in vivo electrochemical measurements in

186
The Pennsylvania State University The Graduate School Department of Chemistry IN VIVO ELECTROCHEMICAL MEASUREMENTS IN DROSOPHILA MELANOGASTER A Dissertation in Chemistry by Monique Adrianne Makos © 2010 Monique Adrianne Makos Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2010

Upload: others

Post on 09-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

The Pennsylvania State University

The Graduate School

Department of Chemistry

IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

DROSOPHILA MELANOGASTER

A Dissertation in

Chemistry

by

Monique Adrianne Makos

© 2010 Monique Adrianne Makos

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

May 2010

Page 2: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

ii

The dissertation of Monique Adrianne Makos was reviewed and approved* by the

following:

Andrew G. Ewing

Professor of Chemistry

J. Lloyd Huck Chair in Natural Science

Dissertation Advisor

Chair of Committee

Mary Elizabeth Williams

Associate Professor of Chemistry

Christine D. Keating

Associate Professor of Chemistry

Richard W. Ordway

Associate Professor of Biology

Michael L. Heien

Assistant Professor of Chemistry

Barbara J. Garrison

Shapiro Professor of Chemistry

Head of the Department of Chemistry

*Signatures are on file in the Graduate School

Page 3: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

iii

Thesis Abstract

Carbon-fiber microelectrodes coupled with electrochemical detection have been

extensively used for the analysis of biogenic amines. In order to determine the functional

role of amines, in vivo studies have primarily used rats and mice as model organisms.

This thesis concerns the development of an electrochemical detection method for in vivo

measurements of dopamine in the nanoliter-sized adult Drosophila melanogaster central

nervous system (CNS). A cylindrical carbon-fiber microelectrode was placed in a fly

brain region containing a dense cluster of dopaminergic neurons while a micropipet

injector was used to exogenously apply dopamine to the area. Changes in dopamine

concentration in the fly were monitored in vivo with background-subtracted fast-scan

cyclic voltammetry (FSCV). Distinct differences were found for the clearance of

exogenously applied dopamine by the dopamine transporter in the brain of a wild-type fly

vs. a mutant fly lacking dopamine transporter function. The measured current response

due to oxidation of dopamine at the electrode surface increased significantly for wild-

type flies following treatment with cocaine which is a known dopamine uptake blocker.

The current remained unchanged for mutant flies under the same conditions. These

results demonstrate the validity of using this novel analytical technique to monitor

dopamine uptake in Drosophila.

The in vivo method described in this thesis has been used to study mechanisms

that underlie drug addiction from a physiological perspective. In addition to being a

valuable tool for the analytical chemistry field, this work is of significant interest to the

neuroscience community. Dopamine neurotransmission is believed to play a critical role

in addiction reinforcing mechanisms of drugs of abuse. Little is known about the in vivo

Page 4: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

iv

nature of drug interactions with invertebrate transporters, mainly because of the lack of

techniques available for quantifying neurochemicals in such small native environments.

Hence, the effects of several psychostimulants on dopamine clearance in the Drosophila

melanogaster CNS have been investigated with in vivo electrochemical detection. FSCV

was used to quantify changes in dopamine concentration in the fly brain when cells were

exposed to cocaine, amphetamine, methamphetamine, or methylphenidate. Clearance of

exogenously applied dopamine was significantly decreased in the wild-type fly following

all drug treatments. In contrast, dopamine uptake remained unchanged when identical

treatments were employed in mutant flies lacking functional dopamine transporters.

Although the understanding of the complex actions of cocaine in the brain has

improved, an effective drug treatment for cocaine addiction has yet to be found. During

the last decade, methylphenidate has been investigated as a potential medication for

cocaine addiction treatment. Methylphenidate binds the dopamine transporter and

increases extracellular dopamine levels in the CNS similar to cocaine but is thought to

elicit fewer addictive and reinforcing effects. Several studies that have investigated the

effects of oral methylphenidate taken by cocaine users have reported mixed results. I

utilized the Drosophila model system to investigate the mechanism behind treating

cocaine addiction with methylphenidate. The results suggested oral consumption of

methylphenidate sufficiently blocks the Drosophila dopamine transporter, and further

inhibition of the transporter by cocaine applied directly to the brain was undetectable.

These data highlight the possibility that methylphenidate could be used as a treatment for

cocaine addiction and demonstrate the great potential of Drosophila as a model system

for future drug abuse research.

Page 5: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

v

Chemical, electrical, and optogenetic methods to stimulate dopamine release in

the adult Drosophila CNS with FSCV detection were investigated. The results suggested

that the noninvasive optogenetic stimulation method is capable of initiating targeted

neurochemical release in the Drosophila CNS. Dopamine release has been shown to

cause pH fluctuations in the rat brain which can interfere with electrochemically

measured signals; therefore, a pH sensor was developed for use in the fly.

The fabrication and characterization of a novel voltammetric pH microelectrode

sensor is described. This sensor has been used to detect pH changes in Drosophila

associated with in vivo neurotransmitter release. Voltammetric pH sensors measure

changes in the redox-potential of a surface-bound, electrochemically active species as a

function of pH. While this approach to measuring pH has been demonstrated with a

variety of quinone-modified electrodes, up until now, none have been developed with

biocompatible materials that exhibit activity on a physiological time scale in a relevant

pH range.

Voltammetric reduction of the commercially available diazonium salt Fast Blue

RR (FBRR) onto the carbon-fiber surface provided a one-step, reagentless procedure for

surface modification of a carbon-fiber microelectrode. This produced a 5-µm diameter

sensor with a pH-sensitive quinone molecule covalently bonded to the carbon surface.

FSCV was used to probe the redox activity of the FBRR molecule as a function of pH.

Calibration of the sensor in solutions ranging from pH 6.5 to 8.0 resulted in a linear pH-

dependent anodic peak potential response. Flow-injection analysis was used to

characterize the modified microelectrode which responded to acidic and basic changes as

low as 0.005 pH units in < 2 s. The long-term stability of the FBRR microelectrode pH

Page 6: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

vi

sensor was tested by continuously applying potential to electrodes in pH 7.5

physiological saline solution for 2.5 h (corresponding to 45,000 voltammetric sweeps).

This is an ample time window for in vivo electrochemical measurements in Drosophila

melanogaster. Furthermore, the pH sensor was successfully used to measure dynamic pH

fluctuations in vivo following dopamine release in the nanoliter-sized CNS of

Drosophila.

The results obtained from the analytical tools developed for in vivo detection of

dopamine and pH changes in the fly suggest the validity of using Drosophila as a model

system to study neurotransmission.

Page 7: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

vii

Table of Contents

List of Figures...................................................................................................................ix

List of Tables...................................................................................................................xii

List of Schemes...............................................................................................................xiii

Abbreviations..................................................................................................................xiv

Acknowledgements.........................................................................................................xvi

Chapter 1: Chemical Measurements in Drosophila.......................................................1

Introduction.............................................................................................................2

Detection Methods for the Analysis of Drosophila Homogenates........................6

Analytical Techniques for Measuring the Physiology of Intact Flies..................16

Scope of the Thesis................................................................................................23

References..............................................................................................................26

Chapter 2: In Vivo Electrochemical Measurements of Exogenously Applied

Dopamine in Drosophila melanogaster...........................................................................31

Introduction............................................................................................................32

Methods..................................................................................................................34

Results and Discussion...........................................................................................37

Conclusions.............................................................................................................51

References...............................................................................................................52

Chapter 3: Using In Vivo Electrochemistry to Study the Physiological Effects of

Cocaine and Other Stimulants on the Drosophila melanogaster Dopamine

Transporter......................................................................................................................55 Introduction...........................................................................................................56

Methods.................................................................................................................58

Results and Discussion..........................................................................................60

Conclusions............................................................................................................77

References..............................................................................................................79

Page 8: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

viii

Chapter 4: Oral Administration of Methylphenidate Blocks the Effect of Cocaine on

Uptake at the Drosophila Dopamine Transporter........................................................83

Introduction...........................................................................................................84

Methods.................................................................................................................87

Results and Discussion..........................................................................................89

Conclusions............................................................................................................97

References..............................................................................................................98

Chapter 5: Methods for Stimulating Dopamine Release in the Drosophila CNS....103

Introduction.........................................................................................................104

Methods...............................................................................................................109

Results and Discussion........................................................................................112

Conclusions..........................................................................................................123

References............................................................................................................124

Chapter 6: Development and Characterization of a Voltammetric Carbon-Fiber

Microelectrode pH Sensor.............................................................................................129

Introduction..........................................................................................................130

Methods................................................................................................................132

Results and Discussion........................................................................................135

Conclusions..........................................................................................................149

References............................................................................................................152

Chapter 7: Future Directions for Quantifying Neurochemicals in Drosophila Using

Electrochemical Detection.............................................................................................155

Investigating Alcohol Addiction with Drosophila..............................................156

Quantifying the Kinetics of Dopamine Uptake in Drosophila............................160

Improving the Detection of Stimulated Dopamine Release in Drosophila.........163

References............................................................................................................166

Appendix.........................................................................................................................169

Page 9: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

ix

List of Figures

Figure 1.1. Drosophila brain regions.................................................................................3

Figure 1.2. MEKC-EC separations of homogenates from Drosophila..............................9

Figure 1.3. Mass spectrometric measurements of the Drosophila proteome...................13

Figure 1.4. Microfluidic device for the analysis of Drosophila embryos........................18

Figure 1.5. Measurements in Drosophila larvae following optogenetic stimulation.......21

Figure 1.6. Investigating dopamine transporter function in adult Drosophila.................24

Figure 2.1. Images of Drosophila taken during microsurgery.........................................39

Figure 2.2. Confocal fluorescence micrographs of intact brains from adult transgenic

TH-GAL4/UAS-GFP flies.................................................................................................41

Figure 2.3. Exogenously applied 1.0 mM dopamine detected in vivo in an adult wild-

type fly...............................................................................................................................43

Figure 2.4. Voltammetric detection of exogenously applied dopamine solutions in the

PAM area of an adult Drosophila brain.............................................................................45

Figure 2.5. Effect of cocaine on dopamine uptake..........................................................47

Figure 2.6. Effect of TTX on dopamine uptake...............................................................50

Figure 3.1. In vivo detection of exogenously applied 1.0 mM dopamine in the adult

Drosophila brain................................................................................................................62

Figure 3.2. Effect of 1.0 mM cocaine treatment on uptake of an exogenously applied 1.0

mM dopamine solution......................................................................................................64

Figure 3.3. Investigating dopamine transporter function.................................................66

Figure 3.4. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before and after 1.0 mM cocaine treatment....................................67

Figure 3.5. Determining the physiological APAP concentration in the Drosophila CNS

from a 1.0 mM APAP bath application.............................................................................69

Figure 3.6. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before and after 10 min of various concentrations of cocaine

treatments...........................................................................................................................71

Page 10: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

x

Figure 3.7. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before and after 1.0 stimulant treatment..........................................74

Figure 4.1. Effect of orally consumed methylphenidate on cocaine inhibition of the

dopamine transporter in the adult Drosophila brain..........................................................91

Figure 4.2. Effect of orally consumed methylphenidate on Drosophila dopamine

transporter function............................................................................................................93

Figure 4.3. Comparison of dopamine concentration in the Drosophila CNS following

drug treatments...................................................................................................................96

Figure 5.1. Cartoon depiction of the effects of blue light exposure on neurons expressing

Channelrhodopsin-2.........................................................................................................108

Figure 5.2. Schematic comparing three methods for stimulating neurotransmitter release

in adult Drosophila..........................................................................................................114

Figure 5.3. Effect of blue light stimulation on flies with genetically altered dopamine

neurons.............................................................................................................................118

Figure 5.4. Voltammograms obtained during blue and red light stimulation of a TH-

GAL4/UAS:ChR2 mutant fly..........................................................................................120

Figure 5.5. Spontaneous release of an electroactive species from a TH-

GAL4/UAS:ChR2 mutant fly..........................................................................................122

Figure 6.1. Cyclic voltammograms of a carbon-fiber microelectrode before and after

FBRR attachment.............................................................................................................137

Figure 6.2. Electrochemical characterization of the FBRR microelectrode pH sensor in

pH 7.5 AHL saline solution.............................................................................................143

Figure 6.3. Cyclic voltammograms of a microelectrode modified with FBRR in AHL

saline solutions of different pH........................................................................................145

Figure 6.4. The anodic peak potential as a function of AHL saline solution pH for

FBRR-modified electrodes..............................................................................................146

Figure 6.5. Plot of anodic peak potential vs. time during flow injection changes of 0.2

pH units in AHL saline....................................................................................................148

Figure 6.6. Physiological pH measurements in an adult Drosophila CNS...................150

Page 11: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xi

Figure 7.1. The fly inebriometer....................................................................................159

Figure 7.2. Modeling dopamine uptake.........................................................................162

Figure 7.3. Voltammetric measurements of dopamine using an applied waveform of 1.0

V vs. a waveform extended to 1.4 V................................................................................165

Page 12: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xii

List of Tables

Table 3.1. Change in [DA]max for four drugs of abuse.....................................................76

Table 5.1. Eliciting dopamine release via chemical stimulation....................................106

Table 6.1. Effect of varying voltammetric deposition parameters for FBRR reduction

onto a carbon-fiber surface..............................................................................................139

Table 7.1. Drosophila mutants that display altered behavioral responses to ethanol....158

Page 13: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xiii

List of Schemes

Scheme 6.1. Electrochemical deposition of FBRR salt onto the carbon-fiber

microelectrode surface.....................................................................................................136

Scheme 6.2. Proposed mechanism for the oxidation-reduction reaction of the surface-

bound quinone derivative of FBRR.................................................................................141

Page 14: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xiv

Abbreviations

ACN: acetonitrile

ADHD: attention deficit hyperactivity disorder

AHL: adult-hemolymph like

ANOVA: analysis of variance

APAP: N-acetyl-p-aminophenol, acetaminophen

CAT: catechol

CE: capillary electrophoresis

ChR2: Channelrhodopsin-2

CNS: central nervous system

[DA]max: peak dopamine concentration

DHBA: dihydroxylbenzylamine

dsRNA: double-stranded RNA

Epa: anodic peak potential

Epc: cathodic peak potential

FBRR: Fast Blue RR

fmn: fumin (Drosophila mutant)

FSCV: fast-scan cyclic voltammetry

GABA: γ-aminobutyric acid

GAL4/UAS: galactosidase-4-upstream activating sequence

GFP: green fluorescent protein

HPLC: high-performance liquid chromatography

IC50: half maximal inhibitory concentration

ISMs: ion-selective microelectrodes

LC: liquid chromatography

LC-IMS-MS: liquid chromatography-ion mobility spectrometry-mass spectrometry

L-DOPA: L-3,4-dihydroxyphenylalanine

LED: light-emitting diode

LOD: limit of detection

MALDI-TOF: matrix-assisted laser desorption ionization time-of-flight

MB: mushroom body

MEKC: micellar electrokinetic chromatography

MEKC-EC: micellar electrokinetic chromatography with electrochemical detection

mRNA: messenger RNA

na5-HT: N-acetyl serotonin

naDA: N-acetyl dopamine

naOA: N-acetyl octopamine

naTA: N-acetyl tyramine

NMDA: N-methyl-D-aspartate

QTOF: quadrupole time-of-flight

PAM: protocerebral anterior medial

PBS: phosphate-buffered saline

PI: propidium iodide

Page 15: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xv

RISC: RNA-induced silencing complex

RNA: ribonucleic acid

RNAi: RNA interference

SDS: sodium dodecyl sulfate

SEM: standard error of the mean

siRNA: small interfering RNA

S/N: signal-to-noise

TEABF4: tetraethylammonium tetrafluoroborate

TES: N-tris(hydroxymethyl)methyl-2-aminoethanesulfonic acid

TH: tyrosine hydroxylase

TTX: tetrodotoxin

UAS: upstream activating sequence

VNC: ventral nerve cord

Page 16: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

xvi

Acknowledgments

Graduate school has been a marathon, and I would like to acknowledge several

people for their contributions to my run. While I have come into contact with many great

scientists during the last five years, without two scientists in particular this thesis work

would not have been accomplished. I would like to thank Andy Ewing for the standards

he set for me. He expected no less than my personal best work, and I am a better scientist

for it. In addition, he has been an example of how to work with people to attain a desired

goal, which is knowledge I will carry with me long after I forget how to dissect a fruit fly.

I would like to thank Michael Heien for his significant contributions to my graduate

school education. Without his technical skills in lab and his assistance with writing draft

after draft (after draft) of papers, much of this thesis simply would not exist. My research

was financially supported by National Institutes of Health Grant 5R01GM078385-02.

On a personal note, I would like to thank my Dad, my Sister, and my Grandma for

their encouragement. My Mom especially has been a role model for my educational

endeavors, as well as for all aspects of my life. If one day I can possess just half her

ability to overcome life’s ups and downs, then I will consider myself successful indeed.

My education at PSU has given me the opportunity to meet two special people.

Donna Omiatek started as the labmate whose chair I was constantly backing up into with

my own, but has since become my best friend. She has picked me up on many a rough

day both inside and outside of the lab. Matthew Pond began as one of my many

classmates, but will continue to be part of my life where ever I go. I appreciate his

interest in my work here at PSU, as well as his support of my future goals.

Lastly, to Drosophilia...may you rest in peace.

Page 17: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 1: Chemical Measurements in Drosophila*

*Reproduced with permission from Makos, M. A., Kuklinski, N. J., Berglund, E. C.,

Heien, M. L., and Ewing, A. G. (2009) Chemical Measurements in Drosophila, TrAC,

Trends Anal. Chem. 28, 1223-1234. © 2009 Elsevier.

1

Page 18: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

Drosophila melanogaster has been extensively used as a model organism in

genetics research and significantly contributed to the molecular, cellular, and

evolutionary understandings of human behavior. Originally pioneered by Thomas H.

Morgan at the beginning of the last century, research utilizing the fruit fly has led to

important insights into the mechanisms of human developmental and physiological

processes. Recently, research has focused on developing analytical methods to obtain

highly sensitive chemical quantification along with spatiotemporal information of

Drosophila (1). The fly matures relatively quickly, developing from an embryo, to larva

(divided into 1st, 2

nd, and 3

rd instar larva stages), to pupa, to a sexually mature adult in a

span of ~12 days. An adult fly brain is approximately 5 nL in volume and comprised of

several distinct structures which control specific tasks (Figure 1.1A) (2). The small

dimensions are a challenge for researchers attempting chemical quantification in the fly

and necessitate the use of techniques capable of handling mass-limited samples.

Although the adult fly has a more simple nervous system when compared to vertebrates,

it is capable of higher-order brain functions including both aversive and appetitive

learning and recalling learned information from prior experiences (3, 4). In addition,

Drosophila larvae can be used as a model for investigating basic neurotransmission and

chemosensory pathways (5). Conservation between the Drosophila and mammalian

proteomes is high with approximately half the protein sequences in the fly having similar

counterparts in the human sequence (6). Many central nervous system (CNS) pathways

are evolutionarily conserved between the two species because of the genetic similarity.

2

Page 19: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.1. Drosophila brain regions. (A) A polygonal model of the Drosophila

melanogaster brain. Major neuropil regions are highlighted in color (brown = mushroom

body; beige = lateral horn; blue = antennal lobe; green = central complex; red = medulla;

orange = lobula; yellow = lobula plate). (B) Tyrosine hydroxylase immunolabeling

showing dopaminergic neuron patterns in multifocal confocal views of adult fly brain.

(Reprinted from (7, 8), with permission from Elsevier and the Society for Neuroscience).

3

Page 20: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Neurochemical basis for observed behaviors. Studies of neurotransmitters in the CNS

are underway in Drosophila to elucidate the roles of neurochemicals in human behavior.

Biogenic amines, namely dopamine, serotonin, and tyramine, are known to be involved in

physiological processes found in both mammalian and Drosophila systems (9-11). For

example, dopamine has been implicated in human and fly behaviors such as reward and

motivation, sleep cycles, alcohol tolerance, and sensitivity to addictive drugs (12-14). In

addition, the neurotransmitter octopamine is thought to control many behaviors in the fly

that norepinephrine regulates in mammals (15). This evidence suggests many of the

neurotransmitter systems that regulate behavior are comparable between mammals and

Drosophila.

Genetic manipulation for chemical analysis and behavioral studies. The Drosophila

proteome was one of the first species with a fully sequenced genome (16). The process

of producing mutants to display a desired behavior via genetic manipulation is a

relatively straightforward task with the fruit fly. The Drosophila genome contains little

genetic redundancy, or multiple genes performing the same biochemical function, which

facilitates identification of individual genes and molecules that influence a particular

behavior (2, 17). Many complex behavioral patterns found in mammalian systems with

regards to learning and memory, courtship, alcohol tolerance, and circadian rhythms have

been studied in the fruit fly through the use of genetic mutants (17-20).

Controlling genetic mutations in Drosophila is possible with the galactosidase-4-

upstream activating sequence (GAL4/UAS) system. The landmark development of the

GAL4/UAS system by Brand and Perrimon in 1993 allows for the rapid generation of

flies containing targeted gene expression (21). Briefly, GAL4 is a gene that encodes for

4

Page 21: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

the yeast transcription activator protein Gal4 and can be expressed in various subsets of

fly tissue. Thousands of GAL4 driver or enhancer lines have been created that direct

transcription to different regions and/or types of cells in the fly (22). For instance, the

TH-GAL4 driver line produces flies with GAL4 only in neurons where tyrosine

hydroxylase (TH) is present (23). GAL4 remains inactive in the fly until it binds an

UAS. Many UAS responder lines have been created to contain the UAS region by a

desired protein like the green fluorescent protein (GFP). For example, the UAS-GFP

responder line can be crossed with the TH-GAL4 driver line to produce flies with GFP

transcription in their TH-containing neurons (24). Because TH is the enzyme involved in

the rate limiting step of dopamine synthesis, TH-GAL4 targets GFP expression in

dopamine neurons. This allows for the visualization of dopaminergic neurons in

Drosophila. Tools for fluorescent immunodetection of specific neuron clusters in the fly

brain are available as well. Figure 1.1B is an example of a fly brain image utilizing TH

immunolabeling.

Drosophila mutants have been successfully used to model several human

neurodegenerative diseases, including Alzheimer’s disease, Parkinson’s disease, and

Huntington’s disease. These diseases are characterized by the late onset of progressive

neurodegeneration and/or formation of abnormal neuronal inclusions or protein

aggregates (25, 26). While genetic mutants have helped in linking particular genes to a

specific disease, little is known about the mechanisms leading up to these pathologies.

The ability to quantify all neuropeptides, amino acids, and neurotransmitters in

Drosophila is a goal researchers are moving towards. Obtaining spatiotemporal

information along with chemical quantification will provide a more analytical view of

5

Page 22: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Drosophila and could lead to a better understanding of the physiological mechanisms that

underlie human behaviors, addictions, and neurodegenerative diseases.

Detection Methods for the Analysis of Drosophila Homogenates

Techniques that are used to separate and quantify mass-limited samples include

capillary electrophoresis (CE), high-performance liquid chromatography (HPLC), and

mass spectrometry. Indeed, these methods are sensitive and selective making them

capable of measuring and identifying multiple compounds in a complex biological

sample. This ability allows the determination of different neurochemicals that are within

the brain which is crucial to understanding changes throughout disease states.

Historically, analytical techniques have used homogenate methods to contend

with the hard, exterior cuticle of the fly. Whole fly heads are easily pulverized using

small tissue grinders; however, significant matrix effects from whole fly head samples

can interfere with quantification. Another approach is to dissect the brains from the head

by hand prior to homogenization. Unwanted signals from the fly head matrix are

reduced, but preparation is more time consuming and requires knowledge of dissection

techniques.

Capillary electrophoresis. CE separates ionic species according to their electrophoretic

mobilities by applying a voltage over a narrow capillary filled with electrolytic solution.

The small injection volumes associated with CE (nanoliters to femtoliters) make it an

excellent method to study volume-limited samples such as Drosophila (27, 28).

Moreover, CE has high resolving power due to its plug-like flow and minimal diffusion.

Neutral molecules can be separated with CE by utilizing a surfactant to carry out micellar

6

Page 23: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

electrokinetic chromatography (MEKC). In MEKC, adding the surfactant sodium

dodecyl sulfate (SDS) to the running buffer at levels above the critical micelle

concentration results in the formation of micelles. The interaction of neutral molecules

with the charged micelles causes retention in the capillary which can lead to the

separation of neutral molecules.

The Ewing laboratory has developed a procedure using MEKC to measure and

quantify biogenic amines, their metabolites, and their precursors in Drosophila. End-

column amperometry is used to selectively detect electroactive species providing a

simple and sensitive detection method without the need for derivatization. Using MEKC

coupled to electrochemical detection (MEKC-EC), different anatomical regions of

Drosophila have been investigated including whole body homogenates (29), whole head

homogenates (29), single head homogenates (27), and more recently dissected brains.

Pioneering work by Ream et al. attempted to identify neurotransmitters in

Drosophila using MEKC which resulted in the identification of four species: dopamine,

tyramine, serotonin, and the dopamine precursor L-3,4-dihydroxyphenylalanine (L-

DOPA) (29). Migration times from standards obtained both before and after the fly

sample were used for peak identification as well as normalization to the migration time of

an internal standard, dihydroxylbenzylamine (DHBA), to compensate for possible peak

drifting. A collection of either heads or bodies (thoraces and abdomens) were

homogenized and separated using TES (N-tris(hydroxymethyl)methyl-2-

aminoethanesulfonic acid)/SDS buffer. A higher abundance of dopamine in samples

from the body was observed when compared to the head only. This may be a

consequence of dopamine being a main component in sclerotization (hardening) of the

7

Page 24: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

cuticle of the fly body. The levels of L-DOPA remained unaltered between the two fly

preparations. Furthermore, the levels of serotonin and tyramine were found to be higher

in the head with tyramine levels close to the limit of detection in the body.

The type of buffer used affects the resolution of separations of the biogenic

amines. By using MEKC and a 25 mM borate/SDS buffer instead of TES/SDS,

additional monoamines and metabolites were separated and identified (30). Borate (at

basic pH) forms a complex with analytes possessing vicinal hydroxyl groups, imparting

negative charge to the complex, which made it possible to separate a standard of 14

neurochemicals. Here, catechol (CAT) was used for the internal standard. In addition to

previously identified molecules, dopamine, tyramine, L-DOPA, and octopamine were

identified in homogenized samples from Drosophila heads. The N-acetylated

metabolites N-acetyl dopamine (naDA), N-acetyl octopamine (naOA), and N-acetyl

serotonin (na5-HT) were identified as well. The excellent separation ability of the

borate/SDS buffer with MEKC-EC was demonstrated by comparing electropherograms

from wild-type Drosophila to a mutant form, inactive, which expresses lower levels of

octopamine and tyramine. As expected, the amounts of naOA, tyramine, and octopamine

were reduced in the mutant vs. the wild-type fly, with tyramine being present at levels

below the limit of detection in the mutant (Figure 1.2A).

The small sample volumes that can be analyzed using CE allow the study of the

variability within a population of flies that arises from individual fly-to-fly differences.

This is accomplished by analyzing one fly head at a time (27). Following

homogenization of a single fly head in 250 nL of perchloric acid, three individual fly

8

Page 25: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.2. MEKC-EC separations of homogenates from Drosophila. (A) Enlarged

portion of an electropherogram (left) includes peaks naOA (2), naDA (3), na5-HT (5),

octopamine (6), dopamine (8). Electropherogram (right) compares wild-type (WT, black

trace) and mutant (inactive, blue trace) head homogenates emphasizing the internal

standard CAT (11) and tyramine (9). Separation was run with borate buffer. There is not

a detectable level of tyramine in the mutant form. (B) Electropherogram of a single head

with TES running buffer highlighting L-DOPA (1), naOA (2), naDA (3), naTA (4), na5-

HT (5), octopamine (6), DHBA (7), dopamine (8), serotonin (10). Tyramine (9) is not

visible on this scale. (C) Electropherogram of hand dissected brain where naTA (4) and

CAT (11) are visible. The working electrode was held at +750 mV vs. a Ag/AgCl

reference electrode for all separations. (Reprinted from (27, 30), with permission from

the American Chemical Society).

9

Page 26: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

heads were analyzed and compared. This procedure resulted in reproducible

identification of nine neurochemicals (Figure 1.2B), including N-acetyl tyramine (naTA).

Despite the resolving power of MEKC, unidentified electroactive species coelute

with some neurochemicals. To reduce this problem, dissected Drosophila brains can be

separated using borate/SDS buffer (Figure 1.2C). By removal of fly components thought

to contain electroactive molecules (e.g., the cuticle, antennae, and eyes) the

electropherogram becomes easier to interpret. This is observed when comparing Figure

1.2A, B with Figure 1.2C, where the number of large, overloading, unidentified peaks is

reduced in the single brain electropherogram. In addition, the dopamine from the cuticles

is not measured which allows the amount of dopamine in the CNS of the fly to be

determined.

High-performance liquid chromatography. HPLC has been used to quantify the

amount of biogenic amines, their metabolites, and their precursors in the Drosophila

CNS. The aim of these studies was to determine the function of molecules and their

localization within the fly head. HPLC is an improved form of column chromatography

where solvent is pushed though the column under high pressures (up to 40 MPa). The

high pressure allows for faster separation times and smaller column particles, yielding

improved resolution. Typically, a C-18 column with an acidic mobile phase and

electrochemical detector has been used to separate and detect compounds (31-35).

Early reports using HPLC demonstrated the separation and quantification of

dopamine, L-DOPA, and α-methyldopa in 1-4 week old brains and retinas of wild-type

flies and ebony mutant flies, which have a darker pigment and impaired vision (31).

Although the levels of all three analytes were variable over time, the authors did report

10

Page 27: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

that these analytes were more abundant in the retina than in the brain and more abundant

in the heads of mutant ebony flies than the heads of wild-type flies.

Hardie and Hirsh expanded the number of neurotransmitters analyzed with HPLC

by quantifying dopamine, octopamine, tyramine, and serotonin in the brains and whole

heads of Drosophila white-eyed (white) mutants (32). They noted that nearly 75% of the

total dopamine within the white mutant head is located outside of the brain. In contrast,

the percentages of octopamine, tyramine, and serotonin present outside of the brain range

from only 1 to 37% when compared to the amount in the brain. The quantitative nature

of HPLC has been utilized to examine the role of tyramine in cocaine sensitization

studies of the inactive and the TβHM18

Drosophila mutants (10). The inactive mutant,

named for the low activity level of the mutant flies, was found to have approximately

60% less tyramine than wild-type flies, despite similar levels of dopamine. While these

mutants displayed expected behavioral responses to cocaine upon their first exposure,

with repeated cocaine exposure minimal behavioral sensitization to cocaine was

observed. The TβHM18

line has a null mutation in the gene that codes for tyramine β-

hydroxylase, the enzyme used to convert tyramine into octopamine. TβHM18

mutant flies

were found to have almost an order of magnitude greater amount of tyramine and near-

normal cocaine sensitization when compared to the wild-type fly, ruling out octopamine

as the biogenic amine contributing to this cocaine sensitization. These two comparisons

showed that tyramine plays a critical role in cocaine sensitization and later helped to

confirm the identity of two tyrosine decarboxylase genes (33).

The location and quantification of biogenic amines within the brain of genetic fly

mutants has been further investigated by the Meinertzhagen laboratory. They developed

11

Page 28: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

a method in the fly to quantify histamine (34), a transmitter known to be located in the

eyes of the fly, and compared it, along with the amount of dopamine and serotonin, in

white, brown, and scarlet mutants which are flies with three different eye-pigment

mutations (35). Since scarlet and brown are the two pigments that control fly eye color,

knocking out one pigment results in a fly with the other eye color, and a knockout of both

pigments results in no eye pigment, white mutants. They measured a significant decrease

(in some cases over 50%) in the neurotransmitters of all three Drosophila mutants when

compared to wild-type flies. Similar trends were observed in comparisons of wild-type

vs. white mutant houseflies, blowflies, and two species of the flesh fly, signifying that

many effects attributed to a mutant gene isolated in a white fly might be from the loss of

pigment itself and not the mutated gene. They also noted that in separations of wild-type

fly head homogenates, 71% of the total dopamine in the head was found in the brain, in

contrast to the results reported by Hardie and Hirsh for white mutant flies.

Mass spectrometry to study proteins and peptides. Mass-spectrometric studies of the

Drosophila proteome have used a variety of methods including matrix-assisted laser

desorption ionization time-of-flight (MALDI-TOF) mass spectrometry (36, 37) and ion-

trap mass spectrometry (38). In addition, a separation step is often added to the analysis

including reverse-phase liquid chromatography (38, 39), ion-mobility spectrometry (40-

42), or strong-cation-exchange chromatograpy (39, 42). The number of genes,

transcripts, and proteins that have been observed within the adult Drosophila are

summarized in Figure 1.3A.

Initial proteomic methods have been used to understand the basic biology of

Drosophila. Figure 1.3B shows a map by Taraszka et al. of the proteomes from three

12

Page 29: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.3. Mass spectrometric measurements of the Drosophila proteome. (A) Venn

diagram of the known adult Drosophila genome (thin black circles and numbers), mRNA

transcripts (thin grey circles and numbers), proteome (thick grey circles and bold

numbers), and the overlap between mRNA transcripts and proteome (bold black italics

numbers). Circle size corresponds to the number of known genes, transcripts, and

proteins listed below the circle. (B) LC-IMS-MS analysis of three digested individual

flies. Many of these features are common within all three individuals but some examples

of the differences have been labeled. Circled features designate peptides found in all

three individuals, boxes only two individuals, and triangles only one individual.

(Reprinted from (40, 41), with permission from the American Chemical Society).

13

Page 30: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

individual Drosophila heads identifying 197 proteins and found at least 101 proteins

present in all three samples (40). The other 96 proteins might not be expressed in every

sample, or the flies could have been using different proteins at the time of sacrifice.

More globally, differences have been observed in the Drosophila proteome lifespan. The

fly proteome has been investigated over sixty days, at seven day increments (42).

Approximately 1700 different proteins were identified and their changes in regulation

compared between three different age groups: young (1-21 day old flies), middle (22-42

day old flies), and old (43-60 day old flies). Of these comparisons, a significant

difference in protein regulation was observed for the young vs. middle-aged groups.

When the proteins experiencing an order of magnitude change or more in abundance

were considered, 30 proteins were down-regulated while 12 proteins were up-regulated in

the middle-aged group. These proteins were found to be associated with metabolism,

development, reproduction, or defense response.

Proteomic methods utilizing Drosophila have yielded insight into Parkinson’s

disease. Flies expressing either mutated A30P (39), mutated A53T (43), or normal

human α-synuclein genes (38) all display symptoms of Parkinson’s disease and have

been investigated. Symptoms include decreased locomotor ability, formation of Lewy

body-like inclusions in the brain, and degeneration of dopaminergic neurons with age.

The symptoms are most severe for the flies with the A30P point mutation, followed by

the A53T point mutated flies, and lastly the normal human α-synuclein mutated flies.

The three mutant fly types have had their proteomes compared to wild-type flies. Of

note, the levels of 49 proteins in the A30P flies and 24 proteins in the A53T flies were

significantly altered. Most of these proteins are associated with the actin cytoskeleton,

14

Page 31: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

mitochondria, and membrane. In the normal human α-synuclein mutant flies, only 12

protein changes were observed, mostly related to metabolism and cellular signaling.

