

Zrinzo L, Zrinzo LV, Hariz M (2007b) The peripeduncular nucleus: a novel target for

deep brain stimulation? Neuroreport 18:1301-1302.

Zweig RM, Jankel WR, Hedreen JC, Mayeux R, Price DL (1989) The pedunculopontine

nucleus in Parkinson's disease. Ann Neurol 26:41-46.