These protein changes correlate with the severity of the Parkinson’s symptoms seen in

the mutated flies and might lead to general insight about the protein alterations associated

with this disease.

Complimentary to the genomic and proteomic work, Drosophila neuropeptides

have been investigated with MALDI-TOF mass spectrometry. Predel et al. characterized

the adult fly peptidome with this technique and were able to identify 32 neuropeptides in

the Drosophila CNS (37). Not only did this reveal the occurrence of these neuropeptides,

but it also depicted their morphological distribution. Recently, Kravitz and coworkers

improved upon this method by combining both MALDI-TOF mass spectrometry and

electrospray ionization quadrupole time-of-flight (QTOF) mass spectrometry. Using the

Drosophila GAL4-UAS system for targeted gene expression, subsets of cells were

genetically labeled to aid in sample preparation. They were able to identify 42

neuropeptides encoded by 18 different genes in adult Drosophila brain extract (44).

The larval Drosophila peptidome has been investigated with both one- and two-

dimensional (1D and 2D) capillary liquid chromatography (LC) followed by QTOF mass

spectrometry. Baggerman et al. identified 38 peptides using the 2D technique vs. 28

peptides using the 1D technique (45, 46). Their results demonstrate the increased

efficiency of 2D LC/QTOF over its 1D counterpart for Drosophila larvae.

Yew, Cody, and Kravitz studied Drosophila cuticular pheromones with real-time

mass spectrometry using atmospheric pressure ionization (47). This technique can

provide near instantaneous analysis of samples, and pheromones can be chemically

15

Page 32: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

investigated over a long period of time from live, awake Drosophila. Flies were

immobilized by a vacuum applied through a pipette tip and probed with a metal pin

attached to a micromanipulator. This allowed the fly to interact behaviorally with

surrounding flies. Pheromone levels were found to be increased in females vs. males, in

females after courtship, and as one moves closer to the genitals of the male fly. While

this work shows the spatial and temporal resolution of atmospheric pressure mass

spectrometry, it does lack the ability to measure analytes from inside of the fly.

Analytical Techniques for Measuring the Physiology of Intact Flies

Recently, analytical methods have been developed to record chemical

measurements in real-time from Drosophila larvae and from adult Drosophila. The

ability to acquire direct physiological information will help bridge the gap between

observed fly behavior and the chemical signaling pathways that underlie those behaviors.

Work has been done to develop technologies for manipulating individual Drosophila

embryos to study development as well. These tools will enable questions about the

functions of an individual organism to be addressed that whole-tissue homogenization

and pooled sampling methods cannot address.

Controlling individual fly embryo development using microfluidics. There is an

increasing interest in using Drosophila embryos to study mechanisms of development

and gene function. One powerful method of silencing a gene of interest is called

ribonucleic acid interference (RNAi). Cells are exposed to specifically designed double-

stranded RNA (dsRNA) that, once inside the cell, is cleaved into smaller dsRNA pieces

(siRNAs) by endogenous enzymes. The siRNA then binds to a RNA-induced silencing

16

Page 33: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

complex (RISC) where it becomes unwound. The unwound siRNA guides the RISC to

the corresponding messenger RNA (mRNA) whereby the RISC destroys the mRNA, thus

eliminating the coding of that particular gene and the gene’s subsequent function (48,

49). While using cells for high-throughput screens is useful, embryos are more ideal

model systems for studying development and gene function because they possess

physiological content with greater biological complexity; however, until recently,

performing RNAi on embryos was a tedious process that required a skilled technician to

individually inject each embryo by hand. Solgaard and colleagues have developed a

microfluidic device coupled with a computer-controlled injection system to inject

Drosophila embryos with dsRNA for high-throughput RNAi screens (50). This

microelectromechanical systems-based device has been automated to detect embryos on a

glass slide, followed by rapid injection of 60 pL RNAi aliquots into each embryo with

98% reliability. Although preliminary prototypes require initial manual injector

alignment to the device, it has potential for future development into a fully automated

process and has already been adapted for various embryo applications where controlled

microinjections of small molecules, such as drugs or proteins, are necessary (51, 52).

Microfluidic technologies have also been utilized to fabricate devices capable of

spatial and temporal control of developing Drosophila embryos. Ismagilov and

coworkers have used a ‘Y’ junction device to investigate a compensatory regulation

mechanism displayed by developing embryos towards external perturbations in

temperature (Figure 1.4 top) (53, 54). When the anterior and posterior sides of an

embryo were exposed to an extreme temperature gradient using two laminar streams held

at different temperatures, the warmer half of the embryo had a higher number of nuclei,

17

Page 34: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.4. Microfluidic device for the analysis of Drosophila embryos. The rate of

development in each half of the embryo exposed to a T-step is affected by temperature, as

demonstrated by the difference in nuclear density (number of nuclei in enlarged areas

shown underneath in yellow numbering). (A, B) Embryos exposed to a T-step of 20

°C/27 °C for 140 min. (A) Anterior half 20 °C, posterior half 27 °C. (B) Anterior half 27

°C, posterior half 20 °C. (C, D) Identical set-up to A and B with embryos exposed to a

greater T-step of 17 °C/27 °C for 150 min. In all images, higher nuclear density was

observed in the warmer half of the embryo. (Reprinted from (53), with permission from

Nature Publishing Group).

18

Page 35: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

and therefore was developing more rapidly, than the cooler half (Figure 1.4A, B). When

the temperature difference between the anterior and posterior sides was increased from

7°C to 10°C, the difference in the rate of development between the two sides increased as

well (Figure 1.4C, D). Also, the Even-skipped gene (a gene that codes for segmentation

during early embryonic development) was expressed sooner in the warmer region of the

embryo causing the usual 7-stripe segmentation pattern to develop in the wrong order.

Interestingly, despite the different developmental rates forced upon the two regions of the

embryo, when allowed to come to room temperature, the embryos displayed the

completed stripe pattern correctly and developed into normal larvae suggesting

Drosophila have a compensation mechanism to counteract extreme environmental

conditions during embryo development. This device has since been adapted to allow

easier attachment of the embryos (55). Continued modifications of the device that

enhance the ability to apply external gradients to an immobilized embryo will enable

future studies on the mechanisms of biochemical networks during development.

Individual larva measurements. There has been recent progress in the development of

techniques for measuring neurotransmitters from individual Drosophila larvae using a

combination of electrochemical detection and optogenetic stimulation methods. Fast-

scan cyclic voltammetry (FSCV) was employed because of its ability to measure rapid

changes in electroactive species like serotonin (56). Channelrhodopsin-2 is a light-

activated cation-selective ion channel that when placed under the control of a GAL4-

UAS system and crossed with flies of a driver line specific to serotonin (Tph-GAL4), will

produce transgenic larvae that release serotonin upon exposure to blue light (24).

Recently, Venton and colleagues utilized the transgenic larvae to measure serotonin

19

Page 36: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

release from neurons located in an isolated larva ventral nerve cord (VNC) using FSCV

with a microelectrode (Figure 1.5A) (57). The extracellular serotonin concentration in

the VNC was found to consistently vary between 280-640 nM during the duration of blue

light exposure (Figure 1.5B, C). Inhibition of the serotonin transporter with cocaine and

fluoxetine confirmed that the removal of serotonin from the extracellular space was due

to transport, and demonstrated the potential use of this model system for studying basic

serotonin signaling mechanisms.

The Drosophila larva model system has potential use in other areas as well. A

novel sampling technique has been developed to obtain nanoliter volumes of hemolymph

from individual Drosophila larvae for chemical analysis (58). Hemolymph contains

amino acids such as glutamate and glutamine that are thought to play a role in

neurodegeneration. This procedure extracts 50-300 nL of hemolymph from a single

Drosophila larva then, following derivatization with fluorescamine, its amino acid

content is quantified using CE with laser-induced fluorescence detection. In a

demonstration of this technique, Shippy and coworkers compared genderblind (gb)

larvae, mutants developed previously by collaborators (59) that contain approximately

half the normal extracellular glutamate concentration, to wild-type larvae. Overall the gb

mutants were found to have 38% lower glutamate levels than the wild-type larvae with 13

amino acids in total successfully separated and quantified from each larva’s hemolymph

(n = 10-17). These initial findings support the continued development of this technique

in quantifying amino acid levels from individual Drosophila hemolymph to better

understand their role in human disease.

20

Page 37: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.5. Measurements in Drosophila larvae following optogenetic stimulation. (A)

Diagram of neuromuscular anatomy of a third-instar larva. (B) Representative traces of

evoked peak serotonin concentration varying with blue light stimulus duration (2, 5, 10,

and 30 s). (C) Pooled data (mean ± SEM, n = 6) shows an increase in peak height with

increasing duration of blue light exposure. Peak height appears to plateau after 10 s; peak

height at 30 s is not significantly different from that at 10 s (Student’s t-test, 2 tailed, p =

0.78). (Reprinted from (57, 59), with permission from the Society for Neuroscience and

Elsevier).

21

Page 38: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Individual adult fly measurements. In addition to Drosophila embryos and larvae,

measurements from intact, whole flies have been accomplished. Calcium imaging has

been employed in conjunction with genetically encoded fluorescent proteins. Fluorescent

proteins that measure calcium changes (an accepted indicator of electrical activity) can be

genetically expressed in specific neurons to target a tissue of interest in the Drosophila

brain using the GAL4-UAS system (60). This methodology has been used to explore

several calcium-sensitive fluorescent proteins including cameleon 2.1, camgaroo 2, and

G-CaMP. Fiala and coworkers have labeled the mushroom body calyx and antennal lobe

structures, which are brain regions in the Drosophila CNS, with cameleon 2.1 and

measured odor-evoked calcium signals in vivo from both regions (61). Moreover, this

technique can be altered to target any brain region of interest for which a GAL4 driver

line exists (60).

Based upon previously published work on dissected mushroom bodies by Davis

and coworkers (62), the GAL4-UAS system was used to label the mushroom bodies with

camgaroo 2, and the intensity changes of the fluorescent Ca2+

reporter in response to

acetylcholine application were recorded in an intact fly (63). In addition, Axel and

colleagues employed two-photon calcium microscopy to image the antennae lobes of

flies that expressed G-CaMP in their projection neurons (64). Using this technique, they

were able to link odor-induced calcium changes to specific areas of the antennal lobe.

Each odor elicited a distinct pattern that appears to be conserved between different

organisms of the same fly genotype. Calcium imaging techniques could potentially be

used for quantitative investigation of olfactory learning and memory in the Drosophila

mushroom bodies and antennae lobes.

22

Page 39: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Scope of the Thesis

Electrochemical detection has been used for in vivo measurements of dopamine in

model systems such as rats, mice, and primates, but until very recently these

measurements were not feasible in an organism as small as Drosophila. My thesis

describes the development of electrochemical techniques for in vivo detection of

dopamine in the nanoliter-sized brain of adult Drosophila.

A method for quantifying the uptake of exogenously applied dopamine by the

Drosophila dopamine transporter is described in Chapter 2 and used in much of the

thesis. FSCV with a carbon-fiber microelectrode is used to monitor changes in dopamine

concentration in the adult fly CNS. Figure 1.6A and Figure 1.6B show in vivo dopamine

concentration traces demonstrating the change in extracellular dopamine before and after

treatment with cocaine, which is known to block uptake by the dopamine transporter. A

wild-type fly and a fumin (fmn) mutant fly that lacks a functional dopamine transporter

have been compared. While the peak dopamine concentration, [DA]max, increased 3-fold

in the wild-type fly following cocaine treatment, dopamine uptake remained unchanged

in the fmn mutant fly. When the [DA]max observed in multiple flies is averaged (Figure

1.6C), the [DA]max of untreated wild-type flies is significantly lower than for fmn mutant

flies. Interestingly, the [DA]max for cocaine treated wild-type flies is not significantly

different from the untreated fmn mutant flies. These measurements support existing

evidence that cocaine effectively blocks the Drosophila dopamine transporter and

validate the use of this in vivo fly method as a model system to study drug addiction

mechanisms.

23

Page 40: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 1.6. Investigating dopamine transporter function in adult Drosophila. (A)

Representative concentration trace of exogenously applied 1.0 mM dopamine in wild-

type Drosophila before (black line) and after (red line) cocaine application. An increase

in dopamine concentration in the adult wild-type fly was observed following a 5 min

exposure to 1.0 mM cocaine. Black arrow corresponds to a 1.0 s dopamine application

beginning at 5.0 s. (B) Representative concentration trace of exogenously applied 1.0

mM dopamine in the fmn mutant before (black line) and after (red line) cocaine

application. No significant change was observed in the adult fmn mutant fly. (C)

Comparison of baseline [DA]max for untreated wild-type and fmn mutant flies (mean ±

SEM; Student’s t-test, p = 0.02 (*), n = 9) and the treated wild-type fly after application

of 1.0 mM cocaine. The difference in [DA]max between wild-type flies treated with

cocaine and untreated fmn mutants or untreated wild-type flies is not significantly

different (mean ± SEM; Student’s t-test, p = 0.3 and 0.08, n = 6-9).

24

Page 41: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 3 describes the application of the in vivo microanalytical technique to

investigate the effects of cocaine, amphetamines, and methylphenidate on dopamine

clearance by the dopamine transporter in the fly. When under the influence of drugs of

abuse, fruit flies exhibit behavioral responses that are amazingly comparable to human

behaviors. The neurotransmitter dopamine has been shown to affect drug addiction

mechanisms. Following drug treatments, elevated levels of extracellular dopamine are

observed. This observation supports behavioral evidence that psychostimulants decrease

dopamine transporter function in Drosophila and is similar to results obtained in

mammalian systems. Furthermore, a study was developed to examine the effects of

methylphenidate on the mechanism of cocaine in the brain using the Drosophila model

system, and this is presented in Chapter 4.

Techniques for stimulating release of endogenous dopamine in the fly are

discussed in Chapter 5 including chemical, electrical, and optogenetic methods. While

the electroactive nature of dopamine makes in vivo electrochemistry an ideal approach for

measuring dopaminergic transmission in the brain, pH fluctuations associated with

dopamine release have been shown to interfere with electrochemically measured signals

in the rat. The fabrication and characterization of a pH microelectrode sensor for use in

the fly brain is described in Chapter 6. The ability of the pH sensor to monitor pH

changes following neurotransmitter release in real-time has been demonstrated in the

Drosophila CNS.

25

Page 42: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Makos, M. A., Kuklinski, N. J., Heien, M. L., Ewing, A. G., and Berglund, E. C.

(2009) Chemical measurements in Drosophila, TrAC, Trends Anal. Chem. 28,

1223-1234.

2. Waddell, S., and Quinn, W. G. (2001) What can we teach Drosophila? What can

they teach us?, Trends Genet. 17, 719-726.

3. Schwaerzel, M., Monastirioti, M., Scholz, H., Friggi-Grelin, F., Birman, S., and

Heisenberg, M. (2003) Dopamine and octopamine differentiate between aversive

and appetitive olfactory memories in Drosophila, J. Neurosci. 23, 10495-10502.

4. Kim, Y.-C., Lee, H.-G., and Han, K.-A. (2007) D1 dopamine receptor dDA1 is

required in the mushroom body neurons for aversive and appetitive learning in

Drosophila, J. Neurosci. 27, 7640-7647.

5. Gerber, B., and Stocker, R. F. (2007) The Drosophila larva as a model for

studying chemosensation and chemosensory learning: a review, Chem. Senses 32,

65-89.

6. Rubin, G. M., Yandell, M. D., Wortman, J. R., Miklos, G. L. G., Nelson, C. R.,

Hariharan, I. K., Fortini, M. E., Li, P. W., Apweiler, R., Fleischmann, W., Cherry,

J. M., Henikoff, S., Skupski, M. P., Misra, S., Ashburner, M., Birney, E.,

Boguski, M. S., Brody, T., Brokstein, P., Celniker, S. E., Chervitz, S. A., Coates,

D., Cravchik, A., Gabrielian, A., Galle, R. F., Gelbart, W. M., George, R. A.,

Goldstein, L. S. B., Gong, F. C., Guan, P., Harris, N. L., Hay, B. A., Hoskins, R.

A., Li, J. Y., Li, Z. Y., Hynes, R. O., Jones, S. J. M., Kuehl, P. M., Lemaitre, B.,

Littleton, J. T., Morrison, D. K., Mungall, C., O'Farrell, P. H., Pickeral, O. K.,

Shue, C., Vosshall, L. B., Zhang, J., Zhao, Q., Zheng, X. Q. H., Zhong, F., Zhong,

W. Y., Gibbs, R., Venter, J. C., Adams, M. D., and Lewis, S. (2000) Comparative

genomics of the eukaryotes, Science 287, 2204-2215.

7. Coulom, H., and Birman, S. (2004) Chronic exposure to rotenone models sporadic

Parkinson's disease in Drosophila melanogaster, J. Neurosci. 24, 10993-10998.

8. Rein, K., Zöckler, M., Mader, M. T., Grübel, C., and Heisenberg, M. (2002) The

Drosophila standard brain, Curr. Biol. 12, 227-231.

9. Monastirioti, M. (1999) Biogenic amine systems in the fruit fly Drosophila

melanogaster, Microsc. Res. Tech. 45, 106-121.

10. McClung, C., and Hirsh, J. (1999) The trace amine tyramine is essential for

sensitization to cocaine in Drosophila, Curr. Biol. 9, 853-860.

11. Bergquist, J., Sciubisz, A., Kaczor, A., and Silberring, J. (2002) Catecholamines

and methods for their identification and quantitation in biological tissues and

fluids, J. Neurosci. Methods 113, 1-13.

12. Bainton, R. J., Tsai, L. T. Y., Singh, C. M., Moore, M. S., Neckameyer, W. S.,

and Heberlein, U. (2000) Dopamine modulates acute responses to cocaine,

nicotine, and ethanol in Drosophila, Curr. Biol. 10, 187-194.

13. Li, H., Chaney, S., Forte, M., and Hirsh, J. (2000) Ectopic G-protein expression in

dopamine and serotonin neurons blocks cocaine sensitization in Drosophila

melanogaster, Curr. Biol. 10, 211-214.

14. Kume, K., Kume, S., Park, S. K., Hirsh, J., and Jackson, F. R. (2005) Dopamine is

a regulator of arousal in the fruit fly, J. Neurosci. 25, 7377-7384.

26

Page 43: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

15. Hoyer, S. C., Eckart, A., Herrel, A., Zars, T., Fischer, S. A., Hardie, S. L., and

Heisenberg, M. (2008) Octopamine in male aggression of Drosophila, Curr. Biol.

18, 159-167.

16. Adams, M. D., Celniker, S. E., Holt, R. A., Evans, C. A., Gocayne, J. D.,

Amanatides, P. G., Scherer, S. E., Li, P. W., Hoskins, R. A., Galle, R. F., George,

R. A., Lewis, S. E., Richards, S., Ashburner, M., Henderson, S. N., Sutton, G. G.,

Wortman, J. R., Yandell, M. D., Zhang, Q., Chen, L. X., Brandon, R. C., Rogers,

Y.-H. C., Blazej, R. G., Champe, M., Pfeiffer, B. D., Wan, K. H., Doyle, C.,

Baxter, E. G., Helt, G., Nelson, C. R., Gabor Miklos, G. L., Abril, J. F., Agbayani,

A., An, H.-J., Andrews-Pfannkoch, C., Baldwin, D., Ballew, R. M., Basu, A.,

Baxendale, J., Bayraktaroglu, L., Beasley, E. M., Beeson, K. Y., Benos, P. V.,

Berman, B. P., Bhandari, D., Bolshakov, S., Borkova, D., Botchan, M. R., Bouck,

J., Brokstein, P., Brottier, P., Burtis, K. C., Busam, D. A., Butler, H., Cadieu, E.,

Center, A., Chandra, I., Cherry, J. M., Cawley, S., Dahlke, C., Davenport, L. B.,

Davies, P., Pablos, B. d., Delcher, A., Deng, Z., Mays, A. D., Dew, I., Dietz, S.

M., Dodson, K., Doup, L. E., Downes, M., Dugan-Rocha, S., Dunkov, B. C.,

Dunn, P., Durbin, K. J., Evangelista, C. C., Ferraz, C., Ferriera, S., Fleischmann,

W., Fosler, C., Gabrielian, A. E., Garg, N. S., Gelbart, W. M., Glasser, K.,

Glodek, A., Gong, F., Gorrell, J. H., Gu, Z., Guan, P., Harris, M., Harris, N. L.,

Harvey, D., Heiman, T. J., Hernandez, J. R., Houck, J., Hostin, D., Houston, K.

A., Howland, T. J., Wei, M.-H., Ibegwam, C., Jalali, M., Kalush, F., Karpen, G.

H., Ke, Z., Kennison, J. A., Ketchum, K. A., Kimmel, B. E., Kodira, C. D., Kraft,

C., Kravitz, S., Kulp, D., Lai, Z., Lasko, P., Lei, Y., Levitsky, A. A., Li, J., Li, Z.,

Liang, Y., Lin, X., Liu, X., Mattei, B., McIntosh, T. C., McLeod, M. P.,

McPherson, D., Merkulov, G., Milshina, N. V., Mobarry, C., Morris, J., Moshrefi,

A., Mount, S. M., Moy, M., Murphy, B., Murphy, L., Muzny, D. M., Nelson, D.

L., Nelson, D. R., Nelson, K. A., Nixon, K., Nusskern, D. R., Pacleb, J. M.,

Palazzolo, M., Pittman, G. S., Pan, S., Pollard, J., Puri, V., Reese, M. G., Reinert,

K., Remington, K., Saunders, R. D., nbsp, C., Scheeler, F., Shen, H., Shue, B. C.,

Sid, eacute, n-Kiamos, I., Simpson, M., Skupski, M. P., Smith, T., Spier, E.,

Spradling, A. C., Stapleton, M., Strong, R., Sun, E., Svirskas, R., Tector, C.,

Turner, R., Venter, E., Wang, A. H., Wang, X., Wang, Z.-Y., Wassarman, D. A.,

Weinstock, G. M., Weissenbach, J., Williams, S. M., Woodage, T., Worley, K. C.,

Wu, D., Yang, S., Yao, Q. A., Ye, J., Yeh, R.-F., Zaveri, J. S., Zhan, M., Zhang,

G., Zhao, Q., Zheng, L., Zheng, X. H., Zhong, F. N., Zhong, W., Zhou, X., Zhu,

S., Zhu, X., Smith, H. O., Gibbs, R. A., Myers, E. W., Rubin, G. M., and Venter,

J. C. (2000) The genome sequence of Drosophila melanogaster, Science 287,

2185-2195.

17. Sokolowski, M. B. (2001) Drosophila: Genetics meets behavior, Nat. Rev. Genet.

2, 879-890.

18. Scholz, H., Ramond, J., Singh, C. M., and Heberlein, U. (2000) Functional

ethanol tolerance in Drosophila, Neuron 28, 261-271.

19. Panda, S., Hogenesch, J. B., and Kay, S. A. (2002) Circadian rhythms from flies

to human, Nature 417, 329-335.

20. Lee, H.-G., Kim, Y.-C., Dunning, J. S., and Han, K.-A. (2008) Recurring ethanol

exposure induces disinhibited courtship in Drosophila, PLoS One 3, e1391.

27

Page 44: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

21. Brand, A. H., and Perrimon, N. (1993) Targeted gene expression as a means of

altering cell fates and generating dominant phenotypes, Development 118, 401-

415.

22. Duffy, J. B. (2002) GAL4 system in Drosophila: A fly geneticist's Swiss army

knife, Genesis 34, 1-15.

23. Friggi-Grelin, F., Coulom, H., Meller, M., Gomez, D., Hirsh, J., and Birman, S.

(2003) Targeted gene expression in Drosophila dopaminergic cells using

regulatory sequences from tyrosine hydroxylase, J. Neurobiol. 54, 618-627.

24. Schroll, C., Riemensperger, T., Bucher, D., Ehmer, J., Voller, T., Erbguth, K.,

Gerber, B., Hendel, T., Nagel, G., Buchner, E., and Fiala, A. (2006) Light-

induced activation of distinct modulatory neurons triggers appetitive or aversive

learning in Drosophila larvae, Curr. Biol. 16, 1741-1747.

25. Marsh, J. L., and Thompson, L. M. (2004) Can flies help humans treat

neurodegenerative diseases?, BioEssays 26, 485-496.

26. Celotto, A. M., and Palladino, M. J. (2005) A "model" model system to study

neurodegeneration, Mol. Interventions 5, 292-303.

27. Powell, P. R., Paxon, T. L., Han, K.-A., and Ewing, A. G. (2005) Analysis of

biogenic amine variability among individual fly heads with micellar electrokinetic

capillary chromatography-electrochemical detection, Anal. Chem. 77, 6902-6908.

28. Edgar, J. S., Pabbati, C. P., Lorenz, R. M., He, M., Fiorini, G. S., and Chiu, D. T.

(2006) Capillary electrophoresis separation in the presence of an immiscible

boundary for droplet analysis, Anal. Chem. 78, 6948-6954.

29. Ream, P. J., Suljak, S. W., Ewing, A. G., and Han, K.-A. (2003) Micellar

electrokinetic capillary chromatography- electrochemical detection for analysis of

biogenic amines in Drosophila melanogaster, Anal. Chem. 75, 3972-3978.

30. Paxon, T. L., Powell, P. R., Lee, H.-G., Han, K.-A., and Ewing, A. G. (2005)

Microcolumn separation of amine metabolites in the fruit fly, Anal. Chem. 77,

5349-5355.

31. Ramadan, H., Alawi, A. A., and Alawi, M. A. (1993) Catecholamines in

Drosophila melanogaster (wild type and ebony mutant) decuticalarized retinas

and brains, Cell Biol. Int. 17, 765-772.

32. Hardie, S. L., and Hirsh, J. (2006) An improved method for the separation and

detection of biogenic amines in adult Drosophila brain extracts by high

performance liquid chromatography, J. Neurosci. Methods 153, 243-249.

33. Cole, S. H., Carney, G. E., McClung, C. A., Willard, S. S., Taylor, B. J., and

Hirsh, J. (2005) Two functional but noncomplementing Drosophila tyrosine

decarboxylase genes: Distinct roles for neural tyramine and octopamine in female

fertility, J. Biol. Chem. 280, 14948-14955.

34. Borycz, J., Vohra, M., Tokarczyk, G., and Meinertzhagen, I. A. (2000) The

determination of histamine in the Drosophila head, J. Neurosci. Methods 101,

141-148.

35. Borycz, J., Borycz, J. A., Kubow, A., Lloyd, V., and Meinertzhagen, I. A. (2008)

Drosophila ABC transporter mutants white, brown, and scarlet have altered

contents and distribution of biogenic amines in the brain, J. Exp. Biol. 211, 3454-

3466.

28

Page 45: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

36. Baldwin, M. A., Medzihradszky, K. F., Lock, C. M., Fisher, B., Settineri, T. A.,

and Burlingame, A. L. (2001) Matrix-assisted laser desorption/ionization coupled

with quadrupole/orthogonal acceleration time-of-flight mass spectrometry for

protein discovery, identification, and structural analysis, Anal. Chem. 73, 1707-

1720.

37. Predel, R., Wegener, C., Russell, W. K., Tichy, S. E., Russell, D. H., and

Nachman, R. J. (2004) Peptidomics of CNS-associated neurohemal systems of

adult Drosophila melanogaster: A mass spectrometric survey of peptides from

individual flies, J. Comp. Neurol. 474, 379-392.

38. Xun, Z., Kaufman, T. C., and Clemmer, D. E. (2008) Proteome response to the

panneural expression of human wild-type a-synuclein: A Drosophila model of

Parkinson's disease, J. Proteome Res. 7, 3911-3921.

39. Xun, Z., Sowell, R. A., Kaufman, T. C., and Clemmer, D. E. (2007) Protein

expression in a Drosophila model of Parkinson’s disease, J. Proteome Res., 348–

357.

40. Taraszka, J. A., Gao, X., Valentine, S. J., Sowell, R. A., Koeniger, S. L., Miller,

D. F., Kaufman, T. C., and Clemmer, D. E. (2005) Proteome profiling for

assessing diversity: Analysis of individual heads of Drosophila melanogaster

using LC-ion mobility-MS, J. Proteome Res. 4, 1238-1247.

41. Taraszka, J. A., Kurulugama, R., Sowell, R. A., Valentine, S. J., Koeniger, S. L.,

Arnold, R. J., Miller, D. F., Kaufman, T. C., and Clemmer, D. E. (2005) Mapping

the proteome of Drosophila melanogaster: Analysis of embryos and adult heads

by LC-IMS-MS methods, J. Proteome Res. 4, 1223-1237.

42. Sowell, R. A., Hersberger, K. E., Kaufman, T. C., and Clemmer, D. E. (2007)

Examining the proteome of Drosophila across organism lifespan, J. Proteome

Res. 6, 3637-3647.

43. Xun, Z., Sowell, R. A., Kaufman, T. C., and Clemmer, D. E. (2008) Quantitative

proteomics of a presymptomatic A53T α-synuclein Drosophila model of

Parkinson disease, Mol. Cell. Proteomics 7, 1191-1203.

44. Yew, J. Y., Wang, Y., Barteneva, N., Dikler, S., Kutz-Naber, K. K., Li, L., and

Kravitz, E. A. (2009) Analysis of neuropeptide expression and localization in

adult Drosophila melanogaster central nervous system by affinity cell-capture

mass spectrometry, J. Proteome Res. 8, 1271-1284.

45. Baggerman, G., Cerstiaens, A., De Loof, A., and Schoofs, L. (2002) Peptidomics

of the larval Drosophila melanogaster central nervous system, J. Biol. Chem. 277,

40368-40374.

46. Baggerman, G., Boonen, K., Verleyen, P., De Loof, A., and Schoofs, L. (2005)

Peptidomic analysis of the larval Drosophila melanogaster central nervous

system by two-dimensional capillary liquid chromatography quadrupole time-of-

flight mass spectrometry, J. Mass Spectrom. 40, 250-260.

47. Yew, J. Y., Cody, R. B., and Kravitz, E. A. (2008) Cuticular hydrocarbon analysis

of an awake behaving fly using direct analysis in real-time time-of-flight mass

spectrometry, Proc. Natl. Acad. Sci. U.S.A. 105, 7135-7140.

48. Fire, A., Xu, S. Q., Montgomery, M. K., Kostas, S. A., Driver, S. E., and Mello,

C. C. (1998) Potent and specific genetic interference by double-stranded RNA in

Caenorhabditis elegans, Nature 391, 806-811.

29

Page 46: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

49. Hannon, G. J. (2002) RNA interference, Nature 418, 244-251.

50. Zappe, S., Fish, M., Scott, M. P., and Solgaard, O. (2006) Automated MEMS-

based Drosophila embryo injection system for high-throughput RNAi screens,

Lab Chip 6, 1012-1019.

51. Wang, W., Liu, X., Gelinas, D., Ciruna, B., and Sun, Y. (2007) A fully automated

robotic system for microinjection of zebrafish embryos, PLoS One 2, e862.

52. Cornell, E., Fisher, W. W., Nordmeyer, R., Yegian, D., Dong, M., Biggin, M. D.,

Celniker, S. E., and Jin, J. (2008) Automating fruit fly Drosophila embryo

injection for high throughput transgenic studies, Rev. Sci. Instrum. 79, e013705.

53. Lucchetta, E. M., Lee, J. H., Fu, L. A., Patel, N. H., and Ismagilov, R. F. (2005)

Dynamics of Drosophila embryonic patterning network perturbed in space and

time using microfluidics, Nature 434, 1134-1138.

54. Lucchetta, E. M., Munson, M. S., and Ismagilov, R. F. (2006) Characterization of

the local temperature in space and time around a developing Drosophila embryo

in a microfluidic device, Lab Chip 6, 185-190.

55. Dagani, G. T., Monzo, K., Fakhoury, J. R., Chen, C. C., Sisson, J. C., and Zhang,

X. J. (2007) Microfluidic self-assembly of live Drosophila embryos for versatile

high-throughput analysis of embryonic morphogenesis, Biomed. Microdevices 9,

681-694.

56. Baur, J. E., Kristensen, E. W., May, L. J., Wiedemann, D. J., and Wightman, R.

M. (1988) Fast-scan voltammetry of biogenic amines, Anal. Chem. 60, 1268-

1272.

57. Borue, X., Cooper, S., Hirsh, J., Condron, B., and Venton, B. J. (2009)

Quantitative evaluation of serotonin release and clearance in Drosophila, J.

Neurosci. Methods 179, 300-308.

58. Augustin, H., Grosjean, Y., Chen, K., Sheng, Q., and Featherstone, D. E. (2007)

Nonvesicular release of glutamate by glial xCT transporters suppresses glutamate

receptor clustering in vivo, J. Neurosci. 27, 111-123.

59. Piyankarage, S. C., Augustin, H., Grosjean, Y., Featherstone, D. E., and Shippy,

S. A. (2008) Hemolymph amino acid analysis of individual Drosophila larvae,

Anal. Chem. 80, 1201-1207.

60. Fiala, A., and Spall, T. (2003) In vivo calcium imaging of brain activity in

Drosophila by transgenic cameleon expression, Sci. STKE 2003, pl6.

61. Fiala, A., Spall, T., Diegelmann, S., Eisermann, B., Sachse, S., Devaud, J.-M.,

Buchner, E., and Galizia, C. G. (2002) Genetically expressed cameleon in

Drosophila melanogaster is used to visualize olfactory information in projection

neurons, Curr. Biol. 12, 1877-1884.

62. Yu, D. H., Baird, G. S., Tsien, R. Y., and Davis, R. L. (2003) Detection of

calcium transients in Drosophila mushroom body neurons with camgaroo

reporters, J. Neurosci. 23, 64-72.

63. Wright, N. J. D. (2006) An in vivo technique for pharmacological manipulation of

Drosophila brain during optical recording, J. Neurosci. Methods 155, 77-80.

64. Wang, J. W., Wong, A. M., Flores, J., Vosshall, L. B., and Axel, R. (2003) Two-

photon calcium imaging reveals an odor-evoked map of activity in the fly brain,

Cell 112, 271-282.

30

Page 47: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 2: In Vivo Electrochemical Measurements of Exogenously

Applied Dopamine in Drosophila melanogaster*

*Reproduced with permission from Makos, M. A., Kim, Y.-C., Han, K.-A., Heien, M. L.,

and Ewing, A. G. (2009) In Vivo Electrochemical Measurements of Exogenously Applied

Dopamine in Drosophila melanogaster, Anal. Chem. 81, 1848-1854. © 2009 American

Chemical Society.

31

Page 48: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

The field of in vivo electrochemistry in the brain began in the 1970’s with Ralph

Adams pioneering the detection of electroactive species. His group measured

neurochemicals in the brains of anesthetized rats with carbon electrodes using cyclic

voltammetry and chronoamperometry (1, 2). Subsequently, background-subtracted fast-

scan cyclic voltammetry (FSCV) coupled with carbon-fiber microelectrodes has been

developed and extensively used as an analytical technique for in vivo measurements of

electroactive neurotransmitters (3-7). In vivo electrochemistry has mainly focused on the

rat as the primary model system to address fundamental questions regarding

neurotransmission mechanisms (8-10). While similar studies have been conducted in

other model systems such as mice and primates, microanalytical methods for in vivo

studies in a model organism as small as Drosophila melanogaster have remained

undeveloped (11-14).

Drosophila has been traditionally used as a model organism for genetic research

because its genetic manipulation is relatively straightforward, and the genome contains

fewer genetic redundancies compared to the mammalian genome, facilitating the

identification of functions of individual genes or molecules (15, 16). Drosophila has a

short life cycle (12-14 days) and thus it is quite feasible to generate mutants that are

genetically homogeneous more quickly in comparison to other model organisms used for

in vivo electrochemistry including rats and mice. Although Drosophila has a relatively

simple nervous system containing approximately 200,000 neurons, it exhibits many of

the same higher-order brain functions as vertebrates at the molecular, cellular, and

behavioral levels. Flies are capable of learning from prior experiences and storing

32

Page 49: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

learned information (15, 16). Many monoamines including dopamine, serotonin,

tyramine, and histamine that regulate human physiological processes are also found in the

Drosophila central nervous system (CNS). In addition, octopamine, specific to

invertebrates, has similar roles to mammalian norepinephrine (17).

The neurotransmitter dopamine has been implicated in physiological human

processes including attention, motivation, emotion, sleep, and addiction (18-21). In

particular, the reinforcing properties of psychostimulants such as cocaine and

amphetamine that block the dopamine transporter or other addictive substances such as

ethanol and nicotine involve an elevated level of extracellular dopamine (18, 22-24).

However, the underlying neuronal mechanisms concerning how dopamine affects

tolerance and addiction remain as yet poorly understood.

Constant-potential amperometry, chronoamperometry, and FSCV are the common

electrochemical techniques that have been used to detect dopamine in vivo using model

systems (25-27). While constant-potential amperometry has the advantage of excellent

temporal resolution over most other electrochemical techniques, its lack of chemical

specificity makes it useful only in a system where the identity of the analyte is known or

when it is combined with a more chemically selective technique (10, 26, 28).

Voltammetry is one of the most widely accepted techniques used to identify single

electrochemical substances. Specifically, background-subtracted FSCV is a dominant

technique used for neurotransmitter detection in vivo because of its chemical selectivity,

relatively high sensitivity, and sub-second temporal resolution (28-30).

This chapter reports on the development of these microanalytical techniques for in

vivo electrochemical detection in the Drosophila CNS. A microsurgery procedure is

33

Page 50: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

explained that allows an electrode to be inserted into an immobilized fly brain while the

fly is kept viable for experimentation. Voltammetry has been carried out to monitor

dopamine in the adult brain of the wild-type fly vs. the mutant fly lacking functional

dopamine transporters. Significant differences are detectable for the clearance of

exogenously applied dopamine by the transporter which supports the validity of the new

method described here.

Methods

Chemicals. All chemicals were used as received and purchased from Sigma (St. Louis,

MO) unless otherwise stated. Adult-hemolymph like (AHL) saline (108 mM NaCl, 5

mM KCl, 2 mM CaCl2, 8.2 mM MgCl2, 4 mM NaHCO3, 1 mM NaH2PO4, 5 mM

trehalose (Fluka BioChemika, Buchs, Switzerland), 10 mM sucrose, 5 mM Trizma

base , pH 7.5) was made using ultrapure (18 MΩ·cm) water and filtered through a 0.2-

μm filter (31). All collagenase, KCl, propidium iodide (PI), dopamine, (+) cocaine, and

tetrodotoxin (TTX) solutions were prepared using AHL saline.

In vivo Drosophila preparation. The Canton-S strain of Drosophila melanogaster was

used for the wild-type fly in this chapter. The transgenic flies carrying tyrosine

hydroxylase TH-GAL4 and UAS-mCD:GFP (membrane tethered green fluorescent

protein) were used to visualize the dopamine neurons (32, 33). The fumin (fmn) mutant

has a genetic lesion abolishing the dopamine transporter function. The genetic

background of the w;fmn mutant was replaced with the Canton-S background (34). All

flies were maintained at 25 °C on a standard cornmeal-agar medium, and 4 to 7 day-old

male flies were used for experiments. For in vivo imaging and voltammetry, the flies

34

Page 51: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

were immobilized on ice and mounted in a homemade collar (38.1 mm diameter concave

plexiglass disk with a 1.0 mm hole in the center) with low melt agarose (Fisher Scientific,

Pittsburgh, PA). Microsurgery was performed on a stereoscope (Olympus SZ60,

Melville, NY) using small dissection scissors and forceps (World Precision Instruments,

Sarasota, FL). After the cuticle was removed from the top portion of the head to expose

the brain, the head was covered with 0.1% collagenase solution for 30 min to relax the

extracellular matrix in the brain and then rinsed and covered with AHL saline. The

images were acquired using an Olympus SZX10 stereomicroscope and an Olympus DP71

digital camera (Figure 2.1A) or a Leica MZ16 stereomicroscope and a Leica DFC290

digital camera (Figure 2.1B and 2.1C; Mannheim, Germany).

Electrochemical measurements. Carbon-fiber microelectrodes were fabricated as

previously described (6). Briefly, a single 5-μm diameter carbon fiber (Amoco,

Greenville, SC) was aspirated into a borosilicate glass capillary (B120-69-10, Sutter

Instruments, Novato, CA), and the capillary was pulled using a regular glass capillary

puller (P-97, Sutter Instruments). Electrical contact was made by back-filling the

capillary with silver paint (4922N DuPont, Delta Technologies Ltd., Stillwater, MN) and

inserting a tungsten wire. To form a cylindrical electrode, the carbon fiber was cut to a

length of 40-50 μm, as measured from the glass junction. Electrode tips were dipped into

epoxy (Epo-Tek, Epoxy Technology, Billerica, MA) for 30 s to ensure a good seal

between the fiber and the glass and then dipped into acetone for 15 s to remove epoxy

from the exposed carbon fiber. A Ag/AgCl reference electrode was made by

chlorodizing a silver wire (0.25 mm diameter, 99.999% purity, Alfa Aesar, Ward Hill,

35

Page 52: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

MA) in bleach overnight. Micropipet injectors were fabricated by pulling glass

capillaries in a glass capillary puller to an opening of approximately 5 μm.

Electrochemical data were collected using an Axopatch 200B Amplifier (Axon

Instruments, Foster City, CA) and two data acquisition boards (PCI-6221, National

Instruments, Austin, TX) run by the TH 1.0 CV program (ESA, Chelmsford, MA) (35).

For amperometric experiments, a constant potential (+750 mV) was first applied to the

working electrode with respect to the reference electrode for at least 15 min to stabilize

the background current. All cyclic voltammograms were obtained using a triangular

waveform (scanned -0.6 V to +1.0 V vs. Ag/AgCl at 200 V/s) repeated every 100 ms

(low pass Bessel filter at 5 kHz). Prior to voltammetric experiments, all electrodes were

cycled (-0.6 V to +1.0 V at 200 V/s) for at least 15 min to stabilize the background

current. Electrochemical responses were plotted and statistical analysis performed using

Prism 3.0 (GraphPad Software, La Jolla, CA).

All electrodes were positioned under a Leica MZ16 stereomicroscope using

micromanipulators (421 series, Newport, Irvine, CA) on top of a Newport BenchTop

Vibration Isolation System. Either a single-barrel glass micropipet or a three-barrel glass

micropipet (3B120F-6, World Precision Instruments) was used to exogenously apply the

dopamine solutions. For the three-barrel micropipet, each barrel was individually

coupled to a microinjection system (Picospritzer II, General Valve Corporation, Fairfield,

NJ) using a PolyFil apparatus (World Precision Instruments).

Sample preparation for confocal imaging. Transgenic TH-GAL4/UAS-GFP flies were

used to visualize the dopamine neurons. The TH-GAL4/UAS-GFP fly brain was exposed

as described above then stained with PI (100 μg/mL) for 20 min. Prior to treatment with

36

Page 53: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

PI, cell death control flies were treated with 1.0 M KCl for 10 min to model the

fluorescence that would occur from PI in a fly brain containing cells that were no longer

viable. After three washes in AHL saline (10 min each), phosphate-buffered saline (PBS)

containing 4% paraformaldehyde was applied for 20 min. Three more washes were done

in PBS only (10 min each) before the brain was dissected out and mounted in Vectashield

mounting media (Vector Laboratories, Burlingame, CA). Fluorescence images of

Drosophila brains stained with PI (λex 536 nm, λem 617 nm) and labeled with GFP (λex

488 nm, λem 507 nm) were acquired using a Leica TCS SP5 laser-scanning confocal

microscope with a 20x objective lens (Figure 2.2).

Results and Discussion

Drosophila preparation and set-up for in vivo measurements. Electrochemical

methods provide a new tool for studying electroactive neurotransmitters in Drosophila. I

am particularly interested in studying dopamine neurotransmission since it plays crucial

roles in numerous CNS functions in Drosophila as in mammals (17). In the Drosophila

brain, multiple clusters of dopamine neuronal cell bodies are spread throughout the outer

layer of the brain cortex and innervate many brain regions. In particular, the dopamine

neuronal cluster in the protocerebral anterior medial (PAM) brain area project to the

nearby mushroom body structure that is crucial for many higher-order neuronal functions

including learning and memory (36-38). Thus, I focused on the PAM neurons for in vivo

analysis of dopamine neurotransmission. To place microelectrodes in the area where the

PAM neurons are located, a microsurgery procedure was developed. A single adult fly

was immobilized in a homemade fly collar using agarose applied to the body and the

37

Page 54: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

bottom portion of the head (Figure 2.1A), leaving the upper portion of the head

uncovered and positioned for dissection. The cuticle was then removed, and the brain

was kept bathed in AHL saline (Figure 2.1B). The salts in the AHL solution were at

physiological concentrations, keeping the immobilized fly viable for 1.5 - 2.5 h which is

sufficient time to perform electrochemical measurements (31). A micromanipulator was

used to guide the cylindrical working electrode into the PAM region. The micropipets

used for dopamine application throughout this chapter were positioned above the PAM

area, approximately 10 μm from the working electrode (Figure 2.1B inset). The

reference electrode was submerged in the AHL saline. Fluorescence microscopy was

used to visualize the location of the PAM dopamine neurons in the brain of the transgenic

TH-GAL4/UAS-GFP fly which expresses GFP in dopamine neurons. The PAM area

represents the largest cluster of dopamine neurons and is easily identifiable (36). Figure

2.1C shows a representative fluorescence image of a dissected brain with GFP-labeled

dopamine neurons. The white box outlines the exposed brain regions where PAM

neurons are clearly visible in green, while the fluorescent cells below the box represent

other dopamine neuronal clusters. Experiments to investigate dopamine uptake were

performed in the PAM dopamine neuronal area.

Viability of Drosophila following microsurgery preparation. Confocal fluorescence

microscopy was used to verify that cells in the brain remain viable following the

microsurgery preparation described above for in vivo electrochemistry in Drosophila.

Following microsurgery, brains were prepared by incubation in PI, a fluorescent dye

which indicates damaged cell membranes. For comparison, control flies were treated

with 1.0 M KCl prior to PI incubation to initiate cell death. A fluorescence image of the

38

Page 55: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.1. Images of Drosophila taken during microsurgery. (A) Fly immobilized in a

homemade fly collar (Scale bar = 500 µm). (B) Fly after cuticle has been removed. The

exposed brain area with the PAM dopamine neurons is outlined by the black box (Scale

bar = 100 µm, electrode and injector not to scale). Inset: Schematic showing relative

electrode and micropipet injector placement for experiments. (C) Fluorescence image

highlighting GFP-labeled dopaminergic neurons. White box outlines the PAM region

(Scale bar = 100 µm).

39

Page 56: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

brain of an adult TH-GAL4/UAS-GFP fly (Figure 2.2A) contains very little red

fluorescence compared to the image of the apoptotic control fly (Figure 2.2B) which

provides one indication that the Drosophila brain does remain viable following the

dissection preparation for in vivo measurements.

Measuring exogenously applied dopamine in Drosophila. In previous studies,

electrochemical detection with FSCV has been used to monitor in vivo dopamine

concentrations in rats (3). Exogenously applied dopamine can be measured at the surface

of a carbon-fiber microelectrode inserted into the PAM area of the Drosophila system.

To further characterize dopamine detection in the PAM area, color plots were used to

display FSCV data. In these experiments, small amounts of a dopamine solution were

ejected in the area near the electrode, and voltammetry was used to quantify the

dopamine changes in the brain and to track its temporal characteristics. Here, 1.0 mM

dopamine was exogenously applied to the adult wild-type brain using a single micropipet

injector, and a microelectrode was used for dopamine detection in the PAM area. A

false-color representation of current (Figure 2.3A) allows one to visualize cyclic

voltammograms over time. The oxidation of dopamine is represented in green while blue

corresponds to the reduction of the orthoquinone, allowing discrimination of a particular

analyte from other species that may be present in the same brain region. Cyclic

voltammetry can be used to identify electroactive species based on the potential at which

oxidation occurs and the overall shape of the wave (10, 28, 29). For example, the cyclic

voltammogram in Figure 2.3B is a background-subtracted average of ten successive

cyclic voltammograms taken at the peak current from the color plot (Figure 2.3A). By

40

Page 57: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.2. Confocal fluorescence micrographs of intact brains from adult transgenic

TH-GAL4/UAS-GFP flies. GFP (green) was used to visualize dopamine neurons; PI

(red) was used to stain damaged cells. Scale bar = 100 μm. (A) Brain demonstrating cell

viability following microsurgery procedure. (B) Brain incubated with 1.0 M KCl as

control model of a brain containing cells that are no longer viable for comparison.

41

Page 58: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

inspection, the shape of the voltammogram and peak potential leads us to conclude that

the increase in current in Figure 2.3A corresponds to the measurement of dopamine.

Finally, the current can be converted to dopamine concentration using in vitro electrode

calibration (Figure 2.3C), and the time required for the concentration to decrease to half

of its maximum value, t1/2, determined. The difference in applied dopamine

concentration vs. that detected at the electrode (millimolar vs. micromolar) is attributed

to reuptake and diffusion of the analyte into the surrounding tissue and solution.

Importantly, the time course of the uptake monitored in the fly brain following

application of exogenous dopamine solution (t1/2 ~ 50 s) is consistent with measurements

of clearance from tissue in other model systems like the rat following exogenous

application of dopamine solution (39). Thus, this method is a valid approach to measure

changes in exogenously applied dopamine concentration occurring in vivo in the adult fly

brain.

Voltammetric vs. amperometric detection of dopamine in vivo. Oxidation of

dopamine produces a current which is dependent on the concentration of applied

dopamine and its diffusion, uptake, and metabolism as it traverses through tissue.

However, the local geometry and position of the micropipet injector also influence the

signal. Specifically, the relative distance of the micropipet to the electrode in the PAM

area (Figure 2.1B) affects the amplitude of the current measured. Because a single

micropipet is difficult to position the same distance from the electrode multiple times, a

pulled triple-barrel capillary was used to exogenously apply three different concentrations

of dopamine to the PAM area in series. The current response from 1.0 mM dopamine,

approximately 150 pmol (Appendix), applied to the PAM region was measured over

42

Page 59: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.3. Exogenously applied 1.0 mM dopamine detected in vivo in an adult wild-

type fly. (A) Successive voltammograms plotted as applied potential vs. time with false

color representation showing current. (B) Background-subtracted fast-scan cyclic

voltammogram of dopamine application (200 V/s, repetition frequency = 10 Hz). (C)

Changes in dopamine concentration over time. Black arrow corresponds to a 1.0 s

dopamine application beginning at 5.0 s. Dopamine concentration was determined by

converting the maximum current from the sampled amperometry plot using the in vitro

calibration average of three electrodes.

43

Page 60: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

time, and repeated with 2.0 mM and 5.0 mM dopamine solutions, with each solution

loaded into a separate barrel of the triple-barrel micropipet injector. Results obtained

using amperometry to measure the dopamine concentration in vivo proved to be variable.

Indeed, the measured concentration at the electrode does not increase linearly with the

applied concentration (r2 = 0.36, n = 4). Hence, FSCV was used for analysis.

Representative data collected using FSCV are shown in Figure 2.4. The measured peak

currents were converted to dopamine concentration by in vitro calibration of the electrode

using standard solutions (Appendix). The plot of normalized measured dopamine

concentration vs. injected dopamine concentration constructed using FSCV

measurements has a slope of 0.73 ± 0.08 (r2 = 0.84, n = 6), close to the expected value of

1. Thus, controlled concentrations of dopamine solutions can be applied locally to the fly

CNS and measured with voltammetry.

The differences observed between amperometry and FSCV are not surprising

when one takes into account the limited sample volume of the Drosophila PAM region.

During amperometric measurements, I hypothesize that local dopamine is “consumed” by

oxidization to the orthoquinone, and the local dopamine concentration is altered, making

the dopamine unavailable for repeated measurements. The orthoquinone might also be

involved in mechanisms of oxidative stress that could affect surrounding tissue in the

local environment. In contrast, voltammetric measurements regenerate the measured

analyte, minimizing the effect on surrounding tissue. Additionally, the diffusion layer,

and thus the volume sampled, with FSCV is smaller than that sampled using

amperometry (~3 pL vs. ~50 pL based on the parameters used in these experiments,

Appendix). Amperometry effectively measures dopamine changes that are averaged over

44

Page 61: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.4. Voltammetric detection of exogenously applied dopamine solutions in the

PAM area of an adult Drosophila brain. A triple-barrel micropipet was used to apply 1.0

mM (black line), 2.0 mM (red line), and 5.0 mM (blue line) dopamine solutions in series

for 1.0 s beginning at 5.0 s (black arrow). Dopamine concentration was determined by

converting the maximum current from the sampled amperometry plot using the in vitro

calibration average of three electrodes.

45

Page 62: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

a larger tissue volume, whereas FSCV measures the dopamine concentration locally

around the electrode. This apparently leads to a more accurate measurement of dopamine

concentration in this system.

Comparison of dopamine uptake in wild-type vs. fmn mutant flies. The fmn mutants

are a Drosophila line where dopamine transporter function has been eliminated through

genetic mutation. Thus, the cells that normally remove dopamine from the extracellular

fluid after it is released cannot do so, or at least not by the normal mechanism, in fmn

mutant flies. I used in vivo voltammetry to investigate the relative magnitude of uptake

of dopamine in the fly brain by comparing the fmn mutants to wild-type flies.

Using the same FSCV parameters described in a previous section, differences in

uptake between the wild-type and fmn mutant brains were first investigated. Dopamine

was exogenously applied to the PAM area (1.0 mM) with a single micropipet injector,

and the current response recorded (baseline measurement). Two baseline measurements

were taken, and the maximum currents averaged together and converted to dopamine

concentration for each fly. Interestingly, comparison of the black traces in Figure 2.5A

and 2.5B shows that the peak dopamine concentration observed after injection, [DA]max,

is considerably smaller in the wild-type fly compared to the fmn mutant fly. When the

average baselines for signals in multiple flies are considered (Figure 2.5C), the [DA]max

was significantly higher in fmn flies compared to wild-type flies (9.5 ± 2.4 µM vs. 3.1 ±

0.8 µM; Student’s t-test, p = 0.02, n = 9). This indicates that less dopamine is detected at

the electrode after exogenous application in the wild-type flies and is likely due to a high

rate of dopamine uptake via the functional dopamine transporter in the PAM neurons in

these flies vs. the nonfunctional dopamine transporter in the fmn flies. Therefore,

46

Page 63: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.5. Effect of cocaine on dopamine uptake. (A) Representative concentration

trace of exogenously applied 1.0 mM dopamine in wild-type Drosophila before (black

line) and after (red line) cocaine application. An increase in dopamine concentration in

the adult wild-type fly was observed following a 5 min exposure to 1.0 mM cocaine.

Black arrow corresponds to a 1.0 s dopamine application beginning at 5.0 s. (B)

Representative concentration trace of exogenously applied 1.0 mM dopamine in the fmn

mutant before (black line) and after (red line) cocaine application. No significant change

was observed in the fmn mutant fly. (C) Baseline comparison of [DA]max for wild-type

and fmn mutant flies (mean ± SEM; Student’s t-test, p = 0.02 (*), n = 9). (D)

Comparison of adult wild-type vs. fmn mutant flies when 1.0 mM dopamine is

exogenously applied after application of 1.0 mM cocaine. The increase in [DA]max is

significantly higher in wild-type flies compared to fmn flies when treated with cocaine

(mean ± SEM; Student’s t-test, p = 0.01 (*), n = 6).

47

Page 64: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

[DA]max can be used to measure changes in dopamine uptake. It is important to point out

that the measurements reported here are highly dependent on electrode and injector

placement, resulting in some variation in the values in different flies of the same

genotype. However, experiments comparing the relative amount of dopamine in different

flies can be carried out by normalization to baseline signals following initial dopamine

application, and temporal changes of uptake in the same fly with different conditions can

be carried out.

The validity of this theory is demonstrated by using a known dopamine uptake

inhibitor, cocaine, to block reuptake of exogenously applied dopamine. To account for

differences in the injector positioning and fly-to-fly variability, the maximum currents of

two baseline measurements were averaged for each fly, and all measurements for that

particular fly were normalized to it. After the baseline measurements, the fly brain was

bathed with 1.0 mM cocaine in AHL saline, and a voltammogram was obtained for

exogenously applied dopamine after five minutes. Representative traces for wild-type

and fmn mutant flies are shown in Figure 2.5A and 2.5B. After the cocaine application,

higher dopamine concentrations were detected at the electrode compared to baseline in

wild-type flies (Figure 2.5A). fmn mutants lacking functional dopamine transporters

showed no change from baseline following the cocaine incubation (Figure 2.5B). When

multiple cocaine-treated flies were considered (Figure 2.5D), the wild-type flies had

significantly increased normalized [DA]max compared to the cocaine-treated fmn mutant

flies (Student’s t-test, p = 0.01, n = 6). This data supports existing evidence that cocaine

blocks dopamine transporter function in Drosophila (24).

48

Page 65: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

The effect of tetrodotoxin (TTX) on dopamine uptake. The effect of neuronal

activities on dopamine uptake was investigated by treating the brains of the two fly

genotypes with TTX. TTX is a neurotoxin that blocks action potentials through the

blockade of voltage-sensitive sodium channels (40-42).

To examine the effects of TTX, the fly brain was bathed with 1.0 μM TTX in

AHL saline after the baseline dopamine measurements, and voltammograms were

obtained for injections of dopamine every five minutes. Representative traces for wild-

type and fmn mutant flies are shown in Figure 2.6A and 2.6B. The fmn mutant clearly

exhibited a different response than the wild-type flies following incubation with TTX.

After TTX treatments in wild-type flies, higher dopamine concentrations were detected at

the electrode compared to baseline (Figure 2.6A). This could be due to several factors.

For example, dopamine uptake in the fly brain may depend on neuronal activity in which

case inhibition of the action potential by TTX would abolish the uptake. Alternatively,

TTX might directly inhibit the uptake process. Both possibilities are supported by the

result that fmn mutants lacking functional dopamine transporters showed no significant

change from baseline following TTX incubation (Figure 2.6B).

Interestingly, the TTX-treated wild-type flies contained significantly increased

normalized [DA]max and t1/2 compared to the TTX-treated fmn mutant flies (Figure 2.6C;

two-way analysis of variance (ANOVA), p < 0.0001 for genotype for [DA]max, p = 0.04

for genotype for t1/2, n = 3). It is possible that the fmn mutant may have a compensatory

increase in the transporter-independent process (i.e., an increased N-methylation) for

inactivating endogenously released as well as exogenously applied dopamine, leading to

decreased dopamine concentrations detected at the electrode. Previous studies have

49

Page 66: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 2.6. Effect of TTX on dopamine uptake. (A) Representative concentration trace

of exogenously applied dopamine in wild-type Drosophila before and after 1.0 µM TTX

application. An increase in dopamine concentration in the adult wild-type fly was

observed following exposure to TTX. Black arrow corresponds to a 1.0 s dopamine

application beginning at 5.0 s. Baseline 2, 10 min, and 20 min traces were omitted for

clarity. (B) Representative concentration trace of exogenously applied dopamine in the

fmn mutant before and after TTX application. No significant change was observed in the

adult fmn mutant fly. (C) Comparison of adult wild-type vs. fmn mutant flies when 1.0

mM dopamine is exogenously applied before and after application of 1.0 μM TTX. The

increases in [DA]max are significantly higher in wild-type flies compared to fmn flies

when treated with 1.0 μM TTX (mean ± SEM; two-way ANOVA, p < 0.0001 (***) for

genotype, n = 3; SEMs for the baseline bars are too small to see).

50

Page 67: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

reported the activity of the dopamine transporter to be dependent on membrane potential

(43). TTX blocks voltage-gated sodium channels, thereby reducing the activity of

neurons via action potentials. Thus, the data suggest that the dopamine transporter is

activity-dependent, as uptake is reduced in the wild-type flies with TTX.

Conclusions

Microanalytical tools have been developed for in vivo electrochemical

measurements in the adult Drosophila CNS. Exogenously applied dopamine is detected

using a cylindrical carbon-fiber microelectrode inserted into the dopamine neuronal

cluster projecting to the mushroom bodies. The signal has been characterized using

FSCV. A known dopamine uptake blocker, cocaine, was used to validate this method for

in vivo measurement of Drosophila dopamine transporter function. Electrochemical

detection with FSCV was used to investigate the effect of TTX on the dopamine

transporter and the peak dopamine concentration measured which is dependent on uptake.

This work presents a new in vivo method for studying electroactive neurotransmitters in

Drosophila which can be used to measure changes in dopamine uptake.

51

Page 68: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Kissinger, P. T., Hart, J. B., and Adams, R. N. (1973) Voltammetry in brain

tissue-a new neurophysiological measurement, Brain Res. 55, 209-213.

2. Adams, R. N. (1976) Probing brain chemistry with electroanalytical techniques,

Anal. Chem. 48, 1126A-1138A.

3. Cheer, J. F., Heien, M. L., Garris, P. A., Carelli, R. M., and Wightman, R. M.

(2005) Simultaneous dopamine and single-unit recordings reveal accumbens

GABAergic responses: Implications for intracranial self-stimulation, Proc. Natl.

Acad. Sci. U.S.A. 102, 19150-19155.

4. Robinson, D. L., Hermans, A., Seipel, A. T., and Wightman, R. M. (2008)

Monitoring rapid chemical communication in the brain, Chem. Rev. 108, 2554-

2584.

5. Phillips, P. E., Stuber, G. D., Heien, M. L., Wightman, R. M., and Carelli, R. M.

(2003) Subsecond dopamine release promotes cocaine seeking, Nature 422, 614-

618.

6. Dayton, M. A., Brown, J. C., Stutts, K. J., and Wightman, R. M. (1980) Faradaic

electrochemistry at micro-voltammetric electrodes, Anal. Chem. 52, 946-950.

7. Millar, J., Armstrong-James, M., and Kruk, Z. L. (1981) Polarographic assay of

iontophoretically applied dopamine and low-noise unit recording using a

multibarrel carbon-fiber microelectrode, Brain Res. 205, 419-424.

8. Burmeister, J. J., Pomerleau, F., Huettl, P., Gash, C. R., Werner, C. E., Bruno, J.

P., and Gerhardt, G. A. (2008) Ceramic-based multisite microelectrode arrays for

simultaneous measures of choline and acetylcholine in CNS, Biosens.

Bioelectron. 23, 1382-1389.

9. Ewing, A. G., Bigelow, J. C., and Wightman, R. M. (1983) Direct in vivo

monitoring of dopamine released from 2 striatal compartments in the rat, Science

221, 169-171.

10. Robinson, D. L., Venton, B. J., Heien, M. L., and Wightman, R. M. (2003)

Detecting subsecond dopamine release with fast-scan cyclic voltammetry in vivo,

Clin. Chem. 49, 1763-1773.

11. Quintero, J. E., Day, B. K., Zhang, Z., Grondin, R., Stephens, M. L., Huettl, P.,

Pomerleau, F., Gash, D. M., and Gerhardt, G. A. (2007) Amperometric measures

of age-related changes in glutamate regulation in the cortex of rhesus monkeys,

Exp. Neurol. 208, 238-246.

12. Giros, B., Jaber, M., Jones, S. R., Wightman, R. M., and Caron, M. G. (1996)

Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking

the dopamine transporter, Nature 379, 606-612.

13. Jones, S. R., Gainetdinov, R. R., Jaber, M., Giros, B., Wightman, R. M., and

Caron, M. G. (1998) Profound neuronal plasticity in response to inactivation of

the dopamine transporter, Proc. Natl. Acad. Sci. U.S.A. 95, 4029-4034.

14. Benoit-Marand, M., Jaber, M., and Gonon, F. (2000) Release and elimination of

dopamine in vivo in mice lacking the dopamine transporter: functional

consequences, Eur. J. Neurosci. 12, 2985-2992.

15. Sokolowski, M. B. (2001) Drosophila: Genetics meets behavior, Nat. Rev. Genet.

2, 879-890.

52

Page 69: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

16. Waddell, S., and Quinn, W. G. (2001) What can we teach Drosophila? What can

they teach us?, Trends Genet. 17, 719-726.

17. Monastirioti, M. (1999) Biogenic amine systems in the fruit fly Drosophila

melanogaster, Microsc. Res. Tech. 45, 106-121.

18. Bainton, R. J., Tsai, L. T. Y., Singh, C. M., Moore, M. S., Neckameyer, W. S.,

and Heberlein, U. (2000) Dopamine modulates acute responses to cocaine,

nicotine, and ethanol in Drosophila, Curr. Biol. 10, 187-194.

19. Panda, S., Hogenesch, J. B., and Kay, S. A. (2002) Circadian rhythms from flies

to human, Nature 417, 329-335.

20. Scholz, H., Ramond, J., Singh, C. M., and Heberlein, U. (2000) Functional

ethanol tolerance in Drosophila, Neuron 28, 261-271.

21. Bergquist, J., Sciubisz, A., Kaczor, A., and Silberring, J. (2002) Catecholamines

and methods for their identification and quantitation in biological tissues and

fluids, J. Neurosci. Methods 113, 1-13.

22. Ritz, M. C., Lamb, R. J., Goldberg, S. R., and Kuhar, M. J. (1987) Cocaine

receptors on dopamine transporters are related to self-administration of cocaine,

Science 237, 1219-1223.

23. Porzgen, P., Park, S. K., Hirsh, J., Sonders, M. S., and Amara, S. G. (2001) The

antidepressant-sensitive dopamine transporter in Drosophila melanogaster: A

primordial carrier for catecholamines, Mol. Pharmacol. 59, 83-95.

24. Kuhar, M. J., Ritz, M. C., and Boja, J. W. (1991) The dopamine hypothesis of the

reinforcing properties of cocaine, Trends Neurosci. 14, 299-302.

25. Zahniser, N. R., Larson, G. A., and Gerhardt, G. A. (1999) In vivo dopamine

clearance rate in rat striatum: regulation by extracellular dopamine concentration

and dopamine transporter inhibitors, J. Pharmacol. Exp. Ther. 289, 266-277.

26. Wightman, R. M., Jankowski, J. A., Kennedy, R. T., Kawagoe, K. T., Schroeder,

T. J., Leszczyszyn, D. J., Near, J. A., Diliberto, E. J., and Viveros, O. H. (1991)

Temporally resolved catecholamine spikes correspond to single vesicle release

from individual chromaffin cells, Proc. Natl. Acad. Sci. U.S.A. 88, 10754-10758.

27. Stamford, J. A., Kruk, Z. L., and Millar, J. (1986) Sub-2nd striatal dopamine

release measured by in vivo voltammetry, Brain Res. 381, 351-355.

28. Clark, R. A., Zerby, S. E., and Ewing, A. G. (1996) Electrochemistry in neuronal

microenvironments, in Electroanalytical Chemistry: A Series of Advances (Bard,

A. J., and Rubinstein, I., Eds.), pp 227-295, Marcel Dekker, New York.

29. Baur, J. E., Kristensen, E. W., May, L. J., Wiedemann, D. J., and Wightman, R.

M. (1988) Fast-scan voltammetry of biogenic amines, Anal. Chem. 60, 1268-

1272.

30. Michael, D., Travis, E. R., and Wightman, R. M. (1998) Color images for fast-

scan CV, Anal. Chem. 70, 586A-592A.

31. Wang, J. W., Wong, A. M., Flores, J., Vosshall, L. B., and Axel, R. (2003) Two-

photon calcium imaging reveals an odor-evoked map of activity in the fly brain,

Cell 112, 271-282.

32. (2003) The FlyBase database of the Drosophila genome projects and community

literature, Nucleic Acids Res. 31, 172-175.

53

Page 70: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

33. Friggi-Grelin, F., Coulom, H., Meller, M., Gomez, D., Hirsh, J., and Birman, S.

(2003) Targeted gene expression in Drosophila dopaminergic cells using

regulatory sequences from tyrosine hydroxylase, J. Neurobiol. 54, 618-627.

34. Kume, K., Kume, S., Park, S. K., Hirsh, J., and Jackson, F. R. (2005) Dopamine is

a regulator of arousal in the fruit fly, J. Neurosci. 25, 7377-7384.

35. Heien, M. L., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman, R.

M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

36. Nassel, D. R., and Elekes, K. (1992) Aminergic neurons in the brain of blowflies

and Drosophila: dopamine- and tyrosine hydroxylase-immunoreactive neurons

and their relationship with putative histaminergic neurons, Cell Tissue Res. 267,

147-167.

37. Davis, R. L. (2005) Olfactory memory formation in Drosophila: from molecular

to systems neuroscience, Annu. Rev. Neurosci. 28, 275-302.

38. Kim, Y.-C., Lee, H.-G., and Han, K.-A. (2007) D1 dopamine receptor dDA1 is

required in the mushroom body neurons for aversive and appetitive learning in

Drosophila, J. Neurosci. 27, 7640-7647.

39. Sabeti, J., Adams, C. E., Burmeister, J., Gerhardt, G. A., and Zahniser, N. R.

(2002) Kinetic analysis of striatal clearance of exogenous dopamine recorded by

chronoamperometry in freely-moving rats, J. Neurosci. Methods 121, 41-52.

40. Narahashi, T., Moore, J. W., and Scott, W. R. (1964) Tetrodotoxin blockage of

sodium conductance increase in lobster giant axons, J. Gen. Physiol. 47, 965-974.

41. Takata, M., Moore, J. W., Kao, C. Y., and Fuhrman, F. A. (1966) Blockage of

sodium conductance increase in lobster giant axon by tarichatoxin (tetrodotoxin),

J. Gen. Physiol. 49, 977-988.

42. Moore, J. W., Blaustein, M. P., Anderson, N. C., and Narahashi, T. (1967) Basis

of tetrodotoxin's selectivity in blockage of squid axons, J. Gen. Physiol. 50, 1401-

1411.

43. Sonders, M. S., Zhu, S. J., Zahniser, N. R., Kavanaugh, M. P., and Amara, S. G.

(1997) Multiple ionic conductances of the human dopamine transporter: the

actions of dopamine and psychostimulants, J. Neurosci. 17, 960-974.

54

Page 71: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 3: Using In Vivo Electrochemistry to Study the

Physiological Effects of Cocaine and Other Stimulants on the

Drosophila melanogaster Dopamine Transporter*

*Reproduced with permission from Makos, M. A., Han, K.-A., Heien, M. L., and Ewing,

A. G. (2010) Using In Vivo Electrochemistry to Study the Physiological Effects of

Cocaine and Other Stimulants on the Drosophila melanogaster Dopamine Transporter,

ACS Chem. Neurosci. 1, 74-83. © 2010 American Chemical Society.

55

Page 72: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

The psychomotor stimulant drugs cocaine, amphetamine, and methylphenidate

bind to the dopamine transporter and alter its function, increasing extracellular dopamine

levels in the brain. The dopamine transporter is the plasma membrane protein primarily

responsible for clearing dopamine from the extracellular space, which leads to the

termination of dopamine neurotransmission (1, 2). Several lines of evidence have

demonstrated that increased extracellular dopamine levels are central to the reinforcing

and addictive properties exhibited by drugs of abuse (3, 4). It is well established that

cocaine blocks dopamine uptake via the dopamine transporter to elevate the extracellular

dopamine concentration (5, 6), and more recently it has been thought to affect the

serotonin and norepinephrine transporters as well (7, 8). Amphetamine has dual effects

on dopamine transport activity, both inhibiting dopamine uptake and inducing reverse

transport through the dopamine transporter (9-11). Methylphenidate, a commonly

prescribed medication for the treatment of attention deficit hyperactivity disorder (12),

blocks the dopamine transporter and increases the synaptic dopamine concentration (13,

14). While methylphenidate is abused by humans and has a similar affinity for the

dopamine transporter as cocaine (3, 6), abuse is not as widespread as that of cocaine. The

pharmacokinetics of the two drugs is thought to contribute to the difference observed in

their addictive properties (15). Neurochemicals in the central nervous system (CNS)

associated with addiction have been investigated for several decades; however, the

mechanisms underlying stimulant addictions and the behaviors they elicit are still not

fully understood.

56

Page 73: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

While animal model systems including rats, mice, and primates have been used

for several decades to study the effects of psychostimulants on dopamine transporter

function (7, 16, 17), recently, there is accumulating evidence for the validity of using

Drosophila melanogaster as a model system for neurotransmission (18, 19). In humans,

dopamine and serotonin play significant roles in regulating diverse physiological

processes including attention, motivation, and addiction, and these two monoamines have

been found to exert similar functions in the fly (20-23). When exposed to cocaine,

nicotine, or ethanol, Drosophila exhibits behavioral responses akin to those displayed by

mammals (24-28). In addition to the above mentioned monoamines, octopamine is a

major neurotransmitter in the CNS of invertebrates. Similar to norepinephrine in

mammals, octopamine dynamics in Drosophila are affected by exposure to cocaine (29).

While behavioral studies are a crucial aspect of investigating psychostimulant actions in

Drosophila, the ability to quantify neurochemicals in vivo would greatly improve

understanding of the molecular and cellular pathways behind the reinforcing and

addictive effects of a drug.

The electroactive nature of several neurotransmitters makes in vivo

electrochemistry an ideal approach for measuring chemical changes in the brain. Uptake

studies on both exogenously applied dopamine and stimulated dopamine release have

been characterized in vivo using voltammetry and chronoamperometry techniques in rats

(30-32). In particular, fast-scan cyclic voltammetry (FSCV) coupled with carbon-fiber

microelectrodes is a valuable method for quantification of biogenic amines in the CNS

because of its chemical selectivity and subsecond temporal resolution (33-35), and it has

been used previously in rats, mice, Drosophila flies, and Drosophila larvae (5, 36-38).

57

Page 74: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Here, I utilized recently developed microanalytical techniques introduced in Chapter 2 to

measure changes in the uptake of exogenously applied dopamine in the CNS of adult

Drosophila with treatments of cocaine, amphetamine, methamphetamine, or

methylphenidate. The physiological stimulant concentration necessary to significantly

block uptake by the dopamine transporter was approximately 10 µM.

Methods

Chemicals. All chemicals were used as received and purchased from Sigma (St. Louis,

MO) unless otherwise stated. Adult-hemolymph like (AHL) saline (108 mM NaCl, 5

mM KCl, 2 mM CaCl2, 8.2 mM MgCl2, 4 mM NaHCO3, 1 mM NaH2PO4, 5 mM

trehalose (Fluka BioChemika, Buchs, Switzerland), 10 mM sucrose, 20 mM Trizma

base , pH 7.5) was made using ultrapure (18 MΩ·cm) water and filtered through a 0.2-

μm filter (18). All collagenase, KCl, dopamine, N-acetyl-p-aminophenol (APAP,

acetaminophen), (+) cocaine, (+) amphetamine, (+) methamphetamine, and

methylphenidate solutions were prepared using AHL saline.

In vivo Drosophila preparation. The Canton-S strain of Drosophila melanogaster was

used for the wild-type fly in this chapter. The fumin (fmn) mutant has a genetic lesion

abolishing the dopamine transporter function. The genetic background of the w;fmn

mutant was replaced with the Canton-S background. All flies were maintained at 25 °C

on a standard cornmeal-agar medium, and 4 to 10 day-old male flies were used for

experiments. The flies were prepared for in vivo voltammetry as described in Chapter 2.

Briefly, flies were immobilized on ice and mounted in a homemade collar (38.1 mm

diameter concave plexiglass disk with a 1.0 mm hole in the center) with low melting

58

Page 75: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

agarose (Fisher Scientific, Pittsburgh, PA). Microsurgery was performed on a

stereoscope (Olympus SZ60, Melville, NY). After the cuticle was removed from the top

portion of the head to expose the brain, the head was covered with 0.1% collagenase

solution for 30 min to relax the extracellular matrix in the brain. The head of the

immobilized fly was then rinsed and bathed with AHL saline (“bath application method”)

with the preparation maintaining its viability for 1.5 - 2.5 h.

Electrochemical measurements. Carbon-fiber microelectrodes were fabricated as

described in Chapter 2 (38). Briefly, a single 5-μm diameter carbon fiber (Amoco,

Greenville, SC) was aspirated into a borosilicate glass capillary (B120-69-10, Sutter

Instruments, Novato, CA), and the capillary was pulled using a regular glass capillary

puller (P-97, Sutter Instruments). Electrical contact was made by back-filling the

capillary with silver composition (4922N DuPont, Delta Technologies Ltd., Stillwater,

MN) and inserting a tungsten wire. To form a cylindrical electrode, the carbon fiber was

cut to a length of 40-50 μm, as measured from the glass junction. Electrode tips were

dipped into epoxy (Epo-Tek, Epoxy Technology, Billerica, MA) for 30 s to ensure a good

seal between the fiber and the glass and then dipped into acetone for 15 s to remove

epoxy from the exposed carbon fiber. Standard dopamine solutions were used for in vitro

electrode calibration (Appendix). A Ag/AgCl reference electrode was made by

chlorodizing a silver wire (0.25 mm diameter, 99.999% purity, Alfa Aesar, Ward Hill,

MA) in bleach overnight. All electrodes were positioned using micromanipulators (421

series, Newport, Irvine, CA). Micropipet injectors were fabricated by pulling glass

capillaries in a glass capillary puller to an opening of approximately 5 μm. Micropipet

59

Page 76: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

injectors were coupled to a microinjection system (Picospritzer II, General Valve

Corporation, Fairfield, NJ) and used to exogenously apply dopamine solutions.

Electrochemical data were collected using an Axopatch 200B Amplifier (Axon

Instruments, Foster City, CA) or a Dagan Chem-Clamp potentiostat (Dagan Corporation,

Minneapolis, MN) and two data acquisition boards (PCI-6221, National Instruments,

Austin, TX) run by the TH 1.0 CV program (ESA, Chelmsford, MA) (36). All cyclic

voltammograms were obtained using a triangular waveform (scanned -0.6 V to +1.0 V vs.

Ag/AgCl at 200 V/s) repeated every 100 ms (low pass Bessel filter at 3-5 kHz). Prior to

voltammetric experiments, all electrodes were cycled (-0.6 V to +1.0 V at 200 V/s) for at

least 15 min to stabilize the background current. Electrochemical responses were plotted

and statistical analysis performed using Prism 5.0 (GraphPad Software, La Jolla, CA).

Results and Discussion

The effect of 1.0 mM cocaine treatment on dopamine uptake. Microanalytical

techniques developed for in vivo electrochemical detection in Drosophila provide a

method for studying the physiological effects of drug treatments on redox-active

neurotransmitters. In Chapter 2, I characterized exogenously applied dopamine uptake

using electrochemical detection with a carbon-fiber microelectrode inserted into the

protocerebral anterior medial (PAM) area of an adult Drosophila brain (38). In this

chapter, I utilize this procedure to explore dopamine neurotransmission in the Drosophila

CNS. Dopamine neuronal cell bodies are clustered together in several distinct areas

throughout the Drosophila brain with the largest neuronal cluster located in the PAM

region projecting to the nearby mushroom body (39-41), a key brain structure for learning

60

Page 77: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

and memory (42). Octopamine levels in this particular brain region are insignificant,

simplifying measurements of dopamine. Thus, my in vivo investigation of dopamine

uptake in Drosophila is focused on the PAM area.

Following microsurgery, a micromanipulator was used to insert the cylindrical

working electrode into the PAM region while the reference electrode was submerged in

the AHL saline bath covering the exposed fly brain. Small amounts of dopamine were

ejected just above the PAM area, approximately 10 μm from the working electrode, with

a single micropipet injector. FSCV was used to monitor changing dopamine levels in the

CNS of both wild-type and fmn mutant flies over time. Voltammetry was performed by

applying potential in a triangular waveform to the electrode while the current response

was recorded. To visualize changes over time, a false-color representation of current is

used (Figure 3.1A) where the green corresponds to the oxidation of dopamine, and the

reduction of the orthoquinone is represented in blue (33). The current response was

converted to dopamine concentration using in vitro electrode calibration (Appendix).

The peak dopamine concentration measured is referred to as [DA]max which is an

established parameter for measuring changes in uptake of extracellular dopamine (17). In

addition to [DA]max, another parameter used to compare dopamine clearance between the

two fly genotypes is t1/2, the full width of time at half maximum of the dopamine

concentration (Figure 3.1B).

The validity of using [DA]max to compare changes in dopamine uptake via the

functional dopamine transporter in wild-type flies vs. the nonfunctional dopamine

61

Page 78: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.1. In vivo detection of exogenously applied 1.0 mM dopamine in the adult

Drosophila brain. (A) Applied potential vs. time gives a visual representation of

successive voltammograms with current viewed in false color. (B) Dopamine

concentration plotted over time. Dopamine concentration was determined from the

measured current using an in vitro calibration average of three electrodes. The black

arrow corresponds to a 1.0 s dopamine application beginning at 5.0 s.

62

Page 79: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

transporter in fmn flies has been demonstrated (38). Here, [DA]max was used to

investigate the effectiveness of a known dopamine uptake inhibitor, cocaine, on blocking

uptake by the Drosophila dopamine transporter in vivo. A 1.0 mM dopamine solution

was exogenously applied to the PAM area for 1.0 s (corresponding to ~150 pmol

dopamine ejected, Appendix), and the current response was recorded for 3 min (Figure

3.2A, B: “baseline 1”). Following three baseline measurements, the fly brain was bathed

with 1.0 mM cocaine in AHL saline for 5 min and then the current response was recorded

over time following dopamine injection (“5 min cocaine”). Cocaine treatment was

continued and dopamine injections were repeated every 5 min while the current response

was recorded.

The representative cyclic voltammogram in Figure 3.2C is a background-

subtracted average of ten successive cyclic voltammograms acquired during an in vivo

dopamine baseline measurement from an adult wild-type fly brain (dashed red line). A

background-subtracted average of ten successive cyclic voltammograms of exogenously

applied dopamine following 15 min of 1.0 mM cocaine treatment is plotted for

comparison (solid black line). Both voltammograms are from the time period when

[DA]max was measured, and by inspection, the voltammetric peaks correspond to the

electrochemical signature of dopamine (35, 43). After a 1.0 mM cocaine treatment, a 3-

fold increase in [DA]max was observed for the adult wild-type fly (Figure 3.2A) while the

[DA]max of the fmn mutant fly (Figure 3.2B) remained unchanged. Notably, comparison

of the baseline measurements in Figure 3.2A, B shows a significant difference between

the two fly types following exogenous dopamine application. Less dopamine is detected

in the wild-type fly vs. the fmn fly, which is likely due to dopamine uptake by the

63

Page 80: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.2. Effect of 1.0 mM cocaine treatment on uptake of an exogenously applied 1.0

mM dopamine solution. (A) Representative concentration trace in the wild-type fly

before (baseline 1, 2) and after cocaine treatment. A significant increase in dopamine

concentration was observed. (B) Representative concentration trace in the fmn mutant fly

before (baseline 1, 2) and after cocaine treatment. Dopamine concentration was

determined by converting the measured current using in vitro electrode calibration. The

black arrow corresponds to a 1.0 s dopamine application beginning at 5.0 s. (C)

Background-subtracted fast-scan cyclic voltammogram of baseline extracellular

dopamine (dashed red line) and extracellular dopamine after 15 min of cocaine treatment

(solid black line) in a wild-type fly (200 V/s, average of 10 scans each).

64

Page 81: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

functional transporter that is present only in the wild-type fly. When the average [DA]max

of multiple flies was considered (Figure 3.3), the [DA]max of untreated wild-type flies

(“baseline 1, 2”) was significantly lower than for fmn mutant flies. Interestingly, the

[DA]max for cocaine treated wild-type flies was not significantly different from the

untreated fmn mutant flies (“baseline 1, 2”) which supports existing evidence that cocaine

blocks the dopamine transporter in Drosophila (44). Upon comparison of the two

genotypes, wild-type flies exhibited a significantly increased normalized [DA]max with

1.0 mM cocaine treatment compared to fmn mutant flies under the same treatment (Figure

3.4; two-way analysis of variance (ANOVA), p < 0.0001 for genotype, p = 0.0008 for

time, p = 0.0002 for interaction, n = 6). To account for slight differences in dopamine

injector positioning between flies, the [DA]max from two of the dopamine baseline

measurements for a fly were averaged together, and all measurements for that fly were

calculated as a percent of the average baseline measurement (i.e., [DA]max normalized)

(38, 45, 46). The maximum effect of the cocaine treatment on the wild-type flies was

observed within 10 min and remained fairly constant for over 20 min of cocaine

treatment while neither genotype experienced a significant change in t1/2. These

observations indicate that cocaine effectively blocks the Drosophila dopamine transporter

function in vivo.

Determining the physiological cocaine concentration in the Drosophila brain. To

estimate the concentration of the 1.0 mM cocaine solution in the PAM area, APAP was

used to mimic the bath application method of the cocaine treatment. APAP was selected

because it is an electroactive molecule that is thought to undergo neither rapid

metabolism nor uptake by monoamine transporters, thus allowing only the oxidation

65

Page 82: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.3. Investigating dopamine transporter function. Comparison of baseline

[DA]max for untreated (no cocaine) wild-type and fmn mutant flies (mean ± SEM;

Student’s t-test, p = 0.02 (*), n = 9) and wild-type flies after 15 min of 1.0 mM cocaine

treatment. The difference in [DA]max between untreated fmn mutants and wild-type flies

treated with cocaine is not significantly different (mean ± SEM; Student’s t-test, p = 0.3,

n = 6-9).

66

Page 83: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.4. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before and after 1.0 mM cocaine treatment. There is a significant

increase in normalized [DA]max for wild-type flies vs. fmn flies with cocaine treatment

(mean ± SEM; two-way ANOVA, p < 0.0001 (***) for genotype, p = 0.0008 (***) for

time, p = 0.0002 (***) for interaction, n = 6). The black arrow corresponds to the

beginning of the cocaine treatment.

67

Page 84: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

current from diffusion of the 1.0 mM bath solution into the brain region to be measured

(47). Furthermore, detection of APAP using voltammetry is well documented (48, 49).

To determine the physiological drug concentration in the Drosophila brain region from a

1.0 mM bath application over the experimental time period, a carbon-fiber

microelectrode was placed in the PAM region of Drosophila, and the fly head was bathed

in 1.0 mM APAP in AHL saline solution. Background-subtracted FSCV was employed

to measure the current in vivo from oxidation of APAP at the surface of the implanted

electrode (Figure 3.5A). The peak oxidation current was converted to APAP

concentration, [APAP], using in vitro electrode calibration with APAP (Figure 3.5B).

The actual [APAP] in the Drosophila brain, or the physiological [APAP], is

approximately 2 orders of magnitude lower (12 ± 5 µM, n = 3 flies) than the applied 1.0

mM bath [APAP]. While the concentration that diffuses into the tissue might differ

slightly between cocaine and APAP due to the distinct properties of the two species, such

as diffusion rate, relative permeability into the tissue, and size (Figure 3.5C, D) this

difference is insignificant compared to the high resistance to diffusion of the brain tissue.

When these calculations are applied to the cocaine solutions, a 1.0 mM cocaine

bath application corresponds to approximately a 12 µM or 0.004 mg/mL cocaine

concentration in the PAM area. This is significantly lower than that used in a study by

Hirsh and colleagues where 0.5 mg/mL cocaine was applied directly to Drosophila nerve

cords (20). Interestingly, my physiological cocaine concentration is consistent with a

recent report by Venton and coworkers that found 10 µM cocaine was sufficient to

effectively block serotonin reuptake by serotonin transporters located in the ventral nerve

cords of Drosophila larvae (37).

68

Page 85: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.5. Determining the physiological APAP concentration in the Drosophila CNS

from a 1.0 mM APAP bath application. (A) Background-subtracted fast-scan cyclic

voltammogram of APAP measured in vivo of an adult wild-type fly with a bath

application of 1.0 mM APAP (average of 10 successive scans). (B) Electrode calibration

plot in standard APAP solutions (mean ± SEM; n = 5). The physiological APAP

concentration in the Drosophila CNS is approximately 2 orders of magnitude lower than

the concentration of the applied bath solution. (C) Structure of APAP. (D) Structure of

cocaine.

69

Page 86: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Drosophila dopamine transporter inhibition as a function of cocaine concentration.

Electrochemical detection with FSCV was used to investigate the effect of three different

concentrations of cocaine (0.05, 0.5, or 1.0 mM) on dopamine uptake by the Drosophila

dopamine transporter. The fly was prepared for in vivo electrochemical measurements

and bathed with 0.05 mM cocaine in AHL saline after the baseline dopamine

measurements were acquired. Voltammograms of 1.0 mM dopamine injections were

obtained every 5 min.

Figure 3.6 is a comparison of the normalized [DA]max for wild-type vs. fmn

mutant flies after separate treatments for 10 min with either 0.05, 0.5, or 1.0 mM cocaine.

A two-way ANOVA was used to analyze the comprehensive data at all doses and a

significant difference in normalized [DA]max was observed for the two fly types (two-way

ANOVA, p < 0.0001 for genotype, concentration, and interaction, n = 6 for each

concentration and genotype). In addition, wild-type flies incubated with 1.0 mM cocaine

had significantly increased normalized [DA]max compared to control measurements of

AHL saline only (one-way ANOVA, p < 0.0001; post hoc Tukey pairwise comparisons,

p < 0.0001, n = 6). Higher dopamine concentrations were detected in wild-type flies

treated with 0.5 mM cocaine as well; however, the effect was not as robust as that

observed with the 1.0 mM cocaine treatment ([DA]max increased ~20% vs. ~125%

compared to AHL treatments). When the applied cocaine concentration was further

decreased to 0.05 mM, there was no significant difference in the normalized [DA]max for

wild-type flies from AHL saline measurements. Neither fly genotype exhibited a

significant change in [DA]max from baseline dopamine measurements when only AHL

saline (no cocaine) was applied in a control experiment (one-way ANOVA, p > 0.05 for

70

Page 87: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.6. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before (baseline) and after 10 min of various concentrations of

cocaine treatments. One of the following treatments was applied: AHL saline only, 0.05

mM, 0.5 mM, or 1.0 mM cocaine solution (mean ± SEM; two-way ANOVA, p < 0.0001

(***) for genotype, concentration, and interaction, n = 6; SEMs for the baseline bars are

too small to see). The bath solutions for the baseline and AHL saline treatment were

identical. The AHL saline treatment was included as a control to ensure the [DA]max

response did not increase from a temporal effect owing to the AHL solution. There is a

significant increase in normalized [DA]max for wild-type flies after cocaine treatments

compared to AHL saline (no cocaine) treatment (one-way ANOVA, p < 0.0001; post hoc

Tukey pairwise comparisons, p < 0.0001 (***) for the 1.0 mM cocaine treatment, n = 6).

No significant change was observed in the fmn mutant flies between AHL saline (no

cocaine) treatment and the three cocaine treatments (one-way ANOVA, p = 0.9, n = 6).

71

Page 88: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

both fly genotypes, n = 6). Only the baseline measurements from these AHL saline

control experiments are plotted for simplicity. There was no significant difference

between baseline measurements for wild-type flies that were later treated with cocaine vs.

baseline measurements for wild-type flies just treated with AHL saline (one-way

ANOVA, p = 0.8, n = 6). Similarly, there was no significant difference between baseline

measurements for fmn flies that were later treated with cocaine vs. baseline

measurements for fmn flies just treated with AHL saline (one-way ANOVA, p = 0.09, n =

6). The fmn mutant flies lacking the dopamine transporter exhibited no change in

extracellular dopamine concentration after 0.05, 0.5, or 1.0 mM cocaine treatment (one-

way ANOVA, p = 0.9, n = 6).

Therefore, at the 1.0 mM concentration, cocaine appears to overcome a threshold

concentration and significantly blocks the Drosophila dopamine transporter in vivo.

These data are consistent with the effect of cocaine on mammalian dopamine transporter

function (5, 8) and with observations previously made with this technique (38). These

findings support the use of Drosophila as a model system for studying pharmacological

effects in vivo. Although the effect of volatilized cocaine on Drosophila behavior has

previously been demonstrated (20), the findings presented here provide the first in vivo

investigation of the effective cocaine concentration needed to block uptake of

exogenously applied dopamine by the dopamine transporter in the adult fly.

The effect of other stimulant treatments on dopamine uptake. In addition to cocaine,

the effects of three other stimulants on Drosophila dopamine transporter function were

investigated. Flies were prepared as for cocaine experiments and then treated with either

1.0 mM amphetamine, methamphetamine, or methylphenidate in AHL saline. Figure 3.7

72

Page 89: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

contains a summary of the normalized [DA]max for adult wild-type flies compared to fmn

mutant flies following each of these drug treatments. When dopamine levels in the flies

treated with amphetamine were examined over time, there was a small, but significant

difference in the amount of dopamine detected in the PAM region of the wild-type brain

compared to the same region in the fmn mutant (Figure 3.7A; two-way ANOVA, p =

0.005 for genotype, n = 5). However, even after 30 min of treatment, the observed

[DA]max for the amphetamine-treated wild-type flies was lower than that of wild-type

flies treated with 1.0 mM cocaine (~25% increase vs. ~125% increase). This finding is

consistent with in vitro inhibition studies demonstrating amphetamine is a less potent

inhibitor of the Drosophila dopamine transporter than cocaine (44).

The Drosophila dopamine transporter was significantly affected by treatment with

1.0 mM methamphetamine as well (Figure 3.7B; two-way ANOVA, p = 0.01 for

genotype, n = 5-6). Methamphetamine-treated wild-type flies exhibited a similar increase

in [DA]max compared to the amphetamine-treated wild-type flies (~30% increase vs.

~25% increase). Interestingly, the trend in time until maximum blocking of dopamine

uptake occurs is later with methamphetamine treatment than with amphetamine or

cocaine treatment. Although the difference between the normalized [DA]max after 5 min

and 20 min of methamphetamine treatment in wild-type flies is not significantly different

(Student’s t-test, p = 0.4, n = 6), the kinetics of the action of methamphetamine on the fly

dopamine transporter could be of interest in future investigations. There is in vitro

evidence that methamphetamine and amphetamine cause internalization of the

mammalian dopamine transporter. These data suggest an additional mechanism that

contributes to the decrease in transporter activity by amphetamines, in addition to

73

Page 90: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 3.7. Comparison of wild-type and fmn mutant flies when 1.0 mM dopamine was

exogenously applied before and after 1.0 mM stimulant treatment. (A) Following

amphetamine treatment, the increases in normalized [DA]max are significantly higher in

wild-type flies compared to fmn mutant flies (mean ± SEM; two-way ANOVA, p = 0.005

(**) for genotype, n = 5). Additionally, the 30 min treatment is significantly different

from baseline 2 for the wild-type flies (one-way ANOVA, p = 0.03 (*); post hoc Tukey

pairwise comparisons, p < 0.05 (*)). (B) The increases in normalized [DA]max are

significantly higher in wild-type vs. fmn flies following methamphetamine treatment

(mean ± SEM; two-way ANOVA, p = 0.01 (*) for genotype, n = 5-6). (C) Following

methylphenidate treatment, the increases in normalized [DA]max for wild-type compared

with fmn flies are significantly higher (mean ± SEM; two-way ANOVA; p = 0.03 (*) for

interaction; p < 0.0001 (***) for genotype, n = 5).

74

Page 91: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

blocking and inducing reverse transport of dopamine through the dopamine transporter

(50-52). While in vitro model systems are often used to predict the effects of

psychostimulants on monoamine uptake, in vitro results are not always an accurate

reflection of the potential of a compound to modulate in vivo function (53-55). Thus,

development of analytical methods capable of in vivo evaluation of drug efficacy plays a

critical role in the neuroscience field. These in vivo measurements confirm

amphetamines do indeed alter Drosophila dopamine transporter function; however, with

the current experimental set-up, it is not possible to speculate on the exact mechanisms of

action occurring in the fly CNS.

Although methylphenidate is commonly studied in mammalian systems, very

little, if any, literature is available on the efficacy of this drug in Drosophila. Because the

fruit fly is becoming a more widely used model system for studying the neurochemical

basis for human behaviors and addictions (19), I chose to examine the effect of this

commonly prescribed drug on dopamine uptake using our in vivo method. Following

methylphenidate treatment, wild-type flies displayed a significantly higher extracellular

dopamine concentration compared to baseline dopamine measurements and the treated

fmn mutant flies (Figure 3.7C; two way ANOVA, p = 0.03 for interaction, p < 0.0001 for

genotype, n = 5). This indicates that methylphenidate blocks dopamine uptake occurring

via the Drosophila dopamine transporter. This finding correlates with the proposed

mechanism of methylphenidate in the human brain (13, 14) and supports the use of

Drosophila in future studies on methylphenidate. Of the four stimulants investigated,

cocaine and methylphenidate displayed the greatest effect on Drosophila dopamine

transporter function in vivo (Table 3.1).

75

Page 92: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Table 3.1. Change in [DA]max for four drugs of abusea.

cocaine amphetamine methamphetamine methylphenidate

IC50 (µM)

Drosophila

dopamine

transporter

6.0b

2.7c

4.9b

(+) 6.6; (-) 34.0c

4.5b 6.8

b

wild type:

[DA]max

normalized (20

min treatment)

223 ± 40 % 117 ± 8 % 129 ± 22 % 174 ± 31 %

fmn mutant:

[DA]max

normalized (20

min treatment)

91 ± 8 % 102 ± 8 % 99 ± 4 % 102 ± 11 %

aMaximum changes for dopamine ([DA]max) values are for (+) amphetamine and (+)

methamphetamine. [DA]max values are mean ± SEM for 1.0 mM drug concentrations (n

= 5-6). Literature IC50 values are included for comparison. bReference (56).

cReference

(44).

76

Page 93: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

In these experiments, exogenously applied dopamine is cleared primarily through

diffusion, metabolism, and uptake by the dopamine transporter. By comparing two fly

genotypes whose diffusion and metabolism are presumably similar since they only differ

in dopamine transporter function, I was able to investigate the uptake component of

dopamine clearance in the presence of various stimulants. All four stimulants tested

caused significantly increased dopamine signal amplitudes ([DA]max), which has been

observed in the cocaine-treated rat CNS where chronoamperometry was employed to

measure exogenously applied dopamine concentrations in vivo (17, 45). In these studies,

Gerhardt and coworkers also reported an increase in the time course of the enhanced

dopamine signal amplitudes, which was not observed in Drosophila. I speculate that

diffusion plays a prominent role in the clearance of dopamine from the Drosophila CNS

due to its reduced size (~ 5 nL) in comparison to the rat CNS which might experience

less diffusion of dopamine away from the electrode (32). A change in t1/2 could be too

minor to observe in the fly system relative to this diffusion factor.

Conclusions

This chapter presents in vivo measurements of dopamine uptake using

exogenously applied dopamine as a function of cocaine concentration in Drosophila. In

addition, physiological effects of amphetamine, methamphetamine, and methylphenidate

are also reported for the adult fly. Cocaine and methylphenidate were found to be more

potent at inhibiting dopamine uptake in vivo by the Drosophila dopamine transporter than

amphetamine and methamphetamine. It is most likely that the variation in the dose-

response results among the four stimulants tested reflects different interactions of the

77

Page 94: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

drugs with the dopamine transporter. Little is known about the in vivo nature of drug

interactions with invertebrate transporters, mainly because of the lack of tools heretofore

available for quantifying neurotransmitters in such small native environments. These

data support continued use of this in vivo Drosophila model system to further investigate

dopamine neurotransmission and enhance understanding of the physiological

mechanisms that underlie human behaviors and addictions.

78

Page 95: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Chen, N., and Reith, M. E. A. (2000) Structure and function of the dopamine

transporter, Eur. J. Pharmacol. 405, 329-339.

2. Mortensen, O. V., and Amara, S. G. (2003) Dynamic regulation of the dopamine

transporter, Eur. J. Pharmacol. 479, 159-170.

3. Di Chiara, G., and Imperato, A. (1988) Drugs abused by humans preferentially

increase synaptic dopamine concentrations in the mesolimbic system of freely

moving rats, Proc. Natl. Acad. Sci. U.S.A. 85, 5274-5278.

4. Kuhar, M. J., Ritz, M. C., and Boja, J. W. (1991) The dopamine hypothesis of the

reinforcing properties of cocaine, Trends Neurosci. 14, 299-302.

5. Giros, B., Jaber, M., Jones, S. R., Wightman, R. M., and Caron, M. G. (1996)

Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking

the dopamine transporter, Nature 379, 606-612.

6. Ritz, M. C., Lamb, R. J., Goldberg, S. R., and Kuhar, M. J. (1987) Cocaine

receptors on dopamine transporters are related to self-administration of cocaine,

Science 237, 1219-1223.

7. Carboni, E., Spielewoy, C., Vacca, C., Nosten-Bertrand, M., Giros, B., and Di

Chiara, G. (2001) Cocaine and amphetamine increase extracellular dopamine in

the nucleus accumbens of mice lacking the dopamine transporter gene, J.

Neurosci. 21, RC141.

8. Sora, I., Hall, F. S., Andrews, A. M., Itokawa, M., Li, X.-F., Wei, H.-B.,

Wichems, C., Lesch, K.-P., Murphy, D. L., and Uhl, G. R. (2001) Molecular

mechanisms of cocaine reward: Combined dopamine and serotonin transporter

knockouts eliminate cocaine place preference, Proc. Natl. Acad. Sci. U.S.A. 98,

5300-5305.

9. Kahlig, K. M., Binda, F., Khoshbouei, H., Blakely, R. D., McMahon, D. G.,

Javitch, J. A., and Galli, A. (2005) Amphetamine induces dopamine efflux

through a dopamine transporter channel, Proc. Natl. Acad. Sci. U.S.A. 102, 3495-

3500.

10. Sulzer, D., Chen, T. K., Lau, Y. Y., Kristensen, H., Rayport, S., and Ewing, A. G.

(1995) Amphetamine redistributes dopamine from synaptic vesicles to the cytosol

and promotes reverse transport, J. Neurosci. 15, 4102-4108.

11. Jones, S. R., Gainetdinov, R. R., Wightman, R. M., and Caron, M. G. (1998)

Mechanisms of amphetamine action revealed in mice lacking the dopamine

transporter, J. Neurosci. 18, 1979-1986.

12. Carrey, N. J., Wiggins, D. M., and Milin, R. P. (1996) Pharmacological treatment

of psychiatric disorders in children and adolescents - Focus on guidelines for the

primary care practitioner, Drugs 51, 750-759.

13. Volkow, N. D., Wang, G.-J., Fowler, J. S., Gatley, S. J., Logan, J., Ding, Y.-S.,

Hitzemann, R., and Pappas, N. (1998) Dopamine transporter occupancies in the

human brain induced by therapeutic doses of oral methylphenidate, Am. J.

Psychiatry 155, 1325-1331.

14. Volkow, N. D., Wang, G.-J., Fowler, J. S., Logan, J., Gerasimov, M., Maynard,

L., Ding, Y.-S., Gatley, S. J., Gifford, A., and Franceschi, D. (2001) Therapeutic

79

Page 96: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

doses of oral methylphenidate significantly increase extracellular dopamine in the

human brain, J. Neurosci. 21, RC121.

15. Volkow, N. D., Ding, Y.-S., Fowler, J. S., Wang, G.-J., Logan, J., Gatley, J. S.,

Dewey, S., Ashby, C., Liebermann, J., Hitzemann, R., and Wolf, A. P. (1995) Is

methylphenidate like cocaine? Studies on their pharmacokinetics and distribution

in the human brain, Arch. Gen. Psychiatry 52, 456-463.

16. Balster, R. L., and Schuster, C. R. (1973) Fixed-interval schedule of cocaine

reinforcement - effect of dose and infusion duration, J. Exp. Anal. Behav. 20, 119-

129.

17. Zahniser, N. R., Larson, G. A., and Gerhardt, G. A. (1999) In vivo dopamine

clearance rate in rat striatum: Regulation by extracellular dopamine concentration

and dopamine transporter inhibitors, J. Pharmacol. Exp. Ther. 289, 266-277.

18. Wang, J. W., Wong, A. M., Flores, J., Vosshall, L. B., and Axel, R. (2003) Two-

photon calcium imaging reveals an odor-evoked map of activity in the fly brain,

Cell 112, 271-282.

19. Makos, M. A., Kuklinski, N. J., Heien, M. L., Ewing, A. G., and Berglund, E. C.

(2009) Chemical measurements in Drosophila, TrAC, Trends Anal. Chem. 28,

1223-1234.

20. Li, H., Chaney, S., Forte, M., and Hirsh, J. (2000) Ectopic G-protein expression in

dopamine and serotonin neurons blocks cocaine sensitization in Drosophila

melanogaster, Curr. Biol. 10, 211-214.

21. Schultz, W. (2000) Multiple reward signals in the brain, Nat. Rev. Neurosci. 1,

199-207.

22. Waddell, S., and Quinn, W. G. (2001) What can we teach Drosophila? What can

they teach us?, Trends Genet. 17, 719-726.

23. Kim, Y.-C., Lee, H.-G., and Han, K.-A. (2007) Classical reward conditioning in

Drosophila melanogaster, Genes Brain Behav. 6, 201-207.

24. Bainton, R. J., Tsai, L. T. Y., Singh, C. M., Moore, M. S., Neckameyer, W. S.,

and Heberlein, U. (2000) Dopamine modulates acute responses to cocaine,

nicotine and ethanol in Drosophila, Curr. Biol. 10, 187-194.

25. McClung, C., and Hirsh, J. (1998) Stereotypic behavioral responses to free-base

cocaine and the development of behavioral sensitization in Drosophila, Curr.

Biol. 8, 109-112.

26. Rothenfluh, A., and Heberlein, U. (2002) Drugs, flies, and videotape: the effects

of ethanol and cocaine on Drosophila locomotion, Curr. Opin. Neurobiol. 12,

639-645.

27. Torres, G., and Horowitz, J. M. (1998) Activating properties of cocaine and

cocaethylene in a behavioral preparation of Drosophila melanogaster, Synapse

29, 148-161.

28. Chang, H. Y., Grygoruk, A., Brooks, E. S., Ackerson, L. C., Maidment, N. T.,

Bainton, R. J., and Krantz, D. E. (2006) Overexpression of the Drosophila

vesicular monoamine transporter increases motor activity and courtship but

decreases the behavioral response to cocaine, Mol. Psychiatry 11, 99-113.

29. Nathanson, J. A., Hunnicutt, E. J., Kantham, L., and Scavone, C. (1993) Cocaine

as a naturally occurring insecticide, Proc. Natl. Acad. Sci. U.S.A. 90, 9645-9648.

80

Page 97: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

30. Ewing, A. G., Bigelow, J. C., and Wightman, R. M. (1983) Direct in vivo

monitoring of dopamine released from 2 striatal compartments in the rat, Science

221, 169-171.

31. Ewing, A. G., and Wightman, R. M. (1984) Monitoring the stimulated release of

dopamine with in vivo voltammetry. II: Clearance of released dopamine from

extracellular fluid, J. Neurochem. 43, 570-577.

32. Sabeti, J., Adams, C. E., Burmeister, J., Gerhardt, G. A., and Zahniser, N. R.

(2002) Kinetic analysis of striatal clearance of exogenous dopamine recorded by

chronoamperometry in freely-moving rats, J. Neurosci. Methods 121, 41-52.

33. Michael, D., Travis, E. R., and Wightman, R. M. (1998) Color images for fast-

scan CV, Anal. Chem. 70, 586A-592A.

34. Baur, J. E., Kristensen, E. W., May, L. J., Wiedemann, D. J., and Wightman, R.

M. (1988) Fast-scan voltammetry of biogenic amines, Anal. Chem. 60, 1268-

1272.

35. Clark, R. A., Zerby, S. E., and Ewing, A. G. (1996) Electrochemistry in Neuronal

Microenvironments, in Electroanalytical Chemistry: A Series of Advances (Bard,

A. J., and Rubinstein, I., Eds.), pp 227-295, Marcel Dekker, New York.

36. Heien, M. L., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman, R.

M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

37. Borue, X., Cooper, S., Hirsh, J., Condron, B., and Venton, B. J. (2009)

Quantitative evaluation of serotonin release and clearance in Drosophila, J.

Neurosci. Methods 179, 300-308.

38. Makos, M. A., Kim, Y.-C., Han, K.-A., Heien, M. L., and Ewing, A. G. (2009) In

vivo electrochemical measurements of exogenously applied dopamine in

Drosophila melanogaster, Anal. Chem. 81, 1848-1854.

39. Coulom, H., and Birman, S. (2004) Chronic exposure to rotenone models sporadic

Parkinson's disease in Drosophila melanogaster, J. Neurosci. 24, 10993-10998.

40. Zhang, K., Guo, J. Z., Peng, Y., Xi, W., and Guo, A. (2007) Dopamine-mushroom

body circuit regulates saliency-based decision-making in Drosophila, Science

316, 1901-1904.

41. Nassel, D. R., and Elekes, K. (1992) Aminergic neurons in the brain of blowflies

and Drosophila: dopamine- and tyrosine hydroxylase-immunoreactive neurons

and their relationship with putative histaminergic neurons, Cell Tissue Res. 267,

147-167.

42. Kim, Y.-C., Lee, H.-G., and Han, K.-A. (2007) D1 Dopamine receptor dDA1 is

required in the mushroom body neurons for aversive and appetitive learning in

Drosophila, J. Neurosci. 27, 7640-7647.

43. Millar, J., Stamford, J. A., Kruk, Z. L., and Wightman, R. M. (1985)

Electrochemical, pharmacological, and electrophysiological evidence of rapid

dopamine release and removal in the rat caudate nucleus following electrical

stimulation of the median forebrain bundle, Eur. J. Pharmacol. 109, 341-348.

44. Porzgen, P., Park, S. K., Hirsh, J., Sonders, M. S., and Amara, S. G. (2001) The

antidepressant-sensitive dopamine transporter in Drosophila melanogaster: A

primordial carrier for catecholamines, Mol. Pharmacol. 59, 83-95.

81

Page 98: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

45. Cass, W. A., Zahniser, N. R., Flach, K. A., and Gerhardt, G. A. (1993) Clearance

of exogenous dopamine in rat dorsal striatum and nucleus accumbens: role of

metabolism and effects of locally applied uptake inhibitors, J. Neurochem. 61,

2269-2278.

46. Church, W. H., Justice Jr, J. B., and Byrd, L. D. (1987) Extracellular dopamine in

rat striatum following uptake inhibition by cocaine, nomifensine, and benztropine,

Eur. J. Pharmacol. 139, 345-348.

47. Morrison, P. F., Bungay, P. M., Hsiao, J. K., Ball, B. A., Mefford, I. N., and

Dedrick, R. L. (1991) Quantitative microdialysis - Analysis of transients and

application to pharmacokinetics in brain, J. Neurochem. 57, 103-119.

48. Logman, M. J., Budygin, E. A., Gainetdinov, R. R., and Wightman, R. M. (2000)

Quantitation of in vivo measurements with carbon-fiber microelectrodes, J.

Neurosci. Methods 95, 95-102.

49. Miner, D. J., Rice, J. R., Riggin, R. M., and Kissinger, P. T. (1981) Voltammetry

of acetaminophen and its metabolites, Anal. Chem. 53, 2258-2263.

50. Saunders, C., Ferrer, J. V., Shi, L., Chen, J., Merrill, G., Lamb, M. E., Leeb-

Lundberg, L. M. F., Carvelli, L., Javitch, J. A., and Galli, A. (2000)

Amphetamine-induced loss of human dopamine transporter activity: An

internalization-dependent and cocaine-sensitive mechanism, Proc. Natl. Acad.

Sci. U.S.A. 97, 6850-6855.

51. Melikian, H. E., and Buckley, K. M. (1999) Membrane trafficking regulates the

activity of the human dopamine transporter, J. Neurosci. 19, 7699-7710.

52. Sandoval, V., Riddle, E. L., Ugarte, Y. V., Hanson, G. R., and Fleckenstein, A. E.

(2001) Methamphetamine-induced rapid and reversible changes in dopamine

transporter function: an in vitro model, J. Neurosci. 21, 1413-1419.

53. Fleckenstein, A. E., Haughey, H. M., Metzger, R. R., Kokoshka, J. M., Riddle, E.

L., Hanson, J. E., Gibb, J. W., and Hanson, G. R. (1999) Differential effects of

psychostimulants and related agents on dopaminergic and serotonergic transporter

function, Eur. J. Pharmacol. 382, 45-49.

54. Kilty, J., Lorang, D., and Amara, S. (1991) Cloning and expression of a cocaine-

sensitive rat dopamine transporter, Science 254, 578-579.

55. Richelson, E., and Pfenning, M. (1984) Blockade by antidepressants and related

compounds of biogenic amine uptake into rat brain synaptosomes: Most

antidepressants selectively block norepinephrine uptake, Eur. J. Pharmacol. 104,

277-286.

56. Chen, R., Wei, H., Hill, E., Chen, L., Jiang, L., Han, D., and Gu, H. (2007) Direct

evidence that two cysteines in the dopamine transporter form a disulfide bond,

Mol. Cell. Biochem. 298, 41-48.

82

Page 99: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 4: Oral Administration of Methylphenidate Blocks the

Effect of Cocaine on Uptake at the Drosophila Dopamine

Transporter*

*In preparation for submission to ACS Chem. Neurosci.

83

Page 100: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

Cocaine addiction is a disease currently estimated to affect 2.1 million users in the

United States alone (1). The molecular and cellular actions of cocaine in the brain are

complex and affect the neurotransmission of several chemicals including dopamine,

serotonin, and norepinephrine through alteration of their transporter function (2-5).

Voltage-gated sodium channels are also blocked by cocaine (6), further supporting the

idea that cocaine works as a nonselective drug in the central nervous system (CNS), and

making it difficult for scientists to develop a suitable drug treatment to combat cocaine

addiction (7-9). The reinforcing and addictive properties of cocaine have been linked to

an increase in extracellular dopamine levels which are caused by blocking the dopamine

transporter (10, 11). It is widely accepted that cocaine decreases dopamine uptake

through binding of the dopamine transporter (12, 13), and it has been demonstrated that

the euphoric feeling experienced by cocaine abusers is associated with the blockade and

subsequent increase in extracellular dopamine (14).

Methylphenidate (Ritalin®), a commonly prescribed medication for the treatment

of attention deficit hyperactivity disorder (ADHD) (15), blocks the dopamine transporter

with a binding affinity similar to that of cocaine and increases the extracellular dopamine

concentration in the human brain (16-18). Methylphenidate has been shown to increase

extracellular norepinephrine by blocking the norepinephrine transporter as well (19).

Although cocaine and methylphenidate undergo similar binding to the dopamine

transporter in the CNS, the abuse potential of the two psychostimulants is different.

Typically, drugs that demonstrate reinforcing effects in laboratory animals are abused by

humans (20), and monkeys will self-administer methylphenidate and cocaine

84

Page 101: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

intravenously at similar rates (21, 22). However, studies have demonstrated that adult

humans do not consistently choose oral methylphenidate over placebo (23-25). While

methylphenidate is abused by humans, its abuse is much more limited than that of

cocaine which is considered one of the most addictive drugs known (26-28).

The route of administration of psychostimulants alters their pharmacokinetic

properties which can influence the abuse potential of a particular drug (29). When three

cocaine administration routes were compared, including smoked (crack), intravenous, and

intranasal, the length of time for each preparation to reach the CNS and for the euphoric

feeling to be experienced by the user was different (30, 31). Although the administered

doses of cocaine caused equivalent levels of blockade of the dopamine transporter with

all three routes, smoked cocaine was found to have a higher abuse potential, greater

reinforcing properties, and to be more addictive than the other two routes. When the

length of time for different routes of methylphenidate administration was investigated,

orally administered methylphenidate took approximately eight times longer than

intravenous administration to reach maximum blockade of the dopamine transporter (17,

32). Indeed, the observation that the shorter the time interval between intake of a drug

and the perceived affects of a drug, the greater the reinforcing properties and therefore

addictive potential of that drug has been documented (33, 34). The slow adsorption of

oral methylphenidate is believed to be an important factor in its limited abuse.

While orally administered methylphenidate in humans has been found to cause

little, if any, euphoric feelings (17, 35), intravenous administration of methylphenidate by

cocaine abusers causes feelings that are similar to those experienced with intravenous

cocaine use (36, 37). Both drugs have a fast adsorption rate in the brain (maximum

85

Page 102: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

concentration occurs in 2-8 min for cocaine and 4-10 min for methylphenidate,

respectively) which is thought to elicit the hedonic feeling associated with drug abuse

(37). The uptake rate of intravenous administration of cocaine and methylphenidate is

similar; however, the clearance rate of the two stimulants differs significantly. The half-

life of methylphenidate in the brain, based on the duration of dopamine transporter

blockade, is longer than that of cocaine (75-90 min vs. 15-25 min, respectively) even

though the initial reinforcing feeling it gives the user disappears just as quickly (~10 min)

as with cocaine use (37). The clearance of methylphenidate from the brain is necessary

before it is possible for an individual to fully experience the reinforcing effects of the

drug again, thus it is speculated that frequent repeated administration and overall abuse of

intravenous methylphenidate is limited in comparison to cocaine.

Over the last decade, methylphenidate has been investigated as a potential

agonist, or replacement medication, for cocaine addiction treatment as a similar approach

has been successful where methadone is used for treating opiate addiction (38, 39).

Several studies have investigated the effects of oral methylphenidate on cocaine users,

and mixed results have been found. Individuals experiencing fewer cravings and reduced

cocaine use with methylphenidate treatment have been reported (40), while other studies

have found no change in cocaine use or cravings for cocaine users after taking

methylphenidate (35, 41). Studies involving the subset of cocaine users who also had

symptoms of ADHD have reported more promising results (42-47). The ADHD cocaine

users experienced fewer cocaine cravings, decreased their cocaine use, and felt some

degree of improvement in their ADHD symptoms. These data suggest methylphenidate

could be a successful agonist medication for cocaine users with ADHD, but they do not

86

Page 103: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

explain the conflicting results for cocaine users without ADHD. A better understanding

of the chemical mechanisms in the CNS during co-administration of methylphenidate and

cocaine is needed to shed light on this potential treatment for cocaine addiction.

Animal models including rats, mice, and primates have been used to investigate

neurochemicals in the CNS associated with drug addiction (10, 13, 21, 22). Techniques

that use invertebrates, such as Drosophila melanogaster (fruit fly) and Apis mellifera

(honey bee), for research involving drugs of abuse have been established as well (48-50).

A method recently developed by the Ewing laboratory (51, 52) utilizes fast-scan cyclic

voltammetry (FSCV) coupled with carbon-fiber microelectrodes to quantify dopamine,

an electroactive neurotransmitter, in the CNS of Drosophila. Here, I apply this

microanalytical technique to study the efficacy of orally consumed methylphenidate on

dopamine uptake in Drosophila and its effect on preventing the actions of cocaine on the

dopamine transporter in vivo.

Methods

Chemicals. All chemicals were used as received and purchased from Sigma (St. Louis,

MO) unless otherwise stated. Adult-hemolymph like (AHL) saline (108 mM NaCl, 5

mM KCl, 2 mM CaCl2, 8.2 mM MgCl2, 4 mM NaHCO3, 1 mM NaH2PO4, 5 mM

trehalose (Fluka BioChemika, Buchs, Switzerland), 10 mM sucrose, 20 mM Trizma

base , pH 7.5) was made using ultrapure (18 MΩ·cm) water and filtered through a 0.2-

μm filter (53). All collagenase, dopamine, (+) cocaine, and methylphenidate bath

treatment solutions were prepared in AHL saline.

87

Page 104: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Drosophila rearing and in vivo preparation. Male flies, 4 to 10 days old, of the

Canton-S strain of Drosophila melanogaster were used for all experiments. Flies were

maintained at 25 °C on a standard cornmeal-agar medium. Some flies received an

additional food supplement consisting of a yeast paste containing 10 mM

methylphenidate aqueous solution that was prepared fresh daily. The flies were reared on

the methylphenidate yeast paste for 3-5 days prior to experimentation. All flies were

prepared for in vivo voltammetry as described in Chapters 2 and 3 (51, 52). Briefly, flies

were mounted in a homemade collar (38.1 mm diameter concave plexiglass disk with a

1.0 mm hole in the center) with low melting agarose (Fisher Scientific, Pittsburgh, PA)

following immobilization with ice. Under a stereoscope (Olympus SZ60, Melville, NY)

the cuticle was removed from the top portion of the head using dissection forceps and

scissors (World Precision Instruments, Sarasota, FL) to expose the brain. Following

microsurgery, 0.1% collagenase solution was applied to the head for 30 min to relax the

extracellular matrix in the brain. The immobilized fly head was then rinsed and bathed

with AHL saline, allowing the preparation to remain viable for 1.5 - 2.5 h.

Electrochemical measurements. The fabrication of the cylindrical carbon-fiber

microelectrodes used for this study has been described in detail previously (51, 54). The

exposed carbon fiber portion of the cylindrical electrodes was 40-50 μm in length. The in

vitro electrode calibration with standard dopamine solutions is shown in the Appendix.

In all experiments, the Ag/AgCl reference electrode used was made by chloridizing a

silver wire (0.25 mm diameter, 99.999% purity, Alfa Aesar, Ward Hill, MA) in bleach

overnight. Electrodes were positioned using micromanipulators purchased from Newport

(421 series, Irvine, CA). Glass capillaries (B120-69-10, Sutter Instruments, Novato, CA)

88

Page 105: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

were pulled using a glass capillary puller (P-97, Sutter Instruments) and cut to an opening

of ~5 μm to form micropipet injectors. The injectors were used to exogenously apply 1.0

mM dopamine solution by coupling them to a microinjection system (Picospritzer II,

General Valve Corporation, Fairfield, NJ).

A Dagan Chem-Clamp potentiostat (Dagan Corporation, Minneapolis, MN) and

two data acquisition boards (PCI-6221, National Instruments, Austin, TX) run by the TH

1.0 CV program (ESA, Chelmsford, MA) were used to collect all electrochemical data

(55). Cyclic voltammograms were obtained by applying a triangular waveform potential

(-0.6 V to +1.0 V vs. a Ag/AgCl reference electrode) repeated every 100 ms at a scan rate

of 200 V/s (low pass Bessel filter at 3 kHz). All electrodes were allowed to cycle for at

least 15 min prior to recording to stabilize the background current. The recorded current

response was converted to dopamine concentration via in vitro electrode calibration

(Appendix). Statistical analysis was accomplished using Prism 5.0 (GraphPad Software,

La Jolla, CA).

Results and Discussion

Dopamine uptake in the Drosophila CNS following cocaine bath treatment. In

Chapter 2, I described the development of a procedure for in vivo electrochemical

detection in adult Drosophila (51), and I demonstrated its use to study the effects of

cocaine and methylphenidate on the clearance of the redox-active neurotransmitter

dopamine in Chapter 3 (52). The Drosophila brain contains dopaminergic neurons

clustered together in several distinct locations with the largest neuronal cluster located in

the protocerebral anterior medial (PAM) region (56). By inserting a cylindrical carbon-

89

Page 106: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

fiber microelectrode into the PAM area of a Drosophila brain, changes in uptake of

exogenously applied dopamine can be quantified. This method is used in this chapter to

monitor the effects of cocaine and methylphenidate on dopamine clearance in the

Drosophila CNS.

Following fly microsurgery (see Methods), a carbon-fiber working electrode was

placed in the PAM region. Dopamine was exogenously applied just above the fly brain

tissue with a micropipet injector, and background-subtracted FSCV was used to measure

the current response in the extracellular fluid of the CNS over time. Using the peak

dopamine concentration, [DA]max, to monitor changes in the uptake of extracellular

dopamine in the CNS has been established (52, 57), and this parameter is utilized here.

Initially, the in vivo baseline current response was recorded for 3 min after a 1.0

mM dopamine solution was exogenously applied to the PAM area for 1.0 s (~150 pmol

dopamine applied). Following three baseline measurements, the fly brain was bathed in

1.0 mM cocaine, which has been shown to inhibit dopamine uptake by the Drosophila

dopamine transporter (52), for 5 min and then dopamine was applied again while the

current response was recorded. Dopamine injections were repeated every 5 min

throughout the 25 min cocaine treatment. Figure 4.1A compares a baseline concentration

trace of dopamine (black line) with a concentration trace obtained after cocaine treatment

(red line). The representative traces demonstrate the effectiveness of a bath application

of cocaine in blocking dopamine uptake via the dopamine transporter.

Although the effect of different administration routes of methylphenidate has been

studied in mammalian systems, to my knowledge no reports have been published on the

efficacy of orally consumed methylphenidate in Drosophila. Flies were orally fed a paste

90

Page 107: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 4.1. Effect of orally consumed methylphenidate on cocaine inhibition of the

dopamine transporter in the adult Drosophila brain. (A) Representative concentration

traces (taken from the maximum anodic peak potential) of exogenously applied 1.0 mM

dopamine in a wild-type fly that did not consume methylphenidate before (baseline, black

line) and after 1.0 mM cocaine bath treatment (red line). A significant increase in

dopamine concentration was observed following cocaine application. Dopamine

concentration was determined from conversion of the measured current using in vitro

electrode calibration. The black arrow corresponds to a 1.0 s dopamine application

beginning at 5.0 s. (B) Representative concentration traces of exogenously applied 1.0

mM dopamine in a wild-type fly that consumed methylphenidate before (baseline, black

line) and after 1.0 mM cocaine bath treatment (red line). No change in dopamine

concentration was observed following cocaine application. (C) Applied potential vs. time

gives a visual representation of successive voltammograms that correspond to the

baseline current (black line) in (A) with current viewed in false color. (D) Applied

potential vs. time gives a visual representation of successive voltammograms that

correspond to the baseline current (black line) in (B) with current viewed in false color.

91

Page 108: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

consisting of a 10 mM methylphenidate solution mixed with yeast for 3-5 days prior to

the cocaine bath application experiment described above. The 1.0 mM cocaine bath

application treatment did not affect the [DA]max following dopamine injection for the flies

that consumed the methylphenidate paste (Figure 4.1B). To visualize changes over time,

a false-color representation of current is used where the green corresponds to the

oxidation of dopamine, and the reduction of the orthoquinone is represented in blue (58).

The color representation for a baseline measurement of current for a fly that did not

consume methylphenidate is shown in Figure 4.1C while Figure 4.1D corresponds to a

fly that consumed methylphenidate.

Effect of orally consumed methylphenidate on dopamine uptake in Drosophila. In

Chapter 3, it was shown that a 1.0 mM bath application of methylphenidate is sufficient

to effectively block dopamine uptake occurring via the dopamine transporter in

Drosophila wild-type flies (52). Here, the results are compared to wild-type flies that

orally consumed methylphenidate prior to administration of the 1.0 mM bath application

of methylphenidate to determine if oral administration of methylphenidate is capable of

blocking the Drosophila dopamine transporter in vivo to a similar degree as the bath

administration.

Flies that consumed a paste consisting of a 10 mM methylphenidate solution

mixed with yeast for 3-5 days prior to the methylphenidate bath treatment were compared

to flies that did not consume the methylphenidate paste. The 1.0 mM methylphenidate

bath application treatment had no effect on the peak current response following dopamine

injection for the flies that consumed methylphenidate (Figure 4.2A). To eliminate

systematic effects, such as slight differences in dopamine injector positioning between

92

Page 109: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 4.2. Effect of orally consumed methylphenidate on Drosophila dopamine

transporter function. (A) The uptake of exogenously applied 1.0 mM dopamine by flies

that orally consumed 10 mM methylphenidate (black) was compared with flies that did

not consume methylphenidate (green). After baseline dopamine measurements, both

groups of flies were treated with bath-applied 1.0 mM methylphenidate for 25 min.

There was a significant increase in normalized [DA]max for the flies that did not consume

methylphenidate prior to the bath methylphenidate treatment (mean ± SEM; two-way

ANOVA, p = 0.05 for interaction, p < 0.0001 for two fly groups, p = 0.03 for bath

treatment, n = 5-6). (B) After baseline dopamine measurements, both groups of flies

were treated with bath-applied 1.0 mM cocaine for 25 min. There was a significant

increase in normalized [DA]max for the flies that did not consume methylphenidate prior

to the bath cocaine treatment (mean ± SEM; two-way ANOVA, p = 0.009 for interaction,

p < 0.0001 for two fly groups, p = 0.002 for bath treatment, n = 6). SEMs for the

baseline bars are too small to see.

93

Page 110: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

flies, the [DA]max was normalized, and the averages of several flies were compared. For

normalization, the [DA]max from two of the dopamine baseline measurements for a fly

were averaged together, and all measurements for that particular fly were calculated as a

percent of the average baseline measurement (i.e., [DA]max normalized). The normalized

[DA]max averages for the two groups of flies, flies that consumed the 10 mM oral

methylphenidate paste and flies that did not, were compared. The flies that did not eat

the methylphenidate paste displayed a significantly higher change in normalized [DA]max

following the 1.0 mM bath application of methylphenidate compared to the flies that did

consume the methylphenidate paste (two-way analysis of variance (ANOVA), p = 0.05

for interaction, p < 0.0001 for two fly groups, p = 0.03 for bath treatment, n = 5-6). A

bath application of methylphenidate does not appear to change uptake by the dopamine

transporter of flies that have previously consumed methylphenidate. This suggests oral

consumption of methylphenidate blocks the Drosophila dopamine transporter in a

manner similar to that of orally consumed methylphenidate in humans (17).

Cocaine effects are undetectable following the oral consumption of methylphenidate.

Bath application of 1.0 mM cocaine was shown to effectively block the Drosophila

dopamine transporter in Chapter 3 (52). To investigate whether prior methylphenidate

consumption is able to affect the action of cocaine on the dopamine transporter, flies

were tested that had been fed the 10 mM methylphenidate paste. Electrochemistry was

used to monitor exogenously applied dopamine clearance before and after application of

a 1.0 mM cocaine bath. Voltammograms were obtained with dopamine exogenously

applied every 5 min for 25 min. Figure 4.2B is a comparison of the normalized [DA]max

for the two groups of flies. The flies that did not consume methylphenidate experienced a

94

Page 111: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

significantly increased normalized [DA]max following the cocaine bath treatment, whereas

flies that had consumed methylphenidate did not exhibit a change in dopamine uptake

with the cocaine treatment (two-way ANOVA, p = 0.009 for interaction, p < 0.0001 for

two fly groups, p = 0.002 for bath treatment, n = 6). Again, this indicates that orally

consumed methylphenidate effectively blocks the Drosophila dopamine transporter

function in vivo, and other effects from the addition of cocaine are not observed. This

result is comparable to the mechanism of action that has been observed in baboons that

were given methylphenidate prior to cocaine administration (37). In addition, the

regional distribution patterns of methylphenidate and cocaine throughout the human brain

have been found to be almost identical using positron emission tomography with similar

in vivo potencies at the human dopamine transporter (37, 59). These data reinforce the

validity of using Drosophila as a model system for studying mechanisms of cocaine

addiction in humans.

Comparison of the extracellular dopamine concentration in the Drosophila CNS

following drug treatments. To further investigate the functionality of the Drosophila

dopamine transporter following drug treatments, non-normalized [DA]max data was

considered. The non-normalized [DA]max was obtained from the cyclic voltammograms

of multiple flies, and the averages for four different groups of flies were compared

(Figure 4.3). The average [DA]max of flies not treated with any form of cocaine or

methylphenidate (“untreated”) was significantly lower than that of flies that had

consumed oral methylphenidate (Student’s t-test, p = 0.009) and flies treated with bath-

applied cocaine only (Student’s t-test, p = 0.006). This confirms the dopamine

transporter in the untreated flies was more functional than that of flies treated with oral

95

Page 112: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 4.3. Comparison of dopamine concentration in the Drosophila CNS following

drug treatments. Untreated flies had a significantly lower extracellular dopamine

concentration than either flies that orally consumed 10 mM methylphenidate (mean ±

SEM; Student’s t-test, p = 0.009 (**)) or flies that were treated with 1.0 mM bath-applied

cocaine for 20 min (mean ± SEM; Student’s t-test, p = 0.006 (**)). There was no

significant difference between the flies that orally consumed methylphenidate and the

flies treated with bath-applied cocaine (mean ± SEM; Student’s t-test, p = 0.9). Flies

treated with 1.0 mM bath-applied methylphenidate for 20 min were not significantly

different from the flies of the other three groups (mean ± SEM; Student’s t-test, p = 0.12-

0.17).

96

Page 113: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

methylphenidate or bath-applied cocaine. The non-normalized data of flies subjected to

only bath-applied methylphenidate were not significantly different from the flies of the

other three groups. Of importance, there was no significant difference between the flies

that had orally consumed methylphenidate and the flies that were treated with the bath-

applied cocaine.

Conclusions

Although there has been improvement in understanding the actions of cocaine in

the brain, an effective drug treatment has yet to be found for cocaine addiction.

Methylphenidate binds the dopamine transporter and increases extracellular dopamine

levels in the CNS similar to cocaine without producing as many of the addictive and

reinforcing properties. In this chapter the Drosophila model system was utilized to

investigate the mechanism behind treating cocaine addiction with methylphenidate. In

vivo electrochemical measurements suggest oral consumption of methylphenidate

sufficiently blocks the Drosophila dopamine transporter thus preventing further

inhibition of the transporter by cocaine applied directly to the CNS. This highlights the

possibility of methylphenidate as a potential treatment for cocaine addiction and the value

of Drosophila as a model system for future drug abuse research.

97

Page 114: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Substance Abuse and Mental Health Services Administration, Office of Applied

Studies (2008) Results from the 2007 national survey on drug use and health:

National findings, DHHS Pub. No. SMA 08-4343, Rockville, MD: SAMHSA.

2. Rocha, B. A., Fumagalli, F., Gainetdinov, R. R., Jones, S. R., Ator, R., Giros, B.,

Miller, G. W., and Caron, M. G. (1998) Cocaine self-administration in dopamine-

transporter knockout mice, Nat. Neurosci. 1, 132-137.

3. Sora, I., Wichems, C., Takahashi, N., Li, X.-F., Zeng, Z., Revay, R., Lesch, K.-P.,

Murphy, D. L., and Uhl, G. R. (1998) Cocaine reward models: Conditioned place

preference can be established in dopamine- and in serotonin-transporter knockout

mice, Proc. Natl. Acad. Sci. U.S.A. 95, 7699-7704.

4. Xu, F., Gainetdinov, R. R., Wetsel, W. C., Jones, S. R., Bohn, L. M., Miller, G.

W., Wang, Y.-M., and Caron, M. G. (2000) Mice lacking the norepinephrine

transporter are supersensitive to psychostimulants, Nat. Neurosci. 3, 465-471.

5. Hall, F. S., Li, X. F., Sora, I., Xu, F., Caron, M., Lesch, K. P., Murphy, D. L., and

Uhl, G. R. (2002) Cocaine mechanisms: enhanced cocaine, fluoxetine, and

nisoxetine place preferences following monoamine transporter deletions,

Neuroscience 115, 153-161.

6. Weidmann, S. (1955) Effects of calcium ions and local anaesthetics on electrical

properties of Purkinje fibers, J. Physiol. 129, 568-582.

7. Lima, M. S., Soares, B. G. O., Reisser, A. A. P., and Farrell, M. (2002)

Pharmacological treatment of cocaine dependence: a systematic review, Addiction

97, 931-949.

8. Gorelick, D. A., Gardner, E. L., and Xi, Z.-X. (2004) Agents in development for

the management of cocaine abuse, Drugs 64, 1547-1573.

9. Karila, L., Gorelick, D., Weinstein, A., Noble, F., Benyamina, A., Coscas, S.,

Blecha, L., Lowenstein, W., Martinot, J. L., Reynaud, M., and Lépine, J. P.

(2008) New treatments for cocaine dependence: A focused review, Int. J.

Neuropsychopharmacolog. 11, 425-438.

10. Di Chiara, G., and Imperato, A. (1988) Drugs abused by humans preferentially

increase synaptic dopamine concentrations in the mesolimbic system of freely

moving rats, Proc. Natl. Acad. Sci. U.S.A. 85, 5274-5278.

11. Kuhar, M. J., Ritz, M. C., and Boja, J. W. (1991) The dopamine hypothesis of the

reinforcing properties of cocaine, Trends Neurosci. 14, 299-302.

12. Ritz, M. C., Lamb, R. J., Goldberg, S. R., and Kuhar, M. J. (1987) Cocaine

receptors on dopamine transporters are related to self-administration of cocaine,

Science 237, 1219-1223.

13. Giros, B., Jaber, M., Jones, S. R., Wightman, R. M., and Caron, M. G. (1996)

Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking

the dopamine transporter, Nature 379, 606-612.

14. Volkow, N. D., Wang, G. J., Fischman, M. W., Foltin, R. W., Fowler, J. S.,

Abumrad, N. N., Vitkun, S., Logan, J., Gatley, S. J., Pappas, N., Hitzemann, R.,

and Shea, C. E. (1997) Relationship between subjective effects of cocaine and

dopamine transporter occupancy, Nature 386, 827-830.

98

Page 115: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

15. Carrey, N. J., Wiggins, D. M., and Milin, R. P. (1996) Pharmacological treatment

of psychiatric disorders in children and adolescents - Focus on guidelines for the

primary care practitioner, Drugs 51, 750-759.

16. Schweri, M. M., Skolnick, P., Rafferty, M. F., Rice, K. C., Janowsky, A. J., and

Paul, S. M. (1985) [3H]Threo-(±)-methylphenidate binding to 3,4-

dihydroxyphenylethylamine uptake sites in corpus striatum: correlation with the

stimulant properties of ritalinic acid esters, J. Neurochem. 45, 1062-1070.

17. Volkow, N. D., Wang, G.-J., Fowler, J. S., Gatley, S. J., Logan, J., Ding, Y.-S.,

Hitzemann, R., and Pappas, N. (1998) Dopamine transporter occupancies in the

human brain induced by therapeutic doses of oral methylphenidate, Am. J.

Psychiatry 155, 1325-1331.

18. Volkow, N. D., Wang, G.-J., Fowler, J. S., Logan, J., Gerasimov, M., Maynard,

L., Ding, Y.-S., Gatley, S. J., Gifford, A., and Franceschi, D. (2001) Therapeutic

doses of oral methylphenidate significantly increase extracellular dopamine in the

human brain, J. Neurosci. 21, RC121.

19. Kuczenski, R., and Segal, D. S. (1997) Effects of methylphenidate on

extracellular dopamine, serotonin, and norepinephrine: comparison with

amphetamine, J. Neurochem. 68, 2032-2037.

20. U.S. Department of Health and Human Services (1989) Testing for abuse liability

of drugs in humans, in National Institute on Drug Abuse Research Monograph

Series No. 92 89-1613 (Fischman, M. W., and Mello, N. K., Eds.), U.S.

Government Printing Office, Washington, DC.

21. Johanson, C. E., and Schuster, C. R. (1975) A choice procedure for drug

reinforcers: cocaine and methylphenidate in the rhesus monkey, J. Pharmacol.

Exp. Ther. 193, 676-688.

22. Bergman, J., Madras, B. K., Johnson, S. E., and Spealman, R. D. (1989) Effects of

cocaine and related drugs in nonhuman primates. III. Self-administration by

squirrel monkeys, J. Pharmacol. Exp. Ther. 251, 150-155.

23. Chait, L. D. (1994) Reinforcing and subjective effects of methylphenidate in

humans, Behav. Pharmacol. 5, 281-288.

24. Roehrs, T., Papineau, K., Rosenthal, L., and Roth, T. (1999) Sleepiness and

reinforcing and subjective effects of methylphenidate, Exp. Clin.

Psychopharmacol. 7, 145-150.

25. Rush, C. R., Essman, W. D., Simpson, C. A., and Baker, R. W. (2001)

Reinforcing and subject-rated effects of methylphenidate and d-amphetamine in

non-drug-abusing humans, J. Clin. Psychopharmacol. 21, 273-286.

26. Koob, G., and Bloom, F. (1988) Cellular and molecular mechanisms of drug

dependence, Science 242, 715-723.

27. Johanson, C. E., and Fischman, M. W. (1989) The pharmacology of cocaine

related to its abuse, Pharmacol. Rev. 41, 3-52.

28. Kollins, S. H., MacDonald, E. K., and Rush, C. R. (2001) Assessing the abuse

potential of methylphenidate in nonhuman and human subjects: A review,

Pharmacol. Biochem. Behav. 68, 611-627.

29. Verebey, K., and Gold, M. S. (1988) From coca leaves to crack: The effects of

dose and routes of administration in abuse liability, Psychiatr. Ann. 18, 513-520.

99

Page 116: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

30. Cone, E. J. (1995) Pharmacokinetics and pharmacodynamics of cocaine, J. Anal.

Toxicol. 19, 459-478.

31. Volkow, N. D., Wang, G. J., Fischman, M. W., Foltin, R., Fowler, J. S.,

Franceschi, D., Franceschi, M., Logan, J., Gatley, S. J., Wong, C., Ding, Y.-S.,

Hitzemann, R., and Pappas, N. (2000) Effects of route of administration on

cocaine induced dopamine transporter blockade in the human brain, Life Sci. 67,

1507-1515.

32. Volkow, N. D., Fowler, J. S., Wang, G. J., Ding, Y. S., and Gatley, S. J. (2002)

Role of dopamine in the therapeutic and reinforcing effects of methylphenidate in

humans: results from imaging studies, Eur. Neuropsychopharmacol. 12, 557-566.

33. Balster, R. L., and Schuster, C. R. (1973) Fixed-interval schedule of cocaine

reinforcement - effect of dose and infusion duration, J. Exp. Anal. Behav. 20, 119-

129.

34. Oldendorf, W. H. (1992) Some relationships between addiction and drug delivery

to the brain, NIDA Res. Monogr. 120, 13-25.

35. Grabowski, J., Roache, J. D., Schmitz, J. M., Rhoades, H., Creson, D., and

Korszun, A. M. (1997) Replacement medication for cocaine dependence:

methylphenidate, J. Clin. Psychopharmacol. 17, 485-488.

36. Parran, T. V., Jr, and Jasinski, D. R. (1991) Intravenous methylphenidate abuse:

prototype for prescription drug abuse, Arch. Intern. Med. 151, 781-783.

37. Volkow, N. D., Ding, Y.-S., Fowler, J. S., Wang, G.-J., Logan, J., Gatley, J. S.,

Dewey, S., Ashby, C., Liebermann, J., Hitzemann, R., and Wolf, A. P. (1995) Is

methylphenidate like cocaine? Studies on their pharmacokinetics and distribution

in the human brain, Arch. Gen. Psychiatry 52, 456-463.

38. Dole, V. P., and Nyswande.M. (1965) A medical treatment for diacetylmorphine

(heroin) addition - a clinical trial with methadone hydrochloride, JAMA-J. Am.

Med. Assoc. 193, 646-650.

39. Johnson, R. E., Jaffe, J. H., and Fudala, P. J. (1992) A controlled trial of

buprenorphine treatment for opioid dependence, JAMA-J. Am. Med. Assoc. 267,

2750-2755.

40. Winhusen, T., Somoza, E., Singal, B. M., Haffer, J., Apparaju, S., Mezinskis, J.,

Desai, P., Elkashef, A., Chiang, C. N., and Horn, P. (2006) Methylphenidate and

cocaine: A placebo-controlled drug interaction study, Pharmacol. Biochem.

Behav. 85, 29-38.

41. Roache, J. D., Grabowski, J., Schmitz, J. M., Creson, D. L., and Rhoades, H. M.

(2000) Laboratory measures of methylphenidate effects in cocaine-dependent

patients receiving treatment, J. Clin. Psychopharmacol. 20, 61-68.

42. Khantzian, E. (1983) An extreme case of cocaine dependence and marked

improvement with methylphenidate treatment, Am. J. Psychiatry 140, 784-785.

43. Khantzian, E. J., Gawin, F., Kleber, H. D., and Riordan, C. E. (1984)

Methylphenidate (Ritalin®) treatment of cocaine dependence: A preliminary

report, J. Subst. Abuse Treat. 1, 107-112.

44. Levin, F. R., Evans, S. M., McDowell, D. M., and Kleber, H. D. (1998)

Methylphenidate treatment for cocaine abusers with adult attention-

deficit/hyperactivity disorder: A pilot study, J. Clin. Psychiatry 59, 300-305.

100

Page 117: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

45. Somoza, E. C., Winhusen, T. M., Bridge, T. P., Rotrosen, J. P., Vanderburg, D.

G., Harrer, J. M., Mezinskis, J. P., Montgomery, M. A., Ciraulo, D. A., Wulsin, L.

R., and Barrett, J. A. (2004) An open-label pilot study of methylphenidate in the

treatment of cocaine dependent patients with adult attention deficit/hyperactivity

disorder, J. Addict. Dis. 23, 77-92.

46. Collins, S. L., Levin, F. R., Foltin, R. W., Kleber, H. D., and Evans, S. M. (2006)

Response to cocaine, alone and in combination with methylphenidate, in cocaine

abusers with ADHD, Drug Alcohol Depend. 82, 158-167.

47. Levin, F. R., Evans, S. M., Brooks, D. J., and Garawi, F. (2007) Treatment of

cocaine dependent treatment seekers with adult ADHD: Double-blind comparison

of methylphenidate and placebo, Drug Alcohol Depend. 87, 20-29.

48. McClung, C., and Hirsh, J. (1998) Stereotypic behavioral responses to free-base

cocaine and the development of behavioral sensitization in Drosophila, Curr.

Biol. 8, 109-112.

49. Bainton, R. J., Tsai, L. T. Y., Singh, C. M., Moore, M. S., Neckameyer, W. S.,

and Heberlein, U. (2000) Dopamine modulates acute responses to cocaine,

nicotine, and ethanol in Drosophila, Curr. Biol. 10, 187-194.

50. Barron, A. B., Maleszka, R., Helliwell, P. G., and Robinson, G. E. (2009) Effects

of cocaine on honey bee dance behavior, J. Exp. Biol. 212, 163-168.

51. Makos, M. A., Kim, Y.-C., Han, K.-A., Heien, M. L., and Ewing, A. G. (2009) In

vivo electrochemical measurements of exogenously applied dopamine in

Drosophila melanogaster, Anal. Chem. 81, 1848-1854.

52. Makos, M. A., Han, K.-A., Heien, M. L., and Ewing, A. G. (2010) Using in vivo

electrochemistry to study the physiological effects of cocaine and other stimulants

on the Drosophila melanogaster dopamine transporter, ACS Chem. Neurosci. 1,

74-83.

53. Wang, J. W., Wong, A. M., Flores, J., Vosshall, L. B., and Axel, R. (2003) Two-

photon calcium imaging reveals an odor-evoked map of activity in the fly brain,

Cell 112, 271-282.

54. Dayton, M. A., Brown, J. C., Stutts, K. J., and Wightman, R. M. (1980) Faradaic

electrochemistry at micro-voltammetric electrodes, Anal. Chem. 52, 946-950.

55. Heien, M. L., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman, R.

M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

56. Nassel, D. R., and Elekes, K. (1992) Aminergic neurons in the brain of blowflies

and Drosophila: dopamine- and tyrosine hydroxylase-immunoreactive neurons

and their relationship with putative histaminergic neurons, Cell Tissue Res. 267,

147-167.

57. Zahniser, N. R., Larson, G. A., and Gerhardt, G. A. (1999) In vivo dopamine

clearance rate in rat striatum: Regulation by extracellular dopamine concentration

and dopamine transporter inhibitors, J. Pharmacol. Exp. Ther. 289, 266-277.

58. Michael, D., Travis, E. R., and Wightman, R. M. (1998) Color images for fast-

scan CV, Anal. Chem. 70, 586A-592A.

101

Page 118: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

59. Volkow, N. D., Wang, G. J., Fowler, J. S., Fischman, M., Foltin, R., Abumrad, N.

N., Gatley, S. J., Logan, J., Wong, C., Gifford, A., Ding, Y.-S., Hitzemann, R.,

and Pappas, N. (1999) Methylphenidate and cocaine have a similar in vivo

potency to block dopamine transporters in the human brain, Life Sci. 65, PL7-

PL12.

102

Page 119: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 5: Methods for Stimulating Dopamine Release in the

Drosophila CNS

103

Page 120: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

Electrochemistry has been used to monitor neurotransmission in the brain for

several decades. A voltammetric technique capable of measuring species in rat brain

tissue was reported by Leland Clark and coworkers in 1965 (1). This was followed by in

vivo cyclic voltammetry work completed in Ralph Adams’s laboratory. They used 1-2

mm diameter electrodes made from carbon paste or solidified graphite-epoxy resin to

measure electroactive neurochemicals in the brains of anesthetized rats (2, 3). The

fabrication and characterization of the carbon-fiber microelectrode by Wightman and

coworkers in 1980 allowed more rapid voltammetric measurements to be carried out in

comparison to the conventional electrodes used at that time (4-6). This tool led to the

progression of electroanalytical techniques to monitor in vivo neurotransmitter dynamics

in the central nervous system (CNS) (7-11). Several voltammetric techniques have been

developed for in vivo measurements including differential pulse voltammetry, normal

pulse voltammetry, linear sweep voltammetry, and cyclic voltammetry (12). Fast-scan

cyclic voltammetry (FSCV) has become a widely used voltammetric technique for in vivo

applications because it is capable of detecting neurotransmitter changes in real-time as

well as providing chemical information regarding the identity of the electroactive species

being measured (12-15).

The neurotransmitter dopamine is of interest because it is known to regulate

several human physiological processes including motivation and addiction. Dopamine

neurotransmission is believed to contribute to the reinforcing and addictive properties of

drugs of abuse such as cocaine and amphetamines (16-18). Monitoring in vivo changes

in dopamine uptake in the Drosophila CNS in the presence of psychostimulants has been

104

Page 121: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

discussed in Chapters 2-4. In addition to uptake, dopamine release is an important

component of neurotransmission. Various electrochemical techniques have been

developed to detect in vivo dopamine release. New approaches to stimulate

neurotransmitter release have been developed as well and used in conjunction with

electrochemical detection including chemical, electrical, and most recently optogenetic

stimulation.

A number of chemicals are capable of evoking neurotransmitter release from a

neuron. A widely used approach is to depolarize the cell membrane using an elevated

concentration of K+ ions (19, 20). This causes voltage-dependent ion channels to open,

whereby an action potential is generated, and that causes neurotransmitter-filled vesicles

to release their contents into the extracellular space. The neurotransmitter dopamine is

released in both the rat and mouse CNS in response to K+ ion stimulation (21-23).

Veratridine is another depolarizing agent that effectively induces dopamine release in

mammals (24-26). Caffeine and nicotine are two stimulants that have been shown to

increase extracellular dopamine in the rat CNS through binding of their respective

receptors, adenosine and nicotinic acetylcholine (27-31). In addition, Ba2+

ions have

been reported to trigger dopamine release in mammalian neuronal preparations (32, 33).

The chemical stimulants listed above are summarized in Table 5.1.

While chemical stimulation is an efficient way to elicit dopamine release in vivo,

a more controlled approach is to utilize the electrical properties of neurons. Brief

electrical pulses generate action potentials which leads to the release of neurotransmitters

(34, 35). Stimulation of a particular pathway in regions where multiple pathways exist is

possible with this method through specific electrode placement, which offers an

105

Page 122: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Table 5.1. Eliciting dopamine release via chemical stimulation.

Stimulant Action References

potassium depolarizes cell membrane

in vivo rat and mouse CNS (21, 22)

in vitro rat brain tissue (23)

veratridine depolarizing agent that acts as a

sodium channel agonist

in vivo rat CNS (24)

in vitro rat brain tissue (25)

in vitro guinea pig cochleae (26)

caffeine adenosine receptor antagonist

in vivo rat CNS (27, 28)

nicotine nicotinic acetylcholine receptor

agonist

in vivo rat CNS (29-31)

barium

thought to induce exocytotic

release by a mechanism similar to

calcium

bovine cell culture (32)

rat neuronal preparation (33)

106

Page 123: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

advantage over chemical stimulation methods. Dopamine release induced by electrical

stimulation was measured in vivo by Ewing et al. in the rat CNS (36). Since then

electrical stimulation has been extensively used to elicit dopamine release in many

applications involving in vivo dopamine neurotransmission (37-44). One drawback of

electrical stimulation is that it causes most neurons within the stimulated region to

undergo neurotransmitter release simultaneously, and this is not an accurate model for the

timing of naturally occurring dopamine release events in the brain.

A targeted approach to manipulate neuronal function that has recently been

investigated utilizes light in place of chemicals or electricity to stimulate cell activity (45,

46). Channelrhodopsin-2 (ChR2) is an ion channel found in the green alga

Chlamydomonas reinhardtii that can be genetically inserted into neurons and optically

controlled. The ChR2 protein is composed of seven trans-membrane domains and

contains the chromophore all-trans retinal. Upon exposure to blue light, all-trans retinal

undergoes isomerization to 13-cis retinal, thus causing a conformational change that

allows the trans-membrane protein to open (47, 48). Na+ ions flow through the channel

and enter the cell as they travel down their electrochemical gradient (45). The increase in

positive charge in the cell causes depolarization of the cell membrane and leads to

vesicular neurotransmitter release into the extracellular space (Figure 5.1). By

genetically expressing ChR2 in a specific type of neuron, blue light stimulation can be

used to elicit release of a particular neurotransmitter of interest (49). Preliminary studies

have demonstrated that ChR2 can be expressed in both adult Drosophila and larvae

dopaminergic neurons and specific release of dopamine initiated with blue light (49-51).

107

Page 124: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 5.1. Cartoon depiction of the effects of blue light exposure on neurons expressing

Channelrhodopsin-2 (ChR2). Upon illumination with blue light, vesicular

neurotransmitter release is stimulated in neurons (light green) genetically altered with

ChR2. The lipid bilayer membrane (light purple) containing the ChR2 ion channel is

enlarged for clarity.

108

Page 125: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

In this chapter, I investigate chemical, electrical, and optogenetic methods to

stimulate dopamine release in the Drosophila CNS with FSCV detection. Chemical and

electrical stimulation tools successfully used in larger mammalian model systems were

modified for and tested in the smaller fly CNS (~5 nL). In addition, I explore using blue

light to noninvasively stimulate neurochemical release through the novel ChR2 ion

channel by genetic insertion of the ChR2 channel in Drosophila dopamine neurons. The

results suggest optogenetic stimulation initiates targeted neuronal release in the

Drosophila CNS.

Methods

Chemicals. All chemicals were used as received and purchased from Sigma (St. Louis,

MO) unless otherwise stated. Adult-hemolymph like (AHL) saline (108 mM NaCl, 5

mM KCl, 2 mM CaCl2, 8.2 mM MgCl2, 4 mM NaHCO3, 1 mM NaH2PO4, 5 mM

trehalose (Fluka BioChemika, Buchs, Switzerland), 10 mM sucrose, 20 mM Trizma

base , pH 7.5) was made using ultrapure (18 MΩ·cm) water and filtered through a 0.2-

μm filter. All collagenase, dopamine, KCl, veratridine (EMD Biosciences, Inc., La Jolla,

CA), caffeine, nicotine, BaCl2, and (+) cocaine solutions were prepared in AHL saline.

In vivo Drosophila preparation. The Canton-S strain of Drosophila melanogaster was

used for the wild-type fly in this chapter. Using the Drosophila galactosidase-4-upstream

activating sequences (GAL4/UAS) gene-targeting system explained in Chapter 1, mutant

flies were bred by crossing female flies carrying ChR2 with male flies expressing

tyrosine hydroxylase (TH) to produce mutant flies of the genotype TH-GAL4/UAS:ChR2

(49, 52). The dopaminergic neurons of the mutant flies can be controlled with blue light

109

Page 126: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

stimulation. Wild-type flies were maintained on a standard cornmeal-agar medium,

while mutant flies were maintained in the dark and fed yeast containing 10 mM all-trans

retinal (light sensitive chemical necessary for ChR2 function) made fresh daily. All flies

were kept at 25 °C, and 4 to 10 day-old male flies were used for experimentation. Flies

were prepared for in vivo FSCV measurements as described in Chapter 2 (53). Briefly,

flies were immobilized on ice and mounted in a homemade collar (38.1 mm diameter

concave plexiglass disk with a 1.0 mm hole in the center) with low melt agarose (Fisher

Scientific, Pittsburgh, PA). Microsurgery was performed on a stereoscope (Olympus

SZ60, Melville, NY) using small dissection scissors and forceps (World Precision

Instruments, Sarasota, FL) to remove the cuticle. The head was covered with 0.1%

collagenase solution for 30 min to relax the extracellular matrix in the brain and then

rinsed and covered with AHL saline to maintain the viability of the preparation for 1.5 -

2.5 h.

Electrochemical measurements. Cylindrical carbon-fiber microelectrodes were

fabricated as described in Chapter 2 (53). Briefly, a single 5-μm diameter carbon fiber

(Amoco, Greenville, SC) was aspirated into a borosilicate glass capillary, and the

capillary was pulled using a regular glass capillary puller (P-97, Sutter Instruments,

Novato, CA). Electrical contact was made by back-filling the capillary with silver paint

(4922N DuPont, Delta Technologies Ltd., Stillwater, MN) and inserting a tungsten wire.

The carbon fiber was cut to a length of 40-50 μm, as measured from the glass junction.

Electrode tips were dipped into epoxy (Epo-Tek, Epoxy Technology, Billerica, MA) for

30 s to ensure a good seal between the fiber and the glass and then dipped into acetone

for 15 s to remove epoxy from the exposed carbon fiber. A Ag/AgCl reference electrode

110

Page 127: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

was made by chlorodizing a silver wire (0.25 mm diameter, 99.999% purity, Alfa Aesar,

Ward Hill, MA) in bleach overnight. All electrodes were positioned using

micromanipulators (421 series, Newport, Irvine, CA).

Electrochemical data were collected using either an Axopatch 200B Amplifier

(Axon Instruments, Foster City, CA) or a Dagan Chem-Clamp potentiostat (Dagan

Corporation, Minneapolis, MN) and two data acquisition boards (PCI-6221, National

Instruments, Austin, TX) run by the TH 1.0 CV program (ESA, Chelmsford, MA) (54).

Cyclic voltammograms were obtained using a triangular waveform (scanned -0.6 V to

+1.0 or +1.2 V vs. Ag/AgCl at 200 V/s) repeated every 100 ms (low pass Bessel filter at

3-5 kHz). Prior to voltammetric experiments, all electrodes were cycled for at least 15

min to stabilize the background current. Electrochemical responses were plotted and

statistical analysis performed using Prism 5.0 (GraphPad Software, La Jolla, CA). After

collection, voltammograms were smoothed (nearest neighbor smooth) and filtered at 2.0

kHz using the TH 1.0 CV program. The current traces were filtered at 0.5-1.0 Hz.

Chemical stimulation equipment. A single-barrel borosilicate glass capillary (B120-

69-10, Sutter Instruments) was used to make injectors to apply KCl (100-500 mM),

veratridine (100 µM), caffeine (1 mM), nicotine (100 µM), and BaCl2 (10 mM)

stimulation solutions. Micropipet injectors were fabricated by pulling the capillaries in a

glass capillary puller to an opening of approximately 5 μm. Stimulation solutions were

pneumatically applied using a Picospritzer II (General Valve Corporation, Fairfield, NJ).

Electrical stimulation equipment. The first type of electrode used to electrically

stimulate dopamine release in the fly was composed of two individually insulated, 75 µm

diameter platinum electrodes (MS303/9-B/SPC, Plastics One Inc., Roanoke, VA). The

111

Page 128: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

second type of electrode used was composed of two tungsten electrodes each with a 125

µm diameter shaft, but tapered at a 12° angle to a point (57720, A-M Systems Inc.,

Carlsborg, WA). A battery operated, constant current stimulus isolator (NeuroLogTM

System NL800, Digitimer Ltd., Holliston, MA) was used to deliver computer controlled

stimulation through the electrodes. A range of monophasic pulses was tested (0.1-10.0

V, 0.5-8.0 ms per phase, 1-24 pulses, 10-60 Hz, 10 µA-10 mA).

Optogenetic stimulation equipment. Blue light was applied through computer control

of a 3-W Luxeon Star LED with a peak intensity of ~470 nm (LXHL-LB3C, Newark,

Chicago, IL). Red light was applied in a similar manner with a 1-W Luxeon Star LED

with a peak intensity of ~625 nm (LXHL-MD1B, Newark).

Results and Discussion

Chemical stimulation of dopamine release in Drosophila. In the fly brain, dopamine

neurons project to the nearby mushroom body (MB) structure which is crucial for many

higher-order functions including learning and memory (55, 56). The neuronal cluster in

the protocerebral anterior medial (PAM) brain area is the largest group of dopamine

neurons in the Drosophila CNS (57), and this is the region where I placed the working

electrode in dopamine uptake experiments reported in Chapters 2-4. Dopamine release

occurs in the region where the dopamine neurons project; therefore, the working

electrode was placed in the MB in this chapter.

After the microsurgery procedure was performed (see Methods), a

micromanipulator was used to guide a 5 µm diameter cylindrical carbon-fiber electrode

into the MB area. FSCV was used to measure changes in current in this brain region, and

112

Page 129: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

the reference electrode was submerged in the surrounding AHL saline. Micropipets were

used to deliver chemical stimulation solutions to the MB area with pneumatic pressure

(Figure 5.2A). The chemicals investigated as potential stimulants in Drosophila included

potassium, veratridine, caffeine, nicotine, and barium. Although all five of the solutions

successfully stimulate dopamine release through a range of mechanisms in mammalian

model systems (Table 5.1), dopamine release was not detected in Drosophila.

Various experimental parameters could account for the data obtained with the

chemical stimulation method. One, diffusion causes a decrease in the concentration of

the chemical stimulant that reaches the brain region surrounding the working electrode

compared with the original concentration in the micropipet injector. However, the

decrease in concentration can be approximated by the micropipet injection of dopamine

described in Chapter 2 (Figure 2.4). The concentration that diffuses into the tissue

depends on the diffusion rate, relative permeability into the tissue, and size of a particular

chemical species. Here, the limiting factor is the high resistance to diffusion of the brain

tissue. Pneumatically applying 1.0 mM dopamine for 1.0 s just above the fly brain results

in a concentration of ~7 µM in the Drosophila tissue. Therefore, the concentration of the

stimulant in the fly brain region is approximately three orders of magnitude lower than

that of the applied solution.

Another possibility is that invertebrate systems respond to the chemical stimulants

differently than mammals, but it is unlikely that the solutions tested would not stimulate

the fly CNS and initiate dopamine release. Released dopamine dissipates from the

extracellular space via uptake by the dopamine transporter (IC50 = 2.9 µM for

Drosophila), metabolism, and diffusion (58). It seems likely that released dopamine is

113

Page 130: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 5.2. Schematic comparing three methods for stimulating neurotransmitter release

in adult Drosophila. The position of the instruments used for chemical, electrical, and

optogenetic stimulation is marked with respect to the cylindrical working electrode (not

drawn to scale).

114

Page 131: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

diluted by the surrounding chemical stimulation solution in the small volume of the fly

brain, thus the concentration of dopamine at the electrode surface is lower than the limit

of detection (LOD) of the working electrode (65 nM). The dopamine concentration

inside vesicles is well above the LOD (~100 mM); however, vesicles are attoliters in

volume. A 1 aL vesicle corresponds to ~0.1 amol of dopamine. Diffusion from the aL

vesicle to the nL surrounding environment results in a decrease of 10-9

. This dilution in

concentration will affect measurements in the ~5 nL fly brain more significantly than a

larger mammalian system, like the rat brain, and suggests an explanation as to why this

method of stimulation was successful in other model systems, but not in the fly.

Electrical stimulation of dopamine release in Drosophila. Electrical stimulation is a

method that has been extensively used to elicit neurotransmitter release in mammalian

systems. Two electrodes are placed on either side of a neuronal pathway of interest (36).

Through application of a voltage, action potentials in neurons are initiated which results

in neurotransmitter release. The dimensions of the fly CNS required modification of

electrical stimulation procedures used in mammalian systems. Following the

microsurgery procedure, I placed two stimulation electrodes on either side of the MB

structure where the cylindrical working electrode was positioned with the tips of all three

electrodes in the same horizontal plane (Figure 5.2B). The working electrode was ~50

µm from each stimulation electrode, and the reference electrode was submerged in the

surrounding AHL saline outside of the fly head. Electrical pulses were applied through

the stimulation electrodes, and a wide range of values for the pulse parameters was tested

(see Methods). In vivo fluctuations in current were measured in the Drosophila CNS

with FSCV detection; however, after inspection of the voltammograms, the current

115

Page 132: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

changes were not attributed to dopamine release. This could be due to several factors.

The parameters used for electrical stimulation must be sufficient to depolarize the

dopamine neurons while not affecting neurotransmission of the entire fly CNS.

Additional neurochemical changes in the system from overstimulation might coincide

with the time scale of dopamine release, which could alter the expected electrochemical

voltammetric signature of dopamine.

Two types of commercially available stimulation electrodes were used for this

investigation: platinum stimulation electrodes 75 µm in diameter and tungsten electrodes

125 µm in diameter but tapered to a sharp point. The size of a fly MB is ~100 µm in

width, and the electrical stimulation electrodes cause damage to the area where they are

placed. A potential solution to alleviating some of the physical destruction caused in the

CNS by the stimulation electrodes is to build an electrical stimulation set-up with smaller

electrodes that have been fabricated by hand. Another approach is to use optogenetic

stimulation of neurons, which eliminates the need to place any stimulation electrodes in

or around the fly brain.

Optogenetic stimulation of dopamine release in Drosophila. Endogenous dopamine

release was evoked in TH-GAL4/UAS:ChR2 mutant flies using optogenetic stimulation.

The dopamine neurons of the mutant flies were genetically altered to express ChR2, a

cation-selective ion channel that can be activated with blue light on a millisecond time

scale (59, 60). The ChR2 protein contains the chromophore all-trans retinal, which

undergoes a conformation change upon exposure to blue light. This causes the ChR2 ion

channel to open and allows Na+ ions to enter the cell. The increased positive charge in

116

Page 133: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

the cell depolarizes the cell membrane and leads to vesicular dopamine release into the

extracellular space (Figure 5.1).

Following fly microsurgery, a blue LED was positioned ~1.5 mm above the

exposed Drosophila MB (Figure 5.2C). FSCV was used to monitor neurochemical

release. A TH-GAL4/UAS:ChR2 mutant fly was illuminated with blue light for 10 s to

stimulate dopamine release. As a control, the experiment was repeated with TH-

GAL4/UAS:ChR2 mutant flies that had not consumed all-trans retinal which is necessary

for ChR2 ion channel function to occur. This is an effective way to eliminate the

response of the ChR2 channel to blue light in the Drosophila CNS system (49, 61).

The results suggest an electroactive species is released in the MB region of

Drosophila with optical stimulation. Figure 5.3A compares the current measured during

blue light stimulation of a mutant fly that consumed all-trans retinal (black line) with a

mutant fly that did not consume all-trans retinal (gray line). While a significant

difference in the two flies is observed, the ~0.07 nA increase in measured current from

the fly that did not consume all-trans retinal is not anticipated. By inspection, the cyclic

voltammogram for this current (Figure 5.3B, gray line) does not resemble a typical wave

shape of dopamine. The non-dopamine like voltammogram confirms that dopamine is

not released in the TH-GAL4/UAS:ChR2 mutant fly that did not consume all-trans

retinal. There is apparently a contribution to the measured signal from factors other than

dopamine oxidation. The voltammogram of the peak signal from the fly that consumed

all-trans retinal (black line) is similar to the shape of a dopamine voltammogram,

suggesting that dopamine is released and measured. However, the formal potential of the

voltammogram is shifted +200 mV compared to a representative voltammogram of

117

Page 134: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 5.3. Effect of blue light stimulation on flies with genetically altered dopamine

neurons. (A) Representative current trace in a TH-GAL4/UAS:ChR2 mutant fly that

consumed all-trans retinal (black line) vs. a mutant fly that did not consume all-trans

retinal (gray line). The black arrow corresponds to a 10-s stimulation with blue light

beginning at 5 s. (B) Background-subtracted fast-scan cyclic voltammograms (average of

2 scans each, 200 V/s) corresponding to the measured peak current during blue light

simulation of a mutant fly that consumed all-trans retinal (black line) and a mutant fly

that did not consume all-trans retinal (gray line). (C) Representative background-

subtracted fast-scan cyclic voltammogram of exogenously applied dopamine measured in

the fly CNS for comparison (average of 5 scans, 200 V/s).

118

Page 135: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

exogenously applied dopamine that has been measured in the fly CNS (Figure 5.3C), thus

necessitating further characterization of the signal. It is possible that another, as yet

unidentified, electroactive compound is released in this brain region.

The possibility of electrical interference from the direct exposure of the fly to a

LED light was examined. A TH-GAL4/UAS:ChR2 mutant fly was exposed to a 10-s

blue light stimulation then 15 min later a 10-s red light stimulation as a control as it has

been shown it does not stimulate ChR2 function in mutant Drosophila larvae (49, 50).

Flies were also tested with the red light first followed by the blue light to ensure the order

of exposure to the two wavelengths of light did not alter the response of the fly. The

cylindrical electrode remained untouched between light stimulations. Figure 5.4

compares two voltammograms obtained in a TH-GAL4/UAS:ChR2 mutant fly that

consumed all-trans retinal. The voltammogram observed following blue light stimulation

(blue line) resembles the wave shape of dopamine, but again the shift in formal potential

is evident. Red light stimulation (red line) does not cause any significant change in the

measured current. Thus the shift in formal potential and the shape of the voltammogram

appear to correspond to dopamine release in the fly and are not due to an electrical

alteration caused from LED illumination.

Of note, two peak currents were measured in a TH-GAL4/UAS:ChR2 mutant fly

that had consumed all-trans retinal which were approximately three times higher than the

typical peak current recorded for the mutant flies (Figure 5.5A). The two increases in

current appear to correspond to spontaneous release as they did not occur during the

stimulation time period with blue light. A 10-s blue light stimulation was applied starting

at 5 s; however, the peaks were recorded 55 s and 105 s later. The height of the two

119

Page 136: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 5.4. Voltammograms obtained during blue and red light stimulation of a TH-

GAL4/UAS:ChR2 mutant fly. Background-subtracted fast-scan cyclic voltammograms

(average of 2 scans each, 200 V/s) in the MB region of a TH-GAL4/UAS:ChR2 mutant

fly that consumed all-trans retinal. A trace obtained during the measured peak current

from a 10-s blue light simulation (blue line) was compared with a trace obtained during a

10-s red light stimulation (red line).

120

Page 137: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

peaks in Figure 5.5A corresponds to ~725 nM and ~945 nM dopamine if compared to in

vitro calibration of the electrode (Appendix). Figure 5.5B is a background-subtracted

cyclic voltammogram from the maximum current measured for the second peak in Figure

5.5A. By inspection, the anodic peak shape indicates dopamine is oxidized at the

electrode surface.

Further characterization of the signal evoked with optogenetic stimulation in

Drosophila is necessary before the measured current can be convincingly attributed to

dopamine. Cyclic voltammograms suggest an aspect of the signal is due to dopamine

oxidation. Because the working electrode was placed in the MB brain region, signal from

oxidation of octopamine, serotonin, and histamine which are electroactive species present

mainly in other regions of the Drosophila brain, is unlikely (62). A possible reason for

the +200 mV shift in formal potential of the voltammograms shown in this chapter is the

difference in placement of the cylindrical electrode in the fly brain. In Chapters 2-4, the

microelectrode was inserted into a cluster of dopamine neurons in the PAM area because

the uptake of applied dopamine was being quantified. Dopamine neurons in the PAM

area project to the MB region of the fly brain, meaning the release of endogenous

dopamine occurs here (55). While the PAM and MB regions are just microns apart, the

density of the two brain regions is slightly different. The measured signal at the working

electrode is low, and the difference in density of the surrounding tissue might cause an

observable effect on the measured current.

121

Page 138: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 5.5. Spontaneous release of an electroactive species from a TH-

GAL4/UAS:ChR2 mutant fly. (A) Current trace from a mutant fly showing two peaks

that do not occur during the blue light stimulation time period. The black arrow

corresponds to a 10-s stimulation with blue light beginning at 5 s. (B) Background-

subtracted fast-scan cyclic voltammogram (average of 2 scans, 200 V/s) from the

maximum measured current of the second peak shown in (A). The wave shape resembles

that of dopamine.

122

Page 139: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Conclusions

Chemical, electrical, and optogenetic methods to induce endogenous dopamine

release in the Drosophila CNS were studied. FSCV detection with a 5-µm carbon-fiber

microelectrode was used to monitor dopamine changes. Methods successfully used in the

rat brain for chemical and electrical stimulation of neurotransmitter release were

modified for the nanoliter-sized fly CNS. ChR2, an emerging optogenetic tool for

controlling neuronal release with blue light, was used to genetically target stimulation of

dopaminergic neurons in Drosophila. Results suggest optogenetic stimulation is a useful

and noninvasive technique for eliciting dopamine release in the fly CNS. Future

investigation of Drosophila mutants genetically altered with the ChR2 ion channel could

lead to progression in the novel field of optogenetic neuronal stimulation.

123

Page 140: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Clark, L. C., and Lyons, C. (1965) Studies of a glassy carbon electrode for brain

polarography with observations on the effect of carbonic andydrase inhibition,

Ala. J. Med. Sci. 2, 353-359.

2. Kissinger, P. T., Hart, J. B., and Adams, R. N. (1973) Voltammetry in brain

tissue-a new neurophysiological measurement, Brain Res. 55, 209-213.

3. Adams, R. N. (1976) Probing brain chemistry with electroanalytical techniques,

Anal. Chem. 48, 1126A-1138A.

4. Dayton, M. A., Brown, J. C., Stutts, K. J., and Wightman, R. M. (1980) Faradaic

electrochemistry at micro-voltammetric electrodes, Anal. Chem. 52, 946-950.

5. Wipf, D. O., Kristensen, E. W., Deakin, M. R., and Wightman, R. M. (1988) Fast-

scan cyclic voltammetry as a method to measure rapid, heterogenous electron-

transfer kinetics, Anal. Chem. 60, 306-310.

6. Wipf, D. O., and Wightman, R. M. (1988) Submicrosecond measurements with

cyclic voltammetry, Anal. Chem. 60, 2460-2464.

7. Salamone, J. D. (1996) The behavioral neurochemistry of motivation:

Methodological and conceptual issues in studies of the dynamic activity of

nucleus accumbens dopamine, J. Neurosci. Methods 64, 137-149.

8. Dale, N., Hatz, S., Tian, F., and Llaudet, E. (2005) Listening to the brain:

microelectrode biosensors for neurochemicals, Trends Biotechnol. 23, 420-428.

9. Fillenz, M. (2005) In vivo neurochemical monitoring and the study of behavior,

Neurosci. Biobehav. Rev. 29, 949-962.

10. Robinson, D. L., Hermans, A., Seipel, A. T., and Wightman, R. M. (2008)

Monitoring rapid chemical communication in the brain, Chem. Rev. 108, 2554-

2584.

11. Wilson, G. S., and Johnson, M. A. (2008) In vivo electrochemistry: What can we

learn about living systems?, Chem. Rev. 108, 2462-2481.

12. Clark, R. A., Zerby, S. E., and Ewing, A. G. (1996) Electrochemistry in neuronal

microenvironments, in Electroanalytical Chemistry: A Series of Advances (Bard,

A. J., and Rubinstein, I., Eds.), pp 227-295, Marcel Dekker, New York.

13. Kuhr, W. G., and Wightman, R. M. (1986) Real-time measurement of dopamine

release in rat brain, Brain Res. 381, 168-171.

14. Baur, J. E., Kristensen, E. W., May, L. J., Wiedemann, D. J., and Wightman, R.

M. (1988) Fast-scan voltammetry of biogenic amines, Anal. Chem. 60, 1268-

1272.

15. Heien, M. L., Khan, A. S., Ariansen, J. L., Cheer, J. F., Phillips, P. E. M.,

Wassum, K. M., and Wightman, R. M. (2005) Real-time measurement of

dopamine fluctuations after cocaine in the brain of behaving rats, Proc. Natl.

Acad. Sci. U.S.A. 102, 10023-10028.

16. Ritz, M. C., Lamb, R. J., Goldberg, S. R., and Kuhar, M. J. (1987) Cocaine

receptors on dopamine transporters are related to self-administration of cocaine,

Science 237, 1219-1223.

17. Kuhar, M. J., Ritz, M. C., and Boja, J. W. (1991) The dopamine hypothesis of the

reinforcing properties of cocaine, Trends Neurosci. 14, 299-302.

124

Page 141: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

18. Giros, B., Jaber, M., Jones, S. R., Wightman, R. M., and Caron, M. G. (1996)

Hyperlocomotion and indifference to cocaine and amphetamine in mice lacking

the dopamine transporter, Nature 379, 606-612.

19. Leszczyszyn, D. J., Jankowski, J. A., Viveros, O. H., Jr., E. J. D., Near, J. A., and

Wightman, R. M. (1991) Secretion of catecholamines from individual adrenal

medullary chromaffin cells, J. Neurochem. 56, 1855-1863.

20. Wightman, R. M., Jankowski, J. A., Kennedy, R. T., Kawagoe, K. T., Schroeder,

T. J., Leszczyszyn, D. J., Near, J. A., Diliberto, E. J., and Viveros, O. H. (1991)

Temporally resolved catecholamine spikes correspond to single vesicle release

from individual chromaffin cells, Proc. Natl. Acad. Sci. U.S.A. 88, 10754-10758.

21. Ewing, A. G., Wightman, R. M., and Dayton, M. A. (1982) In vivo voltammetry

with electrodes that discriminate between dopamine and ascorbate, Brain Res.

249, 361-370.

22. Gerhardt, G. A., Rose, G. M., and Hoffer, B. J. (1986) Release of monoamines

from striatum of rat and mouse evoked by local application of potassium:

Evaluation of a new in vivo electrochemical technique, J. Neurochem. 46, 842-

850.

23. Kant, G. J., and Meyerhoff, J. L. (1978) Release of endogenous norepinephrine

and dopamine from rat brain regions in vitro, Life Sci. 23, 2111-2117.

24. Ahmad, S., Fowler, L. J., and Whitton, P. S. (2005) Effects of combined

lamotrigine and valproate on basal and stimulated extracellular amino acids and

monoamines in the hippocampus of freely moving rats, Naunyn-Schmiedeberg's

Arch. Pharmacol. 371, 1-8.

25. Wightman, R. M., Bright, C. E., and Caviness, J. N. (1981) Direct measurement

of the effect of potassium, calcium, veratridine, and amphetamine on the rate of

release of dopamine from superfused brain tissue, Life Sci. 28, 1279-1286.

26. Halmos, G., Gáborján, A., Lendvai, B., Répássy, G., Szabó, L. Z., and Vizi, E. S.

(2000) Veratridine-evoked release of dopamine from guinea pig isolated cochlea,

Hear. Res. 144, 89-96.

27. Okada, M., Kiryu, K., Kawata, Y., Mizuno, K., Wada, K., Tasaki, H., and

Kaneko, S. (1997) Determination of the effects of caffeine and carbamazepine on

striatal dopamine release by in vivo microdialysis, Eur. J. Pharmacol. 321, 181-

188.

28. Solinas, M., Ferre, S., You, Z.-B., Karcz-Kubicha, M., Popoli, P., and Goldberg,

S. R. (2002) Caffeine induces dopamine and glutamate release in the shell of the

nucleus accumbens, J. Neurosci. 22, 6321-6324.

29. Imperato, A., Mulas, A., and Di Chiara, G. (1986) Nicotine preferentially

stimulates dopamine release in the limbic system of freely moving rats, Eur. J.

Pharmacol. 132, 337-338.

30. Mifsud, J.-C., Hernandez, L., and Hoebel, B. G. (1989) Nicotine infused into the

nucleus accumbens increases synaptic dopamine as measured by in vivo

microdialysis, Brain Res. 478, 365-367.

31. Toth, E., Sershen, H., Hashim, A., Vizi, E. S., and Lajtha, A. (1992) Effect of

nicotine on extracellular levels of neurotransmitters assessed by microdialysis in

various brain regions: Role of glutamic acid, Neurochem. Res. 17, 265-271.

125

Page 142: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

32. Izumi, F., Toyohira, Y., Yanagihara, N., Wada, A., and Kobayashi, H. (1986)

Barium-evoked release of catecholamines from digitonin-permeabilized adrenal

medullary cells, Neurosci. Lett. 69, 172-175.

33. Taglialatela, M., Canzoniero, L. M. T., Amoroso, S., Fatatis, A., Di Renzo, G. F.,

and Annunziato, L. (1989) Cobalt-sensitive and dihydropyridine-insensitive

stimulation of dopamine release from tuberoinfundibular neurons by high

extracellular concentrations of barium ions, Brain Res. 488, 114-120.

34. Portig, P. J., and Vogt, M. (1969) Release into the cerebral ventricles of

substances with possible transmitter function in the caudate nucleus, J. Physiol.

(Lond.) 204, 687-715.

35. Wightman, R. M., Strope, E., Plotsky, P. M., and Adams, R. N. (1976)

Monitoring of transmitter metabolites by voltammetry in cerebrospinal fluid

following neural pathway stimulation, Nature 262, 145-146.

36. Ewing, A. G., Bigelow, J. C., and Wightman, R. M. (1983) Direct in vivo

monitoring of dopamine released from 2 striatal compartments in the rat, Science

221, 169-171.

37. Kuhr, W. G., Ewing, A. G., Caudill, W. L., and Wightman, R. M. (1984)

Monitoring the stimulated release of dopamine with in vivo voltammetry. I:

Characterization of the response observed in the caudate nucleus of the rat, J.

Neurochem. 43, 560-569.

38. Taber, M. T., and Fibiger, H. C. (1995) Electrical stimulation of the prefrontal

cortex increases dopamine release in the nucleus accumbens of the rat:

Modulation by metabotropic glutamate receptors, J. Neurosci. 15, 3896-3904.

39. Chergui, K., Nomikos, G. G., Mathe, J. M., Gonon, F., and Svensson, T. H.

(1996) Burst stimulation of the medial forebrain bundle selectively increases Fos-

like immunoreactivity in the limbic forebrain of the rat, Neuroscience 72, 141-

156.

40. Garris, P. A., Christensen, J. R. C., Rebec, G. V., and Wightman, R. M. (1997)

Real-time measurement of electrically evoked extracellular dopamine in the

striatum of freely moving rats, J. Neurochem. 68, 152-161.

41. Kulagina, N. V., Zigmond, M. J., and Michael, A. C. (2001) Glutamate regulates

the spontaneous and evoked release of dopamine in the rat striatum, Neuroscience

102, 121-128.

42. West, A. R., and Grace, A. A. (2002) Opposite influences of endogenous

dopamine D-1 and D-2 receptor activation on activity states and

electrophysiological properties of striatal neurons: Studies combining in vivo

intracellular recordings and reverse microdialysis, J. Neurosci. 22, 294-304.

43. Robinson, D. L., Venton, B. J., Heien, M. L., and Wightman, R. M. (2003)

Detecting subsecond dopamine release with fast-scan cyclic voltammetry in vivo,

Clin. Chem. 49, 1763-1773.

44. Forster, G. L., and Blaha, C. D. (2003) Pedunculopontine tegmental stimulation

evokes striatal dopamine efflux by activation of acetylcholine and glutamate

receptors in the midbrain and pons of the rat, Eur. J. Neurosci. 17, 751-762.

126

Page 143: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

45. Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., Ollig,

D., Hegemann, P., and Bamberg, E. (2003) Channelrhodopsin-2, a directly light-

gated cation-selective membrane channel, Proc. Natl. Acad. Sci. U.S.A. 100,

13940-13945.

46. Scanziani, M., and Hausser, M. (2009) Electrophysiology in the age of light,

Nature 461, 930-939.

47. Hegemann, P., Gartner, W., and Uhl, R. (1991) All-trans retinal constitutes the

functional chromophore in Chlamydomonas rhodopsin, 60, 1477-1489.

48. Takahashi, T., Yoshihara, K., Watanabe, M., Kubota, M., Johnson, R., Derguini,

F., and Nakanishi, K. (1991) Photoisomerization of retinal at 13-ene is important

for phototaxis of Chlamydomonas reinhardtii: Simultaneous measurements of

phototactic and photophobic responses, Biochem. Biophys. Res. Commun. 178,

1273-1279.

49. Schroll, C., Riemensperger, T., Bucher, D., Ehmer, J., Völler, T., Erbguth, K.,

Gerber, B., Hendel, T., Nagel, G., Buchner, E., and Fiala, A. (2006) Light-

induced activation of distinct modulatory neurons triggers appetitive or aversive

learning in Drosophila larvae, Curr. Biol. 16, 1741-1747.

50. Zhang, W., Ge, W., and Wang, Z. (2007) A toolbox for light control of

Drosophila behaviors through Channelrhodopsin-2 mediated photoactivation of

targeted neurons, Eur. J. Neurosci. 26, 2405-2416.

51. Vickrey, T. L., Condron, B., and Venton, B. J. (2009) Detection of endogenous

dopamine changes in Drosophila melanogaster using fast-scan cyclic

voltammetry, Anal. Chem. 81, 9306-9313.

52. Duffy, J. B. (2002) GAL4 system in Drosophila: A fly geneticist's Swiss army

knife, Genesis 34, 1-15.

53. Makos, M. A., Kim, Y.-C., Han, K.-A., Heien, M. L., and Ewing, A. G. (2009) In

vivo electrochemical measurements of exogenously applied dopamine in

Drosophila melanogaster, Anal. Chem. 81, 1848-1854.

54. Heien, M. L., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman, R.

M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

55. Nassel, D. R., and Elekes, K. (1992) Aminergic neurons in the brain of blowflies

and Drosophila: dopamine- and tyrosine hydroxylase-immunoreactive neurons

and their relationship with putative histaminergic neurons, Cell Tissue Res. 267,

147-167.

56. Kim, Y.-C., Lee, H.-G., and Han, K.-A. (2007) D1 dopamine receptor dDA1 is

required in the mushroom body neurons for aversive and appetitive learning in

Drosophila, J. Neurosci. 27, 7640-7647.

57. Coulom, H., and Birman, S. (2004) Chronic exposure to rotenone models sporadic

Parkinson's disease in Drosophila melanogaster, J. Neurosci. 24, 10993-10998.

58. Porzgen, P., Park, S. K., Hirsh, J., Sonders, M. S., and Amara, S. G. (2001) The

antidepressant-sensitive dopamine transporter in Drosophila melanogaster: A

primordial carrier for catecholamines, Mol. Pharmacol. 59, 83-95.

59. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G., and Deisseroth, K. (2005)

Millisecond-timescale, genetically targeted optical control of neural activity, Nat.

Neurosci. 8, 1263-1268.

127

Page 144: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

60. Li, X., Gutierrez, D. V., Hanson, M. G., Han, J., Mark, M. D., Chiel, H.,

Hegemann, P., Landmesser, L. T., and Herlitze, S. (2005) Fast noninvasive

activation and inhibition of neural and network activity by vertebrate rhodopsin

and green algae channelrhodopsin, Proc. Natl. Acad. Sci. U.S.A. 102, 17816-

17821.

61. Suh, G. S. B., Ben-Tabou de Leon, S., Tanimoto, H., Fiala, A., Benzer, S., and

Anderson, D. J. (2007) Light activation of an innate olfactory avoidance response

in Drosophila, Curr. Biol. 17, 905-908.

62. Monastirioti, M. (1999) Biogenic amine systems in the fruit fly Drosophila

melanogaster, Microsc. Res. Tech. 45, 106-121.

128

Page 145: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 6: Development and Characterization of a Voltammetric

Carbon-Fiber Microelectrode pH Sensor*

*Reproduced with permission from Makos, M. A., Omiatek, D. M., Ewing, A. G., and

Heien, M. L. (2010) Development and Characterization of a Voltammetric Carbon-Fiber

Microelectrode pH Sensor, Langmuir, accepted. © 2010 American Chemical Society.

129

Page 146: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Introduction

Recently, there has been an interest in developing reagentless sensors to detect

small pH changes in non-ideal environments (1). Carbon-based sensing materials are

attractive substrates for this application since they are intrinsically biocompatible,

conductive, and apt for surface modification. Indeed, ion-selective reporter molecules

can be tethered onto a carbon surface through a variety of methods including chemical

oxidation of the surface with corrosive acidic solutions and plasma treatment (2, 3),

physical adsorption of organic precursors (4, 5), and electrochemically-assisted covalent

attachment via the oxidation of amines (6-9) and reduction of diazonium salts (10-16).

Pioneered by Savéant and co-workers in the early 1990s, the reduction of aryl diazonium

salts onto carbon surfaces is a well-characterized method for the selective in situ

attachment of organic molecules (10). This mechanism involves the electrochemical

generation of a solution radical from the diazonium modifier and subsequent covalent

linkage to the carbon surface, which possesses marked stability to external stimuli (13).

Electrochemical measurements in the central nervous system (CNS) can quantify

redox-active chemical messengers such as catecholamines and indolamines, which are

thought to play a fundamental role in the physiological and behavioral aspects of

organisms. In vivo voltammetry with carbon-fiber microelectrodes has been used for

several decades to monitor neurotransmission of these chemicals in the CNS of various

mammalian animal models (17-19). Neurosecretory events are often accompanied by a

flux of endogenous species (e.g., H+, ascorbate) which can interfere with the

voltammetric signature of the electroactive chemical species of interest (5, 20-26). Of

particular interest are pH fluctuations in the surrounding matrix, which are thought to

130

Page 147: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

occur as a result of metabolic processes that follow stimulated neurotransmitter release

(23, 27-29). Wightman and co-workers have reported measuring these small acidic pH

changes in rat brain slices subjected to electrically stimulated secretion with liquid

membrane, ion-selective microelectrodes (ISMs) (23). With the emergence of volume-

limited, CNS-containing animal models such as the fruit fly, Drosophila melanogaster,

comes the need to develop microanalytical tools capable of measuring the pH fluctuations

associated with neurotransmission (30). I have discussed using fast-scan cyclic

voltammetry (FSCV) for quantifying in vivo neurotransmitters in the CNS of Drosophila

in Chapters 2-5 (31, 32); however, the electrode could not be used to measure

fluctuations in the pH of the brain.

Voltammetric pH sensors measure changes in the redox-potential of a surface-

bound, electrochemically active species as a function of pH. This methodology for

measuring pH has been demonstrated with quinone-based surface modifications of

various electrodes (33-36). In a recent study by Tommos and co-workers, the formal

potential of a surface-bound quinone on a gold electrode shifted to more negative

potentials with increasing solvent basicity (35). While a variety of quinone-modified

electrodes have been reported to respond to pH, few have been developed on

biocompatible materials that exhibit activity in a physiologically relevant pH range (1,

37). In this chapter, I will describe a procedure for electrochemically grafting Fast Blue

RR (FBRR) salt, a quinone-containing diazonium derivative, to a cylindrical carbon-fiber

microelectrode. This results in a microelectrode capable of performing real-time,

reagentless pH measurements in biological microenvironments. The redox response of

the FBRR-functionalized electrode is characterized using FSCV in biological media set

131

Page 148: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

to a physiologically relevant pH range. I demonstrate that modification of a carbon-fiber

surface with FBRR is a simple and reproducible method for fabricating a stable pH

sensor that is sensitive enough to measure dynamic physiological pH changes in the CNS

of Drosophila that are associated with stimulated neurotransmitter release.

Methods

Chemicals. All chemicals were used as received unless otherwise stated. 4-

Benzoylamino-2,5-dimethoxybenzenediazonium chloride hemi(zinc chloride) salt (Fast

Blue RR, FBRR, diazonium salt), tetraethylammonium tetrafluoroborate (TEABF4),

NaCl, KCl, CaCl2, MgCl2, NaHCO3, NaH2PO4, sucrose, Trizma base , and acetonitrile

(ACN, anhydrous, 99.8%) were obtained from Sigma Aldrich (St. Louis, MO). Adult-

hemolymph like (AHL) saline (108 mM NaCl, 5 mM KCl, 2 mM CaCl2, 8.2 mM MgCl2,

4 mM NaHCO3, 1 mM NaH2PO4, 5 mM trehalose (Fluka BioChemika, Buchs,

Switzerland), 10 mM sucrose, 20 mM Trizma base , pH 7.5) was made using ultrapure

(18 MΩ cm) water and filtered through a 0.2-μm filter (38). The pH of AHL solutions

was adjusted with 0.5 M NaOH and HCl.

Electrode preparation. Cylindrical carbon-fiber microelectrodes were fabricated as

described in Chapter 2 (31). Briefly, a 5-μm diameter carbon fiber (T-40 12K, Amoco,

Greenville, SC) was aspirated into a borosilicate glass capillary (1B100-4, World

Precision Instruments, Inc., Sarasota, FL) and sealed using a regular glass capillary puller

(P-97, Sutter Instruments, Novato, CA). The carbon fiber was trimmed to a length of

either 50 or 200 μm measured from the glass junction. Electrical contact was made by

back-filling the capillary with a silver composition (4922N DuPont, Delta Technologies

132

Page 149: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Ltd., Stillwater, MN), followed by insertion of a tungsten wire, resulting in a 5 µm

diameter cylindrical carbon-fiber microelectrode. The 200-µm long cylindrical

electrodes were used for all characterization experiments while the 50-µm long electrodes

were used for the in vivo Drosophila application.

Chemical modification of the carbon-fiber microelectrode surface. FBRR salt was

electrochemically grafted onto the carbon-fiber microelectrode surface using diazonium

attachment chemistry. Deposition of the diazonium salt onto the carbon-fiber

microelectrodes was carried out using cyclic voltammetry performed with an Ensman

Instruments EI400 microelectrode potentiostat (Bloomington, IN) operated in the two-

electrode mode. A 2 mM solution of FBRR salt was prepared in ACN containing 0.1 M

TEABF4. Solutions were purged with Ar (g) for 5 min prior to deposition in order to

eliminate signal from the reduction of O2. Electrodes were electrochemically modified

via reduction of the diazonium onto the carbon surface by scanning from +0.4 V to -0.8 V

vs. Ag QRE (3 mm diameter, Bioanalytical Systems, West Lafayette, IN) at 0.5 V/s.

Data were collected and processed using LabView 8.0 software (National Instruments,

Austin, TX) written in-house. Electrode surface coverage was calculated by subtracting

the background current measured for a solution of 0.1 M TEABF4 in ACN from that due

to deposition of the diazonium. The typical surface coverage obtained using the

experimental conditions listed above for a microelectrode with a 200-µm long carbon

fiber was ~20 nmol/cm2.

Electrochemical measurements. Voltammetric responses of the diazonium-modified

electrodes as a function of pH were collected using either a Dagan Chem-Clamp

potentiostat (Dagan Corporation, Minneapolis, MN) or a flow-injection analysis

133

Page 150: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

apparatus with a current amplifier (428, Keithley Instruments, Inc., Cleveland, OH).

Both systems were run by the TH 1.0 CV program (ESA, Chelmsford, MA) (39) coupled

with two data acquisition boards (PCI-6221, National Instruments). A Ag/AgCl

electrode, which served as the reference in all experiments following the initial

deposition of FBRR, was made by chloridizing a silver wire (0.25 mm diameter,

99.999% purity, Alfa Aesar, Ward Hill, MA). Electrodes were positioned using x,y,z-

micromanipulators (421 series, Newport, Irvine, CA). All cyclic voltammograms were

obtained using a triangular waveform scanned from -0.7 to +0.8 V vs. Ag/AgCl at 20 V/s

and repeated every 200 ms unless otherwise noted. Electrochemical responses were

plotted and statistical analysis performed using Prism 5.0 (GraphPad Software, La Jolla,

CA). Anodic peak potentials (Epa) were determined using a fifth order polynomial fit

from LabView 8.0 software written in-house. Cyclic voltammetry was used to estimate

the heterogeneous electron-transfer rate constant, k0, for this system via the method of

Nicholson (40).

In vivo Drosophila preparation. As in Chapter 5, female flies carrying

Channelrhodopsin-2 (ChR2), a light activated ion channel, were crossed with male flies

expressing tyrosine hydroxylase (TH) to produce mutant flies containing dopaminergic

neurons that can be controlled through blue light stimulation (TH-GAL4/UAS:ChR2

genotype) (41). Male mutant flies, 3-7 days old, were maintained at 25 °C in the dark

and fed yeast containing 10 mM all-trans retinal (light sensitive chemical necessary for

ChR2 function) for 2 days prior to experimentation. Blue light was applied through

computer control of a 3-W Luxeon Star LED with a peak intensity of ~470 nm (LXHL-

LB3C, Newark, Chicago, IL). Flies were prepared as described in Chapter 2 for in vivo

134

Page 151: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

FSCV measurements (31). Briefly, ice was used to temporarily immobilize flies before

they were mounted in a homemade collar (38.1 mm diameter concave plexiglass disk

with a 1.0 mm hole in the center) with low melting agarose (Fisher Scientific, Pittsburgh,

PA). Microsurgery was performed on a stereoscope (Olympus SZ60, Melville, NY) to

remove the cuticle from the top portion of the head, thus exposing the brain region. The

head was covered with 0.1% collagenase solution for 30 min to relax the extracellular

matrix in the brain then rinsed and bathed with AHL saline with the preparation

maintaining its viability for 1.5 - 2.5 h.

Results and Discussion

Deposition of FBRR diazonium salt onto a carbon-fiber microelectrode surface.

FBRR was electrochemically reduced onto a carbon-fiber surface using cyclic

voltammetry by scanning from +0.4 V to -0.8 V vs. Ag QRE at a rate of 0.5 V/s in a 2 M

FBRR/0.1 M TEABF4/ACN solution. The proposed mechanism for this deposition is

presented in Scheme 6.1. A representative voltammogram of the diazonium salt

reduction onto a cylindrical carbon-fiber microelectrode is shown in Figure 6.1A (blue

trace). An irreversible reductive wave is observed around -0.5 V which is attributed to

the solution radical formation of the diazonium derivative and its subsequent covalent

linkage to the carbon-fiber surface as reported for a similar molecule (12).

The charge (Q) of the diazonium deposited onto the surface is quantified using the

current-time integral of the voltammetric trace. The slight charge observed from the

solvent background (black trace) has been subtracted from the charge due to diazonium

deposition (blue trace). Faraday’s Law (Q = nNF) is used to convert Q to the number of

135

Page 152: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Scheme 6.1. Electrochemical deposition of FBRR salt onto the carbon-fiber

microelectrode surface.

136

Page 153: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.1. Cyclic voltammograms of a carbon-fiber microelectrode before and after

FBRR attachment. (A) Background charge of the carbon-fiber electrode in solvent only

(black line). Electrochemical reduction of FBRR on the carbon-fiber surface (blue line).

Diazonium concentration = 2 mM in 0.1 M TEABF4/ACN. The potential is scanned +0.4

V to -0.8 V vs. Ag QRE at 0.5 V/s. (B) Cyclic voltammograms (average of 5 scans each)

in AHL saline of a bare carbon-fiber microelectrode (dashed black line) and the same

carbon-fiber microelectrode after modification with FBRR (solid blue line). The

potential is scanned -0.7 V to +0.8 V vs. Ag/AgCl at 20 V/s.

137

Page 154: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

moles of diazonium (N) deposited onto the carbon-fiber surface. In this equation the

number of electrons exchanged in the reduction reaction, n, is 1, and F is Faraday’s

constant (96,485 C/mol). The surface coverage of diazonium on the electrode is

calculated by dividing the number of moles of FBRR by the geometric area of the 200

m cylinder (3.2 x 10-5

cm2). This results in a typical coverage of 20 nmol/cm

2. As a

point of reference, usual monolayer coverage for a small, surface-bound organic

molecule has been reported as 300 pmol/cm2 (42). Therefore, this suggests a multilayer

deposition of FBRR onto the sensor, a result commonly observed for the reduction of aryl

diazonium salts onto carbon surfaces (13). The amount of FBRR deposited onto the

carbon-fiber surface (Table 6.1) is dependent on both the scan rate and potential window

of the voltammetric sweep. The voltammetric deposition of the diazonium is a time-

dependent process; therefore, scanning at slower rates or to an extended negative

waveform potential increases the amount of FBRR deposited onto the electrode surface.

There is no significant effect of varying the concentration of diazonium in solution (0.5-

5.0 mM) on the amount of FBRR deposited onto the electrode.

The presence of FBRR on the surface has been investigated using FSCV. In

Figure 6.1B, cyclic voltammograms recorded at a bare carbon-fiber microelectrode

(dashed black line) and the same microelectrode following modification with FBRR

(solid blue line) show a clear indication of the presence of the electroactive diazonium

salt on the electrode surface. The voltammogram of the FBRR redox system signifies

quasireversible behavior with an apparent formal potential of -0.1 V in AHL saline at

physiological pH. Integration of the oxidative peak area from the redox-active molecule

in Figure 6.1B results in an observed surface coverage of 40 pmol/cm2. This is

138

Page 155: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Table 6.1. Effect of varying voltammetric deposition parameters for FBRR reduction

onto a carbon-fiber surface.

[FBRR] = 2 mM in 0.1 M TEABF4/ACN. Error is SEM with n = 3 electrodes for each

measurement.

Scan rate (V/s) Potential window (V vs. Ag QRE) Γ (nmol/cm2)

0.500 +0.4 → -0.2 2.7 ± 0.9

0.500 +0.4 → -0.4 9.9 ± 3.5

0.500 +0.4 → -0.6 14.5 ± 1.4

0.500 +0.4 → -0.8 21.5 ± 2.7

0.500 +0.4 → -1.0 25.0 ± 2.7

0.050 +0.4 → -0.8 29.4 ± 6.4

0.100 +0.4 → -0.8 24.4 ± 1.2

0.500 +0.4 → -0.8 20.0 ± 2.1

1.0 +0.4 → -0.8 19.7 ± 0.6

5.0 +0.4 → -0.8 9.9 ± 1.4

139

Page 156: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

approximately two orders of magnitude smaller than that calculated from the diazonium

deposition in Figure 6.1A. A proposed mechanism for the oxidation-reduction reaction

of the surface-bound quinone derivative is illustrated in Scheme 6.2. It is thought that

voltammetric cycling of this molecule initially induces a two-electron/two-proton

oxidation to convert the p-methoxy moiety on the conjugated ring to its p-quinone

analogue. The quinone is then chemically reduced in a two-electron exchange to form

the hydroxy derivative of the molecule. Using the method of Nicholson (40), the

heterogeneous electron-transfer rate constant, k°, is determined to be 0.13 cm/s. This

indicates that the FBRR undergoes outer-sphere electron transfer on the carbon-fiber

surface, which is consistent with previous studies that have examined electron transfer

kinetics over a wide insulating layer (43).

Electrochemical characterization of the FBRR microelectrode pH sensor. The effect

of scan rate on the electrochemistry of a FBRR-modified carbon-fiber microelectrode has

been investigated using FSCV. Cyclic voltammograms of a FBRR microelectrode in pH

7.5 AHL saline solution at scan rates of 10, 20, and 50 V/s are plotted in Figure 6.2A.

Because current is directly proportional to scan rate, the current scale on the y-axis has

been divided by scan rate to provide a straightforward comparison of the peak positions

at the different scan rates. Notably, neither the anodic peak potential (Epa) nor the

cathodic peak potential (Epc) significantly shifts in value while varying scan rate in this

range. At scan rates higher than 100 V/s (up to 350 V/s), the Epa becomes more difficult

to identify due to a decrease in the ratio of the faradaic to the capacitive current. By

inspection, the Epa is well resolved from the background current at 20 V/s, a scan rate that

should suffice for monitoring rapidly occurring neurosecretory

140

Page 157: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Scheme 6.2. Proposed mechanism for the oxidation-reduction reaction of the surface-

bound quinone derivative of FBRR.

141

Page 158: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

events during in vivo applications. Therefore, this scan rate has been chosen to monitor

pH changes for the remainder of this chapter. Furthermore, the anodic peak current vs.

scan rate (Figure 6.2B) is linearly dependent for scan rates 10 – 350 V/s (r2 > 0.99). This

confirms that the oxidation and reduction of FBRR is a surface-confined reaction, as

expected, and provides evidence that the diazonium compound is sufficiently tethered to

the carbon-fiber surface.

The long-term stability of the FBRR microelectrode pH sensor has been studied

by continuously cycling modified electrodes in pH 7.5 AHL saline solution for 2.5 h (-0.7

to +0.8 V vs. Ag/AgCl at 5 Hz). This corresponds to 45,000 voltammetric sweeps over

the 2.5 h period. An 8% decrease in peak current is observed during the first 10 min of

cycling (Figure 6.2C). During the remaining 2.5 h, the peak current remains fairly stable,

decreasing by an additional 13%. Therefore, the stability of the FBRR-modified

microelectrode provides an ample time window for monitoring the pH in the CNS of

Drosophila during in vivo electrochemical measurements.

The selectivity of the sensor for H+ has been investigated to determine if alternate

ionic species present in the biological media could interfere with the voltammetric

response. To accomplish this, FBRR microelectrodes (n = 3) have been tested with

FSCV in a series of AHL saline solutions that contained elevated concentrations of

various inorganic cations. When the Na+ concentration in the AHL saline solution is

increased by 40%, the Epa remains unaltered. In addition, increasing the Mg2+

concentration by 45%, Ca2+

concentration by 50%, or K+ concentration by 60% does not

cause a shift in the Epa. These studies validate that changes in the concentration of these

four cations do not contribute to the measured shift in the Epa, which suggests charged

142

Page 159: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.2. Electrochemical characterization of the FBRR microelectrode pH sensor in

pH 7.5 AHL saline solution. (A) Cyclic voltammograms (average of 5 scans each) of the

FBRR redox couple at 3 scan rates. The current scale on the y-axis has been divided by

scan rate so the peak positions can be easily compared between the different scan rates.

In this scan rate range, neither the anodic peak potential (Epa) nor the cathodic peak

potential (Epc) significantly shifts in value. (B) Anodic peak current (ipa) as a function of

scan rate (SEM bars are too small to see). (C) The effect of continuous cycling of the

electrode on ipa. SEM bars are too small to see (n = 3).

143

Page 160: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

species, other than H+ ions, in the AHL saline solution are not affecting the pH response

of the FBRR microelectrode.

pH response of the FBRR microelectrode sensor. To calibrate the voltammetric

response of the sensor, the FBRR-modified carbon-fiber microelectrode has been

investigated in AHL saline solutions of varying pH. The peak characteristics of cyclic

voltammograms recorded over a pH range of 5.0 – 9.0 with a scan rate of 20 V/s have

been examined. Figure 6.3 depicts representative voltammograms of the FBRR

microelectrode in three different pH solutions. The Epa noticeably shifts to more negative

potentials as pH is increased. The Epc follows the same trend with pH as the Epa;

however, the peak becomes difficult to distinguish from the background current in higher

pH solutions ( pH 8), as reported previously for chemically modified electrodes in

physiological media (5). Therefore, the Epa was chosen as the identifier for the sensor

calibration and subsequent in vivo studies instead of the formal potential.

FSCV has been used to determine the response of the sensor to pH changes. In a

pH range of physiological relevance (6.5 – 8.0), a sigmoidal fit best describes the

relationship between Epa and pH (Figure 6.4, n = 9 electrodes). Linear regression of these

data yields a slope of 38 mV/pH unit which is less than the theoretical value of 59

mV/pH unit for a reversible, two-electron/two-proton redox reaction at room temperature

(44). This deviation from the predicted Nernstian value suggests the attachment of FBRR

to the carbon-fiber surface alters the electrochemistry of the quinone couple. pH-

sensitive, glassy carbon electrodes tethered with alternative reporter molecules have been

previously fabricated that exhibit expected Nernstian behavior, but practical limitations,

such as high capacitive currents, larger diameters (millimeter), and lengthy time scales to

144

Page 161: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.3. Cyclic voltammograms of a microelectrode modified with FBRR in AHL

saline solutions of different pH. Asterisk (*) corresponds to the Epa for each

voltammogram (20 V/s, average of 5 scans) with the dashed vertical line included for

comparison purposes. As the pH increases, the Epa visibly shifts to more negative

potentials. (A) pH 6.5 (B) pH 7.5 (C) pH 8.0.

145

Page 162: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.4. The anodic peak potential, Epa, as a function of AHL saline solution pH for

FBRR-modified electrodes. The Epa has a sigmoidal relationship with changing pH in a

physiological relevant pH range (6.5-8.0). Error bars are SEM (n = 9 electrodes).

146

Page 163: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

obtain stable readings, limit their biological usefulness (34, 36). For example, Shiu et. al.

have reported the development of a glassy carbon electrode (3 mm diameter) modified by

adsorption of an anthraquinonesulfonate film that possessed a near Nernstian slope of

56.4 mV/pH unit in aqueous pH buffers (34). However, it would not be feasible to use an

electrode of this size to measure dynamic events associated with in vivo neurosecretion in

volume-limited model systems such as Drosophila.

Microelectrode response time to a pH change. Flow-injection analysis has been used

to study the dynamic response of the sensor by introducing plugs of AHL saline solution

of varying pH to a FBRR-modified microelectrode. Ideally, a fast electrode response

time to a minor change in pH of the surrounding solution would produce a square-shaped

Epa vs. time trace. Figure 6.5 shows the Epa response of the modified electrode sensor to

0.2 pH unit changes. After initial immersion in an AHL saline solution of pH 7.4, the

electrode is exposed to a bolus of AHL saline solution of pH 7.2 (Figure 6.5A).

Likewise, AHL solution of pH 7.6 is introduced to the electrode in pH 7.4 solution in

Figure 6.5B. By inspection, the modified electrode response to a 0.2 change in pH is

square-like and consistent for measurement of either an acidic or a basic pH change.

Indeed, flow-injection calibration of the sensor revealed marked sensitivity for H+ with

the modified electrode capable of detecting pH changes as small as 0.005 (based on S/N

3) with a time response equal to 1.6 s (τ determined from exponential decay).

Measuring dynamic in vivo pH changes in the Drosophila CNS. I utilized the pH

electrode to monitor a dynamic pH change associated with neurotransmitter release in the

fly brain. Optogenetic stimulation using blue light with the mutant fly TH-

GAL4/UAS:ChR2 has been demonstrated to evoke dopamine release in Drosophila

147

Page 164: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.5. Plot of anodic peak potential vs. time during flow injection changes of 0.2

pH units in AHL saline. The electrode is able to consistently measure either an acidic or

a basic pH change. (A) Initial AHL saline solution of pH 7.4 is decreased to pH 7.2. (B)

Initial AHL saline solution of pH 7.4 is increased to pH 7.6.

148

Page 165: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

larvae by Venton and co-workers (45). In addition, a procedure for using optogenetic

stimulation with adult Drosophila was described in Chapter 5. The TH-

GAL4/UAS:ChR2 mutant fly expresses blue light sensitive cation channels which are

specific to dopaminergic neurons, allowing dopamine release to be elicited through timed

blue light stimulations. Following microsurgery, a micromanipulator is used to insert a

cylindrical FBRR-modified electrode into the CNS region of an adult mutant fly. A 5-s

stimulation with blue light is used to induce neurotransmitter release which causes a

change in Epa corresponding to an ~0.034 acidic pH change in the fly (Figure 6.6, solid

red line). This value is in agreement with pH fluctuations observed as a result of

stimulus-induced neurosecretion in rat brain slices (0.047 unit pH change in the cortex)

reported by the Wightman lab using ISMs (23). To ensure the response is due to a

biological change in the fly, the experiment has been repeated with the electrode in the

surrounding solution outside of the fly brain (solid black line). This experiment

demonstrates the high temporal sensitivity of the FBRR sensor and highlights its utility

for real-time analyses of pH fluctuations associated with neurotransmitter release in

volume-limited biological microsystems.

Conclusions

A carbon-fiber microelectrode pH sensor was developed via the voltammetric

reduction of the FBRR diazonium salt. The stability and sensitivity of the sensor for H+

was characterized in biological media set to a physiologically relevant pH range. FSCV

was used to probe the surface-bound diazonium derivative as a function of pH. The peak

corresponding to Epa for the FBRR-modified electrode was correlated to small changes in

149

Page 166: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 6.6. Physiological pH measurements in an adult Drosophila CNS. A

representative trace of a dynamic, acidic pH change associated with neurotransmitter

release being measured with a FBRR-modified electrode in a mutant fly CNS (red line).

A control stimulation of the same electrode in AHL saline solution only is plotted for

comparison (black line). The black arrow corresponds to a 5-s stimulation with blue

light.

150

Page 167: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

pH. Flow-injection analysis was used to characterize the temporal response of the sensor

in solutions of varying pH, resulting in a limit of detection to 0.005 pH units.

Furthermore, direct in vivo measurements of pH were made in the Drosophila CNS after

stimulated neurotransmitter release, revealing an acidic change in a brain region

dominated by dopaminergic neuron innervations. These data demonstrate the utility of

this easily fabricated sensor for measuring dynamic changes in extracellular pH in the fly

and other microanalytical animal models.

151

Page 168: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Kahlert, H. (2008) Functionalized carbon electrodes for pH determination, J.

Solid State Electrochem. 12, 1255-1266.

2. Jankowska, H., Neffe, S., and Swiatkowski, A. (1981) Investigations of the

electrochemical properties of activated carbon and carbon black, Electrochim.

Acta 26, 1861-1866.

3. Li, X., and Horita, K. (2000) Electrochemical characterization of carbon black

subjected to RF oxygen plasma, Carbon 38, 133-138.

4. Chen, P. H., and McCreery, R. L. (1996) Control of electron transfer kinetics at

glassy carbon electrodes by specific surface modification, Anal. Chem. 68, 3958-

3965.

5. Runnels, P. L., Joseph, J. D., Logman, M. J., and Wightman, R. M. (1999) Effect

of pH and surface functionalities on the cyclic voltammetric responses of carbon-

fiber microelectrodes, Anal. Chem. 71, 2782-2789.

6. Barbier, B., Pinson, J., Desarmot, G., and Sanchez, M. (1990) Electrochemical

bonding of amines to carbon-fiber surfaces toward improved carbon-epoxy

composites, J. Electrochem. Soc. 137, 1757-1764.

7. Andrieux, C. P., Gonzalez, F., and Saveant, J. M. (1997) Derivatization of carbon

surfaces by anodic oxidation of arylacetates. Electrochemical manipulation of the

grafted films, J. Am. Chem. Soc. 119, 4292-4300.

8. Buttry, D. A., Peng, J. C. M., Donnet, J. B., and Rebouillat, S. (1999)

Immobilization of amines at carbon-fiber surfaces, Carbon 37, 1929-1940.

9. Adenier, A., Chehimi, M. M., Gallardo, I., Pinson, J., and Vila, N. (2004)

Electrochemical oxidation of aliphatic amines and their attachment to carbon and

metal surfaces, Langmuir 20, 8243-8253.

10. Delamar, M., Hitmi, R., Pinson, J., and Saveant, J. M. (1992) Covalent

modification of carbon surfaces by grafting of functionalized aryl radicals

produced from electrochemical reduction of diazonium salts, J. Am. Chem. Soc.

114, 5883-5884.

11. Allongue, P., Delamar, M., Desbat, B., Fagebaume, O., Hitmi, R., Pinson, J., and

Saveant, J. M. (1997) Covalent modification of carbon surfaces by aryl radicals

generated from the electrochemical reduction of diazonium salts, J. Am. Chem.

Soc. 119, 201-207.

12. Delamar, M., Desarmot, G., Fagebaume, O., Hitmi, R., Pinson, J., and Saveant, J.

M. (1997) Modification of carbon fiber surfaces by electrochemical reduction of

aryl diazonium salts: Application to carbon epoxy composites, Carbon 35, 801-

807.

13. Downard, A. J. (2000) Electrochemically assisted covalent modification of carbon

electrodes, Electroanalysis 12, 1085-1096.

14. Kariuki, J. K., and McDermott, M. T. (2001) Formation of multilayers on glassy

carbon electrodes via the reduction of diazonium salts, Langmuir 17, 5947-5951.

15. Baranton, S., and Belanger, D. (2005) Electrochemical derivatization of carbon

surface by reduction of in situ generated diazonium cations, J. Phys. Chem. B 109,

24401-24410.

152

Page 169: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

16. Pinson, J., and Podvorica, F. (2005) Attachment of organic layers to conductive or

semiconductive surfaces by reduction of diazonium salts, Chem. Soc. Rev. 34,

429-439.

17. Kissinger, P. T., Hart, J. B., and Adams, R. N. (1973) Voltammetry in brain

tissue- new neurophysiological measurement, Brain Res. 55, 209-213.

18. Cass, W. A., Hudson, J., Henson, M., Zhang, Z., Ovadia, A., Hoffer, B. J., and

Gash, D. M. (1996) In vivo electrochemical studies of dopamine overflow and

clearance in the striatum of normal and MPTP-treated rhesus monkeys, J.

Neurochem. 66, 579-588.

19. Zhang, L. F., Doyon, W. M., Clark, J. J., Phillips, P. E. M., and Dani, J. A. (2009)

Controls of tonic and phasic dopamine transmission in the dorsal and ventral

striatum, Mol. Pharmacol. 76, 396-404.

20. Nagy, G., Moghaddam, B., Oke, A., and Adams, R. N. (1985) Simultaneous

monitoring of voltammetric and ion-selective electrodes in mammalian brain,

Neurosci. Lett. 55, 119-124.

21. Rice, M. E., and Nicholson, C. (1989) Measurement of nanomolar dopamine

diffusion using low-noise perfluorinated ionomer coated carbon-fiber

microelectrodes and high-speed cyclic voltammetry, Anal. Chem. 61, 1805-1810.

22. Adams, R. N. (1990) In vivo electrochemical measurements in the CNS, Prog.

Neurobiol. 35, 297-311.

23. Jones, S. R., Mickelson, G. E., Collins, L. B., Kawagoe, K. T., and Wightman, R.

M. (1994) Interference by pH and Ca2+

ions during measurements of

catecholamine release in slices of rat amygdala with fast-scan cyclic voltammetry,

J Neurosci Methods 52, 1-10.

24. Downard, A. J., Roddick, A. D., and Bond, A. M. (1995) Covalent modification

of carbon electrodes for voltammetric differentiation of dopamine and ascorbic

acid, Anal. Chim. Acta 317, 303-310.

25. Heien, M. L. A. V., Johnson, M. A., and Wightman, R. M. (2004) Resolving

neurotransmitters detected by fast-scan cyclic voltammetry, Anal. Chem. 76,

5697-5704.

26. Wilson, G. S., and Johnson, M. A. (2008) In-vivo electrochemistry: What can we

learn about living systems?, Chem. Rev. 108, 2462-2481.

27. Urbanics, R., Lenigerfollert, E., and Lubbers, D. W. (1978) Time course of

changes of extracellular H+ and K

+ activities during and after direct electrical

stimulation of the brain cortex, Pflugers Arch. 378, 47-53.

28. Chen, J. C., and Chesler, M. (1992) pH transients evoked by excitatory synaptic

transmission are increased by inhibition of extracellular carbonic anhydrase, Proc.

Natl. Acad. Sci. U.S.A. 89, 7786-7790.

29. Chesler, M., and Kaila, K. (1992) Modulation of pH by neuronal activity, Trends

Neurosci. 15, 396-402.

30. Piyankarage, S. C., Augustin, H., Grosjean, Y., Featherstone, D. E., and Shippy,

S. A. (2008) Hemolymph amino acid analysis of individual Drosophila larvae,

Anal. Chem. 80, 1201-1207.

31. Makos, M. A., Kim, Y. C., Han, K. A., Heien, M. L., and Ewing, A. G. (2009) In

vivo electrochemical measurements of exogenously applied dopamine in

Drosophila melanogaster, Anal. Chem. 81, 1848-1854.

153

Page 170: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

32. Makos, M. A., Han, K.-A., Heien, M. L., and Ewing, A. G. (2009) Using in vivo

electrochemistry to study the physiological effects of cocaine and other stimulants

on the Drosophila melanogaster dopamine transporter, ACS Chem. Neurosci. 1,

74-83.

33. Katz, E., Lion-Dagan, M., and Willner, I. (1996) pH-switched electrochemistry of

pyrroloquinoline quinone at Au electrodes modified by functionalized

monolayers, J. Electroanal. Chem. 408, 107-112.

34. Shiu, K.-K., Song, F., and Dai, H.-P. (1996) Potentiometric pH sensor with

anthraquinonesulfonate adsorbed on glassy carbon electrodes, Electroanalysis 8,

1160-1164.

35. Hay, S., Westerlund, K., and Tommos, C. (2007) Redox characteristics of a de

novo quinone protein, J. Phys. Chem. B 111, 3488-3495.

36. Holm, A. H., Vase, K. H., Winther-Jensen, B., Pedersen, S. U., and Daasbjerg, K.

(2007) Evaluation of various strategies to formation of pH responsive

hydroquinone-terminated films on carbon electrodes, Electrochim. Acta 53, 1680-

1688.

37. Pandurangappa, M., Lawrence, N. S., and Compton, R. G. (2002) Homogeneous

chemical derivatization of carbon particles: A novel method for functionalizing

carbon surfaces, Analyst 127, 1568-1571.

38. Wang, J. W., Wong, A. M., Flores, J., Vosshall, L. B., and Axel, R. (2003) Two-

photon calcium imaging reveals an odor-evoked map of activity in the fly brain,

Cell 112, 271-282.

39. Heien, M. L. A. V., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman,

R. M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

40. Nicholson, R. S. (1965) Theory and application of cyclic voltammetry for

measurement of electrode reaction kinetics, Anal. Chem. 37, 1351-1355.

41. Schroll, C., Riemensperger, T., Bucher, D., Ehmer, J., Voller, T., Erbguth, K.,

Gerber, B., Hendel, T., Nagel, G., Buchner, E., and Fiala, A. (2006) Light-

induced activation of distinct modulatory neurons triggers appetitive or aversive

learning in Drosophila larvae, Curr. Biol. 16, 1741-1747.

42. Soriaga, M. P., and Hubbard, A. T. (1982) Determination of the orientation of

adsorbed molecules at solid-liquid interfaces by thin-layer electrochemistry-

Aromatic compounds at platinum electrodes, J. Am. Chem. Soc. 104, 2735-2742.

43. Yang, H. H., and McCreery, R. L. (1999) Effects of surface monolayers on the

electron-transfer kinetics and adsorption of methyl viologen and phenothiazine

derivatives on glassy carbon electrodes, Anal. Chem. 71, 4081-4087.

44. Laviron, E. (1983) Electrochemical reactions with protonations at equilibrium:

Part VIII. The 2e, 2H+ reaction (nine-member square scheme) for a surface or for

a heterogeneous reaction in the absence of disproportionation and dimerization

reactions, J. Electroanal. Chem. 146, 15-36.

45. Vickrey, T. L., Condron, B., and Venton, B. J. (2009) Detection of endogenous

dopamine changes in Drosophila melanogaster using fast-scan cyclic

voltammetry, Anal. Chem. 81, 9306-9313.

154

Page 171: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Chapter 7: Future Directions for Quantifying Neurochemicals in

Drosophila Using Electrochemical Detection

155

Page 172: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Drosophila is a useful model system for studying several human physiological

processes including addiction. Many central nervous system (CNS) pathways in flies and

mammals are evolutionarily conserved because of the genetic similarity between the two

species. Research has demonstrated fruit flies exhibit behavioral responses to

psychostimulants that are amazingly comparable to human behaviors. The overall goal

of my thesis was to develop methods capable of quantifying neurochemicals in

Drosophila. Chapters 2-4 were focused on measuring changes in uptake of exogenously

applied dopamine in the fly CNS in the presence of drugs of abuse. Approaches for

stimulating release of endogenous dopamine in Drosophila were investigated in Chapter

5. Chapter 6 described the development of a microelectrode pH sensor for monitoring in

vivo pH fluctuations associated with neurotransmitter release. These methods could lead

to a more analytical view of the basis behind addiction. In this chapter, I will discuss the

future directions of this project with respect to three aspects: biological application,

kinetics of dopamine uptake, and stimulating dopamine release.

Investigating Alcohol Addiction with Drosophila

The majority of the applications discussed in my thesis involved cocaine and

amphetamine addiction. One future application of this project is to utilize the tools I

developed in conjunction with recently identified Drosophila mutants to investigate the

mechanisms underlying alcohol tolerance and abuse.

Addiction is defined as compulsive drug use that has escalated to a level the user

can no longer control with drug use persisting despite significant negative consequences.

One mechanism of addiction is activation of dopaminergic fibers in the brain (1).

156

Page 173: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Additional psychomotor actions occur which vary depending on the particular addictive

substance in question. Ethanol affects γ-aminobutyric acid (GABA) receptors in the CNS

by increasing their usual function (2-4). Because GABA is an inhibitory

neurotransmitter, this leads to a depression in CNS activity. In addition, ethanol interacts

with N-methyl-D-aspartate (NMDA) receptors. This prevents the action of glutamate, an

excitatory neurotransmitter, on NMDA receptors which causes a further decrease in CNS

function (5, 6). In the brain, extracellular dopamine increases with alcohol consumption

because ethanol is thought to stimulate dopamine release in certain regions of the CNS

(7, 8).

While the main actions of ethanol in the brain are known, less is understood

regarding the changes in neuronal activity following short-term ethanol exposure and

their contribution towards alcohol tolerance and sensitivity (9). Alcohol addiction has a

strong genetic correlation, and there remains much to be discovered about the specific

genes that increase the genetic risk of a person developing an addiction to alcohol as well

(10, 11). Because the Drosophila genome possesses little genetic redundancy, it is an

attractive model system for identifying individual genes that influence particular

behaviors (12, 13). Additionally, the behavioral response of flies to ethanol has been

shown to model that of mammals (14-16).

Recently, several mutant fly types have been developed by Heberlein and

coworkers that exhibit unique behavioral responses toward alcohol consumption (17-19).

Table 7.1 is a summary of the modified behaviors the genetically altered flies cheapdate,

tipsy, barfly, and hangover display following exposure to ethanol. The effects of ethanol

on fly behavior were measured using a fly inebriometer (Figure 7.1). This home built

157

Page 174: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Table 7.1. Drosophila mutants that display altered behavioral responses to ethanol.

Fly mutant Modified behavior References

cheapdate increased sensitivity to alcohol

(17)

tipsy increased sensitivity to alcohol

(18)

barfly reduced sensitivity to alcohol

(18)

hangover reduced development of

tolerance to alcohol

(19)

158

Page 175: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 7.1. The fly inebriometer. This device is used to measure changes in fly postural

control upon ethanol exposure. (Reprinted from (18), with permission from John Wiley

and Sons).

159

Page 176: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

apparatus allows the ethanol-induced loss of postural control of flies to be quantified (18,

20). Flies are introduced into the top of a 4-ft glass column where a controlled

concentration of ethanol is circulated. Over time, flies become intoxicated and lose their

postural control. They tumble down the column where they elute out the bottom and are

counted. Fly research labs can analyze the effects of ethanol on the motor control of

hundreds of flies simultaneously with this device.

A future proposal might be to use in vivo electrochemical detection to quantify

uptake of exogenously applied dopamine in the mutant flies cheapdate, tipsy, barfly, and

hangover. The mutants could also be studied following short-term exposure to ethanol in

a fly inebriometer. Comparison of these data to wild-type flies exposed to identical

ethanol concentrations will provide information about the neurochemical changes behind

the altered behavioral response of the mutant flies toward ethanol. As evidence suggests

that neurotransmitter systems affected by ethanol are conserved between flies and

humans (21), this could potentially lead to a better understanding of the role genes

affecting dopamine neurotransmission play in alcohol addiction. While no animal model

is a perfect model for alcoholism, utilizing the genetic advantages of the fruit fly will

allow aspects of this complex disease to be studied and will give insight into the effects

of ethanol on the CNS.

Quantifying the Kinetics of Dopamine Uptake in Drosophila

In vivo uptake has been characterized in the rat brain using experimental data and

has also been modeled by simulations (22-26). Following the work reported in this

160

Page 177: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

thesis, a future investigation might be to model dopamine uptake in the Drosophila CNS,

and to compare it with experimentally obtained results.

Modeling neurotransmitter uptake involves understanding the relative importance

of diffusion vs. uptake processes which can be mathematically examined with the classic

Michaelis-Menten equation (27). Figure 7.2A is a depiction of a typical Michaelis-

Menten plot that can be used to determine kinetic parameters, such as Vmax and Km, for

simple kinetic behavior. Michaelis-Menten kinetics involves assumptions based on

Fickian diffusion, which must be kept in mind when examining physiological processes.

It has been demonstrated that ion diffusion through medium as complex as the

extracellular space of the brain cannot necessarily be assumed to obey Fick’s Laws (28,

29). Tortuosity, or the extent to which diffusing particles are hindered by obstructions in

their path, has been shown to affect small cations moving through extracellular space. In

addition, volume averaging takes into account the variations in extracellular vs.

intracellular space. It has been suggested that equations originally developed to describe

macroscopic problems are valid to describe uptake in the CNS when tortuosity and

volume averaging are taken into account (28, 30, 31).

A suggestion for a future direction is to model dopamine uptake and quantify the

kinetic parameters Vmax and Km for Drosophila to provide a better understanding of in

vivo measurements in the fly CNS. Quantification of the measured dopamine signal can

be divided into two phases: the rising phase and the falling phase (Figure 7.2B). The

rising phase consists of the amount of dopamine that is transported into the tissue and is

oxidized at the electrode surface. The falling phase of the signal is due to the clearance

of dopamine from the tissue. This is a combination of the uptake, metabolism, and

161

Page 178: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 7.2. Modeling dopamine uptake. (A) The classic Michaelis-Menten plot for

determining kinetic parameters Vmax and Km. (B) Representative signal measured during

dopamine uptake.

162

Page 179: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

diffusion of dopamine out of the tissue, making the falling phase of the signal a nontrivial

component to accurately model.

Improving the Detection of Stimulated Dopamine Release in Drosophila

The electrochemical measurements of dopamine reported in my thesis were made

with a 5 µm diameter carbon-fiber electrode. This ~50 µm long cylindrical

microelectrode adequately detected the changes in uptake of exogenously applied

dopamine described in Chapters 2-4; however, the measured signals from dopamine

released via optical stimulation in Chapter 5 were less robust. It is likely that when the

endogenous dopamine released from the fly CNS reaches the microelectrode surface it is

of a lower concentration than the dopamine measured in exogenously applied

experiments. A future step in the analytical development of measuring optically

stimulated dopamine release in Drosophila is to increase the sensitivity of the working

electrode to improve in vivo detection limits for dopamine.

The sensitivity of carbon-fiber microelectrodes is dependent on several properties

of the electrode including the size and the surface roughness or chemistry. Increasing

electrode sensitivity by using a methane/oxygen flame to etch carbon-fiber electrodes

down to < 1 µm diameters has been demonstrated previously by our laboratory (32).

Decreasing the overall size of the working electrode results in lower signal due to

background current since double layer charging current is proportional to surface area, as

well as lower signal arising from the analyte of interest (33). By increasing surface

roughness of a carbon-fiber electrode of a given size, the sensitivity for measuring

dopamine can be improved. Electrochemical over-oxidation of carbon increases the

163

Page 180: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

surface roughness which allows the addition of more oxide groups (34, 35). These oxide

groups act as adsorption sites for cationic species, such as dopamine, thus leading to

increased sensitivity of the electrode. Typically, high positive potentials are avoided

when using carbon electrodes in biological applications to prevent the interference of

water oxidation which has been observed to cause instability and inactivation of pyrolytic

graphite electrodes (36). Several groups have reported nanomolar detection limits for

dopamine using carbon-fiber microelectrodes scanned to a positive potential of 1.4 V

instead of the more traditional 1.0 V (37-39). In addition, carbon-fiber microelectrodes

that were both flame-etched and electrochemically overoxidized have been successfully

used to measure in vivo dopamine concentrations of < 25 nM in the rat CNS (39).

A future direction here would be to use a methane/oxygen flame to etch a 5 µm

diameter working electrode down to ~1 µm diameter. This electrode would then be used

as the working electrode to measure dopamine release in vivo the Drosophila CNS

following optical stimulation. In addition, the anodic scanning potential of the applied

waveform would be increased to 1.4 V. Preliminary work measuring dopamine using

fast-scan cyclic voltammetry with a 5-µm carbon-fiber microelectrode scanned to 1.4 V

has demonstrated a significant increase in signal over dopamine measured with a

waveform scanned to 1.0 V (Figure 7.3A). An improvement in voltammogram shape has

been observed for exogenously applied dopamine measured in the Drosophila mushroom

bodies when the applied waveform is extended to 1.4 V as well (Figure 7.3B).

Decreasing the size of the electrode and extending the waveform to a more positive

potential will improve detection limits for measuring optically stimulated dopamine

release in Drosophila.

164

Page 181: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Figure 7.3. Voltammetric measurements of dopamine using an applied waveform of 1.0

V vs. a waveform extended to 1.4 V. (A) Comparison of an identical dopamine

concentration measured by a 5-µm carbon-fiber microelectrode scanned -0.6 V to 1.0 V

vs. -0.6 V to 1.4 V (mean ± SEM; Student’s t-test, p = 0.0004 (***), n = 4 measurements for

each potential). (B) Cyclic voltammograms (200 V/s, average of 5 scans each) of

exogenously applied dopamine measured in vivo the Drosophila CNS with an applied

waveform of 1.0 V (black line) and a waveform extended to 1.4 V (red line).

165

Page 182: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

References

1. Wise, R. A., and Bozarth, M. A. (1987) A psychomotor stimulant theory of

addiction, Psychol. Rev. 94, 469-492.

2. Hunt, W. A. (1983) The effect of ethanol on GABAergic transmission, Neurosci.

Biobehav. Rev. 7, 87-95.

3. Deitrich, R. A., Dunwiddie, T. V., Harris, R. A., and Erwin, V. G. (1989)

Mechanism of action of ethanol: Initial central nervous system actions,

Pharmacol. Rev. 41, 489-537.

4. Faingold, C. L., N'Gouemo, P., and Riaz, A. (1998) Ethanol and neurotransmitter

interactions-from molecular to integrative effects, Prog. Neurobiol. 55, 509-535.

5. Lovinger, D., White, G., and Weight, F. (1989) Ethanol inhibits NMDA-activated

ion current in hippocampal neurons, Science 243, 1721-1724.

6. Peoples, R. W., and Weight, F. F. (1995) Cutoff in potency implicates alcohol

inhibition of N-methyl-D-aspartate receptors in alcohol intoxication, Proc. Natl.

Acad. Sci. U.S.A. 92, 2825-2829.

7. Imperato, A., and Di Chiara, G. (1986) Preferential stimulation of dopamine

release in the nucleus accumbens of freely moving rats by ethanol, J. Pharmacol.

Exp. Ther. 239, 219-228.

8. Weiss, F., Lorang, M. T., Bloom, F. E., and Koob, G. F. (1993) Oral alcohol self-

administration stimulates dopamine release in the rat nucleus accumbens: Genetic

and motivational determinants, J. Pharmacol. Exp. Ther. 267, 250-258.

9. Kong, E. C., Allouche, L., Chapot, P. A., Vranizan, K., Moore, M. S., Heberlein,

U., and Wolf, F. W. (2010) Ethanol-regulated genes that contribute to ethanol

sensitivity and rapid tolerance in Drosophila, Alcohol. Clin. Exp. Res. 34, 302-

316.

10. Crabbe, J. C. (2002) Genetic contributions to addiction, Annu. Rev. Psychol. 53,

435-462.

11. Le Foll, B., Gallo, A., Le Strat, Y., Lu, L., and Gorwood, P. (2009) Genetics of

dopamine receptors and drug addiction: a comprehensive review, Behav.

Pharmacol. 20, 1-17.

12. Sokolowski, M. B. (2001) Drosophila: Genetics meets behavior, Nat. Rev. Genet.

2, 879-890.

13. Waddell, S., and Quinn, W. G. (2001) What can we teach Drosophila? What can

they teach us?, Trends Genet. 17, 719-726.

14. Wolf, F. W., and Heberlein, U. (2003) Invertebrate models of drug abuse, J.

Neurobiol. 54, 161-178.

15. Scholz, H., Ramond, J., Singh, C. M., and Heberlein, U. (2000) Functional

ethanol tolerance in Drosophila, Neuron 28, 261-271.

16. Devineni, A. V., and Heberlein, U. (2009) Preferential ethanol consumption in

Drosophila models features of addiction, Curr. Biol. 19, 2126-2132.

17. Moore, M. S., DeZazzo, J., Luk, A. Y., Tully, T., Singh, C. M., and Heberlein, U.

(1998) Ethanol intoxication in Drosophila: Genetic and pharmacological evidence

for regulation by the cAMP signaling pathway, Cell 93, 997-1007.

18. Singh, C. M., and Heberlein, U. (2000) Genetic control of acute ethanol-induced

behaviors in Drosophila, Alcohol. Clin. Exp. Res. 24, 1127-1136.

166

Page 183: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

19. Scholz, H., Franz, M., and Heberlein, U. (2005) The hangover gene defines a

stress pathway required for ethanol tolerance development, Nature 436, 845-847.

20. Cohan, F. M., and Hoffmann, A. A. (1986) Genetic divergence under uniform

selection. II. Different responses to selection for knockdown resistance to ethanol

among Drosophila melanogaster populations and their replicate lines, Genetics

114, 145-164.

21. Bainton, R. J., Tsai, L. T. Y., Singh, C. M., Moore, M. S., Neckameyer, W. S.,

and Heberlein, U. (2000) Dopamine modulates acute responses to cocaine,

nicotine, and ethanol in Drosophila, Curr. Biol. 10, 187-194.

22. Justice Jr, J. B., Nicolaysen, L. C., and Michael, A. C. (1988) Modeling the

dopaminergic nerve terminal, J. Neurosci. Methods 22, 239-252.

23. Wightman, R. M., Amatorh, C., Engstrom, R. C., Hale, P. D., Kristensen, E. W.,

Kuhr, W. G., and May, L. J. (1988) Real-time characterization of dopamine

overflow and uptake in the rat striatum, Neuroscience 25, 513-523.

24. Cass, W. A., Zahniser, N. R., Flach, K. A., and Gerhardt, G. A. (1993) Clearance

of exogenous dopamine in rat dorsal striatum and nucleus accumbens: Role of

metabolism and effects of locally applied uptake inhibitors, J. Neurochem. 61,

2269-2278.

25. Jones, S. R., Garris, P. A., and Kilts, C. D. (1995) Comparison of dopamine

uptake in the basolateral amygdaloid nucleus, caudate-putamen, and nucleus

accumbens of the rat, J. Neurochem. 64, 2581-2589.

26. Wu, Q., Reith, M. E. A., Wightman, R. M., Kawagoe, K. T., and Garris, P. A.

(2001) Determination of release and uptake parameters from electrically evoked

dopamine dynamics measured by real-time voltammetry, J. Neurosci. Methods

112, 119-133.

27. Michaelis, L., and Menten, M. L. (1913) The kenetics of the inversion effect,

Biochem. Z. 49, 333-369.

28. Nicholson, C., and Phillips, J. M. (1981) Ion diffusion modified by tortuosity and

volume fraction in the extracellular microenvironment of the rat cerebellum, J.

Physiol. (Lond.) 321, 225-257.

29. Nicholson, C. (1985) Diffusion from an injected volume of a substance in brain

tissue with arbitrary volume fraction and tortuosity, Brain Res. 333, 325-329.

30. Rice, M. E., and Nicholson, C. (1991) Diffusion characteristics and extracellular

volume fraction during normoxia and hypoxia in slices of rat neostriatum, J.

Neurophysiol. 65, 264-272.

31. Nicholson, C. (1995) Interaction between diffusion and Michaelis-Menten uptake

of dopamine after iontophoresis in striatum, Biophys. J. 68, 1699-1715.

32. Strein, T. G., and Ewing, A. G. (1992) Characterization of submicron-sized

carbon electrodes insulated with a phenol-allylphenol copolymer, Anal. Chem. 64,

1368-1373.

33. Bard, A. J., and Faulkner, L. R. (2001) Electrochemical Methods: Fundamentals

and Applications, 2nd ed., John Wiley & Sons, Hoboken, NJ.

34. Gonon, F. G., Fombarlet, C. M., Buda, M. J., and Pujol, J. F. (1981)

Electrochemical treatment of pyrolytic carbon-fiber electrodes, Anal. Chem. 53,

1386-1389.

167

Page 184: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

35. Brajter-Toth, A., El-Nour, K. A., Cavalheiro, E. T., and Bravo, R. (2000)

Nanostructured carbon-fiber disk electrodes for sensitive determinations of

adenosine and uric acid, Anal. Chem. 72, 1576-1584.

36. Goss, C. A., Brumfield, J. C., Irene, E. A., and Murray, R. W. (1993) Imaging the

incipient electrochemical oxidation of highly oriented pyrolytic graphite, Anal.

Chem. 65, 1378-1389.

37. Hafizi, S., Kruk, Z. L., and Stamford, J. A. (1990) Fast cyclic voltammetry:

Improved sensitivity to dopamine with extended oxidation scan limits, J.

Neurosci. Methods 33, 41-49.

38. Heien, M. L., Phillips, P. E. M., Stuber, G. D., Seipel, A. T., and Wightman, R.

M. (2003) Overoxidation of carbon-fiber microelectrodes enhances dopamine

adsorption and increases sensitivity, Analyst 128, 1413-1419.

39. Strand, A. M., and Venton, B. J. (2008) Flame etching enhances the sensitivity of

carbon-fiber microelectrodes, Anal. Chem. 80, 3708-3715.

168

Page 185: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Appendix

1) Calculations for exogenously applied dopamine:

The radius was determined by ejecting the dopamine solution into mineral oil and

measuring the diameter of the bubble that formed. This was tested for both the

single-barrel and three-barrel glass micropipets. Micropipets were manually cut

so each opening would be ~ 5 μm.

volume = V = 4/3πr3 = 4/3π(0.065 cm / 2)

3

V = 1.44x10-4

cm3 ~ 150 nL

150 nL of 1.0 mM DA = 150 pmol dopamine applied

2) Electrode calibration plot:

slope = 0.69 ± 0.04 nA/μM

r2 = 0.98

n = 3

3) FSCV volume sampled:

Diffusion layer radius (δ) = (2Dt)1/2

= (2* 5x10-6

cm2/s * 0.007 s)

1/2 = 2.6x10

-4 cm

δ + radius carbon fiber = r = 2.6x10-4

cm + 2.5x10-4

cm = 5.1x10-4

cm

volume δ and carbon fiber = V = πr2h = π(5.1x10

-4 cm)

2 * 5.0x10

-3 cm = 4.1 pL

volume carbon fiber = V = πr2h = π(2.5x10

-4 cm)

2 * 5.0x10

-3 cm = 0.98 pL

volume sampled = 4.1 pL – 0.98 pL ~ 3 pL

Amperometry volume sampled:

Diffusion layer radius = δ ~ 6r (where r = radius of carbon fiber)

= 6 * 2.5x10-4

cm = 1.5x10-3

cm

δ + radius carbon fiber = r = 1.5x10-3

cm + 2.5x10-4

cm = 1.8x10-3

cm

volume δ and carbon fiber = V = πr2h = π(1.8x10

-3 cm)

2 * 5.0x10

-3 cm = 50.1 pL

volume carbon fiber = V = πr2h = π(2.5x10

-4 cm)

2 * 5.0x10

-3 cm = 0.98 pL

volume sampled = 50.1 pL – 0.98 pL ~ 50 pL

0 10 20 30 40 500

10

20

30

40

[DA] M

i ma

x (

nA

)

169

Page 186: IN VIVO ELECTROCHEMICAL MEASUREMENTS IN

Vita: Monique Adrianne Makos

EDUCATION:

Ph.D. in Chemistry, May 2010, The Pennsylvania State University

B.S. in Chemistry, May 2005, The University of Texas at Austin

AWARDS:

Society for Electroanalytical Chemistry Graduate Student Award, SEAC (2010)

Norma Robinson Award for outstanding graduate research, PSU (2009)

Travel Award for oral presentation, PSU (2007)

The Roberts Award for select incoming graduate students, PSU (2005)

PUBLICATIONS:

Makos MA, Heien ML, Ewing AG “Oral Administration of Methylphenidate

Blocks the Effect of Cocaine on Uptake at the Drosophila Dopamine Transporter”

ACS Chemical Neuroscience, in preparation.

Makos MA, Omiatek DM, Ewing AG, Heien ML “Development and

Characterization of a Voltammetric Carbon-Fiber Microelectrode pH Sensor”

Langmuir, accepted.

Makos MA, Han KA, Heien ML, Ewing AG “Using In Vivo Electrochemistry to

Study the Physiological Effects of Cocaine and Other Stimulants on the

Drosophila melanogaster Dopamine Transporter” ACS Chemical Neuroscience

2010, 1, 74-83.

Makos MA, Kuklinski NJ, Berglund EC, Heien ML, Ewing AG “Chemical

Measurements in Drosophila” TrAC Trends in Analytical Chemistry 2009, 28,

1223-1234.

Makos MA, Kim YC, Han KA, Heien ML, Ewing AG “In Vivo Electrochemical

Measurements of Exogenously Applied Dopamine in Drosophila melanogaster”

Analytical Chemistry 2009, 81, 1848-1854.

ORAL PRESENTATIONS:

Makos MA, Heien ML, Han KA, Ewing AG “Quantifying Real-Time

Neurotransmitter Changes in the Central Nervous System of Drosophila

melanogaster Using Fast-Scan Cyclic Voltammetry” ACS invited session at

Pittcon, Orlando, FL. March 2010.

Makos MA, Kim YC, Han KA, Heien ML, Ewing AG “In Vivo Electrochemical

Monitoring of Dopamine Uptake in Drosophila melanogaster” Pittcon, Chicago,

IL. March 2009.

Makos MA, Kim YC, Han KA, Ewing AG “In Vivo Electrochemistry in the 8-nL

Brain of the Fruit Fly” Pittcon, Chicago, IL. February 2007.